THEORY

Acoustic Levitation Theory

Acoustic levitation is the phenomenon that occurs when an object is kept suspended due to the pressure created by a vibrating plate. When the levitated object is very close to the plate, that being less than half a wavelength of the acoustic waves generated by the plate, the levitation is in the so-called Near-Field Acoustic Levitation, or NFAL, regime. On the other hand, if the object is further away, it is in the Far-Field Acoustic Levitation, or FFAL, regime. The correct theory behind acoustic levitation is not yet completely determined and the phenomenon is often described from different angles and for a specific regime. An effect which in particular is lacking a full explanation is the attractive force. This force appears at certain frequencies and distances from the vibrating surface.

With our experiment we hope to shed some light on the right theory of this phenomenon. You can find out more about our experimental ojectives and our set-up to see how:

Acoustic levitation can be achieved with many different types of objects. A form that is often investigated is that of a small object levitating between a vibrating surface and a fixed reflector [1,2,3]. In our experiment, we will instead focus on objects being levitated solely by a vibrating surface. In this case, it is the levitated object itself which may act as the reflector. The objects levitated in this case tends to be larger and have a more planar surface. It is with this type that we observe not only FFAL but also NFAL as well as attractive forces. With little experimentation under different gravity conditions and a lot of promising applications, this variant can benefit greatly of an investigation in hypergravity.

The conditions of the system, which affect the approximations that can be used and which sets of equations may be utilised, differs per system. These can be parameters including but not limited to: the reflectivity of the object, the elastic deformation of the plate, the sizes of the plate & object and the frequency the plate is driven at. Other approximations can regard the air inbetween the vibrating surface and the object, the material propagating the pressure. Examples of these properties are its viscosity or its thermodynamic properties, such as it being adiabatic [5,6,7] or isothermal [8,9,10].

Near-Field Acoustic Levitation appears when an object with a large surface area is suspended above a vibrating plate at micrometer distances, whether that be a piston, a rigid plate, or a flexural plate moving flexibly with the vibration. As the plate moves up and down it creates a periodically high and low pressure field between the plate & object. The averaged out pressure between the two exceeds the atmospheric pressure and hence the force is pressing the object up repulsively, counteracting the force of gravity [11]. The force pushing the object upwards increases exponentially with smaller levitation distance [12]. The way NFAL is described can largely be classified into two categories: Acoustic Radiation Pressure Theory and Squeeze Film Theory.

The acoustic radiation pressure comes from the second order term of the Taylor expansion of the fluid pressure; it is thus a non-linear effect [5]. Generally, we can see that the force can be approximated in a short-time and long-time regime, looking at the time-average over multiple harmonic cycles for the latter [10]. This theory can both be used for the Near-Field as well as the Far-Field scales. This theory can actually be used to describe both the NFAL as well as the FFAL regime.

According to Chu and Apfel, the acoustic radiation pressure can be divided in two categories depending on whether it is laterally confined or not [13]. In the inviscid limit, it is known as the Rayleigh acoustic pressure whereas in the geometric acoustic limit, it is known as the Langevin acoustic pressure.

The Rayleigh acoustic pressure acting on a partially absorbing target is given by $$ \lt p_{L} - p_{0} \gt = -\frac{\beta}{\alpha} \frac{\rho_{0}c_{0}^{2}}{1+\alpha^{2}\rho_{0}c_{0}^{2}} \lt E \gt,$$ where \( \alpha \) and \( \beta \) are the first two coefficients of the taylor expansion of the velocity response of the object with respect to pressure, \( p_{0} \) is the initial density of the medium, \(\lt E \gt\) is the Eulerian mean energy density, and \( c_{0} \) is the speed of sound in the medium [13].

Conversely, the Langevin acoustic pressure (neglecting the effect of gravity) can be found using $$ \lt p_{L} - p_{0} \gt = \lt E_{L} \gt,$$ where \(\lt E_{L} \gt\) is the Lagrangian mean energy density [13].

In the case of an adiabatic, incompressible, and inviscid liquid, the acoustic levitation force acting on a rigid object in the long-time regime can be written as $$ \boldsymbol{F_{rad}} = \int_{S_{0}} \lt p_{2} \gt \boldsymbol{n} dS, $$ neglecting the gravitational field. Here \(\lt p_{2} \gt\) is the time-averaged 2nd order pressure term [5]. Expanding on this idea by accounting for boundary conditions with respect to the pressure, a numerical model can be produced using COMSOL to show that in the near-field the force increases almost exponentially as the levitation height decreases [12] whereas the force 'spikes' periodically according to the levitation height in the far-field [14].

The squeeze film theory describes the levitation force at short distances as a result of pressure fluctuations wherein the average inwards pressure amplitude is larger than the average outwards pressure amplitude. This causes a difference in the net time-averaged pressure experienced by the levitated object, causing a levitating force [15]. This theory takes viscous effects into account and makes use of the Reynold's equations [16].

One factor to consider for squeeze film theory is the dimensionless so-called ‘squeeze number’. This is given by $$ \sigma = \frac{12 \omega \mu_{0} L}{p_{0} C_{0}^{2} Z^{2}}, $$ where \(\omega\) is the vibration frequency, \(\mu\) is the gas viscosity, \(p\) is the pressure, \(L\) is the object diameter, \(C\) is the speed of sound, and \(Z\) is the levitation height [8].

For an isothermal set-up with a small squeeze number, the Reynold’s solution for the time-averaged levitation force per unit area is given by $$\frac{5\varepsilon^{2}}{4\beta}[\frac{\beta cos(\beta) + \beta cosh(\beta) - sinh(\beta) - sin(\beta)}{cos(\beta) + cosh(\beta)}],$$ where \( \beta \) is the square root of \( \frac{\sigma}{2} \) and \( \varepsilon \) is \(\lt \lt 1 \) [8].

If the set-up has a larger squeeze number or is not isothermal in nature, the levitation force would instead be obtained by solving the equation: $$\frac{\partial}{\partial x} \left(\frac{p^{1/n} Z^{3}}{\mu} \frac{\partial p}{\partial x} \right) = \sigma \frac{\partial}{\partial t} \left( p^{1/n} Z \right),$$ where \(\mu\) is the gas viscosity [8].

Far-Field Acoustic Levitation describes the state of acoustic levitation wherein the levitated object is much larger than the acoustic wavelength. This method of levitation relies on the formation of a standing wave between the object’s surface and one or more transducers. This creates what is called Standing Wave Acoustic Levitation, or SWAL. To accomplish this, typically objects need to be in a resonance state - meaning they have to be located at an integer number of half-wavelengths away from the transducers [17].

For an object experiencing far-field acoustic levitation, the time-averaged radiation pressure is given by $$ \lt p_{r} \gt = \frac{1}{2 \rho_{0} c_{0}^{2}} \lt p^{2} \gt - \frac{\rho_{0} \lt \boldsymbol{u} \cdot \boldsymbol{u} \gt}{2}, $$ where \( \rho_{0} \) is the fluid density and \( c_{0} \) is the sound velocity in the fluid [17].

From the radiation pressure, the total radiation force can be obtained using, $$ \boldsymbol{F_{rad}} = \int_{S_{0}} \lt p_{2} \gt \boldsymbol{n} dS, $$ where \( \boldsymbol{n} \) is the surface normal vector and \( S_{0} \) is the integration surface [17].

In this regime, the shape of the object is known to affect its levitated behaviour. For instance, more vertically curved objects are known to have a higher horizontal stability while too much curvature can result in a lower vertical radiation force [17]. Similarly, objects with large radii of curvature are also known to levitate at slightly lower heights [18].

In contrast to all the previously mentioned forms of acoustic levitation, attractive acoustic levitation is used with the levitated object below rather than above the vibrating surface - resulting in the object being attracted to the surface. While the theory behind this phenomenon is not yet fully understood, several observations have been made that can be used to predict certain behaviours.

For instance, [18,19] have observed that the levitation force seems to be attractive for transducers and objects with diameters much smaller than the acoustic wavelength and vice-versa, indicating the possible existence of a ‘transition diameter’. Similarly, a ‘transition distance’ must also exist since the object never physically touches the transducer. The location of this transition distance depends both on the vibration frequency [21] and on the vibration amplitude, as stated by [19] who were able to model the transition using an axisymmetric CFD simulation rather than an acoustic model.

As is the case for far-field acoustic levitation, a horizontal restoring force can also be observed for attractive acoustic levitation. This force can be attributed to acoustic streaming which causes the time-averaged air pressure between the object and transducer to be lower than its ambient counterpart - essentially causing a self-centering effect [20]. In addition, a self-orienting effect was also observed, which can be calculated using $$ \tau_{rad} = - \int_{S_{0}} p_{rad} \boldsymbol{r} \times \boldsymbol{n} dS, $$ where \( \tau_{rad} \) is the radiation induced torque, \( \boldsymbol{n} \) is the surface normal vector, \( S_{0} \) is the integration surface, and \( \boldsymbol{r} \) is the unit vector pointing from the centre of mass of the object to its surface [20].

References

[1] Xie, W.J., Cao, C.D., Lü, Y.J., Hong, Z.Y., and Wei, B. (2006). Acoustic method for levitation of small living animals, Appl. Phys. Lett. 89, 214102

[2] Sagoff, J. (2012, September 12). No magic show: Real-world levitation to inspire better pharmaceuticals. Retrieved April 04, 2021, from https://www.anl.gov/article/no-magic-show-realworld-levitation-to-inspire-better-pharmaceuticals

[3] Hasegawa, K., Watanabe, A., & Abe, Y. (2019). Acoustic Manipulation of Droplets under Reduced Gravity. Scientific Reports, 9(1). doi:10.1038/s41598-019-53281-4

[4] Langlois, W. E. (1962). Isothermal squeeze films. Quarterly of Applied Mathematics, 20(2), 131-150. doi:10.1090/qam/99963

[5] Andrade, M. A., Pérez, N., & Adamowski, J. C. (2017). Review of Progress in Acoustic Levitation. Brazilian Journal of Physics, 48(2), 190-213. doi:10.1007/s13538-017-0552-6

[6] Li, J., Cao, W., Liu, P., & Ding, H. (2010). Influence of gas inertia and edge effect on squeeze film in near field acoustic levitation. Applied Physics Letters, 96(24), 243507. doi: 10.1063/1.3455896

[7] Nomura, H., Kamakura, T., & Matsuda, K. (2002). Theoretical and experimental examination of near-field acoustic levitation. The Journal of the Acoustical Society of America, 111(4), 1578-1583. doi:10.1121/1.1453452

[8] Minikes, A., & Bucher, I. (2006). Comparing numerical and analytical solutions for squeeze-film levitation force. Journal Of Fluids And Structures, 22(5), 713-719. doi: 10.1016/j.jfluidstructs.2006.02.004

[9] Langlois, W. E. (1962). Isothermal squeeze films. Quarterly of Applied Mathematics, 20(2), 131-150. doi:10.1090/qam/99963

[10] Ilssar, D., Bucher, I., & Flashner, H. (2017). Modeling and closed loop control of near-field acoustically levitated objects. Mechanical Systems and Signal Processing, 85, 367-381. doi:10.1016/j.ymssp.2016.08.011

[11] Hrka, S., & Svenšek, D., Dr. (2015). Acoustic Levitation. University of Ljubljana, Faculty of Mathematics and Physics.

[12] Ueha, S., Hashimoto, Y., & Koike, Y. (2000). Non-contact transportation using near-field acoustic levitation. Ultrasonics, 38(1-8), 26-32. doi:10.1016/s0041-624x(99)00052-9

[13] Chu, B. T., & Apfel, R. E. (1982). Acoustic radiation pressure produced by a beam of sound. The Journal of the Acoustical Society of America, 72(6), 1673–1687. https://doi.org/10.1121/1.388660

[14] Hong, Z., Zhai, W., Yan, N., & Wei, B. (2014). Measurement and simulation of acoustic radiation force on a planar reflector. The Journal Of The Acoustical Society Of America, 135(5), 2553-2558. doi: 10.1121/1.4869678

[15] Shi, M., Feng, K., Hu, J., Zhu, J., & Cui, H. (2019). Near-field acoustic levitation and applications to bearings: A critical review. International Journal of Extreme Manufacturing, 1(3), 032002. doi:10.1088/2631-7990/ab3e54

[16] Reynolds O IV 1886 On the theory of lubrication and its application to Mr Beauchamp tower’s experiments, including an experimental determination of the viscosity of olive oil Phil. Trans. R. Soc. 117 157–234

[17] Andrade, M. A., Okina, F. T., Bernassau, A. L., & Adamowski, J. C. (2017). Acoustic levitation of an object larger than the acoustic wavelength. The Journal of the Acoustical Society of America, 141(6), 4148-4154. doi:10.1121/1.4984286

[18] Xie, W., & Wei, B. (2002). Dependence of acoustic levitation capabilities on geometric parameters. Physical Review E, 66(2). doi: 10.1103/physreve.66.026605

[19] Andrade, M. A., Ramos, T. S., Adamowski, J. C., & Marzo, A. (2020). Contactless pick-and-place of millimetric objects using inverted near-field acoustic levitation. Applied Physics Letters, 116(5), 054104. doi:10.1063/1.5138598

[20] Takasaki, M., Terada, D., Kato, Y., Ishino, Y., & Mizuno, T. (2010). Non-contact ultrasonic support of minute objects. Physics Procedia, 3(1), 1059-1065. doi: 10.1016/j.phpro.2010.01.137

[21] Ueha, S. (2002). Phenomena, theory and applications of nearfield acoustic levitation, Rev. Acústica 33(4)