Next Article in Journal
Planar Printed Structures Based on Matryoshka Geometries: A Review
Next Article in Special Issue
Compliance and Kinetostatics of a Novel 2PRS-2PSS Compliant Parallel Micromanipulator: Modeling and Analysis
Previous Article in Journal
Understanding Quasi-Static and Dynamic Characteristics of Organic Ferroelectric Field Effect Transistors
Previous Article in Special Issue
Controlling the Collective Behaviors of Ultrasound-Driven Nanomotors via Frequency Regulation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Survey of Recent Developments in Magnetic Microrobots for Micro-/Nano-Manipulation

by
Ruomeng Xu
and
Qingsong Xu
*
Department of Electromechanical Engineering, Faculty of Science and Technology, University of Macau, Avenida da Universidade, Taipa, Macau, China
*
Author to whom correspondence should be addressed.
Micromachines 2024, 15(4), 468; https://doi.org/10.3390/mi15040468
Submission received: 12 March 2024 / Revised: 23 March 2024 / Accepted: 28 March 2024 / Published: 29 March 2024
(This article belongs to the Special Issue Advanced Micro-/Nano-Manipulation and Positioning Techniques)

Abstract

:
Magnetically actuated microrobots have become a research hotspot in recent years due to their tiny size, untethered control, and rapid response capability. Moreover, an increasing number of researchers are applying them for micro-/nano-manipulation in the biomedical field. This survey provides a comprehensive overview of the recent developments in magnetic microrobots, focusing on materials, propulsion mechanisms, design strategies, fabrication techniques, and diverse micro-/nano-manipulation applications. The exploration of magnetic materials, biosafety considerations, and propulsion methods serves as a foundation for the diverse designs discussed in this review. The paper delves into the design categories, encompassing helical, surface, ciliary, scaffold, and biohybrid microrobots, with each demonstrating unique capabilities. Furthermore, various fabrication techniques, including direct laser writing, glancing angle deposition, biotemplating synthesis, template-assisted electrochemical deposition, and magnetic self-assembly, are examined owing to their contributions to the realization of magnetic microrobots. The potential impact of magnetic microrobots across multidisciplinary domains is presented through various application areas, such as drug delivery, minimally invasive surgery, cell manipulation, and environmental remediation. This review highlights a comprehensive summary of the current challenges, hurdles to overcome, and future directions in magnetic microrobot research across different fields.

1. Introduction

With the development of disciplines such as biomaterials, engineering, and medicine, microrobots are gaining increasing attention [1,2,3,4,5,6]. Due to their distinctive attributes, including compact dimensions, minimal weight, and exceptional flexibility, these robots possess the capacity to undertake tasks beyond the reach of conventional industrial robots [7]. Significant endeavors have been undertaken to develop microrobotic systems endowed with potent transport and delivery capabilities. They can access intricate and confined regions of the human body with minimal invasiveness, enabling micro-/nano-manipulation tasks, such as targeted delivery, precise surgical procedures, and medical examinations [8,9,10,11,12]. This suggests that microrobots have significant potential for applications in the biomedical field.
Various driving technologies have been widely applied to enable the desired movement of microrobots. The actuation strategies, categorized into chemical or external propulsion, involve electric, thermal, chemical, optical, ultrasonic, and magnetic methods [13,14,15,16]. Due to its property of wireless control and the higher degree of freedom for displacement conferred by magnetic forces, magnetic propulsion distinguishes itself among driving methods [17]. Compared to microrobots propelled by alternative methods, magnetic microrobots demonstrate relative safety for biological tissues and exhibit excellent controllability [18,19,20]. Thus, magnetic microrobots have sparked widespread research interest, leading to rapid development in recent years. Simultaneously, the development of smart materials has endowed magnetic microrobots with multifunctionality, various structural designs, and the possibility of employing different magnetic propulsion methods [21,22,23].
This paper comprehensively investigates the recent progress in the field of magnetic microrobots and delves into various aspects such as materials, propulsion mechanisms, design strategies, fabrication techniques, and applications. Additionally, it summarizes the challenges encountered in current research and offers prospects for future development. Figure 1 displays a schematic illustration of the magnetic microrobots discussed in this review and serves as a visual guide to outline the article’s structure. The review initiates with an analysis of magnetic materials. Given the widespread utilization of microrobots in the biomedical field, it is essential to consider the biocompatibility of materials. Following this, there is a discussion of propulsion methods. Different forms of external magnetic fields can induce various motions or functional attributes in magnetic materials. The design section covers multiple types of microrobots: some are categorized based on structure, such as helical, ciliary, and biohybrid microrobots, while others are classified based on functionality, including surface walkers and scaffold microrobots. Microrobots are significantly smaller than traditional robots, and conventional manufacturing methods cannot meet the precision requirements. Subsequently, there is a detailed presentation of fabrication techniques. Finally, the paper concludes by emphasizing the diverse applications of magnetic microrobots for micro-/nano-manipulation, including drug delivery, minimally invasive surgery, cell manipulation, and environmental remediation.

2. Materials for Magnetic Microrobots

The choice of microrobot fabrication materials plays a crucial role in their locomotion performance and functional properties, which significantly influences their potential application scenarios [24,25,26]. Therefore, investigating the material characteristics of magnetic microrobots is of significant importance for their development.
Microrobots containing magnetic materials can react to external magnetic fields, allowing for directional guidance and the generation of propulsive forces [27,28,29,30]. Furthermore, to preserve additional characteristics of microrobots, such as structural rigidity, biocompatibility, and biodegradability, it is essential to incorporate non-magnetic materials [31,32,33,34].

2.1. Magnetic Materials

For the fabrication of magnetic microrobots, magnetic materials, with their intrinsic property allowing them to be guided by an external magnetic field, have undergone thorough investigation in biomedical applications.
Based on their susceptibility to magnetization in a magnetic field, magnetic materials can be categorized into ferromagnetic [35,36], paramagnetic [37], and antiferromagnetic materials [38]. Ferromagnetic materials are predominantly utilized in magnetic microrobots due to their robust magnetic properties in comparison to paramagnets and antimagnets. Ferromagnetic materials demonstrate significant magnetization in the presence of a magnetic field and can maintain residual magnetism upon the removal of the magnetic field [39]. They exhibit high susceptibility but are inevitably associated with hysteresis effects.
According to the magnitude of coercivity (resistance to demagnetization), ferromagnetic materials can be further classified into hard magnetic (like neodymium—iron—boron, NdFeB) and soft magnetic materials (typically represented by iron and nickel). Both hard and soft magnets exhibit hysteresis behavior, requiring an opposing coercive magnetic field during demagnetization. The relationship between remanent magnetism and coercivity is direct—the higher the remanent magnetism, the higher the coercivity. Consequently, hard magnets possess high coercivity and remanent magnetization, maintaining their magnetization against external fields and making them suitable for permanent magnet applications. In contrast, soft magnetic materials, characterized by low coercivity, can be easily magnetized and demagnetized by external fields [24].
Paramagnetic materials (like ferrite) are also applicable to microrobot fabrication [40]. Unlike ferromagnetic materials, they do not retain magnetization upon removing the external magnetic field. The absence of magnetic hysteresis in paramagnetic materials can reduce potential negative impacts on microorganisms in biological applications. However, similar to antiferromagnetic materials, paramagnetic materials demonstrate reduced sensitivity to magnetic fields compared to their ferromagnetic counterparts. Consequently, their attraction to the same magnetic field is significantly weaker, resulting in limitations in specific applications.
Superparamagnetic materials are characterized by their tiny particle sizes, leading to robust paramagnetism in the presence of an external magnetic field [41]. The coercivity of magnetic materials is significantly influenced by particle size. At sufficiently small sizes, magnetic particles tend to maintain uniform magnetization and resist the formation of magnetic domains, displaying single-domain characteristics with high remanence. Consequently, as the sizes of ferro- or ferrimagnetic particles decrease below a critical size, their remanence diminishes without an external magnetic field (due to decreasing coercivity). However, in the presence of a magnetic field, superparamagnetic particles retain relatively high magnetic susceptibility [42,43].
Superparamagnetic materials are commonly utilized in the fabrication of ferrofluidic robots, which are alternatively referred to as active drops owing to their liquid properties and capacity to maneuver under the influence of an external magnetic field [44,45]. Ferrofluidic robots offer superior capabilities over elastomeric-based soft robots for navigating narrow and confined spaces, thereby minimizing damage to surrounding biological tissues during biomedical procedures and achieving the least invasive approach [46,47].
If preparing droplets at the micrometer and nanometer scales, special methods, such as microfluidic synthesis, are required. Microfluidic synthesis is a method that utilizes microchannels and microfluidic technology to synthesize microparticles, microstructures, or microdevices [48]. On a microfluidic platform, precise control over reaction conditions and substance transport is achieved by manipulating fluid flow, mixing, and reaction processes within microchannels. It allows for precise control and fabrication of microstructures. This method is commonly employed for the preparation of microdevices with specific functionalities and properties, such as microcapsules, nanoparticles, microfluidic chips, etc., and finds wide applications in the biomedical field, chemical engineering, materials science, and other domains [49].
Due to the absence of the aforementioned magnetic hysteresis effect, superparamagnetic materials do not exhibit remanence without an external magnetic field. This significantly reduces the potential side effects of magnetic materials within biological organisms, making them commonly applied in the field of biomedical research.
The magnetic moments of antiferromagnetic materials align along specific directions under an applied magnetic field, but the directions of adjacent magnetic moments are opposite. This arrangement results in the material’s magnetization being zero, even under an applied magnetic field, and it does not exhibit significant magnetism. Due to this characteristic, antiferromagnetic materials are not used in the fabrication of magnetic robots. However, study on antiferromagnetic materials is also significant in the fields of materials science and magnetic materials research, as it holds value for understanding the magnetic behavior of materials and developing new types of magnetic materials.
Most small robots consist of ferromagnetic and superparamagnetic materials, as indicated in Table 1. The widespread use of ferromagnetic compounds, including Ni, Fe, and NdFeB, is credited to their elevated saturation magnetization and their ability to propel and control microrobots under low magnetic field intensities [50,51]. Higher saturation magnetization leads to increased magnetic force in the magnetic field, resulting in enhanced maneuverability.

2.2. Biosafety Materials

Microrobots are extensively employed in biomedical applications due to their small size [41,73,74,75]. The selection of materials should carefully consider both biocompatibility and biodegradability. Otherwise, microrobotic systems may trigger immune responses or inflammation in the body [9,76,77,78]. Before contemplating the actual deployment of microrobots in practical scenarios, it is imperative to meticulously investigate and rectify the biocompatibility, potential biodegradability, and interaction dynamics with the biological elements inherent in such microrobotic systems. Subsequent sections will delineate the endeavors undertaken by the microrobotics community in pursuit of these objectives.

2.2.1. Biocompatible Materials

Biocompatibility is one of the foremost requirements in the design of microrobots for biomedical applications [79]. It not only diminishes the need for microrobot recycling but also prevents inadvertent residues of microrobots [80]. So far, there are two methods for establishing biocompatibility. One approach combines the use of biocompatible materials or the creation of protective coatings [81,82,83,84], while the alternative approach involves the development of biohybrid microstructures. This includes biohybrid microrobots, which consist of artificial constructs affixed to biological cells [85,86], or artificial microrobots enveloped in living cell membranes [87].
Coatings with protective properties—including those derived from biodegradable copolymeric compounds such as PLGA (poly(lactic-co-glycolic acid)) and copolymers containing methacrylic acid, along with natural polymers like chitosan—have exhibited positive biocompatibility in diverse cargo delivery approaches in the gastrointestinal tract (GI) [88,89]. The initial demonstration of in vivo therapy featured microrobots with a magnesium core coated with a layer of antibiotic-loaded PLGA and an outer chitosan covering, ensuring microrobot adherence to the stomach wall through electrostatic interactions. This microrobot design based on PLGA-chitosan composition led to prolonged retention on the stomach wall, thereby augmenting the effectiveness of therapeutic interventions [10].
Furthermore, researchers have explored inorganic coverings, including Ti and Fe. Nelson’s group utilized a burr-like porous spherical structure coated with Ti material to carry and deliver to targeted cells in vivo, ensuring biocompatibility [90]. However, despite ongoing investigations, exposure risks persist during the fabrication process, and the coating layer remains incapable of completely enveloping microrobots.
Derived from polyethylene glycol (PEG), PEGDA (poly(ethylene glycol) diacrylate) is hydrophilic and elastic and boasts favorable photopolymerization characteristics that facilitate the facile fabrication of microrobots with defined configurations for applications in regenerative therapy and tissue regeneration [91,92]. The mechanical properties of the resulting products are modifiable, with the elastic modulus increasing in tandem with the concentration of PEGDA [93]. Liu et al. recently demonstrated the preparation of soft microhelixes with adjustable mechanical characteristics by incorporating PEGDA with alginic acid calcium salt [94]. By manipulating the PEGDA concentration within the initial formulation, one can effortlessly refine the elastic moduli of the pliable microrobots to span from tens of kPa to a few MPa. In achieving the construction of autonomously moving microrobots featuring filamentous hydrogel tails derived from PEGDA, Srivastava et al. utilized live-time, on-site polymerization [95]. The polymeric structure of PEGDA, resembling a thread, possesses low mechanical strength, making it easily bendable and deformable. This characteristic enables the remote capture of cells and microparticles. Furthermore, PDMS and Ecoflex silicone rubbers have been demonstrated to exhibit biocompatibility [96,97,98].
Developing biohybrid microstructures offers an alternative way to establish biocompatibility. The amalgamation of microrobots incorporating components from native cells enhances the overall compatibility of the microscale robotic system with biological organisms, thereby diminishing immune responses. Essentially, the collaborative interaction between the effective mobility of artificial microrobots and the inherent biological characteristics and functions of cellular elements has resulted in the development of microrobots that mimic cells and exhibit inherent biological functionality. This includes capabilities such as toxin and pathogen removal and holds significant promise for therapeutic applications. A recent design involved testing a new magnetic nanopropeller, whereby Fe and Pt were co-deposited onto silicon dioxide, for the delivery of plasmids into cells. The results demonstrated non-toxicity to cells and outstanding biocompatibility [56].

2.2.2. Biodegradable Materials

While biocompatible materials may not completely avoid all safety issues in biomedical applications, the potential presence of residual microrobot components within the body could lead to adverse effects.
Manufactured through an economical production method employing a biological template derived from the cyanobacterium Spirulina, a porous microrobot with a hollow structure is characterized by an outer shell composed of superparamagnetic Fe3O4 and an inner cavity with a helical shape. This microrobot, determined to be compatible with living cells and capable of natural degradation, can be disassembled into smaller fragments under ultrasonic stimulation [99].
While opting for biocompatible or biodegradable materials as the primary constituents for microrobots is a prudent choice for potential clinical applications, the most widely employed strategy still involves using a photosensitive polymer serving as the structural foundation and covered with a coating of materials compatible with living organisms. This approach represents a compromise by considering the manufacturing method’s complexity and the materials’ inherent mechanical robustness. For instance, a magnetic microrobot fabricated using SU-8 (a negative photoresist) with a tailored microstructure created through 3D laser lithography is then covered with a layer of Ni for magnetic manipulation and an extra layer of Ti or Pt to enhance biocompatibility [68].
Since the primary constituents of these microrobots are not subject to natural degradation, there is a risk of accumulation in the body, potentially causing severe inflammation in the absence of an appropriate recycling mechanism. Nevertheless, polymers can be easily fashioned into intricately designed microstructures using 3D laser printing and include both the required hardness and precision of the structure as well as biocompatibility. Consequently, one crucial direction is to develop a 3D manufacturing method based on hydrogel materials to advance the clinical application of cell-transport microrobots.

