Next Article in Journal
Mechanical and Fracture Properties of Steel/GFRP Hybrid Panels for an Improved Moveable Weir after Exposure to Accelerated Natural Environmental Conditions
Next Article in Special Issue
Low-Cost Rapid Fabrication of Conformal Liquid-Metal Patterns
Previous Article in Journal
Effect of Soil Reinforcement on Tunnel Deformation as a Result of Stress Relief
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Rotation of Liquid Metal Droplets Solely Driven by the Action of Magnetic Fields

1
CAS Key Laboratory of Mechanical Behavior and Design of Materials, Department of Precision Machinery and Precision Instrumentation, University of Science and Technology of China, Hefei 230026, China
2
School of Mechanical, Materials, Mechatronic and Biomedical Engineering, University of Wollongong, Wollongong, NSW 2522, Australia
3
School of Mechanical Engineering and Automation, State Key Laboratory of Robotics and System, Harbin Institute of Technology Shenzhen Graduate School, Shenzhen 518055, China
4
College of Mechanical and Electrical Engineering, Soochow University, Suzhou 215006, China
*
Authors to whom correspondence should be addressed.
Appl. Sci. 2019, 9(7), 1421; https://doi.org/10.3390/app9071421
Submission received: 25 February 2019 / Revised: 2 April 2019 / Accepted: 2 April 2019 / Published: 4 April 2019
(This article belongs to the Special Issue Applications of Liquid Metals)

Abstract

:
The self-rotation of liquid metal droplets (LMDs) has garnered potential for numerous applications, such as chip cooling, fluid mixture, and robotics. However, the controllable self-rotation of LMDs utilizing magnetic fields is still underexplored. Here, we report a novel method to induce self-rotation of LMDs solely utilizing a rotating magnetic field. This is achieved by rotating a pair of permanent magnets around a LMD located at the magnetic field center. The LMD experiences Lorenz force generated by the relative motion between the droplet and the permanent magnets and can be rotated. Remarkably, unlike the actuation induced by electrochemistry, the rotational motion of the droplet induced by magnetic fields avoids the generation of gas bubbles and behaves smoothly and steadily. We investigate the main parameters that affect the self-rotational behaviors of LMDs and validate the theory of this approach. We further demonstrate the ability of accelerating cooling and a mixer enabled by the self-rotation of a LMD. We believe that the presented technique can be conveniently adapted by other systems after necessary modifications and enables new progress in microfluidics, microelectromechanical (MEMS) applications, and micro robotics.

1. Introduction

“Liquid metal” such as gallium and its several alloys exhibit a liquid state at room temperature [1,2,3]. Droplets of such liquid metal have shown to be platforms for applications such as stretchable and reconfigurable electronics [3,4,5,6], microfluidics actuators [2,7,8], as well as forming three-dimensional structures [9,10,11]. Liquid metal droplets (LMDs) have demonstrated many unique properties, for instance, large surface tension, favorable thermal and electrical conductivity, strong stability, and extremely excellent biosafety compared with mercury [12,13]. Over the past few years, benefitting from the flexibility and operability of liquid metal surfaces, investigations of formation, actuation, and application of their droplets have gained momentum [2,14,15,16,17,18,19,20,21,22,23]. As has been recently reported, LMDs can act as self-fueled motors, paving the way without human intervention [14,23]; also, a LMD can be used to drive a wheeled robot by changing its center of gravity under electrical fields [22]. Moreover, a potential gradient was applied to explore inducing the Marangoni flow along the surface of LMDs and making microactuators in microfluidics [2,24,25,26,27]. Further studies have demonstrated that LMDs coating modest nanoparticles can be propelled by bubbles generated through photochemical reactions [28,29,30].
Chaotic advection is the key mechanism for enabling applications such as heat transfer, fluid mixing, and chemical reaction improvement [24,31,32,33,34,35], in particular, in some areas including microfluidic systems, chemical and biological transport, etc., smooth and steady methods for inducing chaotic advection are needed, and inexpensive and simple systems are urgently required [36,37,38,39]. Self-rotation of LMDs with negligible solubility in most solvents may be a promising candidate for providing a solution [7]. Nevertheless, to the best knowledge of the authors, there is still a lack of investigations focused on the self-rotational motion of LMDs.
Self-rotation of LMDs has the potential to be widely used in fluid cooling and mixing. Therefore, we have been motivated to explore novel methods for inducing self-rotation of LMDs that are smooth, simple, steady, and, especially, that do not introduce undesired violent chemical reactions. According to our previous report, we introduced a simple and violent chemical reaction-free method in which we utilized Lorenz force induced by magnetic fields to control the locomotion of LMDs [38]. Inspired by this, we report here the self-rotation of EGaIn droplets (consisting of 75% gallium and 25% indium) which are solely driven by magnetic fields without mixing or coating ferromagnetic particles. The relative motion of the magnetic fields and LMDs generates an eddy current in the droplet and further induces the Lorenz force to actuate the self-rotation of the droplet. The theory behind the method was developed and the experiments conducted to validate this approach. Moreover, we demonstrate applications of accelerating cooling and mixing liquids based on the self-rotational LMDs.

