Microscopic in situ observation of electromechanical instability in a dielectric elastomer actuator utilizing transparent carbon nanotube electrodes

Zhen-Qiang Song*ad, Li-Min Wanga, Yongri Lianga, Xiao-Dong Wangb and Shijie Zhu*c
aCenter for Advanced Structural Materials, State Key Laboratory of Metastable Materials Science and Technology, Yanshan University, China. E-mail: zqsong@ysu.edu.cn
bSchool of Civil Engineering and Mechanics, Yanshan University, China
cDepartment of Intelligent Mechanical Engineering, Fukuoka Institute of Technology, Japan. E-mail: zhu@fit.ac.jp
dHebei Key Lab for Optimizing Metal Product Technology and Performance, Yanshan University, China

Received 17th May 2024 , Accepted 12th August 2024

First published on 13th August 2024


Abstract

Electromechanical instability (EMI) restricts the performance of dielectric elastomer actuators (DEAs), leading to premature electrical breakdown at a certain voltage. However, macro-level observations using traditional carbon grease electrodes have failed to capture the detailed features of EMI. In this study, we investigated EMI at the microscopic scale by fabricating transparent and conductive single-walled carbon nanotube (SWCNT) electrodes. Our findings reveal that EMI predominantly occurs in highly localized regions with dimensions on the order of tens of micrometers. This snap-through instability is likely induced by pre-existing defects within the elastomer, such as air voids or conductive particles, which reduce the critical voltage required for EMI in the flawed areas. From the perspective of phase transition principles, these defects act as heterogeneous nucleation sites for new phase embryos, thereby lowering the energy barrier for the electromechanical phase transition (i.e., EMI) compared to homogeneous nucleation in an ideally impurity-free elastomer. This study clarifies the longstanding discrepancy between theoretically predicted deformation bursts and the experimentally observed macroscopic continuous expansion of DEAs under low pre-stretch conditions. Additionally, it underscores the critical importance of material purity in mitigating electromechanical instability.


1. Introduction

As one type of electroactive polymer, dielectric elastomers (DE) exhibit impressive performance to generate large mechanical deformation under electrical stimuli.1–4 Consequently, it is extensively studied as a soft actuator for diverse applications.5–8 The operational principle involves a dielectric elastomer membrane sandwiched between the compliant electrodes on both sides. When a bias voltage is applied through the thickness of the membrane, opposite charges accumulate on the two electrodes, and the electrostatic attractive force causes membrane thinning in thickness and expansion in area, during which the electrical energy is transformed into mechanical motion.9,10 It is reported that the linear strain under electrical actuation is up to 100%,11 with energy density and response speed comparable to those of natural muscles, which inspired significant development of dielectric elastomer actuators (DEAs) as artificial muscles for soft robotics.3,9–14

Enhancing the actuation performance of dielectric elastomer actuators (DEAs) requires optimization of both the electrodes and the elastomer membranes. In particular, the dielectric permittivity and elastic modulus of the membrane are crucial factors influencing the actuation strain. Generally, the stiffness of the elastomers can be lowered via addition of plasticisers or through chemical modifications that tailor their molecular weight and cross-linking density.15 Nevertheless, it is worth noting that very soft elastomers tend to exhibit significant viscoelastic effects, which decrease the energy conversion efficiency of DEAs. A high dielectric permittivity effectively increases the capacitance of the DEA, thereby reducing the electric field intensity required to drive the actuator. High permittivity can be achieved either by chemical modification of the elastomer backbone with a high density of polar groups16 or by using a conventional strategy of elastomer composites composed of certain fillers, such as conductive particles17,18 or ceramic fillers with high permittivity.19,20 Note that the filler concentration should remain below the percolation threshold to avoid the formation of a conductive network through the matrix.

It is well established that the maximum actuation strain of the DEA is also strongly affected by pre-stretching of the membrane.21–24 Optimal pre-stretching enhances actuation performance, whereas insufficient pre-stretching can render DEAs prone to electrical breakdown, severely limiting actuation deformation. This phenomenon is commonly attributed to the occurrence of electromechanical instability (EMI), or snap-through instability, which results from the positive feedback between membrane thinning and an increasing electric field.25–32 Theoretical models predict that the EMI is characterized by the abrupt thinning and sudden area expansion of the membrane at a critical voltage. However, this particular failure feature is seldom observed in the actuation plots from numerous experimental tests conducted under low pre-stretch conditions (typically λpre < 2).25,33–37 Instead, these studies often show a continuous increase in actuation strain with the driving voltage until reaching electrical breakdown. Hence, it raises the question of whether the failure of DEAs can be definitively attributed to the EMI mechanism under low pre-stretch conditions.

To elucidate the process of electromechanical instability of the dielectric elastomer actuator, certain in situ experiments were conducted at the macroscopic scale utilizing a thick (typically hundreds of micrometers) and opaque carbon grease electrode, which conceals the actual morphology of the elastomer membrane.25,33–35,38–42 Recently, researchers resorted to transparent and compliant electrodes to study the actuation process of DEAs at the microscopic scale, such as the liquid electrodes of NaCl solution,43–45 silver nanowires46 and carbon nanotube electrodes.47,48 However, in these tests, only one side of the membrane was coated with a compliant electrode, while the other side was bonded to a rigid substrate, which contrasts with the practical operation of DEA that expands without any constraints on either side of the membrane. Therefore, the failure features observed in these studies, such as the creasing and cratering instability phenomena,43–48 may not be directly applicable to the DEAs. Furthermore, these studies did not consider the impact of varying pre-stretch ratios on surface instability. To the best of our knowledge, the microscopic in situ observations of the electromechanical response of DEAs under unconstrained conditions are still few and far between.

An ideal electrode suitable for the in situ investigation of the DEA requires the maintenance of structural integrity with low electrical resistance, even under maximum actuation stretching. Additionally, the electrodes should be sufficiently thin and compliant to minimize the impedance to the deformation of the dielectric elastomer membrane. Consequently, conductive, flexible and transparent electrodes have been widely explored in recent years, such as carbon nanotubes,49–54 silver nanowires (AgNW),55–59 graphene60,61 and electrolytic hydrogel.62–64 These compliant electrodes are not only capable of sustaining extremely large strains to match the deformation of the DE membrane but also exhibit great optical transmittance akin to the thin metal65–67 and traditional transparent electrodes, such as indium tin oxide (ITO).68 Among these, single-walled carbon nanotube (SWCNT) electrodes show unique self-clearing capability when the DE membrane experiences electrical breakdown.50,69 Therefore, the SWCNT electrode is regarded as an ideal choice for the in situ study of DEA, considering its nanometer-scale thickness, high conductivity, great compliance and transparency.

In this paper, we elaborately prepared a small active region (1 × 1 mm2) coated with a compliant and transparent carbon nanotube electrode on both sides and successively detected the electromechanical instability of DEAs under unconstrained conditions at the microscopic scale. Additionally, the impact of various pre-stretch ratios on the failure mechanism of DEA was examined, and the EMI failure of DEA was analyzed based on the thermodynamic principle and phase transition theory.

2. Experimental methods

2.1 DEA fabrication

The single-walled carbon nanotube provided by the National Institute of Advanced Industrial Science and Technology (AIST, Japan) was utilized for the electrode fabrication. 2.5 mg of SWCNT was dispersed in 45 mL of an aqueous sodium dodecyl sulfate (SDS, Sigma-Aldrich Company) surfactant solution (1% wt), and the dispersion was mixed with a cup-horn sonicator (US-50E, NISSEI Corporation) for 80 min at a power of 50 W. The solution was centrifuged at a rotation speed of 15[thin space (1/6-em)]000 rpm for 6 hours (HITACHI CF15RXII), and the upper 75% supernatant liquid was decanted. This procedure yielded a typical nanotube concentration of approximately 20–25 mg L−1.51,70,71 The uniform SWCNT film was fabricated by vacuum filtration of the solution with a Whatman Nuclepore polycarbonate hydrophilic membrane filter with 50 nm pore size and 47 mm diameter. After filtering the CNT solution, the membrane was rinsed six times with pure water to remove residual SDS. By adjusting the volume of the SWCNT dispersion used in the filtration process, SWCNT films with varying areal densities were obtained. The morphology of the SWCNT networks was characterized using a field emission scanning electronic microscope (JSM-7100F JEOL Ltd), and the electrical properties of the SWCNT film were measured with a 4-point probe resistivity measurement system (RTS-9).

The circular elastomer membrane (VHB 4905, 3M Company) with a thickness of H = 0.5 mm was equi-biaxially pre-stretched at a ratio of λpre = 3 (the ratio of the deformed diameter to the initial diameter). The membrane was then masked by release paper with a circular aperture of about 12 mm in diameter. The SWCNT film was transferred onto the adhesive surface of the VHB 4905 by gently stamping the filter onto the membrane. Finally, the filter was removed using a few drops of ethanol. The DE membrane was fully relaxed for 0.5 hour, and then voltage was applied at a ramping rate of 20 V s−1 using a power supply (Model HAR-30P1, Matsusada Precision Inc.). A digital camera was positioned above the actuator to record the area of the electrode during the electrical actuation process. To examine the microscopic morphological evolution of the electrodes during actuation, the active region was observed in situ with an optical microscope (Keynce VHX-600E).

2.2 Transmittance measurements

The optical transmittance of the DEA coated with SWCNT electrodes at 532 nm was measured using a light source and a digital light intensity lux meter (Model SK-20LX, SATO KEIRYOKI MFG. Co., Ltd). The transmittance, T, was calculated as the ratio of the luminous flux transmitted through the DEA to the incident flux, T = I/I0.72 The optical transmittance at the active region was recorded in real time as the driving voltage increased in steps of 0.1 kV.

2.3 In situ observation of DEA failure

To study the effect of the pre-stretch ratio on the actuation performance of the VHB 4905 actuator, the elastomer membrane was equi-biaxially pre-stretched to various ratios (λpre = 1.5, 3 and 5) before being clamped onto an annular rigid frame. The SWCNT electrodes with a width of 1 mm were oriented perpendicular to each other on the opposite sides of the membrane, with an intersecting area of 1 × 1 mm2, as shown in Fig. 1, with the detailed fabrication process illustrated in Fig. S1 (ESI, Appendix). Real-time observations of the active region were conducted using a Keyence VHX-600E microscope equipped with a VH-Z100 lens, offering magnification from 100× to 1000×. Under the multi-direction illumination variation analysis mode, the incident light reached the surface of the membrane at various angles, facilitating the imaging of the uneven surface of the dielectric elastomer caused by the voltage-driven deformation.
image file: d4sm00596a-f1.tif
Fig. 1 Illustrations of the in situ observation of the actuation process of DEA. (a) Schematic of a pre-stretched dielectric elastomer fixed on a rigid annular frame. (b) Images of intersecting electrodes on the opposite sides with a width of 1 mm. (c) Experimental setup for in situ observation.

3. Results

3.1 SWCNT film characterization

Fig. 2(a)–(d) illustrate the morphology of the SWCNT networks on the filter membrane, employing solution volumes of 80 μL, 150 μL, 250 μL, and 600 μL, respectively. It was found that the SWCNT networks were deposited homogeneously on the filter, indicating an even dispersion of carbon nanotube bundles within the SDS solution, without any noticeable aggregation. In Fig. 2(a), when the solution volume was 80 μL, the sparsely packed nanotube network failed to completely cover the filter surface. As the SWCNT concentration increases, a more densely interconnected network of tubes with proliferating conducting channels is formed, as depicted in Fig. 2(b) and (c). Additionally, the thickness of the compact SWCNT film increases as more carbon nanotubes are deposited, as exhibited in Fig. 2(d).
image file: d4sm00596a-f2.tif
Fig. 2 SEM images of the carbon nanotubes filtered with various solution volumes: (a) 80 μL, (b) 150 μL, (c) 250 μL and (d) 600 μL.

The sheet resistance of the SWCNT film, measured using a collinear four-probe array, is presented as a function of the filtered solution volume in Fig. 3(a). The data reveal that the sheet resistance sharply decreases as the solution volume increases from 80 μL to 300 μL. This decline is attributed to the densification of the SWCNT network through successive deposition, which increases the concentration of the conducting nanotube–nanotube contacts,71 thereby enhancing the film's conductivity. The sheet resistance of the SWCNT film levels off when the filtered solution exceeds 300 μL. It is assumed that, at this point, a compact layer of the SWCNT network is formed, and further deposition of SWCNT primarily increases the film thickness, which accounts for the slow reduction in sheet resistance.


image file: d4sm00596a-f3.tif
Fig. 3 (a) Sheet resistance of the SWCNT film as a function of filtered solution volume. (b) Variations in transmittance as a function of the filtered SWCNT solution volume under open circuit conditions; the dashed line represents the fitting data with the Beer–Lambert law. (c) Transmittance–voltage relationship of DEAs during the voltage-driven actuation process. (d) Actuation plots of DEAs (under λpre = 3) with SWCNT electrodes deposited with different solution volumes and the strain–voltage relationship measured with the carbon grease electrode are presented for comparison.

3.2 Transmittance results of DEA

The transmittance (T) of the DEAs coated with SWCNT electrodes of varying areal densities was measured under both open-circuit and voltage-driving conditions, as shown in Fig. 3(b) and (c), respectively. Under open-circuit conditions, the transmittance of the DEA increases as the amount of deposited SWCNT decreases. The dashed plot in Fig. 3(b) was derived by fitting the data with the Beer–Lambert law, given by A = −lg[thin space (1/6-em)]T = εclεV, where A is the absorbance, T is the transmittance, ε is the absorptivity of the materials, c is the concentration of the species, l is the optical path length and V is the filtered volume of the SWCNT solution. Upon applying a voltage, the transmittance of DEAs with the deposited SWCNT exceeding 250 μL progressively increases, which is attributed to the gradual thinning of both the elastomer membrane and the SWCNT network during the actuation process, as indicated in Fig. 3(c). However, in the case of the DEA with 150 μL of deposited SWCNT, the transmittance shows a non-monotonic relationship with the driving voltage, which will be further elucidated in the following section.

The actuation strain of the DEAs with SWCNT electrodes is plotted as a function of applied voltage in Fig. 3(d). The relative linear actuation strain is defined as εl = 100% × (ll0)/l0, where l is the actuated length (voltage-on) and l0 is the unactuated length (voltage-off).11 For comparison, the carbon grease electrode (MG Chemicals Co., Ltd) was chosen due to its minimal stiffening effects and high conductivity throughout the actuation process. For the DEA with 250 μL of SWCNT deposition, the electromechanical response is comparable to that of the DEA with carbon grease electrodes, exhibiting a maximum actuation strain exceeding 90%. In contrast, for the cases of 100 μL and 150 μL of SWCNT deposition, the actuation plots deviate from the behavior of the carbon grease electrode, causing reduced maximum actuation strains of 57% and 67%, respectively. When the SWCNT deposition exceeds 600 μL, the stiffening effect of the SWCNT network becomes pronounced, shifting the plots towards higher voltage levels. The stiffening effect of the thick SWCNT electrodes was confirmed through a customized test bench under equi-biaxial tensile loading, the details of which are described in Fig. S2 and S3 (ESI, Appendix).

To clarify the real-time morphological evolution of the SWCNT electrodes during the actuation process, the active region was carefully monitored at both macroscopic and microscopic scales as the driving voltage increased. Fig. 4 displays the images of DEA with 150 μL of SWCNT. Fig. 4(a)–(c) present the overall appearance of DEA at voltages of 0 V, 2850 V and 3700 V, respectively, while Fig. 4(d)–(f) exhibit the corresponding microscopic features of the active region. Initially, before applying the voltage, the transmittance of the active region is around 84%, with the SWCNT electrode film uniformly distributed across the membrane surface, as indicated in Fig. 4(a) and (d). As the driving voltage increases to 2850 V, the active region appears frosted and hazy, as shown in Fig. 4(b), which accounts for the transmittance reduction in Fig. 3(c). Microscopically, the initially homogeneous SWCNT film splits due to excessive stretching of the DE membrane, as evidenced in Fig. 4(e). The interconnected nanotube network ruptured with numerous isolated domains exposed to air, thereby diminishing the charge storage capacity of the electrodes. As the voltage further increased, the electromechanical performance continually deteriorated as the cracking area expanded, as shown in Fig. 4(c) and (f), leading to the reduction in the actuation strain, as shown in Fig. 3(d). In contrast, for the DEA with 250 μL SWCNT deposition, the transmittance of the active region consistently increases with the driving voltage, as indicated in Fig. 5(a)–(c) at applied voltages of 0 V, 2980 V and 3500 V, respectively. Microscopic examination revealed that the SWCNT network maintained its structural integrity throughout the actuation process, as seen in Fig. 5(d)–(f). This uniform, thin and transparent SWCNT layer endows DEA with electromechanical performance on par with that of carbon grease electrodes in Fig. 3(d), making it feasible to investigate the failure mechanisms of DEAs in situ at the microscopic scale.


image file: d4sm00596a-f4.tif
Fig. 4 Photographs of DEA (under λpre = 3) coated with 150 μL of the deposited SWCNT electrode. (a)–(c) Global appearance of the DEA at applied voltages of 0 V, 2850 V and 3700 V, respectively, and the corresponding microscopic morphologies of the electrodes are presented in (d)–(f). The inset shows a zoom-in image of the region enclosed by a dashed box.

image file: d4sm00596a-f5.tif
Fig. 5 Photographs of DEA (under λpre = 3) coated with 250 μL of the deposited SWCNT electrode. (a)–(c)Global appearance of the DEA at voltages of 0 V, 2980 V and 3500 V, respectively, and the corresponding microscopic morphologies of the electrodes are presented in (d)–(f).

3.3 Electro-mechanical failure at the microscopic scale

The SWCNT electrode, fabricated with 250 μL of the solution, was employed for the in situ observation in the context of the above results. Fig. 6 shows the actuation behavior of the DEAs with an active area of 1 mm2, subjected to pre-stretches of λpre = 1.5, 3 and 5, respectively. For λpre = 1.5, the membrane initially expands uniformly with an increase in the voltage, and then a localized region (denoted by arrow A) exhibits snap-through deformation, as indicated in Fig. 6(b) and (c). This domain experiences a sudden expansion at nearly constant voltage, causing abrupt thinning of the membrane, which results in electric field surging and ultimately triggers electrical breakdown. Fig. S4(a) and (b) (ESI, Appendix) present the actuation strain plots of the local domain-A and the entire active region, respectively. For the entire active region, the actuation strain exhibits a smooth and monotonic increase as the driving voltage increases, with no discernible snap-through instability identified from the actuation strain plot, which is consistent with the experimental findings in the literature.25,33–37 In contrast, the snap-through instability was apparent in the actuation plot for the local domain-A, as shown in Fig. S4(a) (ESI). These observations suggest that the theoretically predicted electromechanical instability, characterized by discontinuous deformation, occurs within the micro-sized domain and may hardly be detected via conventional global observations. To confirm the reproducibility of these findings, additional tests were conducted under the same conditions, and a similar phenomenon was detected with snap-through instability occurring away from the corner of the square-shaped region, as shown in Fig. S5 (ESI, Appendix). Therefore, the observed localized sudden thinning of the dielectric elastomer is mainly caused by EMI, rather than the stress concentration.
image file: d4sm00596a-f6.tif
Fig. 6 Microscopic in situ observations of the active region for DEA with different pre-stretches. (a)–(d) For λpre = 1.5, with applied voltages of (a) 0 V, (b) 6682 V, (c) 6686 V and (d) 6700 V, respectively. The localized domain showing snap-through instability is denoted by arrows. (e)–(h) For λpre = 3, with applied voltages of (e) 0 V, (f) 3370 V, (g) 3720 V and (h) 3726 V, respectively. (i)–(l) For λpre = 5, with applied voltages of (i) 0 V, (j) 3000 V, (k) 3500 V and (l) 3590 V, respectively.

For the DEA under a pre-stretch of λpre = 3, as the actuation strain increases to approximately 110%, distinct wrinkles are formed due to the excessive planar expansion of the membrane (Fig. 6(f)), which is referred to as the loss of tension (LT) in the literature.42 Although the acrylic elastomer is commonly regarded as an electrical insulator, its conductivity (κ) increases with the applied electric field (E) following an exponential law, κ = κ0[thin space (1/6-em)]exp(E/EC), where κ0 and EC are material constants. Consequently, considerable charge leakage occurs when subjected to extreme stretching. As a result, the membrane elastically contracts and restores the flat morphology due to the reduction in Maxwell stress (Fig. 6(g)), followed immediately by the final electrical breakdown. The corresponding actuation strain plot is shown in Fig. S4(c) (ESI, Appendix).

For the DEA under a λpre = 5, the maximum actuation strain is relatively low due to the pronounced stiffening of the acrylic elastomer under large deformation, as indicated in Fig. S4(d) (ESI, Appendix). Despite a slight increase in the roughness of the membrane at high voltage, the overall morphology of the membrane remains flat throughout the actuation process, as shown in Fig. 6(i)–(l).

4. Discussion

4.1 Mechanisms for localized electromechanical instability

Considering a homogeneous DE membrane subjected to an external force and voltage, with the initial dimensions of L1 (length), L2 (width) and H (thickness), the membrane deforms to the dimensions of l1 = λ1L1, l2 = λ2L2 and h = λ3H due to the combination of mechanical pre-stretch and electrical Maxwell stress. Assuming an ideal dielectric elastomer model, where the material exhibits a purely elastic response and is incompressible (λ1λ2λ3 = 1), the permittivity is independent of deformation. The Helmholtz free energy density of this system can be given as a function of the stretch ratios, λ1 and λ227:
 
image file: d4sm00596a-t1.tif(1)
where Ws is the elastically stored strain energy density for the elastomer, P1 and P2 are the pre-stretched forces in the plane, ε0 is the free space permittivity, εr is the relative dielectric constant (typical value of 3.9 for VHB 490525) and Φ is the driving voltage. For equi-biaxial tensile loading, where P1 = P2 = P and λ1 = λ2 = λ, the thermodynamic principle dictates that an equilibrium state should minimize F, i.e., ∂F/∂λ = 0, and accordingly, the relationship between the stretch λ and the applied voltage Φ can be obtained as
 
image file: d4sm00596a-t2.tif(2)
where the two terms on the left denote the pre-stretched and Maxwell stresses, respectively, and the right term is the elastic recovery stress of the DE membrane. From this point of view, eqn (2) describes the force balance within the system per se.

In this study, the Gent hyperelastic model was employed to characterize the VHB 4905 elastomer under equi-biaxial tensile loading:73

 
image file: d4sm00596a-t3.tif(3)
where μ represents the shear modulus in the small strain range, and Jm denotes a dimensionless parameter that defines the limiting stretch when the polymer chains approach their contour lengths under external loading. The typical values of μ = 40 kPa and Jm = 125 are selected for the commonly used VHB 4905 acrylic elastomer.25,27

For a defect-free DEA under a pre-stretch of λpre = 1.5, the relationship between stretches and the applied voltage was computed based on eqn (1)–(3) and is presented in Fig. 7(a). It is observed that the stretches monotonically increase with the driving voltage at the initial stage (Φ < ΦA), indicating that the membrane reaches an equilibrium state characterized by a specific stretch value at a given voltage. However, as the voltage increases to the values between ΦA and ΦE, three distinct equilibrium stretches are determined on the plot at a certain voltage. Once the driving voltage exceeds ΦE, the equilibrium stretch abruptly jumps to E′, signifying the occurrence of electromechanical instability, given by dΦ(λ)/dλ = 0,74 where ΦE is the critical voltage for the onset of EMI.33,75,76


image file: d4sm00596a-f7.tif
Fig. 7 Calculated actuation plots of the defective domains and defect-free matrix under the pre-stretches of (a) and (b) λpre = 1.5, (c) λpre = 3 and (d) λpre = 5. ΦE, ΦEV and ΦEP denote the critical EMI voltage for the defect-free matrix, air-void domain and conductive-particle domain, respectively. Asterisks correspond to the points of electrical breakdown.

The DEA eventually undergoes electrical breakdown under an excessively high electric field, which can be expressed as

 
Φ = EEB−2 (4)
where EEB is the electrical breakdown field, which is dependent on the stretch of the membrane and can be empirically calculated as EEB = E0λ1.13,74 where E0 is the electrical breakdown field of the unstretched elastomer and is approximately 30.6 MV m−1 for the VHB 4905. Consequently, the operational limit can be determined using eqn (4) and is presented on the voltage–stretch plane in Fig. 7(a). The snap-through instability results in a sharp increase in the stretch at a critical voltage ΦE, ultimately intersecting with the electrical breakdown limit, as indicated by asterisks in Fig. 7(a). It is noteworthy that for a given pre-stretch ratio, the critical voltage for EMI increases linearly with the film thickness, H, while the critical stretch at EMI remains constant.

Although the dielectric elastomer membrane is conventionally considered to be structurally homogeneous, with uniform mechanical and electrical properties across all locations, the presence of impurities during the fabrication process is inevitable, such as microscopic air voids and conductive particles.27,50 In this study, it is assumed that the former defect modifies the mechanical properties of the membrane by reducing the local elastic modulus, whereas the latter defect alters the electrical properties of the elastomer by increasing the local permittivity.

When considering a small domain containing microscopic air voids, it is reasonable to assume that both the small-strain shear modulus (μ) and ultimate limiting stretch (Jm) of this local region would decrease slightly. In this study, these values are assumed to be 5% of those of the defect-free matrix. To simplify the analysis, the influence of air voids on the electrical properties of the elastomer is neglected. The actuation plot of this void-containing domain was calculated and is presented in Fig. 7(a). Initially, the plot for the void-containing domain closely aligns with that of the defect-free matrix. However, when the applied voltage exceeds 4000 V, a noticeable bifurcation emerges, indicating a marginally higher actuation strain for the void-containing domain at a given voltage. Notably, the critical voltage for snap-through instability shifts to a lower level compared to that of the matrix, transitioning from ΦE to ΦEV, which suggests that the void-containing domain is more susceptible to sudden EMI than the defect-free matrix. This finding corroborates the phenomenon depicted in Fig. 6, wherein a burst of deformation occurs in a highly localized domain, while the matrix sustains continuous extension. Although the assumption of a 5% reduction in μ and Jm is somewhat arbitrary, it can be adjusted to other small values, such as 3%, without altering the main conclusion, as illustrated in Fig. S6 (ESI, Appendix).

In addition to microscopic air voids, the elastomer membrane may also contain conductive particles, leading to an increase in the dielectric constant within localized regions. For simplicity, it is hypothesized that micro-sized conductive particles have a negligible impact on the mechanical properties of the elastomer. Fig. 7(b) illustrates the calculated voltage–stretch curve for the particle-containing domain, assuming a 5% increase in the relative dielectric constant vis-à-vis the surrounding matrix. The presence of conductive particles induces a bifurcation in the curves between the particle-containing domain and the surrounding matrix, akin to the effect caused by microscopic air voids, with the critical voltage for EMI shifting towards a lower value ΦEP. In the context of the above analysis, both the microscopic air void and conductive particles contribute to EMI occurring in the defective regions before it occurs in the defect-free matrix.

