Jump to content

Microbotics

From Wikipedia, the free encyclopedia
Jasmine minirobots each smaller than 3 cm (1 in) in width

Microbotics (or microrobotics) is the field of miniature robotics, in particular mobile robots with characteristic dimensions less than 1 mm. The term can also be used for robots capable of handling micrometer size components.

History[edit]

Microbots were born thanks to the appearance of the microcontroller in the last decade of the 20th century, and the appearance of microelectromechanical systems (MEMS) on silicon, although many microbots do not use silicon for mechanical components other than sensors. The earliest research and conceptual design of such small robots was conducted in the early 1970s in (then) classified research for U.S. intelligence agencies. Applications envisioned at that time included prisoner of war rescue assistance and electronic intercept missions. The underlying miniaturization support technologies were not fully developed at that time, so that progress in prototype development was not immediately forthcoming from this early set of calculations and concept design.[1] As of 2008, the smallest microrobots use a scratch drive actuator.[2]

The development of wireless connections, especially Wi-Fi (i.e. in household networks) has greatly increased the communication capacity of microbots, and consequently their ability to coordinate with other microbots to carry out more complex tasks. Indeed, much recent research has focused on microbot communication, including a 1,024 robot swarm at Harvard University that assembles itself into various shapes;[3] and manufacturing microbots at SRI International for DARPA's "MicroFactory for Macro Products" program that can build lightweight, high-strength structures.[4][5]

Microbots called xenobots have also been built using biological tissues instead of metal and electronics.[6] Xenobots avoid some of the technological and environmental complications of traditional microbots as they are self-powered, biodegradable, and biocompatible.

Definitions[edit]

While the "micro" prefix has been used subjectively to mean "small", standardizing on length scales avoids confusion. Thus a nanorobot would have characteristic dimensions at or below 1 micrometer, or manipulate components on the 1 to 1000 nm size range. [citation needed] A microrobot would have characteristic dimensions less than 1 millimeter, a millirobot would have dimensions less than a cm, a mini-robot would have dimensions less than 10 cm (4 in), and a small robot would have dimensions less than 100 cm (39 in). [7]

Many sources also describe robots larger than 1 millimeter as microbots or robots larger than 1 micrometer as nanobots. See also: Category:Micro robots

Design considerations[edit]

The way microrobots move around is a function of their purpose and necessary size. At submicron sizes, the physical world demands rather bizarre ways of getting around. The Reynolds number for airborne robots is less than unity; the viscous forces dominate the inertial forces, so “flying” could use the viscosity of air, rather than Bernoulli's principle of lift. Robots moving through fluids may require rotating flagella like the motile form of E. coli. Hopping is stealthy and energy-efficient; it allows the robot to negotiate the surfaces of a variety of terrains.[8] Pioneering calculations (Solem 1994) examined possible behaviors based on physical realities.[9]

One of the major challenges in developing a microrobot is to achieve motion using a very limited power supply. The microrobots can use a small lightweight battery source like a coin cell or can scavenge power from the surrounding environment in the form of vibration or light energy.[10] Microrobots are also now using biological motors as power sources, such as flagellated Serratia marcescens, to draw chemical power from the surrounding fluid to actuate the robotic device. These biorobots can be directly controlled by stimuli such as chemotaxis or galvanotaxis with several control schemes available. A popular alternative to an onboard battery is to power the robots using externally induced power. Examples include the use of electromagnetic fields,[11] ultrasound and light to activate and control micro robots.[12]

The 2022 study focused on a photo-biocatalytic approach for the "design of light-driven microrobots with applications in microbiology and biomedicine".[13][14][15]

Locomotion of microrobots[edit]

Microrobots employ various locomotion methods to navigate through different environments, from solid surfaces to fluids. These methods are often inspired by biological systems and are designed to be effective at the micro-scale [16]. Several factors need to be maximized (precision, speed, stability), and others have to be minimized (energy consumption, energy loss) in the design and operation of microrobot locomotion in order to guarantee accurate, effective, and efficient movement [17].

When describing the locomotion of microrobots, several key parameters are used to characterize and evaluate their movement, including stride length and transportation costs. A stride refers to a complete cycle of movement that includes all the steps or phases necessary for an organism or robot to move forward by repeating a specific sequence of actions. Stride length (𝞴s) is the distance covered by a microrobot in one complete cycle of its locomotion mechanism. Cost of transport (CoT) defines the work required to move a unit of mass of a microrobot a unit of distance [17]

Surface locomotion[edit]

Microrobots that use surface locomotion can move in a variety of ways, including walking, crawling, rolling, or jumping. These microrobots meet different challenges, such as gravity and friction. One of the parameters describing surface locomotion is the Frounde number, defined as:

Where v is motion speed, g is the gravitational field, and 𝞴s is a stride length. A microrobot demonstrating a low Froude number moves slower and more stable as gravitational forces dominate, while a high Froude number indicates that inertial forces are more significant, allowing faster and potentially less stable movement [17].

Crawling is one of the most typical surface locomotion types. The mechanisms employed by microrobots for crawling can differ but usually include the synchronized movement of multiple legs or appendages. The mechanism of the microrobots' movements is often inspired by animals such as insects, reptiles, and small mammals. An example of a crawling microrobot is RoBeetle. The autonomous microrobot weighs 88 milligrams (approximately the weight of three rice grains). The robot is powered by the catalytic combustion of methanol. The design relies on controllable NiTi-Pt–based catalytic artificial micromuscles with a mechanical control mechanism [18].

Other options for actuating microrobots' surface locomotion include magnetic, electromagnetic, piezoelectric, electrostatic, and optical actuation.

Swimming locomotion[edit]

Swimming microrobots are designed to operate in 3D through fluid environments, like biological fluids or water. To achieve effective movements, locomotion strategies are adopted from small aquatic animals or microorganisms, such as flagellar propulsion, pulling, chemical propulsion, jet propulsion, and tail undulation. Swimming microrobots, in order to move forward, must drive water backward [17].

Microrobots move in the low Reynolds number regime due to their small sizes and low operating speeds, as well as high viscosity of the fluids they navigate. At this level, viscous forces dominate over inertial forces. This requires a different approach in the design compared to swimming at the macroscale in order to achieve effective movements. The low Reynolds number also allows for accurate movements, which makes it good application in medicine, micro-manipulation tasks, and environmental monitoring [16][17].

Dominating viscous (Stokes) drag forces Tdrag on the robot balances the propulsive force Fp generated by a swimming mechanism.

 

Where b is the viscous drag coefficient, v is motion speed, and m is the body mass [17].

One of the examples of a swimming microrobot is a helical magnetic microrobot consisting of a spiral tail and a magnetic head body. This design is inspired by the flagellar motion of bacteria. By applying a magnetic torque to a helical microrobot within a low-intensity rotating magnetic field, the rotation can be transformed into linear motion. This conversion is highly effective in low Reynolds number environments due to the unique helical structure of the microrobot. By altering the external magnetic field, the direction of the spiral microrobot's motion can be easily reversed [19].

At Air-Fluid Interface locomotion[edit]

In the specific instance when microrobots are at the air-fluid interface, they can take advantage of surface tension and forces provided by capillary motion. At the point where air and a liquid, most often water, come together, it is possible to establish an interface capable of supporting the weight of the microrobots through the work of surface tension. Cohesion between molecules of a liquid creates surface tension, which otherwise creates ‘skin’ over the water’s surface, letting the microrobots float instead of sinking. Through such concepts, microrobots could perform specific locomotion functions, including climbing, walking, levitating, floating, and or even jumping, by exploring the characteristics of the air-fluid interface [17][20].

Due to the surface tension ,σ, the buoyancy force, Fb, and the curvature force, Fc, play the most important roles, particularly in deciding whether the microrobot will float or sink on the surface of the liquid. This can be expressed as

Fb is obtained by integrating the hydrostatic pressure over the area of the body in contact with the water. In contrast, Fc is obtained by integrating the curvature pressure over this area or, alternatively, the vertical component of the surface tension, , along the contact perimeter[21] .