2. Materials and Methods

Materials and instrumentation: EGaIn liquid metal and sodium hydroxide (NaOH) were purchased from Santech Materials Co. Ltd, China. An electrolyte solution of NaOH and hydrochloric acid (HCl) were introduced to remove the oxide layer on the surface of LMDs, where NaOH solution was prepared by dissolving solid sodium hydroxide particles in deionized (DI) water, and HCl solution was prepared by diluting concentrated hydrochloric acid with DI water. A small amount (~2 mg) of fine phosphors (Juen technology Co. Ltd, China) was sprinkled on the LMDs to create easily identifiable points for calculating the rotational speed of the LMDs. Blue and yellow dyes were obtained by diluting edible dye (1:5, SUGARMAN) and then dripped into a quartz tube using a syringe pump (LSP02-1B, LONGERPUMP). Self-rotation videos of LMDs were captured using a digital single lens reflex camera (Canon 5D Mark II) equipped with a macro lens (Sigma 105mm 1:2.8 DG Macro HSM). The sequential snapshots were extracted from these videos.
Experimental Setup: The experimental setup is illustrated in Figure 1a. A pair of permanent magnets were fixed to an aluminum frame, and the aluminum frame was connected to the output shaft of a DC motor (Leadshine 57HS09) whose speed and direction were controlled by a microcontroller unit (MCU, Arduino Carduino UNO R3). An EGaIn LMD was placed in a quartz tube (diameter: 8 mm, height: 15 mm) filled with NaOH or HCl solution, and the quartz tube was placed at the center point between the magnets.
Mechanism: The mechanism of self-rotation of LMDs is shown in Figure 1b, in which we hypothesize that by moving two permanent magnets around the droplet within an aqueous solution, an EGaIn droplet can be actuated to self-rotate after the introduction of eddy current. On account of the large surface tension of liquid metal, the EGaIn droplet (diameter < 5 mm) immersed in the aqueous solution can be considered as a sphere. In addition, to simplify the model, the EGaIn droplet is equivalent to several parallel coils located at a different longitude of the droplet. Before the external magnetic field rotates, the magnetic flux through the equivalent coils is almost a constant. When the external magnetic field starts to rotate, relative motion (in other words, phase shift) is formed between the magnetic field and the EGaIn droplet, which further induces the change of the magnetic flux through the equivalent coils φ and eddy current I within the equivalent coils. The eddy current I can be explicitly expressed as
I = d Φ R e d t
where φ is determined by the magnetic flux density B as well as the equivalent coils area S, which can be written as φ =B·S, and Re represents the equivalent resistance of the coils. Subsequently, an induced magnetic field is engendered by the eddy current to hinder the change of magnetic flux, according to the Lenz’s law [40,41]. In other words, assuming that the external magnetic field rotates counterclockwise, the induced magnetic field will generate a clockwise torque to hinder the rotation of the magnetic field, no matter whether the magnetic flux through the coils increases or decreases. It is well known that forces are mutual actions of two bodies [38,42,43,44,45,46], that is, when the equivalent coils impede the external magnetic field, the external magnetic field also exerts a torque Me of the equal magnitude and opposite direction on the EGaIn droplet. According to Ampere’s law, the torque Me can be expressed as
M e = 2 0 π R sin l R I d l × B
where R is the radius of the equivalent coil, and l is the length of the equivalent coil, respectively [42,43,44,45,46]. The torque Me enables the droplet to overcome the friction and commence to self-rotate.

3. Results and Discussions

We established an actuating platform to examine the hypothesis and investigate the self-rotational behaviors of the droplet (Figure 1a). Figure 1c shows the self-rotation of an EGaIn droplet submerged in 0.5 mol/L NaOH solution induced by the magnetic field (also see Video S1); the measured magnetic flux density was ~1 kGs at the center point between the magnets, and we set the DC motor speed to 420 RPM. Then, the EGaIn droplet self-rotated following the same direction of rotation as the magnets (counterclockwise) with a speed of ~100 RPM, which aligns with our analysis given in Figure 1b.
On the basis of successfully demonstrating of magnetic field driven self-rotation of EGaIn droplet, we carried out a set of experiments to ascertain the performance of self-rotation. We found that motor speed, the LMDs size, as well as the NaOH solution concentration are the three main factors that affect the self-rotational speed. In Figure 2a, we can see that the self-rotational speed increases along with the increase of the DC motor rotational speed since larger rates of magnetic flux change dφ can be induced by higher motor speeds. A larger rate of magnetic flux change further induces a larger eddy current which eventually converts into a greater Me (see Equation 2). We also studied the influence of the magnetic flux density B on the self-rotational performance, as shown in Figure 2a. Obviously, a higher B can induce a larger torque and lead to a larger self-rotational speed.
Figure 2b shows that the self-rotational speed increases along with the increase of the size of the EGaIn droplets until the droplet volume reaches ~0.06 mL, and then the speed gradually decreases and eventually remains stable. This is probably due to that, in a larger droplet, the increasing S and decreasing Re enlarge the driving force. When the volume of the EGaIn droplets is larger than 0.06 mL, the friction between the droplet and the sidewall may give a negative effect on the rotational speed of the droplet which has been elucidated by repeating the experiment with a bigger tube, as discussed in Supporting Information S1. As the rotational speed decreases, the relative rotational speed of the droplet and the magnetic field increases (that is, dφ increases), which then can cause the increase of I, the driven force, and the rotational speed. Next, increasing the rotational speed reduces the relative speed, and finally, the driving force and resistance are balanced, which is manifested in the fact that the rotation speed was basically stable during our experiments. Moreover, the depth of EGaIn droplet immersion in the NaOH solution (represented by the percentage of the height of the droplet immersed) also influences the speed of rotation (Figure 2b). For smaller droplets (<0.06 mL), the difference in rotational speed is not obvious for different immersion depths. However, for droplets larger than 0.06 mL, a larger immersion depth leads to a higher speed of rotation. This is probably due to the fact that despite the increase in viscous friction for larger immersion depths, the solid oxide layer formed on the EGaIn droplet surface can be efficiently removed by NaOH solution, thus a slip layer between the droplet and the container wall is formed. This slip layer can reduce the friction like a lubricant, and thus increase the rotational speed. Interestingly, we observed that when the droplet was 100% immersed, the droplet reached its maximum rotational speed when the volume was increased to 0.04 mL, which is smaller than that of other immersion depths (0.06, 0.06, and 0.08 mL). We believe this is due to the fact that the significant increase in rotational speed at 100% immersion flattens the droplets and, therefore, increases the friction between the droplets and the tube.
As shown in Figure 2c, NaOH concentration also affects the self-rotational speed of the LMDs. We found that using NaOH solution with a high concentration can lead to a faster rotational speed until the concentration of NaOH solution reaches 0.03 mol/L. When the concentration of NaOH solution exceeds 0.03 mol/L, the rotational speed of the droplets no longer increases with the increase of concentration and remains almost constant. No self-rotation was observed in our experiments when we reduced the NaOH solution concentration to zero. That might be due to the fact that without NaOH, the oxide layer cannot be removed, and the droplet becomes wrinkled [14,17], so the friction between the droplet, the tube, and the solution eventually increases. With the increase of the concentration of NaOH solution, the oxide layer was gradually removed, the friction gradually decreased, and the rotational speed increased. However, when the concentration reaches the threshold (0.03 mol/L), the oxide layer is almost completely removed, and the speed no longer increases with the concentration increase. Considering HCl solution can also be used to remove the oxide layer on the surface of EGaIn droplets [17,47], as shown in Figure 2d, we further studied the influence of HCl concentration on the rotational speed of EGaIn droplets. We observed the increase in rotational speed when increasing the concentration of HCl solution from 0 to 0.5 mol/L.
Use in cooling systems is an import potential application of liquid metal, and here we demonstrate the ability of liquid metal self-rotation to accelerate liquid cooling. As shown in Figure 3a, we heated the 0.5 mol/L NaOH solution and 0.08 mL EGaIn in a tube with a heat gun (DH-HG2-2000, Delixi Electric). When the temperature of the solution reached about 40 °C, a large number of bubbles were separated from the solution similar to boiling. We stopped heating and rotated the motor at 420 RPM and measured the temperature of the solution every 5 seconds. For comparison, the other tube was tested in the same way except that the permanent magnets in the device were removed, that is, the LMD did not self-rotate as the motor rotated. The temperature change is shown in Figure 3b; the group in which the EGaIn droplet self-rotated cooled significantly faster than the group in which the EGaIn droplet did not rotate. After about 520 s, the group of rotated EGaIn cooled to room temperature (~24.2 °C), and after another 130 s, the other group cooled to room temperature.
Based on the fact that LMDs are negligibly soluble in most liquids, possess great surface tension, and keep almost unmixed with the solvent [7], LMDs have great potential for application as a mixer. Therefore, we further investigated its application as a mixer based on the self-rotation phenomenon of LMDs in a rotating magnetic field, as shown in Figure 4a (see also Video S2). We applied the self-rotation of EGaIn droplets to mix two droplets of the same volume, similar viscosity, and different colors. We added a drop of blue dye and a drop of yellow dye (volume of 0.05 mL, respectively) into the solution while the 0.08 mL EGaIn droplet was rotating in the NaOH (0.5 mol/L) solution, as shown in Figure 4b. The motor speed was 420 RPM, and the magnetic flux density at the center of the droplet was ~0.8 kGs. The self-rotational speed of EGaIn droplet after mixing was ~70 RPM which is almost the same as that when the EGaIn droplet is not acting as a mixer, as shown in Figure 2b. Figure 4c,d demonstrates that the two drops of dye can be mixed along with the rotating EGaIn droplet. This mixer is smooth, quick, and can be easily implemented into a MEMS platform.

4. Conclusions

We have demonstrated a novel method to manipulate the self-rotation of LMDs by solely utilizing magnetic fields which is smooth, simple, steady and, most importantly, does not produce a violent chemical reaction. The relative motion of the magnetic fields and LMDs generates an eddy current in the droplets and further induces the Lorentz force to actuate the self-rotation of the droplets. The motor speed, the LMD sizes, and the concentration of NaOH solution can be easily manipulated to regulate the self-rotational speed of droplet. Moreover, we demonstrated that such a technique can be used for the application of accelerating the cooling and mixing liquids. Nonetheless, bulky rotating magnets and a motor are still required in our current platform; however, we believe that introducing programmed electromagnetic fields can be helpful in resolving this problem in our future work. As such, utilizing magnetic fields to induce the self-rotation of LMDs could widely expand the application of LMDs to be used as MEMS devices.

Supplementary Materials

The following are available online at https://www.mdpi.com/2076-3417/9/7/1421/s1, Figure S1: Self-rotation of EGaIn droplets in different diameter tubes, Video S1: Self-rotation of liquid metal droplet, Video S2: Mixer based on self-rotation of liquid metal droplet.

Author Contributions

S.J., the first author, conceived the idea, performed the experiments, analyzed the data and wrote the paper. S.-Y.T. conceived the idea, analyzed the data and wrote the paper. S.Z. (Sizepeng Zhao) performed the experiments. Z.F. conceived the idea and analyzed the data. H.C. conceived the idea. W.L. analyzed the data and wrote the paper. X.L. conceived the idea, analyzed the data and wrote the paper. S.Z. (Shiwu Zhang) conceived the idea, analyzed the data and wrote the paper.

Funding

The authors acknowledge support from the National Natural Science Foundation of China (No. 51828503, U1713206, 61503270, 61873339), and the Shenzhen Science and Innovation Committee (Reference No. JCYJ20160427183958817). Dr. Shi-Yang Tang is the recipient of the Vice-Chancellor’s Postdoctoral Research Fellowship funded by the University of Wollongong.

Conflicts of Interest

There are no conflicts of interest to declare.

References

  1. Dickey, M.D. Stretchable and Soft Electronics using Liquid Metals. Adv. Mater. 2017, 29, 1606425. [Google Scholar] [CrossRef]
  2. Tang, S.-Y.; Khoshmanesh, K.; Sivan, V.; Petersen, P.; O’Mullane, A.P.; Abbott, D.; Mitchell, A.; Kalantar-Zadeh, K. Liquid metal enabled pump. Proc. Natl. Acad. Sci. USA 2014, 111, 3304–3309. [Google Scholar] [CrossRef] [Green Version]
  3. Ladd, C.; So, J.-H.; Muth, J.; Dickey, M.D. 3D Printing of Free Standing Liquid Metal Microstructures. Adv. Mater. 2013, 25, 5081–5085. [Google Scholar] [CrossRef] [PubMed]
  4. Markvicka, E.J.; Bartlett, M.D.; Huang, X.; Majidi, C. An autonomously electrically self-healing liquid metal–elastomer composite for robust soft-matter robotics and electronics. Nat. Mater. 2018, 17, 618–624. [Google Scholar] [CrossRef]
  5. Reichel, K.S.; Lozada-Smith, N.; Joshipura, I.D.; Ma, J.; Shrestha, R.; Mendis, R.; Dickey, M.D.; Mittleman, D.M. Electrically reconfigurable terahertz signal processing devices using liquid metal components. Nat. Commun. 2018, 9, 4202. [Google Scholar] [CrossRef] [PubMed]
  6. Palleau, E.; Reece, S.; Desai, S.C.; Smith, M.E.; Dickey, M.D. Self-Healing Stretchable Wires for Reconfigurable Circuit Wiring and 3D Microfluidics. Adv. Mater. 2013, 25, 1589–1592. [Google Scholar] [CrossRef] [PubMed]
  7. Khoshmanesh, K.; Tang, S.-Y.; Zhu, J.Y.; Schaefer, S.; Mitchell, A.; Kalantar-Zadeh, K.; Dickey, M.D. Liquid metal enabled microfluidics. Lab Chip 2017, 17, 974–993. [Google Scholar] [CrossRef] [PubMed]
  8. Shao, C.Y.; Kawai, Y.; Esashi, M.; Ono, T. Electrostatic actuator probe with curved electrodes for time-of-flight scanning force microscopy. Rev. Sci. Instrum. 2010, 81, 083702. [Google Scholar] [CrossRef] [PubMed]
  9. Koo, H.J.; So, J.H.; Dickey, M.D.; Velev, O.D. Towards all-soft matter circuits: Prototypes of quasi-liquid devices with memristor characteristics. Adv. Mater. 2011, 23, 3559–3564. [Google Scholar] [CrossRef] [PubMed]
  10. Jeong, S.H.; Hagman, A.; Hjort, K.; Jobs, M.; Sundqvist, J.; Wu, Z. Liquid alloy printing of microfluidic stretchable electronics. Lab Chip 2012, 12, 4657–4664. [Google Scholar] [CrossRef] [PubMed]
  11. Tabatabai, A.; Fassler, A.; Usiak, C.; Majidi, C. Liquid-phase gallium-indium alloy electronics with microcontact printing. Langmuir 2013, 29, 6194–6200. [Google Scholar] [CrossRef]
  12. Liu, T.Y.; Sen, P.; Kim, C.J.C.J. Characterization of Nontoxic Liquid-Metal Alloy Galinstan for Applications in Microdevices. J. Microelectromech. Syst. 2012, 21, 443–450. [Google Scholar] [CrossRef] [Green Version]
  13. Dickey, M.D. Emerging applications of liquid metals featuring surface oxides. ACS Appl. Mater. Interfaces 2014, 6, 18369–18379. [Google Scholar] [CrossRef] [PubMed]
  14. Zhang, J.; Yao, Y.; Sheng, L.; Liu, J. Self-fueled biomimetic liquid metal mollusk. Adv. Mater. 2015, 27, 2648–2655. [Google Scholar] [CrossRef] [PubMed]
  15. Gough, R.C.; Dang, J.H.; Moorefield, M.R.; Zhang, G.B.; Hihara, L.H.; Shiroma, W.A.; Ohta, A.T. Self-Actuation of Liquid Metal via Redox Reaction. ACS Appl. Mater. Interfaces 2016, 8, 6–10. [Google Scholar] [CrossRef] [PubMed]
  16. Junghoon, L.; Chang-Jin, K. Surface-tension-driven microactuation based on continuous electrowetting. J. Microelectromech. Syst. 2000, 9, 171–180. [Google Scholar] [CrossRef]
  17. Khan, M.R.; Eaker, C.B.; Bowden, E.F.; Dickey, M.D. Giant and switchable surface activity of liquid metal via surface oxidation. Proc. Natl. Acad. Sci. USA 2014, 111, 14047–14051. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Tang, S.Y.; Sivan, V.; Khoshmanesh, K.; O’Mullane, A.P.; Tang, X.; Gol, B.; Eshtiaghi, N.; Lieder, F.; Petersen, P.; Mitchell, A.; et al. Electrochemically induced actuation of liquid metal marbles. Nanoscale 2013, 5, 5949–5957. [Google Scholar] [CrossRef] [PubMed]
  19. Krupenkin, T.; Taylor, J.A. Reverse electrowetting as a new approach to high-power energy harvesting. Nat. Commun. 2011, 2, 448. [Google Scholar] [CrossRef] [Green Version]
  20. Bartlett, M.D.; Fassler, A.; Kazem, N.; Markvicka, E.J.; Mandal, P.; Majidi, C. Stretchable, High-k Dielectric Elastomers through Liquid-Metal Inclusions. Adv. Mater. 2016, 28, 3726–3731. [Google Scholar] [CrossRef]
  21. Tang, S.Y.; Joshipura, I.D.; Lin, Y.; Kalantar-Zadeh, K.; Mitchell, A.; Khoshmanesh, K.; Dickey, M.D. Liquid-Metal Microdroplets Formed Dynamically with Electrical Control of Size and Rate. Adv. Mater. 2016, 28, 604–609. [Google Scholar] [CrossRef] [PubMed]
  22. Wu, J.; Tang, S.Y.; Fang, T.; Li, W.; Li, X.; Zhang, S. A Wheeled Robot Driven by a Liquid-Metal Droplet. Adv. Mater. 2018, e1805039. [Google Scholar] [CrossRef] [PubMed]
  23. Mohammed, M.; Sundaresan, R.; Dickey, M.D. Self-Running Liquid Metal Drops that Delaminate Metal Films at Record Velocities. ACS Appl. Mater. Interfaces 2015, 7, 23163–23171. [Google Scholar] [CrossRef] [PubMed]
  24. Tang, S.Y.; Sivan, V.; Petersen, P.; Zhang, W.; Morrison, P.D.; Kalantar-zadeh, K.; Mitchell, A.; Khoshmanesh, K. Liquid Metal Actuator for Inducing Chaotic Advection. Adv. Funct. Mater. 2014, 24, 5851–5858. [Google Scholar] [CrossRef]
  25. Tang, S.Y.; Lin, Y.; Joshipura, I.D.; Khoshmanesh, K.; Dickey, M.D. Steering liquid metal flow in microchannels using low voltages. Lab Chip 2015, 15, 3905–3911. [Google Scholar] [CrossRef] [PubMed]
  26. Wissman, J.; Dickey, M.D.; Majidi, C. Field-Controlled Electrical Switch with Liquid Metal. Adv. Sci. 2017, 4, 1700169. [Google Scholar] [CrossRef]
  27. Russell, L.; Wissman, J.; Majidi, C. Liquid metal actuator driven by electrochemical manipulation of surface tension. Appl. Phys. Lett. 2017, 111, 254101. [Google Scholar] [CrossRef]
  28. Kim, D.; Lee, J.B. Magnetic-field-induced Liquid Metal Droplet Manipulation. J. Korean Phys. Soc. 2015, 66, 282–286. [Google Scholar] [CrossRef]
  29. Tang, X.K.; Tang, S.Y.; Sivan, V.; Zhang, W.; Mitchell, A.; Kalantar-zadeh, K.; Khoshmanesh, K. Photochemically induced motion of liquid metal marbles. Appl. Phys. Lett. 2013, 103, 174104. [Google Scholar] [CrossRef]
  30. Wang, L.; Liu, J. Electromagnetic rotation of a liquid metal sphere or pool within a solution. Proc. R. Soc. A Math. Phys. Eng. Sci. 2015, 471, 20150177. [Google Scholar] [CrossRef]
  31. Chertkov, M.; Lebedev, V. Boundary Effects on Chaotic Advection-Diffusion Chemical Reactions. Phys. Rev. Lett. 2003, 90, 134501. [Google Scholar] [CrossRef] [PubMed]
  32. Priye, A.; Hassan, Y.A.; Ugaz, V.M. Microscale Chaotic Advection Enables Robust Convective DNA Replication. Anal. Chem. 2013, 85, 10536–10541. [Google Scholar] [CrossRef] [PubMed]
  33. Tohidi, A.; Hosseinalipour, S.M.; Taheri, P.; Nouri, N.M.; Mujumdar, A.S. Chaotic advection induced heat transfer enhancement in a chevron-type plate heat exchanger. Heat Mass Transfer 2013, 49, 1535–1548. [Google Scholar] [CrossRef]
