Next Article in Journal
Cause Analysis of Coal Mine Gas Accidents in China Based on Association Rules
Previous Article in Journal
Construction of an Accurate Wheat-Grain Model Based on X-ray Tomography and Bonding Parameters by Discrete Element
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Actuation Mechanisms and Applications for Soft Robots: A Comprehensive Review

1
School of Mechanical Engineering, University of Shanghai for Science and Technology, Shanghai 200093, China
2
School of Transportation and Logistics Engineering, Wuhan University of Technology, Wuhan 430070, China
*
Author to whom correspondence should be addressed.
Appl. Sci. 2023, 13(16), 9255; https://doi.org/10.3390/app13169255
Submission received: 25 July 2023 / Revised: 8 August 2023 / Accepted: 11 August 2023 / Published: 15 August 2023
(This article belongs to the Section Robotics and Automation)

Abstract

:
Soft robots, which exhibit distinguishing features in terms of compliance, adaptability, and safety, have been expansively adopted in various niche applications. For soft robots, innovative actuators have been designed based on smart materials enabling the robots to perform flexible and versatile functions, whereas extra spaces and accessories to accommodate motors and power devices have been eliminated to achieve structural optimisation. Herein, different types of actuation mechanisms for soft robots are summarised to reflect the state-of-the-art research and applications. Major characteristics of the actuation mechanisms are updated. Design methodologies of the actuation mechanisms are discussed in detail. Furthermore, their advantages, disadvantages, and application potential are compared and summarised. In the end, based on our knowledge and understanding, new thoughts and recommendations to further develop the actuation mechanisms are put forward. This review is useful to support the conclusion that, through incorporating actuation mechanisms and advanced intelligent technologies, soft robots tend to create disruptive innovations in applications.

1. Introduction

Smart material-enabled soft robots exhibit greater degrees of freedom and produce muscle-like behaviours [1]. Compared with a rigid robot, a soft robot demonstrates the advantages of reducing the complexity of mechatronic design, minimising the requirement for algorithmic programming, and interacting with unknown environments in a flexible means. Soft robots have been extensively applied in broad prospects, such as industrial production, agriculture, biomedical science, disaster resilience, etc. [2,3,4]. For soft robots, some critical parts are made of smart materials, like polymers, rubbers, silicones, or other soft materials [5,6,7,8,9,10]. The materials are used to design actuation mechanisms to drive soft robots to fulfill complex functions and adapt to changes in complex environments intelligently and autonomously [11,12,13,14]. Some typical soft robots and their actuation mechanisms are illustrated in Figure 1, including shape memory alloy (SMA) (Figure 1a,b), fluid (Figure 1c–f), electro-active polymer (EAP) (Figure 1g–i), magnetic (Figure 1j,k) and origami (Figure 1l).
Almost all soft robots are based on the design principles of bionics [14]. By imitating the behaviours of various creatures in nature, such as reptiles, annelid, molluscs, and humans (bionic hands, arms, etc.), different types of soft robots have been designed, like snake-shaped robots [27,28], inchworm robots [29,30], earthworm robots [31,32], jellyfish robots [33,34,35], robotic hands [36,37,38,39,40], auxiliary gloves [41,42,43], etc. Such a bionic soft robot can achieve a single function, such as multi-gait crawling [44,45,46,47], frictional crawling [48,49,50], jumping [51,52], and rolling [53,54], or it can achieve multiple functions, such as two modes of scrolling and crawling [55] and three modes of straight crawling, steering and rolling [56]. The combination of different actuation mechanisms can make a soft robot more adaptable in different terrain environments, such that it could be used to be underwater, on land, or in the low sky.
Some previous reviews for soft robots (e.g., [2,57]) have focused on the motion modes or the applications of soft robots. In some reviews [58,59,60,61,62], soft robots and their actuation mechanisms have been summarised. For instance, in [61], a survey of magnetically actuated bio/soft robots was conducted, and the future prospects of actuation mechanisms used to improve the performance of different types of soft robots have been discussed. In [62], the actuators of soft robots have been divided into three types, i.e., variable length tendon, fluidic actuation, and electro-active polymer (EAP). In the research, the sensing mechanisms, structures, and applications of soft robots have been introduced. Nevertheless, more comprehensive evaluations and discussions on the latest development of soft robots and their actuation mechanisms are highly imperative. For instance, some newly emerging types of actuations, like light [63,64] and chemical [65], have not been discussed yet.
To bridge the aforementioned knowledge gap, this paper is aimed to update the latest progress of actuation mechanisms and applications for soft robots. The objectives include:
(i)
The actuation materials and mechanisms of soft robots are summarised according to seven categories, i.e., SMA, fluid, EAP, electric and magnetic, light, sound, and chemical;
(ii)
Furthermore, according to the progress of soft robots in recent years, extra categories, including SMA-based wire, spring design, origami-based pneumatic, and dielectric elastomer actuator (part of EAP), are also summed up;
(iii)
The actuations of soft robotics under the above categories are reviewed in terms of characteristics, mechanisms, and typical structures. The advantages and disadvantages of the categories are analysed quantitatively and qualitatively;
(iv)
In addition, challenges for the future development of soft robots are put forward to bring more insights to practitioners and researchers.

2. Actuation Materials and Mechanisms for Soft Robots

In the following, actuation mechanisms are summarised according to the aforementioned seven categories. Simultaneously, smart materials, driving mechanisms, applications, and technical difficulties for the actuation mechanisms are also presented.

2.1. Shape Memory Alloy (SMA)

SMA is a specific alloy that, after being deformed, can return to its initial shape when heated [66,67]. As shown in Figure 2, the effects of the shape changes can be categorised into three types: one-way shape memory effect (OWSME), two-way shape memory effect (TWSME), and pseudo-elasticity [68,69]. Under a certain temperature, stress, and strain, SMA can exhibit two different phases, three different crystal structures (detwinned martensite, austenite, and twinned martensite), and six possible transformations [70]. These phenomena create OWSME, TWSME, and pseudo-elasticity, respectively.
Typical SMAs include Ni-Ti-based (nitinol), Fe-based, Cu-based, etc. Ni-Ti-based SMA is the most commonly used shape memory material, owing to its high thermal stability, strain recovery, and corrosion resistance [71,72]. That is, in terms of heat resistance, transformation, and recovery stress, Fe-based and Cu-based SMAs are not as good as Ni-Ti-based SMAs, so they are less used to design actuations of soft robots [73,74,75]. Ni-Ti-based functional components with shape memory effects can be flexibly produced using additive manufacturing processes [58], which supports the design of small, exquisite, highly automated, and reliable soft robots. Therefore, Ni-Ti-based SMA has attracted increasing attention and has been widely applied in biomedical, aerospace, robotics, mechanical, and other fields [76]. SMA-based actuators can be mainly divided into the spring form and the wire form. More technical details are given below.

2.1.1. SMA Spring

The SMA spring actuator employs a spring as the main body of a soft robot to connect its front and rear moving components. When the spring is deformed and expanded, the movable components operate. As shown in Figure 3a, an earthworm-like micro-robot achieves a locomotion action based on the SMA spring actuation mechanism [77]. As shown in Figure 3b, when the robot is energised, the front foot of this robot produces an electro-adhesive force to ensure the robot is fixed on the ground. Then, upon the electric current is applied, the spring is heated to shrink, making the rear foot move forward. Finally, when the rear foot is anchored, the spring stretches and returns to its original shape, causing the front foot forward. In Figure 3c, the high state of the front foot or the rear foot means the robot is moving forward, and the low state means that it is in an anchored state. That is, the high state of the SMA spring means deformation and contraction, and the low state means elongation. In the first stage (from the 2nd to 4th time interval), the SMA spring contracts, causing the rear foot to move towards, whereas in the second stage (from the 4th to 6th time interval), the front foot is pulled by the elongated SMA spring.
Moreover, the robot with three locomotion actions has eight basic modules and a separation [56]. In this research, SMA springs connect each module, and the stiffness of each SMA spring is 2.92 N/mm [56]. When the temperature rises, the springs will deform to make the robot take a locomotion action. A small crawling robot was developed to imitate the movement of Caenorhabditis elegans [78]. Adjacent motion modules were connected by SMA springs, which could contract up to 50% of their extended lengths. The SMA spring actuator can produce a maximum force of approximately 0.3 N to make the robot move like a real worm. Inspired by the muscular organs and modularisation of sea stars and octopuses, Pan et al. designed a bionic spherical robot [79]. Driven by SMA springs, the robotic feet bend up and down to help the robot crawl on the ground and roll down slopes. With the excitation of 3.8 V, the SMA spring generated a driving force of 0.1 N, and the soft robot was driven at a speed of 0.133–0.147 mm/s with a maximum load of 850 g.

2.1.2. SMA Wire

The general design concept of SMA wire driving is to install an SMA wire into the main body of a soft robot. By applying the electric current, the SMA wire deforms, bending the main body and generating sliding frictions. Kim et al. designed a bionic inchworm robot (shown in Figure 4a) [80]. As shown in Figure 4b, when the temperature of the SMA wire exceeds its phase transition temperature, the SMA wire returns to its initial shape. When the temperature is reduced, the wire is under relaxation. In Figure 4c, during the 2nd time interval, the SMA wire starts to deform and arch up (at the high state), directly driving the rear foot to move forwards. At the same time, the front foot moves back. Then, in the 4th time interval, the SMA wire relaxes and recovers to its original shape, the front and rear legs spread out, and the robot advances to a certain distance.
Wang et al. developed a crawling soft robot to mimic the worm movement [81]. The main body of the robot is a thin tetragonal composite plate with two pairs of SMA wires placed on the four sides of the robot. Periodic currents are applied to deform the SMA wire, enabling the robot to turn and crawl. In a single motion cycle, the moving distance of the robot is 54 mm, which is about one-third of its own body length, and the maximum linear speed can reach 3.6 mm/s. Inspired by the movement of caterpillars, Lin et al. designed a rolling soft robot using silicone materials and SMA wires [82]. The minimum weight of the robot is only 5 g, and the SMA wires installed on both sides of the robot can provide a driving force of 10 mN. Under excitation, the SMA wire deforms and makes the robot curl into a wheel shape, and the rolling speed can reach 0.5 m/s.
From the manufacturing perspective, the spring form is more complicated than the wire form. From the structural perspective, the wire form can be flexibly bent to connect two kinematic joints, whereas the spring form needs to be fixed in a specific position and can only be extended. From the perspective of the bearing capacity, the spring form can produce a greater output force, and the spring form-enabled soft robot can bear a higher load. Therefore, in comparison with the spring form, the wire form can be employed to design a versatile soft robot to play the role of a hand, a body, and a leg to grasp, stretch and move objects [15,83,84].
In addition to spring and wire forms, SMA can also be made into other forms. The research in [85] is a bell-shaped jellyfish robot, which body is composed of eight SMA compound actuators. The deformation of SMA is controlled by voltage changes to simulate the swimming motion of a jellyfish. During 14 cyclic movements, the average thrust and power consumption of the jellyfish robot are 0.0039 N and 16.74 W, respectively, and the maximum speed reaches 5.42 cm/s. RoBeetle, designed by Yang et al., is powered by the catalytic combustion of methanol fuel with high specific energy [86]. Its four legs are made of SMA actuators. Methanol vapor flows out of the container and reaches the surface of the SMA legs. The SMA shrinks when the catalytic surface is exposed to fuel and oxygen, and expands when the fuel flow stops. The robot exhibits a maximum average stride of 2.83 mm, and a maximum average speed of 0.76 mm/s. It can carry a maximum payload of 230 mg (about 2.61 times the weight of the prototype).
As a flexible material that can be deformed and used in widespread applications, SMA still has some technical limitations, no matter whether the form is spring, wire, or other. Firstly, in the repeated phase-changing process, SMA exhibits functional fatigue, which will greatly shorten the service life of the SMA actuator and reduce the driving efficiency [87]. Secondly, most SMA actuators have hysteresis. At present, ferromagnetic shape memory alloys (FSMAs) can be rapidly deformed in a magnetic field owing to the magnetic particles contained [88]. However, the required magnetic field is high, which limits its applications.

2.2. Fluid

Fluid can be designed as an actuation mechanism that converts the pressurised energy of fluid into mechanical energy to realise the motion of a soft robot. The basic principle is to change the pressure in the fluid chamber in a soft robot, expand the channel in the robot, and therefore deform the robotic structure controllably. There are three types of fluid-based actuation mechanisms for soft robots, i.e., pneumatic, hydraulic, and pneumatic–hydraulic [89]. Fluid is a commonly used actuator having the advantages of convenient operation and control, large output power, and easy automation.

2.2.1. Pneumatic

A pneumatic actuator is a practical and versatile pneumatic device. It can operate in an antagonistic way like human muscles, and only generate traction when powered by compressed air [90]. Pneumatic utilises air pressure to generate a driving force to actuate a soft robot. It demonstrates the characteristics of quick response speed, good safety, and strong flexibility [91]. In pneumatic robots, gases are pressurised into embedded channels and elastic cavities [60]. This method generates stress at the gas–solid interface, thereby causing the structure of the soft robot to deform. It is usually assumed that different pressure levels of the gases in the internal cavity of the robot propagate fast enough so that all the gases in the structure have a uniform pressure level [92].
The control and movement mechanism of the pneumatic is shown in Figure 5. The soft robot in Figure 5a consists of an air pump, an air chamber, and two feet [93]. Figure 5b shows that the air pump inflates or deflates the air chamber for elongation or contraction. At the same time, the front and rear feet move to realise the movement of the robot alternately. Figure 5c describes that, during the above entire cycle, the air chamber completes elongation and contraction, and the two feet are anchored or moved collaboratively to realise the movement of the robot. In the 2nd time interval, the pump inflates the chamber to elongate, driving the front foot moving forwards. In the 4th time interval, the front foot is anchored, and then the pump deflates, making the chamber contract and pull the rear foot.
In applications, Ning et al. designed pneumatic devices to imitate worms to achieve a Ω-shaped crawling motion [94]. The robot is composed of eight air chambers, which are in a sawtooth shape. When inflated, the main body of the robot swells and arches upward, and then returns to its original shape after deflation. Then, a crawling process is completed. The robot can shrink the length of its body down to 40 mm (the initial length is 98 mm), the maximum moving distance is 2 mm in one cycle (0.4 s), and the average speed reaches 2.2 mm/s. In terms of materials, Terryn et al. utilised the self-healing ability of Diels–Alder (DA) polymers to design soft-handling robots, soft grippers, and artificial muscles [95]. Under the air pressure of 25 kPa, each actuator produces a gripping force of 0.25–0.32 N and the robotic hand can grab objects up to 92.8 g. If the robot hand is damaged when grasping sharp objects, it can be healed automatically at a high temperature (80 °C) through the DA reaction of the material. This characteristic expands the application field of pneumatic robots. In [96,97], some pneumatic actuation-based soft robots for pipeline-related applications were developed. By integrating a pneumatic device that has telescopic characteristics, a soft robot was designed to monitor and dredge pipelines. When it is possible to make pneumatic actuation-based soft robots smaller, the robots can be applied to medical treatments, such as treating vascular diseases. In addition, researchers leveraged the elongation characteristics of pneumatic actuators to mimic the process of biological navigation [98] and the movement of octopus tentacles [99].