A similar analysis was implemented on the DEAs with pre-stretch ratios of λpre = 3 and λpre = 5, and the voltage–stretch curves for both the defect-free matrix and defective domain are presented in Fig. 7(c) and (d), respectively. It is evident that the stretch ratio monotonically increases with the applied voltage, indicating the elimination of snap-through instability. Although there are slight disparities in the actuation strain between the defective domain and matrix, they are insufficient to trigger instability in the defective region. Thus, the employment of a high pre-stretch not only suppresses electromechanical instability but also mitigates the adverse impacts of foreign impurities on the homogeneous deformation of DEA.

Achieving precise equations to describe the electric-mechanical coupling around imperfections remains challenging due to the complex variations in the stress and electrical fields, especially under large nonlinear deformations, as illustrated by the finite element analysis (FEA) results shown in Fig. S7–S9 (ESI, Appendix). Although the detailed distribution of the stress and electric field was not fully considered in the above analysis, the results obtained were reasonable and did not impact the main conclusion.

4.2 Electromechanical phase transition

The influence of micro-sized defects on the electromechanical instability of DEA can also be analyzed through the lens of phase transition theory. For λpre = 1.5, the free energy surface is calculated using eqn (1) and presented as a function of the stretch (λ) and applied voltage (Φ) in Fig. 8(a). It was found that more than one extreme value exists on the free energy surface. At low voltages, e.g., ΦA = 3860 V, the free energy density demonstrates a single minimum as a function of stretch λ in Fig. 8(c), representing a thermodynamically stable state, as illustrated by point A1 in Fig. 8(b). As the applied voltage increases to ΦB = 3980 V, the free energy plot exhibits two local minima (B1 and B3) and one local maximum (B2) in Fig. 8(d), which are also highlighted in the voltage–stretch curve in Fig. 8(b). The local maximum point (B2) with intermediate stretch is unstable against small perturbations, while the other two states, B1 and B3, are stable against small perturbations, which are termed as “thick-phase” with a smaller stretch and “thin-phase” with a larger stretch, respectively. As the free energy of the “thick-phase” is lower than that of the “thin-phase” at ΦB, the former retains its stability while the latter is in a metastable state, making phase transition unlikely. As the applied voltage reaches ΦC = 4066 V, the free energies of the “thick-phase” and “thin-phase” become equal, indicating the co-existence of two states, with ΦC referred to as the transition voltage, as illustrated in Fig. 8(e). With further voltage increase toΦD, the free energy of the “thin-phase” drops below that of the “thick-phase”, implying that the “thick-to-thin” phase transition is thermodynamically possible at this stage. However, such a phase transition requires overcoming the energy barrier (ΔU), as depicted in Fig. 8(f). Hence, a carefully manufactured DE membrane, free of defects, would remain in the “thick-phase” state until it is overcharged to a voltage of ΦE, where the energy barrier vanishes and only one minimum exists in the free energy plot, as illustrated in Fig. 8(g). At this critical point, the DEA undergoes a snap-through deformation, transitioning from the “thick-phase” to the energetically stable “thin-phase”.
image file: d4sm00596a-f8.tif
Fig. 8 Calculated free energy density of DEA under pre-stretch λpre = 1.5. (a) Free energy density surface as a function of the stretch, λ, and the applied voltage, Φ. The transition voltage ΦC is highlighted by the cross section. (b) Actuation plot marked with typical driving voltages of ΦA = 3860 V, ΦB = 3980 V, ΦC = 4066 V (transition voltage), ΦD = 4300 V and ΦE = 4996 V (critical voltage). (c)–(g) Free energy density plots corresponding to the driving voltages, ΦAΦE, labeled in image (b). (h) Energy barriers for the phase transition in the void-containing domain (ΔUv) and defect-free matrix (ΔU) at an applied voltage of ΦD, respectively.

At the molecular level, distinct configurations can be detected between the “thick-phase” and “thin-phase”. When a voltage is applied, the dipole moments of the polar groups tend to align with the direction of the electric field. During electromechanical instability, the abrupt thinning of the elastomer causes a local surge in the electric field, resulting in a more uniform orientation of polar groups within the “thin-phase” compared to the surrounding “thick-phase”. Additionally, the polymer chains in the “thin-phase” are stretched to a greater length due to the abrupt planar expansion, as schematically shown in Fig. S10 (ESI, Appendix).

The essence of snap-through instability is that more than one extreme value exists on the free energy surface, resulting in two equilibrium states characterized by different stretches at a certain voltage Φ. Therefore, for a dielectric elastomer material (not limited to VHB 4905), as long as its mechanical parameter Ws and dielectric permittivity (εr) give rise to more than one extreme value of F in eqn (1), snap-through instability will occur during the actuation process. Analysis of the free energy surfaces for dielectric elastomer actuators (DEAs) with pre-stretch ratios of λpre = 3 and λpre = 5 revealed only a single minimum throughout the entire operational voltage range, as depicted in Fig. S11 (ESI, Appendix), suggesting the absence of any phase transitions during the actuation process.

The above process is reminiscent of homogeneous nucleation in the water–ice phase transition during cooling, with the “thick-phase” and “thin-phase” corresponding to the water and ice phases, respectively. At a temperature of 0 °C, the free energies of water and ice are equal, analogous to the transition voltage of ΦC in Fig. 8(e). Actually, crystallization rarely occurs spontaneously at temperatures just below 0 °C, despite ice being more energetically stable than water. This is because the liquid must overcome an energy barrier to initiate nucleation, counteracting the increase of the surface free energy due to the formation of a solid–liquid phase boundary during solidification, analogous to the behavior in Fig. 8(f). For pure impurity-free water, a higher activation free energy is required to generate a stable nucleus in the interior of the liquid in the form of atom clusters. This homogeneous nucleation mechanism allows pure water to be supercooled to a temperature well below the equilibrium solidification temperature, even below −41 °C,77 analogous to a perfect (defect-free) DEA overcharged to the critical voltage of ΦE before undergoing phase transition. However, for water containing a mass of impurities, the energy barrier for heterogeneous nucleation is lowered as the nuclei form on preexisting interfaces, enabling the water to readily undergo solidification at temperatures slightly below 0 °C. Similarly, the flaws in the DE membrane, such as microscale air voids and conductive particles, act as preferential nucleation sites that significantly reduce the energy barrier for the “thick-to-thin” phase transition, from ΔU to ΔUv, as shown in Fig. 8(h). Consequently, the defective domain preferentially undergoes localized phase transition prior to the rest of the DE membrane. Particularly, when the critical point of EMI at the defective domain (ΦEV or ΦEP in Fig. 7(a) and (b)) decreases to the transition voltage of the matrix (ΦC in Fig. 8(b)), the DEA will experience a phase transition (i.e., EMI) just above the transition voltage, ΦC, analogous to the crystallization of impure water at temperatures slightly below 0 °C.

