1 Introduction

Artificial muscles, which can transfer other types of energies (such as thermal and electrothermal energy) to mechanical energy, have received extensive attention in multiple fields ranging from material science, mechanical and medical engineering as well as robotics [1]. Ideal artificial muscles should have a performance like high energy conversion efficiency, fast response, scalability, non-hysteretic, long-cycle life and low cost, and so on. Various types of artificial muscles such as pneumatic actuators [19], coating nylon fiber with conductive silver paint [20], and by twisting metal wire and fishing line simultaneously [

2 Experimental Sections

2.1 Fabrication of Self-Coiled Nylon Muscles and Copper Wire-Wrapped Coiled Nylon Muscles

The monofilament nylon fiber was purchased from Statex Productions & Vertriebs GmbH (Shieldex PN# 260151011717 and 260151023534). Self-coiled nylon muscles were made by inserting twist into a nylon fiber. As shown in Fig. S1a, a nylon fiber was hung on a stepper motor, and a constant load was applied to the nylon fiber during the twisting process. To prevent the untwist of nylon yarn during actuation, it was torsionally tethered via a wood stick. By over-twisting, a fully coiled nylon muscle was made. A copper wire-wrapped coiled nylon fiber was made by wrap** a 50-μm-diameter copper wire with the coiled nylon fiber. To avoid the stroke restriction by copper wire, the wrap** direction was opposite to the coil direction (Fig. 2a). This thin copper wire has a negligible mechanical effect on the actuation.

Unless otherwise indicated, the tensile stroke is defined as the length change during actuation divided by the length of the loaded muscle. The applied tensile stress is the constant force normalized to the yarn diameter of the fully twisted yarn. The generated stress was calculated using the generated force divided by the cross-sectional area of the coiled muscle. Spring index C is defined in spring mechanics as C = D/d, where d is the fiber diameter and D is the nominal coil diameter as measured by the fiber centerline.

2.2 Characterizations

The morphology and structure of the coiled nylon yarn and the copper-wire-wrapped nylon yarn were obtained by scanning electron microscopy (SEM, model JSM-7800F, JEOL, Japan). The change in modulus and generated stress for the coiled nylon muscle with temperature was measured by thermomechanical analysis (TMA, TMA 7100, Hitachi, Japan). The change in the pitch of coiled nylon was observed by optical microscopy (DMi8, Leica, Germany). The temperature change of the coiled nylon muscle was recorded using a thermographic high-resolution system (T540, FLIR, United States). The mechanical properties of the nylon fiber were tested by a universal mechanical testing machine (5944, Instron, America).

3 Results and Discussion

3.1 Thermally Powered Self-Coiled Nylon Muscles

For thermally powered Twisted and Coiled Polymer (TCP) artificial muscles, the actuation mechanism is depicted in Fig. 1a. It illustrates that thermal expansion causes the yarn to untwist, which leads to yarn length contraction. Isobaric tensile actuation for thermally powered nylon 6,6 muscle was measured by thermal dynamic analysis (TMA). To achieve reversible actuation, the coiled nylon muscle has been thermally annealed at 150 °C for 0.5 h before actuation, and the working temperature was set from 20 to 180 °C at a rate of 10 °C min−1. As a result, the tensile stroke as a function of temperature for the coiled nylon muscle was obtained (Fig. 1b). Yarn muscle monotonically contracts with increasing temperature for most applied mechanical loads. Notably, if the applied load is smaller than the load used for coiling, adjacent coils are in contact, limiting contraction during actuation. Thus, for the applied load of 5.4 MPa, yarn muscle first contracts with increasing temperature until a peak value of 14.5% was achieved, and keeps no change when continue increasing temperature. Increasing load can separate adjacent coils and achieve larger actuation. Figure 1c shows the load dependence of tensile stroke for the coiled nylon muscles. The tensile stroke increases with increasing applied tensile stress until a maximum value of 20.1% was achieved under an optimal load of 10.7 MPa. With increasing applied load, the muscle’s initial length was stretched, leading to a decreased tensile stroke. The work capacity that nylon muscle generated during contraction monotonically increases with increasing applied tensile stress and reaches 1.3 J g−1 when the applied load was 31 MPa, which is 32 times that of mammal muscle [23].