2.2.2. Origami-Based Pneumatic

Conventional pneumatic actuation-based soft robots are made of soft materials (e.g., silicone), which expand and elongate themselves after being inflated. If the action of intaking air is continued, such a soft robot will deform excessively, which will affect the robot’s movement. Origami is made of folding papers or polymers. Owing to bidirectional stability, origami is a structural innovation of pneumatic soft robots. Origami itself is not flexible and can not deform, but after being folded, it will contract like a spring to avoid unconstrained deformation.
Hong et al. devised a pneumatic origami muscle actuator (POMA) for a soft robotic hand [39]. The chamber of the robot is an origami structure using a 0.2 mm polyethylene film, which can withstand air pressure of 100 kPa. The chamber’s outer layer is a spandex fabric that can be stretched to 200% of its original length, which helps the inner chamber bend. The overall weight of the soft glove is about 200 g, and it can generate a constant driving force and transmit it in various grasping states. Yu et al. presented a pneumatic crawling soft robot based on Miura-Ori [100]. It consists of two pneumatic foldable actuators that act as legs. Each actuator is made of soft materials and a paper skeleton, the total weight is about 2.52 g, and the load can reach 10 g. The working pressure of the PFA is from −10 kPa to 6 kPa, and the maximum strain under positive and negative pressure is 0.19 and 0.43. By controlling the unfolding and folding sequences of the two PFAs, the linear and turning motions of the crawling robot are realized, and the speeds are about 5 mm/s and 15°/s.
Since the origami robot is actuated by using the crease memory of a paper, the movement of the robot can only depend on the folding mode and cannot be changed. Thus, it has low utilisation. Moreover, repeated folding can also make origami vulnerable.

2.2.3. Hydraulic

A hydraulic actuation makes use of liquids to flow directionally with the excitation of the electric currents. Even though this actuator can generate a greater driving force, it has not been sufficiently developed due to the restrictions of the volume and weight of the hydraulic device. In the robotic field, Ueno et al. developed a hydraulic robot that utilises electrically conjugated fluid for actuator design [101]. It consists of a rubber bellow, a pair of needle-ring electrodes, a shrunk water tank, and a grounding foot. The functional fluid undergoes a directional flow with a high voltage of more than 4 kV and an average power consumption of about 77.5 mWs, causing the expansion of the bellow and the movement of the tilting for the soft robot. Its weight is only 1.9 g, but it exhibits a superior migration speed capability of 5.2 mm/s on a horizontal acrylic surface. In [20], a hydraulic robotic arm was developed with the inspiration of an octopus. The robotic arm can simulate the elongating, shortening, bending, and stretching movement of an octopus. The arm has good stretch properties and can be stretched from 100 mm to 160 mm, a 60% increase in length. The diameter is also reduced by 20% (from 20 mm to 16 mm). Additionally, Lu et al. devised a soft strain and pressure sensor based on a liquid conductive metal, which can be easily integrated into a typical soft pneumatic actuator to realise a sensing function [102].

2.2.4. Pneumatic–Hydraulic

The combined pneumatic–hydraulic actuation mechanism has been mostly used for underwater robots. By taking the changes of the internal and external air pressures, a soft robot can absorb or discharge water to make the robot dive or float up. Utilising the elasticity of the dielectric elastomer and pneumatic–hydraulic combination, Godaba et al. designed a gas–liquid joint-driven soft jellyfish robot [103]. The overall weight of the robot is 261 g, and the load capacity is 241 g, so it is equivalent to its own weight. The robot has two chambers. The internal chamber is made of a dielectric elastomer membrane (actuation pressure is 38.97 kPa). When the air pressure in the internal chamber decreases, the water enters the external chamber and the robot descends. When the air pressure in the internal chamber increases, the water is discharged and the robot rises. Through the descending–rising actions of the robot, specific functions are fulfilled. In 20 s, the robot can generate a maximum buoyancy force of about 0.15 N and rise to 18 cm with an average speed of 9 mm/s.
In comparison with other types of actuators, fluid actuators can generally output larger forces and respond to external factors much faster. Nevertheless, fluid actuators require external devices for control and execution. The external devices are large in size and heavy in weight, so the fluid-driven soft robots are difficult for miniaturisation. At present, fluidic actuators are mainly used in semi-automated soft robots like grippers, crawlers, or underwater robots [104,105,106,107], and most of them are pneumatically driven. A trend is to make fluidic actuators to achieve a centimetre level, with which soft robots can be developed to be fully automated.

2.3. Electroactive Polymer (EAP)

EAP is a specific polymer having the ability to change its sizes and shapes by stimulating an electric field. By converting the elastic potential energy into mechanical energy, EAP can enable soft robots to move. EAP exhibits the advantages of large deformation, high energy density, and a lightweight and compact structure. It is therefore a promising driving material to actuate soft robots [57,59,108]. There are different actuation mechanisms of EAP, including polarisation, molecular phase transition, and mass/ion transport. The polarisation mechanism is driven by dielectric elastomers and piezoelectric polymers. The molecular phase transition mechanism is activated by shape memory polymers and liquid crystal elastomers. The mass/ion transport mechanism is mainly used to activate some specific materials, such as gels and conductive polymers [109]. From the material perspective, EAP can be generally divided into two categories according to their actuation mechanisms, that is field-activated [110] and ionic [111]. Dielectric elastomers, liquid crystal elastomers, electro-strictive polymers, polymer electrets, and ferroelectric polymers belong to the field-activated category. Ionic polymer–metal composites, ionic gels, carbon nanotubes, and conductive polymers belong to the ionic category [112,113]. Both of the two categories show remarkable performance in their respective applications. Below are some EAPs that are commonly used in the development of soft robots.

2.3.1. Dielectric Elastomer

Owing to inherent flexibility, large strain, high efficiency, high energy density, and fast response, dielectric elastomer actuators (DEA) have been widely used in designing soft robots [114,115]. DEA is a kind of motion-generating material, which is similar to human muscle in terms of force, strain (displacement per unit length or area), and driving pressure/density [116,117]. This material mainly includes silicone and acrylic. The former has a faster response speed, whereas the latter can produce a greater deformation [118]. In addition to the material preparation produced by the planar method, DEA can also be realised by 3D printing, and its mechanical properties are adjustable [119].
The development of worm-like soft robots is taken as an example for illustrating DEA. Gu et al. leveraged the characteristics of DEA to develop a wall-climbing soft robot (Figure 6a) [120]. In Figure 6b, the elastomeric body of the robot can be expanded and bent at a high voltage (6 kV). When the voltage is lost, the robotic body will contract to achieve a forward motion. The robot can obtain a fast climbing speed owing to the rapid periodic deformation of the elastomeric body. It can be seen from Figure 6c that, when DEA is at a high state, the rear foot is anchored but the front foot moves forward. The two feet exchange the motion state when DEA is at a low state. In a movement cycle, at the first stage (from 2nd to 4th time interval), the front foot is driven by DEA deformation, and at the second stage (from the 4th to 6th time interval), the rear foot is pulled by the restored DEA.
There are various applications of EDA to actuate soft robots. Examples as illustrated in [121,122] are bionic swimming robots that can conduct underwater exploration. At actuation voltages of above 3 kV, their swimming speed can reach more than 1.5 mm/s. The research in [123] shows a micro robot that mimics the annelid. The robot has a small size (20 mm in diameter, and 45 mm in length) and a light weight of 4.7 g. The maximum moving speed of the robot is 2.5 mm/s, and the load capacity is more than twice its own weight. Multi-legged crawling robots in [124,125] are attractive because of their high load capacity (more than 900 g) and fast speed (about 9 mm/s). The rolling robot presented in [126] possesses the advantages of not only a higher rolling speed (36.27 mm/s on average) but also a larger speed–mass ratio (about 41.22 mm/s·g). Therefore, it has good environmental adaptability. Additionally, the bionic caterpillar robot in [127] was designed with an integrated artificial nervous system.

2.3.2. Liquid Crystal Polymer

Liquid crystal polymers are materials that combine the properties of polymer elastomers and liquid crystals. They exhibit large, rapid, and reversible changes in shape after applying certain external stimuli (an example is illustrated in Figure 7 [128]). Meanwhile, they are between ordered crystalline solids and isotropic liquids. While maintaining a certain degree of molecular order, liquid crystals have similar fluidity as liquids [128]. These external stimuli include temperature, light irradiation, or electronic devices (heating by resistance or direct piezoelectric response) [129,130].
The most important application of liquid crystal elastomers for soft robot design is light-driven robots. Rogoz et al. presented a millimetre-scale snail robot with liquid crystal elastomer materials [131]. The position where the elastomer material deforms and bends changes with the position of laser irradiation. After one cycle of laser irradiation, the elastic body moves to a certain distance at a speed of 0.3 mm/s, and then it will proceed to the next cycle. Therefore, the snail can continue to wriggle. A wave-shaped robot, which was reported in [132], curls and crawls like a caterpillar under the light. After the robot absorbs the green light, thermal actuation is triggered and the length shrinks to 20% of the original length, resulting in a total tensile stress of 2.5 kPa. The uniqueness of the robot is its small size (14.8 mm in length) and it can pass through a gap of 0.9 mm. In addition, with a CW power of 2.5 W and 0.4 Hz scanning beam frequency, the robot walks for a distance of 14.2 mm at an average speed of 0.24 mm/s within 59 s. A transporter robot presented in [133] fuelled by light consists of four light-driven liquid crystalline polymer films that play the roles of legs. The robot has an overall length of 2 cm and a weight of 20 mg. In one motion cycle, it can move forward by 4 mm at a speed of 0.5 mm/s. It is a fully light-responsive transporter robot with controlled multi-freedom locomotion and the capability to deliver loads (about 50% of its weight). Wei et al. designed a bionic hand by assembling a photonic film in a liquid crystal elastomer material. It can trigger finger bending and deformation under visible light [134]. It can be imagined that a soft robot that can move quickly and continuously through light-triggered photonic actuators and liquid crystal elastomer photonic films.

2.3.3. Ionic Polymer–Metal Composite (IPMC)

The IPMC material [135] consists of a fluoropolymer ion exchange membrane, which is sandwiched between two conductive precious metal layers (such as platinum, gold, etc.). The mechanism of the IMPC material is illustrated in Figure 8a. The two main characteristics of IPMC are low activation voltage and large bending strain caused by the movement of cations in the polymer matrix. IPMC demonstrates strong performance in a variety of sensing and actuating applications [136]. IPMC actuators have various advantages, such as low driving voltage, relatively fast response, large displacement, the ability to be activated in water or wet conditions, and easy miniaturisation [137].
Carrico et al. printed a crawling robot with IPMC [138]. By applying voltage (<5 V), an IPMC-based actuator stretches and contracts, driving the robot to move along the pipe like a caterpillar or an inchworm (see Figure 8b). IPMC-based actuators have potential applications in the field of underwater robots. The research in [139] presented a fish-shaped robot that can imitate fish swimming in an S shape, and control the speed by the voltage frequency. When the input voltage frequency is 4.0 Hz, the speed is the fastest to reach 7 mm/s. The research in [140] put forward a bionic jellyfish robot, and the curve shape of the IPMC-based actuator can be successfully applied to generate smooth curve motions through heat treatment. When the excitation frequency is 1 Hz and 2 Hz, the vertical floating displacement of the robot is about 0.15 mm and 0.3 mm, and the floating speed is 0.057 mm/s and 0.045 mm/s. Li et al. designed a small capsule underwater robot [141]. The pectoral fin and caudal fin are made of IPMC. Under the condition of a symmetrical square wave with a driving voltage of 3 V and a frequency of 1.5 Hz, various swing combinations of the fins can be realised to drive the movement of the robot. The robot can swim forward at a speed of 4.83 mm/s and turn around at a speed of 4.5–5°/s.

2.3.4. Ionic Gel

Ionic gel is a kind of reticular polymer that can expand in water and is affected by the pH and electric field after being put into aqueous solutions [142] (examples are illustrated in Figure 9). Through the interaction between the polymer and water, the shape and volume of ionic gel are able to change, thereby obtaining the functions of actuators or sensors. According to the formed gel network, gels are divided into two types: chemical gels and physical gels. While in chemical gels, the cross-linked gel network is formed by irreversible covalent bonds, resulting in gels with high elastic modulus and low strain tolerance. In physical gels, cross-linked networks are reversibly formed through non-covalent interactions, such as hydrogen bonding, π stacking, and electrostatic interactions. Since the non-covalent bonds of physical gels can be destroyed by the application of external heat and reformed upon cooling, the gels exhibit sol–gel transitions. At each temperature, the gels experience transition from a gel state to a sol state. However, the sol state is temporary and will return to the gel state after cooling [143]. A soft tactile sensor composed of ionic gel and elastomer was presented and the relationship between its impedance with current frequency was illustrated by Hara et al. [144]. In future work, they will integrate the pressure sensor for soft robot design.
In addition, there are some other polymers applied to make soft robotics. In reference [145], a tank-shape climbing robot (Tankbot) was designed using non-patterned flat and elastomer adhesive treads that operate as tracks. These palm-sized robots are lightweight (60–150 g), fast (up to 12 cm/s), and relatively energy efficient (efficiency up to 0.18). Driven by the motor in the robot, the elastomer tracks enable the Tankbot to climb over obstacles at 16 mm height. Sun et al. demonstrated a kind of artificial muscle polymer called a twisted coil actuator (TCA) [146]. The actuator can provide opportunities for developing untethered soft robots or robotic hands due to its excellent bending performance and programmable motions. A single TCA weighs 0.3 g and has a maximum tensile force of 0.78 N (stress is 8.59 MPa). It can overcome a peak friction force of 0.4 N and achieve a free stroke of 48%. In [147], a rolling robot (210 g total weight) composed of 6 curved fluidic elastomer actuators was put forward. With the catalysed decomposition of hydrogen peroxide, the internal air pressure increases to 24.1 kPa, causing the elastomer to expand and deform. Under the alternate action of six fluidic elastomer actuators, the robot rolls forward. At each step, the robot rolls forward for approximately 40.2 mm.
In general, an EAP-based soft robot exhibits the advantages of large deformation, high energy density, lightweight and compact structure, etc. However, the robot also has the following technical difficulties:
(1)
It requires a high voltage above kilovolts with high power consumption;
(2)
It has a slow response time, which needs much time to deform;
(3)
Because of the softness of these polymers, their load-bearing capacities are less than other actuations.

2.4. Electric and Magnetic

Electric can convert electrical energy into other forms, such as mechanical and magnetic, leading electrical soft robots to move. Electrical robots possess the ability to move and achieve functions under wireless control. An electric motor can be used as a direct actuator, as determined in [148,149]. After receiving motion signals, the motor drives the grippers or limbs of robots directly, and, ultimately, the robots will climb trees or swim under water.
Moreira et al. developed an inchworm robot capable of two-anchor motion that can crawl on flat and inclined surfaces [150]. It is a three-segment body connected with 3D-printed lever arms and two servo motors. This mobile robot is able to carry supplies for search and rescue missions. The speed of the robot can reach 2.54 cm/s, and it can crawl on the horizontal plane and the slope of 19° with a load of 0.45 kg. Another form is a wall-climbing robot designed based on electrostatic adhesion mechanisms [151]. When a high voltage of 3 kV is applied to the pad electrodes of the robot, each pad can generate a holding force of about 3.2 N on the glass surface, supporting the robot to complete straight-line crawling and turning. The maximum speed is 2 cm/s, and the maximum single turning angle is 30°. By controlling the adhesion force of each foot, the robot can climb similarly to a real gecko. In [152,153], electric motors were used to design electric-motor-driven soft grippers.
A magnetic actuator can penetrate most materials and control the direction/scale of the magnetic field efficiently and accurately. In the magnetic field, the magnetic particles on the robot will deform as the magnetic field changes, such as elongation, contraction, bending, or other forms, driving the robot to move. This actuation has the advantage of being wirelessly controlled by a remote untethered magnetic field without physical contact.
In [154], a modular and untethered robot named Limpet II (weight: 450 g; size: 250 × 250 × 140 mm) was developed. It utilises the hybrid electromagnetic module (EMM) as its core module and a 450 mAh 11.1 V Li-Po battery as power, allowing the capabilities of adhesion and locomotion. The maximum force achieved by the EMM is 1.8 N. With an actuation current of 1500 mA, a single module can lift objects with a mass below 184 g. As a sensor, it also can be used for inspecting and monitoring offshore structures and nearby conditions. Niu et al. introduced a 54 mm long magnet worm-like robot actuated by permanent magnetic patches [155]. The magnets are embedded in the robotic feet, and each adjacent magnet has opposite magnetisation directions. When the speed of the driving magnet reaches 150 mm/s or more, the robot has a stride of about 6 mm and an average speed of 38.8 mm/s. If the robot prepares to move, another permanent magnet will generate a moving magnetic field. Joyee and Pan demonstrated an unconstrained full 3D-printed soft robot with a multi-modal motion capability involving linear crawling and turning locomotions [25]. It is very light (only 0.23 g), but has a carrying capacity ten times greater than its own weight. Affected by the external magnetic field, the robot changes its motion state and delivers medicines to a designated location at a speed of 3.1 mm/s.
Electrical and magnetic actuators have the advantages of quick and accurate reactions, low cost, remote control, and easy maintenance. Nevertheless, the technical difficulty of the actuators is the same as their drivers, motors, or coils. Meanwhile, the actuators are not scalable when compared with SMA and EAP. Furthermore, electrical actuators exhibit limited flexibility and high rigidity while magnetic actuators need an external strong magnetic field.