5. Conclusions

In this work, an SWCNT electrode with exceptional optical and electrical properties was fabricated, and its microscopic morphology was monitored in real time throughout the voltage-driven actuation process. Both excessively thin and overly thick SWCNT electrodes degrade the electromechanical performance of DEA, which is attributed to microscopic cracking in the thin electrodes and increased stiffness in the thick electrodes. The impact of pre-stretching on the failure mechanism of the DEA was studied in situ utilizing optimized SWCNT electrodes. Under low pre-stretch conditions, snap-through instability occurs in a highly localized domain, which is ascribed to the heterogeneity of the elastomer materials, such as microscopic air voids or conductive particles that act as nucleation sites for the new phase, thereby reducing the energy barrier for phase transition and shifting the critical voltage of snap-through instability to a lower level. Our findings clarify the discrepancy between the theoretically predicted abrupt deformation and the experimentally observed continuous expansion of DEAs under low pre-stretch conditions. This research highlights the critical role of material purity in the electromechanical response of DEAs, especially those operating under low pre-stretch conditions.

Author contributions

Zhen-Qiang Song: conceptualization, investigation, writing – original draft, and funding acquisition. Li-Min Wang: writing – review & editing. Yongri Liang: methodology and writing – review & editing. Xiao-Dong Wang: validation and formal analysis. Shijie Zhu: writing – review & editing, supervision, and funding acquisition.

Data availability

The data supporting this article have been included as part of the ESI.

Conflicts of interest

There are no conflicts of interest to declare.

Acknowledgements

This work was supported by the National Natural Science Foundation of China under Grant No. 52201191 and the Hebei Natural Science Foundation under Grant No. E2022203108, and funded by the Science and Technology Project of Hebei Education Department under Grant No. QN2022055, and the Innovation Ability Promotion Program of Hebei under Grant No. 22567609H.

References

  1. Y. Chen, H. Zhao, J. Mao, P. Chirarattananon, E. F. Helbling and N. P. Hyun, et al., Controlled flight of a microrobot powered by soft artificial muscles, Nature, 2019, 575(7782), 324–329 CrossRef CAS .
  2. G. Li, X. Chen, F. Zhou, Y. Liang, Y. Xiao and X. Cao, et al., Self-powered soft robot in the Mariana Trench, Nature, 2021, 591(7848), 66–71 CrossRef CAS .
  3. Y. Shi, E. Askounis, R. Plamthottam, T. Libby, Z. Peng and K. Youssef, et al., A processable, high-performance dielectric elastomer and multilayering process, Science, 2022, 377(6602), 228–232 CrossRef CAS .
  4. Y. Jiang, S. Liu, M. Zhong, L. Zhang, N. Ning and M. Tian, Optimizing energy harvesting performance of cone dielectric elastomer generator based on VHB elastomer, Nano Energy, 2020, 71, 104606 CrossRef CAS .
  5. E. Hajiesmaili and D. R. Clarke, Optically addressable dielectric elastomer actuator arrays using embedded percolative networks of zinc oxide nanowires, Mater. Horiz., 2022, 9(12), 3110–3117 RSC .
  6. A. J. Cohen, M. Kollosche, M. C. Yuen, D. Y. Lee, D. R. Clarke and R. J. Wood, Batch-Sprayed and Stamp-Transferred Electrodes: A New Paradigm for Scalable Fabrication of Multilayer Dielectric Elastomer Actuators, Adv. Funct. Mater., 2022, 32(43), 2205394 CrossRef CAS .
  7. E. Hajiesmaili, N. M. Larson, J. A. Lewis and D. R. Clarke, Programmed shape-morphing into complex target shapes using architected dielectric elastomer actuators, Sci. Adv., 2022, 8(28), eabn9198 CrossRef PubMed .
  8. P. Banet, N. Zeggai, J. Chavanne, G. T. M. Nguyen, L. Chikh and C. Plesse, et al., Evaluation of dielectric elastomers to develop materials suitable for actuation, Soft Matter, 2021, 17(48), 10786–10805 RSC .
  9. M. Duduta, E. Hajiesmaili, H. Zhao, R. J. Wood and D. R. Clarke, Realizing the potential of dielectric elastomer artificial muscles, Proc. Natl. Acad. Sci. U. S. A., 2019, 116(7), 2476–2481 CrossRef CAS .
  10. Y. Qiu, E. Zhang, R. Plamthottam and Q. Pei, Dielectric elastomer artificial muscle: materials innovations and device explorations, Acc. Chem. Res., 2019, 52(2), 316–325 CrossRef CAS .
  11. R. Pelrine, R. Kornbluh, Q. Pei and J. Joseph, High-speed electrically actuated elastomers with strain greater than 100%, Science, 2000, 287(5454), 836–839 CrossRef CAS PubMed .
  12. H. B. Fu, Y. Jiang, J. Lv, Y. Huang, Z. P. Gai and Y. Liu, et al., Multilayer Dielectric Elastomer with Reconfigurable Electrodes for Artificial Muscle, Adv. Sci., 2023, 10, 2206094 CrossRef CAS .
  13. J. Li, L. Liu, Y. Liu and J. Leng, Dielectric Elastomer Spring-Roll Bending Actuators: Applications in Soft Robotics and Design, Soft Rob., 2019, 6(1), 69–81 CrossRef .
  14. J. Li, Y. Wang, L. Liu, S. Xu, Y. Liu and J. Leng, et al., A Biomimetic Soft Lens Controlled by Electrooculographic Signal, Adv. Funct. Mater., 2019, 29(36), 1903762 CrossRef .
  15. A. G. Bejenariu, L. Yu and A. L. Skov, Low moduli elastomers with low viscous dissipation, Soft Matter, 2012, 8(14), 3917–3923 RSC .
  16. D. M. Opris, Polar Elastomers as Novel Materials for Electromechanical Actuator Applications, Adv. Mater., 2017, 30, 5 Search PubMed .
  17. G. Kofod, S. Risse, H. Stoyanov, D. N. McCarthy, S. Sokolov and R. Kraehnert, Broad-spectrum enhancement of polymer composite dielectric constant at ultralow volume fractions of silica-supported copper nanoparticles, ACS Nano, 2011, 5(3), 1623–1629 CrossRef CAS PubMed .
  18. B. P. Sahoo, K. Naskar, R. N. P. Choudhary, S. Sabharwal and D. K. Tripathy, Dielectric relaxation behavior of conducting carbon black reinforced ethylene acrylic elastomer vulcanizates, J. Appl. Polym. Sci., 2012, 124(1), 678–688 CrossRef CAS .
  19. X. Huang, L. Xie, Z. Hu and P. Jiang, Influence of BaTiO3 nanoparticles on dielectric, thermophysical and mechanical properties of ethylene-vinyl acetate elastomer/BaTiO3 microcomposites, IEEE Trans. Dielectr. Electr. Insul., 2011, 18, 375–383 CAS .
  20. D. N. Mc Carthy, S. Risse, P. Katekomol and G. Kofod, The effect of dispersion on the increased relative permittivity of TiO2/SEBS composites, J. Phys. D: Appl. Phys., 2009, 42(14), 145406 CrossRef .
  21. Y. P. Su, D. Riccobelli, Y. J. Chen, W. Q. Chen and P. Ciarletta, Tunable morphing of electroactive dielectric-elastomer balloons, Proc. R. Soc. A, 2023, 479(2276), 20230358 CrossRef .
  22. S. Y. Yang, X. H. Zhao and P. Sharma, Avoiding the pull-in instability of a dielectric elastomer film and the potential for increased actuation and energy harvesting, Soft Matter, 2017, 13(26), 4552–4558 RSC .
  23. H. L. Li, L. L. Chen, C. Zhao and S. Y. Yang, Evoking or suppressing electromechanical instabilities in soft dielectrics with deformation-dependent dielectric permittivity, Int. J. Mech. Sci., 2021, 202, 106507 CrossRef .
  24. Y. B. Fu and X. Yu, Axisymmetric necking of a circular electrodes-coated dielectric membrane, Mech. Mater., 2023, 181, 104645 CrossRef .
  25. T. Lu, S. Cheng, T. Li, T. Wang and Z. Suo, Electromechanical Catastrophe, Int. J. Appl. Mech., 2016, 8(07), 1640005 CrossRef .
  26. X. Zhao and Z. Suo, Method to analyze electromechanical stability of dielectric elastomers, Appl. Phys. Lett., 2007, 91(6), 061921 CrossRef .
  27. S. J. A. Koh, C. Keplinger, R. Kaltseis, C.-C. Foo, R. Baumgartner and S. Bauer, et al., High-performance electromechanical transduction using laterally-constrained dielectric elastomers part I: Actuation processes, J. Mech. Phys. Solids, 2017, 105, 81–94 CrossRef CAS .
  28. Z.-Q. Song, K. Ohyama, S. Shian, D. R. Clarke and S. Zhu, Power generation performance of dielectric elastomer generator with laterally-constrained configuration, Smart Mater. Struct., 2020, 29(1), 015018 CrossRef CAS .
  29. W. J. Sun, W. T. Ma, F. Zhang, W. Hong and B. Li, Snap-through path in a bistable dielectric elastomer actuator, J. Appl. Math. Mech., 2022, 43(8), 1159–1170 CrossRef .
  30. H. Conroy Broderick, M. Righi, M. Destrade and R. W. Ogden, Stability analysis of charge-controlled soft dielectric plates, Int. J. Eng. Sci., 2020, 151, 103280 CrossRef .
  31. X. Lv, L. Liu, J. Leng and Y. Liu, Time-dependent electromechanical phase transition and bulging propagation in a viscoelastic dielectric elastomer tube, Smart Mater. Struct., 2020, 29, 015005 CrossRef .
  32. T. Lu, C. Ma and T. Wang, Mechanics of dielectric elastomer structures: A review, Extreme Mech. Lett., 2020, 38, 100752 CrossRef .
  33. J. Huang, T. Li, C. Chiang Foo, J. Zhu, D. R. Clarke and Z. Suo, Giant, voltage-actuated deformation of a dielectric elastomer under dead load, Appl. Phys. Lett., 2012, 100(4), 041911 CrossRef .
  34. S. J. A. Koh, T. Li, J. Zhou, X. Zhao, W. Hong and J. Zhu, et al., Mechanisms of large actuation strain in dielectric elastomers, J. Polym. Sci. Pol. Phys., 2011, 49(7), 504–515 CrossRef CAS .
  35. C. Keplinger, T. Li, R. Baumgartner, Z. Suo and S. Bauer, Harnessing snap-through instability in soft dielectrics to achieve giant voltage-triggered deformation, Soft Matter, 2012, 8(2), 285–288 RSC .
  36. T. Lu, J. Huang, C. Jordi, G. Kovacs, R. Huang and D. R. Clarke, et al., Dielectric elastomer actuators under equal-biaxial forces, uniaxial forces, and uniaxial constraint of stiff fibers, Soft Matter, 2012, 8(22), 6167–6173 RSC .
  37. G. Zurlo, M. Destrade and T. Lu, Fine tuning the electro-mechanical response of dielectric elastomers, Appl. Phys. Lett., 2018, 113(16), 162902 CrossRef .
  38. X. Liu, B. Li, H. Chen, S. Jia and J. Zhou, Voltage-induced wrinkling behavior of dielectric elastomer, J. Appl. Polym. Sci., 2016, 133(14), 43258 CrossRef .
  39. J. Zhu, M. Kollosche, T. Lu, G. Kofod and Z. Suo, Two types of transitions to wrinkles in dielectric elastomers, Soft Matter, 2012, 8(34), 8840–8846 RSC .
  40. J.-S. Plante and S. Dubowsky, Large-scale failure modes of dielectric elastomer actuators, Int. J. Solids Struct., 2006, 43(25), 7727–7751 CrossRef CAS .
  41. Y. Li and T. Zhang, Modeling and Characterizing Two Dielectric Elastomer Folding Actuators for Origami-Inspired Robot, IEEE Rob. Autom. Lett., 2022, 7(4), 11775–11782 Search PubMed .
  42. A. Khurana, M. M. Joglekar and G. Zurlo, Electromechanical stability of wrinkled dielectric elastomers, Int. J. Solids Struct., 2022, 246–247, 111613 CrossRef CAS .
  43. Q. Wang, M. Tahir, J. Zang and X. Zhao, Dynamic electrostatic lithography: multiscale on-demand patterning on large-area curved surfaces, Adv. Mater., 2012, 24(15), 1947–1951 CrossRef CAS .
  44. Q. Wang, L. Zhang and X. Zhao, Creasing to cratering instability in polymers under ultrahigh electric fields, Phys. Rev. Lett., 2011, 106(11), 118301 CrossRef PubMed .
  45. H. S. Park, Q. Wang, X. Zhao and P. A. Klein, Electromechanical instability on dielectric polymer surface: Modeling and experiment, Comput. Methods Appl. Mech. Eng., 2013, 260, 40–49 CrossRef .
  46. S. Shian and D. R. Clarke, Electrically-tunable surface deformation of a soft elastomer, Soft Matter, 2016, 12(13), 3137–3141 RSC .
  47. S. Shian, P. Kjeer and D. R. Clarke, Electric-field induced surface instabilities of soft dielectrics and their effects on optical transmittance and scattering, J. Appl. Phys., 2018, 123, 11 CrossRef .
  48. P. Charoen-Rajapark and D. R. Clarke, Confocal microscopy observations of electrical pre-breakdown of bi-layer elastomer dielectrics, Extreme Mech. Lett., 2021, 49, 101473 CrossRef .
  49. M. Duduta, R. J. Wood and D. R. Clarke, Multilayer dielectric elastomers for fast, programmable actuation without prestretch-Supplementary Material, Adv. Mater., 2016, 28(36), 8058–8063 CrossRef CAS PubMed .
  50. W. Yuan, L. Hu, Z. Yu, T. Lam, J. Biggs and S. M. Ha, et al., Fault-Tolerant Dielectric Elastomer Actuators using Single-Walled Carbon Nanotube Electrodes, Adv. Mater., 2008, 20(3), 621–625 CrossRef CAS .
  51. S. Shian, R. M. Diebold, A. McNamara and D. R. Clarke, Highly compliant transparent electrodes, Appl. Phys. Lett., 2012, 101(6), 061101 CrossRef .
  52. C. Baechler, S. Gardin, H. Abuhimd and G. Kovacs, Inkjet printed multiwall carbon nanotube electrodes for dielectric elastomer actuators, Smart Mater. Struct., 2016, 25(5), 055009 CrossRef .
  53. X. Guo, Y. Huang, Y. Zhao, L. Mao, L. Gao and W. Pan, et al., Highly stretchable strain sensor based on SWCNTs/CB synergistic conductive network for wearable human-activity monitoring and recognition, Smart Mater. Struct., 2017, 26(9), 095017 CrossRef .
  54. J. H. Pu, Y. Meng, Z. X. Xie, Z. H. Peng, J. H. Wu and Y. Shi, et al., A unimorph nanocomposite dielectric elastomer for large out-of-plane actuation, Sci. Adv., 2022, 8(9), eabm6200 CrossRef CAS PubMed .
  55. Y. R. Lee, H. Kwon, D. H. Lee and B. Y. Lee, Highly flexible and transparent dielectric elastomer actuators using silver nanowire and carbon nanotube hybrid electrodes, Soft Matter, 2017, 13(37), 6390–6395 RSC .
  56. J. Lee, J. Y. Woo, J. T. Kim, B. Y. Lee and C.-S. Han, Synergistically enhanced stability of highly flexible silver nanowire/carbon nanotube hybrid transparent electrodes by plasmonic welding, ACS Appl. Mater. Interfaces, 2014, 6(14), 10974–10980 CrossRef CAS PubMed .
  57. W. Hu, X. Niu, L. Li, S. Yun, Z. Yu and Q. Pei, Intrinsically stretchable transparent electrodes based on silver-nanowire-crosslinked-polyacrylate composites, Nanotechnology, 2012, 23(34), 344002 CrossRef PubMed .
  58. L. Zhang, T. Song, L. Shi, N. Wen, Z. Wu and C. Sun, et al., Recent progress for silver nanowires conducting film for flexible electronics, J. Nanostruct. Chem., 2021, 11(3), 323–341 CrossRef CAS PubMed .
  59. D. Tan, C. Jiang, Q. Li, S. Bi and J. Song, Silver nanowire networks with preparations and applications: a review, J. Mater. Sci.: Mater. Electron., 2020, 31(18), 15669–15696 CrossRef CAS .
  60. T. Hwang, H.-Y. Kwon, J.-S. Oh, J.-P. Hong, S.-C. Hong and Y. Lee, et al., Transparent actuator made with few layer graphene electrode and dielectric elastomer, for variable focus lens, Appl. Phys. Lett., 2013, 103(2), 023106 CrossRef .
  61. J. Zang, S. Ryu, N. Pugno, Q. Wang, Q. Tu and M. J. Buehler, et al., Multifunctionality and control of the crumpling and unfolding of large-area graphene, Nat. Mater., 2013, 12(4), 321–325 CrossRef CAS PubMed .
  62. B. Chen, J. J. Lu, C. H. Yang, J. H. Yang, J. Zhou and Y. M. Chen, et al., Highly stretchable and transparent ionogels as nonvolatile conductors for dielectric elastomer transducers, ACS Appl. Mater. Interfaces, 2014, 6(10), 7840–7845 CrossRef CAS PubMed .
  63. C. Keplinger, J.-Y. Sun, C. C. Foo, P. Rothemund, G. M. Whitesides and Z. Suo, Stretchable, transparent, ionic conductors, Science, 2013, 341(6149), 984–987 CrossRef CAS PubMed .
  64. Y. Cao, T. G. Morrissey, E. Acome, S. I. Allec, B. M. Wong and C. Keplinger, et al., A Transparent, Self-Healing, Highly Stretchable Ionic Conductor, Adv. Mater., 2017, 29(10), 1605099 CrossRef PubMed .
  65. S. Rosset and H. R. Shea, Flexible and stretchable electrodes for dielectric elastomer actuators, Appl. Phys. A: Mater. Sci. Process., 2013, 110(2), 281–307 CrossRef CAS .
  66. R. P. Nie, W. B. Tang, Y. Li, L. C. Jia, L. Xu and H. D. Huang, et al., Surfactant-assisted fabrication of room-temperature self-healable dielectric elastomer toward actuation application, Composites, Part B, 2022, 234, 109655 CrossRef CAS .
  67. J. Hubertus, S. Croce, J. Neu, S. Seelecke, G. Rizzello and G. Schultes, Laser Structuring of Thin Metal Films of Compliant Electrodes on Dielectric Elastomers, Adv. Funct. Mater., 2023, 33(16), 2214176 CrossRef CAS .
  68. D. McCoul, W. Hu, M. Gao, V. Mehta and Q. Pei, Recent advances in stretchable and transparent electronic materials, Adv. Electron. Mater., 2016, 2(5), 1500407 CrossRef .
  69. H. Stoyanov, P. Brochu, X. Niu, C. Lai, S. Yun and Q. Pei, Long lifetime, fault-tolerant freestanding actuators based on a silicone dielectric elastomer and self-clearing carbon nanotube compliant electrodes, RSC Adv., 2013, 3(7), 2272–2278 RSC .
  70. M. J. O'connell, S. M. Bachilo, C. B. Huffman, V. C. Moore, M. S. Strano and E. H. Haroz, et al., Band gap fluorescence from individual single-walled carbon nanotubes, Science, 2002, 297(5581), 593–596 CrossRef PubMed .
  71. E. Bekyarova, M. E. Itkis, N. Cabrera, B. Zhao, A. Yu and J. Gao, et al., Electronic properties of single-walled carbon nanotube networks, J. Am. Chem. Soc., 2005, 127(16), 5990–5995 CrossRef CAS PubMed .
  72. ASTM D1003-21, Standard Test Method for Haze and Luminous Transmittance of Transparent Plastics, ASTM International, 2021 Search PubMed .
  73. A. Gent, A new constitutive relation for rubber, Rubber Chem. Technol., 1996, 69(1), 59–61 CrossRef CAS .
  74. S. J. A. Koh, C. Keplinger, T. Li, S. Bauer and Z. Suo, Dielectric elastomer generators: How much energy can be converted?, IEEE/ASME Trans. Mechatron., 2011, 16(1), 33–41 Search PubMed .
  75. S. J. A. Koh, X. Zhao and Z. Suo, Maximal energy that can be converted by a dielectric elastomer generator, Appl. Phys. Lett., 2009, 94(26), 262902 CrossRef .
  76. F. Xin and T. J. Lu, True method to analyze electromechanical stability of dielectric elastomers, EPL, 2016, 114(2), 26002 CrossRef .
  77. A. Hakimian, M. Mohebinia, M. Nazari, A. Davoodabadi, S. Nazifi and Z. Huang, et al., Freezing of few nanometers water droplets, Nat. Commun., 2021, 12(1), 6973 CrossRef CAS PubMed .

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4sm00596a

This journal is © The Royal Society of Chemistry 2024