Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

A universal interface for plug-and-play assembly of stretchable devices

Abstract

Stretchable hybrid devices have enabled high-fidelity implantable1,2,3 and on-skin4,5,6 monitoring of physiological signals. These devices typically contain soft modules that match the mechanical requirements in humans7,8 and soft robots9,10, rigid modules containing Si-based microelectronics11,12 and protective encapsulation modules13,14. To make such a system mechanically compliant, the interconnects between the modules need to tolerate stress concentration that may limit their stretching and ultimately cause debonding failure15,16,17. Here, we report a universal interface that can reliably connect soft, rigid and encapsulation modules together to form robust and highly stretchable devices in a plug-and-play manner. The interface, consisting of interpenetrating polymer and metal nanostructures, connects modules by simply pressing without using pastes. Its formation is depicted by a biphasic network growth model. Soft–soft modules joined by this interface achieved 600% and 180% mechanical and electrical stretchability, respectively. Soft and rigid modules can also be electrically connected using the above interface. Encapsulation on soft modules with this interface is strongly adhesive with an interfacial toughness of 0.24 N mm−1. As a proof of concept, we use this interface to assemble stretchable devices for in vivo neuromodulation and on-skin electromyography, with high signal quality and mechanical resistance. We expect such a plug-and-play interface to simplify and accelerate the development of on-skin and implantable stretchable devices.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: BIND connection for stretchable hybrid device.
Fig. 2: Structural analysis of BIND interface and connection.
Fig. 3: In vivo neuromodulation stretchable device assembled through plug-and-play BIND connections.
Fig. 4: A 21-channel on-skin EMG electrode array assembled using plug-and-play BIND connections.

Similar content being viewed by others

Data availability

The data supporting the findings of this study are available within the article and its Supplementary Information. Other raw data are available from the corresponding authors upon reasonable request. Source data are provided with this paper.

References

  1. Squair, J. W. et al. Neuroprosthetic baroreflex controls haemodynamics after spinal cord injury. Nature 590, 308–314 (2021).

    Article  ADS  CAS  PubMed  Google Scholar 

  2. Park, S. I. et al. Soft, stretchable, fully implantable miniaturized optoelectronic systems for wireless optogenetics. Nat. Biotechnol. 33, 1280–1286 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Choi, S. et al. Highly conductive, stretchable and biocompatible Ag–Au core–sheath nanowire composite for wearable and implantable bioelectronics. Nat. Nanotechnol. 13, 1048–1056 (2018).

    Article  ADS  CAS  PubMed  Google Scholar 

  4. Matsuhisa, N. et al. Printable elastic conductors by in situ formation of silver nanoparticles from silver flakes. Nat. Mater. 16, 834–840 (2017).

    Article  ADS  CAS  PubMed  Google Scholar 

  5. Hua, Q. et al. Skin-inspired highly stretchable and conformable matrix networks for multifunctional sensing. Nat. Commun. 9, 244 (2018).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  6. Koh, A. et al. A soft, wearable microfluidic device for the capture, storage, and colorimetric sensing of sweat. Sci. Transl Med. 8, 366ra165 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  7. Kim, C.-C. et al. Highly stretchable, transparent ionic touch panel. Science 353, 682–687 (2016).

    Article  ADS  CAS  PubMed  Google Scholar 

  8. Kim, D. C. et al. Material-based approaches for the fabrication of stretchable electronics. Adv. Mater. 32, 1902743 (2020).

    Article  CAS  Google Scholar 

  9. Boutry, C. M. et al. A hierarchically patterned, bioinspired e-skin able to detect the direction of applied pressure for robotics. Sci. Robot. 3, eaau6914 (2018).

    Article  PubMed  Google Scholar 

  10. Shih, B. et al. Electronic skins and machine learning for intelligent soft robots. Sci. Robot. 5, eaaz9239 (2020).

    Article  PubMed  Google Scholar 

  11. Huang, Z. et al. Three-dimensional integrated stretchable electronics. Nat. Electron. 1, 473–480 (2018).

    Article  Google Scholar 

  12. Valentine, A. D. et al. Hybrid 3D printing of soft electronics. Adv. Mater. 29, 1703817 (2017).

    Article  Google Scholar 

  13. Graudejus, O. et al. Encapsulating elastically stretchable neural interfaces: yield, resolution, and recording/stimulation of neural activity. Adv. Funct. Mater. 22, 640–651 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Li, K. et al. A generic soft encapsulation strategy for stretchable electronics. Adv. Funct. Mater. 29, 1806630 (2019).

    Article  Google Scholar 

  15. Lu, N., Yoon, J. & Suo, Z. Delamination of stiff islands patterned on stretchable substrates. Int. J. Mater. Res. 98, 717–722 (2007).

    Article  CAS  Google Scholar 

  16. Niu, S. et al. A wireless body area sensor network based on stretchable passive tags. Nat. Electron. 2, 361–368 (2019).

    Article  Google Scholar 

  17. Graz, I. M. et al. Silicone substrate with in situ strain relief for stretchable thin-film transistors. Appl. Phys. Lett. 98, 124101 (2011).

    Article  ADS  Google Scholar 

  18. Fan, J. A. et al. Fractal design concepts for stretchable electronics. Nat. Commun. 5, 3266 (2014).

    Article  ADS  PubMed  Google Scholar 

  19. Miyamoto, A. et al. Inflammation-free, gas-permeable, lightweight, stretchable on-skin electronics with nanomeshes. Nat. Nanotechnol. 12, 907–913 (2017).

    Article  ADS  CAS  PubMed  Google Scholar 

  20. Sun, B. et al. Gas-permeable, multifunctional on-skin electronics based on laser-induced porous graphene and sugar-templated elastomer sponges. Adv. Mater. 30, 1804327 (2018).

    Article  ADS  Google Scholar 

  21. Gerratt, A. P. et al. Elastomeric electronic skin for prosthetic tactile sensation. Adv. Funct. Mater. 25, 2287–2295 (2015).

    Article  CAS  Google Scholar 

  22. Minev, I. R. et al. Electronic dura mater for long-term multimodal neural interfaces. Science 347, 159 (2015).

    Article  ADS  CAS  PubMed  Google Scholar 

  23. Nair, V. et al. Laser writing of nitrogen-doped silicon carbide for biological modulation. Sci. Adv. 6, eaaz2743 (2020).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  24. Wang, W. et al. Strain-insensitive intrinsically stretchable transistors and circuits. Nat. Electron. 4, 143–150 (2021).

    Article  CAS  Google Scholar 

  25. Wang, S. et al. Skin electronics from scalable fabrication of an intrinsically stretchable transistor array. Nature 555, 83–88 (2018).

    Article  ADS  CAS  PubMed  Google Scholar 

  26. Yoon, J. et al. Design and fabrication of novel stretchable device arrays on a deformable polymer substrate with embedded liquid-metal interconnections. Adv. Mater. 26, 6580–6586 (2014).

    Article  CAS  PubMed  Google Scholar 

  27. Kang, J. et al. Modular and reconfigurable stretchable electronic systems. Adv. Mater. Technol. 4, 1800417 (2019).

    Article  Google Scholar 

  28. Tutika, R. et al. Self-healing liquid metal composite for reconfigurable and recyclable soft electronics. Commun. Mater. 2, 64 (2021).

    Article  CAS  Google Scholar 

  29. Dai, X. et al. Self-Healing, flexible, and tailorable triboelectric nanogenerators for self-powered sensors based on thermal effect of infrared radiation. Adv. Funct. Mater. 30, 1910723 (2020).

    Article  MathSciNet  CAS  Google Scholar 

  30. Kadumudi, F. B. et al. The manufacture of unbreakable bionics via multifunctional and self-healing silk-graphene hydrogels. Adv. Mater. 33, 2100047 (2021).

    Article  CAS  Google Scholar 

  31. Son, D. et al. An integrated self-healable electronic skin system fabricated via dynamic reconstruction of a nanostructured conducting network. Nat. Nanotechnol. 13, 1057–1065 (2018).