2.5. Light

Light-driven soft robots are designed based on the light-responsive effect of polymers, such as liquid crystal elastomers, liquid crystal gels, and some ionic polymers, to deform and move their bodies [156,157,158]. This technology, which is to convert light into mechanical energy, is found in possession of some distinct functions, from cleanliness, directional motility, and sensing, to wirelessly controlled interactions between humans and environments [159]. It is foreseeable in the future that the technology will provide novel opportunities for the development of soft robots in space.
Huang et al. fabricated a series of transparent actuators using local surface plasmon resonance semiconductor nanocrystals that respond to infrared light [160]. The actuators have the advantages of wireless and remote control on camouflage soft robots, and significant deformation (curvature up to 0.66 cm−1) under near-infrared light irradiation (200 mW·cm−2). It is realised by preparing three semiconductor nanocrystals as photothermal conversion agents to construct a photo-actuator. Wang et al. demonstrated a smart thin-film composite with perception and motility in the field of light-driven robots [161]. The robot is capable of coordinated actuation, proprio–exteroception, and communication. Under visible light with the irradiation of 100 mW·cm−2, the maximum driving deformations of the light-driven film on tempered glass, printing paper, and sandpaper are 0.49, 0.472, and 0.443. In addition, the temperature and light are controllable by electric signals, driving the robot uninterruptedly. Using the light-driven film, kirigami soft robots with multitasking and information feedback functions, such as walkers, robotic hands, and centipedes, can be designed.
As a new actuation mechanism being developed in recent years, light-driven robots have some aspects for further optimisation, including: (1) this kind of actuator requires a high-intensity laser to make the photosensitive part of the robot heat up and move around, which consumes a lot of electric energy during this period, so the energy conversion efficiency needs to be improved; and (2) due to the light nature and soft material, light-driven robots can be miniaturised, but they can also result in a small output force, small load bearing, and limited application scenarios.

2.6. Sound

A sound-actuated robot senses external sound through the internal voice coil to generate vibrations and performs modular communications for monitoring. This actuator consists of coils, permanent magnets, and circuits. Its working principle is that the coil collects external sound waves and vibrates, cuts the permanent magnet to generate current, and the coil carries the current to drive the corresponding sensor.
Nemitz et al. presented a 160 mm long bionic robot driven by 3D printed voice coil actuators, relying on electromagnetic coils [162]. Powered by a 7.4 V battery, the robot can move 25 body lengths in 12 min. A series of electromagnetic coils are embedded in the robot (just like annelids) to transmit vibration and realise surface movement. It cannot only be used for sound input and output but also to achieve the purpose of communications within functional modules of the robot. This low-cost robot shows great potential in applications, such as pipeline monitoring and search and rescue scenarios.
The issues facing sound actuations are that it is susceptible to external noise, and the stability is low. Moreover, its output driving force is still not high enough, which limits its mobility.

2.7. Chemical

A chemical reaction actuator is to use the mechanisms of chemical combustion or explosion to convert chemical energy into mechanical energy for actuation purposes. These materials are mostly gases with a high-volume energy density, such as alkanes and other hydrocarbons. During the process of reaction, a large amount of gas is generated in such a soft robot. After the expansion of gas, the air pressure rises and ejects the robot to a designated position.
Bartlett et al. designed a soft robot that can be activated by butane combustion to complete unfettered jumping movements [163]. This robot autonomously jumps up to 0.76 m high (six body height) and demonstrates directional jumping of up to 0.15 m (0.5 body length, 20% of jump height) laterally per jump. In addition, Shepherd et al. presented a three-legged robot that burns methane and explodes to jump [164]. It uses electric sparks to ignite methane gas to achieve rapid combustion, releases a large amount of gas (71 kPa), and causes the internal gas network to make a jump action. In each explosion, the robot jumps to a height of 30 cm, and the jumping speed reaches 3.6 m/s. A chemical catalytic decomposition of a pneumatic robot jointly developed by Wehner et al. is controlled by microfluidic logic [165]. It automatically adjusts the fluid flow to catalytically decompose the vehicle’s single propellant fuel supply. The gas (approximately 50 kPa) produced by the decomposition of the fuel expands the fluidic network downstream of the reaction site (expected to expand to 160 times its original volume), thereby causing movements.
There are also other innovative chemical reaction actuation mechanisms applied to soft robots, such as electrolysed water actuators. These types of robots use electrolysed water to decompose hydrogen and oxygen, push the robots to swim forward when exhausting backwards, and change the speed by adjusting their voltage. Nevertheless, the chemical reaction actuation mechanisms exhibit the following limitations:
(1)
Chemical soft robots include reaction devices, which are difficult to miniaturise;
(2)
Some robots driven by explosions are designed for jumping purposes, but the positioning accuracy of the robots is not high.

3. Applications

Compared with traditional rigid robots, soft robots have less complexity of mechanical design or algorithms and more freedom of movement, and thereby they can better adapt to external environments [5,62]. Therefore, soft robots demonstrate more application potential. The early applications of soft robots are mainly on bionics and motion [14]. With more research on deformable materials and increasing uses in soft robot designs, soft robots are increasingly applied to wider applications, such as medicine and rehabilitation, control and exploration, industry and production, and interaction and contact [4].

3.1. Bionic and Motion

Bionic and motion are the basic applications of soft robots. Inspired by observations of animals’ body structures, muscles, and bones, researchers designed bionic soft robots to mimic crawling, jumping, and rolling [14]. Biomimetic objects are generally creatures with simple movement processes that are easy to imitate, such as inchworms, caterpillars, jellyfish, and octopuses. In this way, soft robots can use flexible structures to move in natural environments and quickly adapt or interact with the environments [62].
A type of soft robot can be designed according to the working principle of an inchworm [81]. The inchworm utilises the stretching of abdominal muscles to drive the front and rear legs to cyclically grasp and move to achieve an Ω-shaped continuous crawling motion. The used materials are required to be soft and deformable, and the driving method generally adopts SMA and EAP. When an electric current is applied, the actuating material deforms, mimicking the alternating movement of the front and back legs of the inchworm. SMA wires are longitudinally embedded in a soft polymer to mimic the longitudinal muscle fibers that control the abdominal contraction of the inchworm during movement. Another example is a DEA-driven bionic inchworm robot, which uses reinforcing fibers to suppress the deformation of the dielectric elastomer in other directions [166]. The deformation and bending of the dielectric elastomer can be effectively converted into forward motions. This robot’s average speed can reach 1.1 mm/s, and each motion cycle can advance 3.8 mm.
A type of soft robot can be designed according to the working principle of a caterpillar. This kind of bionic robot simulates the crawling action of caterpillars by driving friction between the moving body and the ground. The materials used are generally polymers, and the actuation methods include fluid, light, and electromagnetic methods. Sheng et al. designed a robot driven by pneumatic actuators. The biological structure and morphological movement of caterpillars are simulated by the synergistic movement of the deformation of the bellows body and the friction feet. The robot can move at a speed of 1.05 cm/s (0.16 body length per second) [167]. A light-driven bionic caterpillar robot realises rubbing motion by light-induced deformation of heated liquid crystal elastomers [23]. This robot can adapt to various unstructured environments, and its movement speed can reach 0.25 mm/s.
A type of soft robot can be designed according to the working principle of a jellyfish. A jellyfish swims by contracting and expanding epidermal muscle fibers as well as the thrust of water currents. Researchers used elastic polymers to mimic the muscle fibers of jellyfish. By applying an electric current or changing the internal pressure, the polymer is deformed to simulate the swimming of the jellyfish. Christianson et al. used DEA materials to imitate the epidermal muscle fibers of jellyfish, and employed fluid electrodes to apply voltage and deform them, so the average swimming speed of such a designed robot could reach 3.2 mm/s, and it could work underwater silently [35].
A type of soft robot can be designed according to the working principle of an octopus. The arms of octopuses can achieve some fine and complex movements by changing the shape of the arms by shortening, elongating, bending, twisting, and deforming the bodies of octopuses [14]. It can change the shape of the bionic octopus robot and grasp objects by using elastic polymers to simulate the muscles of octopus brachiopods and drive them with fluids or electric currents. For instance, a robotic arm was designed to mimic octopus brachiopods with embedded cables that replicate the function of octopus brachiopod muscles for propulsion and object grasping [168].

3.2. Medical and Rehabilitation

The materials used in medical soft robots can achieve good biocompatibility with human bodies and tissues. In medical and rehabilitation, soft robots can be used in minimally invasive surgery, drug delivery, rehabilitation training, and wearable devices [2]. According to application scenarios, medical robots can be divided into two categories: implantable and wearable [13]. Nevertheless, there exist guarantee conditions, such that the human body will not produce allergic, immune, or rejection reactions to the soft robot, and the robot will not cause secondary damage to human tissues while performing its functions. Implantable soft robots can be used in minimally invasive surgery and drug delivery, and the driving method used is often magnetically driven. For instance, a minimally invasive surgical robot, which is with magnetic-driven steering and navigation capabilities, can be reduced to less than a few hundred microns in diameter and can travel through blood vessels [169]. Another example is a drug delivery robot, which can move and turn in the intestinal model under the action of a magnetic field to demonstrate the virtual drug delivery process. Its average speed can reach 3.1 mm/s [25]. Wearable robots can help our limbs perform rehabilitation training or directly act as prosthetic limbs, which require greater output forces and are suitable for fluid-driven methods. In oral rehabilitation, a pneumatically driven robot can achieve an elongation rate of 800–1000% after driving, which is suitable for sensitive areas of the human body and assists in the treatment of patients with mandibular movement disorders [170]. In hand rehabilitation, a low-cost pneumatic hand rehabilitation robot helps fingers perform flexion/extension training with a maximum output force of 110 kPa [38]. In ankle–foot rehabilitation, a wearable robot driven by pneumatic artificial muscle actuators helps humans with orthopedic ankle and foot or assist walking [171].

3.3. Control and Exploration

On the basis of motion, more soft robots will be applied to the motion control of robots and the exploration of unknown environments to perform tasks that rigid robots are unable to accomplish [8]. By changing the direction, speed, and shape of their motion, soft robots can quickly adapt to and explore their environments. They will either be applied to flexible production lines to participate in industrial production or be arranged in unreachable environments as detectors for scientific research. For instance, a pneumatic robotic hand can change the opening and closing of the gripper and the intensity of the gripping force to grip objects [39]. A small bionic robot can disguise itself as a caterpillar, an inchworm, and so on, through the change of shape to pass through a narrow space [172]. A crawling robot with four legs changes the posture of the four legs to realise dynamic turning [173]. Underwater robots change the pressure inside and outside the body to dive into the bottom of the water to explore [107]. Jumping robots use explosion and combustion reactions to advance over obstacles [164].

3.4. Industry and Production

In industrial production, rigid robots are prone to be damaged when colliding with objects due to their high hardness. Nevertheless, the material of a soft robot can be in flexible contact with objects, eliminating stress concentration and reducing potential damage [174]. In addition, soft robots exhibit good expandability and can completely wrap and cover contacted objects. The requirement for the output forces of soft robots by industrial applications is greater, and the actuators to meet the requirement are fluid and electric, and the shape is also mainly based on robotic carriers and robotic hands [175]. For instance, a quadruped soft robot, which uses deep reinforcement learning to train the walking gait of the robot, was designed as a transportation tool in the industry [176]. Another example is a flexible soft robotic gripper, which establishes an extreme learning machine by identifying and grasping shapes [177]. It can help improve the control accuracy in industrial production.

3.5. Human–Robot Collaboration

Soft robots can interact with human operators for collaboration by changing their shapes. A soft-growing pneumatic navigation robot designed by Greer et al. is capable of interacting and turning during human–robot collaboration when encountering obstacles, and finally planning a navigation path [178]. In terms of human interactions, due to the soft and easily deformable materials used, soft robots can safely interact with humans, so humans suffer less damage and load [179]. Soft robots can cooperate with humans as assistive robots (also known as cobots), such as a ten-fingered robotic hand [83]. The robot integrated the embedded SMA and a piezoelectric transducer (PZT) flexure sensor, enabling real-time interaction with human operators for precise control of working objects.