Fig. 1
figure 1

Dual-mode actuation performance for a self-coiled nylon muscle. a Optical images of a coiled nylon muscle at un-actuated and actuated states. b The temperature dependence of tensile stroke for a coiled, 69-μm-diameter nylon muscle under different applied mechanical loads. The spring index, twist density, and twisting load are 1.3, 37 turns cm−1, and 26.7 MPa, respectively. c Tensile stroke and work capacity as a function of applied tensile stress for the muscle used in (b). d The generated stress as a function of temperature under a pre-strain of 10% for the muscle used in (b). e The generated stress and initial stress (σI-min) as a function of applied pre-strain for the muscle used in (d)

From the aspect of natural muscle, the output performance is not only limited to the muscle length contraction, but also means that a force can be generated in isometric conditions without muscle length change [24]. When a nylon artificial muscle was thermally heated, the yarn muscle first absorbs thermal energy and then transmits the energy to the coiled structure. Consequently, an isometric force was generated. Isometric contraction is also an important parameter of a high-performance artificial muscle.

To mimic the isometric contraction, a similar setup was employed (as illustrated in fig. S1c) where the sample length is fixed, and stress evolution could be recorded by the load cell when external stimulation was applied. In each experiment, the specimen was subjected to at least three consecutive thermal cycles where stress can be significantly produced in heating and elapse in cooling. By applying a pre-strain of 10%, nylon muscle produced initial stress (σI-min), and the stress increases with increasing temperature until a maximum value (σI-max) was achieved (Fig. 1d). With an increase in pre-strains, both the maximum and minimum isometric stress increase accordingly (fig. S2), from 4.1 to 10.1 MPa of σI-min. The generated stress was calculated by the difference between σI-max and σI-min. For each pre-strain, σI-min and σI-max stay at the same level despite the number of cycles. With an increase in strains, the value and variation mode of stress is observed to strongly depend on the pre-strains, reflecting the effect of crystallization behavior and thermodynamic properties on contraction performance [25]. The stress generation as a function of the applied pre-strain was shown in Fig. 1e. It is worth noting that nylon muscle can produce more than 28 MPa stress under 25% pre-strain, which is more than 3.5 times than for CNT-NCY hybrid muscle [26] and 100 times than for human muscle [23] (∼0.3 MPa).

3.2 Electrothermally Powered Copper Wire-Wrapped Coiled Nylon Muscles

3.2.1 Isobaric Actuation

Although self-coiled nylon muscles can achieve excellent actuation performance, they cannot be powered by electricity due to their non-conductivity. By wrap** coiled nylon fiber with a copper wire, we made an electrothermally powered nylon muscle. To avoid the stroke restriction by copper wire, the wrap** direction was opposite to the coil direction (Fig. 2a). Also, the winding density (the number of winding turns divided by the total muscle length) of copper wire would affect the uniformity of thermal energy on the yarn muscle that delivered by electrical power. Figure 2b shows the tensile stroke as a function of time under different applied loads. During a heating–cooling cycle, muscle contracts to a maximum value within 4.5 s, and provide a large stroke rate of 2.2% s−1. The load dependence of the tensile stroke and work capacity for electrothermally powered nylon muscle has also been obtained (Fig. 2c). By applying a 0.28 W cm−1 square-wave voltage, nylon muscle can generate a maximum tensile stroke of 12.1% and a work capacity of 0.8 J g−1, respectively. After the wrap** process, we measured the real-time temperature using a thermal camera during the nylon muscle contraction, which is shown in Fig. 2d. It depicts that the nylon yarn muscle was uniformly heated by electrical power. In addition, the real-time temperature of yarn muscle exhibits the same trend with tensile stroke (Fig. 2e).

Fig. 2
figure 2

Isobaric actuation for an electrothermally powered nylon muscle. a Schematic illustration of the fabrication method for a coiled, copper-wire-wrapped nylon muscle. b Tensile stroke as a function of time for a coiled, copper wire-wrapped, 130-μm-diameter nylon muscle under different applied tensile stresses. c Tensile stroke and work capacity as a function of applied tensile stress for the muscle used in (b). Inset shows the SEM image for the copper wire-wrapped nylon muscle. The scale bar is 200 μm. d Thermal camera images show the temperature change for the coiled nylon muscle when driven by a 0.05 Hz, 0.23 W cm−1 square-ware voltage with a 50% duty cycle. e The time dependence of isobaric tensile stroke (left y-axis) at applied tensile stress of 10.5 MPa and the real-time temperature (right y-axis) for the electrothermally driven nylon muscle used in (d)

The actuation performance for electrothermally powered nylon muscle can be improved by varying actuation frequency, duty cycle, and input power. Figure 3a demonstrates the tensile stroke changes with time under different applied frequencies. The applied tensile stress and power were 8.6 MPa and 0.27 W cm−1 square wave, respectively. At a low frequency of 0.02 Hz, nylon muscle generates a maximum tensile stroke of 11.6%. While at a high frequency of 0.5 Hz, nylon muscle contracts insufficiently in the first cycle and quickly increased to the stabilized stroke after a few cycles. Due to the limited time to cool at high frequencies, nylon muscle cannot release back to its original length and only actuated at high-temperature regions. Consequently, muscle stroke decreased with the increase in frequency (Fig. 3b). Besides, the yarn’s diameter also affects the actuation for different frequencies, especially for high frequencies. As shown in Fig. S6b, the cooling time (defined as the time needed to reduce the tensile stroke by 90%. [3]) for an 80-μm-diameter nylon muscle (1.9 s) is much shorter than for a 200-μm-diameter nylon muscle (7.2 s), indicating that muscles with a small diameter can be cycled at a higher frequency than for a larger one (Fig. S6b).

Fig. 3
figure 3

Impact factors of isobaric actuation for an electrothermally powered nylon muscle. a Time dependence of tensile stroke for a coiled, copper-wrapped, 130-μm-diameter nylon muscle under different applied frequencies. The applied voltage and isobaric tensile stress were 0.27 W cm−1, 8.6 MPa respectively. b Tensile stroke as a function of applied square wave frequency for the muscle used in (a). c Tensile stroke versus time under different input power densities with the same frequency of 0.05 Hz for the muscle used in (a). The applied tensile stress was 8.6 MPa. d Tensile stroke as a function of input power at 0.05 Hz, 50% duty cycle square wave voltage. e Tensile stroke versus cycle for the muscle in (a) when driven electrothermally at 0.2 Hz, 50% duty cycle, 0.21 W cm−1 square wave voltage under a 7.5-MPa load. Inset shows the stroke change of the first 10 cycles and the last 10 cycles for the coiled nylon muscle

Note that the contractile rate (tensile stroke divided by the contractile time) and release rate (tensile stroke divided by the release time) are not always matched. Thus, the duty cycle would also affect the actuation performance. Fig. S3a shows the duty cycle dependence of the tensile stroke for a coiled nylon muscle measured at a 0.05 Hz, 0.28 W cm−1 square-wave power. At a low duty cycle of 10%, the heating time is too short, which limited the input electrothermal energy, leading to a small contractile stroke. At a high duty cycle of 90%, the cooling time is too short, which made yarn muscle fail to return to its original state, causing the decrease of absolute tensile stroke. Consequently, an appropriate duty cycle is necessary to achieve the largest tensile actuation. After varying different duty cycles, we found that the maximum tensile stroke is 12% at the optimal duty cycle of 50%. Thus, the tensile stroke increased first with the increment of the duty cycle and then decreased after it reached the maximum value, as shown in Fig. S3b.

More importantly, the input power and cycle stability of muscle are significant for practice use [27]. We have measured the dependence of tensile stroke on the input power density, which is illustrated in Fig. 3c, d. The tensile stroke increases almost linearly and reversibly with the input power density. Typically, the tensile stroke increases from 4.1% at an input power of 0.16 W cm−1 to 12.1% at an input power of 0.27 W cm−1. The cycle stability of electrothermally powered copper wire-wrapped nylon muscle has also been measured. As shown in Fig. 3e, the nylon muscle provides a very stable contraction stroke for at least 1000 cycles when a 0.2 Hz, 50% duty cycle, 0.25 W cm−1 square-wave power was applied. Thus, the electrothermally powered nylon muscle are feasible material for the applications of soft robotics and exosuits. In addition, we have demonstrated the potential application of a copper-wire wrapped nylon artificial muscle lifting heavy weight when electrothermally driven at a 0.05 Hz, 50% duty cycle, and 6.5 V square-wave voltage (shown in Movie S1).

3.2.2 Isometric Actuation

Similar to the self-coiled nylon muscle, the generated isometric stress of a coiled, copper wire-wrapped nylon muscle has been investigated. The muscle was fixed in a tensile dynamometer with two ends clamped and connected by an electrical wire to a power supply (Fig. S1b). Fig. S4a shows the time dependence of the generated stress of the coiled, copper wire-wrapped nylon muscle under different applied preload strains. The input electrical power was 0.05 Hz, 0.27 W cm−1 square-wave with 50% duty cycle. When the voltage was applied to the nylon muscle, the generated isometric stress increased rapidly and then sluggishly increase to the maximum value. When the applied voltage was switched to 0 V, the generated stress gradually decreased to the initial stress. Figure 4a shows that the generated isometric stress first increased with the increase in the pre-strain and then reached a maximum value of 28.4 MPa at the pre-strain of 25%.

Fig. 4
figure 4

Impact factors of isometric actuation for an electrothermally powered nylon muscle. a Stress generation versus applied pre-strains for a coiled, copper wire-wrapped, 130-μm-diameter nylon muscle. b The generated stress as a function of applied frequency for the muscle used in (a). c Stress as a function of input power at 0.05 Hz, 50% duty cycle square-wave voltage. d The generated stress as a function of pre-strain for the muscles with different spring indexes of 1.12, 1.23, and 1.35. e The cycle stability of generated stress for the muscle used in (a) under 10% pre-strain when driven by a 0.1 Hz, 0.3 W cm−1 square wave voltage. f Comparison of tensile stroke and the generated isometric stress with mammalian skeletal muscles (20%, 0.75 MPa) [29], SWNT sheet (33%, ~ 0.75 MPa) [30], CNT-NCY yarn muscle (9%, 12 MPa) [26], CNT composite yarn muscle (12%, 4 MPa) [27], Wool yarn (39%, ~ 2.6 MPa) [31], Flax yarn (17%, ~ 8.5 MPa) [31], Catton yarn (18%, ~ 9 MPa) [31] and poly(ε-caprolactone)-based polyurethane (PCL) muscle (29%, 2.08 MPa) [

Data Availability

The data used to support the results of this study are available from the corresponding author upon reasonable request.

References

  1. Mirvakili, S. M., & Hunter, I. W. (2018). Artificial muscles: Mechanisms, applications, and challenges. Advanced Materials, 30, 1704407.

    Article  Google Scholar 

  2. **ao, W., Hu, D., Chen, W. X., Yang, G., & Han, X. (2021). Design, characterization, and optimization of multi-directional bending pneumatic artificial muscles. Journal of Bionic Engineering, 18, 1358–1368.

    Article  Google Scholar 

  3. Hu, X. H., Jia, J. J., Wang, Y. M., Tang, X. T., Fang, S. L., Wang, Y. L., Baughman, R. H., & Ding, J. N. (2022). Fast large-stroke sheath-driven electrothermal artificial muscles with high power densities. Advanced Functional Materials, 32, 2200591.

    Article  Google Scholar 

  4. Dai, S. P., Zhou, X. S., Hu, X. H., Dong, X., Jiang, Y. Y., Cheng, G. G., Yuan, N. Y., & Ding, J. N. (2021). Carbon nanotube hybrid yarn with mechanically strong healable silicone elastomers for artificial muscle. ACS Applied Nano Materials, 4, 5123–5130.

    Article  Google Scholar 

  5. Chu, H. T., Hu, X. H., Wang, Z., Mu, J. K., Li, N., Zhou, X. S., Fang, S. L., Haines, C. S., Park, J. W., Qin, S., Yuan, N. Y., Xu, J., Tawfick, S., Kim, H. J., Conlin, P., Cho, M., Cho, K., Oh, J. Y., Nielsen, S., … Baughman, R. H. (2021). Unipolar stroke, electroosmotic pump carbon nanotube yarn muscles. Science, 371, 494–498.

    Article  Google Scholar 

  6. Ren, M., Qiao, J., Wang, Y. L., Wu, K. J., Dong, L. Z., Shen, X. F., Zhang, H. C., Yang, W., Wu, Y. L., Yong, Z. Z., Chen, W., Zhang, Y. Y., Di, J. T., & Li, Q. W. (2021). Strong and robust electrochemical artificial muscles by ionic-liquid-in-nanofiber-sheathed carbon nanotube yarns. Small (Weinheim an der Bergstrasse, Germany), 17, 2006181.

    Article  Google Scholar 

  7. Li, S. G., Vogt, D. M., Rus, D., & Wood, R. J. (2017). Fluid-driven origami-inspired artificial muscles. Proceedings of the National Academy of Sciences, 114, 13132–13137.

    Article  Google Scholar 

  8. Guan, Q. H., Sun, J., Liu, Y. J., Wereley, N. M., & Leng, J. S. (2020). Novel bending and helical extensile/contractile pneumatic artificial muscles inspired by elephant trunk. Soft Robotics, 7, 597–614.

    Article  Google Scholar 

  9. Mirvakili, S. M., Sim, D., Hunter, I. W., & Langer, R. (2020). Actuation of untethered pneumatic artificial muscles and soft robots using magnetically induced liquid-to-gas phase transitions. Science Robotics, 5, eaaz4239.

    Article  Google Scholar 

  10. Li, T. F., Li, G. R., Liang, Y. M., Cheng, T. Y., Dai, J., Yang, X. X., Liu, B. Y., Zeng, Z. D., Huang, Z. L., & Luo, Y. W. (2017). Fast-moving soft electronic fish. Science Advances, 3, e1602045.

    Article  Google Scholar 

  11. Kellaris, N., Gopaluni, V. V., Smith, G. M., Mitchell, S. K., & Keplinger, C. (2018). Peano-HASEL actuators: Muscle-mimetic, electrohydraulic transducers that linearly contract on activation. Science Robotics, 3, eaar3276.

    Article  Google Scholar 

  12. Duduta, M., Hajiesmaili, E., Zhao, H. C., Wood, R. J., & Clarke, D. R. (2019). Realizing the potential of dielectric elastomer artificial muscles. Proceedings of the National Academy of Sciences, 116, 2476–2481.

    Article  Google Scholar 

  13. Zhang, Y., Ellingford, C., Zhang, R. N., Roscow, J., Hopkins, M., Keogh, P., McNally, T., Bowen, C., & Wan, C. Y. (2019). Electrical and mechanical self-healing in high-performance dielectric elastomer actuator materials. Advanced Functional Materials, 29, 1808431.

    Article  Google Scholar 

  14. Thomas, S., Germano, P., Martinez, T., & Perriard, Y. (2021). An untethered mechanically-intelligent inchworm robot powered by a shape memory alloy oscillator. Sensors and Actuators A, 332, 113115.

    Article  Google Scholar 

  15. Haines, C. S., Lima, M. D., Li, N., Spinks, G. M., Foroughi, J., Madden, J. D., Kim, S. H., Fang, S. L., Jung de Andrade, M., Göktepe, F., Göktepe, Ö., Mirvakili, S. M., Naficy, S., Lepró, X., Oh, J. Y., Kozlov, M. E., Kim, S. J., Xu, X. R., Swedlove, B. J., … Baughman, R. H. (2014). Artificial muscles from fishing line and sewing thread. Science, 343, 868–872.

    Article  Google Scholar 

  16. Park, J., Yoo, J. W., Seo, H. W., Lee, Y., Suhr, J., Moon, H., Koo, J. C., Choi, H. R., Hunt, R., Kim, K. J., Kim, S. H., & Nam, J. D. (2017). Electrically controllable twisted-coiled artificial muscle actuators using surface-modified polyester fibers. Smart Materials and Structures, 26, 035048.

    Article  Google Scholar 

  17. Tang, X. T., Li, K., Liu, Y. X., Zhou, D., & Zhao, J. G. (2019). A general soft robot module driven by twisted and coiled actuators. Smart Materials and Structures, 28, 035019.

    Article  Google Scholar 

  18. Aziz, S., Naficy, S., Foroughi, J., Brown, H. R., & Spinks, G. M. (2017). Thermomechanical effects in the torsional actuation of twisted nylon 6 fiber. Journal of Applied Polymer Science, 134, 45529.

    Article  Google Scholar 

  19. Liu, Z. F., Fang, S. L., Moura, F., Ding, J. N., Jiang, N., Di, J., Zhang, M., Lepró, X., Galvao, D., Haines, C. S., Yuan, N. Y., Yin, S. G., Lee, D. W., Wang, R., Wang, H. Y., Lv, W., Dong, C., Zhang, R. C., Chen, M. J., … Baughman, R. H. (2015). Hierarchically buckled sheath-core fibers for superelastic electronics, sensors, and muscles. Science, 349, 400–404.

    Article  Google Scholar 

  20. Wu, L. J., Chauhan, I., & Tadesse, Y. (2018). A novel soft actuator for the musculoskeletal system. Advanced Materials Technologies, 3, 1700359.

    Article  Google Scholar 

  21. **ang, C. Q., Yang, H., Sun, Z. Y., Xue, B. C., Hao, L. N., Rahoman, M. A., & Davis, S. (2017). The design, hysteresis modeling and control of a novel SMA-fishing-line actuator. Smart Materials and Structures, 26, 037004.

    Article  Google Scholar 

  22. Helps, T., Taghavi, M., Wang, S. H., & Rossiter, J. (2020). Twisted rubber variable-stiffness artificial muscles. Soft Robotics, 7, 386–395.

    Article  Google Scholar 

  23. Oguntosin, V., & Akindele, A. (2019). Design and characterization of artificial muscles from wedge-like pneumatic soft modules. Sensors and Actuators A: Physical, 297, 111523.

    Article  Google Scholar 

  24. Zhu, S. S., Hu, J. L., & Zhang, Y. C. (2019). A single polymer artificial muscle having dual-mode contractibility, temperature sensibility, and trainability through enthalpy change. Advanced Materials Technologies, 4, 1900017.

    Article  Google Scholar 

  25. Bhatti, M. R. A., Bilotti, E., Zhang, H., Varghese, S., Verpaalen, R. C., Schenning, A. P., Bastiaansen, C. W., & Peijs, T. (2020). Ultra-high actuation stress polymer actuators as light-driven artificial muscles. ACS Applied Materials & Interfaces, 12, 33210–33218.

    Article  Google Scholar 

  26. Xu, L. L., Peng, Q. Y., Zhao, X., Li, P. Y., Xu, J. H., & He, X. D. (2020). A photoactuator based on stiffness-variable carbon nanotube nanocomposite yarn. ACS Applied Materials & Interfaces, 12, 40711–40718.

    Article  Google Scholar 

  27. Xu, L. L., Peng, Q. Y., Zhu, Y., Zhao, X., Yang, M. L., Wang, S. S., Xue, F. H., Yuan, Y., Lin, Z. S., Xu, F., Sun, X. X., Li, J. J., Yin, W. L., Li, Y. B., & He, X. D. (2019). Artificial muscle with reversible and controllable deformation based on stiffness-variable carbon nanotube spring-like nanocomposite yarn. Nanoscale, 11, 8124–8132.

    Article  Google Scholar 

  28. Fukushima, T., Asaka, K., Kosaka, A., & Aida, T. (2005). Fully plastic actuator through layer-by-layer casting with ionic-liquid-based bucky gel. Angewandte Chemie International Edition, 44, 2410–2413.

    Article  Google Scholar 

  29. Baughman, R. H. (2005). Playing nature’s game with artificial muscles. Science, 308, 63–65.

    Article  Google Scholar 

  30. Baughman, R. H., Cui, C., Zakhidov, A. A., Iqbal, Z., Barisci, J. N., Spinks, G. M., Wallace, G. G., Mazzoldi, A., De Rossi, D., Rinzler, A. G., Jaschinski, O., Roth, S., & Kertesz, M. (1999). Carbon nanotube actuators. Science, 284, 1340–1344.

    Article  Google Scholar 

  31. Yang, X. H., Wang, W. H., & Miao, M. H. (2018). Moisture-responsive natural fiber coil-structured artificial muscles. ACS Applied Materials & Interfaces, 10, 32256–32264.

    Article  Google Scholar 

  32. Huang, L. B., **e, X. X., Huang, H., Zhu, J., Yu, J. R., Wang, Y., & Hu, Z. M. (2020). Electrospun polyamide-6 nanofiber for hierarchically structured and multi-responsive actuator. Senors and Actuators A, 302, 111793.

    Article  Google Scholar 

Download references

Acknowledgements

Financial support from the program of the National Natural Science Foundation of China (Grant no. 52105057, 51905222), Natural Science Foundation of Jiangsu Province (Grant no. BK20200916), China Postdoctoral Science Foundation (no. 2021M691307, no. 2022T150274), Jiangsu Postdoctoral Research Foundation (no. 2021K543C), Key Research Project of Zhejiang Lab, and Senior Talent Foundation of Jiangsu University (Grant no. 5501110013) are acknowledged.

Author information

Authors and Affiliations

Authors

Corresponding authors

Correspondence to **nghao Hu or Jianning Ding.

Ethics declarations

Conflict of Interest

The authors declare no conflict of interest.

Additional information

Publisher's Note

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

Supplementary Information

Below is the link to the electronic supplementary material.

Supplementary file1 (DOCX 1789 KB)

Supplementary file2 (MP4 5082 KB)

Supplementary file3 (MOV 24045 KB)

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hu, X., Liu, R., Zhao, K. et al. Coiled Polymer Artificial Muscles Having Dual-Mode Actuation with Large Stress Generation. J Bionic Eng 20, 1626–1634 (2023). https://doi.org/10.1007/s42235-023-00353-x

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s42235-023-00353-x

Keywords