2.3. Current Challenges and Prospects

To obtain materials that are better suited for microscale robot performance, expertise in multidisciplinary fields such as materials science and electromagnetics is required. While existing magnetic materials generally meet current application demands, achieving precise control for medical applications remains a significant challenge. The most widely used ferromagnetic materials exhibit residual magnetism even in the absence of a magnetic field, which is accompanied by hysteresis effects. In applications requiring precise control of magnetic material magnetization, hysteresis effects may result in insufficiently prompt changes in magnetic field intensity, thereby affecting system response speed and accuracy [100,101,102]. Hysteresis effects pose obstacles to achieving precise control with ferromagnetic materials, inevitably leading to energy loss and potentially affecting material lifespan. Paramagnetic materials, although devoid of hysteresis effects, exhibit low sensitivity to magnetic fields, making them less responsive to subtle magnetic field changes. Moreover, they require relatively large magnetic fields for driving and are subject to certain limitations in applications. In the future, exploring magnetic materials that combine the advantages of both materials could facilitate precise control.
The biosafety of microrobots represents a higher pursuit, with current obstacles primarily stemming from fabrication methods. For instance, as mentioned earlier, coating methods face challenges in achieving complete encapsulation without compromising precise magnetic manipulation in a magnetic field.

3. Propulsion of Magnetic Microrobots

When subjected to an external magnetic field, magnetic microrobots undergo both a magnetic force, denoted as F m , and a magnetic torque, T m [103]. The expression for the magnetic force acting on the microrobot is:
F m = V ( M · ) B
where V and M represent the volume and magnetization of the microrobot’s magnetic material, respectively. B signifies the magnetic flux density of the external magnetic field.
Simultaneously, the magnetic torque T m can be mathematically expressed as:
T m = V M × B
As the magnetic microrobot is exposed to the magnetic field, it becomes magnetized, generating magnetization in the process. For a linear isotropic media, the magnetization is influenced by the external field and follows the relationship between M and H :
M = χ a H
where χ a and H denote the susceptibility tensor of the material and the magnetic field strength, respectively.
In the case of a homogeneous magnetic field (where the gradient ▽ vanishes), the magnetic robot does not encounter a gradient force and typically moves parallel to the field. However, T m possesses the capability to force the magnetic microrobot to orient its dipole moment in alignment with the external magnetic field through rotational motion, especially when the microrobot and the magnetic field are not oriented in the same direction [104]. Therefore, magnetic fields employed for propelling microrobots should be time-invariant (such as rotating and oscillating magnetic fields) or inhomogeneous (e.g., a gradient magnetic field).

3.1. Rotating Magnetic Field

Electromagnetic coils are common devices for generating a rotating magnetic field. In comparison to microrobots activated by magnetic field gradients and oscillating magnetic fields, those propelled by rotating magnetic fields exhibit superior maneuverability and precise locomotion [105,106]. In ideal conditions, they can efficiently operate even in magnetic fields with zero field strength.
Generally, Maxwell coils and Helmholtz coils are commonly used devices for generating magnetic fields; both consist of multiple coil sets. Although they do not directly produce rotating magnetic fields, the desired effect of a rotating magnetic field can be indirectly simulated by appropriately adjusting the direction and intensity of the electric current. They are commonly utilized for driving helical robots, whereby the actuation is accomplished by initiating rolling, corkscrew, and spin-top movements [107,108,109]. By engaging in rotation around the helical axis, these microrobots progress in the direction perpendicular to the rotation plane of the microrobot. Much like their helical counterparts, magnetized spherical microrobots can also effortlessly respond to rotating magnetic fields as well [110]. In 1996, Honda et al. introduced a propulsion method utilizing a rotating magnetic field that involved a square-shaped magnet affixed to a coiled copper conductor [111]. Demonstrating successful propelling in silicone oil with elevated viscosity at a low Reynolds number, the magnetically actuated helical wire showcased its suitability for microscale propulsion. The experimental findings of their study indicated the effectiveness of rotating-magnetic-field-based propulsion methods in microscale environments. Derived from the bacterial flagella movement, Zhang et al. introduced a simulated helical flagellum that is propelled by rotational magnetic fields in a microscale fluidic environment [112]. Consisting of a helical appendage and a slim magnetic head formed via the self-scrolling of a metallic double layer, the artificial flagellum mimicked the motion of natural counterparts. It has been reported that rotating magnetic fields not only manipulate magnetic particle aggregates but also control various other intriguing structures [72,113,114].

3.2. Gradient Magnetic Field

The methods for achieving a gradient magnetic field include employing permanent magnets with position control [115], utilizing electromagnetic devices with the ability to adjust their location by manipulating applied electrical currents [116,117,118], and applying stationary electromagnets with controlled electrical signals [119,120]. Using permanent magnets to achieve gradient magnetic fields is relatively simpler compared to the other two methods. By adjusting the position and orientation of permanent magnets, different magnetic field gradients can be produced in space. This can be achieved through mechanical approaches or magnetic positioning systems. If electromagnetic devices or static electromagnetic systems are used, precise position control systems and current regulation devices are required to ensure that the magnetic field intensity and direction can be adjusted and controlled as needed.
In contrast to microrobots propelled by a rotating or an alternating magnetic field, those activated by magnetic field gradients have fewer constraints on their mechanical structures and can travel in alignment with the field gradient. Magnetic-field-gradient-driven microrobots often adopt spherical and cylindrical structures, as these configurations experience minimal surface frictional forces.

3.3. Oscillating Magnetic Field

The methods for generating an oscillating magnetic field include driving electromagnets or coils with alternating current power sources as well as utilizing specific circuits and control systems to induce periodic variations in the magnetic field [110,121].
The generation of non-reciprocal motion in magnetic materials is attributed to oscillating magnetic fields and is identified as magnetic vectors varying over time and moving vertically within a plane [122,123]. This non-reciprocal motion is a crucial feature essential for propulsion methods employing time-varying magnetic torque. In contrast to microrobots propelled by a rotating magnetic field or magnetic field gradients, those propelled by oscillating magnetic fields employ asymmetrical shape deformation to overcome Purcell’s renowned “scallop theorem” [99]. The prevalent design for microrobots driven by oscillating magnetic fields is a multijoint structure, drawing inspiration from the swimming mechanisms observed in fish or bacterial flagella [39,124].

3.4. Current Challenges and Prospects

Due to the increasing popularity of magnetic propulsion research, an expanding array of magnetic propulsion systems has been developed, with many commercially deployed in medical and other fields. Currently, there are two primary types of magnetic propulsion systems: electromagnetic systems and permanent magnet systems. Electromagnetic systems encompass multi-axis Helmholtz coils [125], OctoMag [126], and Maxwell [127]. Permanent magnet systems include single-magnet systems and multi-magnet systems such as the commercially available Niobe system [128].
Although a considerable number of magnetic propulsion systems have been commercialized, several issues remain unresolved. First, the penetrability of these existing magnetic fields is a significant concern. Current magnetic fields used for precise manipulation often require substantial magnetic strength, resulting in high energy consumption and, consequently, elevated costs. Additionally, intense magnetic fields may pose side effects on the human body.
Second, the practicality of current magnetic fields requires improvement. Many commercially available magnetic propulsion systems have limited applications. For instance, the system proposed by Ankon Technologies is currently only employed for propelling magnetic endoscopes [129], while the Niobe system is utilized for guiding catheters in cardiovascular disease treatment.
Microrobots hold vast potential in the fields of cluster control and three-dimensional control. Primarily, cluster control enhances the efficiency and agility of microrobots in task execution [130]. Through collaborative efforts, they can jointly accomplish intricate tasks such as environmental monitoring and search and rescue operations, thus augmenting work efficiency. Additionally, advancements in three-dimensional control technologies empower microrobots to freely navigate in complex three-dimensional spaces, enabling exploration of uncharted territories and unlocking new application potentials in domains like healthcare, construction, and environmental monitoring [131].
Nevertheless, microrobots encounter numerous challenges in both cluster control and three-dimensional control. Firstly, cluster control necessitates addressing issues such as communication, localization, and path planning. The diminutive size and limited computational capabilities of microrobots render cluster cooperation increasingly intricate. Secondly, precise control of microrobots in three-dimensional space requires consideration of factors like gravity and air resistance. Moreover, microrobots are susceptible to environmental disturbances during movement, further complicating control efforts. These challenges, in turn, constrain their practical utility.
In summary, future research endeavors will primarily focus on achieving strong penetrability and multifunctional practicality.

4. Design of Magnetic Microrobots

Due to the constraints imposed by fabrication methods and the distinct physical phenomena dictated by low-Reynolds-number hydrodynamics at the microscale, reducing macroscopic actuation mechanisms for generating microscale motion is often impractical [132].

4.1. Helical Microrobots

In a low-Reynolds-number environment, microrobots achieve effective propulsion by producing anet unidirectional movement through asymmetric motion. Inspired by the bacterial flagellum, a design approach that incorporates magnetic helical structures on microrobots enables propulsion by applying magnetic torques within the body under a weak rotating magnetic field [133]. Typically, a helical microrobot consists of either a magnetic helical tail or one affixed to a magnetic head. This design facilitates a corkscrew-like motion, translating rotational motion around the helical axis into nonreciprocal translational movement [134,135]. Alteration of the external magnetic field’s counterclockwise or clockwise direction makes it easy for helical robots to attain forward or reverse movement [13].
In 2009, Abbott et al. conducted a comparative analysis of three promising methods for microrobot swimming, with the magnetic propulsion of helical structures achieving swimming motion for the first time [110]. Subsequently, helical microrobots have found increasing application in the biomedical field. Proposing a degradable superparamagnetic polymer composite, Peters et al. offered safe degradation of the device in vivo, followed by the non-invasive excretion of degradation products, presenting an alternative to manual or surgical device recovery [83]. The results of in vitro biodegradation experiments are presented in Figure 2a. Subsequently, Ceylan et al. designed a water-based, enzyme-degradable dual-helical microswimmer propelled and controlled by magnetic forces [41]. Under rotational magnetic fields, this microrobot can load cargo volumetrically and exhibit swimming capabilities. The 3D microswimmer was optimized for performance and was 3D-printed using two-photon polymerization. The printing process utilized a gelatin methacryloyl-based magnetic suspension and biofunctionalized magnetic iron nanoparticles with superparamagnetic properties. Figure 2b presents the dual-helical structure of the microrobot. Nowadays, an increasing number of helical microrobots are addressing specific challenges in medical applications. For instance, a microrobotic hierarchical superstructure was proposed in 2023 by Landers et al. [136]. Comprising magnetic helical micromachines intricately connected to a thermally responsive transient magnetic polymer framework, these formations are designed to intelligently navigate and transport small magnetic helical micromachines to intricate small vessels and capillaries, thereby facilitating access to challenging anatomical sites within the human body. As shown in Figure 2c, the microrobot can traverse various shapes of microchannels under the influence of an external magnetic field. Liu et al. successfully proposed a micromotor through microfluidic chips, as illustrated in Figure 2d [137]. Capable of loading neuron stem cells and precisely delivering them to targeted areas under external magnetic fields, the micromotor also contributes to restoring the interconnection of neurons. This research introduces a novel approach for reconstructing neural networks in cases of neuron injuries.

4.2. Surface Microrobots

Surface walking has emerged as a distinctive propulsion method in recent microrobots applications. An illustrative example is the surface-walking Janus microdimer, a microsurface walker activated by magnets and consisting of two magnetically connected Janus spheres made of Ni/Si O 2 [138,139]. When subjected to a flat, undulating magnetic field near a surface, these spheres exhibit an asymmetric rolling motion over each other, resulting in a net displacement. This innovative approach introduces new possibilities in microrobot propulsion, particularly for navigating confined spaces and intricate geometries. Due to its robust navigational capabilities and efficiency in speed modulation, developers envision diverse applications for this device ranging from nanomanipulation to precision medicine treatments.

4.3. Ciliary Microrobots

Magnetic microrobots employing propulsion mechanisms based on traveling waves and metachronal waves draw inspiration from ciliate microscopic organisms in nature, such as Paramecia. This robotic type disrupts temporal symmetry, achieving effective movement through phase discrepancies in the beating motion of neighboring cilia [140].
In 2007, Evans et al. developed a method to create arrays of high-aspect-ratio cantilevered micro- and nanorods using a PDMS-ferrofluid composite material, as shown in Figure 3a [141]. Belardi et al. introduced a novel process for producing large arrays of magnetically driven artificial cilia by utilizing a two-color lithography-based approach in 2011 [142]. The exploration of methods for fabricating cilia has contributed to the development of ciliary microrobots. Furthermore, multi-joint microstructures can generate propagating wave or metachronal undulation propulsion. In 2018, Li et al. designed a magnetic artificial multi-legged millirobot [65]. Showcasing exceptional adaptability in various challenging environments, this robot design demonstrates remarkable features such as rapid mobility, high carrying capacity, and outstanding obstacle-crossing ability. The gait of the microrobot is illustrated in Figure 3b.
Recently, Wei et al. designed a ciliate microrobot using PDMS and NdFeB that mimics an inchworm’s gait and can move on wet surfaces and inclines. The structure of the robot is illustrated in Figure 3c [143]. Using soft miniature devices, Dong et al. investigated the quantitative relationship between metachronal coordination and consequent fluid dynamics [144]. Additionally, they designed fluidic soft devices inspired by cilia with distinctive capabilities for maneuvering and blending thick synthetic and biological fluids in low Reynolds numbers, as depicted in Figure 3d. Xu et al. proposed a swarming method that utilizes physical interactions among magnetic microparticles to arrange them into cilia structures, as shown in Figure 3e [145]. Ciliate microrobots have additional applications. In 2023, Feng et al. incorporated cilia structures into sensor modules, enabling the detection of signals related to joint movements and the ability to sense pressure applied to the skin [146]. The cilia structure, fabricated using carbonyl iron particles, aids with navigation or sensing within the targeted environment. The structure is depicted in Figure 3f.

4.4. Scaffold Microrobots

Drawing inspiration from the porous extracellular matrix, scientists have created 3D microscaffolds with linked pores that act as platforms for cell and tissue cultivation [147,148]. The microscaffolds, which form the basis of magnetically actuated microrobots, offer significant capacity for loading and storing. Playing a vital role in enhancing targeted drug delivery, manipulating cells, and performing minimally invasive surgery, they contribute significantly to efficacy improvement [149,150,151]. In 2017, Go et al. introduced magnetically actuated microscaffolds designed to carry mesenchymal stem cells (MSCs) for repairing joint cartilage. Represented as microspheres with interconnected micro holes, these 3D microscaffolds provide sufficient room to facilitate cell adhesion, provide support, and enable transportation. Besides spherical microscaffold structures, researchers have designed microscaffolds with various shapes by employing state-of-the-art 3D-printing methods. For instance, Jeon et al. created many microscaffold structures characterized by complex shapes and consistent pore sizes. These structures involve cylindrical, helical, and rectangular objects [67,73].

4.5. Biohybrid Microrobots

Over millions of years, biological systems have evolved to operate optimally at the microscale and showcase remarkable movement and functionality. In the past decade, the development of magnetic microrobots has gained attention, with biohybrid systems emerging as attractive approaches. Currently, there is significant interest in biohybrid miniaturized motors, which are characterized by outstanding biocompatibility and minimal toxicity. Consequently, biohybrids present promising alternatives. Comprising two essential elements, biohybrid robots consist of living biological entities or tiny organisms that exhibit optimal biocompatibility and deformability as the first component. The second component involves artificial microstructures or microparticles, which serve as carriers to support these cells [152,153]. These micromotors typically combine synthetic microstructures with non-mobile cells or mobile cells [2,154].

4.5.1. Microrobots Based on Nonmobile Cells

Certain plant cells, like pollen and spores, can be employed to fabricate biohybrid robots. Generally, pollen and spores exhibit excellent biocompatibility features and structural uniformity. Additionally, specific plant cells, such as those with unique architectures like pollen cavities, have the potential to serve as carriers for increased cargo capacity [155,156,157].
Utilizing a vacuum loading technique, Sun et al. successfully embedded drugs into the hollow section of pine pollen grains [72]. The experiments showcased the remarkable drug-encapsulation capability of the pollen-based magnetic microrobots and highlighted their effective containment capability and precision in drug discharge. Driven by programmatically controllable magnetic fields and containing encapsulated magnetic Fe3O4, these micromotors exhibited three different movement modes: rolling, crawling, and rotating. Moreover, under the influence of an external magnetic field, individual pollen-based micromotors could collectively form dynamic phenomena. The dynamic phenomena refer to the controlled release of therapeutic cargo achieved through the manipulation of magnetic nanoparticle aggregation under a powerful magnetic field. This phenomenon is crucial as it enables precise and controlled drug delivery, which is essential for various biomedical applications.
Additionally, spores offer another avenue for the synthesis of magnetically actuated biohybrid microrobots [158,159]. In a study conducted by Zhang et al., magnetic nanoparticles were directly deposited, followed by the encapsulation of functionalized carbon nanodots on porous natural spores, resulting in the synthesis of magnetic spore-based microrobots [159], as shown in Figure 4a.

4.5.2. Microrobots Based on Mobile Cells

In nature, cell structures endowed with mobility not only serve as inspirational sources for the intricate designs of microrobots but also offer numerous functional advantages when integrated into microrobots. These advantages encompass attributes such as biodegradability, ease of absorption, and spontaneous fluorescence. Expanding the biological applications of magnetic robots is achievable by incorporating these intriguing functionalities into their fabrication. The common biohybrid microrobots are depicted in Figure 4.
One method for creating hybrid microrobots involves the combination of active locomotive cells, particularly those possessing flagella, with sperm and bacteria being commonly employed for this purpose [85,165,166,167]. In this approach, the motile cell can either attach to the surface of a synthetic particle or another cell or become ensnared within a specialized microstructure. Under a rotating magnetic field, a self-assembled magnetic nanorobot created by Ali et al. utilizes bacterial flagella attached to a superparamagnetic particle for actuation and steering [160], as depicted in Figure 4b. Due to the high binding affinity of doxorubicin hydrochloride (DOX-HCl), i.e., a hydrophilic anticancer drug, to DNA (nucleus), sperm were found to exhibit significant uptake of DOX-HCl in a study [166]. With the goal of designing a drug-loading microrobot aimed at targeted cancer treatment, Magdanz et al. employed bovine sperm cells as a template to fabricate soft magnetic microrobots through a straightforward electrostatic-based method (Figure 4d) [162].
Another approach involves utilizing live immune cells to engulf entire passive magnetic functional particles, leveraging the phagocytosis processes of immune cells [168]. When macrophages completely engulf the magnetic double-helix micromotor, the microrobot exhibits rolling motion driven by magnetism along a predetermined trajectory, propelling the magnetic spiral components. These robots showcase the capability of continuous swimming even when facing obstacles from cells. Without an external magnetic field, the immunobot autonomously moves by creeping, driven by the self-propulsion locomotion of macrophages in vivo [169].
In addition, Yan et al. discovered that Spirulina platensis, a helical microalgae subspecies, exhibits intrinsic fluorescence, selective cytotoxicity against cancer cells, and natural degradability [170]. By employing a dip-coating method to attach Fe3O4 nanoparticles to the surface of Spirulina platensis, they produced magnetic microrobots collectively known as magnetized Spirulina.

4.6. Current Challenges and Prospects

Researchers have designed various types and sizes of microrobots to accommodate different application scenarios. Considering applications in the medical field, there are still some barriers to overcome. Produced through specific fabrication methods (discussed later), current magnetic microrobots can achieve sizes ranging from nanometers to millimeters. However, a significant challenge arises: how can microrobots of such small dimensions resist fluid flow? Currently, there are two main approaches to address this issue.
Firstly, increasing the external magnetic field is one option, but it comes with high costs and difficulty with providing sufficient driving force due to the small size of the robots. Second, adapting to different flow velocities by designing robots with various sizes presents a compromised solution. However, increasing the size may lead to the issues such as vascular blockages. In the future, hopefully there will be specific structural designs to mitigate these problems.
As microrobots continue to gain widespread adoption in biomedical settings, the importance of biocompatibility is expected to grow significantly as a focal point of future research. Biohybrid microrobots, benefiting from biological components, exhibit enhanced biocompatibility. It can be boldly speculated that there will be an increasing number of biohybrid microrobots tailored to diverse intracorporeal environments in the future. Furthermore, these biohybrid microrobots hold potential to address the challenges related to crossing the blood–brain barrier, which is traditionally considered a formidable obstacle in biomedical interventions.

5. Fabrication of Magnetic Microrobots

After the discussion of various magnetic microrobots, it is also crucial to understand how they achieve their functions through fabrication. Due to the vast differences in structure and materials between magnetic microrobots and industrial robots, the development of suitable fabrication methods for microrobots is urgently needed. Currently, common preparation methods for magnetic microrobots include direct laser writing (DLW), glancing angle deposition (GLAD), biotemplating synthesis (BTS), template-assisted electrochemical deposition (TAED), and magnetic self-assembly (MSA).

5.1. Direct Laser Writing (DLW)

Direct laser writing (DLW) eliminates the need for a mask plate for employing two-photon absorption to induce changes in the solubility of the resist and developer, distinguishing it from conventional lithography techniques [66,171,172]. Utilizing two-photon polymerization (TPP) nanolithography, DLW enables the production of intricate 3D microscopic-scale structures with exceptionally high spatial resolution.
In a study by Wang et al., the locomotion of spiral microrobots with different surface coatings and exhibiting varying degrees of wetting on their surfaces was investigated [173]. The process involved laser-writing of the photoresist on a glass substrate to create helix structures. Subsequently, the polymer microhelices were coated with nickel for magnetic propulsion, followed by a gold coating. This was achieved through physical evaporation to enhance adhesion during subsequent surface modification. Dong et al. utilized 3D laser lithography based on TPP and water dispersion to construct helical microrobots using hydrogel materials [174]. Subsequently, these printed spiral microrobots are coated with magnetic particles on their surfaces, forming motion components driven by an externally rotating magnetic field. In a study by Ceylan et al., the TPP process was employed to modify a superparamagnetic ferrite matrix, significantly increasing the volume-to-surface area ratio [41]. Additionally, Sun et al. designed and fabricated novel burr-like porous spherical structures using the TPP process; the resulting structures showcased superior cellular delivery efficiency in both physiological and external environments [68].

5.2. Glancing Angle Deposition (GLAD)

As an extension of conventional physical vapor deposition methods, glancing angle deposition (GLAD) involves the synchronous manipulation of the substrate with incident vapor deposition, deviating from the typical practice of deposition on a fixed substrate [81].
In the literature, Walker et al. established a monolayer using 500 nm silicon dioxide on a silicon chip as seed nuclei for nucleation points. Subsequently, nickel was deposited onto the silicon particles at a specific angle for magnetostriction purposes [82]. Simultaneously, Venkataramanababu et al. created a 2D array of photoresist pillars through laser interference lithography. Following the deposition of the seed layer, it enhanced the shape customization capability of spiral robots [175].

5.3. Biotemplating Synthesis (BTS)

As mentioned earlier, recent studies demonstrate that biohybrid microrobots, formed through the combination of suitable microorganisms with artificial micromaterials, maintain their outstanding biological characteristics (including biocompatibility and biodegradability) during migration in a low-Reynolds-number fluid. Consequently, the biotemplating synthesis (BTS) approach enables microscale robots to possess healing capacities for various medical uses in a safe and biocompatible fashion. The state-of-the-art literature showcases similar endeavors in microrobots, such as combining with pollen, bacteria, and cells [176,177]. Notably, the natural microalgae Spirulina (Sp.) stands out due to its remarkable 3D helical microarchitecture and frequently serves as a biotemplate for constructing biohybrid helical microrobots for activities such as transporting goods or purifying water [170,178,179].
The magnetic microrobots based on microalgae recently developed by Wang et al. have significantly enhanced swimming speed under a rotating magnetic field [163]. Another potential candidate for constructing cell-based microrobots is red blood cells (RBCs), which lack nuclei, providing a more spacious environment, and are equipped with hemoglobin for oxygen transport [164]. Emphasizing the double-concave structure of RBCs, Gao et al. highlighted their ability to navigate more accurately with the assistance of an external magnetic field and accommodate a greater amount of photosensitizers [71].

5.4. Template-Assisted Electrochemical Deposition (TAED)

Template-assisted electrochemical deposition (TAED) represents a straightforward yet efficient approach to form a firmly adhering layer of specified materials onto a pre-made conductive substrate. Due to its low cost, rapidity, and ease of operation, this deposition technique is considered suitable for mass production of microstructures. With the assistance of well-designed templates, diverse microstructures, such as helical [180], tubular [181], and rod-like [182], have been successfully deposited. Presenting a tri-segmented strategy, Jordi Sort et al. embedded CoPt/Cu/Ni into polycarbonate nanorods through electrochemical deposition [183]. By incorporating semihard magnetic materials to coat nanowires using an anodic aluminum oxide (AAO)-template demolding and premagnetization process, Jang et al. enabled the proposed nanorobots to execute tumbling, procession, and rolling at various rotating frequencies under a rotating magnetic field [184].

5.5. Magnetic Self-Assembly (MSA)

In addition to the operation of magnetically powered microscopic robots as individual units, there are recent explorations into active magnetic particles displaying collective or swarm behaviors for on-demand in vivo applications [185]. Magnetic self-assembly (MSA), recognized for its versatile multifunctionality achieved through controllable structure reconfiguration, is considered an effective alternative for fabricating swarm robots. Leveraging a 2D flat-plane rotating magnetic field, Tasci et al. engineered superparamagnetic beads capable of self-assembling into size-controlled microwheels [186]. In addition to widely employed colloidal particles, Wang and colleagues explored magnetic droplets made from a water suspension containing benzyl–ether and carbonyl iron microparticles [187]. These ferromagnetic particles settled at the bottom of the droplet due to gravity and quickly formed a chain when subjected to a processing field.

5.6. Current Challenges and Prospects

Current fabrication methods have largely met the basic requirements of magnetic microrobots. However, as the fabrication methods are intricately linked with structural design and material selection of robots, future exploration may lead to the discovery of alternative methods to meet the performance demands of novel microrobots. Additionally, there is a prospective direction for future development in the realm of more convenient and automated fabrication processes. While a few self-assembling magnetic microrobots have been reported, the realization of in vivo self-assembly structures has yet to be achieved. The development of in situ manufacturing techniques could effectively mitigate various issues that may arise during the transportation of microrobots. Hence, it represents a promising avenue for future development.

6. Applications of Magnetic Microrobots

In 1959, Richard Feynman first introduced the concept of miniaturized machines during a lecture [188]. Subsequently, in the 1966 film “Fantastic Voyage”, a man ingested a doctor reduced to microscopic size, and the doctor repaired damage in the person’s brain. In recent years, scenes like this have transcended the realm of science fiction. With the advancement of microrobots, their applications have become widespread, particularly in the medical field with micro-/nano-manipulation [189,190,191], as shown in Figure 5.
Discussed in the preceding sections are the design mechanisms and actuation methods of microrobots, which serve as the foundation for their application in various scenarios. The following section will showcase the various applications of microrobots.

6.1. Drug Delivery

Transporting therapeutic cargoes with precision and efficiency to specific sites in the body, especially the complicated and confined environments, poses a challenge for passive drug delivery systems. Compared to traditional passive transport, microrobots can achieve targeted drug delivery, enhancing the percentage of drug reaching the diseased site while simultaneously reducing the side effects on normal tissues. The delivered cargoes can be cells, drugs, or other nanoparticles [99,166,194]. Magnetic microrobots have numerous advantages that make them effective carriers for targeted drug delivery. These advantages include the previously mentioned untethered operation, precision, minimal invasiveness, and the recyclability of the robots to minimize residual effects on the human body as much as possible [195,196,197].
Developed for encapsulating or transporting cargoes, magnetic microrobots encompass a variety of micromaterials. These include various organic or inorganic artificial and biogenic materials, such as hydrogel-based helical microswimmers [41], Janus microparticles [62], bacteria [86,176,198], sperm cells [199], and microalgae [200,201].
In 2008, Burdick et al. designed magnetic nanomotors containing nickel that were capable of magnetically controlled cargo manipulation. These nanomotors can achieve loading, dragging, and release of cargoes along their path under the influence of a magnetic field [202]. In 2015, Denzer et al. conjugated antibodies with drug molecules on the surface of magnetic Janus microspheres. This Janus microsphere can precisely target and deliver drugs to cancer cells [203]. Subsequently, Ceylan et al. designed a double-helical structure microrobot capable of cargo loading under the manipulation of a rotating magnetic field [41]. Recently, Akolpoglu et al. reported a miniature robot based on Escherichia coli (E. coli) that can target and release drugs through a multi-stimulus response [204].

6.2. Minimally Invasive Surgery

Miniature robots can be designed as surgical tools to directly penetrate or retrieve cellular tissues [205]. These freely moving minimally invasive systems can access body tissues beyond the reach of blades and catheters. Moreover, they are expected to reduce the risk of infections and to shorten recovery time [206]. In fact, miniature robots are poised to serve as a complement to existing techniques by enhancing the precision and control capabilities of current surgical robotic tools, thereby augmenting the surgical skills of surgeons. By enhancing the penetrative capability of magnetic fields, miniaturized machines can navigate deep tissues or even capillaries remotely, rendering them a promising approach for minimally invasive surgery [207,208].
Magnetic microrobots equipped with pointed ends or capable of executing corkscrew-like movements under a rotating magnetic field can effectively perform drilling procedures. Leveraging this drilling capability holds tremendous potential for achieving untethered microsurgeries with high precision, particularly for penetrating tissues [209]. Additionally, surface walkers exhibit the capacity to open cell membranes, further expanding the applicability of these microscale systems [210,211].
Wu et al. reported the first microscale propeller capable of penetrating the vitreous humor and reaching the retina. Propelled by an external magnetic field, the spiral propeller with the surface coated by perfluorocarbon reached the retina within 30 min [81]. In 2022, Vyskocil et al. proposed a type of Au/Ag/Ni microrobot that can enter cancer cells and excise a portion of the cytoplasm under the control of a rotating magnetic field [61].

6.3. Cell Manipulation

Manipulating cells involves adjusting their spatial placement, achieving separation from adjacent cells with distinct phenotypes, guiding them to specific target positions, or organizing them in vitro [212,213,214]. Magnetically powered miniaturized robots exhibit the capacity to manipulate a cell in three dimensions and can encompass tasks such as grabbing, conveying, categorizing, secluding, and patterning. They demonstrate outstanding agility and elevated accuracy at the microscale within intricate physiological surroundings while preserving the inherent characteristics of the cells [215]. For example, achieved by magnetically driven micromotors resembling peanuts, creating predefined patterns with cells is entailed using an organized substrate for the cell loading and subsequent delivery process [216]. Additionally, Kim et al. conducted precise manipulation of a microrobot by carrying neurons driven by magnetism, guided it to bridge a discontinuity between two neural clusters, and established connections within fractured neural networks [66].

6.4. Environmental Remediation

In addition to being biocompatible, recoverable, and free from toxins during magnetic manipulation, magnetic microrobots can actively navigate through waterborne pollutants to remove them via capture or degradation [217,218]. For instance, to effectively eliminate both leaked oil and microplastic pollutants simultaneously, a hollow microsubmarine with magnetic properties has been developed using natural sunflower pollen grains as a template [219]. The improved ability of pollutant adsorption in mobile microrobots is attributed to their collective behavior combined with magnetically steered agitation and surpasses that of static microrobots [158].

6.5. Current Challenges and Prospects

From the above discussion, it is evident that magnetic microrobots are primarily applied to the biomedical field, and they have already demonstrated the capability to navigate through various animal tissues, including blood vessels, digestive tracts, lenses, and even neural networks. Currently, most research is focused on functional validation through modeling or in vitro experiments. In the future, there will be increasing emphasis on conducting in vivo experiments to better understand the performance of magnetic microrobots within living organisms, with the ultimate goal of integrating them into clinical medical practice.

7. Conclusions

In conclusion, magnetic microrobots have progressed notably in terms of materials, propulsion, design, fabrication, and application. Figure 6 summarizes the current challenges and future directions in various research areas. The use of biosafe magnetic materials has spurred versatile microrobot development. Various propulsion mechanisms enable precise control and include rotating, gradient, and oscillating magnetic fields. Diverse designs provide tailored solutions, and innovative fabrication techniques showcase creativity. The versatile applications of magnetic microrobots for micro-/nano-manipulation, including drug delivery, highlight their transformative impact in healthcare, environmental management, and beyond.
Despite the successful development of various impressive magnetically actuated microrobots, most are currently limited to the proof-of-concept stage. They can only demonstrate simple functions in artificially designed simulation environments. Practical application in complex biomedical environments remains a significant challenge, indicating a substantial journey ahead.
The promising future applications of magnetic microrobots primarily include several aspects. The first one is cancer therapy. The World Health Organization has classified cancer as the second leading cause of death globally. Microrobots can be designed to locate tumor sites and deliver drugs or therapeutic agents to the tumor area, reducing damage to healthy tissues and improving treatment efficacy. Microrobots can also be designed for in vivo monitoring, diagnosis, or collection of tissue samples for pathological analysis. Additionally, microrobots can play a role in minimally invasive surgery.
The second one is thrombus clearance. Microrobots can be designed to carry drugs or thrombolytics, locate thrombus sites, release drugs, and promote thrombus dissolution. Microrobots can also assist with surgery and can even clean the inner walls of blood vessels regularly to prevent thrombus formation. The third one is the treatment of gastrointestinal diseases. Gastrointestinal (GI) diseases are diverse in type and have varied etiologies, symptoms, and treatment methods, making treatment challenging. Many diseases have similar symptoms, and some require endoscopic or other specialized examinations for diagnosis. Microrobots can be designed as part of an endoscope and can also deliver drugs to specific locations.
Fourth, to achieve the aforementioned applications, microrobots need to be precisely controlled to move through different blood vessels. Therefore, challenges such as resisting blood flow and smoothly passing through capillaries without causing blockages when using magnetic microrobots need to be overcome as part of future development.
Finally, one of the major challenges encountered in the practical application of microrobots is the issue of minimally invasive procedures, particularly concerning the removal of microrobots from biological organisms after they complete their tasks. Besides the requirement for biocompatible materials, designing removal mechanisms poses a complex challenge, especially regarding the localization and detection of microrobots inside the biological organism. It is imperative to ensure that the removal process minimally impacts biological tissues while preserving the integrity of the microrobots.
As we look forward, continued interdisciplinary collaboration and technological innovation are expected to drive further progress in this dynamic field, unlocking new possibilities and refining the capabilities of magnetic microrobots for diverse and impactful applications.

Author Contributions

Conceptualization, R.X. and Q.X.; methodology, R.X. and Q.X.; formal analysis, R.X.; investigation, R.X. and Q.X.; resources, Q.X.; data curation, R.X.; writing—original draft preparation, R.X. and Q.X.; writing—review and editing, R.X. and Q.X.; supervision, Q.X.; project administration, Q.X.; funding acquisition, Q.X. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported in part by the National Natural Science Foundation of China under grant 52175556, the Macao Science and Technology Development Fund under grant 0004/2022/AKP and 0102/2022/A2, the Research Committee of the University of Macau under grants MYRG2022-00068-FST and MYRG-CRG2022-00004-FST-ICI, and the Guangdong Basic and Applied Basic Research Foundation under grant 2023A1515011178.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Miskin, M.Z.; Cortese, A.J.; Dorsey, K.; Esposito, E.P.; Reynolds, M.F.; Liu, Q.; Cao, M.; Muller, D.A.; McEuen, P.L.; Cohen, I. Electronically integrated, mass-manufactured, microscopic robots. Nature 2020, 584, 557–561. [Google Scholar] [CrossRef] [PubMed]
  2. Palagi, S.; Fischer, P. Bioinspired microrobots. Nat. Rev. Mater. 2018, 3, 113–124. [Google Scholar] [CrossRef]
  3. Peyer, K.E.; Zhang, L.; Nelson, B.J. Bio-inspired magnetic swimming microrobots for biomedical applications. Nanoscale 2013, 5, 1259–1272. [Google Scholar] [CrossRef] [PubMed]
  4. Medina-Sánchez, M.; Schmidt, O.G. Medical microbots need better imaging and control. Nature 2017, 545, 406–408. [Google Scholar] [CrossRef] [PubMed]
  5. Yang, G.; Bellingham, J.; Dupont, P.E.; Fischer, P.; Floridi, L.; Full, R.; Jacobstein, N.; Kumar, V.; McNutt, M.; Merrifield, R.; et al. The grand challenges of science robotics. Sci. Robot. 2018, 3, eaar7650. [Google Scholar] [CrossRef]
  6. Lyu, Z.; Cao, Y.; Wang, M.Y.; Xu, Q. Concept design of a monolithic compliant series-elastic actuator with integrated position and two-level force sensing. Mech. Mach. Theory 2024, 193, 105569. [Google Scholar] [CrossRef]
  7. Guo, Z.; Ai, N.; Ge, W.; Xu, Q. Design of an automated robotic microinjection system for batch injection of zebrafish embryos and larvae. Microsyst. Nanoeng. 2024, 10, 20. [Google Scholar] [CrossRef] [PubMed]
  8. Solovev, A.A.; Xi, W.; Gracias, D.H.; Harazim, S.M.; Deneke, C.; Sanchez, S.; Schmidt, O.G. Self-propelled nanotools. ACS Nano 2012, 6, 1751–1756. [Google Scholar] [CrossRef] [PubMed]
  9. Wu, Z.; Chen, Y.; Mukasa, D.; Pak, O.S.; Gao, W. Medical micro/nanorobots in complex media. Chem. Soc. Rev. 2020, 49, 8088–8112. [Google Scholar] [CrossRef]
  10. de Ávila, B.E.; Angsantikul, P.; Li, J.; Lopez-Ramirez, M.A.; Ramirez-Herrera, D.E.; Thamphiwatana, S.; Chen, C.; Delezuk, J.; Samakapiruk, R.; Ramez, V.; et al. Micromotor-enabled active drug delivery for in vivo treatment of stomach infection. Nat. Commun. 2017, 8, 272. [Google Scholar] [CrossRef]
  11. Sitti, M.; Ceylan, H.; Hu, W.; Giltinan, J.; Turan, M.; Yim, S.; Diller, E. Biomedical applications of untethered mobile milli/microrobots. Proc. IEEE 2015, 103, 205–224. [Google Scholar] [CrossRef] [PubMed]
  12. Hoshiar, A.K.; Jeon, S.; Kim, K.; Lee, S.; Kim, J.; Choi, H. Steering algorithm for a flexible microrobot to enhance guidewire control in a coronary angioplasty application. Micromachines 2018, 9, 617. [Google Scholar] [CrossRef] [PubMed]
  13. Zhou, H.; Mayorga-Martinez, C.C.; Pané, S.; Zhang, L.; Pumera, M. Magnetically driven micro and nanorobots. Chem. Rev. 2021, 121, 4999–5041. [Google Scholar] [CrossRef] [PubMed]
  14. Ceylan, H.; Giltinan, J.; Kozielski, K.; Sitti, M. Mobile microrobots for bioengineering applications. Lab Chip 2017, 17, 1705–1724. [Google Scholar] [CrossRef] [PubMed]
  15. Chen, X.; Hoop, M.; Shamsudhin, N.; Huang, T.; Özkale, B.; Li, Q.; Siringil, E.; Mushtaq, F.; Tizio, L.D.; Nelson, B.J.; et al. Hybrid magnetoelectric nanowires for nanorobotic applications: Fabrication, magnetoelectric coupling, and magnetically assisted in vitro targeted drug delivery. Adv. Mater. 2017, 29, 1605458. [Google Scholar] [CrossRef] [PubMed]
  16. Fan, D.; Yin, Z.; Cheong, R.; Zhu, F.Q.; Cammarata, R.C.; Chien, C.L.; Levchenko, A. Subcellular-resolution delivery of a cytokine through precisely manipulated nanowires. Nat. Nanotechnol. 2010, 5, 545–551. [Google Scholar] [CrossRef] [PubMed]
  17. Yang, Z.; Zhang, L. Magnetic actuation systems for miniature robots: A review. Adv. Intell. Syst. 2020, 2, 2000082. [Google Scholar] [CrossRef]
  18. Fischer, P.; Ghosh, A. Magnetically actuated propulsion at low Reynolds numbers: Towards nanoscale control. Nanoscale 2011, 3, 557–563. [Google Scholar] [CrossRef] [PubMed]
  19. Mirzae, Y.; Dubrovski, O.; Kenneth, O.; Morozov, K.I.; Leshansky, A.M. Geometric constraints and optimization in externally driven propulsion. Sci. Robot. 2018, 3, eaas8713. [Google Scholar] [CrossRef]
  20. Heunis, C.; Sikorski, J.; Misra, S. Flexible instruments for endovascular interventions: Improved magnetic steering, actuation, and image-guided surgical instruments. IEEE Robot. Autom. Mag. 2018, 25, 71–82. [Google Scholar] [CrossRef]
  21. Wang, W.; Duan, W.; Ahmed, S.; Mallouk, T.E.; Sen, A. Small power: Autonomous nano-and micromotors propelled by self-generated gradients. Nano Today 2013, 8, 531–554. [Google Scholar] [CrossRef]
  22. Wang, J. Can man-made nanomachines compete with nature biomotors? ACS Nano 2009, 3, 4–9. [Google Scholar] [CrossRef] [PubMed]
  23. Sánchez, S.; Soler, L.; Katuri, J. Chemically powered micro-and nanomotors. Angew. Chem. Int. Ed. 2015, 54, 1414–1444. [Google Scholar] [CrossRef]
  24. Kim, Y.; Zhao, X. Magnetic soft materials and robots. Chem. Rev. 2022, 122, 5317–5364. [Google Scholar] [CrossRef]
  25. Chung, H.-J.; Parsons, A.M.; Zheng, L. Magnetically controlled soft robotics utilizing elastomers and gels in actuation: A review. Adv. Intell. Syst. 2021, 3, 2000186. [Google Scholar] [CrossRef]
  26. Wu, S.; Hu, W.; Ze, Q.; Sitti, M.; Zhao, R. Multifunctional magnetic soft composites: A review. Multifunct. Mater. 2020, 3, 042003. [Google Scholar] [CrossRef]
  27. Xu, Z.; Chen, Y.; Xu, Q. Spreadable magnetic soft robots with on-demand hardening. Research 2023, 6, 0262. [Google Scholar] [CrossRef]
  28. Chen, X.Z.; Jang, B.; Ahmed, D.; Hu, C.; De Marco, C.; Hoop, M.; Mushtaq, F.; Nelson, B.J.; Pané, S. Small-scale machines driven by external power sources. Adv. Mater. 2018, 30, 1705061. [Google Scholar] [CrossRef] [PubMed]
  29. Xu, T.; Gao, W.; Xu, L.-P.; Zhang, X.; Wang, S. Fuel-free synthetic micro-/nanomachines. Adv. Mater. 2017, 29, 1603250. [Google Scholar] [CrossRef]
  30. Zhang, Z.; Wang, X.; Liu, J.; Dai, C.; Sun, Y. Robotic micromanipulation: Fundamentals and applications. Annu. Rev. Control. Robot. Auton. Syst. 2019, 2, 181–203. [Google Scholar] [CrossRef]
  31. Urso, M.; Ussia, M.; Peng, X.; Oral, C.M.; Pumera, M. Reconfigurable self-assembly of photocatalytic magnetic microrobots for water purification. Nat. Commun. 2023, 14, 6969. [Google Scholar] [CrossRef] [PubMed]
  32. Maria-Hormigos, R.; Mayorga-Martinez, C.C.; Pumera, M. Soft magnetic microrobots for photoactive pollutant removal. Small Methods 2023, 7, 2201014. [Google Scholar] [CrossRef] [PubMed]
  33. Mayorga-Martinez, C.C.; Castoralova, M.; Zelenka, J.; Ruml, T.; Pumera, M. Swarming Magnetic Microrobots for Pathogen Isolation from Milk. Small 2023, 19, 2205047. [Google Scholar] [CrossRef]
  34. Bi, C.; Guix, M.; Johnson, B.V.; Jing, W.; Cappelleri, D.J. Design of microscale magnetic tumbling robots for locomotion in multiple environments and complex terrains. Micromachines 2018, 9, 68. [Google Scholar] [CrossRef] [PubMed]
  35. Kim, Y.; Yuk, H.; Zhao, R.; Chester, S.A.; Zhao, X. Printing ferromagnetic domains for untethered fast-transforming soft materials. Nature 2018, 558, 274–279. [Google Scholar] [CrossRef]
  36. Kittel, C. Physical theory of ferromagnetic domains. Rev. Mod. Phys. 1949, 21, 541. [Google Scholar] [CrossRef]
  37. Rikken, R.S.M.; Nolte, R.J.M.; Maan, J.C.; van Hest, J.C.M.; Wilson, D.A.; Christianen, P.C.M. Manipulation of micro- and nanostructure motion with magnetic fields. Soft Matter 2014, 10, 1295–1308. [Google Scholar] [CrossRef]
  38. Jungwirth, T.; Marti, X.; Wadley, P.; Wunderlich, J. Antiferromagnetic spintronics. Nat. Nanotechnol. 2016, 11, 231–241. [Google Scholar] [CrossRef]
  39. Xu, T.; Hao, Z.; Huang, C.; Yu, J.; Zhang, L.; Wu, X. Multimodal locomotion control of needle-like microrobots assembled by ferromagnetic nanoparticles. IEEE/ASME Trans. Mechatron. 2022, 27, 4327–4338. [Google Scholar] [CrossRef]
  40. Yu, J.; Xu, T.; Lu, Z.; Vong, C.I.; Zhang, L. On-demand disassembly of paramagnetic nanoparticle chains for microrobotic cargo delivery. IEEE Trans. Robot. 2017, 33, 1213–1225. [Google Scholar] [CrossRef]
  41. Ceylan, H.; Yasa, I.C.; Yasa, O.; Tabak, A.F.; Giltinan, J.; Sitti, M. 3D-printed biodegradable microswimmer for theranostic cargo delivery and release. ACS Nano 2019, 13, 3353–3362. [Google Scholar] [CrossRef]
  42. Cui, J.; Huang, T.; Luo, Z.; Testa, P.; Gu, H.; Chen, X.; Nelson, B.J.; Heyderman, L.J. Nanomagnetic encoding of shape-morphing micromachines. Nature 2019, 575, 164–168. [Google Scholar] [CrossRef]
  43. Gossuin, Y.; Gillis, P.; Hocq, A.; Vuong, Q.L.; Roch, A. Magnetic resonance relaxation properties of superparamagnetic particles. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol. 2009, 1, 299–310. [Google Scholar] [CrossRef] [PubMed]
  44. Xie, H.; Sun, M.; Fan, X.; Lin, Z.; Chen, W.; Wang, L.; Dong, L.; He, Q. Reconfigurable magnetic microrobot swarm: Multimode transformation, locomotion, and manipulation. Sci. Robot. 2019, 4, eaav8006. [Google Scholar] [CrossRef] [PubMed]
  45. Fan, X.; Dong, X.; Karacakol, A.C.; Xie, H.; Sitti, M. Reconfigurable multifunctional ferrofluid droplet robots. Proc. Natl. Acad. Sci. USA 2020, 117, 27916–27926. [Google Scholar] [CrossRef]
  46. Sun, M.; Yang, S.; Jiang, J.; Jiang, S.; Sitti, M.; Zhang, L. Bioinspired self-assembled colloidal collectives drifting in three dimensions underwater. Sci. Adv. 2023, 9, eadj4201. [Google Scholar] [CrossRef]
  47. Fan, X.; Zhang, Y.; Wu, Z.; Xie, H.; Sun, L.; Chen, T.; Yang, Z. Combined three dimensional locomotion and deformation of functional ferrofluidic robots. Nanoscale 2023, 15, 19499–19513. [Google Scholar] [CrossRef] [PubMed]
  48. Song, Y.; Hormes, J.; Kumar, C.S.S.R. Microfluidic synthesis of nanomaterials. Small 2008, 4, 698–711. [Google Scholar] [CrossRef]
  49. Gu, H.; Duits, M.H.G.; Mugele, F. Droplets formation and merging in two-phase flow microfluidics. Int. J. Mol. Sci. 2011, 12, 2572–2597. [Google Scholar] [CrossRef]
  50. Chiu, W.; Watanabe, Y.; Tahara, M.; Inamura, T.; Hosoda, H. Investigations of Shape Deformation Behaviors of the Ferromagnetic Ni–Mn–Ga Alloy/Porous Silicone Rubber Composite towards Actuator Applications. Micromachines 2023, 14, 1604. [Google Scholar] [CrossRef]
  51. Hou, Y.; Dai, Y.; Zhang, W.; Wang, M.; Zhao, H.; Feng, L. Ultrasound-Based Real-Time Imaging of Hydrogel-Based Millirobots with Volume Change Capability. Micromachines 2023, 14, 422. [Google Scholar] [CrossRef] [PubMed]
  52. Zhao, F.; Rong, W.; Wang, L.; Sun, L. Photothermal-Responsive Shape-Memory Magnetic Helical Microrobots with Programmable Addressable Shape Changes. ACS Appl. Mater. Interfaces 2023, 15, 25942–25951. [Google Scholar] [CrossRef] [PubMed]
  53. Jia, Y.; Zhu, Z.; Jing, X.; Lin, J.; Lu, M. Fabrication and performance evaluation of magnetically driven double curved conical ribbon micro-helical robot. Mater. Des. 2023, 226, 111651. [Google Scholar] [CrossRef]
  54. Zhao, F.; Rong, W.; Li, D.; Wang, L.; Sun, L. Four-dimensional design and programming of shape-memory magnetic helical micromachines. Appl. Mater. Today 2022, 27, 101422. [Google Scholar] [CrossRef]
  55. Lee, H.; Kim, D.; Kwon, S.; Park, S. Magnetically actuated drug delivery helical microrobot with magnetic nanoparticle retrieval ability. ACS Appl. Mater. Interfaces 2021, 13, 19633–19647. [Google Scholar] [CrossRef] [PubMed]
  56. Kadiri, V.M.; Bussi, C.; Holle, A.W.; Son, K.; Kwon, H.; Schütz, G.; Gutierrez, M.G.; Fischer, P. Biocompatible magnetic micro- and nanodevices: Fabrication of FePt nanopropellers and cell transfection. Adv. Mater. 2020, 32, 2001114. [Google Scholar] [CrossRef] [PubMed]
  57. Schwarz, L.; Karnaushenko, D.D.; Hebenstreit, F.; Naumann, R.; Schmidt, O.G.; Medina-Sánchez, M. A rotating spiral micromotor for noninvasive zygote transfer. Adv. Sci. 2020, 7, 2000843. [Google Scholar] [CrossRef] [PubMed]
  58. Wang, X.; Chen, X.; Alcântara, C.C.J.; Sevim, S.; Hoop, M.; Terzopoulou, A.; Marco, C.D.; Hu, C.; de Mello, A.J.; Falcaro, P.; et al. MOFBOTS: Metal–organic-framework-based biomedical microrobots. Adv. Mater. 2019, 31, 1901592. [Google Scholar] [CrossRef] [PubMed]
  59. Park, J.; Jin, C.; Lee, S.; Kim, J.; Choi, H. Magnetically actuated degradable microrobots for actively controlled drug release and hyperthermia therapy. Adv. Healthc. Mater. 2019, 8, 1900213. [Google Scholar] [CrossRef]
  60. Chen, W.; Wen, Y.; Fan, X.; Sun, M.; Tian, C.; Yang, M.; Xie, H. Magnetically actuated intelligent hydrogel-based child-parent microrobots for targeted drug delivery. J. Mater. Chem. B 2021, 9, 1030–1039. [Google Scholar] [CrossRef]
  61. Vyskocil, J.; Mayorga-Martinez, C.C.; Jablonska, E.; Novotny, F.; Ruml, T.; Pumera, M. Cancer cells microsurgery via asymmetric bent surface Au/Ag/Ni microrobotic scalpels through a transversal rotating magnetic field. ACS Nano 2020, 14, 8247–8256. [Google Scholar] [CrossRef] [PubMed]
  62. Alapan, Y.; Bozuyuk, U.; Erkoc, P.; Karacakol, A.C.; Sitti, M. Multifunctional surface microrollers for targeted cargo delivery in physiological blood flow. Sci. Robot. 2020, 5, eaba5726. [Google Scholar] [CrossRef] [PubMed]
  63. Villa, K.; Krejčová, L.; Novotný, F.; Heger, Z.; Sofer, Z.; Pumera, M. Cooperative multifunctional self-propelled paramagnetic microrobots with chemical handles for cell manipulation and drug delivery. Adv. Funct. Mater. 2018, 28, 1804343. [Google Scholar] [CrossRef]
  64. Gu, H.; Boehler, Q.; Cui, H.; Secchi, E.; Savorana, G.; Marco, C.D.; Gervasoni, S.; Peyron, Q.; Huang, T.; Pane, S.; et al. Magnetic cilia carpets with programmable metachronal waves. Nat. Commun. 2020, 11, 2637. [Google Scholar] [CrossRef] [PubMed]
  65. 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]
  66. Kim, E.; Jeon, S.; An, H.; Kianpour, M.; Yu, S.; Kim, J.; Rah, J.; Choi, H. A magnetically actuated microrobot for targeted neural cell delivery and selective connection of neural networks. Sci. Adv. 2020, 6, eabb5696. [Google Scholar] [CrossRef] [PubMed]
  67. Gyak, K.; Jeon, S.; Ha, L.; Kim, S.; Kim, J.; Lee, K.; Choi, H.; Kim, D. Magnetically Actuated SiCN-Based Ceramic Microrobot for Guided Cell Delivery. Adv. Healthc. Mater. 2019, 8, 1900739. [Google Scholar] [CrossRef] [PubMed]
  68. Li, J.; Li, X.; Luo, T.; Wang, R.; Liu, C.; Chen, S.; Li, D.; Yue, J.; Cheng, S.; Sun, D. Development of a magnetic microrobot for carrying and delivering targeted cells. Sci. Robot. 2018, 3, eaat8829. [Google Scholar] [CrossRef] [PubMed]
  69. Zhong, D.; Li, W.; Qi, Y.; He, J.; Zhou, M. Photosynthetic biohybrid nanoswimmers system to alleviate tumor hypoxia for FL/PA/MR imaging-guided enhanced radio-photodynamic synergetic therapy. Adv. Funct. Mater. 2020, 30, 1910395. [Google Scholar] [CrossRef]
  70. Liu, L.; Chen, B.; Liu, K.; Gao, J.; Ye, Y.; Wang, Z.; Qin, N.; Wilson, D.A.; Tu, Y.; Peng, F. Wireless manipulation of magnetic/piezoelectric micromotors for precise neural stem-like cell stimulation. Adv. Funct. Mater. 2020, 30, 1910108. [Google Scholar] [CrossRef]
  71. Gao, C.; Lin, Z.; Wang, D.; Wu, Z.; Xie, H.; He, Q. Red blood cell-mimicking micromotor for active photodynamic cancer therapy. ACS Appl. Mater. Interfaces 2019, 11, 23392–23400. [Google Scholar] [CrossRef] [PubMed]
  72. Sun, M.; Fan, X.; Meng, X.; Song, J.; Chen, W.; Sun, L.; Xie, H. Magnetic biohybrid micromotors with high maneuverability for efficient drug loading and targeted drug delivery. Nanoscale 2019, 11, 18382–18392. [Google Scholar] [CrossRef] [PubMed]
  73. Jeon, S.; Kim, S.; Ha, S.; Lee, S.; Kim, E.; Kim, S.Y.; Park, S.H.; Jeon, J.H.; Kim, S.W.; Moon, C.; et al. Magnetically actuated microrobots as a platform for stem cell transplantation. Sci. Robot. 2019, 4, eaav4317. [Google Scholar] [CrossRef] [PubMed]
  74. Ramcharitar, S.; Patterson, M.S.; Geuns, R.J.V.; Meighem, C.V.; Serruys, P.W. Technology insight: Magnetic navigation in coronary interventions. Nat. Clin. Pract. Cardiovasc. Med. 2008, 5, 148–156. [Google Scholar] [CrossRef] [PubMed]
  75. Schiemann, M.; Killmann, R.; Kleen, M.; Abolmaali, N.; Finney, J.; Vogl, T.J. Vascular guide wire navigation with a magnetic guidance system: Experimental results in a phantom. Radiology 2004, 232, 475–481. [Google Scholar] [CrossRef] [PubMed]
  76. Fernandes, R.; Gracias, D.H. Toward a miniaturized mechanical surgeon. Mater. Today 2009, 12, 14–20. [Google Scholar] [CrossRef]
  77. Ou, J.; Liu, K.; Jiang, J.; Wilson, D.A.; Liu, L.; Wang, F.; Wang, S.; Tu, Y.; Peng, F. Micro-/nanomotors toward biomedical applications: The recent progress in biocompatibility. Small 2020, 16, 1906184. [Google Scholar] [CrossRef] [PubMed]
  78. Wang, S.; Liu, X.; Wang, Y.; Xu, D.; Liang, C.; Guo, J.; Ma, X. Biocompatibility of artificial micro/nanomotors for use in biomedicine. Nanoscale 2019, 11, 14099–14112. [Google Scholar] [CrossRef] [PubMed]
  79. Ongaro, F.; Niehoff, D.; Mohanty, S.; Misra, S. A contactless and biocompatible approach for 3D active microrobotic targeted drug delivery. Micromachines 2019, 10, 504. [Google Scholar] [CrossRef]
  80. Venugopalan, P.L.; Sai, R.; Chandorkar, Y.; Basu, B.; Shivashankar, S.; Ghosh, A. Conformal cytocompatible ferrite coatings facilitate the realization of a nanovoyager in human blood. Nano Lett. 2014, 14, 1968–1975. [Google Scholar] [CrossRef]
  81. Wu, Z.; Troll, J.; Jeong, H.; Wei, Q.; Stang, M.; Ziemssen, F.; Wang, Z.; Dong, M.; Schnichels, S.; Qiu, T.; et al. A swarm of slippery micropropellers penetrates the vitreous body of the eye. Sci. Adv. 2018, 4, eaat4388. [Google Scholar] [CrossRef] [PubMed]
  82. Walker, D.; Käsdorf, B.T.; Jeong, H.; Lieleg, O.; Fischer, P. Enzymatically active biomimetic micropropellers for the penetration of mucin gels. Sci. Adv. 2015, 1, e1500501. [Google Scholar] [CrossRef] [PubMed]
  83. Peters, C.; Hoop, M.; Pané, S.; Nelson, B.J.; Hierold, C. Degradable magnetic composites for minimally invasive interventions: Device fabrication, targeted drug delivery, and cytotoxicity tests. Adv. Mater. 2016, 28, 533–538. [Google Scholar] [CrossRef] [PubMed]
  84. Mair, L.O.; Chowdhury, S.; Paredes-Juarez, G.A.; Guix, M.; Bi, C.; Johnson, B.; English, B.W.; Jafari, S.; Baker-McKee, J.; Watson-Daniels, J.; et al. Magnetically aligned nanorods in alginate capsules (MANiACs): Soft matter tumbling robots for manipulation and drug delivery. Micromachines 2019, 10, 230. [Google Scholar] [CrossRef] [PubMed]
  85. Magdanz, V.; Sanchez, S.; Schmidt, O.G. Development of a sperm-flagella driven micro-bio-robot. Adv. Mater. 2013, 25, 6581–6588. [Google Scholar] [CrossRef] [PubMed]
  86. Felfoul, O.; Mohammadi, M.; Taherkhani, S.; Lanauze, D.D.; Xu, Y.Z.; Loghin, D.; Essa, S.; Jancik, S.; Houle, D.; Lafleur, M.; et al. Magneto-aerotactic bacteria deliver drug-containing nanoliposomes to tumour hypoxic regions. Nat. Nanotechnol. 2016, 11, 941–947. [Google Scholar] [CrossRef] [PubMed]
  87. Wu, Z.; Li, J.; de Ávila, B.E.; Li, T.; Gao, W.; He, Q.; Zhang, L.; Wang, J. Water-powered cell-mimicking Janus micromotor. Adv. Funct. Mater. 2015, 25, 7497–7501. [Google Scholar] [CrossRef]
  88. de Ávila, B.E.; Lopez-Ramirez, M.A.; Mundaca-Uribe, R.; Wei, X.; Ramírez-Herrera, D.E.; Karshalev, E.; Nguyen, B.; Fang, R.H.; Zhang, L.; Wang, J. Multicompartment tubular micromotors toward enhanced localized active delivery. Adv. Mater. 2020, 32, 2000091. [Google Scholar] [CrossRef] [PubMed]
  89. Wei, X.; Beltrán-Gastélum, M.; Karshalev, E.; de Ávila, B.E.; Zhou, J.; Ran, D.; Angsantikul, P.; Fang, R.H.; Wang, J.; Zhang, L. Biomimetic micromotor enables active delivery of antigens for oral vaccination. Nano Lett. 2019, 19, 1914–1921. [Google Scholar] [CrossRef]
  90. Lu, X.; Ambulo, C.P.; Wang, S.; Rivera-Tarazona, L.K.; Kim, H.; Searles, K.; Ware, T.H. 4D-printing of photoswitchable actuators. Angew. Chem. Int. Ed. 2021, 60, 5536–5543. [Google Scholar] [CrossRef]
  91. Gunn, J.W.; Turner, S.D.; Mann, B.K. Adhesive and mechanical properties of hydrogels influence neurite extension. J. Biomed. Mater. Res. Part Off. J. Soc. Biomater. Jpn. Soc. Biomater. Aust. Soc. Biomater. Korean Soc. Biomater. 2005, 72, 91–97. [Google Scholar] [CrossRef] [PubMed]
  92. Zhang, L.; Xu, L.; Wang, Y.; Liu, J.; Tan, G.; Huang, F.; He, N.; Lu, Z. A novel therapeutic vaccine based on graphene oxide nanocomposite for tumor immunotherapy. Chin. Chem. Lett. 2022, 33, 4089–4095. [Google Scholar] [CrossRef]
  93. Nemir, S.; Hayenga, H.N.; West, J.L. PEGDA hydrogels with patterned elasticity: Novel tools for the study of cell response to substrate rigidity. Biotechnol. Bioeng. 2010, 105, 636–644. [Google Scholar] [CrossRef]
  94. Liu, J.; Yu, S.; Xu, B.; Tian, Z.; Zhang, H.; Liu, K.; Shi, X.; Zhao, Z.; Liu, C.; Lin, X.; et al. Magnetically propelled soft microrobot navigating through constricted microchannels. Appl. Mater. Today 2021, 25, 101237. [Google Scholar] [CrossRef]
  95. Srivastava, S.K.; Ajalloueian, F.; Boisen, A. Thread-Like Radical-Polymerization via Autonomously Propelled (TRAP) Bots. Adv. Mater. 2019, 31, 1901573. [Google Scholar] [CrossRef] [PubMed]
  96. Xu, R.; Xu, Q. Design of a Bio-Inspired Untethered Soft Octopodal Robot Driven by Magnetic Field. Biomimetics 2023, 8, 269. [Google Scholar] [CrossRef]
  97. Ziolkowska, K.; Jedrych, E.; Kwapiszewski, R.; Lopacinska, J.; Skolimowski, M.; Chudy, M. PDMS/glass microfluidic cell culture system for cytotoxicity tests and cells passage. Sens. Actuators Chem. 2010, 145, 533–542. [Google Scholar] [CrossRef]
  98. Chen, S.; Huang, S.; Wu, H.; Pan, W.; Wei, S.; Peng, C.; Ni, I.; Murti, B.T.; Tsai, M.; Wu, C.; et al. A Facile, Fabric Compatible, and Flexible Borophene Nanocomposites for Self-Powered Smart Assistive and Wound Healing Applications. Adv. Sci. 2022, 9, 2201507. [Google Scholar] [CrossRef]
  99. Yan, X.; Zhou, Q.; Yu, J.; Xu, T.; Deng, Y.; Tang, T.; Feng, Q.; Bian, L.; Zhang, Y.; Ferreira, A.; et al. Magnetite nanostructured porous hollow helical microswimmers for targeted delivery. Adv. Funct. Mater. 2015, 25, 5333–5342. [Google Scholar] [CrossRef]
  100. Lin, J.; Chiang, M. Hysteresis analysis and positioning control for a magnetic shape memory actuator. Sensors 2015, 15, 8054–8071. [Google Scholar] [CrossRef]
  101. Sadeghzadeh, A.; Asua, E.; Feuchtwanger, J.; Etxebarria, V.; García-Arribas, A. Ferromagnetic shape memory alloy actuator enabled for nanometric position control using hysteresis compensation. Sens. Actuators Phys. 2012, 182, 122–129. [Google Scholar] [CrossRef]
  102. Riccardi, L.; Naso, D.; Turchiano, B.; Janocha, H. Adaptive control of positioning systems with hysteresis based on magnetic shape memory alloys. IEEE Trans. Control. Syst. Technol. 2012, 21, 2011–2023. [Google Scholar] [CrossRef]
  103. Abbott, J.J.; Nagy, Z.; Beyeler, F.; Nelson, B.J. Robotics in the small, part I: Microbotics. IEEE Robot. Autom. Mag. 2007, 14, 92–103. [Google Scholar] [CrossRef]
  104. Fountain, T.W.R.; Kailat, P.V.; Abbott, J.J. Wireless control of magnetic helical microrobots using a rotating-permanent-magnet manipulator. In Proceedings of the 2010 IEEE International Conference on Robotics and Automation, Anchorage, AK, USA, 3–7 May 2010; pp. 576–581. [Google Scholar]
  105. Wu, Z.; Zhang, Y.; Chi, Z.; Xu, Q. Design and development of a new rotating electromagnetic field generation system for driving microrobots. IEEE Trans. Magn. 2021, 58, 1–8. [Google Scholar] [CrossRef]
  106. Wu, Z.; Zhang, Y.; Ai, N.; Chen, H.; Ge, W.; Xu, Q. Magnetic mobile microrobots for upstream and downstream navigation in biofluids with variable flow rate. Adv. Intell. Syst. 2022, 4, 2100266. [Google Scholar] [CrossRef]
  107. Kikuchi, K.; Yamazaki, A.; Sendoh, M.; Ishiyama, K.; Arai, K.I. Fabrication of a spiral type magnetic micromachine for trailing a wire. IEEE Trans. Magn. 2005, 41, 4012–4014. [Google Scholar] [CrossRef]
  108. Gao, W.; Feng, X.; Pei, A.; Kane, C.R.; Tam, R.; Hennessy, C.; Wang, J. Bioinspired helical microswimmers based on vascular plants. Nano Lett. 2014, 14, 305–310. [Google Scholar] [CrossRef] [PubMed]
  109. Fu, Q.; Guo, S.; Zhang, S.; Hirata, H.; Ishihara, H. Characteristic evaluation of a shrouded propeller mechanism for a magnetic actuated microrobot. Micromachines 2015, 6, 1272–1288. [Google Scholar] [CrossRef]
  110. Abbott, J.J.; Peyer, K.E.; Lagomarsino, M.C.; Zhang, L.; Dong, L.; Kaliakatsos, I.K.; Nelson, B.J. How should microrobots swim? Int. J. Robot. Res. 2009, 28, 1434–1447. [Google Scholar] [CrossRef]
  111. Honda, T.; Arai, K.I.; Ishiyama, K. Micro swimming mechanisms propelled by external magnetic fields. IEEE Trans. Magn. 1996, 32, 5085–5087. [Google Scholar] [CrossRef]
  112. Zhang, L.; Abbott, J.J.; Dong, L.; Kratochvil, B.E.; Bell, D.; Nelson, B.J. Artificial bacterial flagella: Fabrication and magnetic control. Appl. Phys. Lett. 2009, 94, 064107. [Google Scholar] [CrossRef]
  113. Ahmed, D.; Baasch, T.; Blondel, N.; Läubli, N.; Dual, J.; Nelson, B.J. Neutrophil-inspired propulsion in a combined acoustic and magnetic field. Nat. Commun. 2017, 8, 770. [Google Scholar] [CrossRef]
  114. Alapan, Y.; Yigit, B.; Beker, O.; Demirörs, A.F.; Sitti, M. Shape-encoded dynamic assembly of mobile micromachines. Nat. Mater. 2019, 18, 1244–1251. [Google Scholar] [CrossRef]
  115. Ryan, P.; Diller, E. Five-degree-of-freedom magnetic control of micro-robots using rotating permanent magnets. In Proceedings of the 2016 IEEE International Conference on Robotics and Automation (ICRA), Stockholm, Sweden, 16–21 May 2016; pp. 1731–1736. [Google Scholar]
  116. Grady, M.S.; Howard, M.A., III; Molloy, J.A.; Ritter, R.C.; Quate, E.G.; Gillies, G.T. Nonlinear magnetic stereotaxis: Three-dimensional, in vivo remote magnetic manipulation of a small object in canine brain. Med. Phys. 1990, 17, 405–415. [Google Scholar] [CrossRef]
  117. Go, G.; Choi, H.; Jeong, S.; Lee, C.; Ko, S.Y.; Park, J.; Park, S. Electromagnetic navigation system using simple coil structure (4 coils) for 3-D locomotive microrobot. IEEE Trans. Magn. 2014, 51, 1–7. [Google Scholar]
  118. Chowdhury, S.; Jing, W.; Cappelleri, D.J. Towards independent control of multiple magnetic mobile microrobots. Micromachines 2015, 7, 3. [Google Scholar] [CrossRef]
  119. Meeker, D.C.; Maslen, E.H.; Ritter, R.C.; Creighton, F.M. Optimal realization of arbitrary forces in a magnetic stereotaxis system. IEEE Trans. Magn. 1996, 32, 320–328. [Google Scholar] [CrossRef]
  120. Diller, E.; Giltinan, J.; Lum, G.Z.; Ye, Z.; Sitti, M. Six-degree-of-freedom magnetic actuation for wireless microrobotics. Int. J. Robot. Res. 2016, 35, 114–128. [Google Scholar] [CrossRef]
  121. Klumpp, S.; Lefèvre, C.T.; Bennet, M.; Faivre, D. Swimming with magnets: From biological organisms to synthetic devices. Phys. Rep. 2019, 789, 1–54. [Google Scholar] [CrossRef]
  122. Wu, Z.; Xu, Z.; Xu, Q. Design and optimization of a new alternating electromagnetic-field-generation system for an inverted microscope. Micromachines 2022, 13, 542. [Google Scholar] [CrossRef]
  123. Cheang, U.K.; Ali, J.; Kim, H.; Rogowski, L.; Kim, M.J. On-surface locomotion of particle based microrobots using magnetically induced oscillation. Micromachines 2017, 8, 46. [Google Scholar] [CrossRef]
  124. Shamsudhin, N.; Zverev, V.I.; Keller, H.; Pane, S.; Egolf, P.W.; Nelson, B.J.; Tishin, A.M. Magnetically guided capsule endoscopy. Med Phys. 2017, 44, e91–e111. [Google Scholar] [CrossRef]
  125. Zhang, Y.; Wu, Z.; Chi, Z.; Xu, Q. Design and Testing of a Rotational Magnetic System for Driving Helical Swimmer in Circular Duct Environment. In Proceedings of the 2021 6th IEEE International Conference on Advanced Robotics and Mechatronics (ICARM), Chongqing, China, 3–5 July 2021; pp. 693–698. [Google Scholar]
  126. Kummer, M.P.; Abbott, J.J.; Kratochvil, B.E.; Borer, R.; Sengul, A.; Nelson, B.J. OctoMag: An Electromagnetic System for 5-DOF Wireless Micromanipulation. IEEE Trans. Robot. 2010, 26, 1006–1017. [Google Scholar] [CrossRef]
  127. Yesin, K.B.; Vollmers, K.; Nelson, B.J. Modeling and Control of Untethered Biomicrorobots in a Fluidic Environment Using Electromagnetic Fields. Int. J. Robot. Res. 2006, 25, 527–536. [Google Scholar] [CrossRef]
  128. Armacost, M.P.; Adair, J.; Munger, T.; Viswanathan, R.R.; Creighton, F.M.; Curd, D.T.; Sehra, R. Accurate and Reproducible Target Navigation with the Stereotaxis Niobe™ Magnetic Navigation System. J. Cardiovasc. Electrophysiol. 2007, 18, S26–S31. [Google Scholar] [CrossRef]
  129. Liao, Z.; Hou, X.; Hu, E.L.; Sheng, J.; Ge, Z.; Jiang, B.; Hou, X.; Liu, J.; Li, Z.; Huang, Q.; et al. Accuracy of Magnetically Controlled Capsule Endoscopy, Compared with Conventional Gastroscopy, in Detection of Gastric Diseases. Clin. Gastroenterol. Hepatol. 2016, 14, 1266–1273. [Google Scholar] [CrossRef] [PubMed]
  130. Chen, Y.; Chen, D.; Liang, S.; Dai, Y.; Bai, X.; Song, B.; Zhang, D.; Chen, H.; Feng, L. Recent advances in field-controlled micro–nano manipulations and micro–nano robots. Adv. Intell. Syst. 2022, 4, 2100116. [Google Scholar] [CrossRef]
  131. Schuerle, S.; Erni, S.; Flink, M.; Kratochvil, B.E.; Nelson, B.J. Three-dimensional magnetic manipulation of micro-and nanostructures for applications in life sciences. IEEE Trans. Magn. 2012, 49, 321–330. [Google Scholar] [CrossRef]
  132. Xu, Z.; Wu, Z.; Yuan, M.; Chen, Y.; Ge, W.; Xu, Q. Versatile magnetic hydrogel soft capsule microrobots for targeted delivery. iScience 2023, 26, 106727. [Google Scholar] [CrossRef] [PubMed]
  133. Mandal, P.; Patil, G.; Kakoty, H.; Ghosh, A. Magnetic active matter based on helical propulsion. Accounts Chem. Res. 2018, 51, 2689–2698. [Google Scholar] [CrossRef]
  134. Samsami, K.; Mirbagheri, S.A.; Meshkati, F.; Fu, H.C. Stability of soft magnetic helical microrobots. Fluids 2020, 5, 19. [Google Scholar] [CrossRef]
  135. Mohammadi, A.; Spong, M.W. Integral line-of-sight path following control of magnetic helical microswimmers subject to step-out frequencies. Automatica 2021, 128, 109554. [Google Scholar] [CrossRef]
  136. Landers, F.C.; Gantenbein, V.; Hertle, L.; Veciana, A.; Llacer-Wintle, J.; Chen, X.; Ye, H.; Franco, C.; Puigmartí-Luis, J.; Kim, M.; et al. On-Command Disassembly of Microrobotic Superstructures for Transport and Delivery of Magnetic Micromachines. Adv. Mater. 2023, e2310084. [Google Scholar] [CrossRef] [PubMed]
  137. Liu, S.; Chen, B.; Feng, Y.; Gao, C.; Du, D.; Jiang, T.; Tu, Y.; Peng, F. Helical hydrogel micromotors for delivery of neural stem cells and restoration of neural connectivity. Chem. Eng. J. 2024, 479, 147745. [Google Scholar] [CrossRef]
  138. Soto, F.; Wang, J.; Ahmed, R.; Demirci, U. Medical micro/nanorobots in precision medicine. Adv. Sci. 2020, 7, 2002203. [Google Scholar] [CrossRef]
  139. Peng, F.; Tu, Y.; Wilson, D.A. Micro/nanomotors towards in vivo application: Cell, tissue and biofluid. Chem. Soc. Rev. 2017, 46, 5289–5310. [Google Scholar] [CrossRef]
  140. Verma, S.K.; Chauhan, R. Nanorobotics in dentistry–A review. Indian J. Dent. 2014, 5, 62–70. [Google Scholar] [CrossRef]
  141. Evans, B.A.; Shields, A.R.; Carroll, R.L.; Washburn, S.; Falvo, M.R.; Superfine, R. Magnetically actuated nanorod arrays as biomimetic cilia. Nano Lett. 2007, 7, 1428–1434. [Google Scholar] [CrossRef]
  142. Belardi, J.; Schorr, N.; Prucker, O.; Rühe, J. Artificial cilia: Generation of magnetic actuators in microfluidic systems. Adv. Funct. Mater. 2011, 21, 3314–3320. [Google Scholar] [CrossRef]
  143. Wei, Y.; Wu, Z.; Dai, Z.; Zhou, B.; Xu, Q. Design of a Magnetic Soft Inchworm Millirobot Based on Pre-strained Elastomer with Micropillars. Biomimetics 2023, 8, 22. [Google Scholar] [CrossRef]
  144. Dong, X.; Lum, G.Z.; Hu, W.; Zhang, R.; Ren, Z.; Onck, P.R.; Sitti, M. Bioinspired cilia arrays with programmable nonreciprocal motion and metachronal coordination. Sci. Adv. 2020, 6, eabc9323. [Google Scholar] [CrossRef]
  145. Xu, Z.; Wu, Z.; Yuan, M.; Chen, H.; Ge, W.; Xu, Q. Multiple Cilia-like Swarms Enable Efficient Microrobot Deployment and Execution. Cell Rep. Phys. Sci. 2023, 4, 101329. [Google Scholar] [CrossRef]
  146. Feng, K.; Lei, M.; Wang, X.; Zhou, B.; Xu, Q. A Flexible Bidirectional Interface with Integrated Multimodal Sensing and Haptic Feedback for Closed-Loop Human–Machine Interaction. Adv. Intell. Syst. 2023, 5, 2300291. [Google Scholar] [CrossRef]
  147. Roseti, L.; Parisi, V.; Petretta, M.; Cavallo, C.; Desando, G.; Bartolotti, I.; Grigolo, B. Scaffolds for bone tissue engineering: State of the art and new perspectives. Mater. Sci. Eng. C 2017, 78, 1246–1262. [Google Scholar] [CrossRef] [PubMed]
  148. Kim, S.; Qiu, F.; Kim, S.; Ghanbari, A.; Moon, C.; Zhang, L.; Nelson, B.J.; Choi, H. Fabrication and characterization of magnetic microrobots for three-dimensional cell culture and targeted transportation. Adv. Mater. 2013, 25, 5863–5868. [Google Scholar] [CrossRef] [PubMed]
  149. Choi, J.; Hwang, J.; Kim, J.; Choi, H. Recent progress in magnetically actuated microrobots for targeted delivery of therapeutic agents. Adv. Healthc. Mater. 2021, 10, 2001596. [Google Scholar] [CrossRef]
  150. Go, G.; Yoo, A.; Song, H.; Min, H.; Zheng, S.; Nguyen, K.T.; Kim, S.; Kang, B.; Hong, A.; Kim, C.; et al. Multifunctional biodegradable microrobot with programmable morphology for biomedical applications. ACS Nano 2020, 15, 1059–1076. [Google Scholar] [CrossRef]
  151. Bernasconi, R.; Cuneo, F.; Carrara, E.; Chatzipirpiridis, G.; Hoop, M.; Chen, X.; Nelson, B.J.; Pané, S.; Credi, C.; Levi, M.; et al. Hard-magnetic cell microscaffolds from electroless coated 3D printed architectures. Mater. Horizons 2018, 5, 699–707. [Google Scholar] [CrossRef]
  152. Carlsen, R.W.; Sitti, M. Bio-hybrid cell-based actuators for microsystems. Small 2014, 10, 3831–3851. [Google Scholar]
  153. Patino, T.; Mestre, R.; Sanchez, S. Miniaturized soft bio-hybrid robotics: A step forward into healthcare applications. Lab Chip 2016, 16, 3626–3630. [Google Scholar] [CrossRef]
  154. Ávila, B.E.D.; Gao, W.; Karshalev, E.; Zhang, L.; Wang, J. Cell-like micromotors. Acc. Chem. Res 2018, 51, 1901–1910. [Google Scholar]
  155. Mundargi, R.C.; Potroz, M.G.; Park, S.; Shirahama, H.; Lee, J.H.; Seo, J.; Cho, N. Natural sunflower pollen as a drug delivery vehicle. Small 2016, 12, 1167–1173. [Google Scholar] [CrossRef] [PubMed]
  156. Prabhakar, A.K.; Potroz, M.G.; Tan, E.; Jung, H.; Park, J.H.; Cho, N. Macromolecular microencapsulation using pine pollen: Loading optimization and controlled release with natural materials. ACS Appl. Mater. Interfaces 2018, 10, 28428–28439. [Google Scholar] [CrossRef]
  157. Maric, T.; Nasir, M.Z.M.; Rosli, N.F.; Budanović, M.; Webster, R.D.; Cho, N.; Pumera, M. Microrobots derived from variety plant pollen grains for efficient environmental clean up and as an anti-cancer drug carrier. Adv. Funct. Mater. 2020, 30, 2000112. [Google Scholar] [CrossRef]
  158. Zhang, Y.; Yan, K.; Ji, F.; Zhang, L. Enhanced removal of toxic heavy metals using swarming biohybrid adsorbents. Adv. Funct. Mater. 2018, 28, 1806340. [Google Scholar] [CrossRef]
  159. Zhang, Y.; Zhang, L.; Yang, L.; Vong, C.I.; Chan, K.F.; Wu, W.K.K.; Kwong, T.N.Y.; Lo, N.W.S.; Ip, M.; Wong, S.H.; et al. Real-time tracking of fluorescent magnetic spore-based microrobots for remote detection of C. diff toxins. Sci. Adv. 2019, 5, eaau9650. [Google Scholar] [CrossRef] [PubMed]
  160. Ali, J.; Cheang, U.K.; Martindale, J.D.; Jabbarzadeh, M.; Fu, H.C.; Kim, M.J. Bacteria-inspired nanorobots with flagellar polymorphic transformations and bundling. Sci. Rep. 2017, 7, 14098. [Google Scholar] [CrossRef] [PubMed]
  161. Han, J.; Go, G.; Zhen, J.; Zheng, S.; Le, V.H.; Park, J.; Park, S. Feasibility Study of Dual-Targeting Paclitaxel-Loaded Magnetic Liposomes Using Electromagnetic Actuation and Macrophages. Sens. Actuators Chem. 2017, 240, 1226–1236. [Google Scholar]
  162. Magdanz, V.; Khalil, I.S.M.; Simmchen, J.; Furtado, G.P.; Mohanty, S.; Gebauer, J.; Xu, H.; Klingner, A.; Aziz, A.; Medina-Sánchez, M.; et al. IRONSperm: Sperm-templated soft magnetic microrobots. Sci. Adv. 2020, 6, eaba5855. [Google Scholar] [CrossRef]
  163. Wang, X.; Cai, J.; Sun, L.; Zhang, S.; Gong, D.; Li, X.; Yue, S.; Feng, L.; Zhang, D. Facile fabrication of magnetic microrobots based on spirulina templates for targeted delivery and synergistic chemo-photothermal therapy. ACS Appl. Mater. Interfaces 2019, 11, 4745–4756. [Google Scholar] [CrossRef]
  164. Guo, J.; Agola, J.O.; Serda, R.; Franco, S.; Lei, Q.; Wang, L.; Minster, J.; Croissant, J.G.; Butler, K.S.; Zhu, W.; et al. Biomimetic rebuilding of multifunctional red blood cells: Modular design using functional components. Acs Nano 2020, 14, 7847–7859. [Google Scholar] [CrossRef] [PubMed]
  165. Xu, H.; Medina-Sánchez, M.; Magdanz, V.; Schwarz, L.; Hebenstreit, F.; Schmidt, O.G. Sperm-hybrid micromotor for targeted drug delivery. ACS Nano 2018, 12, 327–337. [Google Scholar] [CrossRef] [PubMed]
  166. Park, B.; Zhuang, J.; Yasa, O.; Sitti, M. Multifunctional bacteria-driven microswimmers for targeted active drug delivery. ACS Nano 2017, 11, 8910–8923. [Google Scholar] [CrossRef] [PubMed]
  167. Striggow, F.; Medina-Sánchez, M.; Auernhammer, G.K.; Magdanz, V.; Friedrich, B.M.; Schmidt, O.G. Sperm-driven micromotors moving in oviduct fluid and viscoelastic media. Small 2020, 16, 2000213. [Google Scholar] [CrossRef] [PubMed]
  168. Li, J.; Jiang, X.; Li, H.; Gelinsky, M.; Gu, Z. Tailoring materials for modulation of macrophage fate. Adv. Mater. 2021, 33, 2004172. [Google Scholar] [CrossRef] [PubMed]
  169. Yasa, I.C.; Ceylan, H.; Bozuyuk, U.; Wild, A.; Sitti, M. Elucidating the interaction dynamics between microswimmer body and immune system for medical microrobots. Sci. Robot. 2020, 5, eaaz3867. [Google Scholar] [CrossRef] [PubMed]
  170. Yan, X.; Zhou, Q.; Vincent, M.; Deng, Y.; Yu, J.; Xu, J.; Xu, T.; Tang, T.; Bian, L.; Wang, Y.J.; et al. Multifunctional biohybrid magnetite microrobots for imaging-guided therapy. Sci. Robot. 2017, 2, eaaq1155. [Google Scholar] [CrossRef] [PubMed]
  171. Huang, Z.; Tsui, G.C.; Deng, Y.; Tang, C. Two-photon polymerization nanolithography technology for fabrication of stimulus-responsive micro/nano-structures for biomedical applications. Nanotechnol. Rev. 2020, 9, 1118–1136. [Google Scholar] [CrossRef]
  172. Zheng, C.; Jin, F.; Zhao, Y.; Zheng, M.; Liu, J.; Dong, X.; Xiong, Z.; Xia, Y.; Duan, X. Light-driven micron-scale 3D hydrogel actuator produced by two-photon polymerization microfabrication. Sens. Actuators Chem. 2020, 304, 127345. [Google Scholar] [CrossRef]
  173. Wang, X.; Hu, C.; Schurz, L.; Marco, C.D.; Chen, X.; Pané, S.; Nelson, B.J. Surface-chemistry-mediated control of individual magnetic helical microswimmers in a swarm. ACS Nano 2018, 12, 6210–6217. [Google Scholar] [CrossRef]
  174. Dong, M.; Wang, X.; Chen, X.; Mushtaq, F.; Deng, S.; Zhu, C.; Torlakcik, H.; Terzopoulou, A.; Qin, X.; Xiao, X.; et al. 3D-printed soft magnetoelectric microswimmers for delivery and differentiation of neuron-like cells. Adv. Funct. Mater. 2020, 30, 1910323. [Google Scholar] [CrossRef]
  175. Venkataramanababu, S.; Nair, G.; Deshpande, P.; Jithin, M.A.; Mohan, S.; Ghosh, A. Chiro-plasmonic refractory metamaterial with titanium nitride (TiN) core–shell nanohelices. Nanotechnology 2018, 29, 255203. [Google Scholar] [CrossRef] [PubMed]
  176. Alapan, Y.; Yasa, O.; Schauer, O.; Giltinan, J.; Tabak, A.F.; Sourjik, V.; Sitti, M. Soft erythrocyte-based bacterial microswimmers for cargo delivery. Sci. Robot. 2018, 3, eaar4423. [Google Scholar] [CrossRef] [PubMed]
  177. Liu, P.; Wang, Y.; Han, L.; Cai, Y.; Ren, H.; Ma, T.; Li, X.; Petrenko, V.A.; Liu, A. Colorimetric assay of bacterial pathogens based on Co3O4 magnetic nanozymes conjugated with specific fusion phage proteins and magnetophoretic chromatography. ACS Appl. Mater. Interfaces 2020, 12, 9090–9097. [Google Scholar] [CrossRef] [PubMed]
  178. Serrà, A.; Pip, P.; Gómez, E.; Philippe, L. Efficient magnetic hybrid ZnO-based photocatalysts for visible-light-driven removal of toxic cyanobacteria blooms and cyanotoxins. Appl. Catal. Environ. 2020, 268, 118745. [Google Scholar] [CrossRef]
  179. Serrà, A.; Artal, R.; García-Amorós, J.; Sepúlveda, B.; Gómez, E.; Nogués, J.; Philippe, L. Hybrid Ni@ ZnO@ ZnS-microalgae for circular economy: A smart route to the efficient integration of solar photocatalytic water decontamination and bioethanol production. Adv. Sci. 2020, 7, 1902447. [Google Scholar] [CrossRef] [PubMed]
  180. Maeda, Y.; Yasuda, T.; Matsuzaki, K.; Okazaki, Y.; Pouget, E.; Oda, R.; Kitada, A.; Murase, K.; Raffy, G.; Bassani, D.M.; et al. Common mechanism for helical nanotube formation by anodic polymerization and by cathodic deposition using helical pores on silicon electrodes. Electrochem. Commun. 2020, 114, 106714. [Google Scholar] [CrossRef]
  181. Wang, Q.; Wang, Y.; Guo, B.; Shao, S.; Yu, Y.; Zhu, X.; Wan, M.; Zhao, B.; Bo, C.; Mao, C. Novel heparin-loaded mesoporous tubular micromotors formed via template-assisted electrochemical deposition. J. Mater. Chem. B 2019, 7, 2688–2695. [Google Scholar] [CrossRef]
  182. Naderi, L.; Shahrokhian, S. Nickel vanadium sulfide grown on nickel copper phosphide Dendrites/Cu fibers for fabrication of all-solid-state wire-type micro-supercapacitors. Chem. Eng. J. 2020, 392, 124880. [Google Scholar] [CrossRef]
  183. Zhang, J.; Agramunt-Puig, S.; Del-Valle, N.; Navau, C.; Baro, M.D.; Estrade, S.; Peiro, F.; Pane, S.; Nelson, B.J.; Sanchez, A.; et al. Tailoring staircase-like hysteresis loops in electrodeposited trisegmented magnetic nanowires: A strategy toward minimization of interwire interactions. ACS Appl. Mater. Interfaces 2016, 8, 4109–4117. [Google Scholar] [CrossRef]
  184. Jang, B.; Hong, A.; Alcantara, C.; Chatzipirpiridis, G.; Marti, X.; Pellicer, E.; Sort, J.; Harduf, Y.; Or, Y.; Nelson, B.J.; et al. Programmable locomotion mechanisms of nanowires with semihard magnetic properties near a surface boundary. ACS Appl. Mater. Interfaces 2018, 11, 3214–3223. [Google Scholar] [CrossRef] [PubMed]
  185. Shields, C.W.; Velev, O.D. The evolution of active particles: Toward externally powered self-propelling and self-reconfiguring particle systems. Chem 2017, 3, 539–559. [Google Scholar] [CrossRef]
  186. Tasci, T.O.; Herson, P.S.; Neeves, K.B.; Marr, D.W.M. Surface-enabled propulsion and control of colloidal microwheels. Nat. Commun. 2016, 7, 10225. [Google Scholar] [CrossRef] [PubMed]
  187. Wang, Q.; Yang, L.; Wang, B.; Yu, E.; Yu, J.; Zhang, L. Collective behavior of reconfigurable magnetic droplets via dynamic self-assembly. ACS Appl. Mater. Interfaces 2018, 11, 1630–1637. [Google Scholar] [CrossRef] [PubMed]
  188. Toumey, C. Plenty of room, plenty of history. Nat. Nanotechnol. 2009, 4, 783–784. [Google Scholar] [CrossRef] [PubMed]
  189. Feng, J.; Cho, S.K. Mini and micro propulsion for medical swimmers. Micromachines 2014, 5, 97–113. [Google Scholar] [CrossRef]
  190. Hoang, M.C.; Le, V.H.; Nguyen, K.T.; Nguyen, V.D.; Kim, J.; Choi, E.; Bang, S.; Kang, B.; Park, J.; Kim, C. A robotic biopsy endoscope with magnetic 5-DOF locomotion and a retractable biopsy punch. Micromachines 2020, 11, 98. [Google Scholar] [CrossRef] [PubMed]
  191. Li, X.; Guo, S.; Shi, P.; Jin, X.; Kawanishi, M. An endovascular catheterization robotic system using collaborative operation with magnetically controlled haptic force feedback. Micromachines 2022, 13, 505. [Google Scholar] [CrossRef] [PubMed]
  192. Cai, Z.; Fu, Q.; Zhang, S.; Fan, C.; Zhang, X.; Guo, J.; Guo, S. Performance evaluation of a magnetically driven microrobot for targeted drug delivery. Micromachines 2021, 12, 1210. [Google Scholar] [CrossRef]
  193. Wu, Z.; Xu, Q.; Ai, N.; Ge, W. Design of a Novel Magnetically Actuated Biaxial Robot with Compact Structure and Easy Operation. IEEE Robot. Autom. Lett. 2023, 8, 3884–3891. [Google Scholar] [CrossRef]
  194. Gao, W.; Kagan, D.; Pak, O.S.; Clawson, C.; Campuzano, S.; Chuluun-Erdene, E.; Shipton, E.; Fullerton, E.E.; Zhang, L.; Lauga, E.; et al. Cargo-towing fuel-free magnetic nanoswimmers for targeted drug delivery. Small 2012, 8, 460–467. [Google Scholar] [CrossRef] [PubMed]
  195. Iacovacci, V.; Ricotti, L.; Sinibaldi, E.; Signore, G.; Vistoli, F.; Menciassi, A. An intravascular magnetic catheter enables the retrieval of nanoagents from the bloodstream. Adv. Sci. 2018, 5, 1800807. [Google Scholar] [CrossRef] [PubMed]
  196. Luo, M.; Feng, Y.; Wang, T.; Guan, J. Micro-/nanorobots at work in active drug delivery. Adv. Funct. Mater. 2018, 28, 1706100. [Google Scholar] [CrossRef]
  197. Mhanna, R.; Qiu, F.; Zhang, L.; Ding, Y.; Sugihara, K.; Zenobi-Wong, M.; Nelson, B.J. Artificial bacterial flagella for remote-controlled targeted single-cell drug delivery. Small 2014, 10, 1953–1957. [Google Scholar] [CrossRef] [PubMed]
  198. Mostaghaci, B.; Yasa, O.; Zhuang, J.; Sitti, M. Bioadhesive bacterial microswimmers for targeted drug delivery in the urinary and gastrointestinal tracts. Adv. Sci. 2017, 4, 1700058. [Google Scholar] [CrossRef] [PubMed]
  199. Xu, H.; Medina-Sánchez, M.; Maitz, M.F.; Werner, C.; Schmidt, O.G. Sperm micromotors for cargo delivery through flowing blood. ACS Nano 2020, 14, 2982–2993. [Google Scholar] [CrossRef]
  200. Yan, X.; Xu, J.; Zhou, Q.; Jin, D.; Vong, C.I.; Feng, Q.; Ng, D.H.L.; Bian, L.; Zhang, L. Molecular cargo delivery using multicellular magnetic microswimmers. Appl. Mater. Today 2019, 15, 242–251. [Google Scholar] [CrossRef]
  201. Yasa, O.; Erkoc, P.; Alapan, Y.; Sitti, M. Microalga-powered microswimmers toward active cargo delivery. Adv. Mater. 2018, 30, 1804130. [Google Scholar] [CrossRef]
  202. Burdick, J.; Laocharoensuk, R.; Wheat, P.M.; Posner, J.D.; Wang, J. Synthetic nanomotors in microchannel networks: Directional microchip motion and controlled manipulation of cargo. J. Am. Chem. Soc. 2008, 130, 8164–8165. [Google Scholar] [CrossRef]
  203. Denzer, U.W.; Rösch, T.; Hoytat, B.; Abdel-Hamid, M.; Hebuterne, X.; Vanbiervielt, G.; Filippi, J.; Ogata, H.; Hosoe, N.; Ohtsuka, K.; et al. Magnetically guided capsule versus conventional gastroscopy for upper abdominal complaints: A prospective blinded study. J. Clin. Gastroenterol. 2015, 49, 101–107. [Google Scholar] [CrossRef]
  204. Akolpoglu, M.B.; Alapan, Y.; Dogan, N.O.; Baltaci, S.F.; Yasa, O.; Tural, G.A.; Sitti, M. Magnetically steerable bacterial microrobots moving in 3D biological matrices for stimuli-responsive cargo delivery. Sci. Adv. 2022, 8, eabo6163. [Google Scholar] [CrossRef] [PubMed]
  205. Yin, X.; Guo, S.; Song, Y. Magnetorheological fluids actuated haptic-based teleoperated catheter operating system. Micromachines 2018, 9, 465. [Google Scholar] [CrossRef] [PubMed]
  206. Nelson, B.J.; Kaliakatsos, I.K.; Abbott, J.J. Microrobots for minimally invasive medicine. Annu. Rev. Biomed. Eng. 2010, 12, 55–85. [Google Scholar] [CrossRef] [PubMed]
  207. Wang, W.; Wu, Z.; He, Q. Swimming nanorobots for opening a cell membrane mechanically. View 2020, 1, 20200005. [Google Scholar] [CrossRef]
  208. Venugopalan, P.L.; de Avila, B.E.; Pal, M.; Ghosh, A.; Wang, J. Fantastic voyage of nanomotors into the cell. ACS Nano 2020, 14, 9423–9439. [Google Scholar] [CrossRef] [PubMed]
  209. Parmar, J.; Vilela, D.; Pellicer, E.; los Ojos, D.E.; Sort, J.; Sánchez, S. Reusable and long-lasting active microcleaners for heterogeneous water remediation. Adv. Funct. Mater. 2016, 26, 4152–4161. [Google Scholar] [CrossRef]
  210. Jang, B.; Gutman, E.; Stucki, N.; Seitz, B.F.; Wendel-García, P.D.; Newton, T.; Pokki, J.; Ergeneman, O.; Pané, S.; Or, Y.; et al. Undulatory locomotion of magnetic multilink nanoswimmers. Nano Lett. 2015, 15, 4829–4833. [Google Scholar] [CrossRef]
  211. Stanton, M.M.; Park, B.; Miguel-López, A.; Ma, X.; Sitti, M.; Sánchez, S. Biohybrid microtube swimmers driven by single captured bacteria. Small 2017, 13, 1603679. [Google Scholar] [CrossRef]
  212. Feng, L.; Zhou, Q.; Song, B.; Feng, Y.; Cai, J.; Jiang, Y.; Zhang, D. Cell injection millirobot development and evaluation in microfluidic chip. Micromachines 2018, 9, 590. [Google Scholar] [CrossRef]
  213. Zong, Z.; Zhou, X.; Zhang, L.; Tan, Q.; Xiong, J.; Zhang, W. Magnetically propelled soft micromachines with multipatterned fabrications. J. Micromech. Microeng. 2020, 30, 085001. [Google Scholar] [CrossRef]
  214. Feng, L.; Hagiwara, M.; Ichikawa, A.; Arai, F. On-chip enucleation of bovine oocytes using microrobot-assisted flow-speed control. Micromachines 2013, 4, 272–285. [Google Scholar] [CrossRef]
  215. Zhang, L.; Petit, T.; Peyer, K.E.; Nelson, B.J. Targeted cargo delivery using a rotating nickel nanowire. Nanomed. Nanotechnol. Biol. Med. 2012, 8, 1074–1080. [Google Scholar] [CrossRef]
  216. Lin, Z.; Fan, X.; Sun, M.; Gao, C.; He, Q.; Xie, H. Magnetically actuated peanut colloid motors for cell manipulation and patterning. ACS Nano 2018, 12, 2539–2545. [Google Scholar] [CrossRef] [PubMed]
  217. Dekanovsky, L.; Khezri, B.; Rottnerova, Z.; Novotny, F.; Plutnar, J.; Pumera, M. Chemically programmable microrobots weaving a web from hormones. Nat. Mach. Intell. 2020, 2, 711–718. [Google Scholar] [CrossRef]
  218. Jin, D.; Zhang, L. Embodied intelligence weaves a better future. Nat. Mach. Intell. 2020, 2, 663–664. [Google Scholar] [CrossRef]
  219. Sun, M.; Chen, W.; Fan, X.; Tian, C.; Sun, L.; Xie, H. Cooperative recyclable magnetic microsubmarines for oil and microplastics removal from water. Appl. Mater. Today 2020, 20, 100682. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of magnetic microrobots covering five aspects: materials, propulsion, design, fabrication, and applications.
Figure 1. Schematic illustration of magnetic microrobots covering five aspects: materials, propulsion, design, fabrication, and applications.
Micromachines 15 00468 g001
Figure 2. Helical microrobots: (a) The degradation process of superparamagnetic hydrogel swimming microrobots. © 2016 Wiley Online Library. Reprinted with permission from [83]. (b) A double-helical microrobot. © 2019 ACS. Reprinted with permission from [41]. (c) A microrobot through the depicted microfluidic tapered channel. © 2023 Wiley Online Library. Reprinted with permission from [136]. (d) Fabrication process of helical hydrogel micromotors. © 2024 Elsevier. Reprinted with permission from [137].
Figure 2. Helical microrobots: (a) The degradation process of superparamagnetic hydrogel swimming microrobots. © 2016 Wiley Online Library. Reprinted with permission from [83]. (b) A double-helical microrobot. © 2019 ACS. Reprinted with permission from [41]. (c) A microrobot through the depicted microfluidic tapered channel. © 2023 Wiley Online Library. Reprinted with permission from [136]. (d) Fabrication process of helical hydrogel micromotors. © 2024 Elsevier. Reprinted with permission from [137].
Micromachines 15 00468 g002
Figure 3. Ciliary microrobots: (a) Magnetic rod arrays. © 2007 ACS. Reprinted with permission from [141]. (b) The locomotion mode of the microrobot. © 2018 Nature. Reprinted with permission from [65]. (c) A bionic magnetic inchworm robot. © 2023 MDPI. Reprinted with permission from [143]. (d) A cilia array pumping in enclosed channels. © 2020 Science. Reprinted with permission from [144]. (e) Cilia’s generation and controlled movement. © 2023 Elsevier. Reprinted with permission from [145]. (f) Working principle of the sensor with magnetic cilia structure. © 2023 Wiley Online Library. Reprinted with permission from [146].
Figure 3. Ciliary microrobots: (a) Magnetic rod arrays. © 2007 ACS. Reprinted with permission from [141]. (b) The locomotion mode of the microrobot. © 2018 Nature. Reprinted with permission from [65]. (c) A bionic magnetic inchworm robot. © 2023 MDPI. Reprinted with permission from [143]. (d) A cilia array pumping in enclosed channels. © 2020 Science. Reprinted with permission from [144]. (e) Cilia’s generation and controlled movement. © 2023 Elsevier. Reprinted with permission from [145]. (f) Working principle of the sensor with magnetic cilia structure. © 2023 Wiley Online Library. Reprinted with permission from [146].
Micromachines 15 00468 g003
Figure 4. Biohybrid microrobots: (a) Spore-based microrobots. © 2019 Science. Reprinted with permission, from [159]. (b) Bacteria-based microrobots. © 2017 Nature. Reprinted with permission from [160]. (c) Macrophage-based microrobots. © 2017 Elsevier. Reprinted with permission from [161]. (d) Sperm-based microrobots. © 2020 Science. Reprinted with permission from [162]. (e) Spirulina-based microrobots. © 2019 ACS. Reprinted with permission from [163]. (f) RBC (red blood cell)-based microrobots. © 2020 ACS. Reprinted with permission from [164].
Figure 4. Biohybrid microrobots: (a) Spore-based microrobots. © 2019 Science. Reprinted with permission, from [159]. (b) Bacteria-based microrobots. © 2017 Nature. Reprinted with permission from [160]. (c) Macrophage-based microrobots. © 2017 Elsevier. Reprinted with permission from [161]. (d) Sperm-based microrobots. © 2020 Science. Reprinted with permission from [162]. (e) Spirulina-based microrobots. © 2019 ACS. Reprinted with permission from [163]. (f) RBC (red blood cell)-based microrobots. © 2020 ACS. Reprinted with permission from [164].
Micromachines 15 00468 g004
Figure 5. Biomedical applications of magnetic microrobots: (a) Targeted drug delivery. © 2021 MDPI. Reprinted with permission from [192]. (b) Manipulation of the zebrafish embryo. © 2023 IEEE. Reprinted with permission from [193]. (c) Extraction of thorn. © 2023 AAAS. Reprinted with permission from [27].
Figure 5. Biomedical applications of magnetic microrobots: (a) Targeted drug delivery. © 2021 MDPI. Reprinted with permission from [192]. (b) Manipulation of the zebrafish embryo. © 2023 IEEE. Reprinted with permission from [193]. (c) Extraction of thorn. © 2023 AAAS. Reprinted with permission from [27].
Micromachines 15 00468 g005
Figure 6. Challenges and prospects for magnetic microrobots.
Figure 6. Challenges and prospects for magnetic microrobots.
Micromachines 15 00468 g006
Table 1. Comparison of recent magnetic microrobots.
Table 1. Comparison of recent magnetic microrobots.
DesignActuation MethodsMaterialsFabrication MethodsApplicationsSpeed ( μ s 1 )Size ( μ m)Ref.
HelicalRotatingFe3O4, PLA, ATBC, DCM 1TAEDDrug delivery2950 985 × 250 [52]
RotatingPhotoresistDLWDrug delivery128 300 × 10 [53]
RotatingNdFeB, PLADLWDrug delivery≈200–2300 3000 × 109 [54]
RotatingMNPs 2, PEGDA700, ethylenediamineDLWDrug delivery160 100 × 35 [55]
RotatingFe, Pt, Si O 2 GLADCell manipulation241.5[56]
RotatingFe, Ti, Ormocomp, PMMA, PDMS 3DLWCell manipulation≈800130–170[57]
RotatingNi, Zinc-based MOF 4, framework-8 (ZIF-8), zeolitic imidazoleGLADDrug delivery5010[58]
RotatingFe3O4, PEGDA, PETA 5DLWMinimally invasive surgery82120[59]
SurfaceRotatingFe Cl 2 ·4 H 2 O, Fe Cl 3 ·6 H 2 O, PDMSDLWDrug delivery1000–5020[60]
RotatingAu, Ag, NiTAEDMinimally invasive surgery13.19 6 × 0.6 [61]
RotatingAu, Ni, Si O 2 TAEDDrug delivery6003.0–7.8[62]
Oscillating γ -Fe2O3, Pt, sulphonyl esters, PM 6MSACell manipulation 2.0 ± 0.05 ≈4.5[63]
CiliaryRotatingNdFeB, Ecoflex 00-30TAEDDrug delivery834000[64]
GradientFe, PDMSTAEDDrug delivery640 17 × 7 [65]
ScaffoldRotationNi, Ti O 2 , IP-S photoresistDLWCell manipulation 95 × 300 [66]
RotatingMNP, SiCNDLWCell manipulation≈85.56 42 × 15 [67]
GradientNi, Ti, SU-8DLWCell manipulation≈150070–90[68]
BiohybridGradientFe3O4, bacteria, Spirulina platensisBTSMinimally invasive surgery21.7–78.3≈50[69]
RotationFe3O4, BaTi O 3 , S. platensisBTSCell manipulation333.3 22 × 0.6 [70]
OscillatingFe3O4, RBCs, IGG 7BTSDrug delivery56.5 2 [71]
RotationFe3O4, Pine pollenBTSDrug delivery175.1925[72]
1 PLA, polylactic acid; ATBC, Acetyl-tributyl citrate; DCM, dichloromethane; 2 MNPs, magnetic nanoparticles; 3 PMMA, poly methyl methacrylate; PDMS, poly(dimethylsiloxane); 4 MOF, metal–organic framework; 5 PETA, pentaerythritol triacrylate; 6 PM, commercially available superparamagnetic microspheres; 7 RBCs, red blood cells; IGG, indocyanine green.
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

Xu, R.; Xu, Q. A Survey of Recent Developments in Magnetic Microrobots for Micro-/Nano-Manipulation. Micromachines 2024, 15, 468. https://doi.org/10.3390/mi15040468

AMA Style

Xu R, Xu Q. A Survey of Recent Developments in Magnetic Microrobots for Micro-/Nano-Manipulation. Micromachines. 2024; 15(4):468. https://doi.org/10.3390/mi15040468

Chicago/Turabian Style

Xu, Ruomeng, and Qingsong Xu. 2024. "A Survey of Recent Developments in Magnetic Microrobots for Micro-/Nano-Manipulation" Micromachines 15, no. 4: 468. https://doi.org/10.3390/mi15040468

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