    Article  ADS  CAS  PubMed  Google Scholar 

  32. Hwang, H. et al. Stretchable anisotropic conductive film (S-ACF) for electrical interfacing in high-resolution stretchable circuits. Sci. Adv. 7, eabh0171 (2021).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  33. Chen, X. et al. Instant tough bioadhesive with triggerable benign detachment. Proc. Natl Acad. Sci. USA 117, 15497 (2020). 27.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  34. Yang, X. et al. Ultrathin, stretchable, and breathable epidermal electronics based on a facile bubble blowing method. Adv. Electron. Mater. 6, 2000306 (2020).

    Article  CAS  Google Scholar 

  35. Wang, T. et al. Mechanically durable memristor arrays based on a discrete structure design. Adv. Mater. 34, 2106212 (2022).

    Article  CAS  Google Scholar 

  36. Yan, X. et al. Quadruple H-bonding cross-linked supramolecular polymeric materials as substrates for stretchable, antitearing, and self-healable thin film electrodes. J. Am. Chem. Soc. 140, 5280–5289 (2018).

    Article  CAS  PubMed  Google Scholar 

  37. Jeong, J.-W. et al. Materials and optimized designs for human-machine interfaces via epidermal electronics. Adv. Mater. 25, 6839–6846 (2013).

    Article  CAS  PubMed  Google Scholar 

  38. Webb, R. C. et al. Ultrathin conformal devices for precise and continuous thermal characterization of human skin. Nat. Mater. 12, 938–944 (2013).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  39. Park, S. et al. Self-powered ultra-flexible electronics via nano-grating-patterned organic photovoltaics. Nature 561, 516–521 (2018).

    Article  ADS  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

G.L. and Zhiyuan Liu acknowledge support from the National Key Research and Development Program of China (grant no. 2021YFF0501601) and the National Natural Science Foundation of China (grant no. 81927804 and U1913601) and NSFC-Shenzhen Robotics Basic Research Center Program (grant no. U2013207). This project is supported by the National Research Foundation, Singapore (NRF) under NRF’s Medium Sized Centre: Singapore Hybrid-Integrated Next-Generation μ-Electronics (SHINE) Centre funding programme. Any opinions, findings and conclusions or recommendations expressed in this material are those of the author(s) and do not reflect the views of NRF. We thank A. L. Chun for critically reading and editing the manuscript. We thank W. Sun and F. Zhao from Bruker Nano Surface Division for AFM characterization.

Author information

Authors and Affiliations

Authors

Contributions

Y.J. and Zhiyuan Liu conceived and coordinated the project and experiments. S.J. assisted the concept development, sample preparation and manuscript writing. J.S. prepared EMG electrodes and collected data in the EMG experiment. G.Z. and H.G. conducted the molecular dynamics in biphasic network growth model. T.S. conducted the Auger characterization and analysis. C.W. assisted the finite element analysis. W.L., H.J. and Zhihua Liu assisted flexible PCB design. M.Y. and Z. Lu assisted EMG electrode design. H.Z. assisted the EMG experiment. G.L. directed, and J.H., J.S. and Y.L. conducted the in vivo animal experiments except bladder. W.Y.X.P. and S.-C.Y. performed the in vivo bladder experiments. J.X., S.W., T.L. and X.Y. assisted BIND interface fabrication. Y.J. prepared all other samples, conducted all other experiments and wrote the manuscript. Z.B. and X.C. directed the project. All authors read and revised the manuscript.

Corresponding authors

Correspondence to Zhiyuan Liu, Zhenan Bao or Xiaodong Chen.

Ethics declarations

Competing interests

Y.J., Zhiyuan Liu and X.C. are inventors of the international PCT patent (no. PCT/SG2022/050607, priority date 26 August 2021, filing date 25 August 2022, pending) filed by Nanyang Technological University, which covers the BIND interface and BIND connection reported in this paper.