One example of a climbing, walking microrobot that utilizes air-fluid locomotion is the Harvard Ambulatory MicroRobot with Electroadhesion (HAMR-E)[22]. The control system of HAMR-E is developed to allow the robot to function in a flexible and maneuverable manner in a challenging environment. Its features include its ability to move on horizontal, vertical, and inverted planes, which is facilitated by the electro-adhesion system. This uses electric fields to create electrostatic attraction, causing the robot to stick and move on different surfaces [23]. With four compliant and electro-adhesion footpads, HAMR-E can safely grasp and slide over various substrate types, including glass, wood, and metal [22]. The robot has a slim body and is fully posable, making it easy to perform complex movements and balance on any surface.

Flying locomotion[edit]

Flying microrobots are miniature robotic systems meticulously engineered to operate in the air by emulating the flight mechanisms of insects and birds. These microrobots have to overcome the issues related to lift, thrust, and movement that are challenging to accomplish at such a small scale where most aerodynamic theories must be modified. Active flight is the most energy-intensive mode of locomotion, as the microrobot must lift its body weight while propelling itself forward [17].To achieve this function, these microrobots mimic the movement of insect wings and generate the necessary airflow for producing lift and thrust. Miniaturized wings of the robots are actuated with Piezoelectric materials, which offer better control of wing kinematics and flight dynamics [24].

To calculate the necessary aerodynamic power for maintaining a hover with flapping wings, the primary physical equation is expressed as

where m is the body mass, L is the wing length, Φ represents the wing flapping amplitude in radians, ρ indicates the air density, and Vi corresponds to the induced air speed surrounding the body, a consequence of the wings' flapping and rotation movements. This equation illustrates that a small insect or robotic device must impart sufficient momentum to the surrounding air to counterbalance its own weight [25].

One example of a flying microrobot that utilizes flying locomotion is the RoboBee and DelFly Nimble [26][27], which, regarding flight dynamics, emulate bees and fruit flies, respectively. Harvard University invented the RoboBee, a miniature robot that mimics a bee fly, takes off and lands like one, and moves around confined spaces. It can be used in self-driving pollination and search operations for missing people and things. The DelFly Nimble, developed by the Delft University of Technology, is one of the most agile micro aerial vehicles that can mimic the maneuverability of a fruit fly by doing different tricks due to its minimal weight and advanced control mechanisms [26][27].

Types and applications[edit]

Due to their small size, microbots are potentially very cheap, and could be used in large numbers (swarm robotics) to explore environments which are too small or too dangerous for people or larger robots. It is expected that microbots will be useful in applications such as looking for survivors in collapsed buildings after an earthquake or crawling through the digestive tract. What microbots lack in brawn or computational power, they can make up for by using large numbers, as in swarms of microbots.

Potential applications with demonstrated prototypes include:

Medical microbots[edit]

Biohybrid bacterial microswimmers [28]
Biohybrid diatomite microswimmer drug delivery system
Diatom frustule surface functionalised with photoactivable molecules (orange spheres) linked to vitamin B-12 (red sphere) acting as a tumor-targeting tag. The system can be loaded with chemotherapeutic drugs (light blue spheres), which can be selectively delivered to colorectal cancer cells. In addition, diatomite microparticles can be photoactivated to generate carbon monoxide or free radicals inducing tumor cell apoptosis.[29][30]

Biohybrid microswimmers, mainly composed of integrated biological actuators and synthetic cargo carriers, have recently shown promise toward minimally invasive theranostic applications.[31][32][33][34] Various microorganisms, including bacteria,[35][36] microalgae,[37][38] and spermatozoids,[39][40] have been utilised to fabricate different biohybrid microswimmers with advanced medical functionalities, such as autonomous control with environmental stimuli for targeting, navigation through narrow gaps, and accumulation to necrotic regions of tumor environments.[41] Steerability of the synthetic cargo carriers with long-range applied external fields, such as acoustic or magnetic fields,[42][43] and intrinsic taxis behaviours of the biological actuators toward various environmental stimuli, such as chemoattractants,[44] pH, and oxygen,[45][46] make biohybrid microswimmers a promising candidate for a broad range of medical active cargo delivery applications.[41][28]

For example, there are biocompatible microalgae-based microrobots for active drug-delivery in the lungs and the gastrointestinal tract,[47][48][49] and magnetically guided engineered bacterial microbots for 'precision targeting'[50] for fighting cancer[51][52] that all have been tested with mice.

See also[edit]

References[edit]

  1. ^ Solem, J. C. (1996). "The application of microrobotics in warfare". Los Alamos National Laboratory Technical Report LAUR-96-3067. doi:10.2172/369704.
  2. ^ "Microrobotic Ballet". Duke University. June 2, 2008. Archived from the original on 2011-04-03. Retrieved 2014-08-24.
  3. ^ Hauert, Sabine (2014-08-14). "Thousand-robot swarm assembles itself into shapes". Ars Technica. Retrieved 2014-08-24.
  4. ^ Misra, Ria (2014-04-22). "This Swarm Of Insect-Inspired Microbots Is Unsettlingly Clever". io9. Retrieved 2014-08-24.
  5. ^ Temple, James (2014-04-16). "SRI Unveils Tiny Robots Ready to Build Big Things". re/code. Archived from the original on 2014-08-25. Retrieved 2014-08-24.
  6. ^ Kriegman, Sam; Blackiston, Douglas; Levin, Michael; Bongard, Josh (2020). "A scalable pipeline for designing reconfigurable organisms". Proceedings of the National Academy of Sciences. 117 (4): 1853–1859. Bibcode:2020PNAS..117.1853K. doi:10.1073/pnas.1910837117. PMC 6994979. PMID 31932426.
  7. ^ "Microrobotics: Tiny Robots and Their Many Uses | Built In". builtin.com. Retrieved 2024-01-26.
  8. ^ Solem, J. C. (1994). "The motility of microrobots". In Langton, C. (ed.). Artificial Life III: Proceedings of the Workshop on Artificial Life, June 1992, Santa Fe, NM. Proceedings, Santa Fe Institute studies in the sciences of complexity. Vol. 17. Santa Fe Institute Studies in the Sciences of Complexity (Addison-Wesley, Reading, MA). pp. 359–380.
  9. ^ Kristensen, Lars Kroll (2000). "Aintz: A study of emergent properties in a model of ant foraging". In Bedau, M. A.; et al. (eds.). Artificial Life VII: Proceedings of the Seventh International Conference on Artificial Life. MIT Press. p. 359. ISBN 9780262522908.
  10. ^ Meinhold, Bridgette (31 August 2009). "Swarms of Solar Microbots May Revolutionize Data Gathering". Inhabitat.
  11. ^ Ecole Polytechnique Federale de Lausanne (January 18, 2019). "Researchers develop smart micro-robots that can adapt to their surroundings". Phys.org.
  12. ^ Chang, Suk Tai; Paunov, Vesselin N.; Petsev, Dimiter N.; Velev, Orlin D. (March 2007). "Remotely powered self-propelling particles and micropumps based on miniature diodes". Nature Materials. 6 (3): 235–240. Bibcode:2007NatMa...6..235C. doi:10.1038/nmat1843. ISSN 1476-1122. PMID 17293850. S2CID 20558069.
  13. ^ Villa, Katherine; Sopha, Hanna; Zelenka, Jaroslav; Motola, Martin; Dekanovsky, Lukas; Beketova, Darya Chylii; Macak, Jan M.; Ruml, Tomáš; Pumera, Martin (2022-02-05). "Enzyme-Photocatalyst Tandem Microrobot Powered by Urea for Escherichia coli Biofilm Eradication". Small. 18 (36): 2106612. doi:10.1002/smll.202106612. ISSN 1613-6810. PMID 35122470.
  14. ^ Jones, Nicholas. "Revolutionizing Robotics and AGVs with Advanced Drive Control". ds200sdccg4a.com. Retrieved 2024-01-26.
  15. ^ Chemistry, University of; Prague, Technology. "New research into a microrobot powered by urea for E. coli biofilm eradication". phys.org. Retrieved 2022-07-22.
  16. ^ a b Abbott, Jake J.; Peyer, Kathrin E.; Lagomarsino, Marco Cosentino; Zhang, Li; Dong, Lixin; Kaliakatsos, Ioannis K.; Nelson, Bradley J. (November 2009). "How Should Microrobots Swim?". The International Journal of Robotics Research. 28 (11–12) (published July 21, 2009): 1434–1447. doi:10.1177/0278364909341658. ISSN 0278-3649.
  17. ^ a b c d e f g h Sitti, Metin (2017). Mobile microrobotics. Intelligent robotics and autonomous agents. Cambridge, MA: MIT Press. ISBN 978-0-262-03643-6.
  18. ^ Yang, Xiufeng; Chang, Longlong; Pérez-Arancibia, Néstor O. (2020-08-26). "An 88-milligram insect-scale autonomous crawling robot driven by a catalytic artificial muscle". Science Robotics. 5 (45). doi:10.1126/scirobotics.aba0015. ISSN 2470-9476. PMID 33022629.
  19. ^ Liu, Huibin; Guo, Qinghao; Wang, Wenhao; Yu, Tao; Yuan, Zheng; Ge, Zhixing; Yang, Wenguang (2023-01-01). "A review of magnetically driven swimming microrobots: Material selection, structure design, control method, and applications". Reviews on Advanced Materials Science. 62 (1): 119. Bibcode:2023RvAMS..62..119L. doi:10.1515/rams-2023-0119. ISSN 1605-8127.
  20. ^ Koh, Je-Sung; Yang, Eunjin; Jung, Gwang-Pil; Jung, Sun-Pill; Son, Jae Hak; Lee, Sang-Im; Jablonski, Piotr G.; Wood, Robert J.; Kim, Ho-Young; Cho, Kyu-Jin (2015-07-31). "Jumping on water: Surface tension–dominated jumping of water striders and robotic insects". Science. 349 (6247): 517–521. Bibcode:2015Sci...349..517K. doi:10.1126/science.aab1637. ISSN 0036-8075.
  21. ^ Hu, David L.; Chan, Brian; Bush, John W. M. (August 2003). "The hydrodynamics of water strider locomotion". Nature. 424 (6949): 663–666. Bibcode:2003Natur.424..663H. doi:10.1038/nature01793. ISSN 0028-0836. PMID 12904790.
  22. ^ a b de Rivaz, Sébastien D.; Goldberg, Benjamin; Doshi, Neel; Jayaram, Kaushik; Zhou, Jack; Wood, Robert J. (2018-12-19). "Inverted and vertical climbing of a quadrupedal microrobot using electroadhesion". Science Robotics. 3 (25). doi:10.1126/scirobotics.aau3038. ISSN 2470-9476. PMID 33141691.
  23. ^ Rajagopalan, Pandey; Muthu, Manikandan; Liu, Yulu; Luo, Jikui; Wang, Xiaozhi; Wan, Chaoying (July 2022). "Advancement of Electroadhesion Technology for Intelligent and Self-Reliant Robotic Applications". Advanced Intelligent Systems. 4 (7). doi:10.1002/aisy.202200064. ISSN 2640-4567.
  24. ^ Jafferis, Noah T.; Helbling, E. Farrell; Karpelson, Michael; Wood, Robert J. (June 2019). "Untethered flight of an insect-sized flapping-wing microscale aerial vehicle". Nature. 570 (7762): 491–495. Bibcode:2019Natur.570..491J. doi:10.1038/s41586-019-1322-0. ISSN 1476-4687. PMID 31243384.
  25. ^ Shyy, Wei; Lian, Yongsheng; Tang, Jian; Viieru, Dragos; Liu, Hao (2007). Aerodynamics of Low Reynolds Number Flyers. Cambridge Aerospace Series. Cambridge: Cambridge University Press. doi:10.1017/cbo9780511551154. ISBN 978-0-521-88278-1.
  26. ^ a b Wang, S.; den Hoed, M.; Hamaza, S. (2024). "A Low-cost Fabrication Approach to Embody Flexible and Lightweight Strain Sensing on Flapping Wings: 2024 IEEE International Conference onRobotics and Automation". IEEE ICRA 2024 - Workshop on Bioinspired, Soft, and Other Novel Design Paradigms for Aerial Robotics.
  27. ^ a b Chen, Yufeng; Wang, Hongqiang; Helbling, E. Farrell; Jafferis, Noah T.; Zufferey, Raphael; Ong, Aaron; Ma, Kevin; Gravish, Nicholas; Chirarattananon, Pakpong; Kovac, Mirko; Wood, Robert J. (2017-10-25). "A biologically inspired, flapping-wing, hybrid aerial-aquatic microrobot". Science Robotics. 2 (11). doi:10.1126/scirobotics.aao5619. ISSN 2470-9476. PMID 33157886.
  28. ^ a b Buss, Nicole; Yasa, Oncay; Alapan, Yunus; Akolpoglu, Mukrime Birgul; Sitti, Metin (2020). "Nanoerythrosome-functionalized biohybrid microswimmers". APL Bioengineering. 4 (2): 026103. doi:10.1063/1.5130670. PMC 7141839. PMID 32548539. Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License.
  29. ^ Delasoie, Joachim; Schiel, Philippe; Vojnovic, Sandra; Nikodinovic-Runic, Jasmina; Zobi, Fabio (25 May 2020). "Photoactivatable Surface-Functionalized Diatom Microalgae for Colorectal Cancer Targeted Delivery and Enhanced Cytotoxicity of Anticancer Complexes". Pharmaceutics. 12 (5). MDPI AG: 480. doi:10.3390/pharmaceutics12050480. ISSN 1999-4923. PMC 7285135. PMID 32466116. Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License.
  30. ^ Tramontano, Chiara; Chianese, Giovanna; Terracciano, Monica; de Stefano, Luca; Rea, Ilaria (2020-09-28). "Nanostructured Biosilica of Diatoms: From Water World to Biomedical Applications". Applied Sciences. 10 (19). MDPI AG: 6811. doi:10.3390/app10196811. ISSN 2076-3417. Material was copied from this source, which is available under a Creative Commons Attribution 4.0 International License.
  31. ^ Hosseinidoust, Zeinab; Mostaghaci, Babak; Yasa, Oncay; Park, Byung-Wook; Singh, Ajay Vikram; Sitti, Metin (2016). "Bioengineered and biohybrid bacteria-based systems for drug delivery". Advanced Drug Delivery Reviews. 106 (Pt A): 27–44. doi:10.1016/j.addr.2016.09.007. PMID 27641944.
  32. ^ Schwarz, Lukas; Medina-Sánchez, Mariana; Schmidt, Oliver G. (2017). "Hybrid Bio Micromotors". Applied Physics Reviews. 4 (3): 031301. Bibcode:2017ApPRv...4c1301S. doi:10.1063/1.4993441.
  33. ^ Bastos-Arrieta, Julio; Revilla-Guarinos, Ainhoa; Uspal, William E.; Simmchen, Juliane (2018). "Bacterial Biohybrid Microswimmers". Frontiers in Robotics and AI. 5: 97. doi:10.3389/frobt.2018.00097. PMC 7805739. PMID 33500976.
  34. ^ Erkoc, Pelin; Yasa, Immihan C.; Ceylan, Hakan; Yasa, Oncay; Alapan, Yunus; Sitti, Metin (2019). "Mobile Microrobots for Active Therapeutic Delivery". Advanced Therapeutics. 2. doi:10.1002/adtp.201800064. S2CID 88204894.
  35. ^ Park, Byung-Wook; Zhuang, Jiang; Yasa, Oncay; Sitti, Metin (2017). "Multifunctional Bacteria-Driven Microswimmers for Targeted Active Drug Delivery". ACS Nano. 11 (9): 8910–8923. doi:10.1021/acsnano.7b03207. PMID 28873304.
  36. ^ Singh, Ajay Vikram; Hosseinidoust, Zeinab; Park, Byung-Wook; Yasa, Oncay; Sitti, Metin (2017). "Microemulsion-Based Soft Bacteria-Driven Microswimmers for Active Cargo Delivery". ACS Nano. 11 (10): 9759–9769. doi:10.1021/acsnano.7b02082. PMID 28858477.
  37. ^ Weibel, D. B.; Garstecki, P.; Ryan, D.; Diluzio, W. R.; Mayer, M.; Seto, J. E.; Whitesides, G. M. (2005). "Microoxen: Microorganisms to move microscale loads". Proceedings of the National Academy of Sciences. 102 (34): 11963–11967. Bibcode:2005PNAS..10211963W. doi:10.1073/pnas.0505481102. PMC 1189341. PMID 16103369.
  38. ^ Yasa, Oncay; Erkoc, Pelin; Alapan, Yunus; Sitti, Metin (2018). "Microalga-Powered Microswimmers toward Active Cargo Delivery". Advanced Materials. 30 (45): e1804130. Bibcode:2018AdM....3004130Y. doi:10.1002/adma.201804130. PMID 30252963. S2CID 52823884.
  39. ^ Xu, Haifeng; Medina-Sánchez, Mariana; Magdanz, Veronika; Schwarz, Lukas; Hebenstreit, Franziska; Schmidt, Oliver G. (2018). "Sperm-Hybrid Micromotor for Targeted Drug Delivery". ACS Nano. 12 (1): 327–337. arXiv:1703.08510. doi:10.1021/acsnano.7b06398. PMID 29202221.
  40. ^ Chen, Chuanrui; Chang, Xiaocong; Angsantikul, Pavimol; Li, Jinxing; Esteban-Fernández De Ávila, Berta; Karshalev, Emil; Liu, Wenjuan; Mou, Fangzhi; He, Sha; Castillo, Roxanne; Liang, Yuyan; Guan, Jianguo; Zhang, Liangfang; Wang, Joseph (2018). "Chemotactic Guidance of Synthetic Organic/Inorganic Payloads Functionalized Sperm Micromotors". Advanced Biosystems. 2. doi:10.1002/adbi.201700160. S2CID 103392074.
  41. ^ a b Alapan, Yunus; Yasa, Oncay; Yigit, Berk; Yasa, I. Ceren; Erkoc, Pelin; Sitti, Metin (2019). "Microrobotics and Microorganisms: Biohybrid Autonomous Cellular Robots". Annual Review of Control, Robotics, and Autonomous Systems. 2: 205–230. doi:10.1146/annurev-control-053018-023803. S2CID 139819519.
  42. ^ Wu, Zhiguang; Li, Tianlong; Li, Jinxing; Gao, Wei; Xu, Tailin; Christianson, Caleb; Gao, Weiwei; Galarnyk, Michael; He, Qiang; Zhang, Liangfang; Wang, Joseph (2014). "Turning Erythrocytes into Functional Micromotors". ACS Nano. 8 (12): 12041–12048. doi:10.1021/nn506200x. PMC 4386663. PMID 25415461.
  43. ^ Alapan, Yunus; Yasa, Oncay; Schauer, Oliver; Giltinan, Joshua; Tabak, Ahmet F.; Sourjik, Victor; Sitti, Metin (2018). "Soft erythrocyte-based bacterial microswimmers for cargo delivery". Science Robotics. 3 (17). doi:10.1126/scirobotics.aar4423. PMID 33141741. S2CID 14003685.
  44. ^ Zhuang, Jiang; Sitti, Metin (2016). "Chemotaxis of bio-hybrid multiple bacteria-driven microswimmers". Scientific Reports. 6: 32135. Bibcode:2016NatSR...632135Z. doi:10.1038/srep32135. PMC 4995368. PMID 27555465.
  45. ^ Zhuang, Jiang; Wright Carlsen, Rika; Sitti, Metin (2015). "PH-Taxis of Biohybrid Microsystems". Scientific Reports. 5: 11403. Bibcode:2015NatSR...511403Z. doi:10.1038/srep11403. PMC 4466791. PMID 26073316.
  46. ^ Felfoul, Ouajdi; Mohammadi, Mahmood; Taherkhani, Samira; De Lanauze, Dominic; Zhong Xu, Yong; Loghin, Dumitru; Essa, Sherief; Jancik, Sylwia; Houle, Daniel; Lafleur, Michel; Gaboury, Louis; Tabrizian, Maryam; Kaou, Neila; Atkin, Michael; Vuong, Té; Batist, Gerald; Beauchemin, Nicole; Radzioch, Danuta; Martel, Sylvain (2016). "Magneto-aerotactic bacteria deliver drug-containing nanoliposomes to tumour hypoxic regions". Nature Nanotechnology. 11 (11): 941–947. Bibcode:2016NatNa..11..941F. doi:10.1038/nnano.2016.137. PMC 6094936. PMID 27525475.
  47. ^ "Algae micromotors join the ranks for targeted drug delivery". Chemical & Engineering News. Retrieved 19 October 2022.
  48. ^ Zhang, Fangyu; Zhuang, Jia; Li, Zhengxing; Gong, Hua; de Ávila, Berta Esteban-Fernández; Duan, Yaou; Zhang, Qiangzhe; Zhou, Jiarong; Yin, Lu; Karshalev, Emil; Gao, Weiwei; Nizet, Victor; Fang, Ronnie H.; Zhang, Liangfang; Wang, Joseph (22 September 2022). "Nanoparticle-modified microrobots for in vivo antibiotic delivery to treat acute bacterial pneumonia". Nature Materials. 21 (11): 1324–1332. Bibcode:2022NatMa..21.1324Z. doi:10.1038/s41563-022-01360-9. ISSN 1476-4660. PMC 9633541. PMID 36138145.
  49. ^ Zhang, Fangyu; Li, Zhengxing; Duan, Yaou; Abbas, Amal; Mundaca-Uribe, Rodolfo; Yin, Lu; Luan, Hao; Gao, Weiwei; Fang, Ronnie H.; Zhang, Liangfang; Wang, Joseph (28 September 2022). "Gastrointestinal tract drug delivery using algae motors embedded in a degradable capsule". Science Robotics. 7 (70): eabo4160. doi:10.1126/scirobotics.abo4160. ISSN 2470-9476. PMC 9884493. PMID 36170380. S2CID 252598190.
  50. ^ Schmidt, Christine K.; Medina-Sánchez, Mariana; Edmondson, Richard J.; Schmidt, Oliver G. (5 November 2020). "Engineering microrobots for targeted cancer therapies from a medical perspective". Nature Communications. 11 (1): 5618. Bibcode:2020NatCo..11.5618S. doi:10.1038/s41467-020-19322-7. ISSN 2041-1723. PMC 7645678. PMID 33154372.
  51. ^ Thompson, Joanna. "These tiny magnetic robots can infiltrate tumors — and maybe destroy cancer". Inverse. Retrieved 21 November 2022.
  52. ^ Gwisai, T.; Mirkhani, N.; Christiansen, M. G.; Nguyen, T. T.; Ling, V.; Schuerle, S. (26 October 2022). "Magnetic torque–driven living microrobots for increased tumor infiltration". Science Robotics. 7 (71): eabo0665. bioRxiv 10.1101/2022.01.03.473989. doi:10.1126/scirobotics.abo0665. ISSN 2470-9476. PMID 36288270. S2CID 253160428.