DOI:
10.1039/D5MH00453E
(Review Article)
Mater. Horiz., 2025,
12, 5994-6017
Shape-morphing bioelectronic devices
Received
13th March 2025
, Accepted 2nd May 2025
First published on 5th May 2025
Abstract
Shape-morphing bioelectronic devices, which can actively transform their geometric configurations in response to external stimuli (e.g., light, heat, electricity, and magnetic fields), have enabled many unique applications in different areas, ranging from human–machine interfaces to biomedical applications. These devices can not only realize in vivo deformations to execute specific tasks, form conformal contacts with target organs for real-time monitoring, and dynamically reshape their structures to adjust functional properties, but also assist users in daily activities through physical interactions. In this review, we provide a comprehensive overview of recent advances in shape-morphing bioelectronic devices, covering their fundamental working principles, representative deformation modes, and advanced manufacturing methodologies. Then, a broad range of practical applications of shape-morphing bioelectronics are summarized, including electromagnetic devices, optoelectronic devices, biological devices, biomedical devices, and haptic interfaces. Finally, we discuss key challenges and emerging opportunities in this rapidly evolving field, providing insights into future research directions and potential breakthroughs.

Shiwei Xu
| Shiwei Xu obtained his BS degree in engineering mechanics from Huazhong University of Science and Technology in 2020. He is currently pursuing his PhD degree in solid mechanics at Tsinghua University. His research interests include soft robotics, reconfigurable electronics, and mechanics of flexible structures. |

Ruoxi Yang
| Ruoxi Yang obtained her PhD degree in mechanical engineering from Hebei University of Technology in 2023. She is currently a postdoctoral fellow of engineering mechanics at Tsinghua University. Her research interests include the design and fabrication of flexible sensors and actuators, and the development of intelligent control systems based on these electronics. |

Youzhou Yang
| Youzhou Yang obtained both his BS degree in electronic science and technology and ME in microelectronics and solid state electronics from Huazhong University of Science and Technology in 2018 and 2022, respectively. He is currently pursuing his PhD degree in solid mechanics at Tsinghua University. His research interests include mechanically guided 3D assembly for magnetically controlled micro-scale bio-robotics and organoid/tissue engineering. |

Yihui Zhang
| Yihui Zhang obtained his PhD in solid mechanics from the Department of Engineering Mechanics at Tsinghua University in 2011. Then he worked as a postdoctoral fellow from 2011 to 2014 and as a research assistant professor from 2014 to 2015, both at Northwestern University. He is a professor of engineering mechanics and Vice Director of State Key Laboratory of Flexible Electronics Technology at Tsinghua University. His research interests include mechanically guided 3D assembly, stretchable electronics, interface mechanics, microrobots, mechanical metamaterials, and mechanics of flexible structures. |
Wider impact
The active shape-morphing capabilities of bioelectronic devices have significantly expanded the functionalities of conventional devices with fixed geometric configurations. However, existing reviews lack a comprehensive summary of the design concepts, fabrication methods, and application scenarios of shape-morphing bioelectronic devices. This review systematically explores the shape-morphing mechanisms of these devices, covering the working principles of active materials/morphable structures, representative deformation modes, and manufacturing methodologies. Additionally, we highlight their diverse applications across electromagnetic devices, optoelectronic devices, biological devices, biomedical devices, and haptic interfaces. Furthermore, key future directions are discussed, including the development of novel materials, the establishment of inverse design methods and reprogramming strategies for shape morphing, the improvement of actuation accuracy and speed, and the development of biocompatible and comfortable actuation mechanisms. We hope this review provides insights for researchers to drive the advancement of shape-morphing bioelectronic devices through the design of novel materials and advanced structures, the enhancement of shape-morphing capabilities, the innovation of integration/fabrication methods, and the expansion of potential application scenarios.
|
1. Introduction
Flexible bioelectronics represents an area of intensive fundamental and practical interest, because these kinds of transformative technologies have a profound impact on our daily life, spanning areas like healthcare (e.g., physical signal monitoring,1–4 sweat biochemical sensing,5 and others6–10), diagnosis and therapeutics (e.g., ultrasound imaging,11,12 integrated surgical or mucosa-interfacing platforms,13–17 and others18–20), sports protection,21–24 and entertainment.25–29 Recent advances in soft active materials and flexible actuators30–33 have paved the way for flexible bioelectronic devices capable of active shape morphing, substantially extending the functionality of conventional devices with fixed geometric configurations. Typically, shape-morphing bioelectronic devices33 are mainly composed of shape-morphing actuation components and electronic components, where the actuation components leverage active materials or morphable structures to achieve on-demand deformation upon external stimuli (e.g., light, heat, chemistry, electricity, and magnetic fields). Here, the shape-morphing bioelectronic devices are referred to as those devices with active shape-morphing capabilities, noting that the passively morphable devices (i.e., those that can undergo large deformations under external forces) do not belong to this category.
The shape-morphing capabilities of bioelectronic devices have enabled many compelling applications, such as realizing in vivo deformations to perform specific tasks (e.g., therapy and surgery),13,34,35 forming conformal contacts with target objects (e.g., organs, tissues and cells) to allow physiological signal monitoring,36–38 changing configurations to adjust functional properties (e.g., electromagnetic and optical properties),39–41 or assisting users in the daily activities through physical interactions (e.g., rehabilitation and haptic interface).42–45 Despite the remarkable progress, numerous challenges and opportunities exist in this burgeoning area, ranging from designing novel materials/advanced structures, improving shape-morphing capabilities (e.g., reversibility, actuation accuracy, and speed), and developing new integration/fabrication methods, to extending potential application scenarios. The existing reviews that have addressed similar topics primarily concentrated on the fabrication of smart materials and soft actuators,30–32,46 the evaluation of shape-morphing capabilities,33 or specific biomedical applications of soft robots.47,48 Reviews that address the design concepts, fabrication methods, and applications of shape-morphing bioelectronic devices are rare.
Here, this review focuses on the key advances in shape-morphing bioelectronic devices, aiming to provide an overview of the state-of-art status in this exciting area and insights for researchers to drive the advancement of shape-morphing bioelectronic devices. As an essential component of these devices, the actuation component can be constructed using active materials (e.g., dielectric elastomer (DE), shape memory alloys/polymers (SMAs/SMPs), liquid crystal elastomers (LCEs), magnetic elastomers and hydrogels) and/or morphable structures (e.g., bilayer structures, bistable/multistable structures, and pneumatic actuators). According to the different actuation mechanisms, the devices can achieve on-demand shape morphing under various types of stimuli (the inner circle of Fig. 1). Based on the practical needs of different application scenarios, a variety of deformation modes have been realized, such as volumetric expansion and contraction, linear/areal extension and contraction, bending deformation, spiral deformation, complex deformation (e.g., coupled tension, bending and twisting), and transformation between different stable modes (the middle circle of Fig. 1). Leveraging such shape morphing capabilities, these bioelectronic devices are on the verge of revolutionizing multiple domains, such as electromagnetic devices, optoelectronic devices, biological devices for cells and organoids, biomedical devices, and haptic interfaces (the outer circle of Fig. 1). Following the framework of Fig. 1, the present review is outlined as follows. It begins with a summary of the working principles of active materials and morphable structures, representative deformation modes, and manufacturing methodologies of the devices (Section 2). Then, a broad range of practical applications of shape-morphing bioelectronics are overviewed in Sections 3–7. Finally, conclusions and perspectives are provided in Section 8, with a list of promising opportunities in this area, including the development of novel materials, the establishment of inverse design methods and reprogramming strategies for shape morphing, the improvement of actuation accuracy and speed, and the development of biocompatible and comfortable actuation mechanisms.
 |
| Fig. 1 Overview of shape-morphing bioelectronic devices. Inner circle: Representative types of external stimuli that can enable the shape morphing. Middle circle: Typical deformation modes, including volumetric expansion/contraction, extension/contraction, bending deformation, spiral deformation, complex deformation involving coupled tension, bending and torsion, as well as transformation between different stable modes. Outer circle: Applications of shape-morphing bioelectronic devices, spanning electromagnetic devices, optoelectronic devices, biological devices for cells and organoids, biomedical devices and haptic interfaces. | |
2. Overview of the shape-morphing mechanisms
The active shape morphing of bioelectronic devices primarily relies on the utilization of active materials and morphable structures that respond to external stimuli (e.g., light, heat, chemistry, electricity, and external forces). Based on the synergistic material–structure designs, various deformation modes have been realized, enhancing the adaptability across diverse applications. This section provides an overview of active materials and morphable structures used in shape-morphing bioelectronics, their representative deformation modes, and manufacturing methodologies.
2.1 Working principles of active materials and morphable structures
Fig. 2 illustrates the fundamental working principles of representative active materials and morphable structures commonly employed in shape-morphing devices. The dielectric elastomer actuator49 usually adopts a sandwich-like configuration, consisting of a DE membrane (elastomer), where both surfaces are coated with compliant electrodes, as shown in Fig. 2a. When a voltage is applied across the compliant electrodes, the resulting Maxwell stress rapidly induces an expansion in surface area and a contraction in thickness. The multilayer design45,50,51 is widely exploited to achieve an increased actuation force while maintaining the actuation strain. Due to the fast dynamic response, the DE actuators are widely used in haptic devices to provide real-time tactile feedback.45 Additionally, by replacing the elastomer layer with dielectric liquids (e.g., silicone oil), hydraulically amplified electrostatic actuators have been developed, which could offer enlarged levels of deformations.52
 |
| Fig. 2 Working principles of typical active materials and morphable structures in shape-morphing bioelectronic devices. (a) Dielectric elastomer actuated under an electric field. (b) Electrothermal actuation of a SMA spring. (c) Chemically responsive hydrogel capable of significant volumetric expansion/contraction. (d) Typical thermomechanical cycle of a SMP ribbon. (e) Reversible thermal actuation of an LCE ribbon. (f) Magnetic actuation of an elastomer embedded with predefined ferromagnetic domains. (g) Thermal actuation that leverages the strain mismatch in a bilayer structure with different CTEs. (h) Fluidic actuation driven by changes in fluid or air pressure. (i) Mechanically driven transformation of the stable buckling mode in a 3D mesostructure controlled by the loading path. Adapted with permission from ref. 53. Copyright 2018 Springer Nature. | |
Among various electrically controlled actuation mechanisms, the actuation based on SMAs is particularly appealing due to its ability to be electrically activated through Joule (ohmic) heating, achieved by direct electrical current conduction through the material.54 SMAs are commonly fabricated in wire, spring, and membrane configurations, which could be implemented using techniques of laser/FIB cutting and photolithography. The feature size of SMAs used for actuation can vary from a few micrometers to several meters, depending on specific application scenarios. Driven by the phase transition between martensite and austenite, SMA-based actuators exhibit a rich diversity of deformations and a high degree of actuation forces (Fig. 2b).
Hydrogels, consisting of a crosslinked polymeric network infiltrated with water (e.g., hydrogen bonding, covalent/non-covalent bonding, among others), can offer high actuation strains (e.g., >10 times in volume) during hydration and dehydration,55 responding to various external stimuli (e.g., temperature, light, and pH variations) (Fig. 2c). Various processing approaches (e.g., cutting, molding, lithography and printing) have been employed to fabricate hydrogel-based devices with complex geometries and a range of different feature sizes (e.g., tens of micrometers to tens of centimeters),56 showing great potential in applications like drug delivery,57 tissue engineering,58 and health monitoring.38,59
Due to their shape memory effects, shape-locking capability, and highly tunable mechanical properties, SMPs have been extensively employed to enhance the morphing capabilities of bioelectronic devices.60 Conventional SMPs typically consist of polymer chains with different states of motion across the transition temperature, and the softening and hardening of the transition phase in the polymer allows the temporary shape to be locked and unlocked. They can be patterned into different shapes using advanced manufacturing techniques such as 3D printing and laser cutting. SMP-based actuators could be triggered by various external stimuli, including temperature, light, electrical current, and magnetic field. As shown in Fig. 2d, a sequential process of folding deformation and shape recovery is demonstrated during the thermo-mechanical loading. Due to the excellent shape recovery effects, SMPs are widely used in implantable bioelectronic devices.36,61
Liquid crystal elastomers (LCEs), composed of liquid crystal molecules (mesogens) chemically bonded to long-chain organic backbones, represent an emerging class of soft active materials.62 Through various alignment approaches (e.g., surface effect-induced alignment, electric/magnetic field-induced alignment and stress-induced alignment), LCEs could enable programmable actuation via nematic–isotropic phase transition (Fig. 2e). Various fabrication approaches (e.g., 3D printing, cutting, molding, and soft lithography) have been employed to fabricate LCEs with complex geometries and different length scales (from microscale to centimeter scale), holding promise for applications in microrobots63 and biomedical devices.64
Magnetic soft materials are typically composite structures consisting of magnetic particles of varying sizes (ranging from nanometers to millimeters) incorporated into deformable matrices composed of soft elastomers or flexible thin films.65 Under an external magnetic field, these soft materials undergo programmable elastic deformations due to magnetic forces or torques (Fig. 2f). With feature sizes spanning from hundreds of micrometers to several centimeters, the actuators based on magnetic soft materials have enabled many biomedical applications, owing to their deep tissue penetration capabilities and highly controllable magnetic actuation.66–71
In addition to the stimuli-responsive materials, morphable structures have also been extensively explored. Thermally induced strain mismatch, caused by differences in the coefficients of thermal expansion (CTEs) between distinct materials (e.g., graphene, metals, and polymers),72 could enable programmable bending deformations in the bilayer/multilayer architecture (Fig. 2g). By utilizing the microfabrication techniques (e.g., laser cutting and photolithography), bilayer/multilayer structures with feature size ranging from nanometers to a few centimeters can be created.40,73–75 Aside from thermally induced deformations, strain mismatch-driven shape morphing can also be triggered by other external stimuli, such as humidity.76–78
As shown in Fig. 2h, the fluidic actuator, commonly composed of silicone elastomers or plastics, leverages the change of fluid or air pressure inside a closed cavity to realize volumetric expansion or contraction. By rationally engineering the internal channel designs, these actuators can also realize bending and torsional deformations.79 3D printing and molding techniques are usually employed to fabricate these actuators, making them more suitable for large-scale shape-morphing devices (e.g., with centimeter-scale feature sizes). As illustrated in Fig. 2i, buckling-guided assembly methods53,80–88 can harness the bistable/multistable structural designs and/or active materials to realize controlled transformations of 3D mesostructures. The typical process includes (i) design and preparation of 2D precursors with pre-defined geometries, (ii) selective bonding (e.g., typically covalent bonding) of 2D precursors onto a prestrained elastomer substrate with controlled loading magnitudes, and (iii) release of the prestrain of the substrate to complete the assembly. To leverage the bistability/multistability for shape transformation, specially engineered 2D precursors should be designed such that different buckling modes can be stabilized following the distinct release path of the biaxial prestrain (shape I and shape II in Fig. 2i).53 These methods allow access to morphable structures with a rich diversity of 3D geometries with length scales from micrometers to tens of centimeters. Due to the compatibility with mainstream electronic materials (e.g., metals, semiconductors, polymers, 2D materials and oxides), a broad range of applications have been demonstrated, such as electromagnetic,53,89 optoelectronic,90 biological,91 biomedical,13 and haptic devices.92 Additionally, by using soft active materials as the assembly substrates, such as hydrogel,93 LCE,94 SMP95 and DE96 substrates, the assembled 3D mesostructures can be reshaped upon various external stimuli.
2.2 Representative deformation modes
To meet the practical requirements of shape-morphing bioelectronic devices, a diversity of deformation modes have been realized, as summarized in this section.
Fig. 3a illustrates the volume expansion of an endocardial balloon catheter actuated by fluidic actuation (i.e., pneumatic actuation).13 Such actuation deformation could establish a conformal contact of the device with curved tissue surfaces to enable multiplexed sensing and cardiac surgery. Similar expansion deformations were also widely exploited in hydrogel-based devices, such as deformable patches for healing diabetic wounds,58 and artificial bladder detrusors,97 among others.55
 |
| Fig. 3 Representative deformation modes of shape-morphing bioelectronic devices. (a) Volumetric expansion of a device into the predefined shape driven by increased internal pressure. (b) Linear/areal extension and contraction of devices. (c) Bending deformation of a cross-shaped device to wrap around the apex of the heart. (d) Spiral deformation of a ribbon-shaped device for tissue/nerve monitoring. (e) Complex out-of-plane deformation of a network-shaped device driven by electromagnetic forces. (f) Controlled transformation of a 3D device between different stable modes via buckling-guided assembly. | |
The top panel of Fig. 3b demonstrates the linear extension and contraction enabled by a SMA spring-based device, which is employed for treating muscle atrophy.98 In addition to linear deformation, areal expansion and contraction can be achieved through rational microstructure design99 or the use of DEs.49 For instance, flexible network designs composed of LCEs64 or SMAs100 allow relatively high areal contraction upon heating, contributing to wound healing applications (bottom panel of Fig. 3b).
In scenarios where the devices need to form conformal contacts with target objects to realize functionalities, the bending or spiral deformation is usually required. As shown in Fig. 3c, the hydrogel-based device can wrap and subsequently monitor the heart, owing to the programmable bending deformation.38 Since the bending deformation can be easily induced through principles of strain mismatch and shape memory, this deformation mode has been extensively exploited in biomedical/biological devices,101–104 optoelectronic devices,105 and electromagnetic systems.40 Notably, the localized folding deformation can be considered a special case of bending deformation.106–108 When the target objects (e.g., nerves) have slender configurations, the spiral deformation (with coupled bending and twisting) could be employed to enhance conformal contact. Fig. 3d shows a SMP-based twisting electrode designed for nerve stimulation and recording, which can coil around the nerve.36
In certain scenarios (e.g., haptic interfaces), the fabricated devices might need to transition between complex target shapes (e.g., hyperboloids and saddle surfaces). During these transformations, the device may undergo complex shape changes involving tensile, bending, and twisting deformations. A mechanical metasurface, consisting of a matrix of filamentary metal traces, demonstrates complex, dynamic morphing capabilities under reprogrammable Lorentz forces (Fig. 3e).109 Besides, an electromagnetically actuated 3D display screen, capable of both shape morphing and on-demand shape locking, was developed by leveraging the liquid–solid phase transition of liquid metal microfluidic networks.110
Microelectronic devices that can be reversibly switched between two or more geometrically stable 3D configurations hold promise for applications in electromagnetic devices, optoelectronics and haptic interfaces.53,89,92,111 By leveraging smart materials (e.g., LCEs and SMPs) or rational structural designs, the resulting multi-stable devices could maintain their 3D configurations without continuous external energy input and enable rapid shape transitions.112 Taking advantage of sequential and directional control of loading paths in the buckling-guided assembly, a bottom-up design route to geometrically reconfigurable 3D mesostructures was established using ribbon-shaped components as building blocks (Fig. 3f).89
2.3 Manufacturing methodologies
To manufacture the actuation components of shape-morphing bioelectronic devices, various methods could be exploited, such as 3D printing, molding, cutting and photolithography. Two distinct routes have been explored to develop flexible and stretchable electronics, including the route that concentrates on improving the electrical performances of organic electronic materials, and the other one that focuses on creating novel architectures of high-performance inorganic electronic materials to achieve flexibility and stretchability.113 For the fabrication of organic electronic components, techniques such as 3D printing, molding and even lithography can be employed.114–116 Meanwhile, lithography and transfer printing were widely adopted in the fabrication of inorganic flexible/stretchable electronics.80,117,118 Considering that most shape-morphing devices incorporate both actuation and electronic components, their manufacturing can be implemented by two different strategies: (i) in situ fabrication and integration, and (ii) independent processing followed by integration.
The first manufacturing strategy means that the fabrication methods (whether identical or different) for actuation and electronic components are compatible within a unified process, without additional operations. The two components can be prepared using the same techniques, such as 3D printing119,120 and lithography,103,108,121 among others.122 For example, the soft actuation components (e.g., LCEs) and electronic components (e.g., LM circuits) can be patterned simultaneously by direct ink writing (DIW), a 3D multimaterial printing approach.119 In addition, lithography can be utilized to manufacture small-scale biological/biomedical devices. During the manufacture of a micropatterned multielectrode shell for 3D spatiotemporal recording of live cells, the actuation component (i.e., SiO/SiO2 bilayer) and the top electronic component (Cu electrode) are fabricated through a multilayer lithography process.108 Lithographic techniques were also employed to fabricate a reconfigurable antenna comprising Cr and VO2, where the electronic component (patterned Cr) functions both as a passive layer and as a heating electrode for electrothermal deformation, while also serving as a functional component that enables electromagnetic radiation.40 Furthermore, different fabrication techniques for the two components were also utilized to realize in situ fabrication and integration. For instance, in the fabrication of hydrogel-based shell microelectrode arrays (MEAs)122 for brain organoids, the electronic component (PI with patterned Pt) was manufactured using standard lithography. After surface treatment, the top hydrogel layer was molded onto the PI surface, followed by laser cutting to precisely define the device pattern.
In the second manufacturing strategy, the actuation and electronic components are prepared separately and subsequently integrated through operations such as transfer printing and wire bonding. This strategy is typically employed when the fabrication techniques for the two components are not readily compatible within a single unified process. For example, in the manufacture of twisting electrodes36 for peripheral nerve wrapping, the actuation component (i.e., body temperature-activated SMP) was first synthesized and molded, while the electronic component (i.e., Cu/PI pattern) was fabricated using a standard lithography/etching process. To complete the integration, the SMP was heated to enhance interface adhesion, followed by the transfer of the electronic component onto the SMP. To further improve adhesion, bio-adhesive glue was employed. For instance, in the fabrication of hydrogel-based implantable devices, the actuation component (i.e., hydrogel) was prepared via a molding approach, while the electronic component was fabricated via an in situ solution-based method.38 The resulting multi-modal e-skin was then bonded onto the hydrogel using an adhesive glue (e.g., 3M Vetbond 1469C). In certain haptic interface applications, the electronic component (e.g., patterned electrodes) was used to trigger the actuation components to enable controllable deformation (i.e., real-time feedback). For example, during the development of a LCE array-based touch screen, the actuation component (LCE fiber array) was fabricated through molding directly onto a pre-fabricated electronic component (PCB circuit), facilitating the development of an interactive human–machine interface.123
3. Electromagnetic devices
Electromagnetic devices, such as antennas and metasurfaces, hold significant potential for applications in wearable electronics, enabling functions like highly efficient wireless communication for seamless connectivity across various networks, as well as electromagnetic sensing for detecting physiological signals or environmental changes.80,124,125 Shape-morphing electromagnetic devices offer the ability to dynamically adjust their electromagnetic properties in response to varying operational requirements. This section provides an overview of representative electromagnetic devices, including radio frequency (RF) devices39,40,53,81,89,96,125–131 (e.g., antennas, inductors, filters) and others109,110,132–136 (e.g., terahertz metamaterials).
3.1 RF devices
Antennas can be reshaped to achieve tunable electromagnetic properties, such as resonant frequency and radiation patterns. For example, a bottom-up design strategy based on elementary reconfigurable structures of simple ribbon geometries was developed, utilizing the buckling-guided assembly.89 This approach enabled the realization of multimodal antennas with dynamically reconfigurable radiation patterns, as shown in Fig. 4a. Meanwhile, the liquid–gas phase transition facilitates the transformation of an antenna from planar to 3D configuration through fluidic actuation, thereby demonstrating a tunable radiation pattern.128 In the design and application of antennas, an effective strategy to adjust the resonant frequency involves the modification of the electrical length.126,130 For instance, a reconfigurable antenna leveraging thermal mismatch-induced deformation can transition from a planar state to a curved state, demonstrating a tunable operating frequency in the range of 46 GHz to 50 GHz (Fig. 4b).40 Additionally, by utilizing magnetic shape memory polymers with reversible shape morphing and locking capabilities, the resonant frequency of a helical antenna can be readily tuned between 2.15 and 3.26 GHz, as shown in Fig. 4c.127
 |
| Fig. 4 Shape-morphing electromagnetic devices. (a) Morphable 3D antenna with reconfigurable radiation patterns. Left: Optical images of the antenna in different stable operating modes. Right: Corresponding H-plane radiation patterns. Adapted with permission from ref. 89. Copyright 2020 The American Association for the Advancement of Science. (b) Reconfigurable antenna whose resonant frequency can be adjusted by bending deformation. Reproduced with permission from ref. 40. Copyright 2021 Springer Nature. (c) Magnetically actuated antenna capable of continuously tuning working frequency from 2.15 to 3.26 GHz. Reproduced with permission from ref. 127. Copyright 2020 Wiley. (d) Reconfigurable LCR circuit composed of four morphable 3D capacitors that can be individually addressed. The right panel shows the return loss (S11) curves under four different working conditions. Reproduced with permission from ref. 96. Copyright 2019 Oxford University Press on behalf of China Science Publishing & Media Ltd. (e) Band-stop behavior in the deployed state and all-pass behavior in the folded state of a morphable electromagnetic filter. Reproduced with permission from ref. 39. Copyright 2022 Wiley. (f) Morphing process of a reprogrammable surface, illustrating the growth and pinch-off of a droplet dripping from a nozzle. Reproduced with permission from ref. 109. Copyright 2022 Springer Nature. | |
Based on morphable 3D capacitors, reconfigurable inductor-capacitor (LC) RF circuits96,125 can be developed for applications such as sensing, filtering, and frequency selection. A reconfigurable inductor-capacitor-resistor (LCR) circuit was developed, consisting of four morphable 3D capacitors that can be addressed through the different regions in DE substrates (‘A’, ‘B’, ‘C’ and ‘D’). This LCR circuit offers a tunable resonant frequency in the range of 10.7 MHz to 21.3 MHz, as shown in Fig. 4d.96
Engineered electromagnetic metamaterials possess unique capabilities of wave manipulation, filtering, and cloaking. Through structural reconfiguration enabled by magnetic actuation, the developed electromagnetic metamaterial can dynamically switch between wave acceptance (folded state) and wave rejection (deployed state), as demonstrated in Fig. 4e.39
3.2 Other electromagnetic devices
Reconfigurable terahertz metamaterials with tunable electric and magnetic responses have been developed, where the arrays of split-ring resonators on bilayer cantilevers could actively bend upon a thermal stimulus.134Fig. 4f shows a dynamically reprogrammable device driven by reprogrammable and distributed Lorentz forces, which can morph into a wide range of 3D target shapes with high precision (e.g., the growth and pinch-off of a droplet).109 Based on this Lorentz force-driven strategy, many morphable devices have been developed, such as morphable 3D display screens110 and grippers.132,136 Furthermore, by integrating SMP films, a soft electromagnetic gripper with versatile deformation and locking capabilities was fabricated.135 Moreover, the variable-stiffness property of the SMP endows this gripper with significant potential for handling delicate soft objects, such as microscale organs.
Although shape-morphing electromagnetic devices have demonstrated tunable electromagnetic properties such as frequency and radiation pattern, many current designs suffer from relatively large structural size and limited dynamic reconfigurability. Additionally, reversible shape morphing can potentially introduce material/structure fatigue, leading to performance degradation. Further improvements could focus on the integration of fatigue-resistant composites and the development of multistable architectures guided by inverse design strategies.
4. Optoelectronic devices
Shape-morphing optoelectronic devices, including 3D displays, camouflage devices, and biomimetic eyeball cameras, hold great potential for integration into wearable device systems, where their shape-morphing capability enables tunable optical properties such as reconfigurable display patterns,105,137–141 controlled light interactions,41,142–145 and adjustable focus,146,147 thereby enhancing advanced visualization, adaptive concealment, and biomimetic imaging.
4.1 3D displays
Based on the buckling-guided assembly approach, many morphable 3D displays were developed by releasing the pre-stretch strain of the substrate as required.138,139 In addition, morphable 3D displays were also developed by the use of stimuli-responsive substrates.105,137 For example, morphable displays were fabricated by integrating electrically actuated substrates with LED arrays, exhibiting various light patterns during complex deformations.105 Additionally, by incorporating low melting point alloy (LMPA)–graphene nanoplatelets (GNPs)–elastomer composites, 3D morphable displays were fabricated, featuring relatively fast electrothermal actuation capabilities and various 3D light patterns (Fig. 5a).137 By employing complex shaped actuators capable of continuous shape morphing and locking as morphable exoskeletons, the developed 3D displays can be reshaped into a large number of 3D geometric configurations.148
 |
| Fig. 5 Shape-morphing optoelectronic devices. (a) A butterfly-shaped display capable of flapping wings, and a flower-shaped display that can mimic the blooming process. Reproduced with permission from ref. 137. Copyright 2023 Wiley. (b) Chameleon-inspired device capable of dynamically tuning the structural color through the reversible actuation of the LCE component. Reproduced with permission from ref. 142. Copyright 2023 Royal Society of Chemistry. (c) Electromechanically reconfigurable optical nano-kirigami device capable of modulating the optical helicity. Reproduced with permission from ref. 41. Copyright 2021 Springer Nature. (d) Fluidically actuated tunable electronic eye camera with adjustable zoom capability. Reproduced with permission from ref. 147. Copyright 2011 National Academy of Sciences. (e) Adaptive camera system capable of simultaneously adjusting the focal length of the lens and the curvature of the imager via magnetic actuation. Reproduced with permission from ref. 146. Copyright 2021 Springer Nature. | |
4.2 Devices for light manipulation
Structural color-based tunable devices have potential applications in artificial camouflage, optical filters and mechanical sensors. Inspired by the active tunable color system of chameleons, a structural color-based device comprising a nanoscale hole array of photonic crystals, carbon nanotube coatings (heating and sensor), and LCEs was developed (Fig. 5b).142 This device could be activated under a low voltage (e.g., several volts) to undergo linear contraction, enabling a reversible color transition from blue to red. Based on the nano-kirigami method,145 3D optical devices with giant optical chirality can be fabricated. As shown in Fig. 5c, nanophotonic electro-mechanical devices driven by electrostatic forces were demonstrated, exhibiting tunable optical chirality in the near-infrared range with a modulation speed exceeding 10 MHz.41
4.3 Eyeball cameras
Cameras in the form of biomimetic 3D eyeballs can significantly improve imaging performances by offering a broadened field of view, reduced distortions, and decreased chromatic aberrations.146,147,149–151 To improve the accommodation of cameras to the changes in the Petzval surface upon the use of different lenses, tunable hemispherical electronic eye camera systems with adjustable zooming capabilities were developed. For example, the curvature radius of the camera can be precisely controlled from 4.9 mm to 11.5 mm through fluidic actuation, enabling a camera system with continuously adjustable zoom (Fig. 5d).147 A shape-morphing camera that combines a concave imager printed on a magnetic rubber composite with a tunable lens is demonstrated in Fig. 5e, offering focused views of objects at varying distances.146
Reversibility, fast actuation, and precise shape reconfigurability are essential for shape-morphing optoelectronic devices, particularly in applications requiring dynamic light manipulation, tunable optical responses, or adaptive functionalities. Future efforts could focus on the development of active materials and morphable structures that exhibit highly reversible deformations, fast response to low-power stimuli, and programming capabilities. Additionally, incorporating real-time feedback into the device system could further enhance the spatial and temporal deformation precision of shape reconfiguration.
5. Biological devices for cells and organoids
By leveraging the shape morphing process from planar to 3D states, the bioelectronic devices can effectively capture and establish conformal contacts with the target objects (e.g., cells and organoids) at micro- to nano-scale, featuring high signal-to-noise ratio and 3D spatiotemporal recording.91,103,104,108,122,152–155
5.1 Cell devices
Shape-morphing bioelectronic devices serve as powerful tools to capture and read dynamic electrophysiological signals of cellular processing and intercellular communications. As shown in Fig. 6a, by leveraging engineered strain mismatches between SiO2 and SiO layers, the deformation angles of the device can be precisely tuned to establish conformal contact with single cells.152 Based on the gripper-like form, the fabricated device, integrated with a multielectrode array, can conformably wrap around a cardiac cell (tens of micrometers in diameter) to realize 3D spatiotemporal recordings (Fig. 6b).108 Establishing conformal contact between devices and samples with soft and irregular surfaces (e.g., cancer cells) is extremely challenging. Fig. 6c shows an ultrathin biosensing device (graphene composites) capable of mechanical trapping and surface-enhanced Raman spectroscopy (SERS), which has enabled simultaneous capture, profiling, and 3D microscopic mapping of the intrinsic molecular signatures of a single living cell (e.g., cancer cell).154
 |
| Fig. 6 Shape-morphing biological devices for fundamental studies of cells and organoids. (a) Morphable grippers capable of capturing single cells. Reproduced with permission from ref. 152. Copyright 2014 American Chemical Society. (b) Curved shell device with spatially distributed electrodes, designed to wrap around the cardiomyocyte cell for electrophysiological monitoring. Reproduced with permission from ref. 108. Copyright 2018 Wiley. (c) Ultrathin, flexible biosensing platform that can wrap around live breast cancer cells, allowing for 3D molecular spectroscopy via SERS. Reproduced with permission from ref. 154. Copyright 2019 American Chemical Society. (d) Self-rolling cylindrical device for electrical interrogations of human electrogenic spheroids. Reproduced with permission from ref. 104. Copyright 2019 The American Association for the Advancement of Science. (e) SU8 bilayer-based 3D device for spatiotemporal recording of brain organoids. Reproduced with permission from ref. 103. Copyright 2022 The American Association for the Advancement of Science. (f) Hydrogel-based 3D device for brain spheroid electrophysiology. Reproduced with permission from ref. 122. Copyright 2024 The American Association for the Advancement of Science. | |
5.2 Organoid devices
Zooming out from a single-cell perspective, shape-morphing bioelectronics has also significantly enriched the methodologies for organoid studies. A self-rolling biosensor device, built with a metal/polymer bilayer structure, was developed to conformally wrap human cardiac spheroids to provide continuous and stable multiplexed recordings of field potentials with high sensitivity and spatiotemporal resolution (Fig. 6d).104 In addition, the buckling-guided method has been leveraged to construct 3D neural interfaces for cortical spheroids.91 By rationally designing the 2D precursor patterns, the 3D configurations of neural interfaces can be reshaped to accommodate cortical spheroids of varying numbers or shapes. Furthermore, shape-morphing shell microelectrode arrays were developed for studying brain organoids, featuring high signal-to-noise ratio sensing and 3D spatiotemporal recording (Fig. 6e).103 The aforementioned devices primarily utilize external actuation systems or potentially harmful solvents to trigger the shape-morphing process. To address this, Fig. 6f shows a 3D device (hydrogel/PI/Pt trilayer) for brain spheroid electrophysiology, which is driven by the swelling of soft hydrogels and triggered by the addition of the cell culture medium.122
Existing shape-morphing biological devices often face challenges in establishing stable and conformal contact with target 3D cells/organoids. The trigger of the shape morphing for some devices also requires external stimuli (e.g., chemical) that may affect cell/organoid viability. Future studies could focus on integrating inverse design strategies with advanced fabrication techniques to construct 3D interfaces that enable large-area, high-fidelity contact for comprehensive signal acquisition. Additionally, the development of biocompatible actuation materials (e.g., ionic hydrogels or enzymatically triggered SMPs) that respond to gentle physiological stimuli is essential to ensure safe biological applications.
6. Biomedical devices
This section provides an overview of representative shape-morphing biomedical devices for a range of different application scenarios, such as rehabilitation,42,43,66,97,156–158 tissue engineering,58,64,98,100,159–163 surgery,13,35,67,101,102,164–169 drug delivery,34,57,170,171 and health monitoring.36–38,59,61,172–181
6.1 Rehabilitation
Wearable robotic systems, such as flexible hand rehabilitation gloves43 and exoskeletons,182 are emerging rapidly toward the directions of lightweight, safety and comfortability. For example, as illustrated in Fig. 7a, a soft-packaged and portable rehabilitation glove has been developed, enabling a range of fine motor skills (FMSs) with real-time sensory feedback and a closed-loop control system.43 This shape-morphing device supports individuals with hand impairments in performing rehabilitation exercises across various environments (e.g., home, office, and park). Further integration of smart textiles capable of real-time pH and temperature monitoring183 could enhance the functionality of the developed device. In addition to the above rehabilitation devices, in vivo rehabilitation/organ-assisting devices have also been explored.66,156–158 For instance, a pneumatically actuated robotic sleeve was developed, which could act as a cardiac ventricular assisting device.158 A hydrogel-based soft artificial bladder detrusor was demonstrated to be able to shrink an animal bladder upon electrothermal stimulus (Fig. 7b).97 Additionally, a soft robotic ventilator was engineered to restore respiratory function in a pig model of respiratory insufficiency.156
 |
| Fig. 7 Shape-morphing biomedical devices for rehabilitation, tissue engineering, and surgery. (a) Soft rehabilitation glove that assists patients in performing complex FMS rehabilitation exercises and carrying out activities of daily living. Reproduced with permission from ref. 43. Copyright 2023 Springer Nature. (b) Electrothermally responsive soft artificial bladder detrusor capable of shrinking an animal bladder. Reproduced with permission from ref. 97. Copyright 2018 Wiley. (c) Skin temperature-actuated electromechanical wound dressings for accelerated healing of linear and circular wounds. Reproduced with permission from ref. 100. Copyright 2022 The American Association for the Advancement of Science. (d) Thermally actuated device that can facilitate muscle contraction to prevent atrophy. Reproduced with permission from ref. 98. Copyright 2022 Springer Nature. (e) Autonomous biohybrid fish designed with human cardiac biophysics. Reproduced with permission from ref. 163. Copyright 2022 The American Association for the Advancement of Science. (f) Thermally responsive untethered microgripper designed for biopsy. Reproduced with permission from ref. 102. Copyright 2013 Wiley. (g) Origami-inspired miniature device for teleoperated microsurgery. Reproduced with permission from ref. 167. Copyright 2020 Springer Nature. (h) Balloon catheter integrated with an array of multilayered electronic devices for sensing and therapy during cardiac surgery. Reproduced with permission from ref. 13. Copyright 2020 Springer Nature. (i) Millimeter-scale therapeutic device that can deploy and self-stabilize at the entrance to the heart, guiding existing interventional tools toward a target site. Reproduced with permission from ref. 35. Copyright 2023 The American Association for the Advancement of Science. (j) LCE fiber-based device featuring high motion precision for minimally invasive surgery. Reproduced with permission from ref. 169. Copyright 2024 The American Association for the Advancement of Science. | |
6.2 Tissue engineering
To accelerate skin regeneration while avoiding scars and keloids, many shape-morphing patches have been developed.58,64,100,160 For example, Fig. 7c shows a skin temperature-activated electromechanical synergistic wound dressing comprising a SMA-based mechanical metamaterial for wound contraction and an antibacterial electret thin film for electrical stimulation.100 For the treatment of muscle atrophy, a SMA-based device was reported to generate and deliver muscle-contraction-mimicking stimulation to a target tissue with programmed strength and frequency (Fig. 7d).98 Furthermore, bio-hybrid robots based on biological muscle have important implications for tissue engineering and regenerative medicine. Fig. 7e demonstrates an autonomously swimming bio-hybrid fish made of cardiac cells embedded in polymer frames.163 This study shows great potential for a granular analysis of structure–function relationships in cardiovascular physiology.
6.3 Surgery
The shape-morphing capability enables biomedical devices to be highly versatile in surgical applications, including excising tissues for biopsy, performing precise manipulations in teleoperated microsurgery, and establishing conformal contact with target tissues for effective RF ablation. For instance, untethered thermobiochemically actuated microgrippers were developed, which can be remotely controlled by external magnetic fields to assist in vitro biopsies.168 Based on this gripper design, the fabricated device (Fig. 7f), consisting of a SiO2/SiO bilayer and a magnetic layer (iron), can be remotely guided and actuated on-demand to extract tissue samples from real organs (i.e., a porcine liver) and hard-to-reach places within a live animal (i.e., a porcine biliary tree).102 As demonstrated in Fig. 7g, the origami-inspired miniature morphable device can be used for teleoperated microsurgery.167 For applications in cardiac surgeries, many shape-morphing devices with integrated sensing, heating and/or RF components have been developed. Fig. 7h shows a pneumatically actuated endocardial balloon catheter equipped with soft electronic arrays, which can establish conformal contact with curved tissue surfaces to realize high-density spatiotemporal mapping of various parameters, programmable electrical stimulation, radiofrequency ablation, and irreversible electroporation.13 A millimeter-scale device was also developed to deploy and stabilize at the entrance to the heart and guide existing interventional tools toward a target site, as shown in Fig. 7i.35 The above cardiac surgical devices also have the potential to treat Timothy syndrome.184 Precise manipulation of flexible surgical tools is crucial in minimally invasive surgical procedures. An LCE fiber-based device with a low profile (below 2 mm in diameter) was shown to exhibit remarkable motion precision (below 50 μm) and repeatability (Fig. 7j).169
6.4 Drug delivery
The controllable shape-morphing capability endows fabricated devices with significant potential for drug delivery. For instance, magnetically actuated, pH-responsive hydrogel-based devices (in the form of grippers) were developed, featuring controllable locomotion under magnetic fields and on-demand drug delivery in response to pH variations.57 Inspired by this design, untethered thermally actuated microdevices with sharp microtips were fabricated, capable of stably adhering to the mucosal tissue for up to 24 hours before initiating drug release (Fig. 8a).34 To access the diverse 3D surfaces of disease sites within the human body and perform minimally invasive therapies (e.g., drug delivery), an untethered magnetically controlled soft device was introduced, featuring strong adhesion to wet 3D tissue surfaces and on-demand drug delivery in response to pH changes.185
 |
| Fig. 8 Shape-morphing biomedical devices for drug delivery and health monitoring. (a) Submillimeter-scale device for enhanced drug release and retention. The right two panels show prolonged and higher exposures of ketorolac compared to pristine drug. Reproduced with permission from ref. 34. Copyright 2020 The American Association for the Advancement of Science. (b) Conformal in-ear bioelectronics for visual and auditory brain–computer interfaces. Reproduced with permission from ref. 175. Copyright 2023 Springer Nature. (c) Deployable electrode array for minimally invasive large-scale intracranial brain activity mapping. Reproduced with permission from ref. 176. Copyright 2024 Springer Nature. (d) Multifunctional biocompatible devices capable of diagnosis, stimulation, and drug delivery. Reproduced with permission from ref. 38. Copyright 2024 Springer Nature. (e) Water-responsive bioelectronic interface for monitoring a rat heart. Reproduced with permission from ref. 178. Copyright 2023 Springer Nature. (f) Electrochemically actuated nerve cuff for minimally invasive intraoperative monitoring of nerve activity. Reproduced with permission from ref. 37. Copyright 2024 Springer Nature. (g) Hydraulically actuated deployable device for atrial electrophysiological mapping. Reproduced with permission from ref. 179. Copyright 2020 The American Association for the Advancement of Science. | |
6.5 Health monitoring
Based on various actuation mechanisms, the shape-morphing bioelectronic devices can form conformal contacts with captured objects (e.g., organs and tissues) for stable monitoring.36–38,59,61,172–181 For example, leveraging the body temperature-induced shape recovery effect of the SMP, the fabricated device can twin around nerves to realize stimulation and recording.36 In addition, the SMP-based devices can also be employed to develop in-ear visual and auditory brain–computer interfaces, which can adaptively expand and spiral along the auditory meatus under electrothermal actuation to ensure conformal contact (Fig. 8b).175Fig. 8c shows a deployable microelectrode array (MEA) based on an electrothermally actuated SMA (Nitinol alloy), which is capable of large-scale intracranial brain activity mapping.176 Based on the electrothermally actuated hydrogels, a class of shape-morphing devices have been developed, including a robotic cuff for detecting blood pressure, a robotic gripper for tracking bladder volume, an ingestible robot for pH sensing and on-site drug delivery, and a robotic patch for quantifying cardiac function and delivering electrotherapy, as shown in Fig. 8d.38 In addition to the thermal stimulus, the humidity stimulus could also be employed in shape-morphing devices. For example, using water-responsive supercontractile polymers (with more than 50% linear contraction), shape-morphing electrodes have been developed to conformally wrap around nerves, muscles, and hearts of different sizes upon wetting (Fig. 8e).178 The above devices, however, cannot reverse or reprogram their configuration in the body environment, and therefore, are suitable only for single-time use. Electrochemically actuated polymer-based nerve cuffs can actively grip or wrap around delicate nerves with applied voltages as low as a few hundred millivolts (Fig. 8f).37 The fluidic actuation was also used for constructing shape-morphing devices, such as the hydraulically actuated deployable device with an array of stretchable electronic sensors for voltage mapping, which offered a high conformability (∼85 to 90%) with the entire left atrium (Fig. 8g).179
Despite the remarkable progress in the area of biomedical shape-morphing devices, many of the existing devices are limited by single-time use or lack of wireless control. Invasive deployment and insufficient miniaturization also pose limitations on clinical applications. To address these challenges, further efforts could be made to explore wireless magnetic actuation systems and active materials responsive to natural stimuli (e.g., body temperature) for in situ activation.
7. Haptic interfaces
Haptic interfaces can substantially enrich experiences of users in virtual reality (VR) and augmented reality (AR) systems. The shape-morphing capability is particularly useful in haptic interfaces, because it can provide reconfigurable surface textures and diverse vibration sensations.186 This section focuses primarily on two classes of haptic interfaces that can be strengthened by shape morphing capabilities, including touch screens92,123,186–195 and wearable haptic interfaces.25–27,44,45,196,197
7.1 Touch screens
A transparent touch screen based on DE was developed, featuring four distinct surface textures whose touch feeling varies from relatively smooth to rough.195 To increase the complexity of surface textures, haptic arrays based on the touchable-braille-dot design, actuated by various mechanisms, have been reported.187,188,190,194 In addition, Fig. 9a presents a refreshable touch screen capable of forming 3D reconfigurable shapes with desirable load-bearing capabilities, demonstrating significant potential for applications in visually impaired education and VR/AR systems.92 The hydraulically amplified electrostatic actuators can be employed to construct ultra-soft, dynamically reconfigurable touch screens.189,191,193 For example, Fig. 9b demonstrates a touch screen that can offer a rich diversity of haptic feedback, including dynamic tactile patterns and vibrations for localizable surface textures on the morphed shape.193 Furthermore, by integrating sensing and control circuits, the touch screen based on hydraulically amplified electrostatic actuators can realize versatile functionalities, including user interaction, image display, sensing of object mass, and dynamic manipulation of solids and liquids (Fig. 9c).191 Constructed using an array of architected LCE fibers, the touch screen shown in Fig. 9d can offer tactile feedback in multiple directions.123
 |
| Fig. 9 Haptic interfaces. (a) Reconfigurable 3D metasurfaces for assisting visually impaired individuals. Reproduced with permission from ref. 92. Copyright 2024 Springer Nature. (b) Electro-hydraulically actuated tactile display capable of generating dynamic and diverse tactile patterns. Reproduced with permission from ref. 193. Copyright 2024 The American Association for the Advancement of Science. (c) Multifunctional tactile display with high-speed actuation and pressure sensing. Reproduced with permission from ref. 191. Copyright 2023 Springer Nature. (d) Electrothermally controlled tactile display based on an array of vertical LCE fibers. Reproduced with permission from ref. 123. Copyright 2024 The American Association for the Advancement of Science. (e) Wearable haptic device capable of providing normal and shear interaction. Reproduced with permission from ref. 44. Copyright 2020 Wiley. (f) Untethered wearable haptic device that integrates an ultrathin (18 μm in thickness) DE actuator, booster circuit, battery and photodiodes. Reproduced with permission from ref. 196. Copyright 2020 Wiley. (g) DE-based soft wearable haptic device that can provide complex tactile feedback. Reproduced with permission from ref. 45. Copyright 2024 The American Association for the Advancement of Science. (h) Wearable haptic interface integrated with an array of bistable transducers, enabling both dynamic and static stimuli directed as either normal or shear forces. Reproduced with permission from ref. 26. Copyright 2024 Springer Nature. | |
7.2 Wearable haptic interfaces
Fig. 9e describes a wearable device consisting of a 5 × 5 array of hydraulically amplified electrostatic actuators, which could produce both out-of-plane and in-plane motions to generate normal and shear forces.44 An ultra-thin DE-based untethered device is demonstrated in Fig. 9f, featuring a broad vibrotactile feedback spectrum (from 1 Hz to 500 Hz).196 By utilizing DE-based arrays, complex tactile feedbacks with high perception accuracy can be generated, such as patterns of “U”, “C”, “L” and “A” in Fig. 9g.45 Apart from stimuli-responsive soft materials, the integration of rigid functional components (e.g., magnets) with flexible structures can also be employed to construct wearable haptic systems.27 The two types of haptic units in Fig. 9h (left two panels) can deliver both dynamic and static stimuli, directed as either normal or shear forces. By integrating an array of these units, a wireless, skin-conformable haptic interface was developed, serving as a high-density channel capable of rendering input from smartphone-based 3D scanning and inertial sensors.26
While shape-morphing haptic devices have achieved versatile tactile feedback, challenges remain in miniaturization, integration, and multimodal feedback delivery. For instance, some of the existing devices rely on external power sources and lack scalability for skin-level wearables. Future developments could benefit from incorporation of stretchable triboelectric layers for self-powered actuation and modular haptic units with integrated sensing-feedback loops.
8. Conclusions and outlooks
With continuous advancements in materials science, electronics, and mechanical engineering, significant progress has been made in the development of shape-morphing bioelectronic devices. These innovative devices hold the potential to revolutionize various fields, including electromagnetic systems, biomedical engineering, bio-integrated electronics, and haptic interfaces. In particular, key features of representative devices, including their driving mechanisms, key actuation/electronic materials, deformation modes, and reversibility, are summarized in Table 1, which provides a comprehensive view of the diverse family of shape-morphing bioelectronic devices. Despite the remarkable progress summarized herein, rich opportunities exist in this burgeoning area, including the development of novel materials, the establishment of inverse design methods and reprogramming strategies for shape morphing, the improvement of actuation accuracy and speed, and the development of biocompatible and comfortable devices.
Table 1 Summary of representative shape-morphing bioelectronic devices
Applications |
Devices |
Actuation mechanisms |
Key materials |
Deformation modes |
Reversibility |
Ref. |
Electromagnetic devices |
Antenna with reconfigurable radiation patterns |
Buckling-guided assembly |
Silicone; Cu |
Transformation of stable modes |
Yes |
89
|
Antenna with reconfigurable radiation patterns |
Pneumatic actuation |
Silicone; LM |
Volumetric change |
Yes |
128
|
Antenna with tunable resonant frequency |
Strain mismatch |
Cr/VO2; Cr |
Bending deformation |
Yes |
40
|
Antenna with tunable resonant frequency |
Magnetic actuation |
Magnetic SMP; Ag |
Complex deformation |
Yes |
127
|
Reconfigurable electromagnetic filters |
Magnetic actuation |
Magnetic silicone; Cu |
Areal extension and contraction |
Yes |
39
|
Reconfigurable terahertz metamaterials |
Strain mismatch |
Au/SiNx; Au |
Bending deformation |
Yes |
134
|
Reprogrammable metasurface |
Electromagnetic actuation |
PI/Au/PI; Au |
Complex deformation |
Yes |
109
|
|
Optoelectronic devices |
3D display |
LM-based actuation |
LM composite; ZnS:Cu/PDMS |
Complex deformation |
No |
137
|
Tunable structural color system |
LCE-based actuation |
LCE; an array of photonic crystals |
Linear extension and contraction |
Yes |
142
|
Optical nano-kirigami |
Electrostatic actuation |
Au/Si; Au |
Complex deformation |
Yes |
41
|
Tunable eye camera |
Hydraulic actuation |
Silicone; silicon photodiodes and blocking diodes |
Volumetric change |
Yes |
147
|
|
Biological devices |
Single-cell multielectrode shell |
Strain mismatch |
SiO/SiO2; Au |
Bending deformation |
No |
108
|
Ultrathin skin for single cell |
Strain mismatch |
Graphene; Ag NCs |
Complex deformation |
No |
154
|
Tubular microelectrode arrays for cardiac spheroids |
Strain mismatch |
SU8 bilayer; Au |
Bending deformation |
No |
104
|
Gripped-shaped arrays for brain spheroids |
Strain mismatch |
SU8 bilayer; Au |
Bending deformation |
No |
103
|
Flower-shaped microelectrode arrays for brain spheroids |
Hydrogel-based actuation |
Hydrogel/PI; Pt |
Bending deformation |
No |
122
|
|
Biomedical devices |
Rehabilitation glove |
SMA-based actuation |
SMA; SMA |
Bending deformation |
Yes |
43
|
Artificial bladder detrusor |
Hydrogel-based actuation |
Hydrogel; Cu |
Volumetric change |
No |
97
|
Artificial bladder detrusor |
Magnetic actuation |
Magnetic silicone; NdFeB |
Volumetric change |
Yes |
66
|
Wound healing patch |
SMA-based actuation |
SMA; PTFE |
Areal contraction |
No |
100
|
Active tissue adhesive |
SMA-based actuation |
SMA; SMA |
Areal contraction |
Yes |
98
|
Biopsy microgripper |
Strain mismatch |
Thermo-sensitive polymer; Cr/Au |
Bending deformation |
No |
102
|
Endocardial balloon catheters |
Pneumatic actuation |
Silicone; Au |
Volumetric change |
Yes |
13
|
Robotic fibers |
LCE-based actuation |
LCE; stainless-steel wires |
Bending deformation |
Yes |
169
|
Microgrippers for extending drug release |
Strain mismatch |
Wax/Au/Cr; Au/Cr |
Bending deformation |
No |
34
|
Wireless devices for drug delivery |
Magnetic actuation |
Magnetic silicone; NdFeB |
Complex deformation |
Yes |
185
|
Conformal in-ear bioelectronics |
SMP-based actuation |
SMP; Au |
Spiral deformation |
No |
175
|
Deployable electrode array for brain activity mapping |
SMA-based actuation |
SMA; CNT/Au |
Complex deformation |
No |
176
|
Nerve cuff electrodes |
SMP-based actuation |
SMP; Au |
Spiral deformation |
No |
36
|
Shape-adaptive electrode arrays |
Water-responsive polymer-based actuation |
Water-responsive polymer; Au |
Spiral deformation |
No |
178
|
Nerve cuff electrodes |
Electrochemically actuated polymer |
PPy(DBS)/PaC; Au |
Spiral deformation |
Yes |
37
|
Electrocorticography system |
Fluidic actuation |
Silicone; Pt |
Complex deformation |
Yes |
181
|
|
Haptic interfaces |
Touch screen with tunable surface textures |
DE-based actuation |
Silicone; AgNW |
Areal extension and contraction |
Yes |
195
|
Dynamically reconfigurable tactile display |
Electro-hydraulic actuation |
PVC gel/dielectric liquid; MWCNT |
Volumetric change |
Yes |
193
|
Dynamically reconfigurable tactile display |
Electro-hydraulic actuation |
Polyester lidding film/dielectric liquid; carbon glue |
Linear extension and contraction |
Yes |
191
|
LCE fiber-based tactile display |
LCE-based actuation |
LCE; Cu |
Bending deformation |
Yes |
123
|
Wearable haptics with multimodal feedback |
Electro-hydraulic actuation |
PET/silicone/dielectric liquid; Al |
Volumetric change |
Yes |
44
|
Untethered feel-through wearable haptics |
DE-based actuation |
Silicone; SWCNT |
Areal extension and contraction |
Yes |
196
|
Wearable haptic artificial muscle skin |
DE-based actuation |
Synthetic elastomer; SWCNT |
Volumetric change |
Yes |
45
|
8.1 Novel materials
To advance the development of shape-morphing bioelectronic devices, novel materials are highly desired to improve the performances of actuation and electronic components.
Several key factors should be considered when designing novel active materials for actuation components. For example, the development of new soft active materials could involve molecular designs that can enable highly stable and consistent actuation deformations after thousands and even millions of cycles.198,199 Additionally, to ensure scalability, novel active materials compatible with advanced microfabrication techniques should be developed, such as light-responsive LCEs that can be patterned using soft lithography.200,201 Furthermore, active materials responsive to natural stimuli (e.g., body temperature) are ideal for implantable shape-morphing devices, but lack control over morphing onset during implantation. Therefore, development of naturally triggered yet actively controllable materials is crucial for shape-morphing bioelectronics. For instance, the shape memory hydrogels with programmable recovery onset have been reported to precisely control the initial occurrence of deformation.202
For biological and biomedical applications, devices with enhanced flexibility and stretchability are particularly desirable, as they enable conformal contact with ultra-soft cells, tissues, or organs, while posing negligible mechanical constraints. Organic electronic materials, synthesized through advanced chemical approaches, offer promising options to minimize the mechanical constraints, especially when the devices are designed for single-cell studies.115,203–205 Besides, actuation components, typically composed of soft materials or flexible structures, can form more reliable and seamless interfaces when integrated with organic electronic components, with potential to improve the overall device performance. Rich opportunities lie in developing a broader range of organic electronic elements (e.g., sensors, transistors, and others) through advanced processing techniques (e.g., laser processing, molding, and lithography)206 to further enhance the functionality and performance of shape-morphing bioelectronic devices.
8.2 Inverse design methods for shape morphing
In practical applications, the predefined deformed 3D shapes, governed by structural design, material distribution, and external stimuli, are essential for realizing specific functions of shape-morphing devices. Once fabricated, the structural design and the material distribution of shape-morphing devices are determined, and so is the accessible range of possible final 3D shapes. Powerful inverse design methods that can rapidly map target 3D geometries onto optimized 2D precursor patterns and give the required level of stimuli are foundational to the widespread utility of shape morphing.207–212 For example, at the materials level, by programming continuous levels of shape-recovery strain in SMPs and combining theoretical models with numerical algorithms, diverse 3D shapes can be achieved, such as 3D face masks.207 In addition, at the structural level, by rationally designing the network of airways embedded inside the pneumatic-actuated devices, versatile 3D shapes can be programmed using an analytic theoretical model.208 Plenty of open opportunities lie in devising novel and universal inverse design strategies for devices based on various actuation mechanisms with the assistance of artificial intelligence. Furthermore, most of the previous works focused on the inverse design of the actuation components, without considering the influence of the stiffness contribution from electronic components, which may pose an additional challenge in the inverse design.
8.3 Reprogrammability
To adapt to growing tissues of varying shapes, the bioelectronic devices need to morph from their initial shape into many different shapes on demand, driven by stimuli with a controlled spatial distribution. However, the existing shape-morphing biological/biomedical devices are mostly suitable only for single-time use, and have limited ranges of achievable geometries and functions. Future opportunities lie in the development of flexible actuators to establish new design strategies for reprogrammable devices.37,109,110,212 For example, under reprogrammable Lorentz forces, a mechanical metasurface constructed from a matrix of filamentary metal traces demonstrates complex, dynamic morphing capabilities.109 Utilizing network-shaped structures consisting of two-layered LCE ribbons, reversible and complex shape morphing can be achieved by controlling the input voltages.212 By integrating electrochemical actuators and microelectrodes, the developed peripheral nerve interfaces can form conformal contacts with nerves of diverse shapes (e.g., diameters).37 Besides, based on the buckling-guided assembly methods and loading-path-based strategies, the fabricated antennas can be reshaped into a dozen of stable 3D configurations with different radiation patterns.53,89
8.4 Actuation accuracy and speed
The accuracy of device deformation is crucial for biological/biomedical devices that require fully conformal contact with targets and reconfigurable electromagnetic/optical devices that are sensitive to their geometric configurations. To ensure a high deformation accuracy, the external stimuli should be controllable with a sufficient accuracy, and the device should be manufactured with a high precision. Besides, feedback modules and optimization algorithms can be incorporated to enhance the control accuracy.
The actuation speed is a critical metric for applications in haptic devices and diagnostics/surgery devices, where a balance between rapid response and mechanical gentleness is essential to ensure both functionality and safety. In haptic interfaces, a rapid actuation speed is key to real-time interactions that can ensure seamless and immersive experiences. Differently, this metric may be of lesser significance for applications of in vitro cell/organoid culturing, in which the duration of cell growth typically spans tens of hours to several days. Moreover, in biomedical applications involving interaction with delicate or fragile tissues (e.g., nerves, retina, and cortical surfaces), a fast actuation speed might pose mechanical risks, such as inducing unintended stress or damage. In these cases, a slow yet controllable actuation speed is preferred. For example, swelling-induced deformation in hydrogels or low-temperature SMP-based transformations can enable gentle conformal contact. Consequently, the actuation mechanisms must be carefully selected to match the mechanical and physiological requirements of specific application scenarios. Existing actuation mechanisms, such as DE actuation, electrostatic actuation, and magnetic actuation, exhibit faster response speeds (<1 s) than thermally responsive actuation, including LCE, SMP, and SMA actuation. By harnessing the rapid energy release during the switching process of bi-stable or multi-stable structures, ultrafast responses can be achieved, such as snap-through transitions.112 Opportunities lie in the development of fast-response devices through synergistic material–structure designs by combining active materials with bi-stable (or multi-stable) structures.
8.5 Biocompatibility and comfortability
While various actuation mechanisms can be exploited to induce the same pre-designed deformations, environmental constraints in specific application scenarios are important to consider in the selection of active materials/morphable structures, as well as the corresponding stimuli.
The selection of actuation and electronic materials is critical to ensure long-term safety and functionality of bioelectronic devices. Representative actuation materials include SMAs, hydrogels, SMPs, LCEs, and silicone elastomers (typically employed in pneumatic or magnetic actuation systems). SMAs and SMPs provide strong actuation forces and durability, but often require surface treatment or encapsulation to meet long-term biocompatibility requirements. Hydrogels offer excellent biocompatibility due to their high water content, although their long-term stability in in vivo applications remains a concern. LCEs show excellent responsiveness, but their biocompatibility remains to be fully validated. Silicone elastomers, especially PDMS, are both biocompatible and highly stable, making them ideal for long-term applications, though they often require further functional enhancement for active uses.
Typical electronic materials used in bioelectronic devices for sensing, signal transmission and stimulation include metals (e.g., Au, Pt, Cu, and Mg), liquid metals (e.g., EGaIn), carbon-based materials (e.g., graphene and carbon nanotubes), conductive polymers (e.g., PEDOT:PSS, PPy, and PANI), inorganic semiconductor materials (e.g., Si, Ge, and GaAs), semiconductive polymers (e.g., P3HT, PBDT-3T), and piezoelectric materials (e.g., PVDF), among others. Regarding biocompatibility and biostability, noble metals such as Au and Pt exhibit excellent biocompatibility and chemical inertness, ensuring long-term stability in physiological environments. However, metals like Cu may trigger cytotoxic effects due to ion release. Biodegradable metals like Mg offer favorable biocompatibility and could be designed to gradually degrade in vivo, but their rapid corrosion in aqueous environments poses challenges for precise stability control. The soft and stretchable liquid metals offer excellent conductivity but should be encapsulated to prevent ion release. Carbon-based materials demonstrate promising biocompatibility and chemical resilience, although their oxidative degradation and accumulation effects require careful monitoring for chronic use. Conductive polymers like PEDOT:PSS offer tunable biocompatibility and moderate biostability, but are prone to environmental degradation over time and can be stabilized through crosslinking or surface treatments. Traditional semiconductor materials such as Si and GaAs have limited inherent biocompatibility and biostability, but their inherent rigidity leads to certain degree of mechanical mismatch with cells or tissues during their integration. Semiconductive polymers offer good flexibility and biocompatibility, but their limited chemical stability under physiological conditions requires the use of encapsulation strategies. Piezoelectric materials like PVDF are widely recognized for their biocompatibility, along with excellent chemical and mechanical stability, making them well-suited for chronic implantation applications.
Moreover, the stimuli should also be well selected for different application scenarios, considering the biocompatibility and comfortability. For example, in in vivo scenarios, potentially harmful stimuli, such as chemical solvents, high temperatures, and high voltages, should be avoided. Additionally, for devices responsive to humidity or thermal stimuli, strategies to prevent unintended shape morphing during implantation should be devised. In this regard, development of magnetically actuated devices capable of both controlled locomotion and shape morphing could enable novel biomedical applications.
For skin-mounted devices (e.g., rehabilitation devices and haptic interfaces) that require softness, small thickness, lightweight characteristic, and the ability to generate relatively large output forces, the actuation mechanisms like SMA-based actuation, pneumatic actuation, hydraulically amplified electrostatic actuation, and DE-based actuation offer distinct advantages. Although significant progress has been made in the development of relevant wearable devices and systems, challenges remain in achieving thinner and lighter forms of wirelessly controlled rehabilitation devices and haptic interfaces. For instance, portable air pumps with enhanced compactness are required for pneumatic actuation. Lightweight and integrated power supply systems (i.e., high-voltage supply and control) could be developed to enable hydraulically amplified electrostatic actuation and DE-based actuation in a wireless manner.
Author contributions
Shiwei Xu: conceptualization, investigation, visualization, writing – original draft; Ruoxi Yang: investigation; Youzhou Yang: investigation; Yihui Zhang: conceptualization, supervision, project administration, funding acquisition, writing – review & editing.
Data availability
No primary research results, software or code have been included and no new data were generated or analysed as part of this review.
Conflicts of interest
The authors declare no conflict of interest.
Acknowledgements
Y. Z. acknowledges support from the National Key R&D Program of China (Grant No. 2024YFB4707300), the National Natural Science Foundation of China (Grant No. 12225206), the New Cornerstone Science Foundation through the XPLORER PRIZE, the Beijing Natural Science Foundation (Grant No. L242069), and the Tsinghua National Laboratory for Information Science and Technology.
References
- D. H. Kim, N. Lu, R. Ma, Y. S. Kim, R. H. Kim, S. Wang, J. Wu, S. M. Won, H. Tao, A. Islam, K. J. Yu, T. I. Kim, R. Chowdhury, M. Ying, L. Xu, M. Li, H. J. Chung, H. Keum, M. McCormick, P. Liu, Y. W. Zhang, F. G. Omenetto, Y. Huang, T. Coleman and J. A. Rogers, Science, 2011, 333, 838–843 CrossRef CAS PubMed.
- H. U. Chung, B. H. Kim, J. Y. Lee, J. Lee, Z. Xie, E. M. Ibler, K. Lee, A. Banks, J. Y. Jeong, J. Kim, C. Ogle, D. Grande, Y. Yu, H. Jang, P. Assem, D. Ryu, J. W. Kwak, M. Namkoong, J. B. Park, Y. Lee, D. H. Kim, A. Ryu, J. Jeong, K. You, B. Ji, Z. Liu, Q. Huo, X. Feng, Y. Deng, Y. Xu, K.-I. Jang, J. Kim, Y. Zhang, R. Ghaffari, C. M. Rand, M. Schau, A. Hamvas, D. E. Weese-Mayer, Y. Huang, S. M. Lee, C. H. Lee, N. R. Shanbhag, A. S. Paller, S. Xu and J. A. Rogers, Science, 2019, 363, eaau0780 CrossRef CAS PubMed.
- H. Li, P. Tan, Y. Rao, S. Bhattacharya, Z. Wang, S. Kim, S. Gangopadhyay, H. Shi, M. Jankovic, H. Huh, Z. Li, P. Maharjan, J. Wells, H. Jeong, Y. Jia and N. Lu, Chem. Rev., 2024, 124, 3220–3283 CrossRef CAS PubMed.
- H. Jang, K. Sel, E. Kim, S. Kim, X. Yang, S. Kang, K.-H. Ha, R. Wang, Y. Rao, R. Jafari and N. Lu, Nat. Commun., 2022, 13, 6604 CrossRef CAS PubMed.
- Y. Yang, Y. Song, X. Bo, J. Min, O. S. Pak, L. Zhu, M. Wang, J. Tu, A. Kogan, H. Zhang, T. K. Hsiai, Z. Li and W. Gao, Nat. Biotechnol., 2020, 38, 217–224 CrossRef CAS PubMed.
- Y. Zhang, X. Zhou, N. Zhang, J. Zhu, N. Bai, X. Hou, T. Sun, G. Li, L. Zhao, Y. Chen, L. Wang and C. F. Guo, Nat. Commun., 2024, 15, 3048 CrossRef CAS PubMed.
- S. H. Kim, A. Basir, R. Avila, J. Lim, S. W. Hong, G. Choe, J. H. Shin, J. H. Hwang, S. Y. Park, J. Joo, C. Lee, J. Choi, B. Lee, K.-S. Choi, S. Jung, T.-I. Kim, H. Yoo and Y. H. Jung, Nature, 2024, 629, 1047–1054 CrossRef CAS PubMed.
- B. Zhang, J. Li, J. Zhou, L. Chow, G. Zhao, Y. Huang, Z. Ma, Q. Zhang, Y. Yang, C. K. Yiu, J. Li, F. Chun, X. Huang, Y. Gao, P. Wu, S. Jia, H. Li, D. Li, Y. Liu, K. Yao, R. Shi, Z. Chen, B. L. Khoo, W. Yang, F. Wang, Z. Zheng, Z. Wang and X. Yu, Nature, 2024, 628, 84–92 CrossRef CAS PubMed.
- C. Shi, Z. Zou, Z. Lei, P. Zhu, W. Zhang and J. Xiao, Sci. Adv., 2020, 6, eabd0202 CrossRef CAS PubMed.
- Z. Zou, C. Zhu, Y. Li, X. Lei, W. Zhang and J. Xiao, Sci. Adv., 2018, 4, eaaq0508 CrossRef PubMed.
- H. Hu, Y. Ma, X. Gao, D. Song, M. Li, H. Huang, X. Qian, R. Wu, K. Shi, H. Ding, M. Lin, X. Chen, W. Zhao, B. Qi, S. Zhou, R. Chen, Y. Gu, Y. Chen, Y. Lei, C. Wang, C. Wang, Y. Tong, H. Cui, A. Abdal, Y. Zhu, X. Tian, Z. Chen, C. Lu, X. Yang, J. Mu, Z. Lou, M. Eghtedari, Q. Zhou, A. Oberai and S. Xu, Nat. Biomed. Eng., 2023, 7, 1321–1334 CrossRef PubMed.
- C. Wang, X. Chen, L. Wang, M. Makihata, H.-C. Liu, T. Zhou and X. Zhao, Science, 2022, 377, 517–523 CrossRef CAS PubMed.
- M. Han, L. Chen, K. Aras, C. Liang, X. Chen, H. Zhao, K. Li, N. R. Faye, B. Sun, J. H. Kim, W. Bai, Q. Yang, Y. Ma, W. Lu, E. Song, J. M. Baek, Y. Lee, C. Liu, J. B. Model, G. Yang, R. Ghaffari, Y. Huang, I. R. Efimov and J. A. Rogers, Nat. Biomed. Eng., 2020, 4, 997–1009 CrossRef CAS PubMed.
- K. Nan, S. Babaee, W. W. Chan, J. L. P. Kuosmanen, V. R. Feig, Y. Luo, S. S. Srinivasan, C. M. Patterson, A. M. Jebran and G. Traverso, Nat. Biomed. Eng., 2022, 6, 1092–1104 CrossRef CAS PubMed.
- K.-C. Agno, K. Yang, S.-H. Byun, S. Oh, S. Lee, H. Kim, K. Kim, S. Cho, W.-I. Jeong and J.-W. Jeong, Nat. Biomed. Eng., 2023, 8, 963–976 CrossRef PubMed.
- K. Nan, V. R. Feig, B. Ying, J. G. Howarth, Z. Kang, Y. Yang and G. Traverso, Nat. Rev. Mater., 2022, 7, 908–925 CrossRef PubMed.
- J. Shi, S. Xie, Z. Liu, M. Cai and C. F. Guo, Natl. Sci. Rev., 2024, 11, nwae351 CrossRef PubMed.
- H. Liu, S. Ding, X. Lin, S. Wang, Y. Wang, Z. Feng and J. Song, ACS Nano, 2024, 18, 19549–19560 CAS.
- P. Liu, C. Lee, N. Carusetta and R. Avila, J. Appl. Mech., 2025, 92, 051007 CrossRef.
- S. E. Lee, H. Lee, J. H. Kim, J. C. Park, S. Kyung, H. Choi, S. H. Baek, J. H. Park, S. Park, J.-M. Kim, H.-J. Jo, S. H. Cho, J. Kim, H. Kim, S. H. Han, J. K. Oh and B. H. Kim, npj Flexible Electron., 2024, 8, 65 CrossRef.
- G. Chen, X. Xiao, X. Zhao, T. Tat, M. Bick and J. Chen, Chem. Rev., 2022, 122, 3259–3291 CrossRef CAS PubMed.
- G. Gu, N. Zhang, H. Xu, S. Lin, Y. Yu, G. Chai, L. Ge, H. Yang, Q. Shao, X. Sheng, X. Zhu and X. Zhao, Nat. Biomed. Eng., 2021, 7, 589–598 CrossRef PubMed.
- G. Gu, N. Zhang, C. Chen, H. Xu and X. Zhu, ACS Nano, 2023, 17, 9661–9672 CrossRef CAS PubMed.
- C. Xu, Y. Wang, J. Zhang, J. Wan, Z. Xiang, Z. Nie, J. Xu, X. Lin, P. Zhao, Y. Wang, S. Zhang, J. Zhang, C. Liu, N. Xue, W. Zhao and M. Han, Sci. Adv., 2024, 10, eadp6094 CrossRef CAS PubMed.
- J. Park, Y. Lee, S. Cho, A. Choe, J. Yeom, Y. G. Ro, J. Kim, D.-H. Kang, S. Lee and H. Ko, Chem. Rev., 2024, 124, 1464–1534 CrossRef CAS PubMed.
- M. T. Flavin, K.-H. Ha, Z. Guo, S. Li, J.-T. Kim, T. Saxena, D. Simatos, F. Al-Najjar, Y. Mao, S. Bandapalli, C. Fan, D. Bai, Z. Zhang, Y. Zhang, E. Flavin, K. E. Madsen, Y. Huang, L. Emu, J. Zhao, J.-Y. Yoo, M. Park, J. Shin, A. G. Huang, H.-S. Shin, J. E. Colgate, Y. Huang, Z. Xie, H. Jiang and J. A. Rogers, Nature, 2024, 635, 345–352 CrossRef CAS PubMed.
- X. Yu, Z. Xie, Y. Yu, J. Lee, A. Vazquez-Guardado, H. Luan, J. Ruban, X. Ning, A. Akhtar, D. Li, B. Ji, Y. Liu, R. Sun, J. Cao, Q. Huo, Y. Zhong, C. Lee, S. Kim, P. Gutruf, C. Zhang, Y. Xue, Q. Guo, A. Chempakasseril, P. Tian, W. Lu, J. Jeong, Y. Yu, J. Cornman, C. Tan, B. Kim, K. Lee, X. Feng, Y. Huang and J. A. Rogers, Nature, 2019, 575, 473–479 CrossRef CAS PubMed.
- H. Song, G. Luo, Z. Ji, R. Bo, Z. Xue, D. Yan, F. Zhang, K. Bai, J. Liu, X. Cheng, W. Pang, Z. Shen and Y. Zhang, Sci. Adv., 2022, 8, eabm3785 CrossRef CAS PubMed.
- C. Wang, W. Hong, Y. Deng, L. Lan, S. Zhang, J. Ping, Y. Ying, C. Yu, J. Luo, W. Chen, Z. Chen and J. Song, Adv. Funct. Mater., 2025, 35, 2422980 CrossRef.
- W. Pang, S. Xu, L. Liu, R. Bo and Y. Zhang, Adv. Intell. Syst., 2023, 5, 2300060 CrossRef.
- M. Ilami, H. Bagheri, R. Ahmed, E. O. Skowronek and H. Marvi, Adv. Mater., 2021, 33, 2003139 CrossRef CAS PubMed.
- I. Apsite, S. Salehi and L. Ionov, Chem. Rev., 2022, 122, 1349–1415 CrossRef CAS PubMed.
- J. Wang and A. Chortos, Nat. Rev. Mater., 2024, 9, 738–751 CrossRef.
- A. Ghosh, L. Li, L. Xu, P. Dash Ranjeet, N. Gupta, J. Lam, Q. Jin, V. Akshintala, G. Pahapale, W. Liu, A. Sarkar, R. Rais, H. Gracias David and M. Selaru Florin, Sci. Adv., 2020, 6, eabb4133 CrossRef CAS PubMed.
- J. Rogatinsky, D. Recco, J. Feichtmeier, Y. Kang, N. Kneier, P. Hammer, E. O’Leary, D. Mah, D. Hoganson, N. V. Vasilyev and T. Ranzani, Sci. Adv., 2023, 9, eadi5559 CrossRef PubMed.
- Y. C. Zhang, N. Zheng, Y. Cao, F. L. Wang, P. Wang, Y. J. Ma, B. W. Lu, G. H. Hou, Z. Z. Fang, Z. W. Liang, M. K. Yue, Y. Li, Y. Chen, J. Fu, J. Wu, T. Xie and X. Feng, Sci. Adv., 2019, 5, eaaw1066 CrossRef CAS PubMed.
- C. Dong, A. Carnicer-Lombarte, F. Bonafè, B. Huang, S. Middya, A. Jin, X. Tao, S. Han, M. Bance, D. G. Barone, B. Fraboni and G. G. Malliaras, Nat. Mater., 2024, 23, 969–997 CrossRef CAS PubMed.
- L. Zhang, S. Xing, H. Yin, H. Weisbecker, H. T. Tran, Z. Guo, T. Han, Y. Wang, Y. Liu, Y. Wu, W. Xie, C. Huang, W. Luo, M. Demaesschalck, C. McKinney, S. Hankley, A. Huang, B. Brusseau, J. Messenger, Y. Zou and W. Bai, Nat. Commun., 2024, 15, 4777 CrossRef CAS PubMed.
- S. Wu, J. Eichenberger, J. Dai, Y. Chang, N. Ghalichechian and R. R. Zhao, Adv. Intell. Syst., 2022, 4, 2200106 CrossRef.
- Z. Tian, B. Xu, G. Wan, X. Han, Z. Di, Z. Chen and Y. Mei, Nat. Commun., 2021, 12, 509 CrossRef CAS PubMed.
- S. Chen, Z. Liu, H. Du, C. Tang, C. Y. Ji, B. Quan, R. Pan, L. Yang, X. Li, C. Gu, X. Zhang, Y. Yao, J. Li, N. X. Fang and J. Li, Nat. Commun., 2021, 12, 1299 Search PubMed.
- V. Sanchez, C. J. Walsh and R. J. Wood, Adv. Funct. Mater., 2020, 31, 2008278 CrossRef.
- M. Sui, Y. Ouyang, H. Jin, Z. Chai, C. Wei, J. Li, M. Xu, W. Li, L. Wang and S. Zhang, Nat. Mach. Intell., 2023, 5, 1149–1160 CrossRef.
- E. Leroy, R. Hinchet and H. Shea, Adv. Mater., 2020, 32, 2002564 CrossRef CAS PubMed.
- Y. Guo, Y. Luo, R. Plamthottam, S. Pei, C. Wei, Z. Han, J. Fan, M. Possinger, K. Liu, Y. Zhu, Z. Fei, I. Winardi, H. Hong, Y. Zhang, L. Jin and Q. Pei, Sci. Adv., 2024, 10, eadr1765 CrossRef PubMed.
- S. Wei and T. K. Ghosh, Adv. Mater. Technol., 2022, 7, 2101521 CrossRef.
- M. Cianchetti, C. Laschi, A. Menciassi and P. Dario, Nat. Rev. Mater., 2018, 3, 143–153 CrossRef.
- S. Yin, D. R. Yao, Y. Song, W. Heng, X. Ma, H. Han and W. Gao, Chem. Rev., 2024, 124, 11585–11636 CrossRef CAS PubMed.
- Y. Guo, L. Liu, Y. Liu and J. Leng, Adv. Intell. Syst., 2021, 3, 2000282 CrossRef.
- H. Zhao, A. M. Hussain, M. Duduta, D. M. Vogt, R. J. Wood and D. R. Clarke, Adv. Funct. Mater., 2018, 28, 1804328 CrossRef.
- C. Tang, B. Du, S. Jiang, Q. Shao, X. Dong, X.-J. Liu and H. Zhao, Sci. Rob., 2022, 7, eabm8597 CrossRef PubMed.
- P. Rothemund, N. Kellaris, S. K. Mitchell, E. Acome and C. Keplinger, Adv. Mater., 2021, 33, 2003375 CrossRef CAS PubMed.
- H. Fu, K. Nan, W. Bai, W. Huang, K. Bai, L. Lu, C. Zhou, Y. Liu, F. Liu, J. Wang, M. Han, Z. Yan, H. Luan, Y. Zhang, Y. Zhang, J. Zhao, X. Cheng, M. Li, J. W. Lee, Y. Liu, D. Fang, X. Li, Y. Huang, Y. Zhang and J. A. Rogers, Nat. Mater., 2018, 17, 268–276 CrossRef CAS PubMed.
- M. S. Kim, J. K. Heo, H. Rodrigue, H. T. Lee, S. Pane, M. W. Han and S. H. Ahn, Adv. Mater., 2023, 35, 2208517 CrossRef CAS PubMed.
- O. Erol, A. Pantula, W. Liu and D. H. Gracias, Adv. Mater. Technol., 2019, 4, 1900043 CrossRef.
- A. P. Dhand, M. D. Davidson and J. A. Burdick, Nat. Rev. Bioeng., 2025, 3, 108–125 CrossRef CAS.
- H. Li, G. Go, S. Y. Ko, J.-O. Park and S. Park, Smart Mater. Struct., 2016, 25, 027001 CrossRef.
- G. Theocharidis, H. Yuk, H. Roh, L. Wang, I. Mezghani, J. Wu, A. Kafanas, M. Contreras, B. Sumpio, Z. Li, E. Wang, L. Chen, C. F. Guo, N. Jayaswal, X. L. Katopodi, N. Kalavros, C. S. Nabzdyk, I. S. Vlachos, A. Veves and X. Zhao, Nat. Biomed. Eng., 2022, 6, 1118–1133 CrossRef CAS PubMed.
- X. P. Hao, C. Y. Li, C. W. Zhang, M. Du, Z. Ying, Q. Zheng and Z. L. Wu, Adv. Funct. Mater., 2021, 31, 2105481 CrossRef CAS.
- Y. Xia, Y. He, F. Zhang, Y. Liu and J. Leng, Adv. Mater., 2021, 33, e2000713 CrossRef PubMed.
- S. Wang, Q. Jiang, H. Liu, C. Yu, P. Li, G. Pan, K. Xu, R. Xiao, Y. Hao, C. Wang and J. Song, Proc. Natl. Acad. Sci. U. S. A., 2024, 121, e2403380121 CrossRef CAS PubMed.
- J. Zhao, L. Zhang and J. Hu, Adv. Intell. Syst., 2021, 4, 2100065 CrossRef.
- W. Pang, S. Xu, J. Wu, R. Bo, T. Jin, Y. Xiao, Z. Liu, F. Zhang, X. Cheng, K. Bai, H. Song, Z. Xue, L. Wen and Y. Zhang, Proc. Natl. Acad. Sci. U. S. A., 2022, 119, e2215028119 CrossRef CAS PubMed.
- J. Wu, S. Yao, H. Zhang, W. Man, Z. Bai, F. Zhang, X. Wang, D. Fang and Y. Zhang, Adv. Mater., 2021, 33, 2106175 CrossRef CAS PubMed.
- Y. Kim and X. Zhao, Chem. Rev., 2022, 122, 5317–5364 CrossRef CAS PubMed.
- Y. Z. Yang, J. X. Wang, L. Wang, Q. Y. Wu, L. Ling, Y. Y. Yang, S. Ning, Y. Xie, Q. L. Cao, L. Li, J. H. Liu, Q. Ling and J. F. Zang, Sci. Adv., 2022, 8, eabq1456 CrossRef PubMed.
- Y. Kim, G. A. Parada, S. Liu and X. Zhao, Sci. Rob., 2019, 4, eaax7329 CrossRef PubMed.
- H. Gu, E. Hanedan, Q. Boehler, T.-Y. Huang, A. J. T. M. Mathijssen and B. J. Nelson, Nat. Mach. Intell., 2022, 4, 678–684 CrossRef.
- H.-W. Huang, F. E. Uslu, P. Katsamba, E. Lauga, M. S. Sakar and B. J. Nelson, Sci. Adv., 2019, 5, eaau1532 CrossRef PubMed.
- X. Liang, Y. Zhao, D. Liu, Y. Deng, T. Arai, M. Kojima and X. Liu, Cyborg Bionic Syst., 2023, 4, 0060 CrossRef PubMed.
- X. Liang, Z. Chen, Y. Deng, D. Liu, X. Liu, Q. Huang and T. Arai, Cyborg Bionic Syst., 2023, 4, 0009 CrossRef CAS PubMed.
- X. Ni, X. Guo, J. Li, Y. Huang, Y. Zhang and J. A. Rogers, Adv. Mater., 2019, 31, e1905405 CrossRef PubMed.
- Y. Ling, W. Pang, X. Li, S. Goswami, Z. Xu, D. Stroman, Y. Liu, Q. Fei, Y. Xu, G. Zhao, B. Sun, J. Xie, G. Huang, Y. Zhang and Z. Yan, Adv. Mater., 2020, 32, e1908475 CrossRef PubMed.
- Y. Zhu, M. Birla, K. R. Oldham and E. T. Filipov, Adv. Funct. Mater., 2020, 30, 2003741 CrossRef CAS.
- Z. Tian, Y. Wang, Y. Chen, B. Xu, Z. Di and Y. Mei, Adv. Intell. Syst., 2020, 2, 1900092 CrossRef.
- L. Y. Yang, J. Cui, L. Zhang, X. R. Xu, X. Chen and D. P. Sun, Adv. Funct. Mater., 2021, 31, 2101378 CrossRef CAS.
- M. Wang, X. L. Tian, R. H. A. Ras and O. Ikkala, Adv. Mater. Interfaces, 2015, 2, 1500080 CrossRef.
- W. Wang, L. N. Yao, C. Y. Cheng, T. Zhang, H. Atsumi, L. D. Wang, G. Y. Wang, O. Anilionyte, H. Steiner, J. F. Ou, K. Zhou, C. Wawrousek, K. Petrecca, A. M. Belcher, R. Karnik, X. H. Zhao, D. I. C. Wang and H. Ishii, Sci. Adv., 2017, 3, e1601984 CrossRef PubMed.
- B. Gorissen, D. Reynaerts, S. Konishi, K. Yoshida, J. W. Kim and M. De Volder, Adv. Mater., 2017, 29, 1604977 CrossRef PubMed.
- R. Bo, S. Xu, Y. Yang and Y. Zhang, Chem. Rev., 2023, 123, 11137–11189 CrossRef CAS PubMed.
- S. Xu, Z. Yan, K. I. Jang, W. Huang, H. Fu, J. Kim, Z. Wei, M. Flavin, J. McCracken, R. Wang, A. Badea, Y. Liu, D. Xiao, G. Zhou, J. Lee, H. U. Chung, H. Cheng, W. Ren, A. Banks, X. Li, U. Paik, R. G. Nuzzo, Y. Huang, Y. Zhang and J. A. Rogers, Science, 2015, 347, 154–159 CrossRef CAS PubMed.
- X. Cheng, Z. Fan, S. Yao, T. Jin, Z. Lv, Y. Lan, R. Bo, Y. Chen, F. Zhang, Z. Shen, H. Wan, Y. Huang and Y. Zhang, Science, 2023, 379, 1225–1232 CrossRef CAS PubMed.
- Z. Liu, X. Hu, R. Bo, Y. Yang, X. Cheng, W. Pang, Q. Liu, Y. Wang, S. Wang, S. Xu, Z. Shen and Y. Zhang, Science, 2024, 384, 987–994 CrossRef CAS PubMed.
- S. Xu, Z. Tang, R. Yang, Z. Xue, Y. Huang and Y. Zhang, Eur. J. Mech. A-Solids, 2025, 111, 105549 CrossRef.
- Y. Zhang, F. Zhang, Z. Yan, Q. Ma, X. Li, Y. Huang and J. A. Rogers, Nat. Rev. Mater., 2017, 2, 17019 CrossRef CAS.
- Y. Zhang, Z. Yan, K. Nan, D. Xiao, Y. Liu, H. Luan, H. Fu, X. Wang, Q. Yang, J. Wang, W. Ren, H. Si, F. Liu, L. Yang, H. Li, J. Wang, X. Guo, H. Luo, L. Wang, Y. Huang and J. A. Rogers, Proc. Natl. Acad. Sci. U. S. A., 2015, 112, 11757–11764 CrossRef CAS PubMed.
- X. Ning, H. Wang, X. Yu, J. Soares, Z. Yan, K. Nan, G. Velarde, Y. Xue, R. Sun, Q. Dong, H. Luan, C. M. Lee, A. Chempakasseril, M. Han, Y. Wang, L. Li, Y. Huang, Y. Zhang and J. Rogers, Adv. Funct. Mater., 2017, 27, 1605914 CrossRef PubMed.
- X. Ning, X. Yu, H. Wang, R. Sun, R. E. Corman, H. Li, C. M. Lee, Y. Xue, A. Chempakasseril, Y. Yao, Z. Zhang, H. Luan, Z. Wang, W. Xia, X. Feng, R. H. Ewoldt, Y. Huang, Y. Zhang and J. A. Rogers, Sci. Adv., 2018, 4, eaat8313 CrossRef PubMed.
- K. Bai, X. Cheng, Z. Xue, H. Song, L. Sang, F. Zhang, F. Liu, X. Luo, W. Huang, Y. Huang and Y. Zhang, Sci. Adv., 2020, 6, eabb7417 CrossRef CAS PubMed.
- W. Lee, Y. Liu, Y. Lee, B. K. Sharma, S. M. Shinde, S. D. Kim, K. Nan, Z. Yan, M. Han, Y. Huang, Y. Zhang, J. H. Ahn and J. A. Rogers, Nat. Commun., 2018, 9, 1417 CrossRef PubMed.
- Y. Park, C. K. Franz, H. Ryu, H. Luan, K. Y. Cotton, J. U. Kim, T. S. Chung, S. Zhao, A. Vazquez-Guardado, D. S. Yang, K. Li, R. Avila, J. K. Phillips, M. J. Quezada, H. Jang, S. S. Kwak, S. M. Won, K. Kwon, H. Jeong, A. J. Bandodkar, M. Han, H. Zhao, G. R. Osher, H. Wang, K. Lee, Y. Zhang, Y. Huang, J. D. Finan and J. A. Rogers, Sci. Adv., 2021, 7, eabf9153 CrossRef CAS PubMed.
- S. An, X. Li, Z. Guo, Y. Huang, Y. Zhang and H. Jiang, Nat. Commun., 2024, 15, 7340 CrossRef CAS.
- C. Zhang, H. Deng, Y. Xie, C. Zhang, J. W. Su and J. Lin, Small, 2019, 15, e1904224 CrossRef PubMed.
- Y. Li, C. Luo, K. Yu and X. Wang, ACS Appl. Mater. Interfaces, 2021, 13, 8929–8939 CrossRef CAS PubMed.
- J. K. Park, K. Nan, H. Luan, N. Zheng, S. Zhao, H. Zhang, X. Cheng, H. Wang, K. Li, T. Xie, Y. Huang, Y. Zhang, S. Kim and J. A. Rogers, Adv. Mater., 2019, 31, e1905715 CrossRef PubMed.
- W. Pang, X. Cheng, H. Zhao, X. Guo, Z. Ji, G. Li, Y. Liang, Z. Xue, H. Song, F. Zhang, Z. Xu, L. Sang, W. Huang, T. Li and Y. Zhang, Natl. Sci. Rev., 2019, 7, 342–354 CrossRef PubMed.
- X. Yang, C. An, S. Liu, T. Cheng, V. Bunpetch, Y. Liu, S. Dong, S. Li, X. Zou, T. Li, H. Ouyang, Z. Wu and W. Yang, Adv. Healthcare Mater., 2018, 7, 1701014 CrossRef PubMed.
- S. Nam, B. R. Seo, A. J. Najibi, S. L. McNamara and D. J. Mooney, Nat. Mater., 2022, 22, 249–259 CrossRef PubMed.
- Q. Ma, H. Cheng, K. I. Jang, H. Luan, K. C. Hwang, J. A. Rogers, Y. Huang and Y. Zhang, J. Mech. Phys. Solids, 2016, 90, 179–202 CrossRef PubMed.
- G. Yao, X. Mo, C. Yin, W. Lou, Q. Wang, S. Huang, L. Mao, S. Chen, K. Zhao, T. Pan, L. Huang and Y. Lin, Sci. Adv., 2022, 8, eabl8379 CrossRef CAS PubMed.
- J. C. Breger, C. Yoon, R. Xiao, H. R. Kwag, M. O. Wang, J. P. Fisher, T. D. Nguyen and D. H. Gracias, ACS Appl. Mater. Interfaces, 2015, 7, 3398–3405 CrossRef CAS.
- E. Gultepe, J. S. Randhawa, S. Kadam, S. Yamanaka, F. M. Selaru, E. J. Shin, A. N. Kalloo and D. H. Gracias, Adv. Mater., 2013, 25, 514–519 CrossRef CAS PubMed.
- Q. Huang, B. Tang, J. C. Romero, Y. Yang, S. K. Elsayed, G. Pahapale, T.-J. Lee, I. E. Morales Pantoja, F. Han, C. Berlinicke, T. Xiang, M. Solazzo, T. Hartung, Z. Qin, B. S. Caffo, L. Smirnova and D. H. Gracias, Sci. Adv., 2022, 8, eabq5031 CrossRef CAS PubMed.
- A. Kalmykov, C. Huang, J. Bliley, D. Shiwarski, J. Tashman, A. Abdullah, S. K. Rastogi, S. Shukla, E. Mataev, A. W. Feinberg, K. J. Hsia and T. Cohen-Karni, Sci. Adv., 2019, 5, eaax0729 CrossRef CAS PubMed.
- H. Ju, H. Park, N. Kim, J. Lim, D. Jung and J. Lee, Soft Rob., 2022, 9, 767–775 CrossRef PubMed.
- J.-H. Na, A. A. Evans, J. Bae, M. C. Chiappelli, C. D. Santangelo, R. J. Lang, T. C. Hull and R. C. Hayward, Adv. Mater., 2015, 27, 79–85 CrossRef CAS PubMed.
- Q. Liu, W. Wang, M. F. Reynolds, M. C. Cao, M. Z. Miskin, T. A. Arias, D. A. Muller, P. L. McEuen and I. Cohen, Sci. Rob., 2021, 6, eabe6663 CrossRef PubMed.
- J. Cools, Q. Jin, E. Yoon, D. A. Burbano, Z. Luo, D. Cuypers, G. Callewaert, D. Braeken and D. H. Gracias, Adv. Sci., 2018, 5, 1700731 CrossRef PubMed.
- Y. Bai, H. Wang, Y. Xue, Y. Pan, J. T. Kim, X. Ni, T. L. Liu, Y. Yang, M. Han, Y. Huang, J. A. Rogers and X. Ni, Nature, 2022, 609, 701–708 CrossRef CAS PubMed.
- X. Ni, H. Luan, J. T. Kim, S. I. Rogge, Y. Bai, J. W. Kwak, S. Liu, D. S. Yang, S. Li, S. Li, Z. Li, Y. Zhang, C. Wu, X. Ni, Y. Huang, H. Wang and J. A. Rogers, Nat. Commun., 2022, 13, 5576 CrossRef CAS PubMed.
- S. Patel, Z. Rao, M. Yang and C. Yu, Adv. Funct. Mater., 2025, 35, 2417906 CrossRef.
- Y. Chi, Y. Li, Y. Zhao, Y. Hong, Y. Tang and J. Yin, Adv. Mater., 2022, 34, 2110384 CrossRef CAS.
- T. Rafeedi and D. J. Lipomi, Science, 2022, 378, 1174–1175 CrossRef CAS.
- H. Ling, S. Liu, Z. Zheng and F. Yan, Small Methods, 2018, 2, 1800070 CrossRef.
- Y.-Q. Zheng, Y. Liu, D. Zhong, S. Nikzad, S. Liu, Z. Yu, D. Liu, H.-C. Wu, C. Zhu, J. Li, H. Tran, J. B. H. Tok and Z. Bao, Science, 2021, 373, 88–94 CrossRef CAS PubMed.
- R. Yang, A. Dutta, B. Li, N. Tiwari, W. Zhang, Z. Niu, Y. Gao, D. Erdely, X. Xin, T. Li and H. Cheng, Nat. Commun., 2023, 14, 2907 CrossRef CAS PubMed.
- Z. Xue, H. Song, J. A. Rogers, Y. Zhang and Y. Huang, Adv. Mater., 2019, 32, 1902254 CrossRef PubMed.
- J. H. Kim, S. E. Lee and B. H. Kim, Soft Sci., 2023, 3, 16 CrossRef CAS.
- A. Kotikian, J. M. Morales, A. Lu, J. Mueller, Z. S. Davidson, J. W. Boley and J. A. Lewis, Adv. Mater., 2021, 33, 2101814 CrossRef CAS PubMed.
- M. R. Vinciguerra, D. K. Patel, W. Zu, M. Tavakoli, C. Majidi and L. Yao, ACS Appl. Mater. Interfaces, 2023, 15, 24777–24787 CrossRef CAS PubMed.
- W. Wang, Q. Liu, I. Tanasijevic, M. F. Reynolds, A. J. Cortese, M. Z. Miskin, M. C. Cao, D. A. Muller, A. C. Molnar, E. Lauga, P. L. McEuen and I. Cohen, Nature, 2022, 605, 681–686 CrossRef CAS PubMed.
- E. Martinelli, O. Akouissi, L. Liebi, I. Furfaro, D. Maulà, N. Savoia, A. Remy, L. Nikles, A. Roux, L. Stoppini and S. P. Lacour, Sci. Adv., 2024, 10, eadp8054 CrossRef CAS PubMed.
- S. A. M. Weima, R. Norouzikudiani, J. Baek, J. A. Peixoto, T. K. Slot, D. J. Broer, A. DeSimone and D. Liu, Sci. Adv., 2024, 10, eadp0421 CrossRef CAS PubMed.
- L. Bao and T. J. Cui, Microwave Opt. Technol. Lett., 2019, 62, 9–32 CrossRef.
- J. Lee, S. J. Ihle, G. S. Pellegrino, H. Kim, J. Yea, C.-Y. Jeon, H.-C. Son, C. Jin, D. Eberli, F. Schmid, B. L. Zambrano, A. F. Renz, C. Forró, H. Choi, K.-I. Jang, R. Küng and J. Vörös, Nat. Electron., 2021, 4, 291–301 CrossRef CAS.
- F. Liu, X. Cheng, F. Zhang, Y. Chen, H. Song, Y. Huang and Y. Zhang, Adv. Electron. Mater., 2019, 5, 1900256 CrossRef.
- Q. Ze, X. Kuang, S. Wu, J. Wong, S. M. Montgomery, R. Zhang, J. M. Kovitz, F. Yang, H. J. Qi and R. Zhao, Adv. Mater., 2020, 32, 1906657 CrossRef CAS PubMed.
- Y. F. Yin, Y. N. Yu and Y. H. Li, Mech. Adv. Mater. Struct., 2021, 29, 3903–3910 CrossRef.
- J. Ye, F. Zhang, Z. Shen, S. Cao, T. Jin, X. Guo, Z. Li, L. Lin and Y. Zhang, npj Flexible Electron., 2021, 5, 28 CrossRef.
- Z. Xue, T. Jin, S. Xu, K. Bai, Q. He, F. Zhang, X. Cheng, Z. Ji, W. Pang, Z. Shen, H. Song, Y. Shuai and Y. Zhang, Sci. Adv., 2022, 8, eabm6922 CrossRef CAS PubMed.
- F. Liu, Y. Chen, H. Song, F. Zhang, Z. Fan, Y. Liu, X. Feng, J. A. Rogers, Y. Huang and Y. Zhang, Small, 2019, 15, e1804055 CrossRef PubMed.
- T. N. Do, H. Phan, T. Q. Nguyen and Y. Visell, Adv. Funct. Mater., 2018, 28, 1800244 CrossRef.
- G. Mao, D. Schiller, D. Danninger, B. Hailegnaw, F. Hartmann, T. Stockinger, M. Drack, N. Arnold and M. Kaltenbrunner, Nat. Commun., 2022, 13, 4456 CrossRef CAS PubMed.
- H. Tao, A. C. Strikwerda, K. Fan, W. J. Padilla, X. Zhang and R. D. Averitt, Phys. Rev. Lett., 2009, 103, 147401 CrossRef PubMed.
- B. Aksoy and H. Shea, Sci. Adv., 2022, 8, eabk0543 CrossRef CAS PubMed.
- G. Mao, M. Drack, M. Karami-Mosammam, D. Wirthl, T. Stockinger, R. Schwödiauer and M. Kaltenbrunner, Sci. Adv., 2020, 6, eabc0251 CrossRef CAS PubMed.
- S. Oh, S. Lee, S. H. Byun, S. Lee, C. Y. Kim, J. Yea, S. Chung, S. Li, K. I. Jang, J. Kang and J. W. Jeong, Adv. Funct. Mater., 2023, 33, 2214766 CrossRef CAS.
- Y. Lee, B. J. Kim, L. Hu, J. Hong and J.-H. Ahn, Mater. Today, 2022, 53, 51–57 CrossRef CAS.
- S.-B. Kim, D. Lee, J. Kim, T. Kim, J. H. Sim, J.-H. Yang, S. J. Oh, S. Hahn, W. Lee, D. Choi, T.-S. Kim, H. Moon and S. Yoo, Nat. Commun., 2024, 15, 7802 CrossRef CAS PubMed.
- D. C. Kim, H. Seung, J. Yoo, J. Kim, H. H. Song, J. S. Kim, Y. Kim, K. Lee, C. Choi, D. Jung, C. Park, H. Heo, J. Yang, T. Hyeon, M. K. Choi and D.-H. Kim, Nat. Electron., 2024, 7, 365–374 CrossRef CAS.
- D. C. Kim, H. Yun, J. Kim, H. Seung, W. S. Yu, J. H. Koo, J. Yang, J. H. Kim, T. Hyeon and D.-H. Kim, Nat. Electron., 2021, 4, 671–680 CrossRef CAS.
- W. Zhang, H. Tian, T. Liu, H. Liu, F. Zhao, X. Li, C. Wang, X. Chen and J. Shao, Mater. Horiz., 2023, 10, 2024–2034 RSC.
- M. Zhang, A. Pal, Z. Zheng, G. Gardi, E. Yildiz and M. Sitti, Nat. Mater., 2023, 22, 1243–1252 CrossRef CAS PubMed.
- H. Zhao, K. Li, M. Han, F. Zhu, A. Vazquez-Guardado, P. Guo, Z. Xie, Y. Park, L. Chen, X. Wang, H. Luan, Y. Yang, H. Wang, C. Liang, Y. Xue, R. D. Schaller, D. Chanda, Y. Huang, Y. Zhang and J. A. Rogers, Proc. Natl. Acad. Sci. U. S. A., 2019, 116, 13239–13248 CrossRef CAS PubMed.
- Z. Liu, H. Du, J. Li, L. Lu, Z.-Y. Li and N. X. Fang, Sci. Adv., 2018, 4, eaat4436 CrossRef PubMed.
- Z. Rao, Y. Lu, Z. Li, K. Sim, Z. Ma, J. Xiao and C. Yu, Nat. Electron., 2021, 4, 513–521 CrossRef.
- I. Jung, J. Xiao, V. Malyarchuk, C. Lu, M. Li, Z. Liu, J. Yoon, Y. Huang and J. A. Rogers, Proc. Natl. Acad. Sci. U. S. A., 2011, 108, 1788–1793 CrossRef CAS PubMed.
- S. Xu, X. Hu, R. Yang, C. Zang, L. Li, Y. Xiao, W. Liu, B. Tian, W. Pang, R. Bo, Q. Liu, Y. Yang, Y. Lai, J. Wu, H. Zhao, L. Wen and Y. Zhang, Nat. Mach. Intell., 2025 DOI:10.1038/s42256-025-01028-4.
- H. C. Ko, M. P. Stoykovich, J. Song, V. Malyarchuk, W. M. Choi, C. J. Yu, J. B. Geddes, 3rd, J. Xiao, S. Wang, Y. Huang and J. A. Rogers, Nature, 2008, 454, 748–753 CrossRef CAS PubMed.
- Y. M. Song, Y. Xie, V. Malyarchuk, J. Xiao, I. Jung, K.-J. Choi, Z. Liu, H. Park, C. Lu, R.-H. Kim, R. Li, K. B. Crozier, Y. Huang and J. A. Rogers, Nature, 2013, 497, 95–99 CrossRef CAS PubMed.
- Q. Liu and Y. Zhang, Sci. Rob., 2024, 9, eadp5682 CrossRef PubMed.
- K. Malachowski, M. Jamal, Q. Jin, B. Polat, C. J. Morris and D. H. Gracias, Nano Lett., 2014, 14, 4164–4170 CrossRef CAS PubMed.
- W. Xi, C. K. Schmidt, S. Sanchez, D. H. Gracias, R. E. Carazo-Salas, S. P. Jackson and O. G. Schmidt, Nano Lett., 2014, 14, 4197–4204 CrossRef CAS PubMed.
- W. Xu, S. K. Paidi, Z. Qin, Q. Huang, C.-H. Yu, J. V. Pagaduan, M. J. Buehler, I. Barman and D. H. Gracias, Nano Lett., 2019, 19, 1409–1417 CrossRef CAS PubMed.
- X. Cheng, Z. Shen and Y. Zhang, Natl. Sci. Rev., 2024, 11, nwad314 CrossRef PubMed.
- L. Hu, J. Bonnemain, M. Y. Saeed, M. Singh, D. Quevedo Moreno, N. V. Vasilyev and E. T. Roche, Nat. Biomed. Eng., 2022, 7, 110–123 CrossRef PubMed.
- J. Davies, M. T. Thai, B. Sharma, T. T. Hoang, C. C. Nguyen, P. T. Phan, T. N. A. M. Vuong, A. Ji, K. Zhu, E. Nicotra, Y.-C. Toh, M. Stevens, C. Hayward, H.-P. Phan, N. H. Lovell and T. N. Do, Sci. Rob., 2024, 9, eado4553 CrossRef PubMed.
- T. Roche Ellen, A. Horvath Markus, I. Wamala, A. Alazmani, S.-E. Song, W. Whyte, Z. Machaidze, J. Payne Christopher, C. Weaver James, G. Fishbein, J. Kuebler, V. Vasilyev Nikolay, J. Mooney David, A. Pigula Frank and J. Walsh Conor, Sci. Transl. Med., 2017, 9, eaaf3925 CrossRef PubMed.
- L. Merces, L. M. M. Ferro, A. Thomas, D. D. Karnaushenko, Y. Luo, A. I. Egunov, W. Zhang, V. K. Bandari, Y. Lee, J. S. McCaskill, M. Zhu, O. G. Schmidt and D. Karnaushenko, Adv. Mater., 2024, 36, 2313327 CrossRef CAS PubMed.
- Y. Dong, L. Wang, N. Xia, Z. Yang, C. Zhang, C. Pan, D. Jin, J. Zhang, C. Majidi and L. Zhang, Sci. Adv., 2022, 8, eabn8932 CrossRef CAS PubMed.
- Y. Kim, Y. Yang, X. Zhang, Z. Li, A. Vázquez-Guardado, I. Park, J. Wang, A. I. Efimov, Z. Dou, Y. Wang, J. Park, H. Luan, X. Ni, Y. S. Kim, J. Baek, J. J. Park, Z. Xie, H. Zhao, M. Gazzola, J. A. Rogers and R. Bashir, Sci. Rob., 2024, 8, eadd1053 CrossRef PubMed.
- H. Tetsuka, S. Gobbi, T. Hatanaka, L. Pirrami and S. R. Shin, Sci. Rob., 2024, 9, eado0051 CrossRef PubMed.
- K. Y. Lee, S.-J. Park, D. G. Matthews, S. L. Kim, C. A. Marquez, J. F. Zimmerman, H. A. M. Ardoña, A. G. Kleber, G. V. Lauder and K. K. Parker, Science, 2022, 375, 639–647 CrossRef CAS PubMed.
- Y. Piskarev, J. Shintake, C. Chautems, J. Lussi, Q. Boehler, B. J. Nelson and D. Floreano, Adv. Funct. Mater., 2022, 32, 2107662 CrossRef CAS.
- Q. Jin, Y. Yang, J. A. Jackson, C. Yoon and D. H. Gracias, Nano Lett., 2020, 20, 5383–5390 CrossRef CAS PubMed.
- R. H. Soon, Z. Yin, M. A. Dogan, N. O. Dogan, M. E. Tiryaki, A. C. Karacakol, A. Aydin, P. Esmaeili-Dokht and M. Sitti, Nat. Commun., 2023, 14, 3320 CrossRef CAS PubMed.
- H. Suzuki and R. J. Wood, Nat. Mach. Intell., 2020, 2, 437–446 CrossRef.
- T. G. Leong, C. L. Randall, B. R. Benson, N. Bassik, G. M. Stern and D. H. Gracias, Proc. Natl. Acad. Sci. U. S. A., 2009, 106, 703–708 CrossRef CAS PubMed.
- M. E. M. K. Abdelaziz, J. Zhao, B. Gil Rosa, H.-T. Lee, D. Simon, K. Vyas, B. Li, H. Koguna, Y. Li, A. A. Demircali, H. Uvet, G. Gencoglan, A. Akcay, M. Elriedy, J. Kinross, R. Dasgupta, Z. Takats, E. Yeatman, G.-Z. Yang and B. Temelkuran, Sci. Adv., 2024, 10, eadj1984 CrossRef CAS PubMed.
- S. Babaee, Y. Shi, S. Abbasalizadeh, S. Tamang, K. Hess, J. E. Collins, K. Ishida, A. Lopes, M. Williams, M. Albaghdadi, A. M. Hayward and G. Traverso, Nat. Mater., 2021, 20, 1085–1092 CrossRef CAS PubMed.
- X. L. Chang, P. S. Chee, E. H. Lim and W. C. Chong, Smart Mater. Struct., 2019, 28, 015024 CrossRef CAS.
- Z. Zhu, Z. Yan, S. Ni, H. Yang, Y. Xie, X. Wang, D. Zou, C. Tao, W. Jiang, J. Jiang, Z. Su, Y. Xia, Z. Zhou, L. Sun, C. Fan, T. H. Tao, X. Wei, Y. Qian and K. Liu, Adv. Mater., 2024, 36, 2405892 CrossRef CAS PubMed.
- Y. Cui, L. Li, C. Liu, Y. Wang, M. Sun, B. Jia, Z. Shen, X. Sheng and Y. Deng, Nano Lett., 2023, 23, 11693–11701 CrossRef CAS PubMed.
- T. J. Oxley, N. L. Opie, S. E. John, G. S. Rind, S. M. Ronayne, T. L. Wheeler, J. W. Judy, A. J. McDonald, A. Dornom, T. J. H. Lovell, C. Steward, D. J. Garrett, B. A. Moffat, E. H. Lui, N. Yassi, B. C. V. Campbell, Y. T. Wong, K. E. Fox, E. S. Nurse, I. E. Bennett, S. H. Bauquier, K. A. Liyanage, N. R. van der Nagel, P. Perucca, A. Ahnood, K. P. Gill, B. Yan, L. Churilov, C. R. French, P. M. Desmond, M. K. Horne, L. Kiers, S. Prawer, S. M. Davis, A. N. Burkitt, P. J. Mitchell, D. B. Grayden, C. N. May and T. J. O'Brien, Nat. Biotechnol., 2016, 34, 320–327 CrossRef CAS PubMed.
- Z. Wang, N. Shi, Y. Zhang, N. Zheng, H. Li, Y. Jiao, J. Cheng, Y. Wang, X. Zhang, Y. Chen, Y. Chen, H. Wang, T. Xie, Y. Wang, Y. Ma, X. Gao and X. Feng, Nat. Commun., 2023, 14, 4213 CrossRef CAS PubMed.
- S. Wei, A. Jiang, H. Sun, J. Zhu, S. Jia, X. Liu, Z. Xu, J. Zhang, Y. Shang, X. Fu, G. Li, P. Wang, Z. Xia, T. Jiang, A. Cao and X. Duan, Nat. Commun., 2024, 15, 715 CrossRef CAS PubMed.
- J.-Y. Bae, G.-S. Hwang, Y.-S. Kim, J. Jeon, M. Chae, J.-W. Kim, S. Lee, S. Kim, S.-H. Lee, S.-G. Choi, J.-Y. Lee, J.-H. Lee, K.-S. Kim, J.-H. Park, W.-J. Lee, Y.-C. Kim, K.-S. Lee, J. Kim, H. Lee, J. K. Hyun, J.-Y. Kim and S.-K. Kang, Nat. Electron., 2024, 7, 815–828 CrossRef.
- J. Yi, G. Zou, J. Huang, X. Ren, Q. Tian, Q. Yu, P. Wang, Y. Yuan, W. Tang, C. Wang, L. Liang, Z. Cao, Y. Li, M. Yu, Y. Jiang, F. Zhang, X. Yang, W. Li, X. Wang, Y. Luo, X. J. Loh, G. Li, B. Hu, Z. Liu, H. Gao and X. Chen, Nature, 2023, 624, 295–302 CrossRef CAS PubMed.
- V. Kashyap, A. Caprio, T. Doshi, S.-J. Jang, C. F. Liu, B. Mosadegh and S. Dunham, Sci. Adv., 2020, 6, eabc6800 CrossRef CAS PubMed.
- B. J. Woodington, V. F. Curto, Y.-L. Yu, H. Martínez-Domínguez, L. Coles, G. G. Malliaras, C. M. Proctor and D. G. Barone, Sci. Adv., 2021, 7, eabg7833 CrossRef PubMed.
- S. Song, F. Fallegger, A. Trouillet, K. Kim and S. P. Lacour, Sci. Rob., 2023, 8, eadd1002 CrossRef PubMed.
- Y. Chen, W. Yu, A. Benali, D. Lu, S. Y. Kok and R. Wang, Front. Aging Neurosci., 2024, 16, 1327397 CrossRef PubMed.
- Y. Yin, C. Guo, Q. Mu, W. Li, H. Yang and Y. He, Chem. Eng. J., 2024, 500, 157115 CrossRef CAS.
- Z. Zhang, K. Wu, Z. Wu, Y. Xiao, Y. Wang, Q. Lin, C. Wang, Q. Zhu, Y. Xiao and Q. Liu, BMC Pediatr., 2024, 24, 729 CrossRef PubMed.
- Y. Wu, X. Dong, J.-K. Kim, C. Wang and M. Sitti, Sci. Adv., 2022, 8, eabn3431 CrossRef CAS PubMed.
- Ankit, T. Y. K. Ho, A. Nirmal, M. R. Kulkarni, D. Accoto and N. Mathews, Adv. Intell. Syst., 2021, 4, 2100061 CrossRef.
- Y. Qiu, Z. Lu and Q. Pei, ACS Appl. Mater. Interfaces, 2018, 10, 24807–24815 CrossRef CAS PubMed.
- X. C. Qu, X. Ma, B. J. Shi, H. Li, L. Zheng, C. Wang, Z. Liu, Y. B. Fan, X. Y. Chen, Z. Li and Z. L. Wang, Adv. Funct. Mater., 2021, 31, 2006612 CrossRef CAS.
- A. Firouzeh, A. Mizutani, J. Groten, M. Zirkl and H. Shea, Adv. Mater., 2024, 36, 2307636 CrossRef CAS PubMed.
- N. Besse, S. Rosset, J. J. Zarate and H. Shea, Adv. Mater. Technol., 2017, 2, 1700102 CrossRef.
- B. K. Johnson, M. Naris, V. Sundaram, A. Volchko, K. Ly, S. K. Mitchell, E. Acome, N. Kellaris, C. Keplinger, N. Correll, J. S. Humbert and M. E. Rentschler, Nat. Commun., 2023, 14, 4516 CrossRef CAS PubMed.
- Z. Zhang, Z. Xu, L. Emu, P. Wei, S. Chen, Z. Zhai, L. Kong, Y. Wang and H. Jiang, Nat. Mach. Intell., 2023, 5, 643–655 CrossRef.
- S.-Y. Jang, M. Cho, H. Kim, M. Choi, S. Mun, J.-H. Youn, J. Park, G. Hwang, I. Hwang, S. Yun and K.-U. Kyung, Sci. Adv., 2024, 10, eadq2024 CrossRef CAS PubMed.
- G.-H. Feng and S.-Y. Hou, Sens. Actuators, A, 2018, 275, 137–147 CrossRef CAS.
- K. Jun, J. Kim and I. K. Oh, Small, 2018, 14, e1801603 CrossRef PubMed.
- X. Ji, X. Liu, V. Cacucciolo, Y. Civet, A. El Haitami, S. Cantin, Y. Perriard and H. Shea, Adv. Funct. Mater., 2020, 31, 2006639 CrossRef.
- Q. Xiong, X. Liang, D. Wei, H. Wang, R. Zhu, T. Wang, J. Mao and H. Wang, IEEE Trans. Rob., 2022, 38, 3450–3462 Search PubMed.
- Z. Han, Y. Lu and S. Qu, Adv. Funct. Mater., 2024, 34, 2313498 CrossRef CAS.
- X. Li and J. P. Gong, Nat. Rev. Mater., 2024, 9, 380–398 CrossRef CAS.
- S. Li, M. M. Lerch, J. T. Waters, B. Deng, R. S. Martens, Y. Yao, D. Y. Kim, K. Bertoldi, A. Grinthal, A. C. Balazs and J. Aizenberg, Nature, 2022, 605, 76–83 CrossRef CAS PubMed.
- Y. Yao, J. T. Waters, A. V. Shneidman, J. Cui, X. Wang, N. K. Mandsberg, S. Li, A. C. Balazs and J. Aizenberg, Proc. Natl. Acad. Sci. U. S. A., 2018, 115, 12950–12955 CrossRef CAS PubMed.
- C. Ni, D. Chen, Y. Yin, X. Wen, X. Chen, C. Yang, G. Chen, Z. Sun, J. Wen, Y. Jiao, C. Wang, N. Wang, X. Kong, S. Deng, Y. Shen, R. Xiao, X. Jin, J. Li, X. Kong, Q. Zhao and T. Xie, Nature, 2023, 622, 748–753 CrossRef CAS PubMed.
- J. Liu, X. Zhang, Y. Liu, M. Rodrigo, P. D. Loftus, J. Aparicio-Valenzuela, J. Zheng, T. Pong, K. J. Cyr, M. Babakhanian, J. Hasi, J. Li, Y. Jiang, C. J. Kenney, P. J. Wang, A. M. Lee and Z. Bao, Proc. Natl. Acad. Sci. U. S. A., 2020, 117, 14769–14778 CrossRef CAS PubMed.
- Y. Liu, J. Li, S. Song, J. Kang, Y. Tsao, S. Chen, V. Mottini, K. McConnell, W. Xu, Y.-Q. Zheng, J. B. H. Tok, P. M. George and Z. Bao, Nat. Biotechnol., 2020, 38, 1031–1036 CrossRef CAS PubMed.
- H. Shim, K. Sim, B. Wang, Y. Zhang, S. Patel, S. Jang, T. J. Marks, A. Facchetti and C. Yu, Nat. Electron., 2023, 6, 349–359 CrossRef CAS.
- R. Yang, W. Zhang, N. Tiwari, H. Yan, T. Li and H. Cheng, Adv. Sci., 2022, 9, 2202470 CrossRef PubMed.
- S. J. Ahn, J. Byun, H. J. Joo, J. H. Jeong, D. Y. Lee and K. J. Cho, Adv. Mater. Technol., 2021, 6, 2100133 CrossRef.
- E. Siefert, E. Reyssat, J. Bico and B. Roman, Nat. Mater., 2019, 18, 24–28 CrossRef CAS PubMed.
- H. Aharoni, Y. Xia, X. Zhang, R. D. Kamien and S. Yang, Proc. Natl. Acad. Sci. U. S. A., 2018, 115, 7206–7211 CrossRef CAS PubMed.
- J. W. Boley, W. M. van Rees, C. Lissandrello, M. N. Horenstein, R. L. Truby, A. Kotikian, J. A. Lewis and L. Mahadevan, Proc. Natl. Acad. Sci. U. S. A., 2019, 116, 20856–20862 CrossRef CAS PubMed.
- E. Hajiesmaili, M. Larson Natalie, A. Lewis Jennifer and R. Clarke David, Sci. Adv., 2022, 8, eabn9198 CrossRef PubMed.
- K. Liu, F. Hacker and C. Daraio, Sci. Rob., 2021, 6, eabf5116 CrossRef PubMed.
|
This journal is © The Royal Society of Chemistry 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.