Peer review

Peer review information

Nature thanks Jadranka Travas-Sejdic, Yihui Zhang and Rebecca Kramer-Bottiglio for their contribution to the peer review of this work.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data figures and tables

Extended Data Fig. 1 Electro-mechanical performance of soft modules connected by conventional pastes or BIND interfaces.

a-d, Schematic (a) and photographs (b-d) of two soft PDMS/Au modules connected using commercial ACF. Stress concentration at connection region results in electrical failure at 50% strain (b, c) and mechanical failure at 76% strain (d). e-h, Schematic (e) and photographs (f-h) of two BIND interfaces connected without any pastes display conductivity up to 180% strain (f, g) and remain connected even at > 600% strain (h). i, Unlike connections using commercial pastes, BIND connection remained conductive (as shown by the low resistance change) even at 180% strain (left). Magnified graph of the low strain region (0–50%) on the right shows BIND connection experienced < 4 times change in relative resistance at 50% strain.

Extended Data Fig. 2 Electro-mechanical performance of BIND interface depends on evaporation rate and thickness.

a, b, Graph (a) and contour map (b) show the electrical stretchability of a single BIND interface as a function of evaporation rate and thickness. At the lowest evaporation rate (0.1–0.2 Å/s) and thickness (45 nm), a non-conductive interface is obtained. As evaporation rate or thickness increases, electrical stretchability peaks before decreasing to ~40%. c, d, Graph (c) and contour map (d) show the electrical stretchability of soft-soft BIND connection as a function of evaporation rate and thickness is similar to the single BIND interface in a and b. Only a limited combination of evaporation rate (0.5–1.0 Å/s) and thickness (45–60 nm) forms the BIND connection. Lower rates and thicknesses result in non-conductive connection while higher ones form non-adhesive interfaces. e, f, Graph (e) and contour map (f) show the mechanical stretchability of soft-soft BIND connection decreasing monotonically with increasing evaporation rate and thickness. Error bars in a, c, e are s.d. from 3–5 samples.

Extended Data Fig. 3 Soft-soft BIND connection is easy to use and has anti-tearing properties.

a-c, Pressing time (a), applied pressure (b) and peeling direction (c) do not influence the adhesion strength of BIND connection very much, indicating that BIND interfaces are easy to use. Simply pressing for 1 s at a low pressure of 0.001 MPa (i.e., a normal finger press) is enough to connect BIND interfaces together. Consistent with the study in Extended Data Fig. 2, non-adhesive connections (120 nm thick Au layer) cannot adhere regardless of pressure. In all other characterization, lap shear test was employed because it’s more common in practical application. d-g, Unlike ACF connection which breaks easily, BIND connection is resistant to cuts. Lap shear test on ACF and BIND connections that were initially cut (~1 mm) in the middle shows BIND connection withstood tearing forces >1.4 N (d) and remained stretchable even at > 500% strain (e). Photos show ACF connection with an initial cut broke easily at the cut location at 50% strain (f), while the initial cut in a BIND connection could still be stretched at 500% strain (g). Error bars in a-c are s.d. from 3-4 samples. Error bars in e are s.d. from 3 samples.

Extended Data Fig. 4 Advantages of BIND connections.

a, b, Schematic (a) and photo (b) showing two patterned BIND interfaces (on module A and B) are joined face-to-face via BIND connection. Inset in b: The BIND connection is electrically conductive to light an LED. c, BIND connections expanded to other conductive materials such as silver/silver and silver/gold connections also display robust electrical and mechanical stretchability. d, Photo of a soft module connected to a flexible PI module patterned with 6 electrode channels, via soft-rigid BIND connection. e, Soft-rigid BIND connection involving rigid or flexible substrates like PI, PET, glass, and metal show higher electrical (~200%) and mechanical (~800%) stretchability than conventional connection via various commercial pastes. f, Interfacial toughness of SEBS encapsulation layer on a BIND interface (0.24 N/mm) is much larger than various other types of encapsulation layer bonded on a conventional PDMS/Au interface. All encapsulation layers are ~100 μm thick. g, SEM image of BIND encapsulation (~300 nm) on a pair of electrodes, exposing two pads for signal collection (SEM shown in magnified view). Scale bar in g: 100 µm. h-j, Resolution of single BIND interface, BIND connection, and BIND encapsulation can achieve 100 µm. The electrical stretchability of both BIND interface (h) and BIND connection (i) decreases with reduced line width, while the mechanical stretchability was kept relatively stable. The overall width was kept as 5 mm. Inset: SEM image of BIND encapsulation with exposed area of 100 × 100 µm2. Scale bar: 50 µm. Error bars are s.d. from 3-4 samples.

Extended Data Fig. 5 Surface AFM mapping shows decreasing ratio of exposed polymer to metal phases, on non-conductive, BIND and non-adhesive interfaces.

a-c, AFM modulus mapping show the polymer/metal ratio on the surface decreases from non-conductive interface (a), to BIND interface (b), to non-adhesive interface (c), which is consistent with the adhesion mapping (Fig. 2a). Due to the large modulus difference, polymer and metal phase can be easily distinguished and are labeled in blue and yellow color, respectively. d-f, AFM height mapping shows similar height variation on non-conductive interface (d) and BIND interface (e), because their height variation comes from the half-immersed gold nanoparticles. For non-adhesive interface (f), the height variation comes from the stacking gold nanoparticles on top of the polymer, so the variation range is slightly different from the non-conductive and BIND interfaces. Scale bar: 100 nm.

Extended Data Fig. 6 Cross-sectional AFM mapping reveals decreasing penetrating Au nanoparticles inside non-conductive, BIND and non-adhesive interfaces.

a, b, AFM height (a) and adhesion (b) mapping of a BIND connection show polymer and metal phases have an interpenetrating structure. Scale bar: 100 nm. c-k, AFM height, adhesion, and current mapping show that the non-conductive connection (c-e) has more interpenetrating Au nanoparticles than the BIND connection (f-h), while there is nearly no interpenetrating Au inside non-adhesive interface (i-k). Non-conductive connection and BIND connection were formed by pressing two interfaces together, while non-adhesive connection was formed by adhering epoxy to non-adhesive interface. White dash line in i shows the epoxy boundary, beyond which the epoxy region was removed for clearer view in j-k. Scale bar in c-k: 400 nm.

Extended Data Fig. 7 In vivo neuromodulation BIND device is compatible with tissues of different sizes, with mechanical robustness against touching and pulling.

a-e, BIND electrode conformally wraps around the common peroneal nerve (a), sciatic nerve (b) and peroneus longus muscle (c) without suturing. Conformal contact is also achieved when placed on the cerebral cortex (d), or sutured onto a bladder wall (e). The bladder experiment can be further improved in the future, to wrap the BIND electrode around the bladder to avoid suture. f-i, Mechanical interference such as touching (f), cathode pulling (g) and anode pulling (h) during sciatic nerve stimulation does not affect the stimulation performance very much. Here, the ultrathin part of the BIND electrode is wrapped around a sciatic nerve while the thick wiring part relays signals via a BIND connection. Stimulated effectiveness was measured by the moving distance of rat ankles upon applied interference (i). Error bars are s.d. from 7–10 movements. j-k, 3D map shows the impedance of BIND interface is typical of a metal-based stretchable electrode (j). Within the frequency range of interest (e.g., ECoG 10-50 Hz, EMG 10-500 Hz), impedance remained low and nearly unchanged up to ~70% strain (k). Interfacial impedance of ultrathin electrode can be improved in the future by various methods (e.g., coating the interface with low impedance materials such as PEDOT:PSS, or increasing the surface area). Because BIND connections construct electrical pathway by Ohmic contact between Au nanoparticles, its influence on impedance is the same as its electrical resistances in DC tests.

