Abstract
Architected flexible electronic devices with rationally designed 3D geometries have found essential applications in biology, medicine, therapeutics, sensing/imaging, energy, robotics, and daily healthcare. Mechanically-guided 3D assembly methods, exploiting mechanics principles of materials and structures to transform planar electronic devices fabricated using mature semiconductor techniques into 3D architected ones, are promising routes to such architected flexible electronic devices. Here, we comprehensively review mechanically-guided 3D assembly methods for architected flexible electronics. Mainstream methods of mechanically-guided 3D assembly are classified and discussed on the basis of their fundamental deformation modes (i.e., rolling, folding, curving, and buckling). Diverse 3D interconnects and device forms are then summarized, which correspond to the two key components of an architected flexible electronic device. Afterward, structure-induced functionalities are highlighted to provide guidelines for function-driven structural designs of flexible electronics, followed by a collective summary of their resulting applications. Finally, conclusions and outlooks are given, covering routes to achieve extreme deformations and dimensions, inverse design methods, and encapsulation strategies of architected 3D flexible electronics, as well as perspectives on future applications.
1. Introduction
The pioneering work on bendable polymer transistors1 in the 1990s marked the arrival of the stretchable electronics era.2 Evolving from conventional rigid forms of electronic devices, stretchable electronics represent a new generation of device forms, featuring capabilities of withstanding extreme deformations involving high degrees of stretching, compression, bending, and twisting, while maintaining excellent electrical performances. Over the past two decades, advances in stretchable electronics have drastically transformed the landscape of today’s functional devices and led to the current prosperities of the field of flexible and stretchable electronics. Owing to their outstanding deformability, flexible and stretchable electronics have found numerous practical applications, spanning many different aspects of daily life and industry, such as long-term healthcare,3−5 diagnosis and therapeutics,6,7 sports protection8−10 and athletic analysis,11−13 nondestructive testing of large equipment,14 robotics,15−19 cosmetics,20,21 Internet of things,22,23 among others. Commercialization of flexible/stretchable electronics is also under way, represented by a series of startups, such as MC10 (Medidata), iRhythm, VitalConnect, Chero, LifeSignals, BioIntelliSense, Sibel Health, c3nano, and Sonica.
The specific terms “stretchable electronics” and “flexible electronics” were gradually nominated over the history. In this review, to avoid confusion, we define “stretchable electronics” as electronics devices that can not only withstand large degrees of bending and twisting but also high levels of stretch. “Flexible electronics” are defined as electronic devices with low bending/twisting rigidities, such that they can bear large degrees of bending and twisting, without specific restrictions on the stretchability. In general, two distinct routes have been exploited to develop flexible and stretchable electronics, including chemical engineering routes2,24−30 that concentrate to improve electrical performances of organic electronic materials, and structural engineering routes2,31,32 that focus on creating novel architectures of high-performance inorganic electronic materials to render the flexibility and stretchability. There are plenty of literature reviews2,32−35 addressing different aspects of stretchable and flexible electronics through both chemical and structural perspectives. Here, this review focuses on flexible/stretchable electronics achieved by means of structural engineering.
Structurally engineered flexible/stretchable electronics were originated from observations of the spontaneous formation of wavy metal thin films on elastomer substrates (i.e., polydimethylsiloxane (PDMS)).36 The development of stretchable single-crystal Si architectures31 (i.e., “wavy” Si) pioneered the structural engineering of flexible/stretchable electronics. Early flexible/stretchable electronics were mostly in the forms of wavy architectures and “island-bridge” designs.37−40 Serpentine designs were later introduced to offer an ultrahigh stretchability, without evidence reduction of the density of functional elements.41−45 Other structural design concepts, such as kirigami,46−52 spiral,53−61 and fractal designs,62−67 were also proposed to render high levels of stretchability for a limited prescribed space of the structural layout. Based on these design concepts, various planar flexible/stretchable electronic devices were fabricated, achieving applications in diverse areas, such as daily healthcare,3,11,12,68 robotic machines,69−75 multifunctional sensors,73−75 solar cells,76−78 and biomedical devices.6,7,79 Despite the significant progress achieved to endow high degrees of stretchability and flexibility in electronic devices, most of these devices are still restricted in planar forms, leaving the gap of geometric shapes and structural functionalities between flexible/stretchable electronics and natural species unfilled. Consequently, it is essential to develop 3D architected flexible electronics that can better conform to complexly shaped biological objects and/or mimic 3D structural forms of natural species to realize more unique functionalities than planar counterparts.
The appearance of 3D forms in flexible electronics naturally conformed to the logic of industrial upgrading, also marked by the establishment of many different 3D manufacturing approaches, including direct ones (e.g., printing techniques,80−84 photolithographic methods,85−87 molding88−90) and indirect ones (e.g., various mechanically-guided assembly methods91−107). Printing and photolithographic methods can form a wide range of 3D structures in almost arbitrary geometrical shapes, with feature dimensions ranging from centimeter scales down to micro- and nanoscales.83,84,87 However, it is challenging to simultaneously manufacture 3D architectures and integrated circuits using 3D printing and photolithographic methods because they both suffer from limitations of applicable materials (e.g., not applicable to crystalline semiconductors). While molding methods are not restricted to the accessible types of inks, their achievable structures are constrained to dissolvable or meltable polymeric and metallic materials. Additionally, multiple postmolding procedures are often required to fabricate 3D functional devices.80,88,90 Mechanically-guided 3D assembly methods bypass the difficulties faced by direct manufacturing approaches, through exploiting mechanics principles of materials and structures (e.g., residual stress,94,108−110 capillary force,111−114 actuation force,115−119 compressive buckling,103,104,120,121 and the rest119,122,123) to transform planar electronic devices fabricated using mature semiconductor techniques (Figure 1, central panel) into 3D architected ones. Such indirect manufacturing approaches feature broad applicable materials (including high-performance single-crystal semiconductors),103 postassembly deformability (e.g., structural multistability and shape morphing),16,17,123,124 and comparable dimensions with direct ones (i.e., from tens of centimeters down to few hundreds of nanometers).104,106,107 These characteristics make mechanically-guided assembly a promising route to 3D flexible electronics with a diversity of structure-induced functionalities (Figure 1, third circle), thereby enabling many previously hard-to-achieve applications (Figure 1, outer circle), including single-cell monitoring,125,126 growing biosensors,79 3D conformal electronics for organoid culture,127,128 photodetectors with spatial resolution, eyeball cameras,39,101,129,130 electronic fliers,22,131 robotics with extreme dimensions and multiple locomotion modes,16,17 among others.132−135 Although mechanically-guided assembly approaches do not offer the same high-level of geometrical complexity of resulting 3D architectures when comparing to direct manufacturing methods, significant efforts have been devoted to strengthen their assembly capabilities from different perspectives, spanning materials engineering (e.g., growth of heterogeneous materials),91,92,136,137 structural designs (e.g., precursors designs),16,103,138 controlled loadings (e.g., magnitudes, directions, applied regions)116,124,137,139 and so on.
Figure 1.
Overview of mechanically-guided 3D assembly methods for architected flexible electronics. Inner circle: mechanically-guided 3D methods (i.e., rolling, folding, curving induced, and bucking-guided assembly approaches) that can transform devices manufactured by mature planar semiconductor techniques into architected 3D structures; Middle circle: a variety of structure-induced functionalities of architected flexible electronics manufactured using mechanically-guided 3D assembly methods; Outer circle: applications of architected flexible electronics spanning biological devices, biomedical devices, electromagnetic devices, energy devices, optoelectronic devices, and robots.
The review provides a comprehensive overview of the field of mechanically-guided 3D assembly for architected flexible electronics. Different from the existing reviews on 3D assembly methods,35,104−107,140,141 the present review is constructed centering on a unique topic in the field–the structure-induced functionalities. We begin with summarizing existing methods of mechanically-guided 3D assembly based on their distinct deformation modes (Figure 1, second circle) (i.e., rolling, folding, curving, and buckling) (Section 2). Then, the 3D flexible electronic devices are dissected by two different classes of key structural components, and their various geometric forms are discussed separately in Section 3 (for interconnects) and Section 4 (for device forms), followed by their structure-induced functionalities (Section 5). A broad picture presenting diverse applications of 3D flexible electronics is provided in Section 6, covering biological devices, biomedical devices, electromagnetic devices, optoelectronic devices, energy devices, and robotics. Finally, conclusions and perspectives are given in Section 7, addressing current challenges and future directions of mechanically-guided 3D assembly methods and their resulting flexible electronics.
2. Mechanically-Guided 3D Assembly
Mechanically-guided 3D assembly methods exploit controlled mechanical deformations to transform planar electronic devices into 3D ones with various geometric configurations. Based on loading schemes and deformation characteristics, mechanically-guided 3D assembly methods are classified into rolling assembly,15−17,91−94,108,109,115−119,123,136,137,141−164 folding assembly,95−98,110−114,122,125,134,165−185 curving-induced assembly,39,99−102,129,130,135,186−208 and buckling-guided assembly.16,17,103,104,106,120,121,124,132,138,139,209−244 This section provides an overview of these four assembly approaches, emphasizing their distinct design principles.
2.1. Rolling Assembly
Through engineered strain mismatches in heterogeneous structures, rolling assembly approaches typically rely on bending deformations to form 3D architectures, such as tubes, helices, and others. Based on the mechanisms to generate strain mismatches, rolling assembly approaches are divided into two different subfamilies, i.e., residual-stress-induced rolling assembly and responsive-material-induced rolling assembly (Figure 2a–c). In this review, responsive materials denote materials that response to different types of physical (e.g., heating, illumination)16,17,109,117,118,162,163 and chemical (e.g., redox reactions) stimuli.119,123,158
Figure 2.
Assembly based on rolling deformations. (a,b) Representative residual stress-induced rolling of heteroepitaxial crystalline bilayers (a) and nonepitaxially deposited nanomembranes (b). (c) Self-actuated rolling through use of stimuli-responsive materials. (d) SEM image of a rolled SiGe nanotube (530 nm in diameter). Reproduced with permission from ref (91). Copyright 2001 Springer Nature. (e) SEM images of rolled SiGe/Si/Cr helices with different widths and orientations. Reproduced with permission from ref (137). Copyright 2006 American Chemical Society. (f) Rolled tubes of heterogeneously layered structures using different materials (i.e., Pt, Pd/Fe/Pd, TiO2, ZnO, Al2O3, SixNy, SixNy/Ag, diamond-like carbon, and SiO/SiO2). Reproduced with permission from ref (108). Copyright 2008 Wiley. (g) Complex 3D hydrogel “flowers” formed by swelling-induced rolling. Reproduced with permission from ref (115). Copyright 2016 Springer Nature. (h) Artificial cilia formed by residual stress-induced rolling. Reproduced with permission from ref (123). Copyright 2022 Springer Nature.
The residual-stress-induced rolling assembly is typically accomplished through releases of heteroepitaxial crystalline bilayers91,92,136,137 (Figure 2a), prestressed nonepitaxial nanomembranes108,162,163 (Figure 2b), and other layered heterogeneous structures.94Figure 2a illustrates the rolling of a typical heteroepitaxial bilayer crystalline film (consisting of materials 1 and 2 with different sets of lattice constants a1 and a2) by the use of photoresist as the sacrificial layer. In this case, the residual stress that rolls up the bilayer crystalline film roots in the different stress states (i.e., tension and compression in the different layers) caused by lattice mismatch (i.e., a1 > a2). Figure 2b shows the rolling assembly of a prestressed nonepitaxial nanomembrane through the use of sacrificial layer. Removal of the sacrificial layer releases the prestress and rolls up the nanomembrane. Responsive-material-induced rolling assembly approaches are achieved by rationally creating stress gradients in heterogeneous films composed of stimuli-responsive and nonresponsive materials (Figure 2c).
2.1.1. Residual-Stress-Induced Rolling Assembly
Using lattice mismatch-induced residual stresses, epitaxially grown heterogeneous membrane structures consisting of various crystalline materials can be rolled-up to form 3D tubular or helical structures in high precision.91,92,137,142 The geometrical parameters of the prepared tubular or helical microstructures (e.g., tube diameter, bending angle, helical pitch, helicity angle and etc.) can be controlled by changing processing parameters (e.g., etching time),91,92 tuning material properties (e.g., elastic modulus, lattice constants)136,137 and selecting different geometries of precursors.137 For example, through controlled release of a SiGe-based (Si/Ge bilayer) heterogeneous thin film (by adjusting the etching time of the sacrificial layer), a nanotube with diameter of 530 nm was formed along the [010] direction (Figure 2d).91 Harnessing the anisotropic stiffness of a InGaAs/GaAs bilayer, 3D helical structures with engineered pitches were prepared by controlling the angles between the stripes and their crystal orientation (i.e., [100] direction).144 Parameters (e.g., the orientation angles of planar precursors) that control the helicity angle, chirality, diameter, and pitch of resulting nanohelices were investigated.142Figure 2e showcases SiGe/Si/Cr helical nanoribbons with well-defined widths (from 1.30 μm to 0.70 nm) and orientation angles (i.e., between stripes and crystal orientation), which were fabricated using the residual-stress-induced rolling.137 Through controlled release of sacrificial layers, the rolling assembly can also be accomplished on nonepitaxially grown nanomembranes (e.g., with single or multiple layers).108,162 For instance, based on rolling assembly approaches, nanotubes with controlled diameters and lengths were prepared using nanomembranes consisting of a variety of functional materials, including Pt, Pd/Fe/Pd, TiO2, ZnO, Al2O3, SixNy, SixNy/Ag, diamond-like carbon, and SiO/SiO2 (Figure 2f).
Additionally, other strategies have also been reported to introduce residual stresses for the rolling assembly. For example, taking advantages of grain coalescence-induced capillary forces,93 the rolling assembly of tubular and arc-shaped microstructures was realized by melting patterned Sn films on top of the planar precursors. The release of mismatched tensile stresses in layered elastomeric composites was exploited to form arc or helical shapes through rolling.94 Recently, the engineered plastic strains during peeling processes were harnessed to allow controlled rolling of planar films into various 3D architectures.164
2.1.2. Responsive-Material-Induced Rolling Assembly
The second class of rolling assembly approaches mainly leverage engineered stress gradients created by the utility of specific characteristics of various stimuli-responsive materials, such as actuation of smart materials (e.g., hydrogels, liquid crystal elastomers, shape memory polymers/alloys),109,115−118 varied thermal expansion coefficients,245 material phase transitions,119,162,163 among others.123 Hydrogel-based composites are capable of forming complex 3D configurations (Figure 2g) by swelling-driven rolling assembly.115 By programming the actuation of thin liquid crystal elastomer (LCE) sheets through ordered alignments of mesogens, complex 3D shapes were demonstrated by rolling assembly.116 In addition to the above single-step rolling, a two-stage rolling assembly can also be achieved to allow developments of actuators capable of reversible 3D-to-3D deformations. For instance, utilizing U-shaped shape memory alloy (SMA) wires embedded in a bilayer elastomer with an engineered strain mismatch (i.e., one layer is with prestretch, the other is unstretched), 3D structures were fabricated, featuring a fast 3D-to-3D actuation capability.109,161 Particularly, the predetermined strain mismatch of the two elastomeric layers led to the first rolling assembly stage that formed an arc-shaped initial configuration of the actuator. Harnessing the shape memory effect of the embedded SMA wires, secondary reversible rolling was achieved to generate driving forces. A similar two-stage rolling assembly was also realized, taking advantage of the photothermally induced phase transition of VO2 nanomembranes.163 In detail, tubular VO2 nanomembranes with predefined geometries were first assembled using residual-stress-induced rolling. Then, the photothermally induced phase transition led to the second order of reversible rolling assembly, featuring precisely controlled curvature regulation. An active metasurface composed of artificial cilia (i.e., arc-shaped Pt/Ti bilayer ribbons, Figure 2h)123 was fabricated following the same sequence of two-stage rolling assembly but enabled by electrochemical redox reactions of Pt/PtOx.
2.2. Folding Assembly
Folding assembly approaches exploit localized bending deformations to fold planar precursors with predefined creases into various 3D geometries, such as polyhedral or origami structures. The forces that lead to folding can be classified as external forces111−114,176 (e.g., capillary forces) and internal forces110,122,167,174,178 (e.g., forces induced by engineered strain mismatch). The external-capillary-force-induced folding relies on the surface tension of liquid materials such as melted metals and droplets to drive the assembly process (Figure 3a). Meanwhile, internal forces can be introduced by using prestressed nanomembranes or responsive materials, as demonstrated in Figure 3b,c. This section reviews folding assembly methods through capillary-force-induced, residual-stress-induced, and responsive-material-induced routes.
Figure 3.
Assembly based on folding deformations. (a–c) Representative folding strategies using capillary forces, residual stresses, and stimuli-responsive materials. (d) Optical images of folded PDMS pyramids, cubes, and quasispheres formed using capillary forces. Reproduced with permission from ref (111). Copyright 2007 Springer AIP Publishing. (e) Folded 3D polyhedral nanostructures by the reflow of Sn within hinges. Reproduced with permission from ref (113). Copyright 2009 American Chemical Society. (f) Folding based on residual stresses introduced during focused ion beam (FIB) irradiation. Reproduced with permission from ref (177). Copyright 2015 Springer Nature. (g) Reversible folding of a multilayered origami structure in hydrogels. Reproduced with permission from ref (178). Copyright 2014 Wiley. (h) Magnetically controlled folding of a flapping “bird”. Reproduced with permission from ref (183). Copyright 2019 Springer Nature. (i) Miura-origami microstructure formed through electrochemical actuation. Reproduced with permission from ref (122). Copyright 2021 The American Association for the Advancement of Science.
2.2.1. Capillary-Force-Induced Folding Assembly
The most straightforward way to leverage capillary forces is by using droplets. For example, pyramids, cubes, and quasi-spheres were formed using PDMS precursors with predesigned triangular, crossing, or flower-like shapes (Figure 3d).111,112 Particularly, the area reduction of the liquid–air interface via evaporation indicates the consumption of total surface energy of the original droplet. Apart from the dissipated energy (e.g., in forms of heat), the rest of the consumed surface energy transfers to the elastic strain energy of the patterned precursors, allowing the controlled folding of these structures into 3D configurations. Analytical models that capture mechanisms of folding assembly approaches through capillary forces were established.112,168 Reversible folding assembly routes were later developed to form deformable 3D structures.176
In addition to droplets, building meltable hinges using materials with relatively low melting points (e.g., metallic materials such as In and Sn) can also accomplish folding in a well-controlled manner.113,114 Because of the compatibility with mature planar microfabrication methods, the folding of 3D structures can be implemented at nanoscale. For instance, 3D folding microcontainers capable of chemical encapsulation, guided delivery, and spatially controlled chemical reactions were demonstrated by using melting of the patterned Sn solder hinges.114 Similarly, Figure 3e demonstrates the folding assembly of stable 3D polyhedral nanostructures driven by the reflow of Sn within the hinges.113
2.2.2. Residual-Stress-Induced Folding Assembly
Advanced processing techniques of thin films can accurately confine residual stresses at hinged regions to enable controlled folding deformations. Commonly used approaches for this purpose include ion-beam irradiation,177,246 as well as heteroepitaxial and nonepitaxial deposition processes.95,97,98,167
Focused ion beams (FIBs) can create voids at hinges, and, then, fill them with energized ions to simultaneously release compressive stresses and fold planar structures into many desired 3D configurations (Figure 3f).177 Through heteroepitaxial thin-film deposition, complex 3D micro-origami structures were demonstrated.95 In addition, harnessing residual stress during nonepitaxial growth of Cr layer, 3D microgrippers that could be thermally or chemically actuated were fabricated using the folding assembly.167
2.2.3. Responsive-Material-Induced Folding Assembly
Folding assembly using responsive materials can form 3D architectures with high levels of geometric complexity. For example, an origami “bird” structure was assembled using trilayer polymer films with different degrees of photo-cross-linking, capable of swelling-induced reversible folding (Figure 3g).178 Heterogenous layered structures composed of shape memory polymers (SMPs), paper, and resistive circuits, were exploited to form thin film-shaped flexible actuators (TFFAs).174 Harnessing the folding assembly of these TFFAs, a facile fabrication route to shape-shifting crawling robots with complex 3D configurations was established.175 By introducing magnetically controlled single-domain nanomagnet arrays in the folding assembly process, an origami “bird” that could flap its wings under an out-of-plane magnetic field was fabricated (Figure 3h).183 Through the use of residual-stress-induced folding and the introduction of electrochemically responsive Pt/Ti hinges, 3D micro-origami structures capable of reversible deformation were prepared and demonstrated as actuators (Figure 3i).122
2.3. Curving-Induced Assembly
Curving-induced assembly approaches can allow formation of 3D flexible electronics that conform to arbitrary curvilinear surfaces.102 Curving-induced assembly relies on various transfer printing techniques to integrate planar electronic devices with complex 3D curved surfaces.106 This section reviews two typical transfer printing strategies for curving-induced assembly, including planar39,99 and nonplanar strategies.195,196,201,205,207,247,248
2.3.1. Curving-Induced Assembly Based on Planar Transfer Printing
The curving-induced assembly39 based on planar transfer printing (Figure 4a) involves multiple steps that include: (i) fabrication of a 3D intermediate elastomer substrate (i.e., hemispherical PDMS substrate) and the planar device (i.e., focal array); (ii) flattening the as-prepared 3D intermediate substrate followed by the transfer printing of the as-fabricated device onto the flattened substrate; and (iii) relaxation of the substrate followed by the transfer printing of the curved device from substrate to the target curved surface. Electronic devices that conform to complex curved surfaces, such as golf balls, pyramidal substrates, convex paraboloid substrates, and even human heart models were prepared through this type of curving-induced assembly.99
Figure 4.
Curving-induced assembly using transfer printing techniques. (a) Assembly of a hemispherical silicon focal array through the use of a biaxial prestretch that flattens the substrate to facilitate the transfer printing. Reproduced with permission from ref (39). Copyright 2008 Springer Nature. (b) Assembly of electrical circuits on wavy substrates via hydro-printing. Reproduced with permission from ref (196). Copyright 2017 Wiley. (c) Vacuum-assisted transfer printing strategy. Reproduced with permission from ref (186). Copyright 2008 Elsevier. (d) Transfer printing using pneumatically inflated elastomeric balloons as conformal stamps. Reproduced with permission from ref (201). Copyright 2019 Springer Nature. (e) Optical images of the serpentine metal mesh on the pyramid surface and the convex-shaped kirigami imager. Reproduced with permission from ref (201). Copyright 2019 Springer Nature. Reproduced with permission from ref (130). Copyright 2021 Springer Nature. (f) Assembly of conformal 3D electronics through thermoforming of a patterned 2D precursor structure. Reproduced with permission from ref (207) under CC BY. Copyright 2021 The American Association for the Advancement of Science. (g) Microscale transfer printing on curved substrates using reflowable stamps. Reproduced with permission from ref (203) under CC BY. Copyright 2022 The American Association for the Advancement of Science.
2.3.2. Curving-Induced Assembly Based on Nonplanar Transfer Printing
Various nonplanar transfer printing techniques, such as hydroprinting,195,196 vacuum-assisted transfer printing,186 transfer printing using specifically architected 3D flexible stamps190,201 (e.g., hierarchical stamps and pneumatically inflatable balloon stamps) have been developed to enable curving-induced assembly. These techniques feature the direct formation of conformable electronics on complex curved 3D substrates.
Hydroprinting (Figure 4b (i–iv)) contains four steps: (i) printing patterned electrical circuits on a water-soluble substrate, dissolving the substrate to release the circuits, and floating them on the air–water interface; (ii) letting the curved target substrate approach the floating circuits on the interface; (iii) immersing the target substrate to form strong conformal contact with flexible electronic circuits; and (iv) flipping and drying the substrate with integrated circuits.196 Vacuum-assisted transfer printing (Figure 4c) leverages low rigidities of PDMS stamps to form conformal contact with the target substrate.186 Hierarchical perfluoropolyether stamps capable of adaptive deformations were adopted to allow curving-induced assembly of 3D conformal electronic devices on curved substrate with micropatterns.190 By the use of a pneumatically inflated elastomeric balloon as a conformal stamp for both pickup and delivery (Figure 4d), nonplanar transfer printing of flexible electrical circuits can also be accomplished.201 The high level of deformability of the above balloon stamp allows the assembly of flexible electronics on many complex 3D surfaces, such as the pyramid surface shown in the top panel of Figure 4e.201 The concept of kirigami was also introduced to improve the conformability of planar electronic circuits to nondevelopable curvy surfaces.130,206,208 For example, an array of ultrathin silicon optoelectronic pixels with kirigami design was formed on a spherical cap by curving-induced assembly (Figure 4e, bottom panel).130 Recently, a microscale transfer printing approach (Figure 4f) was developed using reflowable materials (e.g., sugar mixture) that can stretch to conform to surfaces with complex topographies and small curvature radii.207 Apart from the above, nonplanar direct printing method harnessing thermosplasticity was employed to assemble liquid metal circuits on irregular 3D surfaces (Figure 4g).203
In summary, the curving assembly methods based on nonplanar transfer printing are more straightforward and effective, when compared to those via planar transfer printing techniques. However, the accurate positioning and alignment of electronic components with respect to the target substrates often stand as technical issues hindering their widespread practical applications, especially on small scales. In particular, the hydroprinting technique features fast and low-cost manufacturing, but the process can be completed only under aqueous environments. The nonplanar transfer printing through the use of elastomeric stamps (e.g., PDMS) stands as the most common and facile method. However, the limited deformability of bulky stamps hinders their application in the fabrication of electronic devices on surfaces with large curvatures as well as the preparation of devices consisting of fragile elements. When comparing with bulky stamps, the flexible balloon stamps present better deformability, which can be utilized to prepare electronic devices on surfaces with complex curvatures. Transfer printing techniques taking advantages of reflowable materials are capable of fabrication of micro- to nanoscale devices on curved complex surfaces. However, the underlying mechanisms of the associated interface mechanics (e.g., adhesion and delamination) need to be better understood. The nonplanar transfer printing based on thermoplasticity features low-cost, rapid fabrication. However, due to the requirement of the heating process, it might not be ideal for the manufacturing of electronic devices consisting of temperature sensitive elements.
2.4. Buckling-Guided Assembly
Buckling-guided assembly approaches103,106 are usually accomplished by sequential processes that include (i) design and preparation of 2D precursors with predefined geometries, (ii) selective bonding (e.g., typically covalent bonding) of 2D precursors onto a prestrained elastomer substrate with controlled loading magnitudes, and (iii) release the prestrain of the substrate to finish the assembly. The detailed schematic illustrations of an exemplary buckling-guided assembly procedure are shown in Figure 5a. Buckling-guided assembly methods allow access to a wide range of complex 3D geometries with dimensions from tens of centimeters down to a few hundreds of nanometers. The excellent compatibility with well-established planar manufacturing techniques endows the applicability to a broad family of functional materials, spanning both organic and inorganic materials, such as polymers (e.g., polyimide (PI), PDMS, SU-8, polyvinylidene fluoride (PVDF), SMP and etc.), metals (e.g., Cu, Au, Ag, SMA and etc.), semiconductors (e.g., Si, Ge, GaAs), ceramics (e.g., lead zirconium titanate (PZT)) and other electronic materials (e.g., graphene, carbon nanotube (CN) and etc.). This section reviews buckling-guided assembly approaches from perspectives that control such a process, including 2D precursor designs, substrate designs, loading-path-based strategies, schemes for interface control, and strategies for freestanding 3D mesostructures.
Figure 5.
Buckling-guided assembly: schematic illustrations, 2D precursor designs, and substrate designs. (a) Schematic illustrations of the process of buckling-guided assembly. (b) Typical 3D mesostructures formed using various 2D precursor designs (filamentary, kirigami, origami, multilayer, and microlattice designs). Reproduced with permission from ref (236). Copyright 2021 The American Association for the Advancement of Science. Reproduced with permission from ref (120). Copyright 2015 National Academy of Sciences. Reproduced with permission from ref (210). Copyright 2016 Wiley. Reproduced with permission from ref (121) under CC BY. Copyright 2016 The American Association for the Advancement of Science. Reproduced with permission from ref (138). Copyright 2023 The American Association for the Advancement of Science. (c) 3D mesostructures with a diversity of composing materials (e.g., doped silicon, graphene, PVDF and magnetic PDMS elastomer). Reproduced with permission from ref (249). Copyright 2018 American Chemical Society. Reproduced with permission from ref (233). Copyright 2020 Wiley. Reproduced with permission from ref (132). Copyright 2019 Springer Nature. Reproduced with permission from ref (237). Copyright 2021 Wiley. (d) Typical substrate designs used in the buckling-guided assembly, including substrates with engineered thickness/modulus, kirigami substrates, and curved substrates. Reproduced with permission from ref (212). Copyright 2017 Wiley. Reproduced with permission from ref (224). Copyright 2019 American Chemical Society. Reproduced with permission from ref (232). Copyright 2019 National Academy of Sciences. Reproduced with permission from ref (240) under CC BY. Copyright 2022 The American Association for the Advancement of Science. (e) Assembly based on substrates composed of different active materials (e.g., hydrogel, LCE, DE, and SMP). Reproduced with permission from ref (230). Copyright 2019 Wiley. Reproduced with permission from ref (234). Copyright 2021 American Chemical Society. Reproduced with permission from ref (225). Copyright 2019 Oxford University Press. Reproduced with permission from ref (226). Copyright 2019 Wiley.
2.4.1. 2D Precursor Designs
Various 2D precursor designs, involving filamentary,103,209,223,236 kirigami,120 origami,210,213 multilayer,121 and microlattice designs,138 were developed to enable formation of 3D architectures with high degrees of complexities and diversities (Figure 5b). Filamentary designs (Figure 5b, filamentary)236 typically consist of slender thin ribbons to ensure that large out-of-plane bending and twisting deformations can be induced to realize their transformation into desired 3D configurations such as helices209 and frameworks.223 Kirigami designs (Figure 5b, kirigami) feature the reduction of stress concentration during buckling-guided assembly by introducing strategic cuts, thereby enabling access to diverse 3D membrane mesostructures (e.g., bionic “tiger”).120 Origami designs (Figure 5b, origami) usually harness engineered folding creases in 2D precursors to form various 3D mesostructures by buckling-guided assembly.210 Multilayer designs are realized by the use of patterned 2D precursors with predefined multilayer layouts, rendering densely distributed complex 3D multilayer configurations (Figure 5b, multilayer).121 Recently, inspired by cellular biological surfaces, a rational microlattice design strategy was developed as a powerful route to achieve desired stiffness distribution of 2D microfilms, thereby allowing their transformation into programmable 3D curved mesosurfaces through the buckling-guided assembly.138 A wide range of materials were incorporated with the above divergent 2D precursor designs, enabling the manufacturing of many different architected multifunctional flexible electronic devices (Figure 5c).132,233,237,249
2.4.2. Substrate Designs
Various substrate design strategies, including substrates with engineered thickness212/modulus,224 kirigami designs,232 substrates with curved 3D configurations (i.e., for ordered buckling-guided assembly)240 and hierarchical designs,241 are devised to enable different forms of forces applied to the 2D precursor, thereby enriching the family of accessible 3D architectures through the buckling-guided assembly. Using substrates with engineered distributions of stiffness (e.g., Figure 5d, first and second column, by changing thickness212 and modulus,224 respectively), varied compressive strains are applied to different local regions of the 2D precursor. As such, the spatially varying 2D-to-3D transformation can be realized, leading to the formation of 3D mesostructures with a desired gradient of the height (i.e., the dimension along the out-of-plane direction). Substrates with kirigami designs allow large rotational motions, thus, enabling additional local twisting deformations of 2D precursor structures (Figure 5d, third column).232 By the use of substrates with curved 3D configurations, an ordered buckling-guided assembly can be accomplished, allowing transformation of 2D precursors into sophisticated 3D structures on diverse curved surfaces, such as the microscale helical networks assembled on a brain-like surface (Figure 5d, rightmost).240 The hierarchical design strategy of the substrate renders the buckling-guided assembly of not only the 2D precursors but also the secondary layer of the substrate during the same process, allowing the creation of 3D mesostructures mounted at multiple-level 3D frameworks with complex configurations.241
Additionally, the use of responsive materials in the substrate allows more flexible loading forms for the buckling-guided assembly, when comparing to substrates consisting of conventional elastomeric materials (Figure 5e).234 For example, the use of thermally responsive hydrogels230 or LCEs234,235 can enable thermal-mechanically controlled buckling-guided assembly in a reversible manner. Dielectric elastomers were also used as substrates for electro-mechanically controlled buckling-guided assembly, achieving sequential and local loading with desired strain distributions.225 Furthermore, by the use of SMP substrates, the buckling-guided assembly can be exploited to form 3D structures with certain level of reconfigurability.226
2.4.3. Loading-Path-Based Strategies
By controlling loading paths and forms, the deformation modes of patterned 2D precursors can be enriched during the buckling-guided assembly to enhance the geometric diversity of the resulting 3D architectures. As an example, the tensile loading simplifies the process of buckling-guided assembly, and extends the accessible range of 3D topologies by avoiding the prestretch of elastomeric substrates (Figure 6a, flexible LED array assembled by tensile buckling).216 By selecting the time sequences or release path of biaxial prestrain in the substrate, specially engineered 2D precursors can be assembled into different stable 3D configurations, thereby allowing the formation of reconfigurable 3D mesostructures.124,139,250 For instance, as shown in Figure 6b (left), the simultaneous release of equal 100% biaxial strain results in a “pop-up” structure of Shape I, while the sequential release of biaxial strain (i.e., first release the prestrain in x-axis, followed by the y-axis) reshapes the 2D cross-ribbon structure into a “pop-down” structure of Shape II.124 Based on this concept, many reconfigurable flexible devices, such as 3D radiofrequency (RF) circuits, can be formed by the buckling-guided assembly (Figure 6b (right)). Taking advantages of such sequential and directional control of loading paths, a bottom-up design route to geometrically reconfigurable 3D mesostructures was established using ribbon-formed components as building blocks.139 Furthermore, by buckling-guided assembly with ordered loading paths, various 3D mesostructures are formed on complex curved surfaces. Taking the ordered loading240 using helicoid substrates as an example (Figure 6c), the substrate was first flattened by applying torsional loadings (720°), followed by an additional uniaxial prestretch (20%). After selective bonding of the 2D precursor, release of the prestretch (20% uniaxial prestrech) enabled the first-order assembly of leaf-like 3D structures on both sides of the flattened substrate, and further release of the torsion allowed transformation of the substrate into the helicoid surface, resulting in the spatial rearrangement of two leaf-like 3D structures during the second-order assembly.
Figure 6.
Buckling-guided assembly: loading-path-based strategies and schemes for interface control. (a) Tensile buckling strategy for the assembly of stretchable LED arrays. Reproduced with permission from ref (216). Copyright 2018 Springer Nature. (b) Loading-path controlled strategy for the assembly of bistable 3D mesostructures and radiofrequency (RF) circuits. Reproduced with permission from ref (124). Copyright 2018 Springer Nature. (c) Ordered assembly process of 3D leaf-like structures on a helicoidal substrate with FEA predictions and an optical image. Reproduced with permission from ref (240) under CC BY. Copyright 2022 The American Association for the Advancement of Science. (d) Interface weakening through use of sacrificial layers. Reproduced with permission from ref (210). Copyright 2016 Wiley. (e) Wrinkling-assisted strategy for controlled interface delamination. Reproduced with permission from ref (244). Copyright 2023 Elsevier. (f) Interface control through use of elastomer substrates with microstructured surfaces. Reproduced with permission from ref (242) under CC BY. Copyright 2023 Springer Nature. (g) Electroadhesion-mediated strategy for controlled interface adhesion. Reproduced with permission from ref (243). Copyright 2023 ASME INTERNATIONAL.
2.4.4. Schemes for Interface Control
The buckling-guided assembly approach requires simultaneous and selective formations of weak interfaces that ensure full delamination of desired parts of precursors and strong interfaces that firmly bond the selective sites of precursor on the substrate (i.e., bonding sites).222 At the macroscale, the structural stiffnesses of 2D precursors are usually sufficient to drive the delamination during buckling-guided assembly, and therefore, the interface control strategies are focused on building strong interfaces that bond the precursor with the substrate. Differently, at the micro- to nanoscale, the unavoidable influences by van der Waals forces and drastically reduced structural stiffnesses make the delamination very difficult. In this case, engineered weak interfaces are necessary to ensure a successful assembly process.
Strategies relying on the use of sacrificial layers to create gaps between 2D precursors and substrates are promising to build weak interfaces, rendering the successful assembly of an origami structure (i.e., with relatively large surface contact) by the buckling-guided approach (Figure 6d).210 In addition, the use of assisting layers to introduce initial disturbance to the buckling-guided assembly can also effectively facilitate the delamination of precursors.132,244 For instance, a patterned PI ribbon was adopted as an underlying supporting layer for the assembly of thin PVDF structures with ultralow stiffnesses (e.g., serpentines).132 Recently, a wrinkling-assisted interface control strategy was developed to evidently facilitate the delamination at desired regions of the film/substrate system.244 In particular, an additional assisting layer was exploited such that the original film–substrate interface was replaced by a weaker film-assisting-layer interface, and the formed wrinkles in the assisting layer induce additional driving forces to separate the film-assisting-layer interface (Figure 6e). Besides, the microtexturing of substrate surfaces is found effective to control the interface during the buckling-guided assembly.231,242 As shown in Figure 6f, by the use of substrates with patterned voids, the direction (i.e., pop-up and pop-down), magnitude, and mode of buckling deformations can be accurately controlled by varying the pillar strain (εpillar), trench strain (εtrench), and boundary constraints (Wpillar).242
Apart from creating weak interfaces to facilitate delamination, building strong interfaces in a controlled manner was also explored to enrich the obtainable 3D architectures through buckling-guided assembly. For example, the electroadhesion-mediated strategy was developed to achieve controlled adhesion by varying the applied voltages during the buckling-guided assembly, giving rise to various reconfigurable 3D mesostructures (Figure 6g).243
2.4.5. Strategies for Freestanding 3D Mesostructures
Based on the buckling-guided assembly, several approaches have been proposed to achieve freestanding 3D mesostructures after separation from the substrate, including the use of photo-cross-linkable bases,214 introducing plastic deformations,213,239 creating portable mechanical constraints,17,138,227 and exploiting materials with shape memory effects.228,235 For example, by drop casting of SU-8 and photo-cross-linking from the backside, a freestanding 3D mesostructure with a very thin SU-8 base of a similar size can be prepared (Figure 7a).214 Plastic deformations of metallic213,214 and thermoplastic materials239 were also employed to maintain the assembled 3D configurations. For example, the plastic deformations of the copper layer were utilized to maintain the 3D configuration of a cubic structure after the release from the substrate, as shown in Figure 7b.214 In addition, mechanical constraints were also utilized to fix the assembled 3D configurations after removing the substrate, such as building mechanical interlocks227 (e.g., elements composed of lugs and hooks) or creating stiffer conformal encapsulating shells17 (e.g., SiO2 in Figure 7c). Furthermore, materials with shape memory effects, such as SMPs228 and LCEs,234,235 can also be exploited to fix the as-assembled configurations and form freestanding 3D mesostructures, as demonstrated in Figure 7d.
Figure 7.
Buckling-guided assembly: strategies for freestanding 3D mesostructures. (a) Process of forming freestanding 3D mesostructures by the use of the SU-8 base. Reproduced with permission from ref (214) under CC BY. Copyright 2017 National Academy of Sciences. (b) 3D shape fixation enabled by controlled plasticity of the metal layer. Reproduced with permission from ref (214) under CC BY. Copyright 2017 National Academy of Sciences. (c) Peekytoe crab-like freestanding 3D mesostructure formed by introducing a conformal SiO2 shell after the assembly. Reproduced with permission from ref (17). Copyright 2022 The American Association for the Advancement of Science. (d) SMP-based freestanding 3D electronic devices. Reproduced with permission from ref (228). Copyright 2018 Wiley.
By means of rolling assembly, folding assembly, curving-induced assembly, and buckling-guided assembly, 3D architectures with numerous geometric configurations, such as tubular, helical, polyhedral, origami, hemispherical, irregularly curved, and complex forms, can be manufactured using various functional materials ranging from polymers, metals, high-performance single-crystal semiconductors, to 2D materials and other electronic materials. A detailed comparison table (Table 1) is presented, summarizing typical geometries, achievable scales and reversibility of the four main mechanically-guided assembly methods. In particular, rolling assembly methods are usually exploited to prepare tubular or helical 3D configurations, with the accessible curvature radius down to sub-10 nm.136 The rolling process can be either reversible or irreversible, depending on the applied mechanisms/materials (e.g., the use of hydrogels, LCEs, and other stimuli-responsive materials could result in reversible rolling). The typical geometries achievable by the use of folding assembly methods include polyhedron and other origami forms, and the minimum structural dimension of such folded configurations (i.e., in the range of hundreds of nanometers114) is slightly higher than the rolled ones, due to their higher levels of structural complexities. Curving-induced assembly usually requires the use of predesigned substrates to provide a base for the deformable device precursor to form desired 3D configurations through various transfer printing techniques, and its minimum achievable structural dimension could be tens of micrometers.207 The buckling-guided assembly methods require the use of elastomer substrates to transform planar device precursors with well-designed geometries into a broad family of complex 3D configurations. Such methods are highly scalable, and can be exploited to fabricate 3D architected devices ranging from submicrometers to tens of centimeters.221,242 To summarize, these mechanically-guided assembly methods have clearly outlined the routes to the constructions of architected flexible electronic devices, by first building the structures of devices (i.e., interconnects and main bodies) and second incorporating structural functionalities.
Table 1. Comparison of Four Main Types of Mechanically-Guided 3D Assembly Methods.
Assembly methods | Typical geometries | Achievable minimum scales | Reversibility | Reference |
---|---|---|---|---|
Rolling assembly | Tubes, helices, etc. | Sub-10 nm | Irreversible/reversible | (91, 136, 156, 162, 163) |
Folding assembly | Polyhedron, origami forms, etc. | Hundreds of nanometers | Irreversible/reversible | (112, 114, 122, 171) |
Curving-induced assembly | Curvilinear forms | Tens of micrometers | Irreversible | (39, 99, 102, 130, 207) |
Buckling-guided assembly | A wide range of complex 3D geometries | A few micrometers | Reversible | (103, 104, 106, 221, 242) |
3. 3D Interconnects
Conventional interconnects mainly use inorganic electronic materials, such as metals (e.g., Au, Ag, and Cu) and semiconductors (e.g., Si), that do not stretch much (i.e., the fracture strain of silicon is only ca. 1%).2 In order to build architected flexible electronic devices, the conductivity and stretchability of the exploited interconnects are required simultaneously. Therefore, constructing interconnects with flexible structures using high performance inorganic electronic materials is complementary. Thinning membranes to extreme makes brittle inorganic materials bendable.251 For example, the Si nanomembrane with a thickness down to ca. 2 nm exhibits flexural rigidities far smaller (e.g., by 15 orders of magnitude) than those of bulk wafers (with thickness of ca. 200 μm). Based on this principle, 3D interconnects in the forms of arc-shaped,31,37−40,99,251−276 serpentine,40,277,278 and helical103,209,279−282 are designed and fabricated to provide high levels of stretchability, flexibility, and conformability that fulfill the requirements of 3D architected flexible electronic devices.
3.1. 3D Arc-Shaped Interconnects
This section reviews 3D arc-shaped interconnects based on their different structural designs, including wavy designs, island-bridge designs, and other designs.
3.1.1. Wavy Designs
Given that there are plenty of reviews252,264,265,268 addressing wavy designs in forms surface wrinkles (Figure 8a, shape I), here, we review only the wavy designs in forms of out-of-plane 3D configurations (Figure 8a, shape II).38 Such wavy designs formed by the buckling-guided assembly (Figure 8b,c) represent early forms of stretchable electronics consisting of rigid functional components.37 Analytical models of such wavy structures have been developed to predict the buckling configuration and peak strain during the assembly.37
Figure 8.
3D arc-shaped interconnects. (a) Schematic illustration of the fabrication strategy of wavy structures based on the buckling-guided assembly. Reproduced with permission from ref (38). Copyright 2006 Wiley. (b) SEM image of a wavy structure array. Reproduced with permission from ref (37). Copyright 2006 Springer Nature. (c) Schematic illustration of the process for fabricating wavy structures on PDMS substrates (Top) and GaAs ribbons formed on PDMS substrates with different prestrains and different widths (Win) of the inactivated region (Bottom). Reproduced with permission from ref (37). Copyright 2006 Springer Nature. (d) Schematic illustrations of the representative fabrication process for island-bridge structures on an elastomeric substrate via buckling-guided assembly. Reproduced with permission from ref (263). Copyright 2009 AIP Publishing. (e) Deformed configuration of the island-bridge structure under bending. Reproduced with permission from ref (40). Copyright 2008 National Academy of Sciences. (f) Island-bridge structures on a curved surface. Reproduced with permission from ref (99). Copyright 2009 Wiley. (g) Flexible electronics consisting of arc-shaped liquid metal interconnects. The left panel shows a flexible LED array with arc-shaped interconnects, and the right panel illustrates the preparation process. Reproduced with permission from ref (270). Copyright 2023 Springer Nature.
The buckling profile can be expressed by a sinusoidal function as
![]() |
1 |
where w is the deflection along the y axis, A is the buckling amplitude to be determined, x1 is the ribbon direction, 2L1 = Win/(1 + εpre) is the buckling wavelength, 2L2 = Win /(1 + εpre) + Wact is the sum of activated (with strong chemical bonds) and inactivated (with unmodified surface chemistry) regions after relaxation, and εpre is the prestrain of the elastomeric substrate (Figure 8c).
Minimization of the total energy (Utot) consisting of bending energy and membrane energy with respect to the buckling amplitude A gives
![]() |
2 |
where is the critical strain for buckling, in
which h denotes the film thickness. The experimental
and theoretical profiles of buckled GaAs thin films formed on the
PDMS substrate with different εpre and Win values agree well (Figure 8c). The peak strain in the buckled thin film
is
, which is typically much smaller than the
prestrain.
3.1.2. Island-Bridge Designs
Evolving from wavy structures, island-bridge designs were later developed and widely used for flexible electronics.39,40,256,263,274,275 Similar to arc-shaped wavy designs, arc-shaped island-bridge designs are also formed through the buckling-guided assembly. Figure 8d illustrates the fabrication process of a typical island-bridge structure, where discrete islands (e.g., containing rigid functional components) adhered to an elastomer substrate are connected by stretchable bridges (e.g., electrically conductive interconnects).263 Notably, owing to low levels of strain in islands (i.e., mechanical strain isolation), such designs can simultaneously provide high stretchability and excellent protection of functional electronic components. Therefore, flexible electronic devices using island-bridge designs can withstand high-degrees of bending and twisting deformations (Figure 8e),40 and conform to curved 3D surfaces such as hemispheres (Figure 8f).99
Many theoretical models were developed to guide the designs of arc-shaped island-bridge interconnects in 3D forms.256,263,274,275 For example, a mechanical model to predict the buckling configurations was developed by assuming the ribbon as a clamped beam of sinusoidal form.263 However, this model is not accurate enough in the regime of large compressions (e.g., compressive strain >40%). To improve the accuracy, a finite deformation model was later established, enabling precise predictions of deflections and peak strains in island-bridge designs.274 Additionally, the mechanics of island-bridge interconnects in curvilinear electronic devices were also investigated to allow a rational design.256,275
Arc-shaped interconnects based on the curving-induced assembly were also proposed.270 In particular, by harnessing the fluidity and high conductivity of liquid metal alloys, stretchable 3D interconnects were prepared for uses in electronic devices. As shown in Figure 8g, exploiting the solid–liquid phase transition and plastic deformations of the liquid metal alloy (i.e., Ga-based liquid metal alloys), 3D arc-shaped interconnects were fabricated through the curving-induced assembly.270
3.2. 3D Serpentine Interconnects
Compared to arc-shaped 3D interconnects, serpentine interconnects in 3D forms exhibit higher levels of stretchability.40,278 As shown in Figure 9a, a stretchability of 70% was achieved in flexible electronic devices by using 3D serpentine interconnects formed by the buckling-guided assembly. In particular, as shown in Figure 9b, for a system assembled with 35% prestrain, when the uniaxial stretching gradually increased to 40%, the 3D serpentine interconnects almost returned to their planar configurations. Through in-plane rotation and out-of-plane buckling of 3D serpentine interconnects, the whole system can accommodate up to 70% tensile strain without failure. By tuning the assembly parameters (e.g., prestrains) and geometric layouts (e.g., the number of arcs, Figure 9c), the stretchability of 3D serpentine interconnects can be further improved.278
Figure 9.
3D serpentine interconnects. (a) SEM image of a stretchable CMOS inverter array using 3D serpentine interconnects. Reproduced with permission from ref (40). Copyright 2008 National Academy of Sciences. (b) Optical images of 3D serpentine interconnects under stretch. Reproduced with permission from ref (40). Copyright 2008 National Academy of Sciences. (c) Optical images and FEA results of deformed configurations of 3D serpentine interconnects under uniaxial stretching. The color in the FEA results denotes the magnitude of the maximum principal strain. Reproduced with permission from ref (278). Copyright 2009 Wiley.
3.3. 3D Helical Interconnects
Interconnects with 3D helical designs, commonly generated through the buckling-guided assembly of 2D filamentary serpentine precursors,103,209,279,280,282 present higher elastic stretchability, when comparing to arc-shaped 3D interconnects and 3D serpentine interconnects. This section highlights the design principles and encapsulation strategies of the 3D helical interconnects.
As shown in Figure 10a, by varying the geometries of 2D precursors and distributions of bonding sites, divergent helical interconnects can be fabricated, including single helices, dual helices, nested coaxial structures, among others.103 The average curvature components and mode ratio (κtwist/κbend) can be calculated using finite element analyses (FEA), as shown in Figure 10a (right) for a typical 3D single helix. To understand the influences of the prestrain and geometric designs of 2D precursors on the resulting 3D helical configurations, an analytical model of compressive buckling based on the energy method was established. This model gives accurate predictions of deformed configurations and maximum strains of a single 3D helix (Figure 10b, left), which agreed well with FEA and experimental results.209 In particular, when a helical interconnect is uniaxially stretched (εstretch ≤ εpre), the out-of-plane displacement (U1), the curvature components (κ̂2 and κ̂3), as well as the maximum principal strain (εM) follow a similar square root scaling
![]() |
3 |
Figure 10.
3D helical interconnects. (a) Helical interconnects with three typical precursor designs. The left panel shows SEM images of single-helix, dual helix, and nested helix designs with their corresponding FEA predictions. The right panel demonstrates the dependences of average curvature components and the mode ratio of a 3D helical interconnect on the prestrain. Reproduced with permission from ref (103). Copyright 2015 The American Association for the Advancement of Science. (b) Illustration of the mechanics model for the buckling-guided assembly of helical interconnects, along with the predictions and FEA results on the distribution of dimensionless pop-up displacement and twist angle for a typical helical mesostructure (with an arc angle of θ0 = 150° for the 2D precursor). Reproduced with permission from ref (209). Copyright 2016 Wiley. (c) Assembled 3D helical coils: optical images (left), deformations and Mises-stress distributions of a 3D coil based on FEA (middle), and distribution of maximum Mises stress for each cross section along the natural coordinate normalized by the arc length (right). Reproduced with permission from ref (280) under CC BY. Copyright 2017 Springer Nature. (d) Optical images and FEA results of both undeformed and deformed 3D helical interconnects with Mises-stress distributions in the conditions of one-stage (left) and two-stage encapsulating processes (right). Reproduced with permission from ref (279). Copyright 2019 Wiley.
Uniformly distributed maximum principal strains along the whole structure, ensure the exceptionally large stretchability and mechanical robustness of 3D helical interconnects (Figure 10c).280
In practical applications, the solid encapsulation usually results in an obvious reduction of the interconnect stretchability due to the mechanical constraints. Specifically, when compared to 3D arc-shaped and serpentine designs, the helical designs show more uniform distributions of strains at lower magnitudes, thereby rendering higher elastic stretchability. In this case, the mechanical constraints induced by solid encapsulation are more significant for interconnects with helical designs. To offer a stretchability higher than that of the conventional strategy of direct encapsulation (or one-stage encapsulation), a two-stage encapsulation strategy that involved prestretching the helical interconnects and then adding in encapsulation materials under the prestretched state was developed. Theoretical and experimental results showed that this two-stage encapsulation strategy can delay the occurrence of stress concentration by preunwinding the knot structural features. As such, a significantly enhanced elastic stretchability after encapsulation was achieved, when comparing with conventional one-stage encapsulation methods (Figure 10d).279
To summarize, in order to fabricate architected electronic devices with high levels of stretchability, the development of 3D interconnects with predesigned deployable geometries are essential. When compared with planar interconnects, the arc-shaped interconnects could provide moderate levels of extra stretchability owing to their out-of-plane structural configurations. The helical interconnects feature a relatively uniform curvature distribution, thereby allowing uniformly distributed strains along the whole structure under stretching, which could achieve higher levels of stretchability than arc-shaped interconnects.
4. 3D Device Forms
Flexible electronics with architected 3D configurations can better mimic and conform to the structural forms of natural species with complex spatial geometries, when compared to those in planar device structures. This section reviews diverse 3D forms of flexible electronics manufactured using mechanically-guided assembly methods, including 3D arc-shaped forms,126,132,133,240,242,283−287 3D helical forms,103,145,160,236,281,288,289 tubular forms,91,147,162,290−293 polyhedral forms,172,181,294−298 hemispherical forms,39,101,102,129,130,194,197,198,217,240,299−301 conformally wrapping forms97,125,127,128,167,184,287,302−310 and other complex 3D forms,17,99,122,138,191,203,240,311 through summarizing and discussing the representative functional devices that benefit from their unique structural characteristics.
4.1. 3D Arc-Shaped Forms
Flexible electronic devices in 3D arc-shaped forms are usually fabricated through buckling-guided assembly and curving-induced assembly approaches. Harnessing the discrete design concept, the geometric configuration and deformation of each consisting segment of a device in 3D arc-shaped form can be individually engineered, exhibiting many structural characteristics that induce various functionalities, such as decoupled sensing of physical signals (e.g., mechanics and electromagnetic fields).133,284
For example, a device in the form of 3D herringbone pattern consisting of two concave arcs (Figure 11a) was fabricated for bidirectional measurements of flow rates.286 In particular, a strain sensor was integrated on one of the two arcs (Figure 11a, left one in the top right panel). Upon the reception of incoming flows in the forward direction (as shown in Figure 11a, top right), the strain sensor underwent tensile deformations. Under a reversed flow direction, the strain sensor experienced compressive deformations. Thereby, harnessing such a specifically designed arc-shaped form, the device was capable of bidirectional quantitative measurements of flow velocity. As shown in Figure 11b, an arc-shaped 3D structure, assembled using buckling-guided approaches from a planar crossing ribbon, was adopted as the device configuration.133 Due to the four deformable arcs that can provide multiple channels for the collection of mechanically driven resistance changes, decoupled quantitative sensing of both in-plane and out-of-plane applied forces (e.g., pressure and shear force) was achieved. Furthermore, a flexible electronic device with complex 3D arc-shaped form that incorporates spatially distributed strain sensors, was fabricated to monitor the contractile forces of engineered optogenetic muscle tissue rings in high precision (Figure 11c).287 In this particular case, the two arc-shaped ribbons in the center of the device (Figure 11c, top right), with flat tops and tiny prominent cantilevers (i.e., as hooks), served as a compressible support for the incorporated strain sensors on the outer surfaces of the arc-shaped ribbons to accomplish the measurements of vertically distributed contractile deformations. Figure 11d shows a flexible device in a very simple 3D arc-shaped configuration formed using curving-induced assembly.126 Particularly, the microbars designed to curve the nanowires were also utilized to support the on-top arc-shaped structures with low bending stiffness, rendering a conformally contacted interface between nanowires and cell walls at the initial state. Meanwhile, upon cell contractions, such arc-shaped 3D nanowires allowed postdeformations that can be leveraged for simultaneous measurements of both electrical and mechanical cellular responses. In another case, mechanically triggered switches were devised for 3D RF antennas using a complex arc-shaped 3D configuration formed by buckling-guided assembly (Figure 11e).284 In detail, the “on/off” states were realized by the shape morphing controlled through loading strains. As shown in Figure 11f, by the use of arc-shaped 3D ribbons, flexible graphene microsupercapacitors were fabricated, featuring an high stretchability of 100% (i.e., uniaxial).285
Figure 11.
Devices in 3D arc-shaped forms. (a) Schematic and exploded-view illustrations of a 3D flow sensor. Reproduced with permission from ref (286). Copyright 2023 Springer Nature. (b) Assembly of a 3D piezoresistive sensor capable of normal/shear force measurements: deformed configurations using the compressive buckling (left) and optical images of the circuit design (right). Reproduced with permission from ref (133). Copyright 2019 American Chemical Society. (c) Design of compliant 3D frameworks instrumented with arc-shaped strain sensors for monitoring millimeter-scale muscle tissues. Reproduced with permission from ref (287). Copyright 2021 National Academy of Sciences. (d) Fabrication process of a 3D nanotransistor for simultaneous measurements of electrical and mechanical cellular responses. Reproduced with permission from ref (126) under CC BY. Copyright 2022 The American Association for the Advancement of Science. (e) Design and fabrication of a ribbon-like dipole antenna with tunable frequency. Reproduced with permission from ref (284). Copyright 2019 Wiley. (f) Fabrication of a stretchable microsupercapacitor composed of wavy-structured electrode arrays. Reproduced with permission from ref (285). Copyright 2015 Wiley.
4.2. 3D Helical Forms
Flexible electronic devices in 3D helical forms can be assembled through approaches harnessing rolling, curving, and buckling. Similar to helical interconnects, devices in 3D helical forms render excellent stretchability and bendability. The structural configuration of 3D helices can be utilized to manufacture devices with novel structure-induced functionalities such as actuations in liquids145,160 and spiral growth.310
For example, as shown in Figure 12a, bioinspired 3D helical structures prepared through swelling-induced rolling assembly were exploited to achieve controlled motions for swimming robots.160 The large out-of-plane deformations (i.e., vertical displacements) of 3D helical architectures were harnessed to create high temperature gradients for efficient thermoelectrical energy harvesting (Figure 12b).289 In particular, the use of 3D helical structures significantly increased the altitude intercept between the heat receiving surface (i.e., the device surface that was in contact with the heat supplier) and the heat dissipation surface (i.e., the suspended top surface of the device) of the device (Figure 12b, top right), thereby generating an enlarged temperature gradient within the 3D device when compared with those in planar forms. As shown in Figure 12c,103 the spatial characteristics of 3D helical structures were exploited to reduce the substrate parasitic capacitance of an inductor, resulting in improved electrical performances (e.g., the maximum Q factors and resonant frequencies increased from 1.7 to 2.2 and from 6.8 to 9.5 GHz, respectively). Harnessing the tunable structural gaps between two interwoven helices upon stretching (Figure 12d) (i.e., consisting of conductive polyurethane-based stretchable fibers embedded with Ag nanoparticles), capacitive strain sensors were fabricated through curving-induced assembly.281
Figure 12.
Devices in 3D helical forms. (a) Mass production procedure of soft 3D helical microswimmers through swelling-induced rolling. Reproduced with permission from ref (160) under CC BY. Copyright 2019 The American Association for the Advancement of Science. (b) Fabrication of 3D thermoelectric coils through the buckling-guided assembly. Reproduced with permission from ref (289) under CC BY. Copyright 2018 The American Association for the Advancement of Science. (c) A stretchable 3D toroidal helical inductor with tunable electrical properties. Reproduced with permission from ref (103). Copyright 2015 The American Association for the Advancement of Science. (d) Schematic illustrations of a wireless strain-sensing system and the fabrication process of capacitive fiber strain sensors. Reproduced with permission from ref (281) under CC BY. Copyright 2021 Spring Nature.
4.3. Tubular Forms
Flexible electronic devices in tubular forms are primarily fabricated by rolling assembly approaches.91 Owing to their structural characteristics (e.g., hollow tubular configurations, micro- to nanoscale diameters, multilayered shells and etc.), tubular flexible electronic devices exhibit unique physical properties, such as controlled light interactions (e.g., scattering and absorption293,312), fluidic controls,147 high density material integrations,291,292,313 and so on.
For instance, tubular 3D quantum well infrared photodetectors (QWIPs) could render enhanced responsivity and detectivity (Figure 13a).293 In particular, the tubular form of the photodetector provided efficient pathways for light coupling and enabled the elimination of the requirement for external light coupling structures, featuring a wide detection angle (i.e., ± 70°) for infrared lights. Additionally, by variation of the structural configurations of such photodetectors, tunable photocurrents and responsivities were also achieved.
Figure 13.
Devices in 3D tubular forms. (a) Schematic diagram and optical image of a 3D tubular quantum well infrared photodetector (QWIP). Reproduced with permission from ref (293) under CC BY. Copyright 2016 The American Association for the Advancement of Science. (b) A self-assembled microfluidic device for flow control and visualization. Reproduced with permission from ref (147). Copyright 2011 Springer Nature. (c) Tubular MEMS actuators and switches using VO2 nanomembranes. Reproduced with permission from ref (162) under CC BY. Copyright 2021 Springer Nature. (d) Schematic illustrations of the ion diffusion and charge transport path of a tubular micro-lithium-ion battery manufactured by residual-stress-induced rolling (right) and the corresponding titled cross-sectional SEM image of an as-prepared spiral electrode (left). Reproduced with permission from ref (313). Copyright 2020 Wiley. (e) A “Swiss-roll” nano-biosupercapacitor (nBSC) used in biological electrolytes. Reproduced with permission from ref (292) under CC BY. Copyright 2021 Springer Nature. (f) Self-assembly of a “Swiss-roll” structure assisted by external magnetic fields. Reproduced with permission from ref (291) under CC BY. Copyright 2019 Springer Nature.
The tubular form can also be utilized to facilitate microfluidic controls.147 For example, through rolling assembly, a tubular SU-8 structure was exploited to reshape planar microfluidic networks into a 3D configuration with controlled curvatures (Figure 13b). In addition, through the use of a tubular configuration (Figure 13c),162 a bifunctional device was manufactured and utilized for the actuation of a microelectromechanical system (MEMS) as well as electrically/optically controlled switching.
The layered shell structures of tubular configurations were also exploited to integrate more functional materials for flexible energy devices to achieve energy densities/capacities comparable to rigid ones.291,292,313−315 For example, a tubular microbattery fabricated by rolling assembly of Au/Ge/Si trilayer nanomembranes was demonstrated to obtain high capacity and excellent cycling performance (Figure 13d).313 Using “Swiss-roll” structures, robust nano-biosupercapacitors (nBSCs) consisting of multilayered shells (as shown in Figure 13e) were prepared and functioned as in vivo smart dusts as well as the microrobotic systems for healthcare.292 Additionally, by incorporating thin magnetic films in a similar “Swiss-roll” configuration, the rolling assembly process could be precisely controlled using magnetic fields, resulting in various sophisticated 3D geometries (Figure 13f).291
4.4. Polyhedral Forms
Flexible electronic devices in polyhedral forms are usually manufactured using folding assembly methods.172,181 Polyhedral forms, such as pyramids and cubes, stand as a typical type of spatial configuration that can facilitate structural functionalities involving capturing,125,134,184,295,316−318 sensing,294,296,298,319 light emitting,297,320 and the rest.
Through the folding of patterned bilayer thin films, microgrippers were manufactured enabling controlled captures (e.g., folding angles) of single cells (Figure 14a).295 Foldable polyhedral displays consisting of ultrathin quantum dot light-emitting diodes (QLEDs) were fabricated (Figure 14b), featuring stable operations during transformations among various structural configurations (e.g., the airplane, butterfly, pyramid, and cube).297 In addition, a cubic monolayer graphene device (Figure 14c) was manufactured, achieving enhanced volumetric light confinements on both the surfaces and the designed hinges/edges.294 Similarly, by folding MoS2-Au-SU-8 photodetectors into cubes and dodecahedrons, an enhanced optical absorption within visible range (400–700 nm) was also achieved through the structure-induced light interactions (e.g., scattering and trapping) at the boundaries and corners (Figure 14d).296
Figure 14.
Devices in 3D polyhedral forms. (a) Fabrication and operation of single-cell grippers made of the SiO/SiO2 bilayer. Reproduced with permission from ref (295). Copyright 2014 American Chemical Society. (b) Design and fabrication of various 3D architectures by folding preprogrammed ultrathin quantum-dot light-emitting diodes (QLEDs). Reproduced with permission from ref (297). Copyright 2021 Springer Nature. (c) Assembly of a monolayer graphene-based cubic structure. Reproduced with permission from ref (294). Copyright 2017 American Chemical Society. (d) Folding of hinged 3D MoS2-Au-SU-8 photodetectors with enhanced optical absorption. Reproduced with permission from ref (296). Copyright 2019 American Chemical Society.
4.5. Hemispherical Forms
Flexible electronic devices in hemispherical forms are usually fabricated through curving-induced and buckling-guided assembly methods. Hemispherical forms with structural expansions in vertical directions can enhance the performances of electromagnetic devices, such as antennas, by offering more surfaces to integrate electronic circuits with longer paths while maintaining a small overall projection dimension of the device.299,301 For example, as shown in Figure 15a, through the buckling-guided assembly, an electrically small antenna (ESA) was manufactured in a hemispherical form with significantly enhanced electrical paths, featuring a high Q factor and tunable resonant frequencies.301
Figure 15.
Devices in hemispherical forms. (a) Fabrication of electrically small antennas (ESAs) by buckling-guided assembly. Reproduced with permission from ref (301). Copyright 2018 Wiley. (b) Schematic illustrations of components and integration schemes for a digital camera in the form of a hemispherical, apposition compound eye. Reproduced with permission from ref (129). Copyright 2013 Springer Nature. (c) Design and fabrication of hemispherical devices for photodetection and imaging. Reproduced with permission from ref (217) under CC BY. Copyright 2018 Springer Nature.
In addition, hemispherical forms can also serve as an isotropic spatial integration platform, thereby enabling flexible optoelectronic devices with omnidirectional sensing/imaging capabilities and reduced geometric distortions.39,101,129,130,194,197,198,217 For instance, as shown in Figure 15b, through the curving-induced assembly, an arthropod-inspired digital camera was developed, featuring a wider field of view and a more homogeneous intensity when comparing to nonhemispherical cameras. Notably, origami and kirigami designs were also exploited to enhance the conformability of the planar structures to the hemispherical surfaces.130,197 In another case, a hemispherical photodetector was manufactured (Figure 15c),217 where the omnidirectional 3D integration enabled by the isotropic configuration of a hemispherical form allowed the simultaneous measurements of both light intensities and incident angles.
4.6. Conformally Wrapping Forms
The formations of conformally wrapping devices usually involve complex deformation modes (refs (97, 125, 127, 128, 167, 184, 287, 302−310)). Similar to polyhedral forms, devices in conformally wrapping forms can also facilitate capturing, owing to their semiclosed/closed structural configurations. An obvious difference between conformally wrapped configurations and polyhedral forms lies in the conformability of constructed interfaces between the captured object and a device. To be noted, in this review, conformally wrapping forms denote the assembled structures of architected flexible electronic devices that feature wrapping capabilities to form conformal contacts with their captured objects at the micro- to nanoscale (e.g., cells, organoids, and the rests), which should be distinguished from those macroscale 3D flexible electronic devices.
Microgrippers capable of controlled capture and release were demonstrated in conformally wrapping forms for various in vivo applications (e.g., surgery).97,167,303 For instance, thermally actuated wrapping grippers with different dimensions ranging from 300 μm to 1.5 mm (Figure 16a) were fabricated using rolling assembly, enabling in vivo tissue extractions from living organisms.303 Harnessing the shape memory effect of SMPs, twining neural interfaces were generated in forms of conformally wrapping helices, featuring body-temperature-driven postassembly climbing motions (Figure 16b).310 The devices in conformally wrapping forms can also facilitate the in situ monitoring of living organisms. As shown in Figure 16c, through the residual-stress-induced rolling assembly, conformally wrapping electronic interfaces were manufactured, capable of single-cell monitoring (i.e., cardiomyocytes).125 In particular, taking advantage of the engineered strain mismatches between SiO2 and SiO layers (i.e., by varying the widths and thicknesses), the deformation angles of the structure can be tuned to achieve a conformal contact with cellular surfaces. Devices in conformally wrapping forms can also enable the spatial integration of biosensors, thereby allowing 3D electrical mapping and monitoring for neurons.127,305,309 For example, through the buckling-guided assembly, a 3D device consisting of a high-density electrode array that gently wrapped around the surface of a spheroid was fabricated, capable of in situ monitoring of cellular growth and intercellular communications (Figure 16d).127
Figure 16.
Conformally wrapping devices. (a) Fabrication and actuation of microgrippers driven by thermally induced residual stresses. Reproduced with permission from ref (303). Copyright 2012 Wiley. (b) Fabrication of a twining electrode with optical image showing its climbing process on a glass tube. Reproduced with permission from ref (310) under CC BY. Copyright 2019 The American Association for the Advancement of Science. (c) SEM images of a wrapping shell with multiple electrodes and FEA results showing the dependence of folding angle on the SiO/SiO2 thicknesses. Reproduced with permission from ref (125). Copyright 2018 Wiley. (d) Buckling-guided assembly of a 3D multifunctional neural interface wrapping a neural spheroid. Reproduced with permission from ref (127) under CC BY. Copyright 2021 The American Association for the Advancement of Science.
4.7. Other Complex Forms
Apart from the above, many other 3D configurations are also exploited for architected flexible electronics, such as compliant configurations,191,203,240 complex origami/kirigami configurations,17,120,122 and architectures with rationally designed microlattices138,321 that are capable of replicating curved biological surfaces.
Figure 17a shows an electronic interface in a curvilinear form, manufactured by the use of curving-induced assembly (associated with local buckling), which was able to conform to many curved surfaces, such as a golf ball.99 An ordered two-stage buckling-guided assembly strategy was developed to transform planar electronic devices into sophisticated 3D forms that can conform to a variety of curved surfaces, such as the apex of a heart model (Figure 17b).240Figure 17c demonstrates a flexible electronic device in the form of a 3D integumentary thin membrane manufactured by the curving-induced assembly.191 In particular, the heart-shaped 3D thin-film configuration allowed the device to form conformal contact with the beating heart and function as an artificial pericardium. Furthermore, using the curving-induced assembly based on the thermoforming,203 mountain-shaped and ear-shaped 3D electronics were also fabricated (Figure 17d).
Figure 17.
Devices in other complex 3D forms. (a) Schematic illustration of curvilinear silicon mesh circuits that wrap around a golf ball. Reproduced with permission from ref (99). Copyright 2009 Wiley. (b) A 3D electronic device capable of conformal attachment on a heart model. Reproduced with permission from ref (240) under CC BY. Copyright 2022 The American Association for the Advancement of Science. (c) Design and fabrication of 3D multifunctional integumentary membranes (MIMs) for spatiotemporal cardiac measurements. Reproduced with permission from ref (191) under CC BY. Copyright 2014 Springer Nature. (d) 3D electronics capable of conformal contact with complexly curved surfaces. Reproduced with permission from ref (203) under CC BY. Copyright 2021 The American Association for the Advancement of Science. (e) Origami microactuators with various folding configurations. Reproduced with permission from ref (122). Copyright 2021 The American Association for the Advancement of Science. (f) 3D mesotructures with complexly curved surfaces using microlattice designs: schematic illustrations (left) and inverse designs of blueberry flower and P. philadelphica berry (right). Reproduced with permission from ref (138). Copyright 2023 The American Association for the Advancement of Science.
Origami designs were also utilized to further enhance the shape-morphing capabilities of devices in forms of complex 3D geometries. As shown in Figure 17e, a series of origami-inspired 3D structures were fabricated using the folding assembly (i.e., induced by electrochemically driven redox reactions), enabling the development of microscale 3D shape-memory actuators.122
Recently, a microlattice design strategy that featured programmable control of structural stiffnesses (i.e., by tuning the porosity distribution of microlattices) was developed, enabling the fabrication of devices in almost arbitrarily curved 3D architectures that can replicate complex biological surfaces (Figure 17f), such as the blueberry flower and Philadelphica berry (Figure 17f, right panel).138
In summary, there are numerous 3D forms in which each demonstrates unique structural functionalities throughout the broad family of architected flexible electronics. The core functionalities of a conventional rigid device are mainly achieved by circuit designs, while for architected flexible electronics, their functionalities are achieved by designs of both electrical circuits and structural configurations. Among different 3D device forms, due to their elementary structural configurations, the overall 3D geometries of arc-shaped devices could be easily shaped to fit the requirements of mechanical sensing, which have found essential applications in strain sensors. Owing to their unique biomimetic hydrodynamic characteristics, the helical forms are ideal for swimming microrobots, which can be applied for drug delivery. Thanks to their unique micro- to nanoscale hollow configurations, the tubular forms hold great potential in photodetectors and microfluidic devices. Notably, through the use of multilayered shells, tubular structures were also exploited to fabricate devices (e.g., microbatteries) with high energy densities. The polyhedral forms are suitable for object capturing at the micro- to nanoscale (e.g., microgrippers) and light interaction (e.g., photodetectors and solar cells), owing to their open 3D cavities and nanoscale structural features at hinges. The hemispherical forms, featuring omnidirectional structural expansions in vertical directions, represent an ideal class of 3D configurations for electromagnetic devices (e.g., antennas) and optoelectronic devices (e.g., photodetectors). The conformally wrapping devices with semiclosed/closed structural configurations are suitable for applications in biological/biomedical devices for cell/tissue/organ monitoring.
5. Structure-Induced Functionalities
Conventionally, the functionalities of electronic devices usually indicate the electrical roles that integrated circuits play. Since the 20th century, rapid developments of human society have emerged technological revolutions in many aspects, such as Internet of Things (IoT), personalized medicine, human-machine interactions and others. As a consequence, the accessible functionalities of electronic devices have been drastically expanded. Different 3D configurations have been developed to induce unique structural functionalities, thereby adding more dimensions to the design of flexible electronics. Such structural revolutions have transformed the roles of electronic devices from solely electrical to mechanoelectrical,22,132,280 optoelectrical,130,217 magnetoelectrical,322 thermoelectrical,61,289 chemoelectrical,236,242,323,324 bioelectrical,127,138,325−327 and physically intelligent.18,22,328,329
In many of the reviewed cases in previous sections, structures do play vital roles in the achievement of their functionalities. However, when it comes to function-driven structural designs of devices, the routes become blurry. Therefore, it is essential to draw clear guidelines to navigate the function-driven structural designs of architected flexible electronic devices. In this section, various structure-induced functionalities, including high areal density 3D integration with spatial resolution, high-efficiency energy harvesting, 3D compliant electronic interfaces, reconfiguration, growth, and structural evolution, are reviewed and discussed, each with exemplary sketches illustrating their structural configurations and functionalities.
5.1. High Areal Density 3D Integration with Programmable Spatial Resolution
Electronic devices in planar forms utilize only in-plane spaces to integrate functional circuits, usually with a limited areal density in the cases of stretchable designs.68,330−335 Various 3D configurations enabled by mechanically-guided assembly methods stand as promising platforms to achieve higher areal densities in flexible electronics, owing to their capabilities of expanding the distributions of functional circuits/elements into out-of-plane spaces with programmable spatial resolutions. In terms of achieving higher areal densities, both the multilayer flexible PCB (FPCB) and mechanically-guided assembly methods can be exploited.121,333,336 In comparison to mechanically-guided assembly methods, the fabrication of multilayer FPCB is more simple and low cost. However, in order to fabricate architected 3D flexible electronics with specific structure-induced functionalities, mechanically-guided assembly methods are more suitable. Notably, the construction of 3D architectures would result in unavoidable spatial density loss of the device due to the spatial voids generated by the mechanically-guided assembly.
As illustrated in Figure 18, the idea can be simply analogized as building flexible/stretchable “flyovers” consisting of high-density electrical circuits. Harnessing 3D structural characteristics, more surface area could be utilized for device integration when compared with planar configurations of the same projection area. In particular, the original in-plane surfaces for integrations (i.e., the projection area) are retained, while extra surfaces are created through the structural pop-up.127,181,217,249,337 Furthermore, the use of multilayer designs could also increase the areal density of the resulting flexible electronic devices, by simply adding more layers for integration while keeping the in-plane projection area unchanged.121
Figure 18.
Schematic illustrations of typical cases for 3D integrations. By the use of a complex arc-shaped design (as a representative form), functional electronic components are spatially integrated, while providing significantly enhanced biaxial stretchability and flexibility.
Apart from high areal densities, the 3D integration also enabled a variety of architected flexible electronic devices with programmable spatial resolutions for sensing of various physical/chemical signals. For instance, a comprehensive photodetection should involve not only the intrinsic physics of incident light, such as wavelength, light intensity, and polarization, but also the external information on light sources, for instance, the incident angles and source locations. As shown in Figure 19, the sensing with desired spatial resolutions can be realized by creating a 3D dome structure (or other 3D structures that can provide spatial distributions of functional materials/electronic elements) through the use of the buckling-guided assembly. Furthermore, through the integrations of various functional semiconductors (e.g., MoS2,194,217,296,338 MoO2,339 ZnO,340−344 TiO2,345,346 perovskites,347,348 etc.) at the preset spatial locations, mechanically-guided methods can also facilitate the manufacturing of diverse sensors with spatial resolutions, enabling detections of different physical/chemical signals in 3D spaces, such as sound fields, thermal fields, magnetic fields, volumetric distributions of volatile organic compounds (VOCs), dispersions of toxic nanoparticles, among others.
Figure 19.
Schematic illustrations of a typical case of spatial resolution. Harnessing the hemispherical configuration integrated with arrays of photodetectors, both the intensities and incident angles can be simultaneously measured. Notably, photodetection is only shown as a representative case. Sensing of other directional dependent signals with spatial resolution, such as sound, thermal radiations, vibration, and the rest, can also be achieved through specific 3D structural designs in other forms.
5.2. High-Efficiency Energy Harvesting
Vibrational energies are abundant in nature. To harvest such form of energies, well-defined 3D architectures with engineered low stiffness regions that can vibrate without mechanical constraints are often required.349 In planar devices, to achieve such functionality, cantilevers made of advanced electronic materials (e.g., graphene, PZT or III-V semiconductors) are usually manufactured with delicate multistep processing procedures. Such complicated manufacturing routes have set a series of grand challenges for their mass productions. In stark contrast, harnessing mechanically-guided structural design and 3D assembly methods (e.g., buckling-guided assembly), devices that sense and harvest vibrational energies can be facilely fabricated.132,219 For instance, 3D architected flexible electronic devices with specifically designed local stiffness (i.e., the use of suspended serpentines with ultralow stiffness at the vibrational regions of the device) can vibrate and resonate upon receiving external stimuli so as to convert and generate energies in forms of electrical currents or voltages (Figure 20). Additionally, the use of mesostructures with ultralow stiffnesses would significantly reduce the resonant frequencies at small scales, thereby enabling efficient energy harvesting of low-frequency vibrations using miniaturized devices. Since the motions of human body and organs are usually in the range of several to tens of Hz, these 3D architected energy harvesters can be used to power biointegrated electronics. Furthermore, by incorporating multistable components in the 3D architecture, the bandwidth can be widened to improve the efficiency of vibration energy harvesting.
Figure 20.
Schematic illustrations of a typical case of energy harvesting. The ultralow stiffness of a serpentine design can be utilized to fabricate an energy harvesting device (through vibrations) only by the formation of 3D architectures that allow fully suspended serpentines under operation modes.
Aside from the vibrational energies, photogenerated energies (often in form of solar energies) are also ubiquitous in nature. The existing devices for photoenergy harvesting or conversion are mostly in planar configurations, such as solar cells,76,350,351 photoelectrochemical water splitting (PEC) cells,352−354 devices for artificial photosynthesis,355 photocatalytic CO2 reductions cells,348,356 and so on. Due to the angle-dependent optical properties of 2D structures (e.g., absorption, reflection, transmittance, scattering, and other forms of optical interactions), planar devices are not able to take full use of solar energies (e.g., shadowing caused by low incident angles of lights). The existing forms of domes or hemispherical configurations in architected flexible electronics manufactured through mechanically-guided assembly methods can be leveraged to resolve the above issue. For example, a 3D electronic device capable of 360° light harvesting and light tracking can be fabricated, representing a promising form of photoenergy-conversion devices.357
5.3. 3D Compliant Electronic Interfaces
Complex 3D surfaces with sophisticated curvature distributions are found in both natural creatures and artificial objects, spanning cells, organoids, human bodies, machines, vehicles, and robots. Flexible electronic devices that can conform to such curved surfaces can serve as compliant electronic interfaces that induce minimal mismatch stresses, thereby allowing long-term monitoring of either biological or artificial objects (e.g., a single cell, organoids, tissues, organs, auto instruments, and so on). By the use of mechanically-guided assembly methods, many 3D compliant electronic interfaces are enabled for such purposes (e.g., twining neural electrode,310 multifunctional electronic interface wrapping a spheroid,127 complex interfaces between heart and thin electronic membrane,191 and so on128,302,309). For instance, through the buckling-guided assembly, flexible electronic devices with 3D concave structures can be manufactured, holding promising potential to achieve controlled capture of a single cell or organoid and conformal interfaces between the electronic surface and cellular membrane surface (Figure 21).
Figure 21.
Schematic illustrations of a typical case for 3D compliant electronic interfaces. A flexible cell device showcases the formation of a compliant conformal interface between the surfaces of an electronic device and a cellular membrane assisted by mechanically-guided assembly.
5.4. Growth, Reconfiguration, and Structural Evolution
Structural evolutions (not limited by the concepts of morphogenesis, organogenesis, or others), such as growth, shape changing, and complex structural transformations, are of pivotal importance for biology. By an artless biomimetic thinking, morphable devices that can reconfigure, grow, reshape, or even structurally evolve represent a more advanced form than those with fixed 3D geometries.
For instance, 3D architectures with multistable steady states can be reshaped by adopting different strategic loading paths (Figure 22, first row) during the buckling-guided assembly. Such reconfigurable flexible electronics have emerged with various new applications in the fields of electromagnetic devices and soft robotics.16,139,358
Figure 22.
Schematic illustrations of a typical case for reconfiguration, growth, and morphological evolution. A representative reconfigurable device is shown in the top row, enabling reversible shape transformation between two stable shapes (shape I and shape II, respectively). The second row shows a flexible device that can first form conformal contact with a muscle tendon by body temperature actuated shape morphing. Then, taking advantage of the 3D helical structure, such device can climb along the tendon. Last but not least, harnessing the flexibility and stretchability, such device can grow with the tendon, showing no oblivious mechanical constrains to the tendon during growth. The third row shows a morphable electronic device (assisted by machine leaning) that can sense the structure and self-evolve to the target 3D configurations by electromagnetically controlled actuation.
As illustrated in Figure 22 (second row), electronic devices with highly flexible designs, such as helical structures310,359 are able to (i) spontaneously conform to the curved surface of a certain biological tissue (e.g., muscle tendon); (ii) climb along the 3D geometries of their attached tissue; and (iii) grow together with the tissue. Such growable designs could facilitate fundamental biological studies of living organisms (e.g., organs, muscles, tendons, neurons, among others) by providing morphable electronic platforms for long-term in situ monitoring of their biological behaviors during growths.
Very recently, a new type of electromagnetically controlled metasurfaces capable of dynamic morphological evolutions has been demonstrated.360 Using discrete morphable serpentine units consisting of polymeric materials and patterned electronic circuits, centimeter-scale morphable surfaces (Figure 22, third row) that can rapidly self-evolve to the target 3D configuration are developed through mechanically-guided methods. Particularly, their structural evolutions are achieved by programming the spatially distributed Lorenz forces applied to the various units. Such self-evolving electronics have already fuzzed up the boundaries of 2D and 3D devices.
In conclusion, structure-induced functionalities (i.e., enabled by mechanically-guided assembly) are diverse and sometimes case-by-case, which opens up a new perspective for the design of architected flexible electronics. Devices that can vibrate and harvest mechanical energies, omnidirectionally absorb and convert incident photoenergies, spatially receive and sense physical/chemical signals, deform and conform to curved natural/artificial surfaces, and grow and evolve spontaneously were just scientific fictions decades ago. Despite this exciting progress, enriched device functionalities induced by structural designs that mostly originated from intuitive inspirations, due to the lack of rational function-driven structural designs. Open opportunities lie in the developments of interdisciplinary function-driven design paradigms that could rationally link diverse functionalities to the corresponding class of 3D architectures.
6. Applications
The intriguing structure-induced functionalities enabled by the mechanically-guided 3D assembly have significantly pushed the research frontiers of flexible electronics, and led to the developments of many new devices,5,361−363 enabling a broad spectrum of applications.104,105,364−366 This section reviews applications of architected flexible electronics spanning various emerging fields including biology, biomedicine, electromagnetics, optoelectronics, energy, and robotics.
6.1. Biological Devices
6.1.1. Cell Devices
Architected flexible electronic devices stand as powerful tools to capture and read dynamic electrophysiological signals of cellular processing (e.g., growth, differentiation, replication, and apoptosis) and intercellular communications from the perspective of a single cell.125,182,295,367−371 As shown in Figure 23a, a flexible electronic platform consisting of a 3D field-effect transistor (FET) array (i.e., with up to 128 FETs as demonstrated) was developed by the use of buckling-guided approaches, with capabilities of recording transmembrane potentials in cardiomyocytes. In particular, the measurements based on this platform indicated an intracellular signal conduction velocity of 0.182 m s–1, which was around five times higher than that of intercellular signal conduction.325
Figure 23.
3D biological electronics for fundamental studies of cells and organoids. (a) Design and construction of 3D FET arrays using compressive buckling and the intracellular recording of electrophysiological signals from neonatal rat cardiomyocytes. Reproduced with permission from ref (325). Copyright 2022 Springer Nature. (b) An ultrathin and flexible biosensing platform that wraps around a micro-object, allowing for 3D molecular spectroscopy via SERS. Reproduced with permission from ref (316). Copyright 2019 American Chemical Society. (c) Curved shell with spatially distributed electrodes, which wrap around the cardiomyocyte cell for electrophysiological monitoring. Reproduced with permission from ref (125) under CC BY. Copyright 2018 Wiley. (d) 3D neural interface with a microelectrode array, which can wrap a neural spheroid for spatiotemporal mapping. Reproduced with permission from ref (127) under CC BY. Copyright 2021 The American Association for the Advancement of Science.
Using the mechanically-guided assembly, various microgrippers were also fabricated for single-cell capturing and manipulations (such as extraction, delivery, and controlled release). For instance, arrayed biocompatible grippers, harnessing engineered actuation through SiO/SiO2 bilayers, were fabricated for precise extraction and delivery of single cells.184,295 Later, by further introduction of responsive materials, such as magnetic materials or thermally responsive materials,184 remote controls of such grippers were enabled. Such microgrippers show great promise for high-throughput biopsies. For example, a flexible electronic device, capable of mechanical trapping and surface-enhanced Raman spectroscopy (MT-SERS),372 was developed as a new tool for simultaneous capturing, profiling, and 3D microscopic mapping of intrinsic molecular signatures of a single living cell (Figure 23b).316 By further integration of microelectrodes on top of the microgrippers, electronic platforms that allowed spatiotemporally recordings of living cells were also developed.125 In particular, utilizing residual-stress-induced rolling, such grippers were able to conformally wrap around a single cell to record the propagation of action potentials (Figure 23c). Flexible arrays of tubular electronics consisting of rolled-up transparent oxides (i.e., SiO/SiO2) were also exploited as platforms for cultures369 and encapsulations of living mammalian cells (i.e., HeLa cells),370 resulting in different cellular assemblies and chromosomal instability during cell divisions. Later on, by further incorporating impedimetric microfluidic sensors into the above platform, simultaneous analyses of single human monocytes and their mediated activations were achieved.373
6.1.2. Organoid Devices
Zooming out from the single-cell perspective, architected flexible electronics have also significantly enriched the methodologies for organoid studies.127,128,147,184,236,305 The development of multifunctional compliant 3D electronic frameworks stood as a representative achievement in this field.236 Equipped with sensors that respond to various physical/chemical signals (Figure 23d),127 such frameworks enabled both in vitro and in vivo biological investigations, such as the recording of coordinated bursting events, regrowth of bridging tissues in cortical spheroids and mechanical characteristics of organoids.305 In addition, shell microelectrode arrays (MEAs) were developed to study brain organoids,128 featuring high signal-to-noise ratio sensing and 3D spatiotemporal recording. Furthermore, compliant 3D arc-shaped structures integrated with strain sensors were exploited to characterize engineered muscle tissues at the millimeter-scale,287 offering high-sensitivity measurements of contractile forces and motions with temporal resolutions. Complex 3D microfluidic architectures were also manufactured for artificial vascular networks, indicating potential applications in 3D cell cultures, engineered tissues, and artificial organs.147,236 Self-powered micro-oscillating systems with low power consumptions were also developed to mimic and replace nature’s propulsion and pumping units at low Reynolds numbers,328 offering open opportunities for the constructions of artificial lives.
6.1.3. Electronic Tissue Scaffolds
Mechanically-guided assembly methods have also driven the development of tissue engineering,214,287,374,375 by creating electronic tissue scaffolds that are capable of in situ monitoring and stimulations. For example, through the buckling-guided assembly, a 3D flexible electronic device with a double-layer-cage configuration was fabricated and utilized as electronic cellular scaffolds for the engineering of dorsal root ganglion (DRG) neural networks (Figure 24a).214 Particularly, such electronic scaffolds facilitated the growth of organized lamellae by providing electrical simulations in terms of capacitive charge injections. Photopatterned tubular biomimetic microvessels were prepared,374 featuring enhanced endothelial longevity and nitric oxide production, making them promising platforms for investigations of microvascular pathobiology in human diseases like pulmonary hypertension (Figure 24b). Transparent tubular scaffolds, namely bioartificial endocrine pancreas (BAEP), were also developed to enable an efficient mass transport.376 In addition, examinations of human mesenchymal stem cell migration, adhesion, and osteogenic differentiation were achieved through the use of flexible electronic tissue scaffolds (Figure 24c).375 Furthermore, in situ studies of retinal pigment epithelium (RPE) cells were enabled by the development of spherical cap-shaped electronic cell scaffolds, featuring noninvasive monitoring of spatially distributed physiological activities, such as growth and apoptosis (Figure 24d).138
Figure 24.
3D biological electronics for tissue engineering. (a) 3D electronic scaffolds for DRG neural networks with uniform dispersions of cells allowing study of the electrophysiological behaviors of the growing. Reproduced with permission from ref (214) under CC BY. Copyright 2017 National Academy of Sciences. (b) Biomimetic tubular architectures that are similar to human small muscular distal acinar pulmonary arteries enables higher activation of cells. Reproduced with permission from ref (374) under CC BY. Copyright 2020 The American Association for the Advancement of Science. (c) In vitro 3D electronic platforms for studying bone cell–material interactions during osteogenic cell differentiation. Reproduced with permission from ref (375) under CC BY. Copyright 2021 Wiley. (d) Spherical cap-shaped electronic cell scaffold with integrated sensing capabilities. Reproduced with permission from ref (138). Copyright 2023 The American Association for the Advancement of Science.
6.2. Biomedical Devices
6.2.1. In Situ Monitoring Devices
3D flexible electronics manufactured through mechanically-guided methods have emerged as promising tools for in vivo monitoring of physiological signals for living organisms, such as tissues and organs.132,191,280,286,302,310,377−384 For example, 3D piezoelectric microsystems (Figure 25a) were fabricated for implantable monitoring of muscle activities (e.g., trotting and climbing), providing platforms for in vivo biological interactions without obvious irritations or mechanical constraints.132 Other 3D forms with high stretchabilities were exploited in wireless skin-compatible electronic sensors, enabling the tracking of spatial motions and the monitoring of electrophysiological signals.280 Climbing-inspired twining electrodes based on shape memory effect of helical SMP structures,310 were fabricated for peripheral neuromodulation (Figure 25b), which can reduce substantially mechanical and geometrical mismatches at the electrode-nerve interfaces. 3D multifunctional membranes191 were designed for spatiotemporal cardiac measurements and stimulations across the entire epicardium. Such membranes could maintain a stable biotic/abiotic interface during cardiac cycles, standing as promising platforms for cardiac studies (Figure 25c). Additionally, other flexible electronics assembled through the use of mechanically-guided methods, such as biosupercapacitors292 and physical sensors,135,378 were also exploited for in situ monitoring.
Figure 25.
3D biomedical electronics for in vivo monitoring of physiological signals. (a) 3D piezoelectric implants capable of voltage measurements during different actions. Reproduced with permission from ref (132). Copyright 2019 Springer Nature. (b) 3D deformable twining electrode for in vivo vagus nerve stimulation (VNS). Reproduced with permission from ref (310) under CC BY. Copyright 2019 The American Association for the Advancement of Science. (c) 3D multifunctional integumentary membranes (3D-MIMs) integrated with a rabbit heart for spatiotemporal measurements of electrical signaling of both pH and temperature. Reproduced with permission from ref (191) under CC BY. Copyright 2014 Springer Nature.
6.2.2. In Situ Therapeutic Devices
Architected flexible electronic devices have also found immense potential in therapeutic applications, such as drug delivery134,317,385−387 and pain management.304,388−392 Many morphable microdevices with specific structural designs were utilized to fabricate microgrippers for in situ therapeutics, which can efficiently deliver drugs within biological lumens (e.g., the gastrointestinal (GI) tract).134 For example, inspired by hookworms, microgrippers that can autonomously latch onto the mucosal tissue were manufactured (Figure 26a, left panel), rendering in vivo drug delivery performances of an almost 2-fold exposure (i.e., oral ketorolac tromethamine without the use of microgrippers, Figure 26a, middle panel) as well as a significantly enhanced retention (i.e., a 6-fold increase in the elimination half-life of a model analgesic, ketorolac tromethamine) (Figure 26a, right panel). Flexible electronic devices with rationally designed 3D shapes were also developed for pain management by blocking peripheral nerve activities.304,388,393 For instance, bioresorbable curling flexible microfluidic devices, featuring precisely controlled delivery of cooling power at arbitrary depths in living tissues, stood as promising tools for pain blocking (Figure 26b).388 Architected flexible electronic devices in 3D helical forms were also manufactured, featuring spontaneous constructions of conformal neural interfaces around small living nerves during the insertion.304
Figure 26.
3D biomedical electronics for therapeutics and surgeries. (a) Submillimeter-scale bioinspired latching tools for enhanced drug release and retention. The right two panels show extended and higher exposures of ketorolac compared to pristine drug. Reproduced with permission from ref (134) under CC BY. Copyright 2020 The American Association for the Advancement of Science. (b) 3D bioresorbable peripheral nerve-cooling and temperature-sensing device for pain blocking. Reproduced with permission from ref (388). Copyright 2022 The American Association for the Advancement of Science. (c) Multifunctional 3D electronic balloon catheter for in vivo electrophysiological/temperature mapping and RF ablation. Reproduced with permission from ref (188). Copyright 2011 Springer Nature. (d) 3D electronic catheters for diagnosis and treatments, including endocardial electrophysiological studies, electrogram measurements, and RF ablation. Reproduced with permission from ref (135). Copyright 2020 Springer Nature.
6.2.3. Surgical Instruments
Owing to the excellent conformability with the biological surfaces of organs, instrumented flexible devices for surgery were developed.135,188,219,394−397 For instance, the pressing demand of functional integration for catheters in clinics pushed the development of miniaturized flexible electronic devices with various 3D architectures.398−402 A multifunctional flexible surgical instrument in the form of a balloon catheter was developed,188 capable of temperature sensing, mapping of cardiac electrophysiological signals with a high signal-to-noise (SNR) ratio of 60 dB, as well as in situ treatments using RF electrodes (maximum power 600 mW) for controlled, localized tissue ablations (temperature rise ∼10 °C) (Figure 26c). In particular, the use of such instruments in live animal models of rabbits was demonstrated. Harnessing a similar device configuration, integrated catheters for cardiac surgeries were further developed by researchers,135 enabling high-density spatiotemporal mappings of temperature, pressure, and electrophysiological signals by three layers of 8 × 8 electrodes. Such an instrumented catheter is capable of performing programmable electrical stimulation, RF ablation, as well as electroporation (Figure 26d).
6.3. Electromagnetic Devices
6.3.1. 3D Antennas
Flexible 3D antennas with various structural configurations were developed through mechanically-guided methods.121,139,301,322,403,404 For example, a 3D spiral inductor for near-field communication (NFC) was fabricated, exhibiting significantly enhanced Q factors and almost doubled voltages over a wide range of working angles (i.e., from 0° to 50°) when compared to devices in planar forms (Figure 27a).121 Flexible hemispherical ESAs were also manufactured to offer ideal communication performances (Figure 27b).301 In particular, the assembled ESA offered a high radiation efficiency of 83% and a tunable center frequencies upon stretching (i.e., 1.08–0.935 GHz). A morphable device in the form of folded 3D serpentines was also manufactured by the buckling-guided assembly, which can serve as robust antennas for telecommunications (i.e., very slight frequency change from 5.2 to 5.32 GHz during the 2D-to-3D transformation).403 In addition, a bottom-up design strategy based on elementary reconfigurable structures of simple ribbon geometries was also developed, enabling demonstration of multimodal antennas with reconfigurable radiation patterns.139 Furthermore, 3D RF/microwave transformers, consisting of tubular membrane arrays, were fabricated by rolling assembly, featuring an ultracompact device footprint and a performance-scalability that is in stark contrast with conventional 2D devices (i.e., the performance of 3D transformer increases with scaling, while 2D devices do not).322
Figure 27.
3D electromagnetic devices. (a) 3D near-field communication (NFC) device. Reproduced with permission from ref (121) under CC BY. Copyright 2016 National Academy of Sciences. (b) Hemispherical electrically small antenna (ESA) with an outstanding quality factor. Reproduced with permission from ref (301). Copyright 2018Wiley. (c) Rolling assembly of tunable 3D magnetic micro-optics for the manipulation of high-energy electron beam. Reproduced with permission from ref (405) under CC BY. Copyright 2022 Springer Nature. (d) 3D micro-origami sensors capable of static mapping of the magnetic flux and dynamic tracking of magnetic objects in 3D space. Reproduced with permission from ref (98) under CC BY. Copyright 2022 Springer Nature. (e) Fully integrated tubular giant magnetoresistance (GMR) sensor that acts as a fluidic channel for in-flow detection of magnetic particles with high signal-to-noise ratio and sensitivity. Reproduced with permission from ref (148). Copyright 2011 American Chemical Society.
6.3.2. Other Electromagnetic Devices
Other applications of mechanically assembled electromagnetic devices include micro/nano-magnets,405−407 spectroscopies,408 and electromagnetic sensors.98,148,319,409 For instance, 3D magnetically charged particle optics in tubular forms were developed. Particularly, such particle optics can generate alternating magnetic fields of about ±100 mT with frequencies up to a hundred MHz, thereby providing adequate optical powers required for applications like electron beam deflection, focusing, and wavefront shaping (Figure 27c).405 Furthermore, microcoils formed by rolling assembly were also integrated into microfluidic circuits for miniaturized nuclear magnetic resonance (NMR). Specifically, such tubular devices resulted in a reduction of background noise (by effective encapsulation) and a high filling factor of samples in the measurement domain, thereby rendering high-resolution chemical analysis of samples with tiny quantities.408 In another case, folding assembly was utilized to fabricate high-density arrays of cubic magnetic sensors for the measurements of 3D magnetic vector fields.98 Such arrays were also integrated into e-skins with embedded magnetic “hairs”, enabling real-time multidirectional tactile perceptions (Figure 27d). Tubular giant magnetoresistance (GMR) devices were integrated with electronic fluidic systems,319 enabling counting of magnetic objects by sensing their weak magnetic fields (Figure 27e)148 and measurements of viscosity/velocity of fluids.409
6.4. Optoelectronic Devices
6.4.1. 3D Displays
Various flexible 3D displays were manufactured through mechanically-guided assembly.45,297,410−416Figure 28a presents a 3D foldable quantum dot light-emitting diodes (QLEDs) display that showed negligible illumination degradation under cycled folding deformations with extreme small bending radii.297 Ultrathin displays of inorganic light-emitting diodes (LEDs) (i.e., AlInGaP) were manufactured, capable of integration with various unusual substrates, such as papers, Al foils, catheter balloons, and glass/plastic tubes and fibers (Figure 28b).45 Furthermore, a 3D display consisting of a 7 × 7 micro-LED pixel array was manufactured, showing unaffected performance upon biaxial stretching up to 100%.412
Figure 28.
3D optoelectronic devices. (a) Schematic illustration of the 3D foldable QLED display and experimental demonstrations. Reproduced with permission from ref (297). Copyright 2021 Springer Nature. (b) 3D flexible waterproof display composed of a stacked array of inorganic LEDs. Reproduced with permission from ref (45). Copyright 2010 Springer Nature. (c) 3D photodetection and imaging system that can simultaneously measure the light direction, intensity, and the incident angle. Reproduced with permission from ref (217) under CC BY. Copyright 2018 Springer Nature. (d) Residual-stress-induced assembly of 3D pinwheel-like nanostructures for the manipulation of the optical chirality. Reproduced with permission from ref (246) under CC BY. Copyright 2018 The American Association for the Advancement of Science.
Origami and kirigami designs were also exploited to enhance the conformability of 3D flexible displays.410 For example, a kirigami-based wrapping method was developed for the fabrication of conformal electroluminescent (EL) devices on various curved surfaces.320
Morphable 3D displays were also achieved by the use of responsive substrates.411,413 For example, a morphable display was manufactured by integrating electrically actuated substrates with LED arrays/EL materials, exhibiting unvaried illuminations during complex deformations.411 Additionally, by the use of low melting point alloy (LMPA)-graphene nanoplatelets (GNPs)-elastomer composites, 3D morphable displays were also fabricated, featuring rapid electrothermal actuation capabilities and stable illuminations under various 3D configurations.413
6.4.2. 3D Devices for Photodetection and Light Manipulation
Various 3D flexible photodetectors were fabricated by the use of mechanically-guided assembly methods.331,417,418 For example, as shown in Figure 28c, photodetectors in the forms of 3D domes exhibited unique structural superiorities, featuring the simultaneous measurement of both light intensity and incident angles.217 Specifically, given the spatially distributed photoresponsive MoS2 films, an incident light could illuminate different MoS2 films from both the entry and exit sites, thereby resulting in increased photocurrents in the corresponding regions. Through imaging of these photoresponsive regions, the directions of incoming lights can be determined. Additionally, the spatial integration of various optoelectronic elements (e.g., n-channel Si NM MOSFETs, Si NM diodes, and p-channel silicon MOSFETs) with a variety of 3D architectures (e.g., interconnected bridges, coils, and chiral structures) was also demonstrated, showing great potential for multimodal photodetection.249 Furthermore, using folded cubic structures consisting of monolayer MoS2, Au electrodes, and SU-8 supports, photodetectors capable of 3D angle-resolved photodetections were reported.296 Tubular QWIPs were also developed, featuring omnidirectional detections of infrared radiations (e.g., with absorption wavelengths peaking at 3.6 and 6.5 μm) and imaging capabilities through photocurrent mappings.293
Additionally, architected flexible devices were also adopted to manipulate light interactions.312,419,420 For example, a collective coupling of 3D confined optical modes for photodetections was enabled by monolithic twin microtube cavities formed through the rolling assembly of nanomembranes.312 Nanoscale light confinements were also achieved at the hinges within folded cubic or polyhedral structures consisting of poly(methyl methacrylate) (PMMA) supported monolayer graphene.294,421 By the use of residual-stress-induced folding assembly, flexible and morphable nano-kirigami devices were developed for light manipulation (Figure 28d).246
Mechanotunable 3D photonic structures, such as tunable optical filters422 and metasurfaces,419,423 were also enabled by mechanically-guided assembly methods, offering dynamic platforms for light manipulations. For example, through nonlinear buckling mechanics, a large-scale 3D array of plasmonic nanodisks (several square centimeters) on elastomeric substrates was shown to be capable of reversibly shifting optical resonances over 600 nm within near-infrared range.419 Nanophotonic electro-mechanical metasurfaces driven by electrostatic forces were demonstrated, featuring a tunable optical chirality in the near-infrared range with a modulation speed of more than 10 MHz.423
6.4.3. Eyeball Cameras
Cameras in forms of biomimetic 3D eyeballs can significantly improve the imaging performance by offering broadened field of views, reduced distortions, and decreased chromatic aberrations.198,424 Various eyeball cameras manufactured using mechanically-guided assembly methods were reported.100,101,129,130,187,197,246,414,425−427
For example, paraboloid electronic eyeball cameras were fabricated through curving-induced assembly of flexible photodetector arrays on top of hemispherical surfaces, demonstrating improved uniformity of illuminations and well-performed focusing capabilities across a wide field of view, when compared with planar ones.100 Inspired by the exceptionally wide fields of view, infinite depth of field, and high sensitivity to motions of a natural arthropod eye,129 digital cameras in forms of apposition compound eyes were manufactured (Figure 29a), achieving a wider field of view (160°) with much fewer eye units (i.e., 180 units) than those of dragonflies and worker ants. Additionally, different superposition compound eyes (RSCEs) were designed to provide wide spectra of artificial reflections (from infrared to X-ray) with minimum chromatic aberrations, thereby enabling enhanced motion-tracking capabilities and high imaging qualities (i.e., the intensity ratio of the focused image to the background pattern can reach ∼5).425 Notably, flexible photodetector arrays with other structural designs were also exploited to improve the conformability so as to facilitate the manufacturing of eyeball cameras.187
Figure 29.
3D electronic eye cameras. (a) Bioinspired designs of a hemispherical, apposition compound eye camera with wide field of view similar to typical insect. Reproduced with permission from ref (129). Copyright 2013 Springer Nature. (b) A bioinspired camera mimicking aquatic vision with improved sensitivity to incident light. Reproduced with permission from ref (426). Copyright 2020 Springer Nature. (c) Curvy shape-adaptive imagers based on ultrathin Si optoelectronic pixel arrays with a stretchable kirigami design. Reproduced with permission from ref (130). Copyright 2020 Springer Nature.
Apart from electronic compound eyes, eyeball cameras inspired by aquatic visions were reported as well.426,427 For example, inspired by the vision of cichlid fish, miniaturized eyeball cameras (with a diameter of 11.5 mm) were developed, offering a large field of view (120°), a deep depth of field (20 cm), a low optical aberration, and a simple visual accommodation capability (Figure 29b).426 Furthermore, a 3D eyeball camera inspired by the cuttlefish eye, consisting of a W-shaped pupil, a single ball lens, a surface-integrated flexible polarizer, and a highly integrated cylindrical silicon photodiode array that can maintain an average degree of polarization of ∼78% in the visible wavelength range, was able to perform high-quality imaging under uneven illumination conditions.427
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 through curving-induced assembly.101 In particular, these cameras employed deformable photodetector arrays on thin elastomeric membranes, which are capable of pneumatically controlled dynamic curvature adjustments. Additionally, origami197 and kirigami130 designs were also exploited for the manufacturing of hemispherical electronic eyes. For example, shape-adaptive imagers with kirigami designs were fabricated featuring high pixel fill factors (i.e., 78%), unvaried electrical performances under expansion, and focused views of objects at different distances with low aberrations. The adjustable optical focusing power of these electronic eyes ranges from 22.9 to 34.7 dioptres (dpt), surpassing the capability of the human eye (Figure 29c).130
6.5. Energy Devices
6.5.1. Microbatteries
Mechanically-guided 3D assembly has driven the development of various microbatteries313−315,428−431 and supercapacitors.291,432,433 For instance, spiral microelectrodes were prepared by residual-stress-induced rolling for lithium-ion microbatteries.313 In detail, such spiral microbatteries featured tiny dimensions (i.e., a footprint area of around 1 mm2), a relatively high maximum area capacity of 1053 μAh cm–2, an energy density of 12.6 mWh cm–3, as well as a retention of 67% after 50 cycles. By the use of the rolling assembly, MnO2-based microbatteries were prepared (Figure 30a), rendering enhanced overall performances of a footprint area of 0.75 mm2, a reversible area capacity of 1 mAh cm–2, as well as a retention of 50% after 600 cycles.314
Figure 30.
3D energy devices. (a) Microbattery composed of 3D metal layer current collectors formed by a rolling assembly. Reproduced with permission from ref (314) under CC BY. Copyright 2022 Wiley. (b) A “Swiss-roll” microelectrode platform for in situ study of electrical conductivity, electrochemical reactions, and morphology evolution in battery electrodes. Reproduced with permission from ref (430) under CC BY. Copyright 2022 The American Association for the Advancement of Science. (c) 3D piezoelectric microdevice capable of efficiently harvesting low-frequency vibrational energies. Reproduced with permission from ref (132). Copyright 2019 Springer Nature. (d) Wearable energy harvesting system composed of 3D thermoelectric coils formed using buckling-guided assembly. Reproduced with permission from ref (289) under CC BY. Copyright 2018 The American Association for the Advancement of Science. (e) Spherical solar cells. Reproduced with permission from ref (357). Copyright 2009 National Academy of Sciences.
Apart from microbatteries, microelectrodes were also manufactured and utilized as platforms for in situ characterizations of batteries (Figure 30b).430 The elucidations of the role that Fe substitution played for conversion-type NiO battery were showcased as a model study through the monitoring of Fe2O3 evolution and solid electrolyte interphase layer, demonstrating great potentials of using such microelectrodes to probe electrochemistry within batteries.
Tubular supercapacitors were also fabricated by the use of mechanically-guided assembly approaches. For instance, a tubular microsupercapacitor (TMSC)291 was demonstrated, featuring excellent overall performances of a footprint area of less than 0.8 mm2, a capacitance of 88.6 mF cm–2, an energy density of 28.69 mWh cm–2, and a retention of 91.8% after 12,000 cycles.
6.5.2. Energy Harvesters
Transforming piezoelectric microsystems into 3D configurations using mechanically-guided assembly methods can enrich their operational modes133,434,435 as well as offer improved energy harvesting performance.132,436 For example, the integration of piezoelectric elements (i.e., PVDF) with 3D structures of ultralow stiffnesses437,438 allowed for energy harvesters with complex modes of vibrations, thereby achieving an enhanced energy harvesting performance (Figure 30c).132 In particular, different vibrational modes, such as out-of-plane (vertical) and in-plane (lateral) modes, were achieved by the use of suspended 3D serpentine arrays, generating root-mean-square (RMS) voltages ranging from 1 to 2.02 mV under a frequency ranging from 600 to 700 Hz. In this work, a buckled bistable serpentine structure equipped with a proof mass was also manufactured, featuring the generation of electrical powers across a wide range of frequencies, spanning 2 orders of magnitude (from 5 to 500 Hz). Integrating thin-film thermoelectric materials (i.e., doped Si) with compliant 3D architectures allowed for efficient thermal impedance matching and increased power conversion efficiencies.289,350,439 For example, a 3D flexible helical array (i.e., an 8 × 8 helical coils array) was prepared for thermoelectric energy harvesting (Figure 30d).289 In particular, the open-circuit voltage generated by this thermoelectric harvester reached 51.3 mV subjected to a temperature difference of only 19 K, and the measured output power was 2 nW. In another case, a folded 3D photovoltaic (PV) device was capable of efficient light trapping, with a short-circuit current density of 3.6 mA cm–2 and a fill factor of 0.49 (Figure 30e).357 Last by not least, graphene-based hydroelectric generators (GHEGs) in various 3D configurations could realize good energy harvesting performances by harnessing humidity gradients (i.e., with generated voltages of up to 1.5 V under the humidity variation of atmosphere).229
6.6. Robotics
3D electronic fliers hold great potential for environmental monitoring, population surveillance, and sensing applications that require volumetric coverage in 3D spaces. Inspired by wind-dispersed seeds, 3D macro-, meso-, and microscale electronic fliers were developed through buckling-guided assembly approaches, featuring controlled, rotational kinematics, and low terminal velocities.22 Incorporations of active electronic and colorimetric materials on top of such 3D fliers allowed battery-free, wireless devices and colorimetric sensors for environmental analysis, such as measurements of particle matter (PM) pollution and pH values in the atmosphere (Figure 31a). Derived from the above, biodegradable 3D colorimetric fliers were also fabricated, allowing remote assessments of multiple environmental parameters such as pH values, heavy metal concentrations, ultraviolet exposures, humidity levels, as well as temperatures.23 Furthermore, by integrating responsive materials (e.g., SMP) with such electronic fliers, morphable 3D mesofliers were developed, featuring large degrees of actuation deformations with a fast response time of 1.08 s.131
Figure 31.
3D fliers and aquatic robots. (a) Seed-inspired 3D electronic fliers with an integrated microsystem or a transistor, which exhibit stable rotational falling in the air. Reproduced with permission from ref (22). Copyright 2021 Springer Nature. (b) Wirelessly controlled swimming microrobot composed of rolled twin-jet-engines. Reproduced with permission from ref (15). Copyright 2020 Springer Nature. (c) Laser controlled microrobots consisting of 3D electrochemical actuators formed by using the rolling assembly. Reproduced with permission from ref (18). Copyright 2020 Springer Nature.
6.6.1. Electronic Fliers
6.6.2. Aquatic Robots
Untethered aquatic robotic systems have important applications in biomedical fields, such as lab-on-a-chip devices, minimally invasive surgical interventions, among others.318 For instance, through the rolling assembly associated with asymmetric designs of tilted head-flagellum geometries, arc-shaped microswimmers were manufactured, featuring bidirectional motions without the requirement of a reversal in the direction of the applied rotational magnetic field.440 Additionally, by the use of morphable SMP substrates, aquatic robots were manufactured,226 featuring remote control of switchable swimming modes through thermomechanical stimulations. Wirelessly powered flexible microjet systems with controlled locomotion modes were also developed (Figure 31b).15 In particular, such systems can be remotely powered by inductive coupling, and their motions were driven by controlled catalytic reactions. Additionally, through the development of a novel type of voltage-driven foldable surface electrochemical actuators, a series of subhundred-micrometer robots were manufactured (Figure 31c).18,441 Remarkably, the fabrication process was highly compatible with planar techniques, rendering the mass production of over one million robots on a 4 in. Si wafer.
6.6.3. Terrestrial Robots
Various terrestrial robots with diverse locomotion modes were developed by using mechanically-guided assembly methods. For example, submillimeter robots of various 3D geometries were developed17 through the assembly of a patterned heterogeneous membrane consisting of a PI skeleton, SMA actuators, and a SiO2 elastic encapsulation shell (Figure 32a). Owing to the rich diversity of structural designs, such robots could enable different locomotion modes under laser control, including crawling, walking, turning, and jumping. Soft microrobots, with a broad spectrum of architected 3D configurations and stiffness adjusting capabilities, were also manufactured through the buckling-guided assembly of LCE composites (i.e., enabling large degrees of 3D-to-3D shape morphing) (Figure 32b).16 In particular, such soft microrobots were equipped with morphable electroadhesive footpads (Figure 32b, bottom left) and stiffness-variable smart joints, enabling different locomotion modes in a single microrobot, including climbing on complexly curved walls (e.g., wedge-shaped, cylindrical, spherical, and wavy surfaces) as well as transitioning between two distinct surfaces. Through the use of the roll assembly, electromagnetically controlled soft robots were manufactured, capable of walking, running, swimming, jumping, steering, and transporting cargos (Figure 32c).442 In addition, origami-inspired designs were also introduced to fabricate robots through the folding assembly, rendering controlled transformations from flat sheets into functional robots.175,443 Furthermore, insect-scale fast-moving and ultrarobust soft robots were also developed by curving-induced assembly of unimorph piezoelectric structures, exhibiting a high moving speed of 20 body lengths per second.19
Figure 32.
3D terrestrial robots. (a) Submillimeter-scale terrestrial robots with heterogeneously integrated 3D mesostructures with a patterned SMA, a PI skeleton, and a SiO2 shell. Reproduced with permission from ref (17). Copyright 2022 The American Association for the Advancement of Science. (b) Soft climbing microrobot capable of climbing on and transitioning between different curved surfaces. Reproduced with permission from ref (16) under CC BY. Copyright 2022 National Academy of Sciences. (c) Amphibious soft electromagnetic robot capable of walking, running, jumping, and swimming. Reproduced with permission from ref (442) under CC BY. Copyright 2022 Springer Nature.
7. Conclusions and Outlooks
During the last two decades, the field of flexible/stretchable electronics has been growing very rapidly, sparking significant interests in the contemporary society, spanning diverse industrial and research areas (e.g., IoT, daily healthcare, robotics, fundamental biology, medicine, energy, sensors, etc.). Originated by pioneer works in organic electronics1 and structurally engineered inorganic electronics,103 flexible/stretchable electronics well pronounce the future needs and directions of electronic devices from divergent points of views. Fruitful research works have been conducted, demonstrating plentiful 3D flexible electronic devices,34,35,106 fabrication protocols for functional circuits on 3D curved surfaces,102,138,201 stretchable standalone device platforms,17,203,228 and so on. Among the various promising directions of the vibrant field, 3D flexible electronics represent a core branch, owing to their unique structure-induced functionalities, expanded design freedoms, and capabilities of better conforming to or replicating complexly shaped biological objects, when compared with planar counterparts. Compatible with well-established fabrication techniques of planar inorganic flexible electronics,35,104,106 mechanically-guided 3D assembly methods stand as powerful tools for the manufacturing of architected 3D flexible electronics with precisely engineered structural configurations and functionalities.138,300 Though the primary goal of mechanically-guided assembly methods (i.e., rolling, folding, curving and buckling) is to manufacture miniaturized flexible electronic devices from nano- to millimeter-scale, many examples have showcased that the mechanically-guided assembly methods, such as the rolling, folding and buckling, are capable of parallel manufacturing of multiple devices/systems/structures in one round.96,119,249,444 Notably, herein, to establish a clear material–fabrication–performance–application relationship, a detailed comparison table (Table 2) summarizing the used assembly methods, device forms, structure-induced functionalities, superiorities comparing to their planar counterparts, and applications of various architected 3D flexible devices is presented. Despite significant progress summarized above, rich opportunities exist in this burgeoning area of architected 3D flexible electronics, as detailed below.
Table 2. Summary of Representative Architected Flexible Devices Prepared through Mechanically-Guided 3D Assembly Methods.
Applications | Devices | Assembly methods | Device forms | Structure-induced functionalities | Superiorities comparing to planar counterparts | Reference |
---|---|---|---|---|---|---|
Cell devices | Inter/intracellular recorders | Buckling | 3D arc-shaped | Vertically directed sensing diodes with predesigned locations | High density measurements of transmembrane potential at multiple locations | (325) |
Single cell microgrippers | Folding | Polyhedral | Conformal cell capture enabled by folded gripper arms | Nondestructive and real-time analysis of 3D surface of cell | (184, 372) | |
Organoid devices | 3D electronic frameworks | Buckling | Conformally wrapping | Compliant 3D electronic neural interface in conformal geometries | 3D spatiotemporal mapping of spontaneous neural activities | (127) |
Shell microelectrode arrays | Folding | Polyhedral | 3D conformal shells with tunable contact area between electrode and target | 3D spatiotemporal recording of encapsulated organoids with large contact area | (128) | |
Electronic tissue scaffolds | Spherical cap-shaped electronic cell scaffold | Buckling | Hemispherical | Creation of biomimetic 3D microenvironments | Noninvasive real-time investigations of cell activities | (138) |
Tubular scaffolds | Rolling | Tubular | Biomimetic tubular structures with multilayer walls | Enhanced viability and precise multicellular layering of artificial arteries | (375) | |
In situ monitoring devices | Piezoelectric microsystems | Buckling | 3D arc-shaped | 3D compliant thin piezoelectric structure with high sensitivity to external forces | Enhanced signals with high reproducibility for differentiating various motion states | (132) |
Climbing-inspired twining electrodes | Rolling | 3D helical | Twining electrode capable of self-climbing along nerves | Conformal neural interface for electrical stimulation and recording | (310) | |
3D multifunctional membranes | Curving | Conformal curvilinear | 3D conformal electronic membranes | Epicardial mapping and stimulation with a mechanically stable interface during normal cardiac cycles | (191) | |
In situ therapeutic devices | “Hookworms” microgrippers | Folding | Worms-mouthparts-shaped | Foldable sharp microtips that ensure effective latching onto the GI mucosa | Enhanced drug delivery performances enabled by controlled long-term residency | (134) |
Pain-blocking microfluidic devices | Rolling | 3D helical | Soft, curled, nerve-clasping microfluidic channels with secure sutureless attachment to nerves | Long-term in situ pain relief through neural cooling | (388) | |
Surgical instruments | Instrumented catheter | Curving | Complex curvilinear | 3D conformal electronic interfaces | High-density spatiotemporal mappings of multiple biological signals and surgical operations | (135, 158) |
3D electromagnetic devices | Hemispherical/spherical ESAs | Buckling | Hemispherical | High volume occupation of the Chu-sphere | Wireless communication with enhanced Q factors and broadened working angles | (301, 404) |
3D magnetically charged particle optics | Rolling | Tubular | Self-rolling tubular microsized electromagnets | Fast beam manipulation | (405) | |
Cubic magnetic sensors | Folding | Polyhedral | Spatial rearrangement of sensing elements through self-folding | Simultaneous measurements of three magnetic field components in 3D space | (98) | |
3D displays | Foldable quantum dot light-emitting | Folding | Polyhedral | 3D foldable QLEDs in origami forms enabled by customized creases | 3D reconfigurable displays | (297) |
Ultrathin inorganic light-emitting diodes | Curving | Complex curvilinear | 3D arrangement of multilayer LED arrays | 3D multilayer displays that can be integrated on various unusual substrates | (45) | |
3D devices for photodetection and Light manipulation | 3D photodetectors | Buckling | Hemispherical | Spatial arrangement of photodetectors | Simultaneous measurements of the direction, intensity and angular divergence of incident light | (217) |
Nano-kirigami metasurface | Folding | Kirigami | Out-of-plane configurations of plasmonic materials | Giant intrinsic optical chirality | (246) | |
Eyeball cameras | Compound eyes cameras | Curving | Hemispherical | Bioinspired spatial arrangements of optical elements on hemispherical surfaces | Wide field of view, low levels of chromatic aberrations and deep depth of field | (129) |
Aquatic eyeball cameras | Curving | Hemispherical | (426, 427) | |||
Tunable hemispherical electronic eye camera | Curving | Hemispherical | Tunable ultrathin kirigami membranes consisting of optoelectronic pixels | Imaging of objects at different distances with low aberration | (130) | |
Microbatteries | Spiral microbatteries and microelectrodes | Rolling | Tubular | 3D stacking of electrode materials | High loadings of electrode materials with reduced footprint area | (313, 314) |
Energy harvesters | Piezoelectric element 3D structures | Buckling | 3D arc-shaped | 3D compliant piezoelectric structure with multiple vibration modes | High power output under wide range of frequencies | (132) |
3D flexible thermoelectric helical array | Buckling | 3D helical | Enlarged temperature gradients through creations of altitude intercept | Enhanced heat exchange efficiency and output power | (289) | |
Robotics | Electronic fliers | Buckling | Shapes resembling wind-dispersed seeds | Rotational falling and low terminal velocities | Flying with controlled rotational kinematics and enhanced stability | (22) |
Aquatic robots | Rolling | Tubular | Construction of tubular jet propulsion engines | Untethered multimodal swimming in a controlled manner | (15) | |
Wall-climbing robot | Buckling | 3D arc-shaped | Reconfigurable 3D structures capable of large deformations | Multiple locomotion modes | (16) |
7.1. Routes to Extremes: Deformations and Dimensions
Through creative uses of the mechanics of materials and structures, mechanically-guided assembly methods transform the as-prepared planar electronic devices into deterministic 3D architectures, achieving many previously inaccessible (or hard-to-achieve) structure-induced functionalities such as the 3D integration with programmable spatial resolution, omnidirectional photoenergy harvesting, and morphological/structural evolution.
Two core aspects of mechanically-guided methods for 3D device assembly are how complicated the accessible deformations can be (i.e., complexity and diversity of deformations) and to what extent each deformation mode can reach (i.e., magnitudes of deformations). Although many basic deformation modes (i.e., bending, folding, twisting, shearing, and buckling) are already exploited in mechanically-guided assembly methods, combined deformation modes applied in desired sequences are rarely achieved due to technical challenges to implement the mechanical loadings experimentally. In addition, the magnitudes of each deformation mode are not yet pushed to extreme. For example, an extreme rolling process might result in tubular structures with ultracompact walls and supersmall diameters of only several manometers; a buckling-guided assembly with huge biaxial prestrains (e.g., 1000%) might generate a device with extreme expansion ratios. To achieve extreme deformations in mechanically-guided 3D assembly, developments of experimental platforms with powerful loading capabilities as well as artificial intelligence boosted topology optimization methods of 2D precursors and substrates are promising to explore.
The manufacturing of well-defined 3D nanostructures with unique physical and chemical properties stands as the long-term goal for nanofabrication methods, and mechanically-guided assembly methods are no exception. Although several initial trials have been conducted using mechanically-guided assembly, only 3D nanostructures with very simple geometric configurations were formed.126,221,242 Therefore, developments of nanoscale assembly approaches that allow the manufacturing of 3D nanoarchitectures with complex geometric configurations stand as a vital direction in this field. Such fabrication capabilities with nanoscale precision are expected to give rise to architected flexible electronic devices with fundamentally new functions.
7.2. Inverse Design Methods
Powerful inverse design methods that can rapidly map target 3D geometries onto optimized 2D precursor patterns and give required loading forms/magnitudes are foundational to the widespread utility of mechanically-guided assembly methods. Very recently, by introducing microlattice designs and tuning their local stiffnesses through a tailored porosity distribution, an inverse design method assisted by machine learning algorithms was developed for the buckling-guided assembly, capable of replicating many 3D curved surfaces in biology.300 However, the nature of the 2D-to-3D assembly and in-plane loadings set certain limitations to the range of accessible 3D geometries. Plenty of open opportunities lie in devising novel inverse design strategies with the assistance of artificial intelligence under the framework of mechanically-guided methods. For example, the two-level inverse design that involves concurrent optimization of kirigami cuts at a global level and microlattice patterns at a local level can well expand the range of inversely designed 3D geometries and is very exciting to explore.
7.3. Encapsulation Strategies
Encapsulation is the key to ensuring the secure and stable operation of flexible electronic devices. The existing encapsulation methods of flexible electronics mostly exploit polymeric materials such as PI and silicone elastomers (e.g., PDMS). These encapsulations are simple and efficient for planar flexible electronics. When turning to those emerging architected 3D flexible electronics that have certain components popping up from the substrate surface,238,240,280 such encapsulation methods are no longer applicable. For example, to allow free out-of-plane deformations of serpentine interconnects, an encapsulation approach harnessing the excellent stretchability of soft network materials (i.e., both uni- and biaxial) can be exploited.333,445−447 However, the reported encapsulation strategies are just the tip of the whole iceberg. Exciting opportunities lie in the developments of new encapsulation strategies that pose significantly reduced mechanical constraints to the functional 3D architectures while providing effective protection.
7.4. Applications
Flexible electronics assembled using mechanically-guided approaches have already found applications in diverse fields, including biology, medicine, sensing, energy, robotics, and so on. On the one hand, the exploration of new application scenarios and areas (e.g., catalysis, deep space exploration, and nonlinear optics) of architected 3D flexible devices stands as a mainstream research direction in the field. On the other hand, the discovery of more structure-induced functionalities is instrumental to expand the applications of architected flexible electronics. For example, the incorporation of rationally distributed local metasurfaces in a designed 3D architecture to achieve controlled coupling of electromagnetic waves into desired functional elements is worthy of investigation.
Acknowledgments
Y.Z. acknowledges support from the National Natural Science Foundation of China (Grant Nos. 12225206, 12050004, and 11921002), the Tsinghua National Laboratory for Information Science and Technology, a grant from the Institute for Guo Qiang, Tsinghua University (Grant No. 2021GQG1009), and the New Cornerstone Science Foundation through the XPLORER PRIZE. R.B. gratefully acknowledges support from the National Natural Science Foundation of China (Grant No. 12102221), and China Postdoctoral Science Foundation for the fellowships (Grant Nos. 2021M691795 and 2022T150361).
Glossary
List of Abbreviations
- PDMS
Polydimethylsiloxane
- LCE
Liquid crystal elastomer
- SMA
Shape memory alloy
- FIB
Focused ion beam
- SMP
Shape memory polymer
- TFFA
Thin film-shaped flexible actuator
- PI
Polyimide
- PVDF
Polyvinylidene fluoride
- PZT
Lead zirconium titanate
- CN
Carbon nanotube
- RF
Radiofrequency
- FEA
Finite element analyse
- QWIP
Quantum well infrared photodetector
- nBSC
Nano-biosupercapacitor
- QLED
Quantum dot light-emitting diode
- ESA
Electrically small antenna
- IoT
Internet of Things
- VOC
Volatile organic compound
- PEC
Photoelectrochemical
- FET
Field-effect transistor
- MT-SERS
Mechanical trapping and surface-enhanced Raman spectroscopy
- MEA
Microelectrode array
- DRG
Dorsal root ganglion
- BAEP
Bioartificial endocrine pancreas
- RPE
Retinal pigment epithelium
- GI
Gastrointestinal
- NFC
Near-field communication
- NMR
Nuclear magnetic resonance
- GMR
Giant magnetoresistance
- EL
Electroluminescent
- LMPA
Low melting point alloy
- GNP
Graphene nanoplatelet
- PMMA
Poly(methyl methacrylate)
- RSCE
Reflecting superposition compound eye
- TMSC
Tubular microsupercapacitor
- RMS
Root-mean-square
- PV
Photovoltaic
- GHEG
Graphene-based hydroelectric generator
- PM
Particle matter
Biographies
Renheng Bo obtained his Ph.D. (2020) in Materials Science and Nanotechnology from Research School of Engineering at the Australian National University. He is currently a postdoctoral fellow at Tsinghua University, with research interests involving multifunctional architected flexible electronics, biomedical devices, and mechanically-guided 3D assembly.
Shiwei Xu obtained his B.S. degree in Engineering Mechanics from Huazhong University of Science and Technology, China, in 2020. He is currently pursuing his Ph.D. degree in Solid Mechanics at Tsinghua University, focusing on mechanically-guided 3D assembly for soft robotics and reconfigurable electronics, under the supervision of Prof. Yihui Zhang.
Youzhou Yang obtained both his B.S. degree in Electronic Science and Technology and M.E. in Microelectronics and Solid State Electronics from Huazhong University of Science and Technology, China, in 2018 and 2022, respectively. Under the supervision of Prof. Yihui Zhang, he is currently pursuing his Ph.D. degree in Solid Mechanics at Tsinghua University, focusing on mechanically-guided 3D assembly for magnetically controlled microscale biorobotics and organoid/tissues engineering.
Yihui Zhang obtained his Ph.D. (2011) in Solid Mechanics from the Department of Engineering Mechanics at Tsinghua University. 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 the Laboratory of Flexible Electronics Technology at Tsinghua University. His research interests include mechanically-guided 3D assembly, unusual soft composite materials, and stretchable electronics.
Author Contributions
§ R.B., S.X., and Y.Y contributed equally to this work. CRediT: Renheng Bo visualization, writing-original draft, writing-review & editing; Shiwei Xu visualization, writing-original draft; Youzhou Yang visualization, writing-original draft; Yihui Zhang supervision, writing-review & editing.
The authors declare no competing financial interest.
Special Issue
Published as part of the Chemical Reviewsvirtual special issue “Wearable Devices”.
References
- Garnier F.; Hajlaoui R.; Yassar A.; Srivastava P. All-Polymer Field-Effect Transistor Realized by Printing Techniques. Science 1994, 265, 1684–1686. 10.1126/science.265.5179.1684. [DOI] [PubMed] [Google Scholar]
- Rogers J. A.; Someya T.; Huang Y. Materials and Mechanics for Stretchable Electronics. Science 2010, 327, 1603–1607. 10.1126/science.1182383. [DOI] [PubMed] [Google Scholar]
- Chung H. U.; Kim B. H.; Lee J. Y.; Lee J.; Xie Z.; Ibler E. M.; Lee K.; Banks A.; Jeong J. Y.; Kim J.; et al. Binodal, Wireless Epidermal Electronic Systems with in-Sensor Analytics for Neonatal Intensive Care. Science 2019, 363, eaau0780 10.1126/science.aau0780. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen G.; Li Y.; Bick M.; Chen J. Smart Textiles for Electricity Generation. Chem. Rev. 2020, 120, 3668–3720. 10.1021/acs.chemrev.9b00821. [DOI] [PubMed] [Google Scholar]
- Chen G.; Xiao X.; Zhao X.; Tat T.; Bick M.; Chen J. Electronic Textiles for Wearable Point-of-Care Systems. Chem. Rev. 2022, 122, 3259–3291. 10.1021/acs.chemrev.1c00502. [DOI] [PubMed] [Google Scholar]
- Wang C.; Chen X.; Wang L.; Makihata M.; Liu H.-C.; Zhou T.; Zhao X. Bioadhesive Ultrasound for Long-Term Continuous Imaging of Diverse Organs. Science 2022, 377, 517–523. 10.1126/science.abo2542. [DOI] [PubMed] [Google Scholar]
- Hu H.; Ma Y.; Gao X.; Song D.; Li M.; Huang H.; Qian X.; Wu R.; Shi K.; Ding H.; et al. Stretchable Ultrasonic Arrays for the Three-Dimensional Mapping of the Modulus of Deep Tissue. Nat. Biomed. Eng. 2023. 10.1038/s41551-023-01038-w [DOI] [PubMed] [Google Scholar]
- Kim J.-H.; Marcus C.; Ono R.; Sadat D.; Mirzazadeh A.; Jens M.; Fernandez S.; Zheng S.; Durak T.; Dagdeviren C. A Conformable Sensory Face Mask for Decoding Biological and Environmental Signals. Nat. Electron. 2022, 5, 794–807. 10.1038/s41928-022-00851-6. [DOI] [Google Scholar]
- Jiang S.; Liu J.; Xiong W.; Yang Z.; Yin L.; Li K.; Huang Y. A Snakeskin-Inspired, Soft-Hinge Kirigami Metamaterial for Self-Adaptive Conformal Electronic Armor. Adv. Mater. 2022, 34, 2204091. 10.1002/adma.202204091. [DOI] [PubMed] [Google Scholar]
- Awad L. N.; Bae J.; O’Donnell K.; De Rossi S. M. M.; Hendron K.; Sloot L. H.; Kudzia P.; Allen S.; Holt K. G.; Ellis T. D.; et al. A Soft Robotic Exosuit Improves Walking in Patients after Stroke. Sci. Transl. Med. 2017, 9, eaai9084 10.1126/scitranslmed.aai9084. [DOI] [PubMed] [Google Scholar]
- Yang Y.; Song Y.; Bo X.; Min J.; Pak O. S.; Zhu L.; Wang M.; Tu J.; Kogan A.; Zhang H.; et al. A Laser-Engraved Wearable Sensor for Sensitive Detection of Uric Acid and Tyrosine in Sweat. Nat. Biotechnol. 2020, 38, 217–224. 10.1038/s41587-019-0321-x. [DOI] [PubMed] [Google Scholar]
- Wang M.; Yang Y.; Min J.; Song Y.; Tu J.; Mukasa D.; Ye C.; Xu C.; Heflin N.; McCune J. S.; et al. A Wearable Electrochemical Biosensor for the Monitoring of Metabolites and Nutrients. Nat. Biomed. Eng. 2022, 6, 1225–1235. 10.1038/s41551-022-00916-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wu S.; Moody K.; Kollipara A.; Zhu Y. Highly Sensitive, Stretchable, and Robust Strain Sensor Based on Crack Propagation and Opening. ACS Appl. Mater. Interfaces. 2023, 15, 1798–1807. 10.1021/acsami.2c16741. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hu H.; Zhu X.; Wang C.; Zhang L.; Li X.; Lee S.; Huang Z.; Chen R.; Chen Z.; Wang C.; et al. Stretchable Ultrasonic Transducer Arrays for Three-Dimensional Imaging on Complex Surfaces. Sci. Adv. 2018, 4, eaar3979 10.1126/sciadv.aar3979. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bandari V. K.; Nan Y.; Karnaushenko D.; Hong Y.; Sun B.; Striggow F.; Karnaushenko D. D.; Becker C.; Faghih M.; Medina-Sánchez M.; et al. A Flexible Microsystem Capable of Controlled Motion and Actuation by Wireless Power Transfer. Nat. Electron. 2020, 3, 172–180. 10.1038/s41928-020-0384-1. [DOI] [Google Scholar]
- Pang W.; Xu S.; Wu J.; Bo R.; Jin T.; Xiao Y.; Liu Z.; Zhang F.; Cheng X.; Bai K.; et al. A Soft Microrobot with Highly Deformable 3d Actuators for Climbing and Transitioning Complex Surfaces. Proc. Natl. Acad. Sci. U.S.A. 2022, 119, e2215028119 10.1073/pnas.2215028119. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Han M.; Guo X.; Chen X.; Liang C.; Zhao H.; Zhang Q.; Bai W.; Zhang F.; Wei H.; Wu C.; et al. Submillimeter-Scale Multimaterial Terrestrial Robots. Sci. Robot. 2022, 7, eabn0602 10.1126/scirobotics.abn0602. [DOI] [PubMed] [Google Scholar]
- Reynolds M. F.; Cortese A. J.; Liu Q.; Zheng Z.; Wang W.; Norris S. L.; Lee S.; Miskin M. Z.; Molnar A. C.; Cohen I.; et al. Microscopic Robots with Onboard Digital Control. Sci. Robot. 2022, 7, eabq2296 10.1126/scirobotics.abq2296. [DOI] [PubMed] [Google Scholar]
- Wu Y.; Yim J. K.; Liang J.; Shao Z.; Qi M.; Zhong J.; Luo Z.; Yan X.; Zhang M.; Wang X.; et al. Insect-Scale Fast Moving and Ultrarobust Soft Robot. Sci. Robot. 2019, 4, eaax1594 10.1126/scirobotics.aax1594. [DOI] [PubMed] [Google Scholar]
- Li S.; Xu J.; Li R.; Wang Y.; Zhang M.; Li J.; Yin S.; Liu G.; Zhang L.; Li B.; et al. Stretchable Electronic Facial Masks for Sonophoresis. ACS Nano 2022, 16, 5961–5974. 10.1021/acsnano.1c11181. [DOI] [PubMed] [Google Scholar]
- Yu C.-C.; Shah A.; Amiri N.; Marcus C.; Nayeem M. O. G.; Bhayadia A. K.; Karami A.; Dagdeviren C. A Conformable Ultrasound Patch for Cavitation-Enhanced Transdermal Cosmeceutical Delivery. Adv. Mater. 2023, 35, 2300066. 10.1002/adma.202370160. [DOI] [PubMed] [Google Scholar]
- Kim B. H.; Li K.; Kim J. T.; Park Y.; Jang H.; Wang X.; Xie Z.; Won S. M.; Yoon H. J.; Lee G.; et al. Three-Dimensional Electronic Microfliers Inspired by Wind-Dispersed Seeds. Nature 2021, 597, 503–510. 10.1038/s41586-021-03847-y. [DOI] [PubMed] [Google Scholar]
- Yoon H.-J.; Lee G.; Kim J.-T.; Yoo J.-Y.; Luan H.; Cheng S.; Kang S.; Huynh H. L. T.; Kim H.; Park J.; et al. Biodegradable, Three-Dimensional Colorimetric Fliers for Environmental Monitoring. Sci. Adv. 2022, 8, eade3201 10.1126/sciadv.ade3201. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Savagatrup S.; Printz A. D.; O’Connor T. F.; Zaretski A. V.; Lipomi D. J. Molecularly Stretchable Electronics. Chem. Mater. 2014, 26, 3028–3041. 10.1021/cm501021v. [DOI] [Google Scholar]
- Root S. E.; Savagatrup S.; Printz A. D.; Rodriquez D.; Lipomi D. J. Mechanical Properties of Organic Semiconductors for Stretchable, Highly Flexible, and Mechanically Robust Electronics. Chem. Rev. 2017, 117, 6467–6499. 10.1021/acs.chemrev.7b00003. [DOI] [PubMed] [Google Scholar]
- Choi S.; Han S. I.; Kim D.; Hyeon T.; Kim D. H. High-Performance Stretchable Conductive Nanocomposites: Materials, Processes, and Device Applications. Chem. Soc. Rev. 2019, 48, 1566–1596. 10.1039/C8CS00706C. [DOI] [PubMed] [Google Scholar]
- Park S.; Heo S. W.; Lee W.; Inoue D.; Jiang Z.; Yu K.; Jinno H.; Hashizume D.; Sekino M.; Yokota T.; et al. Self-Powered Ultra-Flexible Electronics Via Nano-Grating-Patterned Organic Photovoltaics. Nature 2018, 561, 516–521. 10.1038/s41586-018-0536-x. [DOI] [PubMed] [Google Scholar]
- Wang S.; Xu J.; Wang W.; Wang G. N.; Rastak R.; Molina-Lopez F.; Chung J. W.; Niu S.; Feig V. R.; Lopez J.; et al. Skin Electronics from Scalable Fabrication of an Intrinsically Stretchable Transistor Array. Nature 2018, 555, 83–88. 10.1038/nature25494. [DOI] [PubMed] [Google Scholar]
- Zheng Y.; Zhang S.; Tok J. B.; Bao Z. Molecular Design of Stretchable Polymer Semiconductors: Current Progress and Future Directions. J. Am. Chem. Soc. 2022, 144, 4699–4715. 10.1021/jacs.2c00072. [DOI] [PubMed] [Google Scholar]
- Nawaz A.; Merces L.; Ferro L. M. M.; Sonar P.; Bufon C. C. B. Impact of Planar and Vertical Organic Field-Effect Transistors on Flexible Electronics. Adv. Mater. 2023, 35, e2204804 10.1002/adma.202370077. [DOI] [PubMed] [Google Scholar]
- Khang D. Y.; Jiang H.; Huang Y.; Rogers J. A. A Stretchable Form of Single-Crystal Silicon for High-Performance Electronics on Rubber Substrates. Science 2006, 311, 208–212. 10.1126/science.1121401. [DOI] [PubMed] [Google Scholar]
- Kim D. H.; Xiao J.; Song J.; Huang Y.; Rogers J. A. Stretchable, Curvilinear Electronics Based on Inorganic Materials. Adv. Mater. 2010, 22, 2108–2124. 10.1002/adma.200902927. [DOI] [PubMed] [Google Scholar]
- Rafeedi T.; Lipomi D. J. Multiple Pathways to Stretchable Electronics. Science 2022, 378, 1174–1175. 10.1126/science.adf2322. [DOI] [PubMed] [Google Scholar]
- Wang C.; Wang C.; Huang Z.; Xu S. Materials and Structures toward Soft Electronics. Adv. Mater. 2018, 30, 1801368 10.1002/adma.201801368. [DOI] [PubMed] [Google Scholar]
- Xue Z.; Song H.; Rogers J. A.; Zhang Y.; Huang Y. Mechanically-Guided Structural Designs in Stretchable Inorganic Electronics. Adv. Mater. 2020, 32, 1902254. 10.1002/adma.201902254. [DOI] [PubMed] [Google Scholar]
- Bowden N.; Brittain S.; Evans A. G.; Hutchinson J. W.; Whitesides G. M. Spontaneous Formation of Ordered Structures in Thin Films of Metals Supported on an Elastomeric Polymer. Nature 1998, 393, 146–149. 10.1038/30193. [DOI] [Google Scholar]
- Sun Y.; Choi W. M.; Jiang H.; Huang Y. Y.; Rogers J. A. Controlled Buckling of Semiconductor Nanoribbons for Stretchable Electronics. Nat. Nanotechnol. 2006, 1, 201–207. 10.1038/nnano.2006.131. [DOI] [PubMed] [Google Scholar]
- Sun Y. G.; Kumar V.; Adesida I.; Rogers J. A. Buckled and Wavy Ribbons of Gaas for High-Performance Electronics on Elastomeric Substrates. Adv. Mater. 2006, 18, 2857–2862. 10.1002/adma.200600646. [DOI] [Google Scholar]
- Ko H. C.; Stoykovich M. P.; Song J.; Malyarchuk V.; Choi W. M.; Yu C. J.; Geddes J. B. 3rd; Xiao J.; Wang S.; Huang Y.; et al. A Hemispherical Electronic Eye Camera Based on Compressible Silicon Optoelectronics. Nature 2008, 454, 748–753. 10.1038/nature07113. [DOI] [PubMed] [Google Scholar]
- Kim D.-H.; Song J.; Choi W. M.; Kim H.-S.; Kim R.-H.; Liu Z.; Huang Y. Y.; Hwang K.-C.; Zhang Y.-w.; Rogers J. A. Materials and Noncoplanar Mesh Designs for Integrated Circuits with Linear Elastic Responses to Extreme Mechanical Deformations. Proc. Natl. Acad. Sci. U.S.A. 2008, 105, 18675–18680. 10.1073/pnas.0807476105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brosteaux D.; Axisa F.; Gonzalez M.; Vanfleteren J. Design and Fabrication of Elastic Interconnections for Stretchable Electronic Circuits. IEEE Electron Device Lett. 2007, 28, 552–554. 10.1109/LED.2007.897887. [DOI] [Google Scholar]
- Fan Z.; Zhang Y.; Ma Q.; Zhang F.; Fu H.; Hwang K. C.; Huang Y. A Finite Deformation Model of Planar Serpentine Interconnects for Stretchable Electronics. Int. J. Solids Struct. 2016, 91, 46–54. 10.1016/j.ijsolstr.2016.04.030. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jeong J. W.; Yeo W. H.; Akhtar A.; Norton J. J.; Kwack Y. J.; Li S.; Jung S. Y.; Su Y.; Lee W.; Xia J.; et al. Materials and Optimized Designs for Human-Machine Interfaces Via Epidermal Electronics. Adv. Mater. 2013, 25, 6839–6846. 10.1002/adma.201301921. [DOI] [PubMed] [Google Scholar]
- Kim D.-H.; Kim Y.-S.; Wu J.; Liu Z.; Song J.; Kim H.-S.; Huang Y. Y.; Hwang K.-C.; Rogers J. A. Ultrathin Silicon Circuits with Strain-Isolation Layers and Mesh Layouts for High-Performance Electronics on Fabric, Vinyl, Leather, and Paper (Adv. Mater. 36/2009). Adv. Mater. 2009, 21, NA–NA. 10.1002/adma.200990139. [DOI] [Google Scholar]
- Kim R. H.; Kim D. H.; Xiao J.; Kim B. H.; Park S. I.; Panilaitis B.; Ghaffari R.; Yao J.; Li M.; Liu Z.; et al. Waterproof Alingap Optoelectronics on Stretchable Substrates with Applications in Biomedicine and Robotics. Nat. Mater. 2010, 9, 929–937. 10.1038/nmat2879. [DOI] [PubMed] [Google Scholar]
- Shyu T. C.; Damasceno P. F.; Dodd P. M.; Lamoureux A.; Xu L.; Shlian M.; Shtein M.; Glotzer S. C.; Kotov N. A. A Kirigami Approach to Engineering Elasticity in Nanocomposites through Patterned Defects. Nat. Mater. 2015, 14, 785–789. 10.1038/nmat4327. [DOI] [PubMed] [Google Scholar]
- Bertoldi K.; Vitelli V.; Christensen J.; van Hecke M. Flexible Mechanical Metamaterials. Nat. Rev. Mater. 2017, 2, 17066. 10.1038/natrevmats.2017.66. [DOI] [Google Scholar]
- Blees M. K.; Barnard A. W.; Rose P. A.; Roberts S. P.; McGill K. L.; Huang P. Y.; Ruyack A. R.; Kevek J. W.; Kobrin B.; Muller D. A.; et al. Graphene Kirigami. Nature 2015, 524, 204–207. 10.1038/nature14588. [DOI] [PubMed] [Google Scholar]
- Coulais C.; Sabbadini A.; Vink F.; van Hecke M. Multi-Step Self-Guided Pathways for Shape-Changing Metamaterials. Nature 2018, 561, 512–515. 10.1038/s41586-018-0541-0. [DOI] [PubMed] [Google Scholar]
- Eidini M. Zigzag-Base Folded Sheet Cellular Mechanical Metamaterials. EXTREME MECH. LETT. 2016, 6, 96–102. 10.1016/j.eml.2015.12.006. [DOI] [Google Scholar]
- Lamoureux A.; Lee K.; Shlian M.; Forrest S. R.; Shtein M. Dynamic Kirigami Structures for Integrated Solar Tracking. Nat. Commun. 2015, 6, 8092. 10.1038/ncomms9092. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tang Y. C.; Yin J. Design of Cut Unit Geometry in Hierarchical Kirigami-Based Auxetic Metamaterials for High Stretchability and Compressibility. EXTREME MECH. LETT. 2017, 12, 77–85. 10.1016/j.eml.2016.07.005. [DOI] [Google Scholar]
- Alcheikh N.; Shaikh S. F.; Hussain M. M. Ultra-Stretchable Archimedean Interconnects for Stretchable Electronics. EXTREME MECH. LETT. 2018, 24, 6–13. 10.1016/j.eml.2018.08.005. [DOI] [Google Scholar]
- Cavazos Sepulveda A. C.; Diaz Cordero M. S.; Carreño A. A. A.; Nassar J. M.; Hussain M. M.. Stretchable and Foldable Silicon-Based Electronics. Appl. Phys. Lett. 2017, 110. 10.1063/1.4979545 [DOI] [Google Scholar]
- Lv C.; Yu H.; Jiang H. Archimedean Spiral Design for Extremely Stretchable Interconnects. EXTREME MECH. LETT. 2014, 1, 29–34. 10.1016/j.eml.2014.12.008. [DOI] [Google Scholar]
- Mamidanna A.; Song Z.; Lv C.; Lefky C. S.; Jiang H.; Hildreth O. J. Printing Stretchable Spiral Interconnects Using Reactive Ink Chemistries. ACS Appl. Mater. Interfaces 2016, 8, 12594–12598. 10.1021/acsami.6b03922. [DOI] [PubMed] [Google Scholar]
- Qaiser N.; Khan S. M.; Hussain M. M.. In-Plane and out-of-Plane Structural Response of Spiral Interconnects for Highly Stretchable Electronics. J. Appl. Phys. 2018, 124. 10.1063/1.5031176 [DOI] [Google Scholar]
- Qaiser N.; Khan S. M.; Nour M.; Rehman M. U.; Rojas J. P.; Hussain M. M.. Mechanical Response of Spiral Interconnect Arrays for Highly Stretchable Electronics. Appl. Phys. Lett. 2017, 111. 10.1063/1.5007111 [DOI] [Google Scholar]
- Rehman M. U.; Rojas J. P. Optimization of Compound Serpentine-Spiral Structure for Ultra-Stretchable Electronics. EXTREME MECH. LETT. 2017, 15, 44–50. 10.1016/j.eml.2017.05.004. [DOI] [Google Scholar]
- Rojas J. P.; Arevalo A.; Foulds I. G.; Hussain M. M.. Design and Characterization of Ultra-Stretchable Monolithic Silicon Fabric. Appl. Phys. Lett. 2014, 105. 10.1063/1.4898128 [DOI] [Google Scholar]
- Rojas J. P.; Singh D.; Conchouso D.; Arevalo A.; Foulds I. G.; Hussain M. M. Stretchable Helical Architecture Inorganic-Organic Hetero Thermoelectric Generator. Nano Energy 2016, 30, 691–699. 10.1016/j.nanoen.2016.10.054. [DOI] [Google Scholar]
- Fan J. A.; Yeo W. H.; Su Y.; Hattori Y.; Lee W.; Jung S. Y.; Zhang Y.; Liu Z.; Cheng H.; Falgout L.; et al. Fractal Design Concepts for Stretchable Electronics. Nat. Commun. 2014, 5, 3266. 10.1038/ncomms4266. [DOI] [PubMed] [Google Scholar]
- Hattori Y.; Falgout L.; Lee W.; Jung S. Y.; Poon E.; Lee J. W.; Na I.; Geisler A.; Sadhwani D.; Zhang Y.; et al. Multifunctional Skin-Like Electronics for Quantitative, Clinical Monitoring of Cutaneous Wound Healing. Adv. Healthc. Mater. 2014, 3, 1597–1607. 10.1002/adhm.201400073. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Huang K. H.; Lin C. T.; Chen Y. T.; Yang Y. J. J.. Study of Fractal Electrode Designs for Buckypaper-Based Micro-Supercapacitors. J. Appl. Phys. 2019, 125. 10.1063/1.5051702 [DOI] [Google Scholar]
- Xu L.; Gutbrod S. R.; Ma Y.; Petrossians A.; Liu Y.; Webb R. C.; Fan J. A.; Yang Z.; Xu R.; Whalen J. J. 3rd; et al. Materials and Fractal Designs for 3d Multifunctional Integumentary Membranes with Capabilities in Cardiac Electrotherapy. Adv. Mater. 2015, 27, 1731–1737. 10.1002/adma.201405017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Xu S.; Zhang Y.; Cho J.; Lee J.; Huang X.; Jia L.; Fan J. A.; Su Y.; Su J.; Zhang H.; et al. Stretchable Batteries with Self-Similar Serpentine Interconnects and Integrated Wireless Recharging Systems. Nat. Commun. 2013, 4, 1543. 10.1038/ncomms2553. [DOI] [PubMed] [Google Scholar]
- Zhang Y. H.; Fu H. R.; Xu S.; Fan J. A.; Hwang K. C.; Jiang J. Q.; Rogers J. A.; Huang Y. G. A Hierarchical Computational Model for Stretchable Interconnects with Fractal-Inspired Designs. J. Mech. Phys. Solids 2014, 72, 115–130. 10.1016/j.jmps.2014.07.011. [DOI] [Google Scholar]
- Kim D. H.; Lu N.; Ma R.; Kim Y. S.; Kim R. H.; Wang S.; Wu J.; Won S. M.; Tao H.; Islam A.; et al. Epidermal Electronics. Science 2011, 333, 838–843. 10.1126/science.1206157. [DOI] [PubMed] [Google Scholar]
- Yan W.; Noel G.; Loke G.; Meiklejohn E.; Khudiyev T.; Marion J.; Rui G.; Lin J.; Cherston J.; Sahasrabudhe A.; et al. Single Fibre Enables Acoustic Fabrics Via Nanometre-Scale Vibrations. Nature 2022, 603, 616–623. 10.1038/s41586-022-04476-9. [DOI] [PubMed] [Google Scholar]
- Yang Q.; Jin W.; Zhang Q.; Wei Y.; Guo Z.; Li X.; Yang Y.; Luo Q.; Tian H.; Ren T.-L. Mixed-Modality Speech Recognition and Interaction Using a Wearable Artificial Throat. Nat. Mach. Intell. 2023, 5, 169–180. 10.1038/s42256-023-00616-6. [DOI] [Google Scholar]
- Byun J.; Lee Y.; Yoon J.; Lee B.; Oh E.; Chung S.; Lee T.; Cho K.-J.; Kim J.; Hong Y. Electronic Skins for Soft, Compact, Reversible Assembly of Wirelessly Activated Fully Soft Robots. Sci. Robot. 2018, 3, eaas9020 10.1126/scirobotics.aas9020. [DOI] [PubMed] [Google Scholar]
- Xiong W.; Zhu C.; Guo D.; Hou C.; Yang Z.; Xu Z.; Qiu L.; Yang H.; Li K.; Huang Y. Bio-Inspired, Intelligent Flexible Sensing Skin for Multifunctional Flying Perception. Nano Energy 2021, 90, 106550. 10.1016/j.nanoen.2021.106550. [DOI] [Google Scholar]
- Yu Y.; Li J.; Solomon S. A.; Min J.; Tu J.; Guo W.; Xu C.; Song Y.; Gao W. All-Printed Soft Human-Machine Interface for Robotic Physicochemical Sensing. Sci. Robot. 2022, 7, eabn0495 10.1126/scirobotics.abn0495. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu Y. M.; Yiu C. K.; Song Z.; Huang Y.; Yao K. M.; Wong T.; Zhou J. K.; Zhao L.; Huang X. C.; Nejad S. K.; et al. Electronic Skin as Wireless Human-Machine Interfaces for Robotic Vr. Sci. Adv. 2022, 8. 10.1126/sciadv.abl6700 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kim K. K.; Kim M.; Pyun K.; Kim J.; Min J.; Koh S.; Root S. E.; Kim J.; Nguyen B.-N. T.; Nishio Y.; et al. A Substrate-Less Nanomesh Receptor with Meta-Learning for Rapid Hand Task Recognition. Nat. Electron. 2023, 6, 64–75. 10.1038/s41928-022-00888-7. [DOI] [Google Scholar]
- Yoon J.; Li L.; Semichaevsky A. V.; Ryu J. H.; Johnson H. T.; Nuzzo R. G.; Rogers J. A. Flexible Concentrator Photovoltaics Based on Microscale Silicon Solar Cells Embedded in Luminescent Waveguides. Nat. Commun. 2011, 2, 343. 10.1038/ncomms1318. [DOI] [PubMed] [Google Scholar]
- Sheng X.; Bower C. A.; Bonafede S.; Wilson J. W.; Fisher B.; Meitl M.; Yuen H.; Wang S.; Shen L.; Banks A. R.; et al. Printing-Based Assembly of Quadruple-Junction Four-Terminal Microscale Solar Cells and Their Use in High-Efficiency Modules. Nat. Mater. 2014, 13, 593–598. 10.1038/nmat3946. [DOI] [PubMed] [Google Scholar]
- Liu R.; Wang Z. L.; Fukuda K.; Someya T. Flexible Self-Charging Power Sources. Nat. Rev. Mater. 2022, 7, 870–886. 10.1038/s41578-022-00441-0. [DOI] [Google Scholar]
- Liu Y.; Li J.; Song S.; Kang J.; Tsao Y.; Chen S.; Mottini V.; McConnell K.; Xu W.; Zheng Y.-Q.; et al. Morphing Electronics Enable Neuromodulation in Growing Tissue. Nat. Biotechnol. 2020, 38, 1031–1036. 10.1038/s41587-020-0495-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hensleigh R.; Cui H.; Xu Z.; Massman J.; Yao D.; Berrigan J.; Zheng X. Charge-Programmed Three-Dimensional Printing for Multi-Material Electronic Devices. Nat. Electron. 2020, 3, 216–224. 10.1038/s41928-020-0391-2. [DOI] [Google Scholar]
- Ge Q.; Chen Z.; Cheng J.; Zhang B.; Zhang Y.-F.; Li H.; He X.; Yuan C.; Liu J.; Magdassi S.; et al. 3d Printing of Highly Stretchable Hydrogel with Diverse Uv Curable Polymers. Sci. Adv. 2021, 7, eaba4261 10.1126/sciadv.aba4261. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cheng J.; Wang R.; Sun Z.; Liu Q.; He X.; Li H.; Ye H.; Yang X.; Wei X.; Li Z.; et al. Centrifugal Multimaterial 3d Printing of Multifunctional Heterogeneous Objects. Nat. Commun. 2022, 13, 7931. 10.1038/s41467-022-35622-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Duraivel S.; Laurent D.; Rajon D. A.; Scheutz G. M.; Shetty A. M.; Sumerlin B. S.; Banks S. A.; Bova F. J.; Angelini T. E. A Silicone-Based Support Material Eliminates Interfacial Instabilities in 3d Silicone Printing. Science 2023, 379, 1248–1252. 10.1126/science.ade4441. [DOI] [PubMed] [Google Scholar]
- Jung W.; Jung Y. H.; Pikhitsa P. V.; Feng J.; Yang Y.; Kim M.; Tsai H. Y.; Tanaka T.; Shin J.; Kim K. Y.; et al. Three-Dimensional Nanoprinting Via Charged Aerosol Jets. Nature 2021, 592, 54–59. 10.1038/s41586-021-03353-1. [DOI] [PubMed] [Google Scholar]
- Kawata S.; Sun H.-B.; Tanaka T.; Takada K. Finer Features for Functional Microdevices. Nature 2001, 412, 697–698. 10.1038/35089130. [DOI] [PubMed] [Google Scholar]
- Meza L. R.; Das S.; Greer J. R. Strong, Lightweight, and Recoverable Three-Dimensional Ceramic Nanolattices. Science 2014, 345, 1322–1326. 10.1126/science.1255908. [DOI] [PubMed] [Google Scholar]
- Han F.; Gu S.; Klimas A.; Zhao N.; Zhao Y.; Chen S.-C. Three-Dimensional Nanofabrication Via Ultrafast Laser Patterning and Kinetically Regulated Material Assembly. Science 2022, 378, 1325–1331. 10.1126/science.abm8420. [DOI] [PubMed] [Google Scholar]
- Li S.; Lerch M. M.; Waters J. T.; Deng B.; Martens R. S.; Yao Y.; Kim D. Y.; Bertoldi K.; Grinthal A.; Balazs A. C.; et al. Self-Regulated Non-Reciprocal Motions in Single-Material Microstructures. Nature 2022, 605, 76–83. 10.1038/s41586-022-04561-z. [DOI] [PubMed] [Google Scholar]
- Wehner M.; Truby R. L.; Fitzgerald D. J.; Mosadegh B.; Whitesides G. M.; Lewis J. A.; Wood R. J. An Integrated Design and Fabrication Strategy for Entirely Soft, Autonomous Robots. Nature 2016, 536, 451–455. 10.1038/nature19100. [DOI] [PubMed] [Google Scholar]
- Saccone M. A.; Gallivan R. A.; Narita K.; Yee D. W.; Greer J. R. Additive Manufacturing of Micro-Architected Metals Via Hydrogel Infusion. Nature 2022, 612, 685–690. 10.1038/s41586-022-05433-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schmidt O. G.; Eberl K. Thin Solid Films Roll up into Nanotubes. Nature 2001, 410, 168–168. 10.1038/35065525. [DOI] [PubMed] [Google Scholar]
- Schmidt O. G.; Jin-Phillipp N. Y. Free-Standing Sige-Based Nanopipelines on Si (001) Substrates. Appl. Phys. Lett. 2001, 78, 3310–3312. 10.1063/1.1373408. [DOI] [Google Scholar]
- Cho J.-H.; James T.; Gracias D. H. Curving Nanostructures Using Extrinsic Stress. Adv. Mater. 2010, 22, 2320–2324. 10.1002/adma.200904410. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Armon S.; Efrati E.; Kupferman R.; Sharon E. Geometry and Mechanics in the Opening of Chiral Seed Pods. Science 2011, 333, 1726–1730. 10.1126/science.1203874. [DOI] [PubMed] [Google Scholar]
- Vaccaro P. O.; Kubota K.; Fleischmann T.; Saravanan S.; Aida T. Valley-Fold and Mountain-Fold in the Micro-Origami Technique. Microelectron. J. 2003, 34, 447–449. 10.1016/S0026-2692(03)00070-3. [DOI] [Google Scholar]
- Leong T.; Gu Z.; Koh T.; Gracias D. H. Spatially Controlled Chemistry Using Remotely Guided Nanoliter Scale Containers. J. Am. Chem. Soc. 2006, 128, 11336–11337. 10.1021/ja063100z. [DOI] [PubMed] [Google Scholar]
- Bassik N.; Brafman A.; Zarafshar A. M.; Jamal M.; Luvsanjav D.; Selaru F. M.; Gracias D. H. Enzymatically Triggered Actuation of Miniaturized Tools. J. Am. Chem. Soc. 2010, 132, 16314–16317. 10.1021/ja106218s. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Becker C.; Bao B.; Karnaushenko D. D.; Bandari V. K.; Rivkin B.; Li Z.; Faghih M.; Karnaushenko D.; Schmidt O. G. A New Dimension for Magnetosensitive E-Skins: Active Matrix Integrated Micro-Origami Sensor Arrays. Nat. Commun. 2022, 13, 2121. 10.1038/s41467-022-29802-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ko H. C.; Shin G.; Wang S.; Stoykovich M. P.; Lee J. W.; Kim D.-H.; Ha J. S.; Huang Y.; Hwang K.-C.; Rogers J. A. Curvilinear Electronics Formed Using Silicon Membrane Circuits and Elastomeric Transfer Elements. Small 2009, 5, 2703–2709. 10.1002/smll.200900934. [DOI] [PubMed] [Google Scholar]
- Jung I.; Shin G.; Malyarchuk V.; Ha J. S.; Rogers J. A. Paraboloid Electronic Eye Cameras Using Deformable Arrays of Photodetectors in Hexagonal Mesh Layouts. Appl. Phys. Lett. 2010, 96, 021110. 10.1063/1.3290244. [DOI] [Google Scholar]
- Jung I.; Xiao J.; Malyarchuk V.; Lu C.; Li M.; Liu Z.; Yoon J.; Huang Y.; Rogers J. A. Dynamically Tunable Hemispherical Electronic Eye Camera System with Adjustable Zoom Capability. Proc. Natl. Acad. Sci. U.S.A. 2011, 108, 1788–1793. 10.1073/pnas.1015440108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Huang Y.; Wu H.; Xiao L.; Duan Y.; Zhu H.; Bian J.; Ye D.; Yin Z. Assembly and Applications of 3d Conformal Electronics on Curvilinear Surfaces. Mater. Horizons 2019, 6, 642–683. 10.1039/C8MH01450G. [DOI] [Google Scholar]
- Xu S.; Yan Z.; Jang K.-I.; Huang W.; Fu H.; Kim J.; Wei Z.; Flavin M.; McCracken J.; Wang R.; et al. Assembly of Micro/Nanomaterials into Complex, Three-Dimensional Architectures by Compressive Buckling. Science 2015, 347, 154–159. 10.1126/science.1260960. [DOI] [PubMed] [Google Scholar]
- Zhang Y.; Zhang F.; Yan Z.; Ma Q.; Li X.; Huang Y.; Rogers J. A. Printing, Folding and Assembly Methods for Forming 3d Mesostructures in Advanced Materials. Nat. Rev. Mater. 2017, 2, 17019. 10.1038/natrevmats.2017.19. [DOI] [Google Scholar]
- Huang G.; Mei Y. Assembly and Self-Assembly of Nanomembrane Materials-from 2d to 3d. Small 2018, 14, 1703665. 10.1002/smll.201703665. [DOI] [PubMed] [Google Scholar]
- Cheng X.; Zhang Y. Micro/Nanoscale 3d Assembly by Rolling, Folding, Curving, and Buckling Approaches. Adv. Mater. 2019, 31, 1901895. 10.1002/adma.201901895. [DOI] [PubMed] [Google Scholar]
- Karnaushenko D.; Kang T.; Bandari V. K.; Zhu F.; Schmidt O. G. 3d Self-Assembled Microelectronic Devices: Concepts, Materials, Applications. Adv. Mater. 2020, 32, 1902994. 10.1002/adma.201902994. [DOI] [PubMed] [Google Scholar]
- Mei Y.; Huang G.; Solovev A. A.; Urena E. B.; Moench I.; Ding F.; Reindl T.; Fu R. K. Y.; Chu P. K.; Schmidt O. G. Versatile Approach for Integrative and Functionalized Tubes by Strain Engineering of Nanomembranes on Polymers. Adv. Mater. 2008, 20, 4085–4090. 10.1002/adma.200801589. [DOI] [Google Scholar]
- Huang X. N.; Kumar K.; Jawed M. K.; Nasab A. M.; Ye Z. S.; Shan W. L.; Majidi C. Chasing Biomimetic Locomotion Speeds: Creating Untethered Soft Robots with Shape Memory Alloy Actuators. Sci. Robot. 2018, 3, eaau7557 10.1126/scirobotics.aau7557. [DOI] [PubMed] [Google Scholar]
- Jamal M.; Bassik N.; Cho J.-H.; Randall C. L.; Gracias D. H. Directed Growth of Fibroblasts into Three Dimensional Micropatterned Geometries Via Self-Assembling Scaffolds. Biomaterials 2010, 31, 1683–1690. 10.1016/j.biomaterials.2009.11.056. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Py C.; Reverdy P.; Doppler L.; Bico J.; Roman B.; Baroud C. Capillary Origami. Phys. Fluids 2007, 19, 091104. 10.1063/1.2775288. [DOI] [PubMed] [Google Scholar]
- Py C.; Reverdy P.; Doppler L.; Bico J.; Roman B.; Baroud C. N. Capillary Origami: Spontaneous Wrapping of a Droplet with an Elastic Sheet. Phys. Rev. Lett. 2007, 98, 156103. 10.1103/PhysRevLett.98.156103. [DOI] [PubMed] [Google Scholar]
- Cho J.-H.; Gracias D. H. Self-Assembly of Lithographically Patterned Nanoparticles. Nano Lett. 2009, 9, 4049–4052. 10.1021/nl9022176. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cho J.-H.; Azam A.; Gracias D. H. Three Dimensional Nanofabrication Using Surface Forces. Langmuir 2010, 26, 16534–16539. 10.1021/la1013889. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gladman A. S.; Matsumoto E. A.; Nuzzo R. G.; Mahadevan L.; Lewis J. A. Biomimetic 4d Printing. Nat. Mater. 2016, 15, 413. 10.1038/nmat4544. [DOI] [PubMed] [Google Scholar]
- Aharoni H.; Xia Y.; Zhang X.; Kamien R. D.; Yang S. Universal Inverse Design of Surfaces with Thin Nematic Elastomer Sheets. Proc. Natl. Acad. Sci. U.S.A. 2018, 115, 7206–7211. 10.1073/pnas.1804702115. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hu W.; Lum G. Z.; Mastrangeli M.; Sitti M. Small-Scale Soft-Bodied Robot with Multimodal Locomotion. Nature 2018, 554, 81–85. 10.1038/nature25443. [DOI] [PubMed] [Google Scholar]
- Wang C.; Sim K.; Chen J.; Kim H.; Rao Z.; Li Y.; Chen W.; Song J.; Verduzco R.; Yu C. Soft Ultrathin Electronics Innervated Adaptive Fully Soft Robots. Adv. Mater. 2018, 30, 1706695 10.1002/adma.201706695. [DOI] [PubMed] [Google Scholar]
- Xu B.; Tian Z.; Wang J.; Han H.; Lee T.; Mei Y. Stimuli-Responsive and on-Chip Nanomembrane Micro-Rolls for Enhanced Macroscopic Visual Hydrogen Detection. Sci. Adv. 2018, 4, eaap8203 10.1126/sciadv.aap8203. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang Y.; Yan Z.; Nan K.; Xiao D.; Liu Y.; Luan H.; Fu H.; Wang X.; Yang Q.; Wang J.; et al. A Mechanically Driven Form of Kirigami as a Route to 3d Mesostructures in Micro/Nanomembranes. Proc. Natl. Acad. Sci. U.S.A. 2015, 112, 11757–11764. 10.1073/pnas.1515602112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yan Z.; Zhang F.; Liu F.; Han M.; Ou D.; Liu Y.; Lin Q.; Guo X.; Fu H.; Xie Z.; et al. Mechanical Assembly of Complex, 3d Mesostructures from Releasable Multilayers of Advanced Materials. Sci. Adv. 2016, 2, e1601014 10.1126/sciadv.1601014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu Q.; Wang W.; Reynolds M. F.; Cao M. C.; Miskin M. Z.; Arias T. A.; Muller D. A.; McEuen P. L.; Cohen I. Micrometer-Sized Electrically Programmable Shape-Memory Actuators for Low-Power Microrobotics. Sci. Robot. 2021, 6, eabe6663 10.1126/scirobotics.abe6663. [DOI] [PubMed] [Google Scholar]
- Wang W.; Liu Q.; Tanasijevic I.; Reynolds M. F.; Cortese A. J.; Miskin M. Z.; Cao M. C.; Muller D. A.; Molnar A. C.; Lauga E.; et al. Cilia Metasurfaces for Electronically Programmable Microfluidic Manipulation. Nature 2022, 605, 681–686. 10.1038/s41586-022-04645-w. [DOI] [PubMed] [Google Scholar]
- Fu H.; Nan K.; Bai W.; Huang W.; Bai K.; Lu L.; Zhou C.; Liu Y.; Liu F.; Wang J.; et al. Morphable 3d Mesostructures and Microelectronic Devices by Multistable Buckling Mechanics. Nat. Mater. 2018, 17, 268–276. 10.1038/s41563-017-0011-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cools J.; Jin Q.; Yoon E.; Burbano D. A.; Luo Z.; Cuypers D.; Callewaert G.; Braeken D.; Gracias D. H. A Micropatterned Multielectrode Shell for 3d Spatiotemporal Recording from Live Cells. Adv. Sci. 2018, 5, 1700731. 10.1002/advs.201700731. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gao H.; Yang F.; Sattari K.; Du X.; Fu T.; Fu S.; Liu X.; Lin J.; Sun Y.; Yao J. Bioinspired Two-in-One Nanotransistor Sensor for the Simultaneous Measurements of Electrical and Mechanical Cellular Responses. Sci. Adv. 2022, 8, eabn2485 10.1126/sciadv.abn2485. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Park Y.; Franz C. K.; Ryu H.; Luan H.; Cotton K. Y.; Kim J. U.; Chung T. S.; Zhao S.; Vazquez-Guardado A.; Yang D. S.; et al. Three-Dimensional, Multifunctional Neural Interfaces for Cortical Spheroids and Engineered Assembloids. Sci. Adv. 2021, 7, eabf9153 10.1126/sciadv.abf9153. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Huang Q.; Tang B.; Romero J. C.; Yang Y.; Elsayed S. K.; Pahapale G.; Lee T.-J.; Morales Pantoja I. E.; Han F.; Berlinicke C.; et al. Shell Microelectrode Arrays (Meas) for Brain Organoids. Sci. Adv. 2022, 8, eabq5031 10.1126/sciadv.abq5031. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Song Y. M.; Xie Y.; Malyarchuk V.; Xiao J.; Jung I.; Choi K.-J.; Liu Z.; Park H.; Lu C.; Kim R.-H.; et al. Digital Cameras with Designs Inspired by the Arthropod Eye. Nature 2013, 497, 95–99. 10.1038/nature12083. [DOI] [PubMed] [Google Scholar]
- Rao Z.; Lu Y.; Li Z.; Sim K.; Ma Z.; Xiao J.; Yu C. Curvy, Shape-Adaptive Imagers Based on Printed Optoelectronic Pixels with a Kirigami Design. Nat. Electron. 2021, 4, 513–521. 10.1038/s41928-021-00600-1. [DOI] [Google Scholar]
- Ji Z.; Zhao J.; Song H.; Xu S.; Pang W.; Hu X.; Zhang F.; Jin T.; Shuai Y.; Lan Y.; et al. Morphable Three-Dimensional Electronic Mesofliers Capable of on-Demand Unfolding. Sci. China Mater. 2022, 65, 2309–2318. 10.1007/s40843-022-2007-8. [DOI] [Google Scholar]
- Han M.; Wang H.; Yang Y.; Liang C.; Bai W.; Yan Z.; Li H.; Xue Y.; Wang X.; Akar B.; et al. Three-Dimensional Piezoelectric Polymer Microsystems for Vibrational Energy Harvesting, Robotic Interfaces and Biomedical Implants. Nat. Electron. 2019, 2, 26–35. 10.1038/s41928-018-0189-7. [DOI] [Google Scholar]
- Won S. M.; Wang H.; Kim B. H.; Lee K.; Jang H.; Kwon K.; Han M.; Crawford K. E.; Li H.; Lee Y.; et al. Multimodal Sensing with a Three-Dimensional Piezoresistive Structure. ACS Nano 2019, 13, 10972–10979. 10.1021/acsnano.9b02030. [DOI] [PubMed] [Google Scholar]
- Ghosh A.; Li L.; Xu L.; Dash R. P.; Gupta N.; Lam J.; Jin Q.; Akshintala V.; Pahapale G.; Liu W.; et al. Gastrointestinal-Resident, Shape-Changing Microdevices Extend Drug Release in Vivo. Sci. Adv. 2020, 6, eabb4133 10.1126/sciadv.abb4133. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Han M.; Chen L.; Aras K.; Liang C.; Chen X.; Zhao H.; Li K.; Faye N. R.; Sun B.; Kim J. H.; et al. Catheter-Integrated Soft Multilayer Electronic Arrays for Multiplexed Sensing and Actuation During Cardiac Surgery. Nat. Biomed. Eng. 2020, 4, 997–1009. 10.1038/s41551-020-00604-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Prinz V. Y.; Seleznev V. A.; Gutakovsky A. K.; Chehovskiy A. V.; Preobrazhenskii V. V.; Putyato M. A.; Gavrilova T. A. Free-Standing and Overgrown Ingaas/Gaas Nanotubes, Nanohelices and Their Arrays. Physica E 2000, 6, 828–831. 10.1016/S1386-9477(99)00249-0. [DOI] [Google Scholar]
- Zhang L.; Ruh E.; Gruetzmacher D.; Dong L.; Bell D. J.; Nelson B. J.; Schoenenberger C. Anomalous Coiling of Sige/Si and Sige/Si/Cr Helical Nanobelts. Nano Lett. 2006, 6, 1311–1317. 10.1021/nl052340u. [DOI] [PubMed] [Google Scholar]
- Cheng X.; Fan Z.; Yao S.; Jin T.; Lv Z.; Lan Y.; Bo R.; Chen Y.; Zhang F.; Shen Z.; et al. Programming 3d Curved Mesosurfaces Using Microlattice Designs. Science 2023, 379, 1225–1232. 10.1126/science.adf3824. [DOI] [PubMed] [Google Scholar]
- Bai K.; Cheng X.; Xue Z.; Song H.; Sang L.; Zhang F.; Liu F.; Luo X.; Huang W.; Huang Y.; et al. Geometrically Reconfigurable 3d Mesostructures and Electromagnetic Devices through a Rational Bottom-up Design Strategy. Sci. Adv. 2020, 6, eabb7417 10.1126/sciadv.abb7417. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Taylor J. M.; Luan H.; Lewis J. A.; Rogers J. A.; Nuzzo R. G.; Braun P. V. Biomimetic and Biologically Compliant Soft Architectures Via 3d and 4d Assembly Methods: A Perspective. Adv. Mater. 2022, 34, 2108391 10.1002/adma.202108391. [DOI] [PubMed] [Google Scholar]
- Xu B.; Zhang B.; Wang L.; Huang G.; Mei Y. Tubular Micro/Nanomachines: From the Basics to Recent Advances. Adv. Funct. Mater. 2018, 28, 1705872. 10.1002/adfm.201705872. [DOI] [Google Scholar]
- Bell D. J.; Dong L. X.; Nelson B. J.; Golling M.; Zhang L.; Grutzmacher D. Fabrication and Characterization of Three-Dimensional Ingaas/Gaas Nanosprings. Nano Lett. 2006, 6, 725–729. 10.1021/nl0525148. [DOI] [PubMed] [Google Scholar]
- Songmuang R.; Rastelli A.; Mendach S.; Schmidt O. G. Siox/Si Radial Superlattices and Microtube Optical Ring Resonators. Appl. Phys. Lett. 2007, 90, 091905. 10.1063/1.2472546. [DOI] [Google Scholar]
- Chun I. S.; Li X. Controlled Assembly and Dispersion of Strain-Induced Ingaas/Gaas Nanotubes. IEEE Trans. Nanotechnol. 2008, 7, 493–495. 10.1109/TNANO.2008.926372. [DOI] [Google Scholar]
- Zhang L.; Abbott J. J.; Dong L.; Peyer K. E.; Kratochvil B. E.; Zhang H.; Bergeles C.; Nelson B. J. Characterizing the Swimming Properties of Artificial Bacterial Flagella. Nano Lett. 2009, 9, 3663–3667. 10.1021/nl901869j. [DOI] [PubMed] [Google Scholar]
- Chun I. S.; Challa A.; Derickson B.; Hsia K. J.; Li X. Geometry Effect on the Strain-Induced Self-Rolling of Semiconductor Membranes. Nano Lett. 2010, 10, 3927–3932. 10.1021/nl101669u. [DOI] [PubMed] [Google Scholar]
- Jamal M.; Zarafshar A. M.; Gracias D. H. Differentially Photo-Crosslinked Polymers Enable Self-Assembling Microfluidics. Nat. Commun. 2011, 2, 527. 10.1038/ncomms1531. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mönch I.; Makarov D.; Koseva R.; Baraban L.; Karnaushenko D.; Kaiser C.; Arndt K.-F.; Schmidt O. G. Rolled-up Magnetic Sensor: Nanomembrane Architecture for in-Flow Detection of Magnetic Objects. ACS Nano 2011, 5, 7436–7442. 10.1021/nn202351j. [DOI] [PubMed] [Google Scholar]
- Smith E. J.; Schulze S.; Kiravittaya S.; Mei Y.; Sanchez S.; Schmidt O. G. Lab-in-a-Tube: Detection of Individual Mouse Cells for Analysis in Flexible Split-Wall Microtube Resonator Sensors. Nano Lett. 2011, 11, 4037–4042. 10.1021/nl1036148. [DOI] [PubMed] [Google Scholar]
- Huang W.; Yu X.; Froeter P.; Xu R.; Ferreira P.; Li X. On-Chip Inductors with Self-Rolled-up Sinx Nanomembrane Tubes: A Novel Design Platform for Extreme Miniaturization. Nano Lett. 2012, 12, 6283–6288. 10.1021/nl303395d. [DOI] [PubMed] [Google Scholar]
- Grimm D.; Bufon C. C. B.; Deneke C.; Atkinson P.; Thurmer D. J.; Schaeffel F.; Gorantla S.; Bachmatiuk A.; Schmidt O. G. Rolled-up Nanomembranes as Compact 3d Architectures for Field Effect Transistors and Fluidic Sensing Applications. Nano Lett. 2013, 13, 213–218. 10.1021/nl303887b. [DOI] [PubMed] [Google Scholar]
- Martinez-Cisneros C. S.; Sanchez S.; Xi W.; Schmidt O. G. Ultracompact Three-Dimensional Tubular Conductivity Microsensors for Ionic and Biosensing Applications. Nano Lett. 2014, 14, 2219–2224. 10.1021/nl500795k. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sharma R.; Bufon C. C. B.; Grimm D.; Sommer R.; Wollatz A.; Schadewald J.; Thurmer D. J.; Siles P. F.; Bauer M.; Schmidt O. G. Large-Area Rolled-up Nanomembrane Capacitor Arrays for Electrostatic Energy Storage. Adv. Electron. Mater. 2014, 4, 1301631. 10.1002/aenm.201301631. [DOI] [Google Scholar]
- Magdanz V.; Medina-Sanchez M.; Chen Y.; Guix M.; Schmidt O. G. How to Improve Spermbot Performance. Adv. Funct. Mater. 2015, 25, 2763–2770. 10.1002/adfm.201500015. [DOI] [Google Scholar]
- Yu X.; Huang W.; Li M.; Comberiate T. M.; Gong S.; Schutt-Aine J. E.; Li X. Ultra-Small, High-Frequency, and Substrate-Immune Microtube Inductors Transformed from 2d to 3d. Sci. Rep. 2015, 5, 9661. 10.1038/srep09661. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Deng J.; Lu X.; Liu L.; Zhang L.; Schmidt O. G. Introducing Rolled-up Nanotechnology for Advanced Energy Storage Devices. Adv. Electron. Mater. 2016, 6, 1600797. 10.1002/aenm.201600797. [DOI] [Google Scholar]
- Jalil A. R.; Chang H.; Bandari V. K.; Robaschik P.; Zhang J.; Siles P. F.; Li G.; Buerger D.; Grimm D.; Liu X.; et al. Fully Integrated Organic Nanocrystal Diode as High Performance Room Temperature No2 Sensor. Adv. Mater. 2016, 28, 2971–2977. 10.1002/adma.201506293. [DOI] [PubMed] [Google Scholar]
- Zhao Z.; Kuang X.; Yuan C.; Qi H. J.; Fang D. Hydrophilic/Hydrophobic Composite Shape-Shifting Structures. ACS Appl. Mater. Interfaces 2018, 10, 19932–19939. 10.1021/acsami.8b02444. [DOI] [PubMed] [Google Scholar]
- Deng T.; Zhang Z.; Liu Y.; Wang Y.; Su F.; Li S.; Zhang Y.; Li H.; Chen H.; Zhao Z.; et al. Three-Dimensional Graphene Field-Effect Transistors as High-Performance Photodetectors. Nano Lett. 2019, 19, 1494–1503. 10.1021/acs.nanolett.8b04099. [DOI] [PubMed] [Google Scholar]
- Huang H. W.; Uslu F. E.; Katsamba P.; Lauga E.; Sakar M. S.; Nelson B. J. Adaptive Locomotion of Artificial Microswimmers. Sci. Adv. 2019, 5, eaau1532 10.1126/sciadv.aau1532. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Huang X.; Kumar K.; Jawed M. K.; Mohammadi Nasab A.; Ye Z.; Shan W.; Majidi C. Highly Dynamic Shape Memory Alloy Actuator for Fast Moving Soft Robots. Adv. Mater. Technol. 2019, 4, 1800540. 10.1002/admt.201800540. [DOI] [Google Scholar]
- Tian Z.; Xu B.; Wan G.; Han X.; Di Z.; Chen Z.; Mei Y.. Gaussian-Preserved, Non-Volatile Shape Morphing in Three-Dimensional Microstructures for Dual-Functional Electronic Devices. Nat. Commun. 2021, 12. 10.1038/s41467-020-20843-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li X.; Cao C.; Liu C.; He W.; Wu K.; Wang Y.; Xu B.; Tian Z.; Song E.; Cui J.; et al. Self-Rolling of Vanadium Dioxide Nanomembranes for Enhanced Multi-Level Solar Modulation. Nat. Commun. 2022, 13, 7819. 10.1038/s41467-022-35513-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang F.; Li D.; Wang C.; Liu Z.; Yang M.; Cui Z.; Yi J.; Wang M.; Jiang Y.; Lv Z.; et al. Shape Morphing of Plastic Films. Nat. Commun. 2022, 13, 7294. 10.1038/s41467-022-34844-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- In H. J.; Lee H.; Nichol A. J.; Kim S. G.; Barbastathis G. Carbon Nanotube-Based Magnetic Actuation of Origami Membranes. J. Vac. Sci. Technol. B 2008, 26, 2509–2512. 10.1116/1.3002559. [DOI] [Google Scholar]
- Bassik N.; Stern G. M.; Gracias D. H. Microassembly Based on Hands Free Origami with Bidirectional Curvature. Appl. Phys. Lett. 2009, 95, 091901. 10.1063/1.3212896. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Leong T. G.; Randall C. L.; Benson B. R.; Bassik N.; Stern G. M.; Gracias D. H. Tetherless Thermobiochemically Actuated Microgrippers. Proc. Natl. Acad. Sci. U.S.A. 2009, 106, 703–708. 10.1073/pnas.0807698106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li H.; Guo X.; Nuzzo R. G.; Hsia K. J. Capillary Induced Self-Assembly of Thin Foils into 3d Structures. J. Mech. Phys. Solids 2010, 58, 2033–2042. 10.1016/j.jmps.2010.09.011. [DOI] [Google Scholar]
- Randhawa J. S.; Gurbani S. S.; Keung M. D.; Demers D. P.; Leahy-Hoppa M. R.; Gracias D. H. Three-Dimensional Surface Current Loops in Terahertz Responsive Microarrays. Appl. Phys. Lett. 2010, 96, 191108. 10.1063/1.3428657. [DOI] [Google Scholar]
- Azam A.; Laflin K. E.; Jamal M.; Fernandes R.; Gracias D. H. Self-Folding Micropatterned Polymeric Containers. Biomed. Microdevices 2011, 13, 51–58. 10.1007/s10544-010-9470-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cho J.-H.; Keung M. D.; Verellen N.; Lagae L.; Moshchalkov V. V.; Van Dorpe P.; Gracias D. H. Nanoscale Origami for 3d Optics. Small 2011, 7, 1943–1948. 10.1002/smll.201100568. [DOI] [PubMed] [Google Scholar]
- Pandey S.; Ewing M.; Kunas A.; Nguyen N.; Gracias D. H.; Menon G. Algorithmic Design of Self-Folding Polyhedra. Proc. Natl. Acad. Sci. U.S.A. 2011, 108, 19885–19890. 10.1073/pnas.1110857108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Randall C. L.; Kalinin Y. V.; Jamal M.; Manohar T.; Gracias D. H. Three-Dimensional Microwell Arrays for Cell Culture. Lab Chip 2011, 11, 127–131. 10.1039/C0LC00368A. [DOI] [PubMed] [Google Scholar]
- Felton S. M.; Tolley M. T.; Shin B.; Onal C. D.; Demaine E. D.; Rus D.; Wood R. J. Self-Folding with Shape Memory Composites. Soft Matter 2013, 9, 7688. 10.1039/c3sm51003d. [DOI] [Google Scholar]
- Felton S.; Tolley M.; Demaine E.; Rus D.; Wood R. A Method for Building Self-Folding Machines. Science 2014, 345, 644–646. 10.1126/science.1252610. [DOI] [PubMed] [Google Scholar]
- Legrain A.; Janson T. G.; Berenschot J. W.; Abelmann L.; Tas N. R. Controllable Elastocapillary Folding of Three-Dimensional Micro-Objects by through-Wafer Filling. J. Appl. Phys. 2014, 115, 214905. 10.1063/1.4878460. [DOI] [Google Scholar]
- Cui A.; Liu Z.; Li J.; Shen T. H.; Xia X.; Li Z.; Gong Z.; Li H.; Wang B.; Li J.; et al. Directly Patterned Substrate-Free Plasmonic ″Nanograter’’ Structures with Unusual Fano Resonances. Light-Sci. Appl. 2015, 4, e308 10.1038/lsa.2015.81. [DOI] [Google Scholar]
- Na J.-H.; Evans A. A.; Bae J.; Chiappelli M. C.; Santangelo C. D.; Lang R. J.; Hull T. C.; Hayward R. C. Programming Reversibly Self-Folding Origami with Micropatterned Photo-Crosslinkable Polymer Trilayers. Adv. Mater. 2015, 27, 79–85. 10.1002/adma.201403510. [DOI] [PubMed] [Google Scholar]
- Anacleto P.; Gultepe E.; Gomes S.; Mendes P. M.; Gracias D. H. Self-Folding Microcube Antennas for Wireless Power Transfer in Dispersive Media. Technology 2016, 4, 120–129. 10.1142/S2339547816500047. [DOI] [Google Scholar]
- Mao Y.; Pan Y.; Zhang W.; Zhu R.; Xu J.; Wu W. Multi-Direction-Tunable Three-Dimensional Meta-Atoms for Reversible Switching between Midwave and Long-Wave Infrared Regimes. Nano Lett. 2016, 16, 7025–7029. 10.1021/acs.nanolett.6b03210. [DOI] [PubMed] [Google Scholar]
- Pandey S.; Macias N. J.; Ciobanu C.; Yoon C.; Teuscher C.; Gracias D. H. Assembly of a 3d Cellular Computer Using Folded E-Blocks. Micromachines 2016, 7, 78. 10.3390/mi7050078. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jin Q.; Li M.; Polat B.; Paidi S. K.; Dai A.; Zhang A.; Pagaduan J. V.; Barman I.; Gracias D. H. Mechanical Trap Surface-Enhanced Raman Spectroscopy for Three-Dimensional Surface Molecular Imaging of Single Live Cells. Angew. Chem., Int. Ed. 2017, 56, 3822–3826. 10.1002/anie.201700695. [DOI] [PubMed] [Google Scholar]
- Cui J. Z.; Huang T. Y.; Luo Z. C.; Testa P.; Gu H. R.; Chen X. Z.; Nelson B. J.; Heyderman L. J. Nanomagnetic Encoding of Shape-Morphing Micromachines. Nature 2019, 575, 164–168. 10.1038/s41586-019-1713-2. [DOI] [PubMed] [Google Scholar]
- Jin Q.; Yang Y.; Jackson J. A.; Yoon C.; Gracias D. H. Untethered Single Cell Grippers for Active Biopsy. Nano Lett. 2020, 20, 5383–5390. 10.1021/acs.nanolett.0c01729. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li C.; Xue Y.; Han M.; Palmer L. C.; Rogers J. A.; Huang Y.; Stupp S. I. Synergistic Photoactuation of Bilayered Spiropyran Hydrogels for Predictable Origami-Like Shape Change. Matter 2021, 4, 1377–1390. 10.1016/j.matt.2021.01.016. [DOI] [Google Scholar]
- Xu X.; Davanco M.; Qi X.; Forrest S. R. Direct Transfer Patterning on Three Dimensionally Deformed Surfaces at Micrometer Resolutions and Its Application to Hemispherical Focal Plane Detector Arrays. Org. Electron. 2008, 9, 1122–1127. 10.1016/j.orgel.2008.07.011. [DOI] [Google Scholar]
- Shin G.; Jung I.; Malyarchuk V.; Song J.; Wang S.; Ko H. C.; Huang Y.; Ha J. S.; Rogers J. A. Micromechanics and Advanced Designs for Curved Photodetector Arrays in Hemispherical Electronic-Eye Cameras. Small 2010, 6, 851–856. 10.1002/smll.200901350. [DOI] [PubMed] [Google Scholar]
- Kim D.-H.; Lu N.; Ghaffari R.; Kim Y.-S.; Lee S. P.; Xu L.; Wu J.; Kim R.-H.; Song J.; Liu Z.; et al. Materials for Multifunctional Balloon Catheters with Capabilities in Cardiac Electrophysiological Mapping and Ablation Therapy. Nat. Mater. 2011, 10, 316–323. 10.1038/nmat2971. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kim J.; Lee M.; Shim H. J.; Ghaffari R.; Cho H. R.; Son D.; Jung Y. H.; Soh M.; Choi C.; Jung S.; et al. Stretchable Silicon Nanoribbon Electronics for Skin Prosthesis. Nat. Commun. 2014, 5, 5747. 10.1038/ncomms6747. [DOI] [PubMed] [Google Scholar]
- Park H.; Cho H.; Kim J.; Bang J. W.; Seo S.; Rahmawan Y.; Lee D. Y.; Suh K.-Y. Multiscale Transfer Printing into Recessed Microwells and on Curved Surfaces Via Hierarchical Perfluoropolyether Stamps. Small 2014, 10, 52–59. 10.1002/smll.201300772. [DOI] [PubMed] [Google Scholar]
- Xu L.; Gutbrod S. R.; Bonifas A. P.; Su Y.; Sulkin M. S.; Lu N.; Chung H. J.; Jang K. I.; Liu Z.; Ying M.; et al. 3d Multifunctional Integumentary Membranes for Spatiotemporal Cardiac Measurements and Stimulation across the Entire Epicardium. Nat. Commun. 2014, 5, 3329. 10.1038/ncomms4329. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang Y.; Huang Y.; Rogers J. A. Mechanics of Stretchable Batteries and Supercapacitors. Curr. Opin Solid St M 2015, 19, 190–199. 10.1016/j.cossms.2015.01.002. [DOI] [Google Scholar]
- Park J.; Choi S.; Janardhan A. H.; Lee S.-Y.; Raut S.; Soares J.; Shin K.; Yang S.; Lee C.; Kang K.-W.; et al. Electromechanical Cardioplasty Using a Wrapped Elasto-Conductive Epicardial Mesh. Sci. Transl. Med. 2016, 8, 344ra386. 10.1126/scitranslmed.aad8568. [DOI] [PubMed] [Google Scholar]
- Choi C.; Choi M. K.; Liu S.; Kim M. S.; Park O. K.; Im C.; Kim J.; Qin X.; Lee G. J.; Cho K. W.; et al. Human Eye-Inspired Soft Optoelectronic Device Using High-Density Mos2-Graphene Curved Image Sensor Array. Nat. Commun. 2017, 8, 1664. 10.1038/s41467-017-01824-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Le Borgne B.; De Sagazan O.; Crand S.; Jacques E.; Harnois M. Conformal Electronics Wrapped around Daily Life Objects Using an Original Method: Water Transfer Printing. ACS Appl. Mater. Interfaces 2017, 9, 29424–29429. 10.1021/acsami.7b07327. [DOI] [PubMed] [Google Scholar]
- Saada G.; Layani M.; Chernevousky A.; Magdassi S. Hydroprinting Conductive Patterns onto 3d Structures. Adv. Mater. Technol. 2017, 2, 1600289. 10.1002/admt.201600289. [DOI] [Google Scholar]
- Zhang K.; Jung Y. H.; Mikael S.; Seo J.-H.; Kim M.; Mi H.; Zhou H.; Xia Z.; Zhou W.; Gong S.; et al. Origami Silicon Optoelectronics for Hemispherical Electronic Eye Systems. Nat. Commun. 2017, 8, 1782. 10.1038/s41467-017-01926-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee G. J.; Choi C.; Kim D.-H.; Song Y. M. Bioinspired Artificial Eyes: Optic Components, Digital Cameras, and Visual Prostheses. Adv. Funct. Mater. 2018, 28, 1705202. 10.1002/adfm.201705202. [DOI] [Google Scholar]
- Ershad F.; Sim K.; Thukral A.; Zhang Y. S.; Yu C. Emerging Soft Bioelectronics for Cardiac Health Diagnosis and Treatment. APL Mater. 2019, 7, 031301. 10.1063/1.5060270. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fan D.; Lee B.; Coburn C.; Forrest S. R. From 2d to 3d: Strain- and Elongation-Free Topological Transformations of Optoelectronic Circuits. Proc. Natl. Acad. Sci. U.S.A. 2019, 116, 3968–3973. 10.1073/pnas.1813001116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sim K.; Chen S.; Li Z.; Rao Z.; Liu J.; Lu Y.; Jang S.; Ershad F.; Chen J.; Xiao J.; et al. Three-Dimensional Curvy Electronics Created Using Conformal Additive Stamp Printing. Nat. Electron. 2019, 2, 471–479. 10.1038/s41928-019-0304-4. [DOI] [Google Scholar]
- Tian L.; Zimmerman B.; Akhtar A.; Yu K. J.; Moore M.; Wu J.; Larsen R. J.; Lee J. W.; Li J.; Liu Y.; et al. Large-Area Mri-Compatible Epidermal Electronic Interfaces for Prosthetic Control and Cognitive Monitoring. Nat. Biomed. Eng. 2019, 3, 194–205. 10.1038/s41551-019-0347-x. [DOI] [PubMed] [Google Scholar]
- Choi J.; Han C.; Cho S.; Kim K.; Ahn J.; Del Orbe D.; Cho I.; Zhao Z. J.; Oh Y. S.; Hong H.; et al. Customizable, Conformal, and Stretchable 3d Electronics Via Predistorted Pattern Generation and Thermoforming. Sci. Adv. 2021, 7. 10.1126/sciadv.abj0694 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu J.; Jiang S.; Xiong W.; Zhu C.; Li K.; Huang Y.. Self-Healing Kirigami Assembly Strategy for Conformal Electronics. Adv. Funct. Mater. 2022. 10.1002/adfm.202109214 [DOI] [Google Scholar]
- Liu X.; Cao Y.; Zheng K.; Zhang Y.; Wang Z.; Chen Y.; Chen Y.; Ma Y.; Feng X.. Liquid Droplet Stamp Transfer Printing. Adv. Funct. Mater. 2021, 31. 10.1002/adfm.202105407 [DOI] [Google Scholar]
- Jiang S.; Liu J.; Xiong W.; Yang Z.; Yin L.; Li K.; Huang Y. A Snakeskin-Inspired, Soft-Hinge Kirigami Metamaterial for Self-Adaptive Conformal Electronic Armor. Adv. Mater. 2022, 34, 2204091 10.1002/adma.202204091. [DOI] [PubMed] [Google Scholar]
- Zabow G. Reflow Transfer for Conformal Three-Dimensional Microprinting. Science 2022, 378, 894–898. 10.1126/science.add7023. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu S.; He J.; Rao Y.; Dai Z.; Ye H.; Tanir J. C.; Li Y.; Lu N. Conformability of Flexible Sheets on Spherical Surfaces. Sci. Adv. 2023, 9, eadf2709 10.1126/sciadv.adf2709. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu Y.; Yan Z.; Lin Q.; Guo X.; Han M.; Nan K.; Hwang K. C.; Huang Y.; Zhang Y.; Rogers J. A. Guided Formation of 3d Helical Mesostructures by Mechanical Buckling: Analytical Modeling and Experimental Validation. Adv. Funct. Mater. 2016, 26, 2909–2918. 10.1002/adfm.201505132. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yan Z.; Zhang F.; Wang J.; Liu F.; Guo X.; Nan K.; Lin Q.; Gao M.; Xiao D.; Shi Y.; et al. Controlled Mechanical Buckling for Origami-Inspired Construction of 3d Microstructures in Advanced Materials. Adv. Funct. Mater. 2016, 26, 2629–2639. 10.1002/adfm.201504901. [DOI] [PMC free article] [PubMed] [Google Scholar]
- McCracken J. M.; Xu S.; Badea A.; Jang K. I.; Yan Z.; Wetzel D. J.; Nan K.; Lin Q.; Han M.; Anderson M. A.; et al. Deterministic Integration of Biological and Soft Materials onto 3d Microscale Cellular Frameworks. Adv. Biosyst. 2017, 1, 1700068. 10.1002/adbi.201700068. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nan K.; Luan H.; Yan Z.; Ning X.; Wang Y.; Wang A.; Wang J.; Han M.; Chang M.; Li K.; et al. Engineered Elastomer Substrates for Guided Assembly of Complex 3d Mesostructures by Spatially Nonuniform Compressive Buckling. Adv. Funct. Mater. 2017, 27, 1604281. 10.1002/adfm.201604281. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shi Y.; Zhang F.; Nan K.; Wang X.; Wang J.; Zhang Y.; Zhang Y.; Luan H.; Hwang K.-C.; Huang Y.; et al. Plasticity-Induced Origami for Assembly of Three Dimensional Metallic Structures Guided by Compressive Buckling. Extreme Mech. Lett. 2017, 11, 105–110. 10.1016/j.eml.2016.11.008. [DOI] [Google Scholar]
- Yan Z.; Han M.; Shi Y.; Badea A.; Yang Y.; Kulkarni A.; Hanson E.; Kandel M. E.; Wen X.; Zhang F.; et al. Three-Dimensional Mesostructures as High-Temperature Growth Templates, Electronic Cellular Scaffolds, and Self-Propelled Microrobots. Proc. Natl. Acad. Sci. U.S.A. 2017, 114, e9455–e9464. 10.1073/pnas.1713805114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fan Z.; Hwang K.-C.; Rogers J. A.; Huang Y.; Zhang Y. A Double Perturbation Method of Postbuckling Analysis in 2d Curved Beams for Assembly of 3d Ribbon-Shaped Structures. J. Mech. Phys. Solids 2018, 111, 215–238. 10.1016/j.jmps.2017.10.012. [DOI] [Google Scholar]
- Guo X.; Wang X.; Ou D.; Ye J.; Pang W.; Huang Y.; Rogers J. A.; Zhang Y. Controlled Mechanical Assembly of Complex 3d Mesostructures and Strain Sensors by Tensile Buckling. npj Flex. Electron. 2018, 2, 14. 10.1038/s41528-018-0028-y. [DOI] [Google Scholar]
- Lee W.; Liu Y.; Lee Y.; Sharma B. K.; Shinde S. M.; Kim S. D.; Nan K.; Yan Z.; Han M.; Huang Y.; et al. Two-Dimensional Materials in Functional Three-Dimensional Architectures with Applications in Photodetection and Imaging. Nat. Commun. 2018, 9, 1417. 10.1038/s41467-018-03870-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ling Y.; Zhuang X.; Xu Z.; Xie Y.; Zhu X.; Xu Y.; Sun B.; Lin J.; Zhang Y.; Yan Z. Mechanically Assembled, Three-Dimensional Hierarchical Structures of Cellular Graphene with Programmed Geometries and Outstanding Electromechanical Properties. ACS Nano 2018, 12, 12456–12463. 10.1021/acsnano.8b06675. [DOI] [PubMed] [Google Scholar]
- Ning X.; Yu X.; Wang H.; Sun R.; Corman R. E.; Li H.; Lee C. M.; Xue Y.; Chempakasseril A.; Yao Y.; et al. Mechanically Active Materials in Three-Dimensional Mesostructures. Sci. Adv. 2018, 4, eaat8313 10.1126/sciadv.aat8313. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang F.; Liu F.; Zhang Y. Analyses of Mechanically-Assembled 3d Spiral Mesostructures with Applications as Tunable Inductors. Sci. China Technol. Sci. 2019, 62, 243–251. 10.1007/s11431-018-9368-y. [DOI] [Google Scholar]
- Liu W.; Zou Q.; Zheng C.; Jin C. Metal-Assisted Transfer Strategy for Construction of 2d and 3d Nanostructures on an Elastic Substrate. ACS Nano 2019, 13, 440–448. 10.1021/acsnano.8b06623. [DOI] [PubMed] [Google Scholar]
- Liu Y.; Wang X.; Xu Y.; Xue Z.; Zhang Y.; Ning X.; Cheng X.; Xue Y.; Lu D.; Zhang Q.; et al. Harnessing the Interface Mechanics of Hard Films and Soft Substrates for 3d Assembly by Controlled Buckling. Proc. Natl. Acad. Sci. U.S.A. 2019, 116, 15368–15377. 10.1073/pnas.1907732116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu Y.; Xu Z.; Hwang K. C.; Huang Y.; Zhang Y. Postbuckling Analyses of Frame Mesostructures Consisting of Straight Ribbons for Mechanically Guided Three-Dimensional Assembly. Proc.R.Soc.A 2019, 475, 20190012. 10.1098/rspa.2019.0012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Luan H.; Cheng X.; Wang A.; Zhao S.; Bai K.; Wang H.; Pang W.; Xie Z.; Li K.; Zhang F.; et al. Design and Fabrication of Heterogeneous, Deformable Substrates for the Mechanically Guided 3d Assembly. ACS Appl. Mater. Interfaces 2019, 11, 3482–3492. 10.1021/acsami.8b19187. [DOI] [PubMed] [Google Scholar]
- Pang W.; Cheng X.; Zhao H.; Guo X.; Ji Z.; Li G.; Liang Y.; Xue Z.; Song H.; Zhang F.; et al. Electro-Mechanically Controlled Assembly of Reconfigurable 3d Mesostructures and Electronic Devices Based on Dielectric Elastomer Platforms. Natl. Sci. Rev. 2020, 7, 342–354. 10.1093/nsr/nwz164. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Park J. K.; Nan K.; Luan H.; Zheng N.; Zhao S.; Zhang H.; Cheng X.; Wang H.; Li K.; Xie T.; et al. Remotely Triggered Assembly of 3d Mesostructures through Shape-Memory Effects. Adv. Mater. 2019, 31, 1905715 10.1002/adma.201905715. [DOI] [PubMed] [Google Scholar]
- Park Y.; Luan H.; Kwon K.; Zhao S.; Franklin D.; Wang H.; Zhao H.; Bai W.; Kim J. U.; Lu W.; et al. Transformable, Freestanding 3d Mesostructures Based on Transient Materials and Mechanical Interlocking. Adv. Funct. Mater. 2019, 29, 1903181. 10.1002/adfm.201903181. [DOI] [Google Scholar]
- Wang X.; Guo X.; Ye J.; Zheng N.; Kohli P.; Choi D.; Zhang Y.; Xie Z.; Zhang Q.; Luan H.; et al. Freestanding 3d Mesostructures, Functional Devices, and Shape-Programmable Systems Based on Mechanically Induced Assembly with Shape Memory Polymers. Adv. Mater. 2019, 31, 1805615 10.1002/adma.201805615. [DOI] [PubMed] [Google Scholar]
- Yang C.; Huang Y.; Cheng H.; Jiang L.; Qu L. Rollable, Stretchable, and Reconfigurable Graphene Hygroelectric Generators. Adv. Mater. 2019, 31, 1805705 10.1002/adma.201805705. [DOI] [PubMed] [Google Scholar]
- Zhang C.; Deng H.; Xie Y.; Zhang C.; Su J. W.; Lin J. Stimulus Responsive 3d Assembly for Spatially Resolved Bifunctional Sensors. Small 2019, 15, 1904224 10.1002/smll.201904224. [DOI] [PubMed] [Google Scholar]
- Zhang Y.; Wang F.; Ma Y.; Feng X. Buckling Configurations of Stiff Thin Films Tuned by Micro-Patterns on Soft Substrate. Int. J. Solids Struct. 2019, 161, 55–63. 10.1016/j.ijsolstr.2018.11.004. [DOI] [Google Scholar]
- Zhao H.; Li K.; Han M.; Zhu F.; Vazquez-Guardado A.; Guo P.; Xie Z.; Park Y.; Chen L.; Wang X.; et al. Buckling and Twisting of Advanced Materials into Morphable 3d Mesostructures. Proc. Natl. Acad. Sci. U.S.A. 2019, 116, 13239–13248. 10.1073/pnas.1901193116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lim S.; Luan H.; Zhao S.; Lee Y.; Zhang Y.; Huang Y.; Rogers J. A.; Ahn J.-H. Assembly of Foldable 3d Microstructures Using Graphene Hinges. Adv. Mater. 2020, 32, 2001303. 10.1002/adma.202001303. [DOI] [PubMed] [Google Scholar]
- Li Y.; Luo C.; Yu K.; Wang X. Remotely Controlled, Reversible, on-Demand Assembly and Reconfiguration of 3d Mesostructures Via Liquid Crystal Elastomer Platforms. ACS Appl. Mater. Interfaces 2021, 13, 8929–8939. 10.1021/acsami.0c21371. [DOI] [PubMed] [Google Scholar]
- Li Y.; Yu H.; Yu K.; Guo X.; Wang X. Reconfigurable Three-Dimensional Mesotructures of Spatially Programmed Liquid Crystal Elastomers and Their Ferromagnetic Composites. Adv. Funct. Mater. 2021, 31, 2100338. 10.1002/adfm.202100338. [DOI] [Google Scholar]
- Luan H.; Zhang Q.; Liu T.-L.; Wang X.; Zhao S.; Wang H.; Yao S.; Xue Y.; Kwak J. W.; Bai W.; et al. Complex 3d Microfluidic Architectures Formed by Mechanically Guided Compressive Buckling. Sci. Adv. 2021, 7, eabj3686 10.1126/sciadv.abj3686. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Miao L.; Song Y.; Ren Z.; Xu C.; Wan J.; Wang H.; Guo H.; Xiang Z.; Han M.; Zhang H. 3d Temporary-Magnetized Soft Robotic Structures for Enhanced Energy Harvesting. Adv. Mater. 2021, 33, 2102691 10.1002/adma.202102691. [DOI] [PubMed] [Google Scholar]
- Zhang F.; Li S.; Shen Z.; Cheng X.; Xue Z.; Zhang H.; Song H.; Bai K.; Yan D.; Wang H.; et al. Rapidly Deployable and Morphable 3d Mesostructures with Applications in Multimodal Biomedical Devices. Proc. Natl. Acad. Sci. U.S.A. 2021, 118. 10.1073/pnas.2026414118 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li R.; Zhang C.; Li J.; Zhang Y.; Liu S.; Hu Y.; Jiang S.; Chen C.; Xin C.; Tao Y.; et al. Magnetically Encoded 3d Mesostructure with High-Order Shape Morphing and High-Frequency Actuation. Natl. Sci. Rev. 2022, 9, nwac163. 10.1093/nsr/nwac163. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Xue Z.; Jin T.; Xu S.; Bai K.; He Q.; Zhang F.; Cheng X.; Ji Z.; Pang W.; Shen Z.; et al. Assembly of Complex 3d Structures and Electronics on Curved Surfaces. Sci. Adv. 2022, 8, eabm6922 10.1126/sciadv.abm6922. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhao H.; Cheng X.; Wu C.; Liu T. L.; Zhao Q.; Li S.; Ni X.; Yao S.; Han M.; Huang Y.; et al. Mechanically Guided Hierarchical Assembly of 3d Mesostructures. Adv. Mater. 2022, 34, 2109416 10.1002/adma.202109416. [DOI] [PubMed] [Google Scholar]
- Ahn J.; Ha J. H.; Jeong Y.; Jung Y.; Choi J.; Gu J.; Hwang S. H.; Kang M.; Ko J.; Cho S.; et al. Nanoscale Three-Dimensional Fabrication Based on Mechanically Guided Assembly. Nat. Commun. 2023, 14, 833. 10.1038/s41467-023-36302-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pang W.; Liu L.; Xu S.; Shuai Y.; Zhao J.; Zhang Y.; Electroadhesion-Mediated Interface Delamination for Assembly of Reconfigurable 3d Mesostructures. J. Appl. Mech. 2023, 90. 10.1115/1.4056861 [DOI] [Google Scholar]
- Shuai Y.; Zhao J.; Bo R.; Lan Y.; Lv Z.; Zhang Y. A Wrinkling-Assisted Strategy for Controlled Interface Delamination in Mechanically-Guided 3d Assembly. J. Mech. Phys. Solids 2023, 173, 105203. 10.1016/j.jmps.2023.105203. [DOI] [Google Scholar]
- Ling Y.; Pang W.; Li X.; Goswami S.; Xu Z.; Stroman D.; Liu Y.; Fei Q.; Xu Y.; Zhao G.; et al. Laser-Induced Graphene for Electrothermally Controlled, Mechanically Guided, 3d Assembly and Human-Soft Actuators Interaction. Adv. Mater. 2020, 32, 1908475 10.1002/adma.201908475. [DOI] [PubMed] [Google Scholar]
- Liu Z.; Du H.; Li J.; Lu L.; Li Z.-Y.; Fang N. X. Nano-Kirigami with Giant Optical Chirality. Sci. Adv. 2018, 4, eaat4436 10.1126/sciadv.aat4436. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jiang J.; Zhang S.; Wang B.; Ding H.; Wu Z. Hydroprinted Liquid-Alloy-Based Morphing Electronics for Fast-Growing/Tender Plants: From Physiology Monitoring to Habit Manipulation. Small 2020, 16, 2003833. 10.1002/smll.202003833. [DOI] [PubMed] [Google Scholar]
- Liu Y.; Zheng M.; O’Connor B.; Dong J.; Zhu Y. Curvilinear Soft Electronics by Micromolding of Metal Nanowires in Capillaries. Sci. Adv. 2022, 8, eadd6996 10.1126/sciadv.add6996. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kim B. H.; Lee J.; Won S. M.; Xie Z.; Chang J. K.; Yu Y.; Cho Y. K.; Jang H.; Jeong J. Y.; Lee Y.; et al. Three-Dimensional Silicon Electronic Systems Fabricated by Compressive Buckling Process. ACS Nano 2018, 12, 4164–4171. 10.1021/acsnano.8b00180. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Luo G.; Fu H.; Cheng X.; Bai K.; Shi L.; He X.; Rogers J. A.; Huang Y.; Zhang Y. Mechanics of Bistable Cross-Shaped Structures through Loading-Path Controlled 3d Assembly. J. Mech. Phys. Solids 2019, 129, 261–277. 10.1016/j.jmps.2019.05.007. [DOI] [Google Scholar]
- Rogers J. A.; Lagally M. G.; Nuzzo R. G. Synthesis, Assembly and Applications of Semiconductor Nanomembranes. Nature 2011, 477, 45–53. 10.1038/nature10381. [DOI] [PubMed] [Google Scholar]
- Khang D. Y.; Rogers J. A.; Lee H. H. Mechanical Buckling: Mechanics, Metrology, and Stretchable Electronics. Adv. Funct. Mater. 2009, 19, 1526–1536. 10.1002/adfm.200801065. [DOI] [Google Scholar]
- Jiang H.; Sun Y.; Rogers J. A.; Huang Y.. Mechanics of Precisely Controlled Thin Film Buckling on Elastomeric Substrate. Appl. Phys. Lett. 2007, 90. 10.1063/1.2719027 [DOI] [Google Scholar]
- Wang S.; Song J.; Kim D.-H.; Huang Y.; Rogers J. A.. Local Versus Global Buckling of Thin Films on Elastomeric Substrates. Appl. Phys. Lett. 2008, 93. 10.1063/1.2956402 [DOI] [Google Scholar]
- Yan Z. G.; Wang B. L.; Wang K. F. Stretchability and Compressibility of a Novel Layout Design for Flexible Electronics Based on Bended Wrinkle Geometries. Compos. B. Eng. 2019, 166, 65–73. 10.1016/j.compositesb.2018.11.123. [DOI] [Google Scholar]
- Li Z. W.; Wang Y.; Xiao J. L. Mechanics of Curvilinear Electronics and Optoelectronics. Curr. Opin Solid St M 2015, 19, 171–189. 10.1016/j.cossms.2015.01.003. [DOI] [Google Scholar]
- Cheng H. Y.; Zhang Y. H.; Hwang K. C.; Rogers J. A.; Huang Y. G. Buckling of a Stiff Thin Film on a Pre-Strained Bi-Layer Substrate. Int. J. Solids Struct. 2014, 51, 3113–3118. 10.1016/j.ijsolstr.2014.05.012. [DOI] [Google Scholar]
- Jiang H.; Sun Y.; Rogers J. A.; Huang Y. Post-Buckling Analysis for the Precisely Controlled Buckling of Thin Film Encapsulated by Elastomeric Substrates. Int. J. Solids Struct. 2008, 45, 2014–2023. 10.1016/j.ijsolstr.2007.11.007. [DOI] [Google Scholar]
- Song J.; Jiang H.; Liu Z. J.; Khang D. Y.; Huang Y.; Rogers J. A.; Lu C.; Koh C. G. Buckling of a Stiff Thin Film on a Compliant Substrate in Large Deformation. Int. J. Solids Struct. 2008, 45, 3107–3121. 10.1016/j.ijsolstr.2008.01.023. [DOI] [Google Scholar]
- Yoder M. A.; Yan Z.; Han M.; Rogers J. A.; Nuzzo R. G. Semiconductor Nanomembrane Materials for High-Performance Soft Electronic Devices. J. Am. Chem. Soc. 2018, 140, 9001–9019. 10.1021/jacs.8b04225. [DOI] [PubMed] [Google Scholar]
- Cheng H.; Song J.. A Simply Analytic Study of Buckled Thin Films on Compliant Substrates. J. Appl. Mech. 2014, 81. 10.1115/1.4025306 [DOI] [Google Scholar]
- Wang A.; Avila R.; Ma Y. J.. Mechanics Design for Buckling of Thin Ribbons on an Elastomeric Substrate without Material Failure. J. Appl. Mech. 2017, 84. 10.1115/1.4037149 [DOI] [Google Scholar]
- Song J.; Huang Y.; Xiao J.; Wang S.; Hwang K. C.; Ko H. C.; Kim D. H.; Stoykovich M. P.; Rogers J. A.. Mechanics of Noncoplanar Mesh Design for Stretchable Electronic Circuits. J. Appl. Phys. 2009, 105. 10.1063/1.3148245 [DOI] [Google Scholar]
- Wang Y.; Li Z. W.; Xiao J. L.. Stretchable Thin Film Materials: Fabrication, Application, and Mechanics. J. Electron. Packag. 2016, 138. 10.1115/1.4032984 [DOI] [Google Scholar]
- Sun Y. G.; Rogers J. A. Structural Forms of Single Crystal Semiconductor Nanoribbons for High-Performance Stretchable Electronics. J. Mater. Chem. 2007, 17, 832–840. 10.1039/b614793c. [DOI] [Google Scholar]
- Jiang H. Q.; Khang D. Y.; Fei H. Y.; Kim H.; Huang Y. G.; Xiao J. L.; Rogers J. A. Finite Width Effect of Thin-Films Buckling on Compliant Substrate: Experimental and Theoretical Studies. J. Mech. Phys. Solids 2008, 56, 2585–2598. 10.1016/j.jmps.2008.03.005. [DOI] [Google Scholar]
- Zhang Q. T.; Yin H. Spontaneous Buckling-Driven Periodic Delamination of Thin Films on Soft Substrates under Large Compression. J. Mech. Phys. Solids 2018, 118, 40–57. 10.1016/j.jmps.2018.05.009. [DOI] [Google Scholar]
- Song J.; Jiang H.; Huang Y.; Rogers J. A. Mechanics of Stretchable Inorganic Electronic Materials. J. Vac. Sci. Technol. A 2009, 27, 1107–1125. 10.1116/1.3168555. [DOI] [Google Scholar]
- Choi W. M.; Song J.; Khang D. Y.; Jiang H.; Huang Y. Y.; Rogers J. A. Biaxially Stretchable ″Wavy″ Silicon Nanomembranes. Nano Lett. 2007, 7, 1655–1663. 10.1021/nl0706244. [DOI] [PubMed] [Google Scholar]
- Li G.; Zhang M.; Liu S.; Yuan M.; Wu J.; Yu M.; Teng L.; Xu Z.; Guo J.; Li G.; et al. Three-Dimensional Flexible Electronics Using Solidified Liquid Metal with Regulated Plasticity. Nat. Electron. 2023, 6, 154. 10.1038/s41928-022-00914-8. [DOI] [Google Scholar]
- Song J.; Feng X.; Huang Y. Mechanics and Thermal Management of Stretchable Inorganic Electronics. Natl. Sci. Rev. 2016, 3, 128–143. 10.1093/nsr/nwv078. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jiang H.; Khang D. Y.; Song J.; Sun Y.; Huang Y.; Rogers J. A. Finite Deformation Mechanics in Buckled Thin Films on Compliant Supports. Proc. Natl. Acad. Sci. U.S.A. 2007, 104, 15607–15612. 10.1073/pnas.0702927104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ma Y.; Xue Y.; Jang K. I.; Feng X.; Rogers J. A.; Huang Y. Wrinkling of a Stiff Thin Film Bonded to a Pre-Strained, Compliant Substrate with Finite Thickness. Proc.R.Soc.A 2016, 472, 20160339. 10.1098/rspa.2016.0339. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li R.; Li M.; Su Y. W.; Song J. Z.; Ni X. Q. An Analytical Mechanics Model for the Island-Bridge Structure of Stretchable Electronics. Soft Matter 2013, 9, 8476–8482. 10.1039/c3sm51476e. [DOI] [Google Scholar]
- Wang S. D.; Xiao J. L.; Song J. Z.; Ko H. C.; Hwang K. C.; Huang Y. G.; Rogers J. A. Mechanics of Curvilinear Electronics. Soft Matter 2010, 6, 5757–5763. 10.1039/c0sm00579g. [DOI] [Google Scholar]
- Chen C.; Tao W.; Liu Z. J.; Zhang Y. W.; Song J. Controlled Buckling of Thin Film on Elastomeric Substrate in Large Deformation. Theor. Appl. Mech. Lett. 2011, 1, 021001. 10.1063/2.1102101. [DOI] [Google Scholar]
- Zhang Y.; Wang S.; Li X.; Fan J. A.; Xu S.; Song Y. M.; Choi K.-J.; Yeo W.-H.; Lee W.; Nazaar S. N.; et al. Experimental and Theoretical Studies of Serpentine Microstructures Bonded to Prestrained Elastomers for Stretchable Electronics. Adv. Funct. Mater. 2014, 24, 2028–2037. 10.1002/adfm.201302957. [DOI] [Google Scholar]
- Kim D. H.; Liu Z.; Kim Y. S.; Wu J.; Song J.; Kim H. S.; Huang Y.; Hwang K. C.; Zhang Y.; Rogers J. A. Optimized Structural Designs for Stretchable Silicon Integrated Circuits. Small 2009, 5, 2841–2847. 10.1002/smll.200900853. [DOI] [PubMed] [Google Scholar]
- Li K.; Cheng X.; Zhu F.; Li L. Z.; Xie Z. Q.; Luan H. W.; Wang Z. H.; Ji Z. Y.; Wang H. L.; Liu F.; et al. A Generic Soft Encapsulation Strategy for Stretchable Electronics. Adv. Funct. Mater. 2019, 29. 10.1002/adfm.201806630 [DOI] [Google Scholar]
- Jang K. I.; Li K.; Chung H. U.; Xu S.; Jung H. N.; Yang Y.; Kwak J. W.; Jung H. H.; Song J.; Yang C.; et al. Self-Assembled Three Dimensional Network Designs for Soft Electronics. Nat. Commun. 2017, 8, 15894. 10.1038/ncomms15894. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee J.; Ihle S. J.; Pellegrino G. S.; Kim H.; Yea J.; Jeon C.-Y.; Son H.-C.; Jin C.; Eberli D.; Schmid F.; et al. Stretchable and Suturable Fibre Sensors for Wireless Monitoring of Connective Tissue Strain. Nat. Electron. 2021, 4, 291–301. 10.1038/s41928-021-00557-1. [DOI] [Google Scholar]
- Wang B.; Bao S.; Vinnikova S.; Ghanta P.; Wang S.. Buckling Analysis in Stretchable Electronics. npj Flex. Electron. 2017, 1. 10.1038/s41528-017-0004-y [DOI] [Google Scholar]
- Nan K.; Wang H.; Ning X.; Miller K. A.; Wei C.; Liu Y.; Li H.; Xue Y.; Xie Z.; Luan H.; et al. Soft Three-Dimensional Microscale Vibratory Platforms for Characterization of Nano-Thin Polymer Films. ACS Nano 2019, 13, 449–457. 10.1021/acsnano.8b06736. [DOI] [PubMed] [Google Scholar]
- Liu F.; Cheng X.; Zhang F.; Chen Y.; Song H.; Huang Y.; Zhang Y. Design and Assembly of Reconfigurable 3d Radio-Frequency Antennas Based on Mechanically Triggered Switches. Adv. Electron. Mater. 2019, 5, 1900256. 10.1002/aelm.201900256. [DOI] [Google Scholar]
- Qi D.; Liu Z.; Liu Y.; Leow W. R.; Zhu B.; Yang H.; Yu J.; Wang W.; Wang H.; Yin S.; et al. Suspended Wavy Graphene Microribbons for Highly Stretchable Microsupercapacitors. Adv. Mater. 2015, 27, 5559–5566. 10.1002/adma.201502549. [DOI] [PubMed] [Google Scholar]
- Kwon K.; Kim J. U.; Won S. M.; Zhao J.; Avila R.; Wang H.; Chun K. S.; Jang H.; Lee K. H.; Kim J. H.; et al. A Battery-Less Wireless Implant for the Continuous Monitoring of Vascular Pressure, Flow Rate and Temperature. Nat. Biomed. Eng. 2023. 10.1038/s41551-023-01022-4 [DOI] [PubMed] [Google Scholar]
- Zhao H.; Kim Y.; Wang H.; Ning X.; Xu C.; Suh J.; Han M.; Pagan-Diaz G. J.; Lu W.; Li H.; et al. Compliant 3d Frameworks Instrumented with Strain Sensors for Characterization of Millimeter-Scale Engineered Muscle Tissues. Proc. Natl. Acad. Sci. U.S.A. 2021, 118. 10.1073/pnas.2100077118 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Medina-Sánchez M.; Magdanz V.; Guix M.; Fomin V. M.; Schmidt O. G. Swimming Microrobots: Soft, Reconfigurable, and Smart. Adv. Funct. Mater. 2018, 28, 1707228. 10.1002/adfm.201707228. [DOI] [Google Scholar]
- Nan K.; Kang S. D.; Li K.; Yu K. J.; Zhu F.; Wang J.; Dunn A. C.; Zhou C.; Xie Z.; Agne M. T.; et al. Compliant and Stretchable Thermoelectric Coils for Energy Harvesting in Miniature Flexible Devices. Sci. Adv. 2018, 4, eaau5849 10.1126/sciadv.aau5849. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Deng J.; Ji H.; Yan C.; Zhang J.; Si W.; Baunack S.; Oswald S.; Mei Y.; Schmidt O. G. Naturally Rolled-up C/Si/C Trilayer Nanomembranes as Stable Anodes for Lithium-Ion Batteries with Remarkable Cycling Performance. Angew. Chem., Int. Ed. 2013, 52, 2326–2330. 10.1002/anie.201208357. [DOI] [PubMed] [Google Scholar]
- Gabler F.; Karnaushenko D. D.; Karnaushenko D.; Schmidt O. G. Magnetic Origami Creates High Performance Micro Devices. Nat. Commun. 2019, 10, 3013. 10.1038/s41467-019-10947-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee Y.; Bandari V. K.; Li Z.; Medina-Sánchez M.; Maitz M. F.; Karnaushenko D.; Tsurkan M. V.; Karnaushenko D. D.; Schmidt O. G. Nano-Biosupercapacitors Enable Autarkic Sensor Operation in Blood. Nat. Commun. 2021, 12, 4967. 10.1038/s41467-021-24863-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang H.; Zhen H.; Li S.; Jing Y.; Huang G.; Mei Y.; Lu W. Self-Rolling and Light-Trapping in Flexible Quantum Well-Embedded Nanomembranes for Wide-Angle Infrared Photodetectors. Sci. Adv. 2016, 2, e1600027 10.1126/sciadv.1600027. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Joung D.; Nemilentsau A.; Agarwal K.; Dai C.; Liu C.; Su Q.; Li J.; Low T.; Koester S. J.; Cho J. H. Self-Assembled Three-Dimensional Graphene-Based Polyhedrons Inducing Volumetric Light Confinement. Nano Lett. 2017, 17, 1987–1994. 10.1021/acs.nanolett.6b05412. [DOI] [PubMed] [Google Scholar]
- Malachowski K.; Jamal M.; Jin Q.; Polat B.; Morris C. J.; Gracias D. H. Self-Folding Single Cell Grippers. Nano Lett. 2014, 14, 4164–4170. 10.1021/nl500136a. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Xu W.; Li T.; Qin Z.; Huang Q.; Gao H.; Kang K.; Park J.; Buehler M. J.; Khurgin J. B.; Gracias D. H. Reversible Mos2 Origami with Spatially Resolved and Reconfigurable Photosensitivity. Nano Lett. 2019, 19, 7941–7949. 10.1021/acs.nanolett.9b03107. [DOI] [PubMed] [Google Scholar]
- Kim D. C.; Yun H.; Kim J.; Seung H.; Yu W. S.; Koo J. H.; Yang J.; Kim J. H.; Hyeon T.; Kim D.-H. Three-Dimensional Foldable Quantum Dot Light-Emitting Diodes. Nat. Electron. 2021, 4, 671–680. 10.1038/s41928-021-00643-4. [DOI] [Google Scholar]
- Bao N.; Liu Q.; Reynolds M. F.; Figueras M.; Smith E.; Wang W.; Cao M. C.; Muller D. A.; Mavrikakis M.; Cohen I.; et al. Gas-Phase Microactuation Using Kinetically Controlled Surface States of Ultrathin Catalytic Sheets. Proc. Natl. Acad. Sci. U.S.A. 2023, 120, e2221740120 10.1073/pnas.2221740120. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Adams J. J.; Duoss E. B.; Malkowski T. F.; Motala M. J.; Ahn B. Y.; Nuzzo R. G.; Bernhard J. T.; Lewis J. A. Conformal Printing of Electrically Small Antennas on Three-Dimensional Surfaces. Adv. Funct. Mater. 2011, 23, 1335–1340. 10.1002/adma.201003734. [DOI] [PubMed] [Google Scholar]
- Fan Z.; Yang Y.; Zhang F.; Xu Z.; Zhao H.; Wang T.; Song H.; Huang Y.; Rogers J. A.; Zhang Y. Inverse Design Strategies for 3d Surfaces Formed by Mechanically Guided Assembly. Adv. Mater. 2020, 32, e1908424 10.1002/adma.201908424. [DOI] [PubMed] [Google Scholar]
- Liu F.; Chen Y.; Song H.; Zhang F.; Fan Z.; Liu Y.; Feng X.; Rogers J. A.; Huang Y.; Zhang Y. High Performance, Tunable Electrically Small Antennas through Mechanically Guided 3d Assembly. Small 2019, 15, 1804055 10.1002/smll.201804055. [DOI] [PubMed] [Google Scholar]
- Hao X. P.; Li C. Y.; Zhang C. W.; Du M.; Ying Z.; Zheng Q.; Wu Z. L.. Self-Shaping Soft Electronics Based on Patterned Hydrogel with Stencil-Printed Liquid Metal. Adv. Funct. Mater. 2021, 31. 10.1002/adfm.202105481 [DOI] [Google Scholar]
- Gultepe E.; Randhawa J. S.; Kadam S.; Yamanaka S.; Selaru F. M.; Shin E. J.; Kalloo A. N.; Gracias D. H. Biopsy with Thermally-Responsive Untethered Microtools. Adv. Mater. 2013, 25, 514–519. 10.1002/adma.201203348. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hiendlmeier L.; Zurita F.; Vogel J.; Del Duca F.; Al Boustani G.; Peng H.; Kopic I.; Nikic M.; Teshima T.; Wolfrum B. 4d Printed Soft and Stretchable Self-Folding Cuff Electrodes for Small-Nerve Interfacing. Adv. Mater. 2023, 35, e2210206 10.1002/adma.202210206. [DOI] [PubMed] [Google Scholar]
- Ryu H.; Park Y.; Luan H.; Dalgin G.; Jeffris K.; Yoon H. J.; Chung T. S.; Kim J. U.; Kwak S. S.; Lee G.; et al. Transparent, Compliant 3d Mesostructures for Precise Evaluation of Mechanical Characteristics of Organoids. Adv. Mater. 2021, 33, 2100026 10.1002/adma.202100026. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Du X.; Cui H.; Sun B.; Wang J.; Zhao Q.; Xia K.; Wu T.; Humayun M. S. Photothermally Triggered Shape-Adaptable 3d Flexible Electronics. Adv. Mater. Technol. 2017, 2, 1700120. 10.1002/admt.201700120. [DOI] [Google Scholar]
- Wang J.; Zhao Q.; Wang Y.; Zeng Q.; Wu T.; Du X. Self-Unfolding Flexible Microelectrode Arrays Based on Shape Memory Polymers. Adv. Mater. Technol. 2019, 4, 1900566. 10.1002/admt.201900566. [DOI] [Google Scholar]
- Wang H.; Wang Y.; Tee B. C.; Kim K.; Lopez J.; Cai W.; Bao Z. Shape-Controlled, Self-Wrapped Carbon Nanotube 3d Electronics. Adv. Sci. 2015, 2, 1500103. 10.1002/advs.201500103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kalmykov A.; Huang C.; Bliley J.; Shiwarski D.; Tashman J.; Abdullah A.; Rastogi S. K.; Shukla S.; Mataev E.; Feinberg A. W.; et al. Organ-on-E-Chip: Three-Dimensional Self-Rolled Biosensor Array for Electrical Interrogations of Human Electrogenic Spheroids. Sci. Adv. 2019, 5, eaax0729 10.1126/sciadv.aax0729. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang Y. C.; Zheng N.; Cao Y.; Wang F. L.; Wang P.; Ma Y. J.; Lu B. W.; Hou G. H.; Fang Z. Z.; Liang Z. W.; et al. Climbing-Inspired Twining Electrodes Using Shape Memory for Peripheral Nerve Stimulation and Recording. Sci. Adv. 2019, 5. 10.1126/sciadv.aaw1066 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhu Y.; Birla M.; Oldham K. R.; Filipov E. T. Elastically and Plastically Foldable Electrothermal Micro-Origami for Controllable and Rapid Shape Morphing. Adv. Funct. Mater. 2020, 30, 2003741. 10.1002/adfm.202003741. [DOI] [Google Scholar]
- Wang X.; Wang Z.; Dong H.; Saggau C. N.; Tang H.; Tang M.; Liu L.; Baunack S.; Bai L.; Liu J.; et al. Collective Coupling of 3d Confined Optical Modes in Monolithic Twin Microtube Cavities Formed by Nanomembrane Origami. Nano Lett. 2022, 22, 6692–6699. 10.1021/acs.nanolett.2c02083. [DOI] [PubMed] [Google Scholar]
- Tang H.; Karnaushenko D. D.; Neu V.; Gabler F.; Wang S.; Liu L.; Li Y.; Wang J.; Zhu M.; Schmidt O. G. Stress-Actuated Spiral Microelectrode for High-Performance Lithium-Ion Microbatteries. Small 2020, 16, 2002410 10.1002/smll.202002410. [DOI] [PubMed] [Google Scholar]
- Qu Z.; Zhu M.; Yin Y.; Huang Y.; Tang H.; Ge J.; Li Y.; Karnaushenko D. D.; Karnaushenko D.; Schmidt O. G.. A Sub-Square-Millimeter Microbattery with Milliampere-Hour-Level Footprint Capacity. Adv. Energy Mater. 2022, 12. 10.1002/aenm.202200714 [DOI] [Google Scholar]
- Si W.; Monch I.; Yan C.; Deng J.; Li S.; Lin G.; Han L.; Mei Y.; Schmidt O. G. A Single Rolled-up Si Tube Battery for the Study of Electrochemical Kinetics, Electrical Conductivity, and Structural Integrity. Adv. Mater. 2014, 26, 7973–7978. 10.1002/adma.201402484. [DOI] [PubMed] [Google Scholar]
- Xu W.; Paidi S. K.; Qin Z.; Huang Q.; Yu C.-H.; Pagaduan J. V.; Buehler M. J.; Barman I.; Gracias D. H. Self-Folding Hybrid Graphene Skin for 3d Biosensing. Nano Lett. 2019, 19, 1409–1417. 10.1021/acs.nanolett.8b03461. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ghosh A.; Liu W.; Li L.; Pahapale G. J.; Choi S. Y.; Xu L.; Huang Q.; Zhang R.; Zhong Z.; Selaru F. M.; et al. Autonomous Untethered Microinjectors for Gastrointestinal Delivery of Insulin. ACS Nano 2022, 16, 16211–16220. 10.1021/acsnano.2c05098. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ghosh A.; Liu Y.; Artemov D.; Gracias D. H. Magnetic Resonance Guided Navigation of Untethered Microgrippers. Adv. Healthc. Mater. 2021, 10, 2000869 10.1002/adhm.202000869. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Melzer M.; Kaltenbrunner M.; Makarov D.; Karnaushenko D.; Karnaushenko D.; Sekitani T.; Someya T.; Schmidt O. G. Imperceptible Magnetoelectronics. Nat. Commun. 2015, 6, 6080. 10.1038/ncomms7080. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee Y.-K.; Xi Z.; Lee Y.-J.; Kim Y.-H.; Hao Y.; Choi H.; Lee M.-G.; Joo Y.-C.; Kim C.; Lien J.-M.; et al. Computational Wrapping: A Universal Method to Wrap 3d-Curved Surfaces with Nonstretchable Materials for Conformal Devices. Sci. Adv. 2020, 6, eaax6212 10.1126/sciadv.aax6212. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen I. T.; Poblete F. R.; Bagal A.; Zhu Y.; Chang C. H. Anelasticity in Thin-Shell Nanolattices. Proc. Natl. Acad. Sci. U. S. A. 2022, 119, e2201589119 10.1073/pnas.2201589119. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Huang W.; Zhou J.; Froeter P. J.; Walsh K.; Liu S.; Kraman M. D.; Li M.; Michaels J. A.; Sievers D. J.; Gong S.; et al. Three-Dimensional Radio-Frequency Transformers Based on a Self-Rolled-up Membrane Platform. Nat. Electron. 2018, 1, 305–313. 10.1038/s41928-018-0073-5. [DOI] [Google Scholar]
- Lee H.; Choi T. K.; Lee Y. B.; Cho H. R.; Ghaffari R.; Wang L.; Choi H. J.; Chung T. D.; Lu N.; Hyeon T.; et al. A Graphene-Based Electrochemical Device with Thermoresponsive Microneedles for Diabetes Monitoring and Therapy. Nat. Nanotechnol. 2016, 11, 566–572. 10.1038/nnano.2016.38. [DOI] [PubMed] [Google Scholar]
- Zhang C.; Chen J.; Gao J.; Tan G.; Bai S.; Weng K.; Chen H. M.; Ding X.; Cheng H.; Yang Y.; et al. Laser Processing of Crumpled Porous Graphene/Mxene Nanocomposites for a Standalone Gas Sensing System. Nano Lett. 2023, 23, 3435–3443. 10.1021/acs.nanolett.3c00454. [DOI] [PubMed] [Google Scholar]
- Gu Y.; Wang C.; Kim N.; Zhang J.; Wang T. M.; Stowe J.; Nasiri R.; Li J.; Zhang D.; Yang A.; et al. Three-Dimensional Transistor Arrays for Intra- and Inter-Cellular Recording. Nat. Nanotechnol. 2022, 17, 292–300. 10.1038/s41565-021-01040-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu Y.; Li X.; Yang H.; Zhang P.; Wang P.; Sun Y.; Yang F.; Liu W.; Li Y.; Tian Y.; et al. Skin-Interfaced Superhydrophobic Insensible Sweat Sensors for Evaluating Body Thermoregulation and Skin Barrier Functions. ACS Nano 2023, 17, 5588–5599. 10.1021/acsnano.2c11267. [DOI] [PubMed] [Google Scholar]
- Shirzaei Sani E.; Xu C.; Wang C.; Song Y.; Min J.; Tu J.; Solomon S. A.; Li J.; Banks J. L.; Armstrong D. G.; et al. A Stretchable Wireless Wearable Bioelectronic System for Multiplexed Monitoring and Combination Treatment of Infected Chronic Wounds. Sci. Adv. 2023, 9, eadf7388 10.1126/sciadv.adf7388. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Akbar F.; Rivkin B.; Aziz A.; Becker C.; Karnaushenko D. D.; Medina-Sánchez M.; Karnaushenko D.; Schmidt O. G. Self-Sufficient Self-Oscillating Microsystem Driven by Low Power at Low Reynolds Numbers. Sci. Adv. 2021, 7, eabj0767 10.1126/sciadv.abj0767. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wu S.; Hong Y.; Zhao Y.; Yin J.; Zhu Y. Caterpillar-Inspired Soft Crawling Robot with Distributed Programmable Thermal Actuation. Sci. Adv. 2023, 9, eadf8014 10.1126/sciadv.adf8014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang Y.; Adam M. L.; Zhao Y.; Zheng W.; Gao L.; Yin Z.; Zhao H. Machine Learning-Enhanced Flexible Mechanical Sensing. Nano-Micro Lett. 2023, 15, 55. 10.1007/s40820-023-01013-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hoang A. T.; Hu L.; Katiyar A. K.; Ahn J.-H. Two-Dimensional Layered Materials and Heterostructures for Flexible Electronics. Matter 2022, 5, 4116–4132. 10.1016/j.matt.2022.10.016. [DOI] [Google Scholar]
- Sim K.; Rao Z.; Zou Z.; Ershad F.; Lei J.; Thukral A.; Chen J.; Huang Q.-A.; Xiao J.; Yu C. Metal Oxide Semiconductor Nanomembrane-Based Soft Unnoticeable Multifunctional Electronics for Wearable Human-Machine Interfaces. Sci. Adv. 2019, 5, eaav9653 10.1126/sciadv.aav9653. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Song H.; Luo G.; Ji Z.; Bo R.; Xue Z.; Yan D.; Zhang F.; Bai K.; Liu J.; Cheng X.; et al. Highly-Integrated, Miniaturized, Stretchable Electronic Systems Based on Stacked Multilayer Network Materials. Sci. Adv. 2022, 8, eabm3785 10.1126/sciadv.abm3785. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee G.; Zarei M.; Wei Q.; Zhu Y.; Lee S. G. Surface Wrinkling for Flexible and Stretchable Sensors. Small 2022, 18, 2203491 10.1002/smll.202270225. [DOI] [PubMed] [Google Scholar]
- Yang R.; Zhang W.; Tiwari N.; Yan H.; Li T.; Cheng H. Multimodal Sensors with Decoupled Sensing Mechanisms. Adv. Sci. 2022, 9, 2202470 10.1002/advs.202202470. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang Y.; Xu C.; Yu X.; Zhang H.; Han M. Multilayer Flexible Electronics: Manufacturing Approaches and Applications. Mater. Today Phys. 2022, 23, 100647. 10.1016/j.mtphys.2022.100647. [DOI] [Google Scholar]
- Shi J.; Dai Y.; Cheng Y.; Xie S.; Li G.; Liu Y.; Wang J.; Zhang R.; Bai N.; Cai M.; et al. Embedment of Sensing Elements for Robust, Highly Sensitive, and Cross-Talk-Free Iontronic Skins for Robotics Applications. Sci. Adv. 2023, 9, eadf8831 10.1126/sciadv.adf8831. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hwangbo S.; Hu L.; Hoang A. T.; Choi J. Y.; Ahn J. H. Wafer-Scale Monolithic Integration of Full-Colour Micro-Led Display Using Mos(2) Transistor. Nat. Nanotechnol. 2022, 17, 500–506. 10.1038/s41565-022-01102-7. [DOI] [PubMed] [Google Scholar]
- Wazir N.; Liu R.; Ding C.; Wang X.; Ye X.; Lingling X.; Lu T.; Wei L.; Zou B. Vertically Stacked Mose2/Moo2 Nanolayered Photodetectors with Tunable Photoresponses. ACS Appl. Nano Mater. 2020, 3, 7543–7553. 10.1021/acsanm.0c01195. [DOI] [Google Scholar]
- Bo R.; Nasiri N.; Chen H.; Caputo D.; Fu L.; Tricoli A. Low-Voltage High-Performance Uv Photodetectors: An Interplay between Grain Boundaries and Debye Length. ACS Appl. Mater. Interfaces. 2017, 9, 2606–2615. 10.1021/acsami.6b12321. [DOI] [PubMed] [Google Scholar]
- Nasiri N.; Bo R.; Wang F.; Fu L.; Tricoli A. Ultraporous Electron-Depleted Zno Nanoparticle Networks for Highly Sensitive Portable Visible-Blind Uv Photodetectors. Adv. Mater. 2015, 27, 4336–4343. 10.1002/adma.201501517. [DOI] [PubMed] [Google Scholar]
- Chen H.; Bo R.; Shrestha A.; Xin B.; Nasiri N.; Zhou J.; Di Bernardo I.; Dodd A.; Saunders M.; Lipton-Duffin J.; et al. Nio-Zno Nanoheterojunction Networks for Room-Temperature Volatile Organic Compounds Sensing. Adv. Opt. Mater. 2018, 6, 1800677. 10.1002/adom.201800677. [DOI] [Google Scholar]
- Nasiri N.; Bo R.; Chen H.; White T. P.; Fu L.; Tricoli A. Structural Engineering of Nano-Grain Boundaries for Low-Voltage Uv-Photodetectors with Gigantic Photo- to Dark-Current Ratios. Adv. Opt. Mater. 2016, 4, 1787–1795. 10.1002/adom.201600273. [DOI] [Google Scholar]
- Bo R.; Taheri M.; Liu B.; Ricco R.; Chen H.; Amenitsch H.; Fusco Z.; Tsuzuki T.; Yu G.; Ameloot R.; et al. Hierarchical Metal-Organic Framework Films with Controllable Meso/Macroporosity. Adv. Sci. 2020, 7, 2002368. 10.1002/advs.202002368. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nasiri N.; Bo R.; Hung T. F.; Roy V. A. L.; Fu L.; Tricoli A. Tunable Band-Selective Uv-Photodetectors by 3d Self-Assembly of Heterogeneous Nanoparticle Networks. Adv. Funct. Mater. 2016, 26, 7359–7366. 10.1002/adfm.201602195. [DOI] [Google Scholar]
- Fusco Z.; Rahmani M.; Bo R.; Verre R.; Motta N.; Käll M.; Neshev D.; Tricoli A. Nanostructured Dielectric Fractals on Resonant Plasmonic Metasurfaces for Selective and Sensitive Optical Sensing of Volatile Compounds. Adv. Mater. 2018, 30, 1800931. 10.1002/adma.201800931. [DOI] [PubMed] [Google Scholar]
- Chen H.; Zhang M.; Bo R.; Barugkin C.; Zheng J.; Ma Q.; Huang S.; Ho-Baillie A. W. Y.; Catchpole K. R.; Tricoli A. Superior Self-Powered Room-Temperature Chemical Sensing with Light-Activated Inorganic Halides Perovskites. Small 2018, 14, 1702571. 10.1002/smll.201702571. [DOI] [PubMed] [Google Scholar]
- Chen H.; Zhang M.; Fu X.; Fusco Z.; Bo R.; Xing B.; Nguyen H. T.; Barugkin C.; Zheng J.; Lau C. F. J.; et al. Light-Activated Inorganic Cspbbr2i Perovskite for Room-Temperature Self-Powered Chemical Sensing. Phys. Chem. Chem. Phys. 2019, 21, 24187–24193. 10.1039/C9CP03059J. [DOI] [PubMed] [Google Scholar]
- Yang L.; Liu C.; Yuan W.; Meng C.; Dutta A.; Chen X.; Guo L.; Niu G.; Cheng H. Fully Stretchable, Porous Mxene-Graphene Foam Nanocomposites for Energy Harvesting and Self-Powered Sensing. Nano Energy 2022, 103, 107807. 10.1016/j.nanoen.2022.107807. [DOI] [Google Scholar]
- Han W. B.; Heo S.-Y.; Kim D.; Yang S. M.; Ko G.-J.; Lee G. J.; Kim D.-J.; Rajaram K.; Lee J. H.; Shin J.-W.; et al. Zebra-Inspired Stretchable, Biodegradable Radiation Modulator for All-Day Sustainable Energy Harvesters. Sci. Adv. 2023, 9, eadf5883 10.1126/sciadv.adf5883. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ko J. H.; Kim S. H.; Kim M. S.; Heo S. Y.; Yoo Y. J.; Kim Y. J.; Lee H.; Song Y. M. Lithography-Free, Large-Area Spatially Segmented Disordered Structure for Light Harvesting in Photovoltaic Modules. ACS Appl. Mater. Interfaces. 2022, 14, 44419–44428. 10.1021/acsami.2c12131. [DOI] [PubMed] [Google Scholar]
- Tran-Phu T.; Fusco Z.; Di Bernardo I.; Lipton-Duffin J.; Toe C. Y.; Daiyan R.; Gengenbach T.; Lin C.-H.; Bo R.; Nguyen H. T.; et al. Understanding the Role of Vanadium Vacancies in Bivo4 for Efficient Photoelectrochemical Water Oxidation. Chem. Mater. 2021, 33, 3553–3565. 10.1021/acs.chemmater.0c04866. [DOI] [Google Scholar]
- Tran-Phu T.; Chen H.; Bo R.; Di Bernardo I.; Fusco Z.; Simonov A. N.; Tricoli A. High-Temperature One-Step Synthesis of Efficient Nanostructured Bismuth Vanadate Photoanodes for Water Oxidation. Energy Technol. 2019, 7, 1801052. 10.1002/ente.201801052. [DOI] [Google Scholar]
- Bo R.; Zhang F.; Bu S.; Nasiri N.; Di Bernardo I.; Tran-Phu T.; Shrestha A.; Chen H.; Taheri M.; Qi S.; et al. One-Step Synthesis of Porous Transparent Conductive Oxides by Hierarchical Self-Assembly of Aluminum-Doped Zno Nanoparticles. ACS Appl. Mater. Interfaces. 2020, 12, 9589–9599. 10.1021/acsami.9b19423. [DOI] [PubMed] [Google Scholar]
- Cheng X.; Zhang F.; Bo R.; Shen Z.; Pang W.; Jin T.; Song H.; Xue Z.; Zhang Y. An Anti-Fatigue Design Strategy for 3d Ribbon-Shaped Flexible Electronics. Adv. Mater. 2021, 33, 2102684. 10.1002/adma.202102684. [DOI] [PubMed] [Google Scholar]
- Chen H.; Zhang M.; Xing B.; Fu X.; Bo R.; Mulmudi H. K.; Huang S.; Ho-Baillie A. W. Y.; Catchpole K. R.; Tricoli A. Superior Self-Charged and -Powered Chemical Sensing with High Performance for No2 Detection at Room Temperature. Adv. Opt. Mater. 2020, 8, 1901863. 10.1002/adom.201901863. [DOI] [Google Scholar]
- Guo X.; Li H.; Yeop Ahn B.; Duoss E. B.; Hsia K. J.; Lewis J. A.; Nuzzo R. G. Two- and Three-Dimensional Folding of Thin Film Single-Crystalline Silicon for Photovoltaic Power Applications. Proc. Natl. Acad. Sci. U.S.A. 2009, 106, 20149–20154. 10.1073/pnas.0907390106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ling Y.; Pang W.; Liu J.; Page M.; Xu Y.; Zhao G.; Stalla D.; Xie J.; Zhang Y.; Yan Z. Bioinspired Elastomer Composites with Programmed Mechanical and Electrical Anisotropies. Nat. Commun. 2022, 13, 524. 10.1038/s41467-022-28185-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Xu J.; Wu H.-C.; Zhu C.; Ehrlich A.; Shaw L.; Nikolka M.; Wang S.; Molina-Lopez F.; Gu X.; Luo S.; et al. Multi-Scale Ordering in Highly Stretchable Polymer Semiconducting Films. Nat. Mater. 2019, 18, 594–601. 10.1038/s41563-019-0340-5. [DOI] [PubMed] [Google Scholar]
- Bai Y.; Wang H.; Xue Y.; Pan Y.; Kim J.-T.; Ni X.; Liu T.-L.; Yang Y.; Han M.; Huang Y.; et al. A Dynamically Reprogrammable Surface with Self-Evolving Shape Morphing. Nature 2022, 609, 701–708. 10.1038/s41586-022-05061-w. [DOI] [PubMed] [Google Scholar]
- Liu Y.; He K.; Chen G.; Leow W. R.; Chen X. Nature-Inspired Structural Materials for Flexible Electronic Devices. Chem. Rev. 2017, 117, 12893–12941. 10.1021/acs.chemrev.7b00291. [DOI] [PubMed] [Google Scholar]
- Park Y.; Chung T. S.; Lee G.; Rogers J. A. Materials Chemistry of Neural Interface Technologies and Recent Advances in Three-Dimensional Systems. Chem. Rev. 2022, 122, 5277–5316. 10.1021/acs.chemrev.1c00639. [DOI] [PubMed] [Google Scholar]
- Min J.; Tu J.; Xu C.; Lukas H.; Shin S.; Yang Y.; Solomon S. A.; Mukasa D.; Gao W. Skin-Interfaced Wearable Sweat Sensors for Precision Medicine. Chem. Rev. 2023, 123, 5049. 10.1021/acs.chemrev.2c00823. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Choi S.; Lee H.; Ghaffari R.; Hyeon T.; Kim D.-H. Recent Advances in Flexible and Stretchable Bio-Electronic Devices Integrated with Nanomaterials. Adv. Mater. 2016, 28, 4203–4218. 10.1002/adma.201504150. [DOI] [PubMed] [Google Scholar]
- Li J.; Esteban-Fernandez de Avila B.; Gao W.; Zhang L.; Wang J. Micro/Nanorobots for Biomedicine: Delivery, Surgery, Sensing, and Detoxification. Sci. Robot. 2017, 2, eaam6431 10.1126/scirobotics.aam6431. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ning X.; Wang X.; Zhang Y.; Yu X.; Choi D.; Zheng N.; Kim D. S.; Huang Y.; Zhang Y.; Rogers J. A. Assembly of Advanced Materials into 3d Functional Structures by Methods Inspired by Origami and Kirigami: A Review. Adv. Mater. Interfaces 2018, 5, 1800284. 10.1002/admi.201800284. [DOI] [Google Scholar]
- Cohen-Karni T.; Qing Q.; Li Q.; Fang Y.; Lieber C. M. Graphene and Nanowire Transistors for Cellular Interfaces and Electrical Recording. Nano Lett. 2010, 10, 1098–1102. 10.1021/nl1002608. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tian B.; Cohen-Karni T.; Qing Q.; Duan X.; Xie P.; Lieber C. M. Three-Dimensional, Flexible Nanoscale Field-Effect Transistors as Localized Bioprobes. Science 2010, 329, 830–834. 10.1126/science.1192033. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Huang G.; Mei Y.; Thurmer D. J.; Coric E.; Schmidt O. G. Rolled-up Transparent Microtubes as Two-Dimensionally Confined Culture Scaffolds of Individual Yeast Cells. Lab Chip 2009, 9, 263–268. 10.1039/B810419K. [DOI] [PubMed] [Google Scholar]
- Xi W.; Schmidt C. K.; Sanchez S.; Gracias D. H.; Carazo-Salas R. E.; Jackson S. P.; Schmidt O. G. Rolled-up Functionalized Nanomembranes as Three-Dimensional Cavities for Single Cell Studies. Nano Lett. 2014, 14, 4197–4204. 10.1021/nl4042565. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Duan X.; Gao R.; Xie P.; Cohen-Karni T.; Qing Q.; Choe H. S.; Tian B.; Jiang X.; Lieber C. M. Intracellular Recordings of Action Potentials by an Extracellular Nanoscale Field-Effect Transistor. Nat. Nanotechnol. 2012, 7, 174–179. 10.1038/nnano.2011.223. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fusco Z.; Bo R.; Wang Y.; Motta N.; Chen H.; Tricoli A. Self-Assembly of Au Nano-Islands with Tuneable Organized Disorder for Highly Sensitive Sers. J. Mater. Chem. C 2019, 7, 6308–6316. 10.1039/C9TC01231A. [DOI] [Google Scholar]
- Egunov A. I.; Dou Z.; Karnaushenko D. D.; Hebenstreit F.; Kretschmann N.; Akgün K.; Ziemssen T.; Karnaushenko D.; Medina-Sánchez M.; Schmidt O. G. Impedimetric Microfluidic Sensor-in-a-Tube for Label-Free Immune Cell Analysis. Small 2021, 17, 2002549. 10.1002/smll.202002549. [DOI] [PubMed] [Google Scholar]
- Jin Q.; Bhatta A.; Pagaduan J. V.; Chen X.; West-Foyle H.; Liu J.; Hou A.; Berkowitz D.; Kuo S. C.; Askin F. B.; et al. Biomimetic Human Small Muscular Pulmonary Arteries. Sci. Adv. 2020, 6, eaaz2598 10.1126/sciadv.aaz2598. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Herzer R.; Gebert A.; Hempel U.; Hebenstreit F.; Oswald S.; Damm C.; Schmidt O. G.; Medina-Sánchez M. Rolled-up Metal Oxide Microscaffolds to Study Early Bone Formation at Single Cell Resolution. Small 2021, 17, 2005527. 10.1002/smll.202005527. [DOI] [PubMed] [Google Scholar]
- Park J.; Kalinin Y. V.; Kadam S.; Randall C. L.; Gracias D. H. Design for a Lithographically Patterned Bioartificial Endocrine Pancreas. Artif. Organs 2013, 37, 1059–1067. 10.1111/aor.12131. [DOI] [PubMed] [Google Scholar]
- Kim S. J.; Cho K. W.; Cho H. R.; Wang L.; Park S. Y.; Lee S. E.; Hyeon T.; Lu N.; Choi S. H.; Kim D.-H. Stretchable and Transparent Biointerface Using Cell-Sheet-Graphene Hybrid for Electrophysiology and Therapy of Skeletal Muscle. Adv. Funct. Mater. 2016, 26, 3207–3217. 10.1002/adfm.201504578. [DOI] [Google Scholar]
- Kwak J. W.; Han M.; Xie Z.; Chung H. U.; Lee J. Y.; Avila R.; Yohay J.; Chen X.; Liang C.; Patel M.; et al. Wireless Sensors for Continuous, Multimodal Measurements at the Skin Interface with Lower Limb Prostheses. Sci. Transl. Med. 2020, 12, eabc4327 10.1126/scitranslmed.abc4327. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Guan Y.-S.; Ershad F.; Rao Z.; Ke Z.; da Costa E. C.; Xiang Q.; Lu Y.; Wang X.; Mei J.; Vanderslice P.; et al. Elastic Electronics Based on Micromesh-Structured Rubbery Semiconductor Films. Nat. Electron. 2022, 5, 881–892. 10.1038/s41928-022-00874-z. [DOI] [Google Scholar]
- Kim B.; Soepriatna A. H.; Park W.; Moon H.; Cox A.; Zhao J.; Gupta N. S.; Park C. H.; Kim K.; Jeon Y.; et al. Rapid Custom Prototyping of Soft Poroelastic Biosensor for Simultaneous Epicardial Recording and Imaging. Nat. Commun. 2021, 12, 3710. 10.1038/s41467-021-23959-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang Y.; Li X.; Fan S.; Feng X.; Cao K.; Ge Q.; Gao L.; Lu Y. Three-Dimensional Stretchable Microelectronics by Projection Microstereolithography (Pμsl). ACS Appl. Mater. Interfaces. 2021, 13, 8901–8908. 10.1021/acsami.0c20162. [DOI] [PubMed] [Google Scholar]
- Bang J.; Ahn J.; Zhang J.; Ko T. H.; Park B.; Lee Y. M.; Jung B. K.; Lee S. Y.; Ok J.; Kim B. H.; et al. Stretchable and Directly Patternable Double-Layer Structure Electrodes with Complete Coverage. ACS Nano 2022, 16, 12134–12144. 10.1021/acsnano.2c02664. [DOI] [PubMed] [Google Scholar]
- Kang M.; Jeong H.; Park S.-W.; Hong J.; Lee H.; Chae Y.; Yang S.; Ahn J.-H. Wireless Graphene-Based Thermal Patch for Obtaining Temperature Distribution and Performing Thermography. Sci. Adv. 2022, 8, eabm6693 10.1126/sciadv.abm6693. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dutta A.; Cheng H. Pathway of Transient Electronics Towards Connected Biomedical Applications. Nanoscale 2023, 15, 4236–4249. 10.1039/D2NR06068J. [DOI] [PubMed] [Google Scholar]
- Xu H.; Medina-Sanchez M.; Magdanz V.; Schwarz L.; Hebenstreit F.; Schmidt O. G. Sperm-Hybrid Micromotor for Targeted Drug Delivery. ACS Nano 2018, 12, 327–337. 10.1021/acsnano.7b06398. [DOI] [PubMed] [Google Scholar]
- Wu Z.; Chen Y.; Mukasa D.; Pak O. S.; Gao W. Medical Micro/Nanorobots in Complex Media. Chem. Soc. Rev. 2020, 49, 8088–8112. 10.1039/D0CS00309C. [DOI] [PubMed] [Google Scholar]
- Park W.; Nguyen V. P.; Jeon Y.; Kim B.; Li Y.; Yi J.; Kim H.; Leem J. W.; Kim Y. L.; Kim D. R.; et al. Biodegradable Silicon Nanoneedles for Ocular Drug Delivery. Sci. Adv. 2022, 8, eabn1772 10.1126/sciadv.abn1772. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Reeder J. T.; Xie Z.; Yang Q.; Seo M.-H.; Yan Y.; Deng Y.; Jinkins K. R.; Krishnan S. R.; Liu C.; McKay S.; et al. Soft, Bioresorbable Coolers for Reversible Conduction Block of Peripheral Nerves. Science 2022, 377, 109–115. 10.1126/science.abl8532. [DOI] [PubMed] [Google Scholar]
- Song E.; Li J.; Won S. M.; Bai W.; Rogers J. A. Materials for Flexible Bioelectronic Systems as Chronic Neural Interfaces. Nat. Mater. 2020, 19, 590–603. 10.1038/s41563-020-0679-7. [DOI] [PubMed] [Google Scholar]
- Kim K.; Kim H. J.; Zhang H.; Park W.; Meyer D.; Kim M. K.; Kim B.; Park H.; Xu B.; Kollbaum P.; et al. All-Printed Stretchable Corneal Sensor on Soft Contact Lenses for Noninvasive and Painless Ocular Electrodiagnosis. Nat. Commun. 2021, 12, 1544. 10.1038/s41467-021-21916-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Theocharidis G.; Yuk H.; Roh H.; Wang L.; Mezghani I.; Wu J.; Kafanas A.; Contreras M.; Sumpio B.; Li Z.; et al. A Strain-Programmed Patch for the Healing of Diabetic Wounds. Nat. Biomed. Eng. 2022, 6, 1118–1133. 10.1038/s41551-022-00905-2. [DOI] [PubMed] [Google Scholar]
- Zhang J.; Kim K.; Kim H. J.; Meyer D.; Park W.; Lee S. A.; Dai Y.; Kim B.; Moon H.; Shah J. V.; et al. Smart Soft Contact Lenses for Continuous 24-h Monitoring of Intraocular Pressure in Glaucoma Care. Nat. Commun. 2022, 13, 5518. 10.1038/s41467-022-33254-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kim S.; Oh Y. S.; Lee K.; Kim S.; Maeng W.-Y.; Kim K. S.; Kim G.-B.; Cho S.; Han H.; Park H.; et al. Battery-Free, Wireless, Cuff-Type, Multimodal Physical Sensor for Continuous Temperature and Strain Monitoring of Nerve. Small 2023, 19, 2206839. 10.1002/smll.202206839. [DOI] [PubMed] [Google Scholar]
- Barbot A.; Tan H.; Power M.; Seichepine F.; Yang G.-Z. Floating Magnetic Microrobots for Fiber Functionalization. Sci. Robot. 2019, 4, eaax8336 10.1126/scirobotics.aax8336. [DOI] [PubMed] [Google Scholar]
- Rivkin B.; Becker C.; Singh B.; Aziz A.; Akbar F.; Egunov A.; Karnaushenko D. D.; Naumann R.; Schäfer R.; Medina-Sánchez M.; et al. Electronically Integrated Microcatheters Based on Self-Assembling Polymer Films. Sci. Adv. 2021, 7, eabl5408 10.1126/sciadv.abl5408. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Heng W.; Solomon S.; Gao W. Flexible Electronics and Devices as Human-Machine Interfaces for Medical Robotics. Adv. Mater. 2022, 34, 2107902 10.1002/adma.202107902. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chang T.; Akin S.; Kim M. K.; Murray L.; Kim B.; Cho S.; Huh S.; Teke S.; Couetil L.; Jun M. B.; et al. A Programmable Dual-Regime Spray for Large-Scale and Custom-Designed Electronic Textiles. Adv. Mater. 2022, 34, 2108021 10.1002/adma.202108021. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhou C.; Yang Y.; Wang J.; Wu Q.; Gu Z.; Zhou Y.; Liu X.; Yang Y.; Tang H.; Ling Q.; et al. Ferromagnetic Soft Catheter Robots for Minimally Invasive Bioprinting. Nat. Commun. 2021, 12, 5072. 10.1038/s41467-021-25386-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yang Y.; Wang J.; Wang L.; Wu Q.; Ling L.; Yang Y.; Ning S.; Xie Y.; Cao Q.; Li L.; et al. Magnetic Soft Robotic Bladder for Assisted Urination. Sci. Adv. 2022, 8, eabq1456 10.1126/sciadv.abq1456. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang L.; Kim Y.; Guo C. F.; Zhao X. Hard-Magnetic Elastica. J. Mech. Phys. Solids 2020, 142, 104045. 10.1016/j.jmps.2020.104045. [DOI] [Google Scholar]
- Wang L.; Zheng D.; Harker P.; Patel A. B.; Guo C. F.; Zhao X. Evolutionary Design of Magnetic Soft Continuum Robots. Proc. Natl. Acad. Sci. U. S. A. 2021, 118, e2021922118 10.1073/pnas.2021922118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shao L.-H.; Qu X.; Wang T.; Cui Z.; Liu Y.; Zhu Y. Interfacial Shear Stress Transfer between Elastoplastic Fiber and Elastic Matrix. J. Mech. Phys. Solids 2023, 173, 105218. 10.1016/j.jmps.2023.105218. [DOI] [Google Scholar]
- Zhang L.; Zhang Z.; Weisbecker H.; Yin H.; Liu Y.; Han T.; Guo Z.; Berry M.; Yang B.; Guo X.; et al. 3d Morphable Systems Via Deterministic Microfolding for Vibrational Sensing, Robotic Implants, and Reconfigurable Telecommunication. Sci. Adv. 2022, 8, eade0838 10.1126/sciadv.ade0838. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen X.; Jian W.; Wang Z.; Ai J.; Kang Y.; Sun P.; Wang Z.; Ma Y.; Wang H.; Chen Y.; et al. Wrap-Like Transfer Printing for Three-Dimensional Curvy Electronics. Sci. Adv. 2023, 9, eadi0357 10.1126/sciadv.adi0357. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Huber R.; Kern F.; Karnaushenko D. D.; Eisner E.; Lepucki P.; Thampi A.; Mirhajivarzaneh A.; Becker C.; Kang T.; Baunack S.; et al. Tailoring Electron Beams with High-Frequency Self-Assembled Magnetic Charged Particle Micro Optics. Nat. Commun. 2022, 13, 3220. 10.1038/s41467-022-30703-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Becker C.; Karnaushenko D.; Kang T.; Karnaushenko D. D.; Faghih M.; Mirhajivarzaneh A.; Schmidt O. G. Self-Assembly of Highly Sensitive 3d Magnetic Field Vector Angular Encoders. Sci. Adv. 2019, 5, eaay7459 10.1126/sciadv.aay7459. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Huang W.; Yang Z.; Kraman M. D.; Wang Q.; Ou Z.; Rojo M. M.; Yalamarthy A. S.; Chen V.; Lian F.; Ni J. H.; et al. Monolithic Mtesla-Level Magnetic Induction by Self-Rolled-up Membrane Technology. Sci. Adv. 2020, 6, eaay4508 10.1126/sciadv.aay4508. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lepucki P.; Egunov A. I.; Rosenkranz M.; Huber R.; Mirhajivarzaneh A.; Karnaushenko D. D.; Dioguardi A. P.; Karnaushenko D.; Büchner B.; Schmidt O. G.; et al. Self-Assembled Rolled-up Microcoils for Nl Microfluidics Nmr Spectroscopy. Adv. Mater. Technol. 2021, 6, 2000679. 10.1002/admt.202000679. [DOI] [Google Scholar]
- Bermúdez Ureña E.; Mei Y.; Coric E.; Makarov D.; Albrecht M.; Schmidt O. G. Fabrication of Ferromagnetic Rolled-up Microtubes for Magnetic Sensors on Fluids. J. Phys. D: Appl. Phys. 2009, 42, 055001. 10.1088/0022-3727/42/5/055001. [DOI] [Google Scholar]
- Kim T.; Price J. S.; Grede A.; Lee S.; Choi G.; Guan W.; Jackson T. N.; Giebink N. C. Kirigami-Inspired 3d Organic Light-Emitting Diode (Oled) Lighting Concepts. Adv. Mater. Technol. 2018, 3, 1800067. 10.1002/admt.201800067. [DOI] [Google Scholar]
- Ju H.; Park H.; Kim N.; Lim J.; Jung D.; Lee J. A Locally Actuatable Soft Robotic Film for Actively Reconfiguring Shapes of Flexible Electronics. Soft Robot. 2022, 9, 767–775. 10.1089/soro.2021.0046. [DOI] [PubMed] [Google Scholar]
- Lee Y.; Kim B. J.; Hu L.; Hong J.; Ahn J.-H. Morphable 3d Structure for Stretchable Display. Mater. Today 2022, 53, 51–57. 10.1016/j.mattod.2022.01.017. [DOI] [Google Scholar]
- Oh S.; Lee S.; Byun S. H.; Lee S.; Kim C. Y.; Yea J.; Chung S.; Li S.; Jang K. I.; Kang J.; et al. 3d Shape-Morphing Display Enabled by Electrothermally Responsive, Stiffness-Tunable Liquid Metal Platform with Stretchable Electroluminescent Device. Adv. Funct. Mater. 2023. 10.1002/adfm.202214766 [DOI] [Google Scholar]
- Deng Y.; Liu W.; Cheung Y. K.; Li Y.; Hong W.; Yu H. Curved Display Based on Programming Origami Tessellations. Microsyst. Nanoeng. 2021, 7, 101. 10.1038/s41378-021-00319-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Song J. K.; Kim J.; Yoon J.; Koo J. H.; Jung H.; Kang K.; Sunwoo S. H.; Yoo S.; Chang H.; Jo J.; et al. Stretchable Colour-Sensitive Quantum Dot Nanocomposites for Shape-Tunable Multiplexed Phototransistor Arrays. Nat. Nanotechnol. 2022, 17, 849–856. 10.1038/s41565-022-01160-x. [DOI] [PubMed] [Google Scholar]
- Zhang Q.; Xu M.; Zhou L.; Liu S.; Wang W.; Zhang L.; Xie W.; Yu C. A Flexible Organic Mechanoluminophore Device. Nat. Commun. 2023, 14, 1257. 10.1038/s41467-023-36916-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nasiri N.; Bo R.; Fu L.; Tricoli A. Three-Dimensional Nano-Heterojunction Networks: A Highly Performing Structure for Fast Visible-Blind Uv Photodetectors. Nanoscale 2017, 9, 2059–2067. 10.1039/C6NR08425G. [DOI] [PubMed] [Google Scholar]
- Kang J.; Yoo Y. J.; Ko J. H.; Mahmud A. A.; Song Y. M. Trilayered Gires-Tournois Resonator with Ultrasensitive Slow-Light Condition for Colorimetric Detection of Bioparticles. Nanomaterials 2023, 13, 319. 10.3390/nano13020319. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gao L.; Zhang Y.; Zhang H.; Doshay S.; Xie X.; Luo H.; Shah D.; Shi Y.; Xu S.; Fang H.; et al. Optics and Nonlinear Buckling Mechanics in Large-Area, Highly Stretchable Arrays of Plasmonic Nanostructures. ACS Nano 2015, 9, 5968–5975. 10.1021/acsnano.5b00716. [DOI] [PubMed] [Google Scholar]
- Li T.; Hantusch M.; Qu J.; Bandari V. K.; Knupfer M.; Zhu F.; Schmidt O. G. On-Chip Integrated Process-Programmable Sub-10 Nm Thick Molecular Devices Switching between Photomultiplication and Memristive Behaviour. Nat. Commun. 2022, 13, 2875. 10.1038/s41467-022-30498-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang Y.; Zhou W.; Cao K.; Hu X.; Gao L.; Lu Y. Architectured Graphene and Its Composites: Manufacturing and Structural Applications. Compos. Part A Appl. Sci. Manuf. 2021, 140, 106177. 10.1016/j.compositesa.2020.106177. [DOI] [Google Scholar]
- Ko J. H.; Yoo Y. J.; Lee Y.; Jeong H.-H.; Song Y. M. A Review of Tunable Photonics: Optically Active Materials and Applications from Visible to Terahertz. iScience 2022, 25, 104727. 10.1016/j.isci.2022.104727. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen S.; Liu Z.; Du H.; Tang C.; Ji C.-Y.; Quan B.; Pan R.; Yang L.; Li X.; Gu C.; et al. Electromechanically Reconfigurable Optical Nano-Kirigami. Nat. Commun. 2021, 12, 1299. 10.1038/s41467-021-21565-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liao F.; Zhou Z.; Kim B. J.; Chen J.; Wang J.; Wan T.; Zhou Y.; Hoang A. T.; Wang C.; Kang J.; et al. Bioinspired in-Sensor Visual Adaptation for Accurate Perception. Nat. Electron. 2022, 5, 84–91. 10.1038/s41928-022-00713-1. [DOI] [Google Scholar]
- Huang C. C.; Wu X.; Liu H.; Aldalali B.; Rogers J. A.; Jiang H. Large-Field-of-View Wide-Spectrum Artificial Reflecting Superposition Compound Eyes. Small 2014, 10, 3050–3057. 10.1002/smll.201400037. [DOI] [PubMed] [Google Scholar]
- Kim M.; Lee G. J.; Choi C.; Kim M. S.; Lee M.; Liu S.; Cho K. W.; Kim H. M.; Cho H.; Choi M. K.; et al. An Aquatic-Vision-Inspired Camera Based on a Monocentric Lens and a Silicon Nanorod Photodiode Array. Nat. Electron. 2020, 3, 546–553. 10.1038/s41928-020-0429-5. [DOI] [Google Scholar]
- Kim M.; Chang S.; Kim M.; Yeo J.-E.; Kim M. S.; Lee G. J.; Kim D.-H.; Song Y. M. Cuttlefish Eye-Inspired Artificial Vision for High-Quality Imaging under Uneven Illumination Conditions. Sci. Robot. 2023, 8, eade4698 10.1126/scirobotics.ade4698. [DOI] [PubMed] [Google Scholar]
- Singh D.; Kutbee A. T.; Ghoneim M. T.; Hussain A. M.; Hussain M. M. Strain-Induced Rolled Thin Films for Lightweight Tubular Thermoelectric Generators. Adv. Mater. Technol. 2018, 3, 1700192. 10.1002/admt.201700192. [DOI] [Google Scholar]
- Li Y.; Qu J.; Li F.; Qu Z.; Tang H.; Liu L.; Zhu M.; Schmidt O. G. Advanced Architecture Designs Towards High-Performance 3d Microbatteries. Nano Matter. Sci. 2021, 3, 140–153. 10.1016/j.nanoms.2020.10.004. [DOI] [Google Scholar]
- Liu L.; Huang S.; Shi W.; Sun X.; Pang J.; Lu Q.; Yang Y.; Xi L.; Deng L.; Oswald S.; et al. Single “Swiss-Roll” Microelectrode Elucidates the Critical Role of Iron Substitution in Conversion-Type Oxides. Sci. Adv. 2022, 8, eadd6596 10.1126/sciadv.add6596. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu B.; Bo R.; Taheri M.; Di Bernardo I.; Motta N.; Chen H.; Tsuzuki T.; Yu G.; Tricoli A. Metal-Organic Frameworks/Conducting Polymer Hydrogel Integrated Three-Dimensional Free-Standing Monoliths as Ultrahigh Loading Li-S Battery Electrodes. Nano Lett. 2019, 19, 4391–4399. 10.1021/acs.nanolett.9b01033. [DOI] [PubMed] [Google Scholar]
- Li F.; Wang J.; Liu L.; Qu J.; Li Y.; Bandari V. K.; Karnaushenko D.; Becker C.; Faghih M.; Kang T.; et al. Self-Assembled Flexible and Integratable 3d Microtubular Asymmetric Supercapacitors. Adv. Sci. 2019, 6, 1901051. 10.1002/advs.201901051. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gao L.; Wang Y.; Hu X.; Zhou W.; Cao K.; Wang Y.; Wang W.; Lu Y. Cellular Carbon-Film-Based Flexible Sensor and Waterproof Supercapacitors. ACS Appl. Mater. Interfaces. 2019, 11, 26288–26297. 10.1021/acsami.9b09438. [DOI] [PubMed] [Google Scholar]
- Shi J.; Wang L.; Dai Z.; Zhao L.; Du M.; Li H.; Fang Y. Multiscale Hierarchical Design of a Flexible Piezoresistive Pressure Sensor with High Sensitivity and Wide Linearity Range. Small 2018, 14, 1800819. 10.1002/smll.201800819. [DOI] [PubMed] [Google Scholar]
- Xu H.; Gao L.; Wang Y.; Cao K.; Hu X.; Wang L.; Mu M.; Liu M.; Zhang H.; Wang W.; et al. Flexible Waterproof Piezoresistive Pressure Sensors with Wide Linear Working Range Based on Conductive Fabrics. Nano-Micro Lett. 2020, 12, 159. 10.1007/s40820-020-00498-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhao J.; Li W.; Guo X.; Wang H.; Rogers J. A.; Huang Y. Theoretical Modeling of Tunable Vibrations of Three-Dimensional Serpentine Structures for Simultaneous Measurement of Adherent Cell Mass and Modulus. MRS Bull. 2021, 46, 107–114. 10.1557/s43577-021-00043-1. [DOI] [Google Scholar]
- Zhao J.; Zhang F.; Guo X.; Huang Y.; Zhang Y.; Wang H. Torsional Deformation Dominated Buckling of Serpentine Structures to Form Three-Dimensional Architectures with Ultra-Low Rigidity. J. Mech. Phys. Solids 2021, 155, 104568. 10.1016/j.jmps.2021.104568. [DOI] [Google Scholar]
- Zhao J.Postbuckling Analysis of Ultra-Low Rigidity Serpentine Structures. J. Appl. Mech. 2022, 89. 10.1115/1.4053397 [DOI] [Google Scholar]
- Yang L.; Wang Z.; Wang H.; Jin B.; Meng C.; Chen X.; Li R.; Wang H.; Xin M.; Zhao Z.; et al. Self-Healing, Reconfigurable, Thermal-Switching, Transformative Electronics for Health Monitoring. Adv. Mater. 2023, 35, 2207742 10.1002/adma.202207742. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mohanty S.; Jin Q.; Furtado G. P.; Ghosh A.; Pahapale G.; Khalil I. S. M.; Gracias D. H.; Misra S. Bidirectional Propulsion of Arc-Shaped Microswimmers Driven by Precessing Magnetic Fields. Adv. Intell. Syst. 2020, 2, 2000064. 10.1002/aisy.202000064. [DOI] [Google Scholar]
- Miskin M. Z.; Cortese A. J.; Dorsey K.; Esposito E. P.; Reynolds M. F.; Liu Q.; Cao M.; Muller D. A.; McEuen P. L.; Cohen I. Electronically Integrated, Mass-Manufactured, Microscopic Robots. Nature 2020, 584, 557–561. 10.1038/s41586-020-2626-9. [DOI] [PubMed] [Google Scholar]
- Mao G.; Schiller D.; Danninger D.; Hailegnaw B.; Hartmann F.; Stockinger T.; Drack M.; Arnold N.; Kaltenbrunner M. Ultrafast Small-Scale Soft Electromagnetic Robots. Nat. Commun. 2022, 13, 4456. 10.1038/s41467-022-32123-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yi S.; Wang L.; Chen Z.; Wang J.; Song X.; Liu P.; Zhang Y.; Luo Q.; Peng L.; Wu Z.; et al. High-Throughput Fabrication of Soft Magneto-Origami Machines. Nat. Commun. 2022, 13, 4177. 10.1038/s41467-022-31900-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu Q.; Shen Z.; Liu Z.; Shuai Y.; Lv Z.; Jin T.; Cheng X.; Zhang Y. Probability-Based Analyses of the Snap-through in Cage-Shaped Mesostructures under out-of-Plane Compressions. Acta Mech. Solida Sin 2023, 36, 569–581. 10.1007/s10338-023-00399-8. [DOI] [Google Scholar]
- Yan D.; Chang J.; Zhang H.; Liu J.; Song H.; Xue Z.; Zhang F.; Zhang Y. Soft Three-Dimensional Network Materials with Rational Bio-Mimetic Designs. Nat. Commun. 2020, 11, 1180. 10.1038/s41467-020-14996-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ma Q.; Cheng H.; Jang K. I.; Luan H.; Hwang K. C.; Rogers J. A.; Huang Y.; Zhang Y. A Nonlinear Mechanics Model of Bio-Inspired Hierarchical Lattice Materials Consisting of Horseshoe Microstructures. J. Mech. Phys. Solids 2016, 90, 179–202. 10.1016/j.jmps.2016.02.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu J.; Yan D.; Zhang Y. Mechanics of Unusual Soft Network Materials with Rotatable Structural Nodes. J. Mech. Phys. Solids 2021, 146, 104210. 10.1016/j.jmps.2020.104210. [DOI] [Google Scholar]