  34. Lefevre, A.; Mota, J.P.B.; Rodrigo, A.J.S.; Saatdjian, E. Chaotic advection and heat transfer enhancement in Stokes flows. Int. J. Heat Fluid 2003, 24, 310–321. [Google Scholar] [CrossRef]
  35. Solomon, T.H.; Mezic, I. Uniform resonant chaotic mixing in fluid flows. Nature 2003, 425, 376–380. [Google Scholar] [CrossRef] [PubMed]
  36. Stroock, A.D.; Dertinger, S.K.W.; Ajdari, A.; Mezic, I.; Stone, H.A.; Whitesides, G.M. Chaotic mixer for microchannels. Science 2002, 295, 647–651. [Google Scholar] [CrossRef] [PubMed]
  37. Sundararajan, P.; Stroock, A.D. Transport Phenomena in Chaotic Laminar Flows. Chem. Biomol. Eng. 2012, 3, 473–496. [Google Scholar] [CrossRef] [PubMed]
  38. Shu, J.; Tang, S.Y.; Feng, Z.H.; Li, W.H.; Li, X.P.; Zhang, S.W. Unconventional locomotion of liquid metal droplets driven by magnetic fields. Soft Matter 2018, 14, 7113–7118. [Google Scholar] [CrossRef]
  39. Li, X.; Xie, J.; Tang, S.Y.; Xu, R.; Li, X.; Li, W.; Zhang, S. A Controllable Untethered Vehicle Driven by Electrically Actuated Liquid Metal Droplets. IEEE Trans. Ind. Inform. 2018. [Google Scholar] [CrossRef]
  40. Priede, J.; Gerbeth, G. Spin-up instability of electromagnetically levitated spherical bodies. IEEE. Trans. Magn. 2000, 36, 349–353. [Google Scholar] [CrossRef]
  41. Graf, H.; Lauer, U.A.; Schick, F. Eddy-current induction in extended metallic parts as a source of considerable torsional moment. J. Magn. Reson Imaging 2006, 23, 585–590. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Koo, C.; LeBlanc, B.E.; Kelley, M.; Fitzgerald, H.E.; Huff, G.H.; Han, A. Manipulating Liquid Metal Droplets in Microfluidic Channels With Minimized Skin Residues Toward Tunable RF Applications. J. Microelectromech. Syst. 2015, 24, 1069–1076. [Google Scholar] [CrossRef]
  43. Khan, M.R.; Trlica, C.; So, J.H.; Valeri, M.; Dickey, M.D. Influence of water on the interfacial behavior of gallium liquid metal alloys. ACS Appl. Mater. Interfaces 2014, 6, 22467–22473. [Google Scholar] [CrossRef] [PubMed]
  44. Dang, J.H.; Gough, R.C.; Morishita, A.M.; Ohta, A.T.; Shiroma, W.A. Liquid-metal frequency-reconfigurable slot antenna using air-bubble actuation. Electron. Lett. 2015, 51, 1630–1631. [Google Scholar] [CrossRef]
  45. Beinerts, T.; Bojarevičs, A.; Bucenieks, I.; Gelfgat, Y.; Kaldre, I. Use of permanent magnets in electromagnetic facilities for the treatment of aluminum alloys. Metall. Mater. Trans. B 2016, 47, 1626–1633. [Google Scholar] [CrossRef]
  46. Dzelme, V.; Jakovics, A.; Bucenieks, I. Numerical modelling of liquid metal electromagnetic pump with rotating permanent magnets. Mater. Sci. Eng. 2018, 424, 012046. [Google Scholar] [CrossRef]
  47. Zavabeti, A.; Ou, J.Z.; Carey, B.J.; Syed, N.; Orrell-Trigg, R.; Mayes, E.L.H.; Xu, C.L.; Kavehei, O.; O’Mullane, A.P.; Kaner, R.B.; et al. A liquid metal reaction environment for the room-temperature synthesis of atomically thin metal oxides. Science 2017, 358, 332–335. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Self-rotation behavior of a liquid metal droplet (LMD). (a) Schematic of the self-rotation inducing setup. (b) Forces schematic of the equivalent coils in an EGaIn droplet; the magnetic field directions are indicated by the blue arrows, the magnetic field rotating direction (with a velocity of Vm) is indicated by the green arrow, the self-rotation direction of the EGaIn droplet (with a velocity of Vd) is indicated by the red arrow, and the forces experienced on the equivalent coils are indicated by the black arrows. (c) Continuous captures of the self-rotation of an EGaIn droplet (volume of 0.08 mL).
Figure 1. Self-rotation behavior of a liquid metal droplet (LMD). (a) Schematic of the self-rotation inducing setup. (b) Forces schematic of the equivalent coils in an EGaIn droplet; the magnetic field directions are indicated by the blue arrows, the magnetic field rotating direction (with a velocity of Vm) is indicated by the green arrow, the self-rotation direction of the EGaIn droplet (with a velocity of Vd) is indicated by the red arrow, and the forces experienced on the equivalent coils are indicated by the black arrows. (c) Continuous captures of the self-rotation of an EGaIn droplet (volume of 0.08 mL).
Applsci 09 01421 g001
Figure 2. Change of the self-rotational performance under different operating parameters. (a) Self-rotational speed vs. speed of the motor plot; blue, red, and yellow curves indicate different magnetic field densities at the center of EGaIn droplets, respectively. (b) Self-rotational speed vs. sizes of droplets plot; blue, red, yellow, and purple curves indicate different depths of EGaIn droplet immersion in the NaOH solution, respectively. (c) Self-rotational speed vs. concentration of NaOH plot. (d) Self-rotational speed vs. concentration of HCl plot.
Figure 2. Change of the self-rotational performance under different operating parameters. (a) Self-rotational speed vs. speed of the motor plot; blue, red, and yellow curves indicate different magnetic field densities at the center of EGaIn droplets, respectively. (b) Self-rotational speed vs. sizes of droplets plot; blue, red, yellow, and purple curves indicate different depths of EGaIn droplet immersion in the NaOH solution, respectively. (c) Self-rotational speed vs. concentration of NaOH plot. (d) Self-rotational speed vs. concentration of HCl plot.
Applsci 09 01421 g002
Figure 3. Cooling system based on a self-rotating LMD. (a) Schematic illustration of the cooling setup. (b) Temperature of solution vs. time plot.
Figure 3. Cooling system based on a self-rotating LMD. (a) Schematic illustration of the cooling setup. (b) Temperature of solution vs. time plot.
Applsci 09 01421 g003
Figure 4. Mixer based on a self-rotating LMD. (a) Schematic illustration of the mixing setup. (b–d) Continuous captures of mixing two drops of dye.
Figure 4. Mixer based on a self-rotating LMD. (a) Schematic illustration of the mixing setup. (b–d) Continuous captures of mixing two drops of dye.
Applsci 09 01421 g004

Share and Cite

MDPI and ACS Style

Shu, J.; Tang, S.-Y.; Zhao, S.; Feng, Z.; Chen, H.; Li, X.; Li, W.; Zhang, S. Rotation of Liquid Metal Droplets Solely Driven by the Action of Magnetic Fields. Appl. Sci. 2019, 9, 1421. https://doi.org/10.3390/app9071421

AMA Style

Shu J, Tang S-Y, Zhao S, Feng Z, Chen H, Li X, Li W, Zhang S. Rotation of Liquid Metal Droplets Solely Driven by the Action of Magnetic Fields. Applied Sciences. 2019; 9(7):1421. https://doi.org/10.3390/app9071421

Chicago/Turabian Style

Shu, Jian, Shi-Yang Tang, Sizepeng Zhao, Zhihua Feng, Haoyao Chen, Xiangpeng Li, Weihua Li, and Shiwu Zhang. 2019. "Rotation of Liquid Metal Droplets Solely Driven by the Action of Magnetic Fields" Applied Sciences 9, no. 7: 1421. https://doi.org/10.3390/app9071421

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