Extended Data Fig. 8 The in vivo neuromodulation BIND electrode has various application, including subcutaneous CMAP, cortex ECoG, and bladder, demonstrating its usage universality.

a-f, For simultaneous nerve stimulation and CMAP recording, one BIND electrode is wrapped around the common peroneal nerve and another is wrapped around the peroneus longus muscle (a). Stimulation current was applied to the common peroneal nerve with a parallel resistance of 250 Ω. (b). P-P voltage of CMAP is triggered when stimulation current reaches a threshold, then increases and plateaus (c). Spectroscopy of recorded CMAP shows obvious power density in the typical CMAP frequency range (10–500 Hz) (d). Ankle movement was measured by the distance, d, of rat ankle before (e) and during (f) stimulation. g, h, 4-channel BIND electrode on cerebral cortex for ECoG recording, consisting of ultrathin electrode module, and thick wiring module (g). ECoG signals of healthy and epilepsy rats differ in both amplitude and frequency (h). i-k, For bladder stimulation, two BIND electrodes were sutured onto the bladder wall to transmit electrical pulses, where the bladder contraction is recorded by external pressure sensor (i, j). For bladder recording, one BIND electrode is wrapped around the bladder. Normal saline is injected into the bladder through the catheter, and the corresponding expansion is detected by BIND electrode (k).

Extended Data Fig. 9 The 21-channel EMG electrode assembled via BIND connection shows high signal fidelity and resistance against mechanical interferences, including connection pressing and stretching (50%).

a-b, EMG signals collected from ACF-connected control electrode (a) and BIND electrode (b) show that, pressing/releasing the connection between soft wiring and customized PCB induced noisy signals in control electrode, especially with actions. EMG signals from BIND electrode were less noisy and clearly distinguishable during and after pressure, indicating its resistance to pressure. c-g, Besides pressing, BIND connection also exhibits resistance to 50% stretching interference. One side of the BIND connection between soft wiring and customized PCB was fixed, and the other side was manually stretched to the 50% mark (c). Both electrodes showed noisy signals when stretched but the larger SNR in BIND electrode indicates that it can detect signals under strain (d, e). Releasing strain induced larger noise in control electrode but both electrodes recovered their ability to detect EMG signals after strain is released (f, g). All signals were subjected to a 50 Hz notch filter and a 30 Hz high-pass filter. Error bars in d, e are s.d. from 21 channels.

Extended Data Fig. 10 21-channel EMG electrode assembled via BIND connections maps various hand and finger gestures in air and underwater, with mechanical resistance to interference.

a, BIND electrode detects EMG signals of hand gestures (clench, open, raise and bend), finger movements (stretching of individual fingers) and maximum voluntary contraction (MVC, measured by grip dynamometer) in air. b-d, Even underwater, BIND electrode detected EMG signals with high SNR from different hand and finger gestures (b) and tolerated pressure (c) and 50% strain (d) at connection points. High quality signals with little noise are seen before, during and after pressure or 50% strain. Here, the entire BIND electrode, including two BIND connections, was immersed in water. Signals in a, b were subjected to a 50 Hz notch filter and a 10 Hz high pass filter. Signals in c, d were subjected to a 50 Hz notch filter and a 30 Hz high pass filter.

Supplementary information

Supplementary Information

This file contains Supplementary Notes 1–4, Figs. 1–12, legends for Videos 1–8 and References.

Supplementary Video 1

Robust wiring against mechanical interference. See main Supplementary Information PDF for a full description.

Supplementary Video 2

Simultaneous in vivo CMAP stimulation and recording. See main Supplementary Information PDF for a full description.

Supplementary Video 3

ECoG recording in healthy and epileptic rats. See main Supplementary Information PDF for a full description.

Supplementary Video 4

Plug-and-play assembly of the 21-channel EMG electrode array. See main Supplementary Information PDF for a full description.

Supplementary Video 5

Mechanical interference in 21-channel on-skin EMG detection. See main Supplementary Information PDF for a full description.

Supplementary Video 6

Gesture detection through a 21-channel BIND electrode (in air). See main Supplementary Information PDF for a full description.

Supplementary Video 7

Underwater performance of a 21-channel EMG BIND electrode. See main Supplementary Information PDF for a full description.

Supplementary Video 8

Plug-and-play assembly of a customized multifunctional circuit. See main Supplementary Information PDF for a full description.

Source data

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Jiang, Y., Ji, S., Sun, J. et al. A universal interface for plug-and-play assembly of stretchable devices. Nature 614, 456–462 (2023). https://doi.org/10.1038/s41586-022-05579-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-022-05579-z

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing