arxiv:2004.12980v1 [physics.app-ph] 27 apr 2020 · p= 0; (1b) r v= 0: (1c) the time-dependent term...

14
Nonlinear interaction of acoustic waves with a spheroidal particle: radiation force and torque effects Everton B. Lima, 1 Jos´ e P. Le˜ ao-Neto, 2 Alisson S. Marques, 1 Giclˆ enio C. Silva, 1 Jos´ e H. Lopes, 3 and Glauber T. Silva 1, * 1 Physical Acoustics Group, Instituto de F´ ısica, Universidade Federal de Alagoas, Macei´o, AL 57072-970, Brazil 2 Campus Arapiraca/Unidade de Ensino Penedo, Universidade Federal de Alagoas, Penedo, Alagoas 57200-000, Brazil 3 Grupo de F´ ısica da Mat´ eria Condensada, N´ ucleo de Ciˆ encias Exatas, Universidade Federal de Alagoas, Arapiraca, AL 57309-005, Brazil (Dated: April 28, 2020) The nonlinear interaction of a time-harmonic acoustic wave with an anisotropic particle gives rise to the radiation force and torque effects. These phenomena are at the heart of the acoustofluidics technology, where microparticles such as cells and microorganisms are acoustically manipulated. We present a theoretical model considering a generic acoustic beam interacting with a subwavelength spheroidal particle in a nonviscous fluid. Concise analytical expressions of the radiation force and torque are obtained in the scattering dipole approximation. The radiation force is given in terms of a gradient and scattering force; while the radiation torque has two fundamental contributions, namely, the momentum arm and acoustic spin (spin-torque effect). As a practical example, we use the theory to describe the interaction of two crossed plane waves and a prolate spheroidal particle. The results reveal the particle is transversely trapped in a pressure node and is axially pushed by the radiation force. Also, the momentum arm aligns the particle in the axial direction. At certain specific positions, only the spin-torque occurs. Our findings are remarkably consistent with finite-element simulations. The success of our model enables its use as an investigation tool for the manipulation of anisotropic microparticles in acoustofluidics. I. INTRODUCTION The behavior of microparticles under an ultrasonic acoustic wave has been extensively analyzed in micro- acoustofluidic devices [1, 2]. Notable examples include separation of circulating tumor cells [3], cell and mi- croparticle patterning [4, 5], assess the membrane elas- ticity of cell by acoustic deformation [6, 7], and selec- tive acoustic tweezer [8]. The nonlinear wave-particle interaction gives rise to the acoustic radiation forces and torques phenomena [9]. The careful control of these ef- fects enables particle handling in micrometer-sized cavi- ties or microchannels with a plethora of applications in biotechnology and analytical chemistry. The acoustic radiation forces and torques are com- monly investigated considering isotropic particles, i.e., with a spherical shape. In reality, the morphology of most cells and other microorganisms have a degree of asym- metry. Prominent examples of acoustofluidic systems for manipulation of asymmetric particles include glass fibers [10], Escherichia coli bacterium [11], red blood cells [12], microfibers [13], and alumina microdisks [14]. Other experiments have been performed in acoustic levi- tation systems in air [15–17]. Understanding how acous- tic forces and torques develop on anisotropic micropar- ticles is key to dynamic analysis, as well as to devise new applications of acoustofluidic methods. Addition- ally, these phenomena seem to have a crucial role in the propelling mechanisms of microswimmers under an ul- trasound field [18, 19]. * gtomaz@fis.ufal.br At first glance, the available alternative to model the wave interaction with anisotropic particles is the use of numerical techniques, such as the finite [10, 14, 20–22] and boundary [23] element methods, Born approxima- tion [24], numerical quadrature [25, 26], and T -matrix approach [27]. In general, numerical methods demand high-performance computing and high memory usage for three-dimensional simulations. Moreover, it is also chal- lenging to determine the behavior of the wave-particle system as one or more parameters vary continuously. Early studies involving anisotropic particles dealt with the radiation torque problem on circular disks [28–31]. Some other investigations have been surveyed in Ref. [32]. More recently, efforts have been devoted to describing the acoustic radiation force [33, 34] and torque [35–37] on spheroidal particles. These analyses rely on the partial- wave expansion of the acoustic fields. In this method, the expansion coefficients of the incoming beam (beam-shape coefficients) should be known a priori [38–43]. Numeri- cal schemes can also be employed to compute the beam- shape coefficients [44–48], and even experimental meth- ods [49]. Regardless these studies, the ultrasonic waves produced in acoustofluidic devices have a complex spatial form (structured waves), and the corresponding beam- shape coefficients are generally unknown. In the case of isotropic particles, the radiation force caused by struc- tured waves can be easily computed using the Gorkov’s theory [50], which requires only the incident pressure and fluid velocity, either analytically or numerically. A sim- ilar approach was developed for the radiation torque on a spherical particle [51]. The purpose of this article is to model the radiation force and torque imparted on a subwavelength spheroid arXiv:2004.12980v1 [physics.app-ph] 27 Apr 2020

Upload: others

Post on 18-Jul-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: arXiv:2004.12980v1 [physics.app-ph] 27 Apr 2020 · p= 0; (1b) r v= 0: (1c) The time-dependent term e i!t is omitted for simplic-ity. Equation (1c) is obtained by taking the rotational

Nonlinear interaction of acoustic waves with a spheroidal particle: radiation force andtorque effects

Everton B. Lima,1 Jose P. Leao-Neto,2 Alisson S. Marques,1

Giclenio C. Silva,1 Jose H. Lopes,3 and Glauber T. Silva1, ∗

1Physical Acoustics Group, Instituto de Fısica, Universidade Federal de Alagoas, Maceio, AL 57072-970, Brazil2Campus Arapiraca/Unidade de Ensino Penedo, Universidade Federal de Alagoas, Penedo, Alagoas 57200-000, Brazil

3Grupo de Fısica da Materia Condensada, Nucleo de Ciencias Exatas,Universidade Federal de Alagoas, Arapiraca, AL 57309-005, Brazil

(Dated: April 28, 2020)

The nonlinear interaction of a time-harmonic acoustic wave with an anisotropic particle gives riseto the radiation force and torque effects. These phenomena are at the heart of the acoustofluidicstechnology, where microparticles such as cells and microorganisms are acoustically manipulated. Wepresent a theoretical model considering a generic acoustic beam interacting with a subwavelengthspheroidal particle in a nonviscous fluid. Concise analytical expressions of the radiation force andtorque are obtained in the scattering dipole approximation. The radiation force is given in termsof a gradient and scattering force; while the radiation torque has two fundamental contributions,namely, the momentum arm and acoustic spin (spin-torque effect). As a practical example, we usethe theory to describe the interaction of two crossed plane waves and a prolate spheroidal particle.The results reveal the particle is transversely trapped in a pressure node and is axially pushedby the radiation force. Also, the momentum arm aligns the particle in the axial direction. Atcertain specific positions, only the spin-torque occurs. Our findings are remarkably consistent withfinite-element simulations. The success of our model enables its use as an investigation tool for themanipulation of anisotropic microparticles in acoustofluidics.

I. INTRODUCTION

The behavior of microparticles under an ultrasonicacoustic wave has been extensively analyzed in micro-acoustofluidic devices [1, 2]. Notable examples includeseparation of circulating tumor cells [3], cell and mi-croparticle patterning [4, 5], assess the membrane elas-ticity of cell by acoustic deformation [6, 7], and selec-tive acoustic tweezer [8]. The nonlinear wave-particleinteraction gives rise to the acoustic radiation forces andtorques phenomena [9]. The careful control of these ef-fects enables particle handling in micrometer-sized cavi-ties or microchannels with a plethora of applications inbiotechnology and analytical chemistry.

The acoustic radiation forces and torques are com-monly investigated considering isotropic particles, i.e.,with a spherical shape. In reality, the morphology of mostcells and other microorganisms have a degree of asym-metry. Prominent examples of acoustofluidic systemsfor manipulation of asymmetric particles include glassfibers [10], Escherichia coli bacterium [11], red bloodcells [12], microfibers [13], and alumina microdisks [14].Other experiments have been performed in acoustic levi-tation systems in air [15–17]. Understanding how acous-tic forces and torques develop on anisotropic micropar-ticles is key to dynamic analysis, as well as to devisenew applications of acoustofluidic methods. Addition-ally, these phenomena seem to have a crucial role in thepropelling mechanisms of microswimmers under an ul-trasound field [18, 19].

[email protected]

At first glance, the available alternative to model thewave interaction with anisotropic particles is the use ofnumerical techniques, such as the finite [10, 14, 20–22]and boundary [23] element methods, Born approxima-tion [24], numerical quadrature [25, 26], and T -matrixapproach [27]. In general, numerical methods demandhigh-performance computing and high memory usage forthree-dimensional simulations. Moreover, it is also chal-lenging to determine the behavior of the wave-particlesystem as one or more parameters vary continuously.

Early studies involving anisotropic particles dealt withthe radiation torque problem on circular disks [28–31].Some other investigations have been surveyed in Ref. [32].More recently, efforts have been devoted to describing theacoustic radiation force [33, 34] and torque [35–37] onspheroidal particles. These analyses rely on the partial-wave expansion of the acoustic fields. In this method, theexpansion coefficients of the incoming beam (beam-shapecoefficients) should be known a priori [38–43]. Numeri-cal schemes can also be employed to compute the beam-shape coefficients [44–48], and even experimental meth-ods [49]. Regardless these studies, the ultrasonic wavesproduced in acoustofluidic devices have a complex spatialform (structured waves), and the corresponding beam-shape coefficients are generally unknown. In the case ofisotropic particles, the radiation force caused by struc-tured waves can be easily computed using the Gorkov’stheory [50], which requires only the incident pressure andfluid velocity, either analytically or numerically. A sim-ilar approach was developed for the radiation torque ona spherical particle [51].

The purpose of this article is to model the radiationforce and torque imparted on a subwavelength spheroid

arX

iv:2

004.

1298

0v1

[ph

ysic

s.ap

p-ph

] 2

7 A

pr 2

020

Page 2: arXiv:2004.12980v1 [physics.app-ph] 27 Apr 2020 · p= 0; (1b) r v= 0: (1c) The time-dependent term e i!t is omitted for simplic-ity. Equation (1c) is obtained by taking the rotational

2

by an arbitrary acoustic beam. Our approach is basedon the general expressions of the radiation force [34] andtorque [36, 37] that depend on the expansion coefficientsof the incident and scattering waves., e.g., the beam-shape and scattering coefficients. These expressions areexact to the dipole moment for the scattering wave. Thescattering coefficients are obtained through the bound-ary conditions on the particle surface [34]. Also, the re-lationship between the beam-shape coefficients and theincident acoustic fields (pressure and fluid velocity) isestablished. Strikingly, the final expressions of the radia-tion force and torque are given in terms of the incomingfields evaluated at the geometric center of the particle.Despite being developed for a rigid spheroid, the theorycan be readily adapted to accommodate the elasticityand absorption of the particle, as well as the surroundingfluid viscosity. These effects have already been investi-gated for isotropic particles [52, 53].

Our model is used for the analysis of a spheroidal par-ticle interacting with two crossed plane waves at rightangle. This structured beam forms a transverse standingwave and an axial (perpendicular) traveling wave. Themodel predicts the radiation force pushes the particle toa transverse pressure node or antinode. The radiationtorque aligns the particle at a node in broadside orienta-tion. In contrast, the orientation in an antinode is alongthe axis of the standing wave. Additionally, the axialradiation force pushes the particle in the direction par-allel to the pressure node. Lastly, the model predictionsare verified against finite-element (FE) simulations. Wefind an excellent agreement between the theoretical andnumerical results.

II. THEORETICAL MODEL

A. Governing equations

Consider a fluid of infinite extension characterized byan ambient density ρ0, adiabatic speed of sound c0, andcompressibility β0 = 1/ρ0c

20. Our analysis is restricted

to acoustic fields of time harmonic dependence e−iωt,with ω = 2πf and f being the angular and linear fre-quencies, respectively. The corresponding wavenumberis k = ω/c0 = 2π/λ, where λ is the acoustic wavelength.The acoustic pressure and fluid velocity are expressed us-ing the complex-phase representation, p(r, t) = p(r)e−iωt

and v(r, t) = v(r)e−iωt, respectively. In Cartesian co-ordinates, the position vector and fluid velocity are riei(with i = x, y, z) and v = viei, where the ei is the Carte-sian unit vector. The summation over repeated indexesis automatically assumed hereafter. We also adopt thenotation (rx, ry, rz) = (x, y, z).

In the inviscid limit, the wave dynamics is modeled by

FIG. 1. The acoustic wave scattering by the spheroidalparticle in a fluid medium. The particle has a major andminor semiaxis denoted by a and b. The arbitrary incomingbeam is depicted as a red sinusoidal line. The scattered wavesare represented by red arrows. The geometric center of thespheroid defines the particle frame of reference, r = (x, y, z).The laboratory frame of reference is the primed Cartesianaxes.

the well-known linear acoustic equations

v = − i

ρ0c0k∇p, (1a)(

∇2 + k2)p = 0, (1b)

∇× v = 0. (1c)

The time-dependent term e−iωt is omitted for simplic-ity. Equation (1c) is obtained by taking the rotational ofEq. (1a).

In the presence of an inclusion, such as a particle, anincident wave is scattered by the inclusion. The bound-ary condition for the scattered pressure psc at the farfieldis the Sommerfeld radiation condition. In spherical coor-dinates (r, θ, ϕ), this means

limr→∞

r (∂r − ik) psc = 0. (2)

The radiation condition singles out only the solutionwhich represents “outgoing” waves.

Another boundary condition comes from consideringthe particle as a rigid body. In this case, the normalcomponent of the total velocity of the fluid, i.e. inci-dent plus scattering contributions, should vanish on theparticle surface S0,

n · v|r∈S0 = 0. (3)

Here n is the outward normal unit-vector on S0.

B. Prolate spheroidal particle

A prolate spheroidal particle with a major and minoraxis denoted by 2a and 2b, is placed in the wave pathof an arbitrary incoming beam as depicted in Fig. 1. Its

Page 3: arXiv:2004.12980v1 [physics.app-ph] 27 Apr 2020 · p= 0; (1b) r v= 0: (1c) The time-dependent term e i!t is omitted for simplic-ity. Equation (1c) is obtained by taking the rotational

3

geometric center defines the particle coordinate system(x, y, z). The particle surface is conveniently describedin prolate spheroidal coordinates whose connection theCartesian system is expressed by

x =d

2

√(ξ2 − 1)(1− η2) cosϕ, (4a)

y =d

2

√(ξ2 − 1)(1− η2) sinϕ, (4b)

z =dξη

2, (4c)

where ξ ≥ 1 is the spheroidal radial coordinate, −1 ≤ η ≤1, and 0 ≤ ϕ ≤ 2π is azimuth angle, with the interfocaldistance being

d = 2√a2 − b2. (5)

The particle surface is given by

ξ = ξ0 =

[1−

(b

a

)2]−1/2

, (6)

where ξ0 is regarded as the geometric parameter of theparticle. The particle orientation follows the z-direction,d = d ez. Note that we recover a sphere of radius r0 bysetting

d→ 0, ξ0 →∞,ξ0d

2→ r0. (7)

The particle dynamics is better analyzed in a fixedlaboratory system (x′, y′, z′). In this reference frame,the particle orientation angle α is expressed by d · ez′ =d cosα.

C. Acoustic scattering

The incoming and scattered pressure can be expressedby the partial wave expansion in spherical coordinates(r, θ, ϕ),

pin = p0

∞∑n=0

n∑m=−n

anmjn(kr)Y mn (θ, ϕ), (8a)

psc = p0

∞∑n=0

n∑m=−n

anmsnmhn(kr)Y mn (θ, ϕ), (8b)

in which p0 is the peak pressure, jn is the nth-order spher-ical Bessel function, hn is the spherical Hankel function ofthe first-type, and Y mn is the spherical harmonics of nth-order and mth-degree. The quantity anm is known as thebeam-shape coefficient. The scattering coefficients snmcan be obtained from the boundary condition in Eq. (3).

As we advance to consider a particle much smaller thanthe wavelength, i.e., the so-called long-wavelength limit,only the monopole (n = 0) and dipole (n = 1) modesof the multipole expansion in (8b) suffice to describe the

acoustic scattering [54]. We define the smallness param-eter of the particle as

ε =ξ0kd

2= ka� 1. (9)

We shall use ε as the expansion parameter in the long-wavelength approximation.

Using the partial wave expansion in spheroidal coor-dinates (ξ, η, ϕ), one can show that the monopole anddipole scattering coefficients of a rigid spheroid are givenby [34]

s00 = − iε3

3f00 −

ε6

9f200, (10a)

s10 =iε3

6f10 −

ε6

36f210, (10b)

s1,−1 = s11 =iε3

12f11 −

ε6

144f211, (10c)

with i being the imaginary unit. The dipole coefficientsof the modes perpendicular to the axial direction (s1,−1and s11) are degenerated due to the axial symmetry ofthe particle. The scattering factors f00, f10, and f11 ofRef. [34] are re-written here as

f00 = 1− ξ−20 , (11a)

f10 =2

3ξ30

[ξ0

ξ20 − 1− ln

(ξ0 + 1√ξ20 − 1

)]−1, (11b)

f11 =8

3ξ30

[2− ξ20

ξ0(ξ20 − 1)+ ln

(ξ0 + 1√ξ20 − 1

)]−1. (11c)

These functions depend solely on the geometric factor ξ0.To recover the scattering coefficients of a spherical par-

ticle, we take the limit ξ0 →∞ in (11), which yields

f00 = 1, f10 = 1, f11 = 2. (12)

It immediately follows from (10) that the dipole scat-tering coefficients of a sphere are all degenerated, as ex-pected,

s10 = s1,−1 = s11, (sphere). (13)

D. Multipole expansion of the incident beam

The beam-shape coefficients of an incident beam isobtained as follows. Multiplying Eq. (8a) by Y m∗n ,integrating the result over the unit-sphere, and us-ing the orthogonal relation of the spherical harmonics∫ π0

∫ 2π

0Y mn (θ, ϕ)Y m

′∗n′ (θ, ϕ) sin θ dθ dϕ = δnn′δmm′ (with

δnm being the Kronecker delta function), we find [44]

anm =

1

p0 jn(kr)

∫ π

0

∫ 2π

0

pin(kr, θ, ϕ)Y m∗n (θ, ϕ) sin θ dθ dϕ,

(14)

Page 4: arXiv:2004.12980v1 [physics.app-ph] 27 Apr 2020 · p= 0; (1b) r v= 0: (1c) The time-dependent term e i!t is omitted for simplic-ity. Equation (1c) is obtained by taking the rotational

4

where asterisk denotes complex conjugation. The beam-shape coefficient can be calculated for any radial dis-tance kr. In particular, one can relate them with theacoustic pressure evaluated in the origin of the particleframe, r = 0. We anticipate here that the acoustic ra-diation force depends on the beam-shape coefficients tothe quadrupole approximation [51]. Therefore, we trun-cate the Taylor expansion of the incident pressure aroundr = 0 as

pin(r) = pin(0) + ri∂ipin|r=0 +1

2!rirj∂i∂jpin|r=0, (15)

with i, j = x, y, z. Before proceeding, we need the asymp-totic expansion of the spherical Bessel function for r → 0,

jn(kr) ≈√π

2n+1Γ(n+ 3/2)(kr)n, (16)

in which Γ(n) is the gamma function. Replacing Eqs. (15)and (16) into Eq. (14) and evaluating the integrals withrx = r sin θ cosϕ, ry = r sin θ sinϕ, and rz = r cos θ, wefind beam-shape coefficient up to the quadrupole order,

a00 =

√4π

p0pin, (17a)

a10 = 2i√

3πρ0c0p0

vin,z, (17b)

a1,±1 = i√

6πρ0c0p0

(∓vin,x + ivin,y), (17c)

a20 = −i√

5πρ0c0kp0

(∂xvin,x + ∂yvin,y − 2∂zvin,z) ,

(17d)

a2,±1 = i√

30πρ0c0kp0

(∓∂zvin,x + i∂zvin,y) , (17e)

a2,±2 = i

√15π

2

ρ0c0kp0

(∂xvin,x − ∂yvin,y ∓ 2i∂xvin,y) .

(17f)

In this derivation, we assumed the fluid flow isirrotational–see Eq. (1c). We remark the acoustic fieldsof (17) are evaluated at r = 0.

E. Acoustic radiation force

As the linear momentum flux carried by a wave is ex-changed with the particle, the radiation force (a time-average quantity over the wave period 2π/ω) appears onthe particle [47],

F rad = −∫S1

Re

[(β0|p|2

4− ρ0|v|2

4

)I +

ρ02vv∗

]· er dS,

(18)where S1 is a spherical surface that encloses the parti-cle, I is the unit tensor given in Eq. (C2), and dS isthe area element. The combination of two vectors asvv forms a dyad, i.e., a second-rank tensor–see detailsin Appendix C. The total pressure and fluid velocity arep = pin + psc and v = vin + vsc, respectively.

In the long-wavelength limit ε � 1, the Cartesiancomponents of the acoustic radiation force exerted ona spheroidal particle are expressed by [34]

Fx + iFy =iE0

2k2

[√2

3[a∗00a1,−1 (s∗00 + s11 + 2s∗00s11) + a00a

∗11 (s00 + s∗11 + 2s∗11s00)]

+

√2

5

(a10a

∗21s10 + a2,−1a

∗10s∗10 +

√2[a11a

∗22s11 + a2,−2a

∗1,−1s

∗11])

+

√2

15(a∗20s11a1,−1 + a20a

∗11s∗11)

],

(19a)

Fz =E0

k2Im

[2√15a10a

∗20s10 +

1√5

(a1,−1a∗2,−1 + a11a

∗21)s11 +

1√3a00a

∗10(s00 + s∗10 + 2s00s

∗10)

], (19b)

where the asterisk denotes complex conjugation, andE0 = β0p

20/2 is the characteristic energy density of the

incident wave.

To obtain the radiation force as a function of the in-coming acoustic fields, we replace the beam-shape coef-

ficients from (19) by (17). We also use the Helmholtzequation for the incoming pressure ∂i∂ipin = −k2pin.Details on the derivation can be seen in Appendix A.

Page 5: arXiv:2004.12980v1 [physics.app-ph] 27 Apr 2020 · p= 0; (1b) r v= 0: (1c) The time-dependent term e i!t is omitted for simplic-ity. Equation (1c) is obtained by taking the rotational

5

Accordingly, the radiation force is given by

F rad = −πa3 Re

[β0(Q∗grad − iε3Q∗sca

)· ∇pin

+ ρ0(D∗grad − iε3D∗sca

)· ∇vin

]r=0

,

= F sca + F grad. (20)

The scattering force results from the self-interaction ofthe scattering wave, whereas the gradient force comesfrom the incoming and scattering wave interaction. Fur-thermore, the scattering force is much weaker than thegradient one by a factor of ε3. The monopole tensors(Qgrad and Qsca) and dipole vectors (Dgrad and Dsca)are expressed by

Qgrad =2

3f00pin(0) eiei, (21a)

Qsca = −f009pin(0)

[(2f00 + f11) (exex + eyey)

+ 2(f00 + f10) ezez

], (21b)

Dgrad = −f112

[vin,x(0)ex + vin,y(0)ey]− f10vin,z(0)ez,

(21c)

Dsca = −1

6[f2114

[vin,x(0)ex + vin,y(0)ey] + f210vin,z(0)ez

].

(21d)

Note that eiej is a dyad, which forms the Cartesian basisof a second-rank tensor. It is straightforward to showfrom (20) that the gradient force is expressed by

F grad = −∇U(0), (22a)

U = πa3[β0f00

3|pin|2 −

ρ02

(f112

(|vin,x|2 + |vin,y|2)

+ f10|vin,z|2)]. (22b)

Considering an standing wave, the corresponding am-plitude pin is a real-valued function. Consequently, ∇pinand Q∗sca are also real-valued quantities. We thus con-clude that Re[iQ∗sca · ∇pin] = 0 and Re[iD∗sca · ∇vin] = 0,so only the gradient force remains.

On the contrary, for traveling waves, which possesscomplex amplitude, say, a plane wave eik·r, the gradi-ent operator becomes ∇ → ik, then F grad = 0.

F. Acoustic radiation torque

The acoustic radiation torque exerted on the particleis given by [31]

τ rad = −1

2Re

∫S1

r × ρ0vv∗ · er dS. (23)

The Cartesian components of the radiation torque on aprolate spheroidal particle are given by [36, 37]

τx = − E0

k3√

2Re

[(a1,−1 + a11)(1 + s11)a∗10s

∗10

+ (a∗1,−1 + a∗11)(1 + s10)a10s∗11

], (24a)

τy =E0

k3√

2Re

[i (a1,−1 − a11)(1 + s11)a∗10s

∗10

− i (a∗1,−1 − a∗11)(1 + s10)a10s∗11

], (24b)

τz =E0

k3Re[(|a1,−1|2 − |a11|2)(1 + s11)s∗11

]. (24c)

We shall obtain the radiation torque in terms ofthe momentum flux tensor and the acoustic spin den-sity [55, 56] of the incident wave. Here, they are ex-pressed, respectively, by

P =ρ02

Re[vinv∗in], (25a)

S =ρ02ω

Im[v∗in × vin]. (25b)

The acoustic spin is an intrisic property of a wave. Itgives a measure of the fluid velocity rotation, which isindependent of the reference frame.

To obtain the acoustic radiation torque, we substitutethe scattering coefficients of (10) into (24). The wholederivation is developed in Appendix B. Accordingly, theradiation torque is

τ rad = −πa3[χ [ez ×P(0) · ez]−

ε3

24χ2ωS⊥(0)

], (26)

where χ = f11 − 2f10 is the dipole-difference factor andS⊥ = Sxex + Syey is the transverse spin density. It isworth noticing that (P · ez)i = (ρ0/2) Re[vin,iv

∗in,z].

The first term of Eq. (26) is related to the moment armcaused the linear momentum flux P applied along the ax-ial direction (z axis). This contribution does not dependon the frequency. The second term is the spin-inducedtorque, which is much weaker than the momentum armterm. It also varies with the frequency to the third powerω3.

According to (12), as the particle approaches anisotropic geometry (ξ0 → ∞), the dipole-difference fac-tor becomes χ = 0. Thus, the radiation torque vanishes,as predicted before [57, 58]. In Fig. 2, we show that thedipole-difference factor χ for a rigid spheroid. It is alwayspositive and has a maximum at ξ0 = 1.3181. Besides, χgoes to zero as the particle approaches a thin line, e.g.,ξ0 → 1. Thus, the radiation torque vanishes in this limit.

Page 6: arXiv:2004.12980v1 [physics.app-ph] 27 Apr 2020 · p= 0; (1b) r v= 0: (1c) The time-dependent term e i!t is omitted for simplic-ity. Equation (1c) is obtained by taking the rotational

6

FIG. 2. The dipole-difference factor χ is plotted as a functionof the geometric parameter of the particle. The maximumvalue occurs at ξ0 = 1.3181.

III. RESULTS AND DISCUSSION

A. Two crossed plane waves

Now we applied the developed theory to study the in-teraction of a subwavelength spheroid with two crossedplane waves at right angle. This acoustic beam was cho-sen because it has acoustic spin [55]. Moreover, as weshall see later, this beam forms simultaneously a travel-ing and standing wave. In Fig. 3, we sketch wave-particleinteraction. The wave vectors in the laboratory frame are

k′1 =k√

2

2(ex′ + ez′) , (27a)

k′2 =k√

2

2(−ex′ + ez′) . (27b)

Hence, the corresponding pressure field is

pin =p02

[eik

′1·(r

′+h′) + eik′2·(r

′+h′)],

= p0 cos

[k√2

(x′ + h)

]eikz

′. (28)

We see that the wave interference sets in a standing wavealong the transverse direction (x′ axis). The offset vec-tor h′ = hex′ gives the position of the nearest pressureantinode regarding the particle center in the transversedirection.

To calculate the wave vectors in the particle system,we introduce the (clockwise) rotation matrix around they′ axis,

Ry′(α) =

cosα 0 − sinα0 1 0

sinα 0 cosα

. (29)

From (27), we obtain the wave vectors in the particle

FIG. 3. Two plane waves crossing at right angle and in-teracting with a spheroidal particle in the x′z′ plane. Theparticle frame corresponds to the x and z axes. The angle αgives the particle orientation with respect to the z′-axis. They′- and y-axis are going into the screen. The wavevectors k1

and k2 are depicted by red and green arrows. The vector h(purple arrow) points to a transverse pressure antinode.

system as

k1 = Ry′(α)k′1

=k√2

[(cosα− sinα) ex + (cosα+ sinα) ez] , (30a)

k2 = Ry′(α)k′2

= − k√2

[(cosα+ sinα) ex − (cosα− sinα) ez] .

(30b)

Also, the offset parameter becomes h = h(cosα ex +sinα ez). We thus obtain the incident pressure in theparticle frame,

pin =p02

[eik1·(r+h) + eik2·(r+h)

],

=p02

[eik[(x+z) cosα−(x−z) sinα+h]/

√2

+ e−ik[(x−z) cosα+(x+z) sinα+h]/√2

]. (31)

The related fluid velocity is readily calculated by insert-ing this equation into Eq. (1a),

vin =p0

2ρ0c0k

[k1eik1·(r+h) + k2eik2·(r+h)

]. (32)

The linear momentum flux of the incoming beam isderived by replacing Eq. (32) into (25a),

P(0) =E0

4Re

[k1k1 + k2k2 + k1k2ei(k1−k2)·h

+ k2k1e−i(k1−k2)·h]. (33)

Page 7: arXiv:2004.12980v1 [physics.app-ph] 27 Apr 2020 · p= 0; (1b) r v= 0: (1c) The time-dependent term e i!t is omitted for simplic-ity. Equation (1c) is obtained by taking the rotational

7

But k1 − k2 =√

2k (cosα ex + sinα ez), then

P(0) =E0

4Re

[k1k1+k2k2+k1k2ei

√2kh+k2k1e−i

√2kh

].

(34)The acoustic spin density is obtained by substitutingEq. (32) into (25b),

S(0) = − E0

4ωk2Im[(k2 × k1)

(e−i(k1−k2)·h − ei(k1−k2)·h

)]=E0

2ωsin(√

2kh)ey = S⊥(0). (35)

Here we have utilized the relation k2 × k1 = k2 ey.

B. Radiation torque

The radiation torque in the laboratory frame is derivedby substituting Eqs. (33) and (35) into Eq. (26), andnoting that ey = ey′ . Thus, we obtain

τ rad =

πa3E0

4

[χ cos

(√2kh

)sin 2α+

ε3

12χ2 sin

(√2kh

)]ey′ .

(36)

When the particle is trapped in a pressure node (h =

π/k√

2) or antinode (h = 0), the radiation torque be-comes

τ radnode = −τ rad

anti = −πa3

4E0χ sin 2α ey′ . (37)

These torques are caused exclusively by the momentumarm effect. At the end, the particle will be aligned tobroadside orientation (α = 0) in a pressure node, andparallel to the standing wave axis (α = π/2) in a pressureantinode.

Another interesting field position is h = h± =±√

2π/4k. In this case, only the spin-induced torquearises,

τ radspin(h±) = ± ε

3

48πa3E0χ

2ey′ . (38)

This torque sets the particle to rotate around the minoraxis. The particle spins in clockwise direction at h = h+(spin up), and in counterclockwise manner at h = h−(spin down).

The incoming beam alongside the particle orientationsin a node and antinode are illustrated in Fig. 4. The localfluid velocity polarizabilities are also shown for h = h±.In this case, the velocities rotate circularly.

C. Radiation force

The transformation that furnishes the radiation forcein the laboratory frame is

F rad′ = Ry′(−α)F rad. (39)

FIG. 4. The amplitude (real part) of the incoming wavein Eq. (28). The equilibrium orientation of the spheroidalparticle (gray ellipses) are illustrated in a node (α = 0) andantinode (α = π/2). The fluid velocity polarizabilities areshown when h = h+ =

√2π/4k (spin up) and h = h− =

−√

2π/4k (spin down). Color legend: p0 (bright yellow) and−p0 (dark red).

Here the matrix Ry′(−α) defined in Eq. (29) representsa counterclockwise rotation by an angle α around they′-axis.

In the particle frame, the gradient force is derived bysubstituting Eqs. (31) and (32) into (22). After somestraightforward calculations, we arrive at

F grad =ε

12√

2πa2E0 [8f00 + 3(f11 − 2f10) cos 2α]

sin(√

2kh)

(cosα ex + sinα ez) . (40)

Inserting Eq. (40) into (39), we obtain the transverseradiation force in the laboratory frame,

F grad′ = επa2E0Φac sin(√

2kh) ex′ , (41a)

Φac =1

12√

2[8f00 + 3(f11 − 2f10) cos 2α] . (41b)

The sign of the acoustophoretic factor Φac determineswhether the particle will be trapped in a pressure node(Φac > 0) or antinode (Φac < 0). Moreover, the radia-tion force varies linearly with frequency, ε ∼ ω. We willanalyze to the axial radiation force later.

We plot the acoustophoretic factor Φac versus ξ0 inFig. 5. No significant difference is noted as the orien-tation changes from 0 to π/2. The factor is positive;thus, the particle will be transversely trapped in a pres-sure node. As the particle approaches a spherical shapeξ0 → ∞, then Φac →

√2/3. In the other extreme, as

ξ0 → 1, we have Φac → 0. Hence, no transverse radia-tion force is produced on a thin line particle.

Now we take a closer look at the axial radiation forceas the particle is trapped in a pressure node. This forceis in fact the scattering force defined in (20). Again, we

insert Eqs. (31) and (32) into (20) and set h = π/√

2k

Page 8: arXiv:2004.12980v1 [physics.app-ph] 27 Apr 2020 · p= 0; (1b) r v= 0: (1c) The time-dependent term e i!t is omitted for simplic-ity. Equation (1c) is obtained by taking the rotational

8

FIG. 5. The acoustophoretic factor Φac versus the geometricparameter ξ0 for different particle orientations.

(pressure node) to find

F scax = −πa

2ε4E0

24√

2

[f211 + (f211 − 4f210) cos 2α

]sinα,

(42a)

F scaz =

πa2ε4E0

24√

2

[4f210 + (f211 − 4f210) cos 2α

]cosα.

(42b)

Applying the rotation operator Ry′(−α) on (42) yieldsthe axial radiation force in the laboratory frame,

F sca′ = πa2E0Qrad ez′ , (pressure node), (43a)

Qrad =ε4

48√

2

[4f210 + f211 + (f211 − 4f210) cos 2α

]. (43b)

The quantity Qrad is referred to as the dimensionless ra-diation force efficiency. We remark the scattering forceis related to the linear momentum flux carried away bythe scattered waves. This connection can be understoodfrom the radiation force relationship with the scatter-ing cross-section as discussed in Refs. [42, 59]. In turn,the scattering cross-section of a rigid spheroid scales asπa2(ka)4 [60], which is the same dependence seen in thescattering force of Eq. (43a).

In Fig. 6, we plot the radiation force efficiency dividedby (ka)4 versus with ξ0. The efficiency is positive and

asymptotically approaches Qrad/(ka)4 ∼ 1/6√

2 ≈ 0.118as ξ0 → ∞. Also, it goes to zero as ξ0 → 1. We seethat Qrad is positive with no significant difference as theorientation changes from 0 to π/2. Thereby, the forcepushes the particle along the direction of the travelingcomponent of the incoming wave.

An important aspect of Eq. (43a) is the possibilityto recover the axial radiation force imparted to a rigidspherical particle as obtained in Ref. [61]. By replac-ing the scattering factors of (12) for a spherical par-

ticle of radius r0 into (43a), we encounter F spherez′ =

πr20(kr0)4E0/6√

2. This result agrees with Ref. [61]. Tosee this connection, we multiply Eq. (15b) of [61] byπr20E0/4, and set the parameters, in the paper’s nota-tion, to ρ = 1/2, κ = −1 (rigid sphere) and γ = π/4.

FIG. 6. The radiation force efficiency Qrad normalized toε4 versus the geometric parameter of the particle for differentorientations.

TABLE I. Parameters used in the finite-element simulationsin Comsol at room temperature and pressure.

Parameter ValueSpheroidal particleMajor semiaxis (a) 47.14 µmMinor semiaxis (b) 30.71 µmGeometric parameter (ξ0) 1.3181Size parameter (ε) 0.2Medium (water)Density (ρ0) 999.66 kg m−3

Speed of sound (c0) 1481 m s−1

Compressibility (β0) 0.456 GPa−1

Domain radius 5a = 235.7 µmRadius of S1 a+ b/4 = 54.82 µmMaximum element size inside V1 b/18 = 1.706 µmMaximum element size outside V1 λ/48 = 30.71 µmPML depth (15 layers) 2a = 94.28 µmAcoustic wavePressure peak (p0) 100 kPaAcoustic energy density (E0) 2.28 J m−3

Linear frequency (f) 1 MHzWavenumber (k) 4242.5 m−1

Wavelength (λ0) 1.481 mmComputer systemCPU Xeon E5-2690 3.00GHzOperating system LinuxMaximum memory usage ∼ 128 GBComputational time per dataset ∼ 20 min

D. Finite-element simulations

We performed FE simulations of the wave-particle in-teraction in water. A set of full 3D simulations were de-vised in Comsol Multiphysics (Comsol, Inc., USA). TheFE model used in our simulations is outlined as follows.A spherical region of radius R is defined as the simula-tion domain. The particle is placed at the center of thisregion. We set a spherical surface S1 as the integrationsurface to compute the radiation force and torque as pre-scribed in Eqs. (18) and (23). The volume between theparticle surface S0 and S1 is denoted by V1. To mimicthe Sommerfeld radiation condition given in Eq. (2) for

Page 9: arXiv:2004.12980v1 [physics.app-ph] 27 Apr 2020 · p= 0; (1b) r v= 0: (1c) The time-dependent term e i!t is omitted for simplic-ity. Equation (1c) is obtained by taking the rotational

9

FIG. 7. The sketch of the FE mesh used to compute the ra-diation force and torque on a spheroidal particle. The modelcomprises a spherical region as the simulation domain and thePML. The dashed-line (light blue) depicts the integration sur-face S1. The particle is placed at the center of the simulationdomain. A finer mesh with 1.746×106 elements is defined in-side the volume enclosed by S1. Outside this region, a coarsermesh is used with 2.86 × 105 elements. The PML has 25860elements.

the scattered waves, we use the perfect matched layer(PML). The rigid boundary condition for the particle asgiven in Eq (3) is assumed. The sketch of the FE modelis displayed in Fig. 7. The incident beam is set as thebackground pressure in the particle frame. The totalpressure and fluid velocity fields are then computed andused to calculate the radiation force and torque. Thephysical parameters were inspired on those reported foracoustofluidic devices [62].

To verify the correctness of the numerical solutions,we performed some convergence tests by varying someparameter such as the domain radius, mesh density, andPML depth. The parameter variation is set up to thepoint the solution does not change with any further mod-ification. More details on convergence analysis to radia-tion force problems can be found in [20]. Typical accu-racy achieved in our tests is less than 1 % for effects oforder O(1) and O(ε), and up to 10 % for O(ε3) and O(ε4).The higher-order effects such as the spin-torque and scat-tering force seem to be more sensitive to reflections onthe PML. The simulational parameters summarized inTable I were chosen to warrant the aforementioned nu-merical accuracy.

To compare the obtained numerical results with theory,we use the normalized root-mean-square error (NRMSE),

NRMSE (%) =100

xmax − xmin

√√√√ 1

N

N∑n=1

(xtheoryn − xnumn

)2,

(44)

FIG. 8. Numerical and theoretical calculations of the trans-verse radiation force (F grad

x′ ) on a spheroidal particle as a func-tion of the scaled distance kh. The particle (gray ellipse) hasa geometric parameter of ξ0 = 1.3181 with orientation α = 0.The simulational parameters are presented in Table I.

in which N is the number of sampling points, xmax =max{xtheoryn } and xmin = min{xtheoryn }.

Importantly, both the theoretical and numerical mod-els consider the inviscid approximation. Although, actualfluids may support shear stress within the particle bound-ary layer, which may cause viscous torques [63]. Whenthe boundary layer δ = (2µ0/ρ0ω)1/2 (µ0 is the dynamicviscosity of the fluid), is much smaller than the particlesize, say δ/b � 1, viscous torque can be discarded. Inour simulations, δ/b = 10−2, and thus, we may neglectviscous torque effects here.

In Fig. 8, we plot the transverse radiation force F gradx′

as a function of the scaled distance kh. The particle(gray ellipse) has a geometric parameter of ξ0 = 1.3181,which corresponds to the maximum amplitude of thedipole-difference factor seen in Fig. 2, and consequentlythe maximum of the radiation torque in Eq (36). Theradiation force is evaluated with the equations of (41).We see the particle will be trapped in the pressurenode kh = π/

√2 = 0.71π. Both visual inspection and

NRMSE = 0.412% indicate the theoretical result is inremarkable agreement with the FE data.

The axial radiation force (F scaz′ ) in a pressure node

versus particle orientation is depicted in Fig. 9. Thisforce was calculated with the equations of (43). TheNRMSE = 5.15% is one order of magnitude above theerror of the transverse force in Fig. 8. This is so becausethe scattering force is more sensitive to numerical errors,since it is much weaker than the transverse counterpart.The maximum and minimum force is experienced by theparticle as its orientation is α = 0, π/2, respectively.

In Fig. 10, we show how the radiation torque varieswith the particle orientation in a pressure node. Thetorque is evaluated using Eq. (37), and is caused by themomentum arm only. As previously discussed, the pref-erential orientation of the particle is α = 0, i.e., theparticle will be aligned perpendicularly to the standing

Page 10: arXiv:2004.12980v1 [physics.app-ph] 27 Apr 2020 · p= 0; (1b) r v= 0: (1c) The time-dependent term e i!t is omitted for simplic-ity. Equation (1c) is obtained by taking the rotational

10

FIG. 9. The axial radiation force (F scaz′ ) at a pressure node

versus particle orientation. The particle (gray ellipses) hasa geometric parameter of ξ0 = 1.3181. The simulational pa-rameters are surveyed in Table I.

FIG. 10. The theoretical and numerical results of the ra-diation torque versus orientation angle α. The particle (grayellipses) has a geometric parameter of ξ0 = 1.3181 and istrapped in a pressure node. The simulational parameters aresummarized in Table I.

wave axis. Again, we find an excellent agreement be-tween the theory and numerical results, provided thatNRMSE = 0.57%.

The amplitude of the spin-torque is shown in Fig. 11versus the particle position kh. The particle orientationis fixed to α = 0. Equation (36) is used to calculate theradiation torque. The theoretical and numerical dataare in good match with NRMSE = 4.51%. This error isten times larger than that of the momentum arm torquein Fig. 10. Possibly, because the spin-torque is a weakerphenomenon and then is more sensitive to undesired wavereflections from the PML.

IV. CONCLUDING REMARKS

The acoustic radiation force and torque are the by-products of the nonlinear interaction of an incoming wavewith a spheroidal particle. We develop a theory to de-

FIG. 11. The theoretical and numerical spin-torque on thespheroidal particle with a fixed orientation α = 0. The par-ticle (gray ellipse) has a geometric parameter of ξ0 = 1.3181and its position varies from an antinode to a node. The sim-ulational parameters are shown in Table I.

scribe these phenomena considering subwavelength par-ticles and an incoming wave of arbitrary character. Themain contribution of our work are the laconic expressionsof the radiation force in Eqs. (20) and (22), and radia-tion torque in Eq. (26). In these expressions, the acousticfields can be described either analytically, numerically, orexperimentally. Moreover, the particle anisotropy allowsthe rise of the radiation torque from the momentum armand acoustic spin of the incoming wave. Whereas, the ra-diation force is composed of the gradient and scatteringforces similarly to the spherical particle case [51].

Our theory is successfully employed to analyze theinteraction between a rigid spheroidal particle and twoplane waves crossing at right angle. This composed beamis one of the simplest waves to possess acoustic spin [55].When the particle is trapped in a pressure node, the in-duced radiation torque is solely due to the momentumarm caused by linear momentum flux. However, as theparticle is placed at π

√2/4k along the transverse direc-

tion (i.e., the x axis), a clockwise spin-torque is activated.

The spin flips as the particle changes place to −π√

2/4k.Importantly, the spin-torque is (ka)3 weaker than themomentum arm contribution. The most prominent ef-fect of the acoustic radiation force is to trap the particlein a transverse pressure node. The axial radiation forcein a pressure node follows the traveling wave direction.The theoretical results are verified against FE simula-tions based on a full three-dimensional model. An ex-cellent agreement is found between theory and numericalexperiments.

The next level to be considered in our model is tobring in the elastic properties of particles. These fea-tures will convey the theory closer to acoustofluidic ex-periments with cells and other microorganisms. More-over, energy absorption by the particle can be as wellconsidered through complex wave numbers. The absorp-tion is known to enhance the acoustic radiation force [52].

Page 11: arXiv:2004.12980v1 [physics.app-ph] 27 Apr 2020 · p= 0; (1b) r v= 0: (1c) The time-dependent term e i!t is omitted for simplic-ity. Equation (1c) is obtained by taking the rotational

11

With these extensions, it will be possible to predict theconditions to trap particles in either a pressure node orantinode. Besides, devising a coated layer on spheroidalparticles may render them unresponsive to the radiationforce [64]. Also, getting to know how the elastic proper-ties affect the radiation torque has great importance tocontrol the rotational degree of freedom in microparticlemanipulation.

Our predictions can be tested with experiments ofeither acoustofluidics or acoustic levitation technology.The radiation force and torque derived expressions arevalid for any structured wave that is commonly employedin these techniques. Experimental confirmation of our re-sults may unlock new methods for acoustic manipulationof particles.

The theoretical model of the nonlinear interaction be-tween acoustic waves and anisotropic particles is a timelyresearch contribution. It has a great potential withinboth acoustic manipulation of particles and the investi-gation of fundamental aspects of acoustic waves.

ACKNOWLEDGMENTS

We thank the National Council for Scientific andTechnological Development–CNPq, Brazil (Grant No.401751/2016-3 and No. 307221/2016-4) for financial sup-port.

Appendix A: Radiation force expressions

Inserting the beam-shape coefficients from (17) into(19), using ∂i∂ipin = −k2pin, and rearranging the terms,we arrive at

FxFyFz

= Re

Dxy 0 00 Dxy 00 0 Dzz

v∗in,xv∗in,yv∗in,z

r=0

, (A1a)

Dxy = −2πi

k2

[3ρ0k

(s11vin · ∇⊥ + s10vin,z∂z)

− is00 (1 + 2s∗11)pinc0

], (A1b)

Dzz = −2πi

k2

[3ρ0k

(s11vin · ∇⊥ + s10vin,z∂z)

− is00 (1 + 2s∗10)pinc0

]. (A1c)

Replacing the scattering coefficients of (A1) with (10),we obtain

F rad = −πa3 Re

[β0(Q∗grad − iε3Q∗sca

)· ∇pin

+ ρ0(D∗grad − iε3D∗sca

)· ∇vin

]r=0

,

Qgrad =2

3f00pin(0) eiei, (A2a)

Qsca = −f009pin(0)

[(2f00 + f11) (exex + eyey)

+ 2(f00 + f10) ezez

], (A2b)

Dgrad = −f112

[vin,x(0)ex + vin,y(0)ey]− f10vin,z(0)ez,

(A2c)

Dsca = −1

6[f2114

[vin,x(0)ex + vin,y(0)ey] + f210vin,z(0)ez

].

(A2d)

Appendix B: Radiation torque expressions

Substituting the beam-shape coefficients of (17)into (24), we obtain(τxτy

)=

6πρ0k3

Im

[(s∗10 + s11 + 2s11s

∗10)

(vin,yv

∗in,z

−vin,xv∗in,z

)],

(B1a)

τz =6πρ0k3

(s11 + s∗11 + 2|s11|2

)Im[vin,xv

∗in,y

].

(B1b)

After using the scattering coefficients of (10) intoEq. (B1b), we find the axial radiation torque τz = O(ε12).This contribution is then neglected.

In contrast, the transverse components of the torqueare given by(

τxτy

)=πε3

2k3(f11 − 2f10)ρ0 Re

(vin,yv

∗in,z

vin,xv∗in,z

)+

πε6

24k3(f11 − 2f10)2ρ0 Im

(−vin,yv∗in,zvin,xv

∗in,z

). (B2)

We note that the axial projection of the linear momentumflux is (P·ez)i = Pi,z = (ρ0/2) Re[vin,iv

∗in,z], which allows

us to re-write the O(ε3) term in the right-hand side ofEq. (B2),

ρ02

Re

(vin,yv

∗in,z

−vin,xv∗in,z

)= −

([ez × (P · ez)]x[ez × (P · ez)]y

). (B3)

We turn to the acoustic spin term. The O(ε6) term of

Page 12: arXiv:2004.12980v1 [physics.app-ph] 27 Apr 2020 · p= 0; (1b) r v= 0: (1c) The time-dependent term e i!t is omitted for simplic-ity. Equation (1c) is obtained by taking the rotational

12

Eq. (B2) can be written as

ρ0 Im

(−vin,yv∗in,zvin,xv

∗in,z

)=ρ02

Im

((v∗in × vin)x(v∗in × vin)y

)= ω

(SxSy

)= ωS⊥, (B4)

with S⊥ being the transverse spin density. By combiningEqs. (B3), (B4) into Eq. (B2), we arrive at

τ rad = −πa3[χ [ez ×P(0) · ez]−

ε3

24χ2ωS⊥(0)

], (B5)

where χ = f11 − 2f10.

Appendix C: Basic properties of dyads

Let a = aiei, b = biei, and c = ciei be vectors withai, bi, ci ∈ C. The dyad D is defined as the product

D = ab, (C1a)

Dij = aibj , i, j ∈ {x, y, z}, (C1b)

which is a second-rank tensor. Note that ab 6= ba. Dyadsfollow the distributive rule a(b+ c) = ab+ac. The unitdyad is

I = exex + eyey + ezez = eiei. (C2)

The pre- and post-dot product are defined, respectively,by

c · ab = (c · a)b = ciaibjej , (C3a)

ab · c = a(b · c) = biciajej . (C3b)

We also have

a · I = I · a = a. (C4)

[1] A. Ozcelik, J. Rufo, F. Guo, Y. Gu, P. Li, J. Lata, andT. J. Huang, Acoustic tweezers for the life sciences, Nat.Methods 15, 1021 (2018).

[2] M. Wu, A. Ozcelik, J. Rufo, Z. Wang, R. Fang, and T. J.Huang, Acoustofluidic separation of cells and particles,Microsyst. Nanoeng. 5, 32 (2019).

[3] P. Li, Z. Mao, Z. Peng, L. Zhou, Y. Chen, P.-H. Huang,C. I. Truica, J. J. Drabick, W. S. El-Deiry, M. Dao,S. Suresh, and T. J. Huang, Acoustic separation of circu-lating tumor cells, Proc. Natl. Acad. Sci. USA 112, 4970(2015).

[4] D. J. Collins, B. Morahan, J. Garcia-Bustos, C. D. M.Plebanski, and A. Neild, Two-dimensional single-cell pat-terning with one cell per well driven by surface acousticwaves, Nat. Commun. 6, 8686 (2015).

[5] G. T. Silva, J. H. Lopes, J. P. Leao-Neto, M. K. Nichols,and B. W. Drinkwater, Particle patterning by ultrasonicstanding waves in a rectangular cavity, Phys. Rev. Ap-plied 11, 054044 (2019).

[6] P. Mishra, M. Hill, and P. Glynne-Jones, Deformationof red blood cells using acoustic radiation forces, Biomi-crofluidics 8, 034109 (2014).

[7] G. T. Silva, L. Tian, A. Franklin, X. Wang, X. Han,S. Mann, and B. W. Drinkwater, Acoustic deformationfor the extraction of mechanical properties of lipid vesiclepopulations, Phys. Rev. E 99, 063002 (2019).

[8] M. Baudoin, J.-C. Gerbedoen, A. Riaud, O. B. Matar,N. Smagin, and J.-L. Thomas, Folding a focalized acous-tical vortex on a flat holographic transducer: Miniatur-ized selective acoustical tweezers, Sci. Adv. 5, eaav1967(2019).

[9] M. Baudoin and J.-L. Thomas, Acoustic tweezers forparticle and fluid micromanipulation, Annu. Rev. Fluid

Mech. 52, 205 (2019).[10] T. Schwarz, G. Petit-Pierre, and J. Dual, Rotation of

non-spherical micro-particles by amplitude modulationof superimposed orthogonal ultrasonic modes, J. Acoust.Soc. Am. 133, 1260 (2012).

[11] Y. Ai, C. K. Sanders, and B. L. Marrone, Separation ofescherichia coli bacteria from peripheral blood mononu-clear cells using standing surface acoustic waves, Anal.Chem. 85, 9126 (2013).

[12] O. Jakobsson, M. Antfolk, and T. Laurell, Continuousflow two-dimensional acoustic orientation of nonsphericalcells, Anal. Chem. 86, 6111 (2014).

[13] T. Schwarz, P. Hahn, G. Petit-Pierre, and J. Dual, Ro-tation of fibers and other non-spherical particles by theacoustic radiation torque, Microfluid Nanofluid 18, 65(2015).

[14] A. Garbin, I. Leibacher, P. Hahn, H. L. Ferrand, A. Stu-dart, and J. Dual, Acoustophoresis of disk-shaped mi-croparticles: A numerical and experimental study ofacoustic radiation forces and torques, J. Acoust. Soc. Am.138, 2759 (2015).

[15] D. Foresti and D. Poulikakos, Acoustophoretic contact-less elevation, orbital transport and spinning of matterin air, Phys. Rev. Lett. 112, 024301 (2014).

[16] A. Marzo, S. A. Seah, B. W. Drinkwater, D. R. Sahoo,B. Long, and S. Subramanian, Holographic acoustic ele-ments for manipulation of levitated objects, Nat. Com-mun. 6, 8661 (2015).

[17] Z. Y. Hong, J. F. Yin, W. Zhai, N. Yan, W. L. Wang,J. Zhang, and B. W. Drinkwater, Dynamics of levitatedobjects in acoustic vortex fields, Sci. Rep. 7, 7093 (2017).

[18] W. Wang, L. A. Castro, M. Hoyos, and T. E. Mallouk,Autonomous motion of metallic microrods propelled by

Page 13: arXiv:2004.12980v1 [physics.app-ph] 27 Apr 2020 · p= 0; (1b) r v= 0: (1c) The time-dependent term e i!t is omitted for simplic-ity. Equation (1c) is obtained by taking the rotational

13

ultrasound, ACS Nano 67, 6122 (2012).[19] L. Ren, N. Nama, J. M. McNeill, F. Soto, Z. Yan, W. Liu,

W. Wang, J. Wang, and T. E. Mallouk, 3d steerable,acoustically powered microswimmers for single-particlemanipulation, Sci. Adv. 5, eaax3084 (2019).

[20] P. Glynne-Jones, P. P. Mishra, R. J. Boltryk, and M. Hill,Efficient finite element modeling of radiation forces onelastic particles of arbitrary size and geometry, J. Acoust.Soc. Am. 133, 1885 (2013).

[21] P. Hahn, I. Leibacher, T. Baasch, and J. Dual, Numer-ical simulation of acoustofluidic manipulation by radia-tion forces and acoustic streaming for complex particles,Lab. Chip. 15, 4302 (2015).

[22] D. W. Greve, E. R. Dauson, and I. J. Oppenheim, Forcesand torques on rods in an ultrasonic standing wave, in2018 IEEE International Ultrasonics Symposium (IUS)(2018) pp. 1–4.

[23] F. B. Wijaya and K.-M. Lim, Numerical calculation ofacoustic radiation force and torque acting on rigid non-spherical particles, Acta Acust. united Ac. 101, 531(2015).

[24] T. S. Jerome, Y. A. Ilinskii, E. A. Zabolotskaya, andM. F. Hamilton, Born approximation of acoustic radia-tion force and torque on soft objects of arbitrary shape,J. Acoust. Soc. Am. 145, 36 (2019).

[25] F. G. Mitri, Acoustic radiation force on oblate and pro-late spheroids in Bessel beams, Wave Motion 57, 231(2015).

[26] F. G. Mitri, Radiation forces and torque on a rigidelliptical cylinder in acoustical plane progressive and(quasi)standing waves with arbitrary incidence, Phys.Fluid 28, 077104 (2016).

[27] Z. Gong, P. L. Marston, and W. Li, T -matrix evaluationof three-dimensional acoustic radiation forces on non-spherical objects in Bessel beams with arbitrary orderand location, Phys. Rev. E 99, 063004 (2019).

[28] M. Kotani, An acoustical problem relating to the theoryof Rayleigh disc, Proc. Phys. Math. Soc. Japan 15, 30(1933).

[29] L. V. King, On the theory of the inertia and diffractioncorrections for the Rayleigh disc, Proc. Royal Soc. A 153,17 (1935).

[30] J. B. Keller, Acoustic torques and forces on disks, J.Acoust. Soc. Am. 29, 1085 (1957).

[31] G. Maidanik, Torques due to acoustical radiation pres-sure, J. Acoust. Soc. Am. 30, 620 (1958).

[32] P. L. Marston, Comment on radiation forces and torqueon a rigid elliptical cylinder in acoustical plane progres-sive and (quasi)standing waves with arbitrary incidence[Phys. Fluids 28, 077104 (2016)], Phys. Fluids 29, 029101(2016).

[33] P. L. Marston, W. Wei, and D. B. Thiessen, Acoustic ra-diation force on elliptical cylinders and spheroidal objectsin low frequency standing waves, in AIP Conf. Proc., 838(2006) pp. 495–499.

[34] G. T. Silva and B. W. Drinkwater, Acoustic radiationforce exerted on a small spheroidal rigid particle by abeam of arbitrary wavefront: Examples of traveling andstanding plane waves, J. Acoustic. Soc. Am. 144, EL453(2018).

[35] Z. Fan, D. Mei, K. Yang, and Z. Chen, Acoustic radiationtorque on an irregularly shaped scatterer in an arbitrarysound field, J. Acoust. Soc. Am. 124, 2727 (2008).

[36] J. P. Leao-Neto, J. H. Lopes, and G. T. Silva, Acousticradiation torque exerted on a subwavelength spheroidalparticle by a traveling and standing plane wave, J.Acoust. Soc. Am. 147, 2177 (2020), submitted for pub-lication.

[37] J. H. Lopes, E. B. Lima, J. P. Leao-Neto, and G. T. Silva,Acoustic spin transfer to a subwavelength spheroidal par-ticle, Phys. Rev. E 101, 043102 (2020).

[38] D. Baresch, J.-L. Thomas, and R. Marchiano, Sphericalvortex beams of high radial degree for enhanced single-beam tweezers, J. Appl. Phys. 113, 184901 (2013).

[39] F. G. Mitri and G. T. Silva, Generalization of the ex-tended optical theorem for scalar arbitrary-shape acous-tical beams in spherical coordinates., Phys. Rev. E 90,053204 (2014).

[40] G. T. Silva, A. L. Baggio, J. H. Lopes, and F. G.Mitri, Computing the acoustic radiation force exertedon a sphere using the translational addition theorem,IEEE Trans. Ultrason. Ferroelectr. Freq. Control 62, 576(2015).

[41] Z. Gong and P. L. Marston, Multipole expansion ofacoustical Bessel beams with arbitrary order and loca-tion, J. Acoust. Soc. Am. 141, EL574 (2017).

[42] J. P. Leao-Neto, J. H. Lopes, and G. T. Silva, Extendedoptical theorem in isotropic solids and its application tothe elastic radiation force, J. Appl. Phys. 121, 144902(2017).

[43] L. Zhang, A general theory of arbitrary Bessel beam scat-tering and interactions with a sphere, J. Acoust. Soc. Am.143, 2796 (2018).

[44] G. T. Silva, Off-axis scattering of an ultrasound Besselbeam by a sphere, IEEE Trans. Ultrason. Ferroelectr.Freq. Control 58, 298 (2011).

[45] F. G. Mitri and G. T. Silva, Off-axial acoustic scatteringof a high-order Bessel vortex beam by a rigid sphere,Wave Motion 46, 392 (2011).

[46] G. T. Silva, J. H. Lopes, and F. G. Mitri, Off-axial acous-tic radiation force of repulsor and tractor bessel beamson a sphere., IEEE Trans. Ultrason. Ferroelectr. Freq.Control 60, 1207 (2013).

[47] J. H. Lopes, M. Azarpeyvand, and G. T. Silva, Acous-tic interaction forces and torques acting on suspendedspheres in an ideal fluid, IEEE Trans. Ultrason. Ferro-electr. Freq. Control 63, 186 (2016).

[48] Z. Gong, P. L. Marston, and W. Li, Reversals of acousticradiation torque in bessel beams using theoretical and nu-merical implementations in three dimensions, Phys. Rev.Appl. 11, 064022 (2019).

[49] D. Zhao, J.-L. Thomas, and R. Marchiano, Computationof the radiation force exerted by the acoustic tweezersusing pressure field measurements, J. Acoust. Soc. Am.146, 1650 (2019).

[50] L. P. Gor’kov, On the forces acting on a small particlein an acoustic field in an ideal fluid, Sov. Phys.-Dokl. 6,773 (1962).

[51] G. T. Silva, Acoustic radiation force and torque on anabsorbing compressible particle in an inviscid fluid, J.Acoust. Soc. Am. 136, 2405 (2014).

[52] J. P. Leao-Neto and G. T. Silva, Acoustic radiation forceand torque exerted on a small viscoelastic particle in anideal fluid, Ultrasonics 71, 1 (2016).

[53] L. Zhang and P. Marston, Acoustic radiation torque onsmall objects in viscous fluids and connection with vis-cous dissipation, J. Acoust. Soc. Am. 136, 2917 (2014).

Page 14: arXiv:2004.12980v1 [physics.app-ph] 27 Apr 2020 · p= 0; (1b) r v= 0: (1c) The time-dependent term e i!t is omitted for simplic-ity. Equation (1c) is obtained by taking the rotational

14

[54] A. D. Pierce, Acoustics: An Introduction to Its PhysicalPrinciples and Applications, 3rd ed. (Springer, Switzer-land, 2019) p. 489.

[55] C. Shi, R. Zhao, Y. Long, S. Yang, Y. Wang, H. Chen,J. Ren, and X. Zhang, Observation of acoustic spin, Nat.Sci. Rev. 6, 707 (2019).

[56] K. Y. Bliokh and F. Nori, Spin and orbital angularmomenta of acoustic beams, Phys. Rev. B 99, 174310(2019).

[57] G. T. Silva, T. P. Lobo, and F. G. Mitri, Radiation torqueproduced by an arbitrary acoustic wave, Europhys. Lett.97, 54003 (2012).

[58] L. Zhang and P. L. Marston, Optical theorem for acousticnon-diffracting beams and application to radiation forceand torque, Biomed. Opt. Expr. 4, 1610 (2013).

[59] L. Zhang and P. L. Marston, Geometrical interpretationof negative radiation forces of acoustical Bessel beams onspheres., Phys. Rev. E 84, 035601(R) (2011).

[60] A. Silbiger, Scattering of sound by an elastic prolatespheroid, J. Acoust. Soc. Am. 35, 564 (1963).

[61] S. Xu, C. Qiu, and Z. Liu, Transversally stable acous-tic pulling force produced by two crossed plane waves,Europhys. Lett. 99, 44003 (2012).

[62] O. Dron and J.-L. Aider, Acoustic energy measurementfor a standing acoustic wave in a micro-channel, Euro-phys. Lett. 97, 44011 (2012).

[63] C. P. Lee and T. G. Wang, Near-boundary streamingaround a small sphere due to two orthogonal standingwaves, J. Acoust. Soc. Am. 85, 1081 (1989).

[64] J. P. Leao-Neto, J. H. Lopes, and G. T. Silva, Core-shell particles that are unresponsive to acoustic radiationforce, Phys. Rev. Applied 6, 024025 (2016).