4. Further Discussions

In this paper, the movement mechanisms, characteristics, applications, and technical limitations of actuators for soft robots are discussed. In this session, based on some indicators, these actuators are further discussed to summarise their advantages, disadvantages, and application potentials (to facilitate readability, the advantages, disadvantages, and application potentials of each actuation mechanism are enumerated in Table 1, and key parameters of typical soft robots are shown in Table 2). Meanwhile, each actuation mechanism is evaluated quantitatively in terms of the output force/carrying capacity, miniaturisation, lightweight, efficiency, response speed, and technology maturity (see Figure 10).
In Figure 10, output force/carrying capacity indicates how much driving force a soft robot can generate and how much load it can bear; miniaturisation means how to reduce the size of a soft robot as much as possible; lightweight focuses on whether the robotic weight can be decreased; efficiency reflects the energy conversion rate of a soft robot; response speed shows the maneuverability and sensitivity of a soft robot; and technology maturity describes the current application status of the actuation and whether the technology is mature. For the above indicators, the larger the values, the better the performance of a soft robot. These six indicators interfere with each other, and jointly affect the soft robot’s movement mechanism and function realisation.
SMA, fluid, and EAP are the most widely used actuators in various applications (see Table 3). Compared with SMA and EAP, Fluid has the advantages of a larger output force, stronger load, higher efficiency, and faster response speed. Nevertheless, the above actuators are difficult to realise miniaturisation and lightweight design.
Other actuators are recently developed, but they are making significant progress. Part of them, electric and magnetic, have obvious advantages in miniaturisation and wireless control [61]. In recent years, many soft robots actuated by magnetic are designed, like C-balls, connected and driven via the magnet force [180], thermo-magnetically actuated hybrid soft millirobots based on 3D printing technology [181]. The advantage of miniaturisation and wireless control shows the promising application prospects of magnetic actuation.
For light actuation, the biggest advantage is that its mass and volume can be made smaller. Light-driven robots are within 20 mm in length and weigh around 20 mg. Nevertheless, they need a high-intensity laser as a heat source, which consumes high energy, and the material is soft. The output force, carrying capacity, and deformability are also small. Sound actuation is at a medium average level, and the advantages are not obvious. Nevertheless, in aspects of miniaturisation and lightweight, it has a bit of an advantage. Chemical reaction actuation can provide a great instantaneous force by explosion or combustion, and it also has a good performance in response speed and efficiency.
From the perspective of the output force and carrying capacity, the actuations of fluids and chemicals have obvious advantages. They are able to carry loads more than ten times their own weights and have a fast move speed (see Table 2). In terms of miniaturisation, soft robots actuated by EAP and light like in [23,126,133] have a relatively small size (the length can reach 20 mm).
Response speed and efficiency are the most important factors of soft robots that researchers are concerned about. Because hydrocarbons react violently and release heat quickly during the combustion process, chemical actuation can obtain energy efficiently. For instance, the response speed for such a robot is able to reach 3.6 m/s [164], faster than other actuations. Moreover, the efficiency is also adjustable so that it can also arrive at a middle-to-high level. SMA, fluids, and magnetic require less energy, but can also reach about 5 mm/s in terms of response speed [53,82,95,182]. For EAPs, the voltage required of electrical polymers is relatively high, especially for dielectric elastomers, which could require a high voltage level of more than 3 kV [125,126]. Ionic polymers do not require high-voltage excitation, but the response speed is much lower than that of electrical polymers.
For efficiency, fluids-actuated robots can make full use of air pressure to bear a load ten times greater than their own weights. Simultaneously, the robots have relatively fast response speeds, so the efficiency is higher than other types of actuations. Robots actuated by SMA, electric and magnetic, and sound and chemical, have low energy consumption, so they can be put in the second echelon in the rank of efficiency. Finally, the actuation of EAP needs a high excitation voltage consuming much energy and light-driven has low output power, so the efficiencies of EAP and light are lower than that of others.
Soft robots have more flexible degrees of freedom and continuous deformation capabilities, which enable them to reach special locations untouchable by humans. At the same time, soft robots also demonstrate good environmental adaptability. They can operate in extreme temperatures, high pressure, thin air, heavy pollution, strong radiation, etc. The continuous discovery and application of new flexible materials [183,184], as well as the combination of various sensors [185,186] and actuations, have greatly enriched the types of soft robots. Materials with a softer texture, lighter weight, and faster deformation play an increasingly important role in the shape, volume, function, speed, and aesthetics of the robot. Low-latency sensing technology and sensitive driving speed are beneficial for improving the rapid response of soft robots.
Table 1. Comparison of various actuation mechanisms.
Table 1. Comparison of various actuation mechanisms.
ActuationsAdvantagesDisadvantagesApplications
Shape Memory Alloys (SMAs) [187,188,189,190]High thermal stability.
Recoverable strain.
Strong corrosion resistance.
Easy to automate.
Able to be additively manufactured.
Slow response.
High energy consumption and low efficiency.
Affected by temperature.
Aerospace, automotive, biomedicine, and robotics.
Fluids Grabbing devices such as robotic hands and robotic arms.
Swimming robot.
Pneumatic [191,192,193,194] Fast response.
High security.
Strong flexibility.
Need external air pumps.
Difficult to miniaturize.
Origami-based Pneumatic [100]Low cost.
Light weight.
Quick response.
Able to be programmable.
Easy to be vulnerable.
Low reuse rate.
Hydraulic [107,195]Good driving force.Large weight.
Complex structural elements.
Inconvenient to maintain.
Pneumatic-hydraulic [103] Both pneumatic and hydraulic advantages.Difficult to miniaturize.
Electroactive polymers (EAPs) Robot gripper.
Creeping, underwater, aerial robots.
Soft actuators such as artificial muscles.
Electronic [196,197,198] High energy density.
Quick response.
Long operating time.
Able to perceive itself.
High drive voltage.
Prone to aging and failure.
Ionic [10,199]Low driving voltage.
Large bending displacement.
Response time is longer than electronic.
Small output force.
Electric & Magnetic [200,201,202]Automatic control.
Fast response.
Adapt to enclosed areas.
Large external coil that generates the magnetic field.
High power consumption.
Small controllable area.
Drug delivery and surgery.
Robot crawling device, swimming device, and micro-pump.
Light [128,133]Able to be controlled wirelessly and remotely.
Easy to miniaturize.
Have access to human–computer interaction.
Need high-intensity light.
Low light-to-heat conversion efficiency of actuation material.
Biomedical applications, multi-modal motion of underwater robots
Sound [203]Transfer energy through vibration.
In-module communication.
Vulnerable to external unrequired noise.
Limited mobility.
Sound monitoring, communication, and search and rescue.
Chemical [164,204]High energy density.
Quick response.
Untethered.
Greater obstacle avoidance capability.
Need to be refueled regularly.
High requirements for impact resistance of materials.
Robotic movement and directional jumping.
Table 2. Summary of the characteristics of the different soft robots.
Table 2. Summary of the characteristics of the different soft robots.
ActuationSpeed
(m/s)
Source of EnergyWeight
(g)
Carrying Capacity (g)Output ForceMax Displacement per StepDimensionRefs.
SMARoll robotSMA0.13<10 V DC191.2--50 mm100 × 50 × 8 mm3[53]
GoQBotSMA0.5<10 V DC5–6.2-0.05 N25 cm100 mm (Length)[82]
Inchworm robotFluid0.0022Air pressure-500.5 N2 mm98 × 25 × 17 mm3[94]
Tube climberFluid0.006Air pressure98138113.8 N2.4 cm70–90 mm (Length)[96]
Jellyfish robotFluid and EAP0.009Gas–liquid pressure difference2612410.15 N18 cm140 mm (Height)
59 mm (Radius)
[103]
Annelid robotEAP0.00533 kV DC10.3-2 MPa9 mm170 mm (Length)[125]
RSREAP0.036273.2 kV DC0.88--145.09 mm24.83 mm (Radius)[126]
Multi-material robotMagnetic0.0031Magnetic field0.233-5.5 mm40 × 5 × 2 mm3[25]
C-BallsMagnetic0.046Magnetic field30----[180]
Helical robotMagnetic0.01Magnetic field0.04--19.45 cm30 mm (Length)
4.5 mm (Diameter)
[182]
Inching robotLight0.00025200 mW·cm−2 (Light intensity)---2.7 mm17 mm (Length)[23]
Transporter robotLight0.0005150–250 mW·cm−2 (Light intensity)0.020.0050.05 N4 mm20 mm (Span)[133]
WormbotSound0.056300 mAh
7.4 V (Battery)
---6 mm160 mm (Length)
27 mm (Radius)
[162]
Tripedal robotChemical3.6Methane--71 kPa30 cm130 mm (Length)
10 mm (Height)
[164]
Jumping robotChemical0.43 (Horizontal)
0.86 (Vertical)
Butane510-138 kPa0.6 m150 mm (Radius)
80 mm (Height)
[52]
Table 3. Chronology of SMAs, fluids, and EAP actuation.
Table 3. Chronology of SMAs, fluids, and EAP actuation.
Years SMAsFluidsEAPs
2006Applsci 13 09255 i001Earthworm robot [77] Swimming robot [139]
2007Jellyfish microrobot [205] Annelid robot [123]
2008Six-legged robot [206]Robotic eye [207]
2009Omegabot [29]Miniature soft hand [208]Jellyfish robot [140]
2010GoQBot [82] Tankbot [145]
2011∩-shaped robot [80]Multigait soft robot [209]
2012 Grasping hand [210]
2013Meshworm [211]Snake robot [27]Folding robot [212]
2014Starfish robot [213]Micro inchworm robot [101]Quadruped robot [124]
2015Robotic hand [83]Oral robot [170]Aligned fibers robot [166]
2016Soft morphing hand [214]Tube-climbing robot [96]Versatile soft grippers [196]
2017Curved gripper [215]Self-healing robot [95]Electronic fish [197]
2018Gecko-inspired gripper [216]Caterpillar robot [217]Feedback control robot [218]
2019Spherical robot [79]Soft robotic manipulator [219]Millimetre-scale snail robot [131]
2020RoBeetle [86]Balloon-type robot [99]Capsule underwater robot [141]

5. Conclusions and Outlook

Actuators play crucial roles in the structure design, motion control, and function execution of soft robots in applications. In this paper, the characteristics, mechanisms, and typical structures of soft robots with various actuators are discussed. Furthermore, the advantages and disadvantages of the actuators are investigated in detail.
In the future, more research on soft robots can be carried out in the fields of miniaturisation, intelligence, automation control, and multi-function design. Miniaturisation means designing soft robots with lower density and more flexible materials, reducing the weight and volume of soft robots as much as possible. Research is undergoing to design centimetre-scale, millimetre-scale, or even micrometre-scale soft robots that allow them to fit into cramped environments and reach places out of reach of humans. Intelligence means that, with the applications of integrated circuits and artificial intelligence technologies, soft robots can have exquisite structures, precise positioning, and advanced functions. Automation involves programming, sensor applications, and data processing. When instructed, soft robots can be autonomously controlled to accomplish complex operations, so that some risks and uncertainties caused by human participation can be effectively avoided. At present, the functions of soft robots are relatively single, so they can only realise relatively simple walking, grasping, and communication. In future, multi-functions of soft robots are imperative to be applied in practical situations.
In particular, the design of actuation mechanisms for soft robots has some challenges to be addressed in future works. The first challenge is the response speed and efficiency. After a soft robot is stimulated, there will be a delay time, so that the response lags are unable to provide timely feedback. The second challenge is material choice and design. Soft robots need more flexible materials to adapt to more deformation requirements and faster response speed. At the same time, they need to have a certain load capacity to conduct complex tasks and operations. Moreover, material properties constrain the design of complex modelling and control strategies, making it difficult to control precise positions and execute tasks according to geometrically exact models. Therefore, the development of soft robots is still in the laboratory stage.
Taking these factors into consideration, it is unable to design soft robots to perform some complex tasks in small areas. Nevertheless, as new materials and technologies are under continuous development and applications, it is expected to break through these difficulties by designing more advanced soft robots that are smaller in structure, more precise in movement, more flexible in control, and eco-friendly in operation. Together with the aforementioned features, new actuation mechanisms will be essential for soft robots to support complex application requirements, such as medical treatment, manufacturing, surveying, maintenance, information collection, human–computer interaction, electromechanical control, and wider fields.

Funding

This research was sponsored by the National Natural Science Foundation of China (Project No. 51975444; Project No. 52105396), the Ministry of Science and Technology of China (Project No. G2022013009), the Science and Technology Commission of Shanghai Municipality (Project No. 23010503700), and the Fundamental Research Funds for the Central Universities (Project No. 2021IVA053; Project No. 2021III027JC).

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is available upon request.

Conflicts of Interest

The authors declar no conflict of interest.

References

  1. Rus, D.; Tolley, M.T. Design, fabrication and control of soft robots. Nature 2015, 521, 467–475. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Cianchetti, M.; Laschi, C.; Menciassi, A.; Dario, P. Biomedical applications of soft robotics. Nat. Rev. Mater. 2018, 3, 143–153. [Google Scholar] [CrossRef]
  3. Hann, S.Y.; Cui, H.; Nowicki, M.; Zhang, L.G. 4D printing soft robotics for biomedical applications. Addit. Manuf. 2020, 36, 101567. [Google Scholar] [CrossRef]
  4. Zhang, Y.; Lu, M. A review of recent advancements in soft and flexible robots for medical applications. Int. J. Med. Robot. Comput. Assist. Surg. 2020, 16, e2096. [Google Scholar] [CrossRef] [PubMed]
  5. Whitesides, G.M. Soft Robotics. Angew. Chem. Int. Ed. 2018, 57, 4258–4273. [Google Scholar] [CrossRef]
  6. Wallin, T.J.; Pikul, J.; Shepherd, R.F. 3D printing of soft robotic systems. Nat. Rev. Mater. 2018, 3, 84–100. [Google Scholar] [CrossRef]
  7. Zolfagharian, A.; Kaynak, A.; Kouzani, A. Closed-loop 4D-printed soft robots. Mater. Des. 2020, 188, 108411. [Google Scholar] [CrossRef]
  8. Elango, N.; Faudzi, A.A.M. A review article: Investigations on soft materials for soft robot manipulations. Int. J. Adv. Manuf. Technol. 2015, 80, 1027–1037. [Google Scholar] [CrossRef] [Green Version]
  9. Hines, L.; Petersen, K.H.; Lum, G.Z.; Sitti, M. Soft Actuators for Small-Scale Robotics. Adv. Mater. 2017, 29, 1603483. [Google Scholar] [CrossRef]
  10. Du, X.; Cui, H.; Xu, T.; Huang, C.; Wang, Y.; Zhao, Q.; Xu, Y.; Wu, X. Reconfiguration, Camouflage, and Color-Shifting for Bioinspired Adaptive Hydrogel-Based Millirobots. Adv. Funct. Mater. 2020, 30, 1909202. [Google Scholar] [CrossRef]
  11. Marchese, A.D.; Katzschmann, R.; Rus, D.L. A Recipe for Soft Fluidic Elastomer Robots. Soft Robot. 2015, 2, 7–25. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Rich, S.I.; Wood, R.J.; Majidi, C. Untethered soft robotics. Nat. Electron. 2018, 1, 102–112. [Google Scholar] [CrossRef]
  13. El-Atab, N.; Mishra, R.B.; Al-Modaf, F.; Joharji, L.; Alsharif, A.A.; Alamoudi, H.; Diaz, M.; Qaiser, N.; Hussain, M.M. Soft Actuators for Soft Robotic Applications: A Review. Adv. Intell. Syst. 2020, 2, 2000128. [Google Scholar] [CrossRef]
  14. Kim, S.; Laschi, C.; Trimmer, B. Soft robotics: A bioinspired evolution in robotics. Trends Biotechnol. 2013, 31, 287–294. [Google Scholar] [CrossRef] [PubMed]
  15. Huang, X.; Kumar, K.; Jawed, M.K.; Nasab, A.M.; Ye, Z.; Shan, W.; Majidi, C. Highly Dynamic Shape Memory Alloy Actuator for Fast Moving Soft Robots. Adv. Mater. Technol. 2019, 4, 1800540. [Google Scholar] [CrossRef]
  16. Wang, W.; Rodrigue, H.; Kim, H.-I.; Han, M.-W.; Ahn, S.-H. Soft composite hinge actuator and application to compliant robotic gripper. Compos. Part B Eng. 2016, 98, 397–405. [Google Scholar] [CrossRef]
  17. Olson, G.; Hatton, R.L.; Adams, J.A.; Mengüç, Y. An Euler–Bernoulli beam model for soft robot arms bent through self-stress and external loads. Int. J. Solids Struct. 2020, 207, 113–131. [Google Scholar] [CrossRef]
  18. Chi, Y.; Tang, Y.; Liu, H.; Yin, J. Leveraging Monostable and Bistable Pre-Curved Bilayer Actuators for High-Performance Multitask Soft Robots. Adv. Mater. Technol. 2020, 5, 2000370. [Google Scholar] [CrossRef]
  19. Reinhart, R.F.; Steil, J.J. Hybrid Mechanical and Data-driven Modeling Improves Inverse Kinematic Control of a Soft Robot. Procedia Technol. 2016, 26, 12–19. [Google Scholar] [CrossRef]
  20. Cianchetti, M.; Arienti, A.; Follador, M.; Mazzolai, B.; Dario, P.; Laschi, C. Design concept and validation of a robotic arm inspired by the octopus. Mater. Sci. Eng. C 2011, 31, 1230–1239. [Google Scholar] [CrossRef]
  21. Sun, W.; Li, B.; Zhang, F.; Fang, C.; Lu, Y.; Gao, X.; Cao, C.; Chen, G.; Zhang, C.; Wang, Z.L. TENG-Bot: Triboelectric nanogenerator powered soft robot made of uni-directional dielectric elastomer. Nano Energy 2021, 85, 106012. [Google Scholar] [CrossRef]
  22. Tang, X.; Li, K.; Liu, Y.; Zhou, D.; Zhao, J. A soft crawling robot driven by single twisted and coiled actuator. Sens. Actuators A Phys. 2019, 291, 80–86. [Google Scholar] [CrossRef]
  23. Zeng, H.; Wani, O.M.; Wasylczyk, P.; Priimagi, A. Light-Driven, Caterpillar-Inspired Miniature Inching Robot. Macromol. Rapid Commun. 2018, 39, 1700224. [Google Scholar] [CrossRef]
  24. Lee, K.-M.; Kim, Y.; Paik, J.K.; Shin, B. Clawed Miniature Inchworm Robot Driven by Electromagnetic Oscillatory Actuator. J. Bionic Eng. 2015, 12, 519–526. [Google Scholar] [CrossRef]
  25. Joyee, E.B.; Pan, Y. Additive manufacturing of multi-material soft robot for on-demand drug delivery applications. J. Manuf. Process. 2020, 56, 1178–1184. [Google Scholar] [CrossRef]
  26. Li, J.; Godaba, H.; Zhang, Z.; Foo, C.; Zhu, J. A soft active origami robot. Extreme Mech. Lett. 2018, 24, 30–37. [Google Scholar] [CrossRef]
  27. Onal, C.D.; Rus, D. Autonomous undulatory serpentine locomotion utilizing body dynamics of a fluidic soft robot. Bioinspir. Biomim. 2013, 8, 026003. [Google Scholar] [CrossRef]
  28. Mann, M.P.; Damti, L.; Tirosh, G.; Zarrouk, D. Minimally actuated serial robot. Robotica 2017, 36, 408–426. [Google Scholar] [CrossRef] [Green Version]
  29. Koh, J.-S.; Cho, K.-J. Omegabot: Biomimetic inchworm robot using SMA coil actuator and smart composite microstructures (SCM). In Proceedings of the 2009 IEEE International Conference on Robotics and Biomimetics (ROBIO), Guilin, China, 19–23 December 2009; pp. 1154–1159. [Google Scholar] [CrossRef]
  30. Koh, J.-S.; Cho, K.-J. Omegabot: Crawling robot inspired by Ascotis Selenaria. In Proceedings of the 2010 IEEE International Conference on Robotics and Automation, Anchorage, AK, USA, 3–7 May 2010; pp. 109–114. [Google Scholar] [CrossRef]
  31. Horchler, A.D.; Kandhari, A.; Daltorio, K.A.; Moses, K.C.; Andersen, K.B.; Bunnelle, H.; Kershaw, J.; Tavel, W.H.; Bachmann, R.J.; Chiel, H.J.; et al. Worm-Like Robotic Locomotion with a Compliant Modular Mesh. In Biomimetic and Biohybrid Systems: 4th International Conference, Living Machines 2015, Barcelona, Spain, 28–31 July 2015; Lecture Notes in Computer Science; Wilson, S.P., Verschure, P.F.M.J., Mura, A., Prescott, T.J., Eds.; Springer International Publishing: Cham, Switzerland, 2015; pp. 26–37. [Google Scholar] [CrossRef]
  32. Jiang, F.; Zhang, Z.; Wang, X.; Cheng, G.; Zhang, Z.; Ding, J. Pneumatically Actuated Self-Healing Bionic Crawling Soft Robot. J. Intell. Robot. Syst. 2020, 100, 445–454. [Google Scholar] [CrossRef]
  33. Godaba, H.; Wang, Y.; Cao, J.; Zhu, J. Development of soft robots using dielectric elastomer actuators. In Electroactive Polymer Actuators and Devices; Bar-Cohen, Y., Vidal, F., Eds.; 97981T–97981T-9; SPIE: Bellingham, DC, USA, 2016. [Google Scholar] [CrossRef]
  34. Ren, Z.; Hu, W.; Dong, X.; Sitti, M. Multi-functional soft-bodied jellyfish-like swimming. Nat. Commun. 2019, 10, 2703. [Google Scholar] [CrossRef] [Green Version]
  35. Christianson, C.; Bayag, C.; Li, G.; Jadhav, S.; Giri, A.; Agba, C.; Li, T.; Tolley, M.T. Jellyfish-Inspired Soft Robot Driven by Fluid Electrode Dielectric Organic Robotic Actuators. Front. Robot. AI 2019, 6, 126. [Google Scholar] [CrossRef] [Green Version]
  36. Preechayasomboon, P.; Rombokas, E. Negshell casting: 3D-printed structured and sacrificial cores for soft robot fabrication. PLoS ONE 2020, 15, e0234354. [Google Scholar] [CrossRef]
  37. Alspach, A.; Kim, J.; Yamane, K. Design and fabrication of a soft robotic hand and arm system. In Proceedings of the 2018 IEEE International Conference on Soft Robotics (RoboSoft), Livorno, Italy, 24–28 April 2018; pp. 369–375. [Google Scholar] [CrossRef]
  38. Ariyanto, M.; Setiawan, J.D.; Ismail, R.; Haryanto, I.; Febrina, T.; Saksono, D.R. Design and Characterization of Low-Cost Soft Pneumatic Bending Actuator for Hand Rehabilitation. In Proceedings of the 2018 5th International Conference on Information Technology, Computer, and Electrical Engineering (ICITACEE), Semarang, Indonesia, 27–28 September 2018; pp. 45–50. [Google Scholar] [CrossRef]
  39. Hong, T.H.; Park, S.-H.; Park, J.-H.; Paik, N.-J.; Park, Y.-L. Design of Pneumatic Origami Muscle Actuators (POMAs) for A Soft Robotic Hand Orthosis for Grasping Assistance. In Proceedings of the 2020 3rd IEEE International Conference on Soft Robotics (RoboSoft), New Haven, CT, USA, 15 May–15 July 2020; pp. 627–632. [Google Scholar] [CrossRef]
  40. Shahid, T.; Gouwanda, D.; Nurzaman, S.G.; Gopalai, A.A. Moving toward Soft Robotics: A Decade Review of the Design of Hand Exoskeletons. Biomimetics 2018, 3, 17. [Google Scholar] [CrossRef] [Green Version]
  41. Yap, H.K.; Lim, J.H.; Goh, J.C.H.; Yeow, C.-H. Design of a Soft Robotic Glove for Hand Rehabilitation of Stroke Patients With Clenched Fist Deformity Using Inflatable Plastic Actuators. J. Med. Devices 2016, 10, 044504. [Google Scholar] [CrossRef]
  42. Shintake, J.; Cacucciolo, V.; Floreano, D.; Shea, H. Soft Robotic Grippers. Adv. Mater. 2018, 30, e1707035. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Cappello, L.; Meyer, J.T.; Galloway, K.C.; Peisner, J.D.; Granberry, R.; Wagner, D.A.; Engelhardt, S.; Paganoni, S.; Walsh, C.J. Assisting hand function after spinal cord injury with a fabric-based soft robotic glove. J. Neuroeng. Rehabil. 2018, 15, 59. [Google Scholar] [CrossRef] [PubMed]
  44. Lambrecht, B.; Horchler, A.; Quinn, R. A Small, Insect-Inspired Robot that Runs and Jumps. In Proceedings of the 2005 IEEE International Conference on Robotics and Automation, Barcelona, Spain, 18–22 April 2005; pp. 1240–1245. [Google Scholar] [CrossRef]
  45. Tolley, M.T.; Shepherd, R.F.; Mosadegh, B.; Galloway, K.C.; Wehner, M.; Karpelson, M.; Wood, R.J.; Whitesides, G.M.; Garriga-Casanovas, A.; Collison, I.; et al. A Resilient, Untethered Soft Robot. Soft Robot. 2014, 1, 213–223. [Google Scholar] [CrossRef] [Green Version]
  46. Lu, H.; Zhang, M.; Yang, Y.; Huang, Q.; Fukuda, T.; Wang, Z.; Shen, Y. A bioinspired multilegged soft millirobot that functions in both dry and wet conditions. Nat. Commun. 2018, 9, 3944. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Meng, L.; Kang, R.; Gan, D.; Chen, G.; Chen, L.; Branson, D.T.; Dai, J.S. A Mechanically Intelligent Crawling Robot Driven by Shape Memory Alloy and Compliant Bistable Mechanism. J. Mech. Robot. 2020, 12, 061005. [Google Scholar] [CrossRef]
  48. Qi, Q.; Teng, Y.; Li, X. Design and characteristic study of a pneumatically actuated earthworm-like soft robot. In Proceedings of the 2015 International Conference on Fluid Power and Mechatron. (FPM), Harbin, China, 5–7 August 2015; pp. 435–439. [Google Scholar] [CrossRef]
  49. Hu, W.; Lum, G.Z.; Mastrangeli, M.; Sitti, M. Small-scale soft-bodied robot with multimodal locomotion. Nature 2018, 554, 81–85. [Google Scholar] [CrossRef] [PubMed]
  50. Song, C.-W.; Lee, D.-J.; Lee, S.-Y. Bioinspired segment robot with earthworm-like plane locomotion. J. Bionic Eng. 2016, 13, 292–302. [Google Scholar] [CrossRef]
  51. Jung, G.-P.; Casarez, C.S.; Jung, S.-P.; Fearing, R.S.; Cho, K.-J. An integrated jumping-crawling robot using height-adjustable jumping module. In Proceedings of the 2016 IEEE International Conference on Robotics and Automation (ICRA), Stockholm, Sweden, 16–21 May 2016; pp. 4680–4685. [Google Scholar] [CrossRef]
  52. Tolley, M.T.; Shepherd, R.F.; Karpelson, M.; Bartlett, N.W.; Galloway, K.C.; Wehner, M.; Nunes, R.; Whitesides, G.M.; Wood, R.J. An untethered jumping soft robot. In Proceedings of the 2014 IEEE/RSJ International Conference on Intelligent Robots and Systems, Chicago, IL, USA, 14–18 September 2014; pp. 561–566. [Google Scholar] [CrossRef]
  53. Mansour, N.A.; Jang, T.; Baek, H.; Shin, B.; Ryu, B.; Kim, Y. Compliant closed-chain rolling robot using modular unidirectional SMA actuators. Sens. Actuators A Phys. 2020, 310, 112024. [Google Scholar] [CrossRef]
  54. Kim, K.; Agogino, A.K.; Agogino, A.M.; Feng, R.; Zhang, Y.; Liu, J.; Zhang, Y.; Li, J.; Baoyin, H.; Shah, D.S.; et al. Rolling Locomotion of Cable-Driven Soft Spherical Tensegrity Robots. Soft Robot. 2020, 7, 346–361. [Google Scholar] [CrossRef] [PubMed]
  55. Zhou, J.; Li, X.; Xu, J.; Tian, Y.; Zhao, H.; Liu, X.-J. A Soft Crawling Robot Inspired by Inchworms. In Proceedings of the 2019 IEEE 9th Annual International Conference on CYBER Technology in Automation, Control, and Intelligent Systems (CYBER), Suzhou, China, 29 July–2 August 2019; pp. 209–214. [Google Scholar] [CrossRef]
  56. Du, Y.; Xu, M.; Dong, E.; Zhang, S.; Yang, J. A novel soft robot with three locomotion modes. In Proceedings of the 2011 IEEE International Conference on Robotics and Biomimetics, Karon Beach, Thailand, 7–11 December 2011; pp. 98–103. [Google Scholar] [CrossRef]
  57. Chen, S.; Cao, Y.; Sarparast, M.; Yuan, H.; Dong, L.; Tan, X.; Cao, C. Soft Crawling Robots: Design, Actuation, and Locomotion. Adv. Mater. Technol. 2020, 5, 1900837. [Google Scholar] [CrossRef]
  58. Farber, E.; Zhu, J.-N.; Popovich, A.; Popovich, V. A review of NiTi shape memory alloy as a smart material produced by additive manufacturing. Mater. Today Proc. 2020, 30, 761–767. [Google Scholar] [CrossRef]
  59. Gupta, U.; Qin, L.; Wang, Y.; Godaba, H.; Zhu, J. Soft robots based on dielectric elastomer actuators: A review. Smart Mater. Struct. 2019, 28, 103002. [Google Scholar] [CrossRef]
  60. Polygerinos, P.; Correll, N.; Morin, S.A.; Mosadegh, B.; Onal, C.D.; Petersen, K.; Cianchetti, M.; Tolley, M.T.; Shepherd, R.F. Soft Robotics: Review of Fluid-Driven Intrinsically Soft Devices; Manufacturing, Sensing, Control, and Applications in Human-Robot Interaction: Review of Fluid-Driven Intrinsically Soft Robots. Adv. Eng. Mater. 2017, 19, 1700016. [Google Scholar] [CrossRef]
  61. Ebrahimi, N.; Bi, C.; Cappelleri, D.J.; Ciuti, G.; Conn, A.T.; Faivre, D.; Habibi, N.; Hošovský, A.; Iacovacci, V.; Khalil, I.S.M.; et al. Magnetic Actuation Methods in Bio/Soft Robotics. Adv. Funct. Mater. 2020, 31, 2005137. [Google Scholar] [CrossRef]
  62. Lee, C.; Kim, M.; Kim, Y.J.; Hong, N.; Ryu, S.; Kim, H.J.; Kim, S. Soft robot review. Int. J. Control. Autom. Syst. 2017, 15, 3–15. [Google Scholar] [CrossRef]
  63. Ahn, C.; Liang, X.; Cai, S. Bioinspired Design of Light-Powered Crawling, Squeezing, and Jumping Untethered Soft Robot. Adv. Mater. Technol. 2019, 4, 1900185. [Google Scholar] [CrossRef]
  64. Wang, R.; Han, L.; Wu, C.; Dong, Y.; Zhao, X. Localizable, Identifiable, and Perceptive Untethered Light-Driven Soft Crawling Robot. ACS Appl. Mater. Interfaces 2022, 14, 6138–6147. [Google Scholar] [CrossRef] [PubMed]
  65. Ilievski, F.; Mazzeo, A.D.; Shepherd, R.F.; Chen, X.; Whitesides, G.M. Soft Robotics for Chemists. Angew. Chem. Int. Ed. 2011, 50, 1890–1895. [Google Scholar] [CrossRef] [PubMed]
  66. Suhail, R.; Amato, G.; McCrum, D. Heat-activated prestressing of NiTiNb shape memory alloy wires. Eng. Struct. 2020, 206, 110128. [Google Scholar] [CrossRef]
  67. Kök, M.; Al-Jaf, A.O.A.; Çirak, Z.D.; Qader, I.N.; Özen, E. Effects of heat treatment temperatures on phase transformation, thermodynamical parameters, crystal microstructure, and electrical resistivity of NiTiV shape memory alloy. J. Therm. Anal. Calorim. 2019, 139, 3405–3413. [Google Scholar] [CrossRef]
  68. Cisse, C.; Zaki, W.; Ben Zineb, T. A review of constitutive models and modeling techniques for shape memory alloys. Int. J. Plast. 2016, 76, 244–284. [Google Scholar] [CrossRef]
  69. Kumari, S.; Dinbandhu; Abhishek, K. Study of machinability aspects of shape memory alloys: A critical review. Mater. Today Proc. 2021, 44, 1336–1343. [Google Scholar] [CrossRef]
  70. Mohd-Jani, J.; Leary, M.; Subic, A.; Gibson, M.A. A review of shape memory alloy research, applications and opportunities. Mater. Des. 2014, 56, 1078–1113. [Google Scholar] [CrossRef]
  71. Huang, X.; Ford, M.; Patterson, Z.J.; Zarepoor, M.; Pan, C.; Majidi, C. Shape memory materials for electrically-powered soft machines. J. Mater. Chem. B 2020, 8, 4539–4551. [Google Scholar] [CrossRef]
  72. Karami, M.; Chen, X. Nanomechanics of shape memory alloys. Mater. Today Adv. 2021, 10, 100141. [Google Scholar] [CrossRef]
  73. Wilkes, K.E.; Liaw, P.K. The fatigue behavior of shape-memory alloys. JOM 2000, 52, 45–51. [Google Scholar] [CrossRef]
  74. Huang, W. On the selection of shape memory alloys for actuators. Mater. Des. 2002, 23, 11–19. [Google Scholar] [CrossRef]
  75. Malik, V.; Srivastava, S.; Gupta, S.; Sharma, V.; Vishnoi, M.; Mamatha, T. A novel review on shape memory alloy and their applications in extraterrestrial roving missions. Mater. Today Proc. 2020, 44, 4961–4965. [Google Scholar] [CrossRef]
  76. Clithy, E. Application of Shape Memory Alloy. Sci. Insights 2020, 33, 167–174. [Google Scholar] [CrossRef]
  77. Kim, B.; Lee, M.G.; Lee, Y.P.; Kim, Y.; Lee, G. An earthworm-like micro robot using shape memory alloy actuator. Sens. Actuators A Phys. 2006, 125, 429–437. [Google Scholar] [CrossRef]
  78. Yuk, H.; Kim, D.; Lee, H.; Jo, S.; Shin, J.H. Shape memory alloy-based small crawling robots inspired by C. elegans. Bioinspir. Biomim. 2011, 6, 046002. [Google Scholar] [CrossRef] [PubMed]
  79. Pan, J.; Shi, Z.; Wang, T. Variable-model SMA-driven spherical robot. Sci. China Technol. Sci. 2019, 62, 1401–1411. [Google Scholar] [CrossRef]
  80. Kim, M.-S.; Chu, W.-S.; Lee, J.-H.; Kim, Y.-M.; Ahn, S.-H. Manufacturing of inchworm robot using shape memory alloy (SMA) embedded composite structure. Int. J. Precis. Eng. Manuf. 2011, 12, 565–568. [Google Scholar] [CrossRef]
  81. Wang, W.; Lee, J.-Y.; Rodrigue, H.; Song, S.-H.; Chu, W.-S.; Ahn, S.-H. Locomotion of inchworm-inspired robot made of smart soft composite (SSC). Bioinspir. Biomim. 2014, 9, 046006. [Google Scholar] [CrossRef]
  82. Lin, H.-T.; Leisk, G.G.; Trimmer, B. GoQBot: A caterpillar-inspired soft-bodied rolling robot. Bioinspir. Biomim. 2011, 6, 026007. [Google Scholar] [CrossRef]
  83. She, Y.; Li, C.; Cleary, J.; Su, H.-J. Design and Fabrication of a Soft Robotic Hand With Embedded Actuators and Sensors. J. Mech. Robot. 2015, 7, 021007. [Google Scholar] [CrossRef]
  84. Umedachi, T.; Vikas, V.; Trimmer, B.A. Highly deformable 3-D printed soft robot generating inching and crawling locomotions with variable friction legs. In Proceedings of the 2013 IEEE/RSJ International Conference on Intelligent Robots and Systems, Tokyo, Japan, 3–7 November 2013; pp. 4590–4595. [Google Scholar] [CrossRef]
  85. Villanueva, A.; Smith, C.; Priya, S. A biomimetic robotic jellyfish (Robojelly) actuated by shape memory alloy composite actuators. Bioinspir. Biomim. 2011, 6, 036004. [Google Scholar] [CrossRef]
  86. Yang, X.; Chang, L.; Pérez-Arancibia, N.O. An 88-milligram insect-scale autonomous crawling robot driven by a catalytic artificial muscle. Sci. Robot. 2020, 5, eaba0015. [Google Scholar] [CrossRef]
  87. Rao, A.; Srinivasa, A.R. Experiments on functional fatigue of thermally activated shape memory alloy springs and correlations with driving force intensity. In Proceedings of the Behavior and Mechanics of Multifunctional Materials and Composites 2013, San Diego, CA, USA, 10–14 March 2013; Volume 8689, pp. 198–206. [Google Scholar] [CrossRef]
  88. Jugo, J.; Feuchtwanger, J.; Corres, J. Numerical optimization based control design for a ferromagnetic shape memory alloy actuator. Sens. Actuators A Phys. 2021, 331, 112835. [Google Scholar] [CrossRef]
  89. Ranzani, T.; Russo, S.; Bartlett, N.W.; Wehner, M.; Wood, R.J. Increasing the Dimensionality of Soft Microstructures through Injection-Induced Self-Folding. Adv. Mater. 2018, 30, e1802739. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Belforte, G.; Eula, G.; Ivanov, A.; Sirolli, S. Soft Pneumatic Actuators for Rehabilitation. Actuators 2014, 3, 84–106. [Google Scholar] [CrossRef] [Green Version]
  91. Xu, Q.; Liu, J. Effective enhanced model for a large deformable soft pneumatic actuator. Acta Mech. Sin. 2020, 36, 245–255. [Google Scholar] [CrossRef]
  92. Breitman, P.; Matia, Y.; Gat, A.D. Fluid Mechanics of Pneumatic Soft Robots. Soft Robot. 2021, 8, 519–530. [Google Scholar] [CrossRef]
  93. Ge, J.Z.; Calderon, A.A.; Perez-Arancibia, N.O. An earthworm-inspired soft crawling robot controlled by friction. In Proceedings of the 2017 IEEE International Conference on Robotics and Biomimetics (ROBIO), Macau, China, 5–8 December 2017; pp. 834–841. [Google Scholar] [CrossRef] [Green Version]
  94. Jinqiang, N.; Chaoyang, T.; Yuxiang, L. Inchworm Inspired Pneumatic Soft Robot Based on Friction Hysteresis. J. Robot. Autom. 2017, 1, 54–63. [Google Scholar] [CrossRef]
  95. Terryn, S.; Brancart, J.; Lefeber, D.; Van Assche, G.; Vanderborght, B. Self-healing soft pneumatic robots. Sci. Robot. 2017, 2, eaan4268. [Google Scholar] [CrossRef]
  96. Verma, M.S.; Ainla, A.; Yang, D.; Harburg, D.; Whitesides, G.M.; Ohta, P.; Valle, L.; King, J.; Low, K.; Yi, J.; et al. A Soft Tube-Climbing Robot. Soft Robot. 2018, 5, 133–137. [Google Scholar] [CrossRef]
  97. Zhang, Z.; Wang, X.; Wang, S.; Meng, D.; Liang, B. Design and Modeling of a Parallel-Pipe-Crawling Pneumatic Soft Robot. IEEE Access 2019, 7, 134301–134317. [Google Scholar] [CrossRef]
  98. Hawkes, E.W.; Blumenschein, L.H.; Greer, J.D.; Okamura, A.M. A soft robot that navigates its environment through growth. Sci. Robot. 2017, 2, eaan3028. [Google Scholar] [CrossRef]
  99. Nakajima, T.; Yamaguchi, T.; Wakabayashi, S.; Arie, T.; Akita, S.; Takei, K. Transformable Pneumatic Balloon-Type Soft Robot Using Attachable Shells. Adv. Mater. Technol. 2020, 5, 2000201. [Google Scholar] [CrossRef]
  100. Yu, M.; Yang, W.; Yu, Y.; Cheng, X.; Jiao, Z. A Crawling Soft Robot Driven by Pneumatic Foldable Actuators Based on Miura-Ori. Actuators 2020, 9, 26. [Google Scholar] [CrossRef] [Green Version]
  101. Ueno, S.; Takemura, K.; Yokota, S.; Edamura, K. Micro inchworm robot using electro-conjugate fluid. Sens. Actuators A Phys. 2014, 216, 36–42. [Google Scholar] [CrossRef]
  102. Lu, S.; Chen, D.; Hao, R.; Luo, S.; Wang, M. Design, fabrication and characterization of soft sensors through EGaIn for soft pneumatic actuators. Measurement 2020, 164, 107996. [Google Scholar] [CrossRef]
  103. Godaba, H.; Li, J.; Wang, Y.; Zhu, J. A Soft Jellyfish Robot Driven by a Dielectric Elastomer Actuator. IEEE Robot. Autom. Lett. 2016, 1, 624–631. [Google Scholar] [CrossRef]
  104. Deimel, R.; Brock, O. A novel type of compliant and underactuated robotic hand for dexterous grasping. Int. J. Robot. Res. 2016, 35, 161–185. [Google Scholar] [CrossRef] [Green Version]
  105. Xie, M.; Zhu, M.; Yang, Z.; Okada, S.; Kawamura, S. Flexible self-powered multifunctional sensor for stiffness-tunable soft robotic gripper by multimaterial 3D printing. Nano Energy 2020, 79, 105438. [Google Scholar] [CrossRef]
  106. Jin, G.; Sun, Y.; Geng, J.; Yuan, X.; Chen, T.; Liu, H.; Wang, F.; Sun, L. Bioinspired soft caterpillar robot with ultra-stretchable bionic sensors based on functional liquid metal. Nano Energy 2021, 84, 105896. [Google Scholar] [CrossRef]
  107. Chen, G.; Yang, X.; Zhang, X.; Hu, H. Water hydraulic soft actuators for underwater autonomous robotic systems. Appl. Ocean Res. 2021, 109, 102551. [Google Scholar] [CrossRef]
  108. Duduta, M.; Hajiesmaili, E.; Zhao, H.; Wood, R.J.; Clarke, D.R. Realizing the potential of dielectric elastomer artificial muscles. Proc. Natl. Acad. Sci. USA 2019, 116, 2476–2481. [Google Scholar] [CrossRef] [Green Version]
  109. Kumar, R.; Senthamaraikannan, P.; Saravanakumar, S.; Khan, A.; Ganesh, K.; Ananth, S.V. Electroactive polymer composites and applications. In Polymer Nanocomposite-Based Smart Materials; Elsevier: Amsterdam, The Netherlands, 2020; pp. 149–156. [Google Scholar] [CrossRef]
  110. Cheng, Z.; Zhang, Q. Field-Activated Electroactive Polymers. MRS Bull. 2008, 33, 183–187. [Google Scholar] [CrossRef]
  111. Chang, L.; Liu, Y.; Yang, Q.; Yu, L.; Liu, J.; Zhu, Z.; Lu, P.; Wu, Y.; Hu, Y. Ionic Electroactive Polymers Used in Bionic Robots: A Review. J. Bionic Eng. 2018, 15, 765–782. [Google Scholar] [CrossRef]
  112. O’halloran, A.; O’malley, F.; McHugh, P. A review on dielectric elastomer actuators, technology, applications, and challenges. J. Appl. Phys. 2008, 104, 071101. [Google Scholar] [CrossRef]
  113. Brochu, P.; Pei, Q. Dielectric Elastomers for Actuators and Artificial Muscles. In Electroactivity in Polymeric Materials; Springer: Boston, MA, USA, 2012; pp. 1–56. [Google Scholar]
  114. Anderson, I.A.; Gisby, T.A.; McKay, T.G.; O’brien, B.M.; Calius, E.P. Multi-functional dielectric elastomer artificial muscles for soft and smart machines. J. Appl. Phys. 2012, 112, 041101. [Google Scholar] [CrossRef]
  115. Rosset, S.; Shea, H.R. Flexible and stretchable electrodes for dielectric elastomer actuators. Appl. Phys. A 2013, 110, 281–307. [Google Scholar] [CrossRef] [Green Version]
  116. Brochu, P.; Pei, Q. Advances in Dielectric Elastomers for Actuators and Artificial Muscles. Macromol. Rapid Commun. 2010, 31, 10–36. [Google Scholar] [CrossRef]
  117. Gu, G.-Y.; Zhu, J.; Zhu, L.-M.; Zhu, X. A survey on dielectric elastomer actuators for soft robots. Bioinspir. Biomim. 2017, 12, 011003. [Google Scholar] [CrossRef]
  118. Wang, N.; Cui, C.; Guo, H.; Chen, B.; Zhang, X. Advances in dielectric elastomer actuation technology. Sci. China Technol. Sci. 2018, 61, 1512–1527. [Google Scholar] [CrossRef]
  119. Chortos, A.; Hajiesmaili, E.; Morales, J.; Clarke, D.R.; Lewis, J.A. 3D Printing of Interdigitated Dielectric Elastomer Actuators. Adv. Funct. Mater. 2020, 30, 1907375. [Google Scholar] [CrossRef]
  120. Gu, G.; Zou, J.; Zhao, R.; Zhao, X.; Zhu, X. Soft wall-climbing robots. Sci. Robot. 2018, 3, eaat2874. [Google Scholar] [CrossRef] [PubMed]
  121. Shintake, J.; Shea, H.; Floreano, D. Biomimetic underwater robots based on dielectric elastomer actuators. In Proceedings of the 2016 IEEE/RSJ International Conference on Intelligent Robots and Systems (IROS), Daejeon, Republic of Korea, 9–14 October 2016; pp. 4957–4962. [Google Scholar] [CrossRef] [Green Version]
  122. Christianson, C.; Goldberg, N.N.; Deheyn, D.D.; Cai, S.; Tolley, M.T. Translucent soft robots driven by frameless fluid electrode dielectric elastomer actuators. Sci. Robot. 2018, 3, eaat1893. [Google Scholar] [CrossRef]
  123. Jung, K.; Koo, J.C.; Nam, J.-D.; Lee, Y.K.; Choi, H.R. Artificial annelid robot driven by soft actuators. Biomimetics 2007, 2, S42. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Nguyen, C.T.; Phung, H.; Nguyen, T.D.; Lee, C.; Kim, U.; Lee, D.; Moon, H.; Koo, J.; Nam, J.-D.; Choi, H.R. A small biomimetic quadruped robot driven by multistacked dielectric elastomer actuators. Smart Mater. Struct. 2014, 23, 065005. [Google Scholar] [CrossRef]
  125. Xu, L.; Chen, H.-Q.; Zou, J.; Dong, W.-T.; Gu, G.-Y.; Zhu, L.-M.; Zhu, X.-Y. Bio-inspired annelid robot: A dielectric elastomer actuated soft robot. Bioinspir. Biomim. 2017, 12, 025003. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Li, W.-B.; Zhang, W.-M.; Zou, H.-X.; Peng, Z.-K.; Meng, G. A Fast Rolling Soft Robot Driven by Dielectric Elastomer. IEEE/ASME Trans. Mechatron. 2018, 23, 1630–1640. [Google Scholar] [CrossRef]
  127. Henke, E.-F.M.; Schlatter, S.; Anderson, I.A.; Nasab, A.M.; Sabzehzar, A.; Tatari, M.; Majidi, C.; Shan, W.; Ainla, A.; Verma, M.S.; et al. Soft Dielectric Elastomer Oscillators Driving Bioinspired Robots. Soft Robot. 2017, 4, 353–366. [Google Scholar] [CrossRef] [Green Version]
  128. da Cunha, M.P.; Debije, M.G.; Schenning, A.P.H.J. Bioinspired light-driven soft robots based on liquid crystal polymers. Chem. Soc. Rev. 2020, 49, 6568–6578. [Google Scholar] [CrossRef]
  129. Ohm, C.; Brehmer, M.; Zentel, R. Liquid Crystalline Elastomers as Actuators and Sensors. Adv. Mater. 2010, 22, 3366–3387. [Google Scholar] [CrossRef]
  130. Cui, Y.; Yin, Y.; Wang, C.; Sim, K.; Li, Y.; Yu, C.; Song, J. Transient thermo-mechanical analysis for bimorph soft robot based on thermally responsive liquid crystal elastomers. Appl. Math. Mech. 2019, 40, 943–952. [Google Scholar] [CrossRef]
  131. Rogóż, M.; Dradrach, K.; Xuan, C.; Wasylczyk, P. A Millimeter-Scale Snail Robot Based on a Light-Powered Liquid Crystal Elastomer Continuous Actuator. Macromol. Rapid Commun. 2019, 40, 1900279. [Google Scholar] [CrossRef] [PubMed]
  132. Rogóż, M.; Zeng, H.; Xuan, C.; Wiersma, D.S.; Wasylczyk, P. Light-Driven Soft Robot Mimics Caterpillar Locomotion in Natural Scale. Adv. Opt. Mater. 2016, 4, 1689–1694. [Google Scholar] [CrossRef]
  133. da Cunha, M.P.; Ambergen, S.; Debije, M.G.; Homburg, E.F.G.A.; Toonder, J.M.J.D.; Schenning, A.P.H.J. A Soft Transporter Robot Fueled by Light. Adv. Sci. 2020, 7, 1902842. [Google Scholar] [CrossRef] [Green Version]
  134. Wei, W.; Zhang, Z.; Wei, J.; Li, X.; Guo, J. Phototriggered Selective Actuation and Self-Oscillating in Dual-Phase Liquid Crystal Photonic Actuators. Adv. Opt. Mater. 2018, 6, 1800131. [Google Scholar] [CrossRef]
  135. Bhandari, B.; Lee, G.-Y.; Ahn, S.-H. A review on IPMC material as actuators and sensors: Fabrications, characteristics and applications. Int. J. Precis. Eng. Manuf. 2012, 13, 141–163. [Google Scholar] [CrossRef]
  136. Haq, M.U.; Gang, Z. Ionic polymer–metal composite applications. Emerg. Mater. Res. 2016, 5, 153–164. [Google Scholar] [CrossRef] [Green Version]
  137. Jo, C.; Pugal, D.; Oh, I.-K.; Kim, K.J.; Asaka, K. Recent advances in ionic polymer–metal composite actuators and their modeling and applications. Prog. Polym. Sci. 2013, 38, 1037–1066. [Google Scholar] [CrossRef]
  138. Carrico, J.D.; Kim, K.J.; Leang, K.K. 3D-printed ionic polymer-metal composite soft crawling robot. In Proceedings of the 2017 IEEE International Conference on Robotics and Automation (ICRA), Singapore, 29 May–3 June 2017; pp. 4313–4320. [Google Scholar] [CrossRef]
  139. Guo, S.; Ge, Y.; Li, L.; Liu, S. Underwater Swimming Micro Robot Using IPMC Actuator. In Proceedings of the 2006 International Conference on Mechatron. and Automation, Luoyang, China, 25–28 June 2006; pp. 249–254. [Google Scholar] [CrossRef]
  140. Yeom, S.-W.; Oh, I.-K. A biomimetic jellyfish robot based on ionic polymer metal composite actuators. Smart Mater. Struct. 2009, 18, 085002. [Google Scholar] [CrossRef]
  141. Li, H.; Fan, M.; Yue, Y.; Hu, G.; He, Q.; Yu, M. Motion Control of Capsule-like Underwater Robot Utilizing the Swing Properties of Ionic Polymer Metal Composite Actuators. J. Bionic Eng. 2020, 17, 281–289. [Google Scholar] [CrossRef]
  142. Naga, N.; Ishida, T.; Hayakawa, T. Synthesis of Joint-Linker Type Ionic Gels Containing Imidazole Units in the Network. ECS Trans. 2018, 88, 107–118. [Google Scholar] [CrossRef]
  143. Fukagawa, M.; Koshiba, Y.; Fukushima, T.; Morimoto, M.; Ishida, K. Anomalous piezoelectric properties of poly(vinylidene fluoride–trifluoroethylene)/ionic liquid gels. Jpn. J. Appl. Phys. 2018, 57, 04FL06. [Google Scholar] [CrossRef] [Green Version]
  144. Hara, Y.; Yoshida, K.; Khosla, A.; Kawakami, M.; Hosoda, K.; Furukawa, H. Very Wide Sensing Range and Hysteresis Behaviors of Tactile Sensor Developed by Embedding Soft Ionic Gels in Soft Silicone Elastomers. ECS J. Solid State Sci. Technol. 2020, 9, 061024. [Google Scholar] [CrossRef]
  145. Unver, O.; Sitti, M. Tankbot: A Palm-size, Tank-like Climbing Robot using Soft Elastomer Adhesive Treads. Int. J. Robot. Res. 2010, 29, 1761–1777. [Google Scholar] [CrossRef]
  146. Sun, J.; Tighe, B.; Liu, Y.; Zhao, J. Twisted-and-Coiled Actuators with Free Strokes Enable Soft Robots with Programmable Motions. Soft Robot. 2021, 8, 213–225. [Google Scholar] [CrossRef]
  147. Onal, C.D.; Chen, X.; Whitesides, G.M.; Rus, D. Soft Mobile Robots with On-Board Chemical Pressure Generation. In Robotics Research; Springer Tracts in Advanced Robotics; Springer International Publishing: Cham, Switzerland, 2016; pp. 525–540. [Google Scholar] [CrossRef]
  148. Lam, T.L.; Xu, Y. A flexible tree climbing robot: Treebot—Design and implementation. In Proceedings of the 2011 IEEE International Conference on Robotics and Automation, Shanghai, China, 9–13 May 2011; pp. 5849–5854. [Google Scholar] [CrossRef]
  149. Calisti, M.; Giorelli, M.; Laschi, C. A Locomotion Strategy for an Octopus-Bioinspired Robot. In Biomimetic and Biohybrid Systems: First International Conference, Living Machines 2012, Barcelona, Spain, 9–12 July 2012. Proceedings; Lecture Notes in Computer Science; Springer: Berlin/Heidelberg, Germany, 2012; Volume 7375, pp. 337–338. [Google Scholar] [CrossRef]
  150. Moreira, F.; Abundis, A.; Aguirre, M.; Castillo, J.; Bhounsule, P.A. An Inchworm-inspired Robot Based on Modular Body, Electronics and Passive Friction Pads Performing the Two-anchor Crawl Gait. J. Bionic Eng. 2018, 15, 820–826. [Google Scholar] [CrossRef]
  151. Chen, R.; Liu, R.; Chen, J.; Zhang, J. A gecko inspired wall-climbing robot based on electrostatic adhesion mechanism. In Proceedings of the 2013 IEEE International Conference on Robotics and Biomimetics (ROBIO), Shenzhen, China, 12–14 December 2013; pp. 396–401. [Google Scholar] [CrossRef]
  152. Sun, Y.; Liu, Y.; Pancheri, F.; Lueth, T.C. LARG: A Lightweight Robotic Gripper With 3-D Topology Optimized Adaptive Fingers. IEEE/ASME Trans. Mechatron. 2022, 27, 2026–2034. [Google Scholar] [CrossRef]
  153. Liu, C.-H.; Chung, F.-M.; Chen, Y.; Chiu, C.-H.; Chen, T.-L. Optimal Design of a Motor-Driven Three-Finger Soft Robotic Gripper. IEEE/ASME Trans. Mechatron. 2020, 25, 1830–1840. [Google Scholar] [CrossRef]
  154. Sayed, M.E.; Roberts, J.O.; McKenzie, R.M.; Aracri, S.; Buchoux, A.; Stokes, A.A. Limpet II: A Modular, Untethered Soft Robot. Soft Robot. 2021, 8, 319–339. [Google Scholar] [CrossRef]
  155. Niu, H.; Feng, R.; Xie, Y.; Jiang, B.; Sheng, Y.; Yu, Y.; Baoyin, H.; Zeng, X. MagWorm: A Biomimetic Magnet Embedded Worm-Like Soft Robot. Soft Robot. 2021, 8, 507–518. [Google Scholar] [CrossRef]
  156. Wang, C.; Sim, K.; Chen, J.; Kim, H.; Rao, Z.; Li, Y.; Chen, W.; Song, J.; Verduzco, R.; Yu, C. Soft Ultrathin Electronics Innervated Adaptive Fully Soft Robots. Adv. Mater. 2018, 30, e1706695. [Google Scholar] [CrossRef] [Green Version]
  157. Lu, X.; Zhang, H.; Fei, G.; Yu, B.; Tong, X.; Xia, H.; Zhao, Y. Liquid-Crystalline Dynamic Networks Doped with Gold Nanorods Showing Enhanced Photocontrol of Actuation. Adv. Mater. 2018, 30, e1706597. [Google Scholar] [CrossRef] [PubMed]
  158. Shahsavan, H.; Aghakhani, A.; Zeng, H.; Guo, Y.; Davidson, Z.S.; Priimagi, A.; Sitti, M. Bioinspired underwater locomotion of light-driven liquid crystal gels. Proc. Natl. Acad. Sci. USA 2020, 117, 5125–5133. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. Huang, Y.; Yu, Q.; Su, C.; Jiang, J.; Chen, N.; Shao, H. Light-Responsive Soft Actuators: Mechanism, Materials, Fabrication, and Applications. Actuators 2021, 10, 298. [Google Scholar] [CrossRef]
  160. Huang, F.; Weng, M.; Feng, Z.; Li, X.; Zhang, W.; Chen, L. Transparent photoactuators based on localized-surface-plasmon-resonant semiconductor nanocrystals: A platform for camouflage soft robots. Nanoscale 2020, 12, 11878–11886. [Google Scholar] [CrossRef]
  161. Wang, X.; Chan, K.H.; Cheng, Y.; Ding, T.; Li, T.; Achavananthadith, S.; Ahmet, S.; Ho, J.S.; Ho, G.W. Somatosensory, Light-Driven, Thin-Film Robots Capable of Integrated Perception and Motility. Adv. Mater. 2020, 32, e2000351. [Google Scholar] [CrossRef] [PubMed]
  162. Nemitz, M.P.; Mihaylov, P.; Barraclough, T.W.; Ross, D.; Stokes, A.A.; Kandhari, A.; Wang, Y.; Chiel, H.J.; Quinn, R.D.; Daltorio, K.A.; et al. Using Voice Coils to Actuate Modular Soft Robots: Wormbot, an Example. Soft Robot. 2016, 3, 198–204. [Google Scholar] [CrossRef]
  163. Bartlett, N.W.; Tolley, M.T.; Overvelde, J.T.B.; Weaver, J.C.; Mosadegh, B.; Bertoldi, K.; Whitesides, G.M.; Wood, R.J. A 3D-printed, functionally graded soft robot powered by combustion. Science 2015, 349, 161–165. [Google Scholar] [CrossRef] [Green Version]
  164. Shepherd, R.F.; Stokes, A.A.; Freake, J.; Barber, J.; Snyder, P.W.; Mazzeo, A.D.; Cademartiri, L.; Morin, S.A.; Whitesides, G.M. Using Explosions to Power a Soft Robot. Angew. Chem. Int. Ed. 2013, 52, 2892–2896. [Google Scholar] [CrossRef]
  165. Wehner, M.; Truby, R.L.; Fitzgerald, D.J.; Mosadegh, B.; Whitesides, G.M.; Lewis, J.A.; Wood, R.J. An integrated design and fabrication strategy for entirely soft, autonomous robots. Nature 2016, 536, 451–455. [Google Scholar] [CrossRef] [Green Version]
  166. Shian, S.; Bertoldi, K.; Clarke, D.R. Use of aligned fibers to enhance the performance of dielectric elastomer inchworm robots. In Electroactive Polymer Actuators and Devices 2015; Bar-Cohen, Y., Ed.; SPIE: Bellingham, DC, USA, 2015; Volume 9430, pp. 417–425. [Google Scholar] [CrossRef]
  167. Sheng, X.; Xu, H.; Zhang, N.; Ding, N.; Zhu, X.; Gu, G. Multi-material 3D printing of caterpillar-inspired soft crawling robots with the pneumatically bellow-type body and anisotropic friction feet. Sens. Actuators A Phys. 2020, 316, 112398. [Google Scholar] [CrossRef]
  168. Calisti, M.; Giorelli, M.; Levy, G.; Mazzolai, B.; Hochner, B.; Laschi, C.; Dario, P. An octopus-bioinspired solution to movement and manipulation for soft robots. Bioinspir. Biomim. 2011, 6, 036002. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Kim, Y.; Parada, G.A.; Liu, S.; Zhao, X. Ferromagnetic soft continuum robots. Sci. Robot. 2019, 4, eaax7329. [Google Scholar] [CrossRef] [PubMed]
  170. Sun, Y.; Lim, C.M.; Tan, H.H.; Ren, H. Soft oral interventional rehabilitation robot based on low-profile soft pneumatic actuator. In Proceedings of the 2015 IEEE International Conference on Robotics and Automation (ICRA), Seattle, WA, USA, 26–30 May 2015; pp. 2907–2912. [Google Scholar] [CrossRef]
  171. Park, Y.-L.; Chen, B.-R.; Pérez-Arancibia, N.O.; Young, D.; Stirling, L.; Wood, R.J.; Goldfield, E.C.; Nagpal, R. Design and control of a bio-inspired soft wearable robotic device for ankle–foot rehabilitation. Bioinspir. Biomim. 2014, 9, 016007. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Shi, Z.; Pan, J.; Tian, J.; Huang, H.; Jiang, Y.; Zeng, S. An Inchworm-inspired Crawling Robot. J. Bionic Eng. 2019, 16, 582–592. [Google Scholar] [CrossRef]
  173. Tanaka, H.; Chen, T.-Y.; Hosoda, K. Dynamic Turning of a Soft Quadruped Robot by Changing Phase Difference. Front. Robot. AI 2021, 8, 629523. [Google Scholar] [CrossRef]
  174. Salem, M.E.M.; Wang, Q.; Wen, R.; Xiang, M. Design and Characterization of Soft Pneumatic Actuator for Universal Robot Gripper. In Proceedings of the 2018 International Conference on Control and Robots (ICCR), Hong Kong, China, 15–17 September 2018; pp. 6–10. [Google Scholar] [CrossRef]
  175. Zhang, C.; Zhu, P.; Lin, Y.; Tang, W.; Jiao, Z.; Yang, H.; Zou, J. Fluid-driven artificial muscles: Bio-design, manufacturing, sensing, control, and applications. Bio-Design Manuf. 2021, 4, 123–145. [Google Scholar] [CrossRef]
  176. Ji, Q.; Fu, S.; Tan, K.; Muralidharan, S.T.; Lagrelius, K.; Danelia, D.; Andrikopoulos, G.; Wang, X.V.; Wang, L.; Feng, L. Synthesizing the optimal gait of a quadruped robot with soft actuators using deep reinforcement learning. Robot. Comput. Manuf. 2022, 78, 102382. [Google Scholar] [CrossRef]
  177. Petković, D.; Danesh, A.S.; Dadkhah, M.; Misaghian, N.; Shamshirband, S.; Zalnezhad, E.; Pavlović, N.D. Adaptive control algorithm of flexible robotic gripper by extreme learning machine. Robot. Comput. Manuf. 2016, 37, 170–178. [Google Scholar] [CrossRef]
  178. Greer, J.D.; Blumenschein, L.H.; Alterovitz, R.; Hawkes, E.W.; Okamura, A.M. Robust navigation of a soft growing robot by exploiting contact with the environment. Int. J. Robot. Res. 2020, 39, 1724–1738. [Google Scholar] [CrossRef] [Green Version]
  179. Man, K.; Damasio, A. Homeostasis and soft robotics in the design of feeling machines. Nat. Mach. Intell. 2019, 1, 446–452. [Google Scholar] [CrossRef] [Green Version]
  180. Chen, Z.; Zhao, C.; Zhang, Y.; Zhu, Y.; Fan, J.; Zhao, J. C-Balls: A Modular Soft Robot Connected and Driven via Magnet Forced. J. Physics Conf. Ser. 2019, 1207, 012006. [Google Scholar] [CrossRef] [Green Version]
  181. Hu, X.; Ge, Z.; Wang, X.; Jiao, N.; Tung, S.; Liu, L. Multifunctional thermo-magnetically actuated hybrid soft millirobot based on 4D printing. Compos. Part B Eng. 2022, 228, 109451. [Google Scholar] [CrossRef]
  182. Park, J.E.; Jeon, J.; Cho, J.H.; Won, S.; Jin, H.-J.; Lee, K.H.; Wie, J.J. Magnetomotility of untethered helical soft robots. RSC Adv. 2019, 9, 11272–11280. [Google Scholar] [CrossRef] [PubMed]
  183. Miriyev, A.; Stack, K.; Lipson, H. Soft material for soft actuators. Nat. Commun. 2017, 8, 596. [Google Scholar] [CrossRef] [Green Version]
  184. Gariya, N.; Kumar, P. A review on soft materials utilized for the manufacturing of soft robots. Mater. Today Proc. 2021, 46, 11177–11181. [Google Scholar] [CrossRef]
  185. Thuruthel, T.G.; Shih, B.; Laschi, C.; Tolley, M.T. Soft robot perception using embedded soft sensors and recurrent neural networks. Sci. Robot. 2019, 4, eaav1488. [Google Scholar] [CrossRef]
  186. Jin, X.; Feng, C.; Ponnamma, D.; Yi, Z.; Parameswaranpillai, J.; Thomas, S.; Salim, N.V. Review on exploration of graphene in the design and engineering of smart sensors, actuators and soft robotics. Chem. Eng. J. Adv. 2020, 4, 100034. [Google Scholar] [CrossRef]
  187. Koh, J.-S.; An, S.-M.; Cho, K.-J. Finger-sized climbing robot using artificial proleg. In Proceedings of the 2010 3rd IEEE RAS & EMBS International Conference on Biomedical Robotics and Biomechatronics, Tokyo, Japan, 26–29 September 2010; pp. 610–615. [Google Scholar] [CrossRef]
  188. Koh, J.-S.; Cho, K.-J. Omega-Shaped Inchworm-Inspired Crawling Robot With Large-Index-and-Pitch (LIP) SMA Spring Actuators. IEEE/ASME Trans. Mechatron. 2013, 18, 419–429. [Google Scholar] [CrossRef]
  189. Kadir, M.R.A.; Dewi, D.E.O.; Jamaludin, M.N.; Nafea, M.; Ali, M.S.M. A multi-segmented shape memory alloy-based actuator system for endoscopic applications. Sens. Actuators A Phys. 2019, 296, 92–100. [Google Scholar] [CrossRef]
  190. Wang, W.; Tang, Y.; Li, C. Controlling bending deformation of a shape memory alloy-based soft planar gripper to grip deformable objects. Int. J. Mech. Sci. 2021, 193, 106181. [Google Scholar] [CrossRef]
  191. Hošovský, A.; Piteľ, J.; Židek, K.; Tóthová, M.; Sárosi, J.; Cveticanin, L. Dynamic characterization and simulation of two-link soft robot arm with pneumatic muscles. Mech. Mach. Theory 2016, 103, 98–116. [Google Scholar] [CrossRef]
  192. Miao, Y.; Dong, W.; Du, Z. Design of a Soft Robot with Multiple Motion Patterns Using Soft Pneumatic Actuators. IOP Conf. Series Mater. Sci. Eng. 2017, 269, 012013. [Google Scholar] [CrossRef] [Green Version]
  193. Ohta, P.; Valle, L.; King, J.; Low, K.; Yi, J.; Atkeson, C.G.; Park, Y.-L. Design of a Lightweight Soft Robotic Arm Using Pneumatic Artificial Muscles and Inflatable Sleeves. Soft Robot. 2018, 5, 204–215. [Google Scholar] [CrossRef] [Green Version]
  194. Qin, L.; Liang, X.; Huang, H.; Chui, C.K.; Yeow, R.C.-H.; Zhu, J. A Versatile Soft Crawling Robot with Rapid Locomotion. Soft Robot. 2019, 6, 455–467. [Google Scholar] [CrossRef]
  195. Li, S.; Vogt, D.M.; Rus, D.; Wood, R.J. Fluid-driven origami-inspired artificial muscles. Proc. Natl. Acad. Sci. USA 2017, 114, 13132–13137. [Google Scholar] [CrossRef] [Green Version]
  196. Shintake, J.; Rosset, S.; Schubert, B.; Floreano, D.; Shea, H. Versatile Soft Grippers with Intrinsic Electroadhesion Based on Multifunctional Polymer Actuators. Adv. Mater. 2016, 28, 231–238. [Google Scholar] [CrossRef]
  197. Li, T.; Li, G.; Liang, Y.; Cheng, T.; Dai, J.; Yang, X.; Liu, B.; Zeng, Z.; Huang, Z.; Luo, Y.; et al. Fast-moving soft electronic fish. Sci. Adv. 2017, 3, e1602045. [Google Scholar] [CrossRef] [Green Version]
  198. Terryn, S.; Langenbach, J.; Roels, E.; Brancart, J.; Bakkali-Hassani, C.; Poutrel, Q.-A.; Georgopoulou, A.; Thuruthel, T.G.; Safaei, A.; Ferrentino, P.; et al. A review on self-healing polymers for soft robotics. Mater. Today 2021, 47, 187–205. [Google Scholar] [CrossRef]
  199. Lee, Y.; Song, W.; Sun, J.-Y. Hydrogel soft robotics. Mater. Today Phys. 2020, 15, 100258. [Google Scholar] [CrossRef]
  200. Ijaz, S.; Li, H.; Hoang, M.C.; Kim, C.-S.; Bang, D.; Choi, E.; Park, J.-O. Magnetically actuated miniature walking soft robot based on chained magnetic microparticles-embedded elastomer. Sens. Actuators A Phys. 2020, 301, 111707. [Google Scholar] [CrossRef]
  201. Eshaghi, M.; Ghasemi, M.; Khorshidi, K. Design, manufacturing and applications of small-scale magnetic soft robots. Extreme Mech. Lett. 2021, 44, 101268. [Google Scholar] [CrossRef]
  202. Venkiteswaran, V.K.; Samaniego, L.F.P.; Sikorski, J.; Misra, S. Bio-Inspired Terrestrial Motion of Magnetic Soft Millirobots. IEEE Robot. Autom. Lett. 2019, 4, 1753–1759. [Google Scholar] [CrossRef] [Green Version]
  203. Zhang, P.; Zhou, Q. Voice coil based hopping mechanism for microrobot. In Proceedings of the 2009 IEEE International Conference on Robotics and Automation, Kobe, Japan, 12–17 May 2009; pp. 3001–3006. [Google Scholar] [CrossRef]
  204. Loepfe, M.; Schumacher, C.M.; Lustenberger, U.B.; Stark, W.J. An Untethered, Jumping Roly-Poly Soft Robot Driven by Combustion. Soft Robot. 2015, 2, 33–41. [Google Scholar] [CrossRef]
  205. Yang, Y.; Ye, X.; Guo, S. A new type of jellyfish-like microrobot. In Proceedings of the IEEE International Conference on Integration Technology (ICIT’07), Shenzhen, China, 20–24 March 2007; pp. 673–678. [Google Scholar] [CrossRef]
  206. Chen, Y.-T.; Yen, J.-Y.; Liu, S.-H. Sensor fusion in a Six-legged Bio-mimicking Robot. IFAC Proc. Vol. 2008, 41, 15624–15629. [Google Scholar] [CrossRef] [Green Version]
  207. Wang, X.-Y.; Zhang, Y.; Fu, X.-J.; Xiang, G.-S. Design and Kinematic Analysis of a Novel Humanoid Robot Eye Using Pneumatic Artificial Muscles. J. Bionic Eng. 2008, 5, 264–270. [Google Scholar] [CrossRef]
  208. Wakimoto, S.; Ogura, K.; Suzumori, K.; Nishioka, Y. Miniature soft hand with curling rubber pneumatic actuators. In Proceedings of the 2009 IEEE International Conference on Robotics and Automation, Kobe, Japan, 12–17 May 2009; pp. 556–561. [Google Scholar] [CrossRef]
  209. Shepherd, R.F.; Ilievski, F.; Choi, W.; Morin, S.A.; Stokes, A.A.; Mazzeo, A.D.; Chen, X.; Wang, M.; Whitesides, G.M. Multigait soft robot. Proc. Natl. Acad. Sci. USA 2011, 108, 20400–20403. [Google Scholar] [CrossRef]
  210. Yamaguchi, A.; Takemura, K.; Yokota, S.; Edamura, K. A robot hand using electro-conjugate fluid: Grasping experiment with balloon actuators inducing a palm motion of robot hand. Sens. Actuators A Phys. 2012, 174, 181–188. [Google Scholar] [CrossRef]
  211. Seok, S.; Onal, C.D.; Cho, K.-J.; Wood, R.J.; Rus, D.; Kim, S. Meshworm: A Peristaltic Soft Robot With Antagonistic Nickel Titanium Coil Actuators. IEEE/ASME Trans. Mechatron. 2013, 18, 1485–1497. [Google Scholar] [CrossRef]
  212. Felton, S.M.; Tolley, M.T.; Onal, C.D.; Rus, D.; Wood, R.J. Robot self-assembly by folding: A printed inchworm robot. In Proceedings of the 2013 IEEE International Conference on Robotics and Automation, Karlsruhe, Germany, 6–10 May 2013; pp. 277–282. [Google Scholar] [CrossRef]
  213. Mao, S.; Dong, E.; Jin, H.; Xu, M.; Zhang, S.; Yang, J.; Low, K.H. Gait study and pattern generation of a starfish-like soft robot with flexible rays actuated by SMAs. J. Bionic Eng. 2014, 11, 400–411. [Google Scholar] [CrossRef]
  214. Kim, H.-I.; Han, M.-W.; Song, S.-H.; Ahn, S.-H. Soft morphing hand driven by SMA tendon wire. Compos. Part B Eng. 2016, 105, 138–148. [Google Scholar] [CrossRef]
  215. Rodrigue, H.; Wang, W.; Kim, D.-R.; Ahn, S.-H. Curved shape memory alloy-based soft actuators and application to soft gripper. Compos. Struct. 2017, 176, 398–406. [Google Scholar] [CrossRef]
  216. Modabberifar, M.; Spenko, M. A shape memory alloy-actuated gecko-inspired robotic gripper. Sens. Actuators A Phys. 2018, 276, 76–82. [Google Scholar] [CrossRef]
  217. Zou, J.; Lin, Y.; Ji, C.; Yang, H. A Reconfigurable Omnidirectional Soft Robot Based on Caterpillar Locomotion. Soft Robot. 2018, 5, 164–174. [Google Scholar] [CrossRef] [PubMed]
  218. Cao, J.; Liang, W.; Ren, Q.; Gupta, U.; Chen, F.; Zhu, J. Modelling and Control of a Novel Soft Crawling Robot Based on a Dielectric Elastomer Actuator. In Proceedings of the 2018 IEEE International Conference on Robotics and Automation (ICRA), Brisbane, Australia, 21–25 May 2018; pp. 1–9. [Google Scholar] [CrossRef]
  219. Feng, N.; Wang, H.; Hu, F.; Gouda, M.A.; Gong, J.; Wang, F. A fiber-reinforced human-like soft robotic manipulator based on sEMG force estimation. Eng. Appl. Artif. Intell. 2019, 86, 56–67. [Google Scholar] [CrossRef]
Figure 1. Various types of soft robots and their actuation mechanisms. (a) A centimetre-level untethered soft robot [15]. (b) A compliant robotic gripper [16]. (c) A spatial soft arm [17]. (d) An energy-efficient soft gripper [18]. (e) Bionic Handling Assistant (BHA) [19]. (f) A hydraulic robotic arm inspired by an octopus [20]. (g) A triboelectric nanogenerator-powered soft robot [21]. (h) A soft crawling robot motivated by a single twisted and coiled actuator (TCA) [22]. (i) A light-driven robot made of liquid crystal elastomer membrane [23]. (j) A novel inchworm robot driven by an electromagnetic oscillatory actuator (EOA) [24]. (k) A magnetically driven soft robot with multi-modal motion capability [25]. (l) A flexible active origami robot [26].
Figure 1. Various types of soft robots and their actuation mechanisms. (a) A centimetre-level untethered soft robot [15]. (b) A compliant robotic gripper [16]. (c) A spatial soft arm [17]. (d) An energy-efficient soft gripper [18]. (e) Bionic Handling Assistant (BHA) [19]. (f) A hydraulic robotic arm inspired by an octopus [20]. (g) A triboelectric nanogenerator-powered soft robot [21]. (h) A soft crawling robot motivated by a single twisted and coiled actuator (TCA) [22]. (i) A light-driven robot made of liquid crystal elastomer membrane [23]. (j) A novel inchworm robot driven by an electromagnetic oscillatory actuator (EOA) [24]. (k) A magnetically driven soft robot with multi-modal motion capability [25]. (l) A flexible active origami robot [26].
Applsci 13 09255 g001
Figure 2. SMA phases and crystal structures [68,69].
Figure 2. SMA phases and crystal structures [68,69].
Applsci 13 09255 g002
Figure 3. (a) An SMA spring robot [77]. (b) The movement mechanism (① The initial status of the robot, ② The contraction status of the robot, ③ The elongation status of the robot to enable it to move forward). (c) The moving states of each robot’s part in the time intervals.
Figure 3. (a) An SMA spring robot [77]. (b) The movement mechanism (① The initial status of the robot, ② The contraction status of the robot, ③ The elongation status of the robot to enable it to move forward). (c) The moving states of each robot’s part in the time intervals.
Applsci 13 09255 g003
Figure 4. (a) An SMA wire robot [80]. (b) The movement mechanism. (c) The moving states of each robot’s part in the time intervals.
Figure 4. (a) An SMA wire robot [80]. (b) The movement mechanism. (c) The moving states of each robot’s part in the time intervals.
Applsci 13 09255 g004
Figure 5. (a) A pneumatic robot [93]. (b) The movement mechanism. (c) The moving states of each robot’s part in the time intervals.
Figure 5. (a) A pneumatic robot [93]. (b) The movement mechanism. (c) The moving states of each robot’s part in the time intervals.
Applsci 13 09255 g005
Figure 6. (a) A DEA robot [120]. (b) The movement mechanism. (c) The moving states of each robotic part in the time intervals.
Figure 6. (a) A DEA robot [120]. (b) The movement mechanism. (c) The moving states of each robotic part in the time intervals.
Applsci 13 09255 g006
Figure 7. A liquid crystal polymer-based soft robot actuated via NIR light to mimic a human arm to grasp, lift up, lower down, and release an object [128].
Figure 7. A liquid crystal polymer-based soft robot actuated via NIR light to mimic a human arm to grasp, lift up, lower down, and release an object [128].
Applsci 13 09255 g007
Figure 8. (a) The mechanism of IPMC [135]. (b) An IPMC-made soft robot mimics a caterpillar moving along a pipe [138].
Figure 8. (a) The mechanism of IPMC [135]. (b) An IPMC-made soft robot mimics a caterpillar moving along a pipe [138].
Applsci 13 09255 g008
Figure 9. (a) Laser-cutting device to fabricate ionic gel material [142]. (b) A fabricated biomimetic aqua-bot [142]. (c) A fabricated walking soft robot [142].
Figure 9. (a) Laser-cutting device to fabricate ionic gel material [142]. (b) A fabricated biomimetic aqua-bot [142]. (c) A fabricated walking soft robot [142].
Applsci 13 09255 g009
Figure 10. Radar charts of each actuation’s characteristics. ①: Output force/carrying capacity. ②: Miniaturization. ③: Lightweight. ④: Efficiency. ⑤: Response speed. ⑥: Technology maturity.
Figure 10. Radar charts of each actuation’s characteristics. ①: Output force/carrying capacity. ②: Miniaturization. ③: Lightweight. ④: Efficiency. ⑤: Response speed. ⑥: Technology maturity.
Applsci 13 09255 g010
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Li, W.; Hu, D.; Yang, L. Actuation Mechanisms and Applications for Soft Robots: A Comprehensive Review. Appl. Sci. 2023, 13, 9255. https://doi.org/10.3390/app13169255

AMA Style

Li W, Hu D, Yang L. Actuation Mechanisms and Applications for Soft Robots: A Comprehensive Review. Applied Sciences. 2023; 13(16):9255. https://doi.org/10.3390/app13169255

Chicago/Turabian Style

Li, Weidong, Diangang Hu, and Lei Yang. 2023. "Actuation Mechanisms and Applications for Soft Robots: A Comprehensive Review" Applied Sciences 13, no. 16: 9255. https://doi.org/10.3390/app13169255

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop