1,092
Views
0
CrossRef citations to date
0
Altmetric
Research Article

Numerical modelling and theoretical analysis of the acoustic attenuation in bubbly liquids

, , , , , & show all
Article: 2210193 | Received 01 Mar 2023, Accepted 28 Apr 2023, Published online: 12 May 2023

Abstract

The propagations of acoustic waves in bubbly liquids have been extensively investigated through experimental and theoretical methods, a computational fluid dynamics (CFD) method was introduced to investigate the acoustic attenuation through bubbles in this study. Numerical ‘measurements’ of attenuation coefficient (α) and phase velocity (V) were conducted using a homogeneous cavitation model, which were modeled from experimental schemes and compared with the theoretical results. At extremely small bubble volume fraction (β around 10−11), new formulas of the theoretical α (αtheo) were proposed respectively for linear and transient oscillations, and a new formula of the numerical α (αnum) was proposed for transient oscillations. Results showed that αnum matched precisely with αtheo for linear, nonlinear and transient oscillations. At medium β (around 10−4), the relative difference of αnum between the VOF and present methods was less than 1.6%, while it reached 15.4% after replacing the bounded Keller-Miksis equation (KME) in the present method with the KME. However, the traditional theoretical α and V matched precisely with the predictions by the present method with the KME. Thus new theoretical α and V were proposed based on the bounded KME, and the relative differences between αtheo and αnum were less than 1%. It can be concluded that the bounded KME should be used in both numerical and theoretical predictions.

Nomenclature

α=

attenuation coefficient

αtheo=

theoretical α

αlin=

αtheo of linear oscillations

αnonlin=

αtheo of nonlinear oscillations

αtran=

αtheo of transient oscillations

αm=

αtheo at medium β using the KME

αmb=

αtheo at medium β using the bounded KME

αnum=

numerical α

αnumt=

numerical α of transient oscillations

αexp=

experimental α

αrad=

α due to radiation

β=

bubble volume fraction

σ=

surface tension coefficient

σs=

scattering cross section

δ=

damping coefficient

d=

acoustic path length

n=

bubble number density

pA=

amplitude of the acoustic excitation

1. Introduction

Ultrasonic cavitation has been widely applied in biomedical ultrasonic diagnosis and treatment (Helfield et al., Citation2021; Stride et al., Citation2020). The former application (contrast-enhanced ultrasound imaging) uses ultrasound contrast agents (UCAs) to increase the echo difference between the examination part and the surrounding tissues, so as to obtain clearer ultrasonic images (Yusefi & Helfield, Citation2022). The latter application often introduces UCAs as the cavitation nuclei and needs a high-intensity ultrasound to cause violent collapse, so as to achieve the purpose of drug/gene delivery (Alishiri et al., Citation2021; Yang et al., Citation2020), cancer treatment (Clark et al., Citation2021), sonothrombolysis (Nederhoed et al., Citation2021), and so on. UCAs are coated microbubbles with diameters ranging from 0.7–10 µm. They are made of high-molecular-weight inert gases and coated by lipid, protein, or polymer shells (Frinking et al., Citation2020) to improve their stability.

The Rayleigh-Plesset (RP) equation (Plesset & Prosperetti, Citation1977) provided the theoretical prediction of the dynamics of a spherical uncoated bubble in an unbounded liquid. Several extensions of the RP equation have been proposed to take into account the liquid compressibility (Brenner et al., Citation2002), such as the famous KME (Keller & Miksis, Citation1980). Bubble oscillations exhibit nonlinear behaviours under middle or high-intensity ultrasound, and the nonlinear characteristics are enhanced by the bubble shells of UCAs (Gümmer et al., Citation2021; Omoteso et al., Citation2021). It has been more than two decades of progress in the modelling of UCAs (Helfield, Citation2019; Versluis et al., Citation2020). By adding pressure terms to the RP type equations, many models (Allen et al., Citation2002; Cowley & McGinty, Citation2019; de Jong et al., Citation1994; Marmottant et al., Citation2005; Paul et al., Citation2010) have been proposed to consider the influence of bubble shells. The influence is mainly described by two terms: the shell elasticity and the dilatational viscosity (κs). The Marmottant model (Marmottant et al., Citation2005) is widely used and can well predict the nonlinear oscillation behaviours, such as the ‘compression-only’ and ‘thresholding’ (Versluis et al., Citation2020). In this model, the shell elasticity was embedded in the surface tension (σ), which was discontinuous when the buckling or rupture occurred. Paul et al. (Citation2010) adopted an exponential elasticity model to make a smooth transition from σbreak−up to σL. κs increases with the ambient bubble radius (R0), and it is usually given as a constant in the range of 10−10 kg/s to 10−8 kg/s (Helfield & Goertz, Citation2013; Tu et al., Citation2009; Tu et al., Citation2011).

The measurements of α can be used to obtain the shell parameters of UCAs. UCAs are around resonances under the most common frequencies of medical ultrasound (1-7 MHz) (Gorce et al., Citation2000), thus their dynamics are greatly dependent on R0. For small-amplitude oscillations, RP type equations can be simplified to linear equations of motion, based on which αtheo can be obtained. Gorce et al. (Citation2000) divided UCAs into groups and theoretically analysed the influence of R0 on echogenicity and attenuation based on the simplification of linear oscillation. Segers et al. (Citation2016) measured the scattering and attenuation of sorted UCAs at higher driving pressure, and calculated αtheo by solving the RP type equation to capture the nonlinear behaviours. At medium β, the dependences of α and V on β need to be considered. Silberman (Citation1957) conducted experimental measurements of α and V in a standing wave tube at β above 0.03%. Wijngaarden (Citation1972) proposed a continuum approach to model the wave propagation in bubbly liquids. Commander and Prosperetti (Citation1989) developed a linear model to predict α and V at small driving pressure and medium β, and good agreements were obtained with the experiments by Silberman (Citation1957) et al. Leroy et al. (Citation2009) experimentally measured the transmission coefficients of ultrasound through a single layer of bubbles and found that the resonant frequency increased with the bubble concentration, which was contrary to the conclusion of traditional theoretical analyzes (Ida et al., Citation2007; Yasui et al., Citation2008; Yasui et al., Citation2009).

Numerical simulations have been widely used to investigate cavitation (Fuster, Citation2019; Modarreszadeh et al., Citation2021; Si et al., Citation2023; Subroto et al., Citation2021; Suo et al., Citation2021; Yasui, Citation2021). The interface capturing methods, such as the VOF method (Chen et al., Citation2021; Lee et al., Citation2021; Li et al., Citation2022; Yamamoto & Komarov, Citation2021; Ye et al., Citation2021b), can accurately predict bubble dynamics. However, numerous grids are needed to capture each bubble, and it is difficult to model the bubble shell of UCAs. Homogeneous cavitation models are almost the only way for the predictions of cavitation with numerous bubbles. Ye et al. (Ye et al., Citation2020a; Ye et al., Citation2020b) proposed a homogeneous cavitation model which can accurately predict the acoustic cavitation of the equal-sized spherical bubbles in a stationary liquid. Due to the dependency of bubble dynamics on R0, Ye et al. (Citation2021a) proposed a homogeneous grouping cavitation model which divided bubbles into groups, and the dynamics of each group were separately calculated.

Since the bubble oscillations can be well predicted by the homogeneous cavitation model (Ye et al., Citation2021a), the acoustic attenuation through bubbles should also be well predicted. Numerical simulations can not only investigate the complex cavitating flows, but also help to improve the basic theory. This study conducted theoretical and numerical investigations on the attenuation and phase velocity through uncoated and coated bubbles. At extremely small β, the inter-bubble interaction can be neglected. αtheo caused by oscillating bubbles in the linear (αlin2) and transient (αtran) states were obtained, and several simplified forms of αtheo for linear and nonlinear oscillations were introduced in Section 2. The numerical ‘measurements’ of α were implemented by comparing the recorded pressure with and without bubbles in the acoustic beam, which were modeled from experimental schemes, and comparisons were made with αtheo in Section 3. At medium β, the present method was validated by comparing with the VOF method, and then comparisons were made with the experimental (Silberman, Citation1957) and theoretical (Commander & Prosperetti, Citation1989) results of millimetre bubbles, based on which the theoretical formulas were improved (Section 3.4). At last, α and V through microbubbles were numerically measured at medium β, and the influences of the nonlinearity and transience were numerically analysed.

2. Methods

2.1. Bubble dynamics equation

Neglecting the liquid compressibility, the phase change, and the diffusion of non-condensable gas into liquid, the dynamics of a spherical bubble under far-field acoustic excitation can be described by the RP type equation (Plesset & Prosperetti, Citation1977): (1) RR¨+32R˙2=1ρL[pG2σR4(μL+μth)R˙Rp0pa(t)](1) (2) pG=pG0(R0R)3κ(2) where R is the bubble radius at time t, dots denote time derivatives, ρL is the liquid density, pG is the gas pressure inside the bubble, the subscript 0 denotes the initial equilibrium value without excitation, σ is the surface tension coefficient, µL is the liquid viscosity, µth is the effective thermal viscosity, pa(t) is the acoustic excitation, and κ is the polytropic exponent. The famous KME (Keller & Miksis, Citation1980) considering the liquid compressibility was given as: (3) (1R˙cL)RR¨+32(1R˙3cL)R˙2=1ρL(1+R˙cL+RcLddt)×(pG2σR4(μL+μth)R˙Rp0pa(t))(3) where cL is the speed of sound in liquid. Substituting Eq. (1) into Eq. (3), another form of the KME can be obtained (Zhang, Citation2013): (4) (12R˙cL+M)RR¨+(322R˙cLM)R˙2=1ρL[pG2σR4(μL+μth)R˙R(3κpG2σR)R˙cLp˙a(t)RcLp0pa(t)](4) (5) M=4(μL+μth)ρLcLR(5) M ≪ 1 for most oscillations, and |R˙|/cL≪ 1 for small-amplitude oscillations, thus a popular simplified form can be obtained by further neglecting the time differentials of the surface tension and pa(t) in Eq. (3) (Brenner et al., Citation2002): (6) RR¨+32R˙2=1ρL[pG(13κR˙cL)2σR4(μL+μth)R˙Rp0pa(t)](6) Marmottant et al. (Citation2005) proposed an equation for coated bubbles based on Eq. (6) (Segers et al., Citation2016; Segers et al., Citation2018): (7) RR¨+32R˙2=1ρL[pG(13κR˙cL)2σ(R)R4(μL+μth+κs/R)R˙Rp0pa(t)](7) (8) pG0=p0+2σ(R0)R0(8) (9) σ(R)={χmax(R2Rb21,0)RRrupturedσR>Rruptured(9) where κs is the dilatational viscosity of the shell, χ is the elastic modulus, Rb is the buckling radius, and Rruptured is the rupture radius. Eqs. (6) and (7) are named as the simplified KME in this study. The complete KME for coated bubbles can be obtained by substituting the Marmottant model (Marmottant et al., Citation2005) into Eq. (3): (10) (12R˙cL+M)RR¨+(322R˙cLM4κsρLcLR2)R˙2=1ρL[pG2σ(R)R4(μL+μth+κs/R)R˙R(3κpG+2σ(R)R+4χR)R˙cLp˙a(t)RcLp0pa(t)4κsρLcLR2](10) (11) M=4(μL+μth+κs/R)ρLcLR(11)

2.2. Linear equation and damping coefficients

For the small-amplitude oscillation about the equilibrium radius R = R0(1 + x) (x ≪ 1), the KME (Eq. (4) or (10)) can be linearised to the classic equation by neglecting high-order terms (Leighton, Citation1994; Versluis et al., Citation2020): (12) x¨+δtotω0x˙+ω02x=pAρLR02sin(ωt)(12) where δtot is the total damping coefficient, pA is the amplitude of the acoustic excitation, and ω0 is the angular eigenfrequency. ω0 of uncoated and coated bubbles can respectively be obtained as: (13) ω02=3κp0+2(3κ1)σ/R0ρLR02(1+M(R0))+δtotω0ω2R0cL(13) (14) ω02=3κp0+2(3κ+1)σ(R0)/R0+4χ/R0ρLR02(1+M(R0))+δtotω0ω2R0cL(14) where δtotω0ω2R0/cL is from the linearisation of p˙a(t)R/cL, and δtotω0 is independent on ω0. δtot is a summation of the damping coefficients due to the heat diffusion (δth), the re-radiation of sound (δrad), the viscosities of liquid (δvis) and shell (δshell). The summation of δvis, δth, and δshell can be obtained as: (15) δvis+δth+δshell=4(μL+μth+κs/R0)ρLω0R02(1+M(R0))(15) and δrad of both uncoated and coated bubbles are obtained as (Leighton, Citation1994; Li et al., Citation2018): (16) δrad1=ω0R0cL+(ω2ω02)R0ω0cL=ω2R0ω0cL(16) where (ω2ω02)R0/ω0cL is from the linearisation of p˙a(t)R/cL, and ω0R0/cL is from the linearisation of the other radiation terms on the right side of the KME. If p˙a(t)R/cL in the KME is neglected, another δrad can be given as (Helfield, Citation2019): (17) δrad2=ω0R0cL(17)

If the differential of the surface tension is further neglected, i.e. the linearisation of the simplified KME (Eq. (6) or (7)), one more expression of δrad can be obtained (Paul et al., Citation2010): (18) δrad3=3κpG0ρLω0cLR0(18) where pG0 was replaced by p0 in Refs. (Segers et al., Citation2016; Segers et al., Citation2018; Versluis et al., Citation2020).

2.3. Theoretical attenuation coefficient

2.3.1. Linear analysis

As mentioned above, the attenuation of a plane wave travelling through bubbles is due to the liquid viscosity, heat diffusion, and scattering. The scattered pressure at the bubble wall can be obtained as (Segers et al., Citation2016; Versluis et al., Citation2020): (19) ps(t)=ρL(RR¨+2R˙2)(19) The energy loss due to the scattering of a single bubble is usually expressed by the scattering cross-section σs, which is defined as the ratio of the scattering energy to the acoustic intensity. For the small-amplitude oscillation, the linearised σs is given as (Segers et al., Citation2016; Versluis et al., Citation2020): (20) σs=4πR02|ps|2pA2(20)

The amplitude of the scattered pressure |ps| can be approximately obtained as (Helfield, Citation2019): (21) |ps|ρLω2R02|x|(21) where the relative amplitude |x| can be obtained by solving the analytic solution of the linearised equation of motion (Eq. (12)). Substituting Eq. (21) into Eq. (20), the linearised σs can be rewritten as (Versluis et al., Citation2020): (22) Ωs=4πR02(ω02ω21)2+(δtotω0ω)2(22)

The attenuation by a single bubble can be obtained by multiplying Ωs by δtot/δrad. Neglecting the inter-bubble interaction, αtheo of the small-amplitude wave travelling through bubbles in unit distance can be obtained by the summation of the attenuation caused by the bubbles in the acoustic beam (Versluis et al., Citation2020): (23) αlin1=10ln10R0(n(R0)Ωsδtotδrad)(23) where n(R0) is the bubble number density.

In addition, the energy loss rate of an uncoated bubble also can be obtained by integrating the work of all damping forces (Jamshidi & Brenner, Citation2013; Leighton, Citation1994): (24) W˙disp1=f01/f[(4(μL+μth)RR˙+3κpG2σ/RcLR˙+p˙a(t)RcL)4πR2R˙dt]=2πρLδtot(1+M(R0))ω2R05|x|2(24) Then αtheo of the small-amplitude oscillations also can be obtained by the ratio of W˙disp1 and sound intensity: (25) αlin2=10ln10R0(n(R0)W˙disp1pA2/2ρLcL)=10ln10R0(n(R0)Ωsδtotω0cLR0ω2)(25)

The above result is also applicable for coated bubbles. It can be seen that αlin1 = αlin2 when δrad = δrad1.

2.3.2. Nonlinear analysis

With the increase of pA, the bubble oscillations gradually exhibit nonlinear behaviours, which make αlin1 and αlin2 inapplicable. In order to predict αtheo considering the nonlinearity, Segers et al. (Citation2016) solved the simplified KME numerically using the ode45 solver in MATLAB, and then calculated |ps| and σs, respectively, by Eqs. (19) and (20). Although the damping coefficients are dependent on pA, Segers et al. supposed that δtot/δrad was independent on pA, and αnonlin was obtained as (Segers et al., Citation2016): (26) αnonlin1=10ln10R0(n(R0)σsδtotδrad)(26) It should be noted that Eq. (20) is simplified for small-amplitude oscillations (RR0). If the relative amplitude |x| is considerable, the fluctuation of the bubble surface area needs to be considered in Eq. (20). Moreover, the supposition of δtot/δrad independent on pA is in doubt.

After solving the bubble dynamics equation numerically, as mentioned in Section 2.3.1, the energy loss rate can be obtained by integrating the work of all damping forces (Leighton, Citation1994): (27) W˙disp2=f01/f(4(μL+μth)RR˙+3κpG2σ/RcLR˙+p˙a(t)RcL(2R˙cLM)RR¨(2R˙cL+M)R˙2)4πR2R˙dt(27) The energy loss of a coated bubble can be similarly obtained. Then αnonlin can be obtained by: (28) αnonlin2=10ln10R0(n(R0)W˙disp2pA2/2ρLcL)(28)

In addition, the energy loss should equal the energy input to the bubble from the acoustic excitation, thus W˙disp also can be obtained as (Leighton, Citation1994): (29) W˙disp3=f01/f[pa(t)4πR2R˙dt](29) Then αnonlin also can be obtained by: (30) αnonlin3=10ln10R0(n(R0)W˙disp3pA2/2ρLcL)=10ln10R0(n(R0)pa(t)4πR2R˙¯pA2/2ρLcL)(30) where the overbar denotes the time average value over one cycle.

2.3.3. Transient analysis

The above analyzes are available for bubble oscillations in the steady (periodic) state. During experimental measurements, a dozen cycles are transmitted and then recorded, and α of these cycles is different from that in the steady state. Usually, pA of these cycles gradually increases at the beginning and decreases in the final stage (as shown in Figure ). At the beginning, the energy absorbed by the bubble is not only dissipated but also converted to mechanical energy. In the final stage, the mechanical energy stored by the bubble is released. If the energy loss is calculated by integrating the work of all damping forces (Segers et al., Citation2016), for example by Eq. (27), the mechanical energy is neglected. According to our simulations, the energy loss should be calculated by Eq. (29) to include the mechanical energy. The average sound intensity of these cycles before the attenuation is 0tpa(t)2dt/tρLcL. Thus, αtheo of these cycles can be obtained as: (31) αtran=10ln10R0(n(R0)0tpa(t)4πR2R˙dt0tpa(t)2dt/ρLcL)(31)

2.3.4. Multibubble analysis

The obtained αtheo in Sections 2.3.1–2.3.3 are only applicable at extremely small β. At medium β, the wave number in the mixture of a monodisperse bubble population deduced by Commander and Prosperetti (Citation1989) was given as: (32) km2=ω2cL2+4πω2n(R0)R0ω02ω2+δtotω0ωi11+M(R0)(32) where M(R0) was neglected in all previous researches. αm and Vm in the mixture were respectively given by (Commander & Prosperetti, Citation1989; Trujillo, Citation2018): (33) αm=20logeIm(km)(33) (34) Vm=ωRe(km)(34) where Im and Re are respectively the imaginary and real parts.

On the other hand, for a bubble cluster including several or dozens of bubbles, the bubble dynamics also can be predicted considering the inter-bubble interaction (Ida et al., Citation2007; Yasui et al., Citation2008; Yasui et al., Citation2009). For the bubble oscillation under far-field acoustic excitation, the scattered pressure at a distance of r from the bubble centre was usually obtained by (Ida et al., Citation2007; Yasui et al., Citation2008): (35) ps=prp=ρL(R2R¨+2RR˙2)r(35) where the liquid compressibility was neglected. Assuming that the adjacent bubbles had the same dynamics, the inter-bubble interaction was usually considered by adding the scattered pressure caused by the surrounding bubbles to the right side of RP type equations (Ida et al., Citation2007; Yasui et al., Citation2008; Yasui et al., Citation2009): (36) (iRri)(RR¨+2R˙2)(36) where ri is the inter-bubble distance. Adding Eq. (36) to the right side of Eq. (4), ω0 can be obtained after the linearisation: (37) ω02=3κp0+2(3κ1)σ/R0ρLR02(1+M(R0)+R0ri1)+δtotω0ω2RcL(37) which decreases with the bubble concentration. However, according to the experimental investigation conducted by Leroy et al. (Citation2009), ω0 increased with the bubble concentration. In our opinion, the main reason might be that the acoustic excitation is supposed to be acting at far field in Eq. (4), and the liquid compressibility is neglected in Eq. (35).

2.4. Experimental attenuation coefficient

The experimental α can be measured by two aligned transducers, where one transmits acoustic waves (usually a dozen cycles) and the other records the pressure (Li et al., Citation2018; Segers et al., Citation2016), as shown in Figure . α can be obtained by comparing the recorded pressure with and without bubbles in the acoustic beam (Segers et al., Citation2016): (38) αexp=10dlog10|pref|2|pbub|2(38) where |pbub|2 and |pref|2 are respectively the amplitudes of the power spectra of the recorded pressure with and without bubbles, and d is the acoustic path length over which the wave interacts with the bubbles.

Figure 1. Schematic diagram of the measurements of α.

Figure 1. Schematic diagram of the measurements of α.

2.5. Numerical modelling

2.5.1. Numerical measurements of α and V

The numerical measurements were modelled from the experiments described in Section 2.4. Figure  shows the 2-dimensional computational grids and boundary conditions for the numerical measurements of α and V. The computational domain was 30λ (λ is the wavelength) in both length and height. The line at the bottom was the axis, and the acoustic excitation was applied on the left boundary. The non-reflecting pressure outlet was specified at the top and right boundaries. Bubbles distributed in the region marked red, whose left border was 8λ away from the left boundary. A monitoring line was created to record the pressure with and without bubbles in the acoustic beam. It should be noted that although the non-reflecting pressure outlet was applied, there was still pressure reflecting back from the right boundary at an amplitude of nearly 1%. Thus, the monitoring line was 15λ away from the right boundary, and only the wave before the 30th cycle was recorded to exclude the interference of reflecting pressure. In addition, the scattered pressure from bubbles was considerable at medium β, it reflected from the left boundary and interfered with the recorded pressure at the monitoring line. Therefore, the distance between bubbles and the left boundary was increased to 15λ at medium β (Section 3.4). Based on the recorded pressure, αnum was obtained similar to αexp using Eq. (38), i.e. αnum = αexp. Only V of steady oscillations at medium β were investigated in this study. The recorded pressure (reaching the steady state) with and without bubbles were respectively fitted by sine functions to obtain the time lags tbub and tref, then Vnum was obtained by: (39) 1Vnum=1cLtreftbubd(39) The acoustic excitation was a sine wave multiplied by two coefficients Cy and Cp. Cy was dependent on the ratio of y coordinate to λ, as shown in Figure (a). Figure (b) shows the pressure distribution at pA = 10 Pa after the acoustic wave traveling through the whole computational domain. It can be seen that the acoustic excitation concentrated around the axis, so that only these bubbles around the axis were irradiated, and the influence of scattered pressure from these bubbles on the recorded pressure was negligible. Cp was used to generate a 20-cycle or Gaussian wave when analysing α in the transient state, as shown in Figure . After the grid and time-step independences verification, 242 thousand quadrilateral grids were employed, and the simulation time-step size was 0.004/f. The grid was concentrated in the direction of wave propagation, and the grid number per wavelength was 50.

Figure 2. Computational domain and boundary conditions for the numerical ‘measurements’ of α and V, which were modelled from the experimental schemes. The acoustic excitation was applied on the left boundary. The area-weighted average pressure at the monitoring line were recorded with and without bubbles in the acoustic beam, based on which αnum and Vnum were obtained.

Figure 2. Computational domain and boundary conditions for the numerical ‘measurements’ of α and V, which were modelled from the experimental schemes. The acoustic excitation was applied on the left boundary. The area-weighted average pressure at the monitoring line were recorded with and without bubbles in the acoustic beam, based on which αnum and Vnum were obtained.

Figure 3. (a) Dependence of Cy on y/λ, and (b) the pressure distribution at pA = 10 Pa. The plane wave concentrated around the axis to reduce the scattered pressure from bubbles far from the axis.

Figure 3. (a) Dependence of Cy on y/λ, and (b) the pressure distribution at pA = 10 Pa. The plane wave concentrated around the axis to reduce the scattered pressure from bubbles far from the axis.

Figure 4. Waveforms of the 20-cycle and Gaussian pulses at pA = 10 Pa, which were used to analyse α in the transient state.

Figure 4. Waveforms of the 20-cycle and Gaussian pulses at pA = 10 Pa, which were used to analyse α in the transient state.

In numerical simulations, the far-field pa(t) in the RP equation is unknown, and it needs to be replaced by the local pressure. Thus the bounded RP equation neglecting the liquid compressibility was used in Refs. (Ye et al., Citation2020b; Ye et al., Citation2021a), and the local pressure was supposed to act at the distance of Re away from the bubble. Re was the radius of the region occupied by the bubble, which means R/Re = β1/3. Considering the liquid compressibility, the bounded KME for the numerical simulations of uncoated bubbles was obtained as: (40) (1RRe2R˙cL+M)RR¨+(322RRe2R˙cLM)R˙2=1ρL[pG2σR4(μL+μth)R˙R(3κpG2σR)R˙cLp˙RcLp](40) where p is the local pressure. The bounded KME reduces to the KME when β0 tends to 0. The equation for coated bubbles can be similarly obtained by adding R/Re to Eq. (10).

The present homogeneous method was implemented in the commercial software ANSYS FLUENT. The density and viscosity of the mixture were defined as the volume-weighted average values of the two phases. The slip velocity between phases was neglected. The bubble wall velocity R˙ was recorded by an added transport equation, whose source term was equal to R¨ calculated by Eq. (40). After obtaining the bubble dynamics, the change rate of β was given as (Ye et al., Citation2020b): (41) dβdt=4πnR2R˙(41) The liquid compressibility was modelled by the Tait's equation of state for water (Lauer et al., Citation2012): (42) (ρLρL0)7.15=1+7.15(pp0)K0(42) where ρL0 = 1000 kg/m3 and K0 = 2.2 × 109 Pa. The other physical parameters for microbubbles were as follows: p0 = 100 kPa, ρG = 1 kg/m3, μL= µth = 0.001 Pa·s, σ = 0.072 N/m, χ = 0.5 N/m, κs = 10−8 kg/s, Rb = 0.97R0 (σ(R0) = 0.0314 N/m), Rruptured = 1.2R0, κ = 1.4 for uncoated bubbles and 1.07 for coated bubbles. All flows were considered to be laminar and the gravity was neglected.

2.5.2. Validation at medium β

In addition to the numerical measurements described in Section 2.5.1, comparisons were made with the predictions by the interface capturing method VOF (also using ANSYS FLUENT). Figure  shows the computational grids and boundary conditions of the validation cases. Acoustic excitation was applied on the left boundary and traveled through a pipe. 49 bubbles were distributed around the outlet of the pipe, and they were arranged neatly (7 × 7) on the cross-section parallel to the left boundary at a distance of 8λ. The inter-bubble distance was 10−4 m, which means n = 1012 m−3. A monitoring face was created at a distance of 2λ from bubbles to obtain αnum, and only the wave before the 16th cycle was recorded. Two symmetry planes were applied as shown in Figure , and only one-eighth of the bubbles were inside the computational domain. For the VOF method, special contractions of the mesh were applied around bubble surfaces, about 1 million tetrahedral grids were employed around the bubbles, and 0.5 million hexahedral grids were employed in the other region; the gas inside the bubbles obeyed the ideal-gas law. For the present method, the grid number around bubbles was decreased to 0.1 million, while the grids in the other region were kept the same. It is complex for the VOF method to handle the surface tension of microbubbles, especially for coated microbubbles. It is also complex for the homogeneous models to precisely predict the heat transfer. For a better comparison, the surface tension was neglected, and the flow was assumed to be isothermal, which is the same as in Refs. (Ye et al., Citation2020b; Ye et al., Citation2021a). These two simplifications were only used in this case.

Figure 5. Computational grids and boundary conditions for the measurement of α through 49 bubbles when using the VOF method.

Figure 5. Computational grids and boundary conditions for the measurement of α through 49 bubbles when using the VOF method.

Figure 6. (a) Distribution of bubbles on the cross-section when using the VOF method. (b) Special contractions of the mesh were applied around bubble surfaces.

Figure 6. (a) Distribution of bubbles on the cross-section when using the VOF method. (b) Special contractions of the mesh were applied around bubble surfaces.

In order to further validate the present method at medium β, the experimental measurements of α and V conducted by Silberman (Citation1957) were simulated at β0 = 3.77 × 10−4. Standing waves were established inside a vertical tube filled with air bubbles in the experiments. Since the pressure can be precisely obtained during numerical simulations, there is no need to generate standing waves, and the computational domain shown in Figure  was still used. Both R0 and β0 were medium in these cases, it took a long time for bubbles to reach the steady state, especially around resonance. For simplicity, bubbles were uniformly distributed in the whole computational domain, just as the sound generator and hydrophone were placed inside the bubbly liquid in the experiments (Silberman, Citation1957). And two monitoring points (points 1 and 2) were created close to the left boundary to record the pressure, as sketched in Figure . αnum was obtained by comparing the power spectra of the pressure at these two points. The pressure at these two points were also fitted by sine functions to obtain the time lags t1 and t2, Vnum of these cases were obtained by: (43) Vnum=d12t2t1(43) The thermal damping was significantly dependent on f in these cases, µth and κ were obtained by (Prosperetti et al., Citation1988): (44) μth=pG04ωReΦ(44) (45) κ=ImΦ/3(45) (46) Φ=3γ13(γ1)iX[(i/X)1/2coth(i/X)1/21](46) (47) X=DGωR02(47) where γ = 1.4 is the ratio of specific heats, DG = 2.1 × 10−5 m2/s is the thermal diffusivity of the gas.

3. Results and discussion

The acoustic attenuation caused by uncoated and coated bubbles were numerically ‘measured’ at pA = 10 Pa in Section 3.1, and comparisons were made with the theoretical αlin1 and αlin2. Each αlin was respectively predicted using δrad1 (αlin1 = αlin2), δrad2 (neglecting p˙a(t)), and δrad3 (neglecting p˙a(t) and the differential of surface tension). Then pA was increased to 10 and 20 kPa in Section 3.2, αnum were compared with αnonlin1 (Segers et al., Citation2016), αnonlin2 (work of damping forces), and αnonlin3 (work of pa(t)). In Section 3.3, the 20-cycle and Gaussian pulses were applied, and αnum of these pulses were compared with αtran. αnonlin and αtran were obtained by solving the KME using the ode solver in MATLAB. For better comparisons with αtheo, n2/3Ωs was set to 10−5 in Sections 3.1-3.3, n values were mostly between 104 and 105 m−3 at f = 1 MHz (β0 were around 10−11). In Section 3.4, α and V at medium β0 (the maximum n2/3Ωs exceeded 0.1) were numerically measured, and comparisons were made with the experimental (Silberman, Citation1957) and theoretical results.

3.1. Small-amplitude oscillations in the steady state

Figure  compares αlin and αnum caused by uncoated bubbles at pA = 10 Pa, f = 1 and 3 MHz. Firstly, the difference of αlin1 predicted using δrad1, δrad2, and δrad3 is obvious. Especially when the differential of the surface tension is neglected, αlin1 is obviously under-estimated around resonance, since δrad3 is overpredicted. Secondly, the deviations of αlin2 predicted using δrad2 and δrad3 are much smaller. When δrad2 is used, deviations mainly appear for bubbles far from resonance. When δrad3 is used, the deviations increase for bubbles around resonance, taking f = 1 MHz for example, the relative deviation increases to 1% at resonance. Thirdly, αnum match precisely with αlin2, both by using the KME (comparing with αlin1-δrad1) and the simplified KME (comparing with αlin2-δrad3). It can be concluded that αlin1 is only applicable when δrad1 is used, and if the other δrad is used, the total energy loss should be directly obtained by integrating the work of all damping forces, not by multiplying the scattering energy by δtot/δrad. In other words, αlin2 should be used if the differential of pa(t) or surface tension is neglected.

Figure 7. Comparison of αlin and αnum caused by uncoated bubbles at pA = 10 Pa, f = (a) 1 and (b) 3 MHz. αnum predicted using both the KME and simplified KME match precisely with the corresponding αlin2.

Figure 7. Comparison of αlin and αnum caused by uncoated bubbles at pA = 10 Pa, f = (a) 1 and (b) 3 MHz. αnum predicted using both the KME and simplified KME match precisely with the corresponding αlin2.

Figure 8. Comparison of αlin and αnum caused by coated bubbles at pA = 10 Pa, f = (a) 1 MHz and (b) 3 MHz. αlin1 linearised from the simplified KME (δrad3) is obviously overpredicted, since the differential of the surface tension in the KME cannot be neglected for coated bubbles.

Figure 8. Comparison of αlin and αnum caused by coated bubbles at pA = 10 Pa, f = (a) 1 MHz and (b) 3 MHz. αlin1 linearised from the simplified KME (δrad3) is obviously overpredicted, since the differential of the surface tension in the KME cannot be neglected for coated bubbles.

Figure  compares αlin and αnum caused by coated bubbles at pA = 10 Pa, f = 1 and 3 MHz. αlin1 obtained using δrad3 is overpredicted by more than 100%, the reason is that the differential of the surface tension is dominant in δrad for coated bubbles. αlin2 obtained using δrad3 is also obviously overpredicted around resonance, for example, it is overpredicted by 7.6% at resonance at f = 1 MHz (much larger than uncoated bubbles). αnum predicted using the two equations still match precisely with αlin2. It can be concluded that the differential of the surface tension cannot be neglected for coated bubbles, and the proposed numerical measurements of α match precisely with αtheo for small-amplitude oscillations. In the following sections, αlin2 obtained using δrad1 is used as a reference.

3.2. Nonlinear oscillations in the steady state

Figure  compares αnonlin and αnum caused by uncoated bubbles at f = 1 MHz, pA = 10 and 20 kPa. ω0 decreases with increasing pA (Helfield, Citation2019). αnum match precisely with αnonlin2 and αnonlin3, while αnonlin1 (using δrad1) is obviously overpredicted. There are two reasons leading to the deviation of αnonlin1. Firstly, σs (Eq. (20)) is inapplicable for nonlinear oscillations, since the bubble surface area cannot be treated as a constant. Secondly, the ratio of the total energy loss to scattering energy is dependent on pA. To prove this, the attenuation due to radiation (αrad) is obtained by only keeping the radiation damping force in αnonlin2, and the total attenuation is obtained by multiplying αrad by δtot/δrad. It can be seen from Figure  that although αrad is precisely predicted, the attenuation is still overpredicted, which means that the ratio of the total energy loss to scattering energy is overpredicted by δtot/δrad.

Figure 9. Comparison of αnonlin and αnum caused by uncoated bubbles at f = 1 MHz, pA = (a) 10 kPa and (b) 20 kPa. αnum match precisely with αnonlin2 (work of damping forces) and αnonlin3 (work of pa(t)).

Figure 9. Comparison of αnonlin and αnum caused by uncoated bubbles at f = 1 MHz, pA = (a) 10 kPa and (b) 20 kPa. αnum match precisely with αnonlin2 (work of damping forces) and αnonlin3 (work of pa(t)).

Figure  compares αnonlin and αnum caused by coated bubbles at f = 1 MHz, pA = 10 and 20 kPa. αnum still match precisely with αnonlin2 and αnonlin3. αnonlin1 and αradδtot/δrad are respectively smaller and larger than αnum, which means that the scattering cross-section is under-estimated by Eq. (20), and the ratio of the total energy loss to scattering energy is overpredicted by δtot/δrad.

Figure 10. Comparison of αnonlin and αnum caused by coated bubbles at f = 1 MHz, pA = (a) 10 and (b) 20 kPa.

Figure 10. Comparison of αnonlin and αnum caused by coated bubbles at f = 1 MHz, pA = (a) 10 and (b) 20 kPa.

3.3. Attenuation in the transient state

Figure  compares αtran and αnum of the 20-cycle and Gaussian pulses through uncoated and coated bubbles at f = 1 MHz, pA = 10 Pa. Firstly, both αtran and αnum are smaller than αlin2 around resonance while larger than αlin2 far from resonance. The reason for the former is that the amplitude of R gradually increases for the oscillation around resonance. As shown in Figure , although the pressure amplitudes of cycles 4–10 have reached pA, their relative amplitudes |x| are smaller than that in the steady state, which reduces α. On the other hand, it takes much shorter time to reach the steady state for bubbles far from resonance, and some absorbed energy is converted to mechanical energy, which increases α. αnum predicted using cycles 14–17 agree much better with αlin2 (see Figure ), since these four cycles are close to the steady state. The Gaussian pulse only consists of several cycles as shown in Figure , larger deviations exist between αnum and αlin2, the relative difference was 45% at R0 = 3.72 µm. Secondly, αnum of the 20-cycle pulse does not match well with αtran, especially around resonance, and the relative difference was 4.2% at R0 = 3.72 µm. The reason is that the formula of αnum (Eq. (38)) is applicable for oscillations in the steady state. For transient oscillations, the sound intensities of these 20 cycles with and without bubbles can be directly obtained by integrations, then a new formula can be obtained: (48) αnumt=αexpt=10dlog100tpref(t)2dt0tpbub(t)2dt(48) It can be seen from Figure  that αnumt match well with αtran, and the relative difference decreased to 0.59% at R0 = 3.72 µm. Thus the experimental α of transient oscillations should also be predicted using Eq. (48) instead of Eq. (38). Figure  compares αtran and αnum of the 20-cycle pulse when pA is increased from 10 Pa to 20 kPa, αtran around resonance is still smaller than that in the steady state (αnonlin3), and αnumt still match well with αtran.

Figure 11. αtran and αnum of the 20-cycle and Gaussian pulses through (a) uncoated and (b) coated bubbles at f = 1 MHz, pA = 10 Pa. αnum is smaller than αlin2 around resonance while larger than αlin2 far from resonance, especially for the Gaussian pulse.

Figure 11. αtran and αnum of the 20-cycle and Gaussian pulses through (a) uncoated and (b) coated bubbles at f = 1 MHz, pA = 10 Pa. αnum is smaller than αlin2 around resonance while larger than αlin2 far from resonance, especially for the Gaussian pulse.

Figure 12. Evolution of x of the uncoated bubble with R0 = 3.72 µm at f = 1 MHz, pA = 10 Pa predicted using the ode solver in MATLAB.

Figure 12. Evolution of x of the uncoated bubble with R0 = 3.72 µm at f = 1 MHz, pA = 10 Pa predicted using the ode solver in MATLAB.

Figure 13. αtran and αnum of the 20-cycle pulse through (a) uncoated and (b) coated bubbles at f = 1 MHz, pA = 20 kPa.

Figure 13. αtran and αnum of the 20-cycle pulse through (a) uncoated and (b) coated bubbles at f = 1 MHz, pA = 20 kPa.

3.4. Attenuation and phase velocity at medium β

It has been shown that αnum match precisely with αtheo for linear, nonlinear, and transient oscillations at extremely small β. As described in Section 2.5.2, the present method was validated by compared with the results of the VOF method at medium β. Figure  compares αnum caused by 49 bubbles predicted by the VOF and present methods at f = 1 MHz, pA = 10 Pa, n = 1012 m−3 (β0 around 10−4). It can be seen that good agreements are obtained between the VOF method and the present method with the bounded KME, and the relative differences are all less than 1.6%. Obvious deviations are obtained when replacing the bounded KME with the KME (the maximum relative difference reaches 15.4%), and R0 corresponding to the peak of α decreases. Since the VOF method can precisely predict the dynamics of multibubbles, α was well predicted by the present method at β0 around 10−4.

Figure 14. Comparison of αnum caused by 49 bubbles predicted by the VOF and present methods at f = 1 MHz, pA = 10 Pa, n = 1012 m−3. The relative differences between the VOF method and the present method (with the bounded KME) are all less than 1.6%, while obvious deviations exist using the present method with the KME.

Figure 14. Comparison of αnum caused by 49 bubbles predicted by the VOF and present methods at f = 1 MHz, pA = 10 Pa, n = 1012 m−3. The relative differences between the VOF method and the present method (with the bounded KME) are all less than 1.6%, while obvious deviations exist using the present method with the KME.

After comparing with the VOF method, comparisons with the experimental and theoretical results were made. Figure  compares the experimental (Silberman, Citation1957), theoretical and numerical α and V through air bubbles of R0 = 0.994 and 1.02 mm at β0 = 3.77 × 10−4. It was found that αnum and Vnum predicted using the KME matched precisely with αm and Vm. As concluded in the previous paragraph, the bounded KME should be used in numerical simulations, thus we deduced the theoretical α and V based on the bounded KME (αmb and Vmb). The linearisation of the bounded KME lead to: (49) ω0b2=3κp0+2(3κ1)σ/R0ρLR02(1+M(R0)β01/3)+δtotω0ω2RcL(49) (50) δvis,b+δth,b=4(μL+μth)ρLω0R02(1+M(R0)β01/3)(50)

Figure 15. Comparisons of the experimental (Silberman, Citation1957), theoretical and numerical (a) α and (b) V caused by uncoated bubbles with R0 = 0.994 and 1.02 mm at β0 = 3.77 × 10−4. αnum predicted using the KME match precisely with αm, while those using the bounded KME match precisely with the proposed αmb.

Figure 15. Comparisons of the experimental (Silberman, Citation1957), theoretical and numerical (a) α and (b) V caused by uncoated bubbles with R0 = 0.994 and 1.02 mm at β0 = 3.77 × 10−4. αnum predicted using the KME match precisely with αm, while those using the bounded KME match precisely with the proposed αmb.

The corresponding wave number for the monodisperse bubble population was obtained as: (51) kmb2=ω2cL2+4πω2n(R0)R0ω0b2ω2+δtot,bω0bωi11+M(R0)β01/3(51) αmb and Vmb based on the bounded KME can be obtained by replacing km in Eqs. (33) and (34) with kmb. It can be seen from Figure  that αnum and Vnum predicted using the bounded KME match precisely with αmb and Vmb, the relative differences of α were all less than 1%. It can be concluded that the bounded KME should be used in both numerical and theoretical predictions of α and V. This conclusion is consistent with the actual situation that the acoustic excitation is not at infinity, but near bubbles. f corresponding to the peak of αlin2, αm and αmb are respectively named as: flin2, fm and fmb. fm is 1.19% larger than flin2, while 3.74% smaller than fmb since ω0 is 3.85% smaller than ω0b, which is consistent with the results in the previous paragraph: R0 corresponding to the α peak is smaller using the KME. If the inter-bubble interaction is considered using Eq. (36), ω0 will decrease but not increase. The difference between αm and αmb is negligible far below resonance, so is the difference between Vm and Vmb.

Figure  and Figure  show the influences of β0 on α/n respectively caused by uncoated and coated microbubbles at pA = 10 Pa, f = 1 and 3 MHz. R0 corresponding to the peak of αm increases with β0, and it further increases when using the bounded KME (αmb). αnum predicted using the bounded KME match precisely with αmb. Figure  compares Vm, Vmb and Vnum caused by uncoated and coated microbubbles at β0 = 10−5, f = 1 MHz, pA = 10 Pa. R0 corresponding to the peak of Vmb is larger than that of Vm due to the increase of resonance frequency. Vnum predicted using the bounded KME match precisely with Vmb.

Figure 16. Influences of β0 on α/n caused by uncoated microbubbles at pA = 10 Pa, f = (a) 1 MHz and (b) 3 MHz. αnum predicted using the bounded KME match precisely with αmb.

Figure 16. Influences of β0 on α/n caused by uncoated microbubbles at pA = 10 Pa, f = (a) 1 MHz and (b) 3 MHz. αnum predicted using the bounded KME match precisely with αmb.

Figure 17. Influences of β0 on α/n caused by coated bubbles at pA = 10 Pa, f = (a) 1 MHz and (b) 3 MHz.

Figure 17. Influences of β0 on α/n caused by coated bubbles at pA = 10 Pa, f = (a) 1 MHz and (b) 3 MHz.

Figure 18. Comparisons of Vm, Vmb and Vnum caused by (a) uncoated and (b) coated microbubbles at β0 = 10−5, f = 1 MHz, pA = 10 Pa. Vnum predicted using the bounded KME match precisely with Vmb.

Figure 18. Comparisons of Vm, Vmb and Vnum caused by (a) uncoated and (b) coated microbubbles at β0 = 10−5, f = 1 MHz, pA = 10 Pa. Vnum predicted using the bounded KME match precisely with Vmb.

Figure 19. Influence of d on αnum caused by uncoated bubbles at β0 = 2 × 10−6, f = 1 MHz, pA = 10 Pa.

Figure 19. Influence of d on αnum caused by uncoated bubbles at β0 = 2 × 10−6, f = 1 MHz, pA = 10 Pa.

Figure 20. Influences of d on αnum caused by (a) uncoated and (b) coated bubbles at β0 = 2 × 10−6, f = 1 MHz, pA = 20 kPa.

Figure 20. Influences of d on αnum caused by (a) uncoated and (b) coated bubbles at β0 = 2 × 10−6, f = 1 MHz, pA = 20 kPa.

The advantage of the present CFD method is to deal with the complex conditions combining the nonlinearity, transience, medium β, focused wave and so on, which are difficult to handle by theoretical analyzes. Figure  shows the influence of d on αnum at β0 = 2 × 10−6, f = 1 MHz, pA = 10 Pa. It can be seen that the influence of d on the linear α is negligible. For nonlinear oscillations, α is dependent on pA, therefore, it is also dependent on d at medium β. Figure  shows the influences of d on αnum at β0 = 2 × 10−6, f = 1 MHz, pA = 20 kPa. R0 corresponding to the α peak increases obviously with the increasing d, since pA decreases during the propagation. Especially for uncoated bubbles (Figure a), αnum becomes closer to αlin2 with the increasing d. Figure  shows αnum and αnumt of the 20-cycle pulse at β0 = 10−5, and those at small β0 are repeated from Figure a as references. The transient αnum is 12.5% smaller than αlin2 around resonance (R0 = 3.72 µm) at small β0, while it is only about half of αmb around resonance (R0 = 3.8 µm) at β0 = 10−5. And the differences between αnum and αnumt become greater at β0 = 10−5.

Figure 21. αnum and αnumt of the 20-cycle pulse at β0 = 10−5, pA = 10 Pa. The differences between the steady (αmb) and transient α at medium β are much greater than those at small β. The differences between αnum and αnumt also increase.

Figure 21. αnum and αnumt of the 20-cycle pulse at β0 = 10−5, pA = 10 Pa. The differences between the steady (αmb) and transient α at medium β are much greater than those at small β. The differences between αnum and αnumt also increase.

4. Conclusions and prospects

Numerical measurements of α and V were implemented using a homogeneous cavitation model, which were modelled from experimental schemes, and comparisons were made with αtheo. At extremely small β, αnum match precisely with αtheo for linear, nonlinear, and transient oscillations. For linear oscillations, αlin2 was obtained, which calculated the total energy loss by integrating the work of all damping forces, instead of by multiplying the scattering energy by δtot/δrad. For nonlinear oscillations, both the linearised σs (Eq. (20)) and δtot/δrad are dependent on pA. For transient oscillations, the total energy loss should be obtained by integrating the work of pa(t) when calculating the theoretical α (αtran) to include the mechanical energy, and a new formula was proposed for the calculations of the numerical (αnumt) and experimental (αexpt) α by integrating the sound intensity of recorded pressure.

At medium β, the relative difference of αnum between the VOF and present methods was less than 1.6%, while it reached 15.4% after replacing the bounded KME in the present method with the KME. And simulations towards millimetre bubbles showed that αnum and Vnum predicted using the KME matched precisely with the traditional theoretical results (αm and Vm). Then new theoretical αmb and Vmb were proposed based on the bounded KME, and the relative differences between αmb and αnum (based on the bounded KME) were less than 1% at β0 = 3.77 × 10−4. Therefore, the bounded KME should be used in both numerical and theoretical predictions, and the resonance frequency increases with β (Eq. (49)). In addition, the influences of nonlinearity and transience on α at medium β are quite different from that at small β, and the present method can be applied to investigate complex acoustic cavitation where theoretical analyzes are difficult to apply.

The bubble radius is usually distributed within a range. This study focused on α caused by equal-sized bubbles. The inter-bubble interaction between different-sized bubbles is more complex and needs to be considered by using the grouping cavitation model in the future.

Disclosure statement

No potential conflict of interest was reported by the author(s).

Data availability statement

The data that support the findings of this study are available within the article.

Additional information

Funding

Financial support from the Zhejiang Provincial Natural Science Foundation of China (No. LQ21E060003, LY23E060001, LY22E080006) and Fundamental Research Funds of Zhejiang University of Science and Technology (No. 2021QN029).

References

  • Alishiri, M., Ebrahimi, S., Shamloo, A., Boroumand, A., & Mofrad, M. R. K. (2021). Drug delivery and adhesion of magnetic nanoparticles coated nanoliposomes and microbubbles to atherosclerotic plaques under magnetic and ultrasound fields. Engineering Applications of Computational Fluid Mechanics, 15(1), 1703–1725. https://doi.org/10.1080/19942060.2021.1989042
  • Allen, J. S., May, D. J., & Ferrara, K. W. (2002). Dynamics of therapeutic ultrasound contrast agents. Ultrasound in Medicine & Biology, 28(6), 805–816. https://doi.org/10.1016/S0301-5629(02)00522-7
  • Brenner, M. P., Hilgenfeldt, S., & Lohse, D. (2002). Single-bubble sonoluminescence. Reviews of Modern Physics, 74(2), 425–484. https://doi.org/10.1103/RevModPhys.74.425
  • Chen, Y.-h., Zhan, J.-m., & Li, Y.-t. (2021). Numerical simulation of cavitation-bubble expansion and collapse inside a bottle subjected to impact on its topside. Engineering Applications of Computational Fluid Mechanics, 15(1), 1440–1451. https://doi.org/10.1080/19942060.2021.1976279
  • Clark, A., Bonilla, S., Suo, D., Shapira, Y., & Averkiou, M. (2021). Microbubble-Enhanced heating: Exploring the effect of microbubble concentration and pressure amplitude on high-intensity focused ultrasound treatments. Ultrasound in Medicine & Biology, 47(8), 2296–2309. https://doi.org/10.1016/j.ultrasmedbio.2021.03.035
  • Commander, K. W., & Prosperetti, A. (1989). Linear pressure waves in bubbly liquids: Comparison between theory and experiments. The Journal of the Acoustical Society of America, 85(2), 732–746. https://doi.org/10.1121/1.397599
  • Cowley, J., & McGinty, S. (2019). A mathematical model of sonoporation using a liquid-crystalline shelled microbubble. Ultrasonics, 96, 214–219. https://doi.org/10.1016/j.ultras.2019.01.004
  • de Jong, N., Cornet, R., & Lancée, C. T. (1994). Higher harmonics of vibrating gas-filled microspheres. Part one: Simulations. Ultrasonics, 32(6), 447–453. https://doi.org/10.1016/0041-624X(94)90064-7
  • Frinking, P., Segers, T., Luan, Y., & Tranquart, F. (2020). Three decades of ultrasound contrast agents: A review of the past, present and future improvements. Ultrasound in Medicine & Biology, 46(4), 892–908. https://doi.org/10.1016/j.ultrasmedbio.2019.12.008
  • Fuster, D. (2019). A review of models for bubble clusters in cavitating flows. Flow, Turbulence and Combustion, 102(3), 497–536. https://doi.org/10.1007/s10494-018-9993-4
  • Gorce, J. M., Arditi, M., & Schneider, M. (2000). Influence of bubble size distribution on the echogenicity of ultrasound contrast agents. Investigative Radiology, 35(11), 661–671. https://doi.org/10.1097/00004424-200011000-00003
  • Gümmer, J., Schenke, S., & Denner, F. (2021). Modelling lipid-coated microbubbles in focused ultrasound applications at subresonance frequencies. Ultrasound in Medicine & Biology, 47(10), 2958–2979. https://doi.org/10.1016/j.ultrasmedbio.2021.06.012
  • Helfield, B. (2019). A review of phospholipid encapsulated ultrasound contrast agent microbubble physics. Ultrasound in Medicine & Biology, 45(2), 282–300. https://doi.org/10.1016/j.ultrasmedbio.2018.09.020
  • Helfield, B., Zou, Y., & Matsuura, N. (2021). Acoustically-stimulated nanobubbles: Opportunities in medical ultrasound imaging and therapy. Frontiers in Physics, 9, 654374. https://doi.org/10.3389/fphy.2021.654374
  • Helfield, B. L., & Goertz, D. E. (2013). Nonlinear resonance behavior and linear shell estimates for Definity™ and MicroMarker™ assessed with acoustic microbubble spectroscopy. The Journal of the Acoustical Society of America, 133(2), 1158–1168. https://doi.org/10.1121/1.4774379
  • Ida, M., Naoe, T., & Futakawa, M. (2007). Suppression of cavitation inception by gas bubble injection: A numerical study focusing on bubble-bubble interaction. Physical Review E, 76(4), 046309. https://doi.org/10.1103/PhysRevE.76.046309
  • Jamshidi, R., & Brenner, G. (2013). Dissipation of ultrasonic wave propagation in bubbly liquids considering the effect of compressibility to the first order of acoustical Mach number. Ultrasonics, 53(4), 842–848. https://doi.org/10.1016/j.ultras.2012.12.004
  • Keller, J. B., & Miksis, M. (1980). Bubble oscillations of large amplitude. The Journal of the Acoustical Society of America, 68(2), 628–633. https://doi.org/10.1121/1.384720
  • Lauer, E., Hu, X. Y., Hickel, S., & Adams, N. A. (2012). Numerical modelling and investigation of symmetric and asymmetric cavitation bubble dynamics. Computers & Fluids, 69(11), 1–19. https://doi.org/10.1016/j.compfluid.2012.07.020
  • Lee, J. W., Heo, H., Sohn, D. K., & Ko, H. S. (2021). Development of a numerical method for multiphase flows using an electrostatic model in a wire-mesh sensor. Engineering Applications of Computational Fluid Mechanics, 15(1), 344–362. https://doi.org/10.1080/19942060.2021.1876775
  • Leighton, T. G. (1994). The acoustic bubble. Academic Press.
  • Leroy, V., Strybulevych, A., Scanlon, M. G., & Page, J. H. (2009). Transmission of ultrasound through a single layer of bubbles. The European Physical Journal E, 29(1), 123–130. https://doi.org/10.1140/epje/i2009-10457-y
  • Li, H., Yang, Y., Zhang, M., Yin, L., Tu, J., Guo, X., & Zhang, D. (2018). Acoustic characterization and enhanced ultrasound imaging of long-circulating lipid-coated microbubbles. Journal of Ultrasound in Medicine, 37(5), 1243–1256. https://doi.org/10.1002/jum.14470
  • Li, Q., Ming, D. Z., Lei, M., Guo, X., Liu, J. L., Zhu, H. W., Fang, L., & Wang, Z. B. (2022). Numerical investigation on the coupled mechanisms of bubble breakup in a venturi-type bubble generator. Engineering Applications of Computational Fluid Mechanics, 16(1), 229–247. https://doi.org/10.1080/19942060.2021.2008501
  • Marmottant, P., Sander, V. D. M., Emmer, M., Versluis, M., De Jong, N., Hilgenfeldt, S., & Lohse, D. (2005). A model for large amplitude oscillations of coated bubbles accounting for buckling and rupture. The Journal of the Acoustical Society of America, 118(6), 3499–3505. https://doi.org/10.1121/1.2109427
  • Modarreszadeh, A., Timofeev, E., Merlen, A., & Pernod, P. (2021). Numerical simulation of the interaction of wave phase conjugation with bubble clouds. International Journal of Multiphase Flow, 141, 103638. https://doi.org/10.1016/j.ijmultiphaseflow.2021.103638
  • Nederhoed, J. H., Tjaberinga, M., Otten, R. H. J., Evers, J. M., Musters, R. J. P., Wisselink, W., & Yeung, K. K. (2021). Therapeutic use of microbubbles and ultrasound in acute peripheral arterial thrombosis: A systematic review. Ultrasound in Medicine & Biology, 47(10), 2821–2838. https://doi.org/10.1016/j.ultrasmedbio.2021.06.001
  • Omoteso, K. A., Roy-Layinde, T. O., Laoye, J. A., Vincent, U. E., & McClintock, P. V. E. (2021). Acoustic vibrational resonance in a Rayleigh-Plesset bubble oscillator. Ultrasonics Sonochemistry, 70, 105346. https://doi.org/10.1016/j.ultsonch.2020.105346
  • Paul, S., Katiyar, A., Sarkar, K., Chatterjee, D., Shi, W. T., & Forsberg, F. (2010). Material characterization of the encapsulation of an ultrasound contrast microbubble and its subharmonic response: Strain-softening interfacial elasticity model. The Journal of the Acoustical Society of America, 127(6), 3846–3857. https://doi.org/10.1121/1.3418685
  • Plesset, M. S., & Prosperetti, A. (1977). Bubble dynamics and cavitation. Annual Review of Fluid Mechanics, 9(1), 145–185. https://doi.org/10.1146/annurev.fl.09.010177.001045
  • Prosperetti, A., Crum, L. A., & Commander, K. W. (1988). Nonlinear bubble dynamics. The Journal of the Acoustical Society of America, 83(2), 502–514. https://doi.org/10.1121/1.396145
  • Segers, T., de Jong, N., & Versluis, M. (2016). Uniform scattering and attenuation of acoustically sorted ultrasound contrast agents: Modeling and experiments. The Journal of the Acoustical Society of America, 140(4), 2506–2517. https://doi.org/10.1121/1.4964270
  • Segers, T., Gaud, E., Versluis, M., & Frinking, P. (2018). High-precision acoustic measurements of the nonlinear dilatational elasticity of phospholipid coated monodisperse microbubbles. Soft Matter, 14(47), 9550–9561. https://doi.org/10.1039/C8SM00918J
  • Si, Q., Ali, A., Liao, M., Yuan, J., Gu, Y., Yuan, S., & Bois, G. (2023). Assessment of cavitation noise in a centrifugal pump using acoustic finite element method and spherical cavity radiation theory. Engineering Applications of Computational Fluid Mechanics, 17(1), 2173302. https://doi.org/10.1080/19942060.2023.2173302
  • Silberman, E. (1957). Sound velocity and attenuation in bubbly mixtures measured in standing wave tubes. The Journal of the Acoustical Society of America, 29(8), 925–933. https://doi.org/10.1121/1.1909101
  • Stride, E., Segers, T., Lajoinie, G., Cherkaoui, S., Bettinger, T., Versluis, M., & Borden, M. (2020). Microbubble agents: New directions. Ultrasound in Medicine & Biology, 46(6), 1326–1343. https://doi.org/10.1016/j.ultrasmedbio.2020.01.027
  • Subroto, T., Lebon, G. S. B., Eskin, D. G., Skalicky, I., Roberts, D., Tzanakis, I., & Pericleous, K. (2021). Numerical modelling and experimental validation of the effect of ultrasonic melt treatment in a direct-chill cast AA6008 alloy billet. Journal of Materials Research and Technology, 12, 1582–1596. https://doi.org/10.1016/j.jmrt.2021.03.061
  • Suo, X. Y., Jiang, Y., & Wang, W. J. (2021). Hydraulic axial plunger pump: Gaseous and vaporous cavitation characteristics and optimization method. Engineering Applications of Computational Fluid Mechanics, 15(1), 712–726. https://doi.org/10.1080/19942060.2021.1913232
  • Trujillo, F. J. (2018). A strict formulation of a nonlinear Helmholtz equation for the propagation of sound in bubbly liquids. Part I: Theory and validation at low acoustic pressure amplitudes. Ultrasonics Sonochemistry, 47, 75–98. https://doi.org/10.1016/j.ultsonch.2018.04.014
  • Tu, J., Guan, J., Qiu, Y., & Matula, T. J. (2009). Estimating the shell parameters of SonoVue microbubbles using light scattering. The Journal of the Acoustical Society of America, 126(6), 2954–2962. https://doi.org/10.1121/1.3242346
  • Tu, J., Swalwell, J. E., Giraud, D., Cui, W., Chen, W., & Matula, T. J. (2011). Microbubble sizing and shell characterization using flow cytometry. IEEE Transactions on Ultrasonics, Ferroelectrics and Frequency Control, 58(5), 955–963. https://doi.org/10.1109/TUFFC.2011.1896
  • Versluis, M., Stride, E., Lajoinie, G., Dollet, B., & Segers, T. (2020). Ultrasound contrast agent modeling: A review. Ultrasound in Medicine & Biology, 46(9), 2117–2144. https://doi.org/10.1016/j.ultrasmedbio.2020.04.014
  • Wijngaarden, L. V. (1972). One-dimensional flow of liquids containing small Gas bubbles. Annual Review of Fluid Mechanics, 4(1), 369–396. https://doi.org/10.1146/annurev.fl.04.010172.002101
  • Yamamoto, T., & Komarov, S. V. (2021). Enhancement of oscillation amplitude of cavitation bubble due to acoustic wake effect in multibubble environment. Ultrasonics Sonochemistry, 78, 105734. https://doi.org/10.1016/j.ultsonch.2021.105734
  • Yang, Y., Li, Q., Guo, X., Tu, J., & Zhang, D. (2020). Mechanisms underlying sonoporation: Interaction between microbubbles and cells. Ultrasonics Sonochemistry, 67, 105096. https://doi.org/10.1016/j.ultsonch.2020.105096
  • Yasui, K. (2021). Numerical simulations for sonochemistry. Ultrasonics Sonochemistry, 78, 105728. https://doi.org/10.1016/j.ultsonch.2021.105728
  • Yasui, K., Iida, Y., Tuziuti, T., Kozuka, T., & Towata, A. (2008). Strongly interacting bubbles under an ultrasonic horn. Physical Review E, 77(1), 016609. https://doi.org/10.1103/PhysRevE.77.016609
  • Yasui, K., Lee, J., Tuziuti, T., Towata, A., Kozuka, T., & Lida, Y. (2009). Influence of the bubble-bubble interaction on destruction of encapsulated microbubbles under ultrasound. The Journal of the Acoustical Society of America, 126(3), 973–982. https://doi.org/10.1121/1.3179677
  • Ye, Y., Dong, C., Zhang, Z., & Liang, Y. (2020a). Considering the diffusive effects of cavitation in a homogeneous mixture model. Processes, 8(6), 662. https://doi.org/10.3390/pr8060662
  • Ye, Y., Dong, C., Zhang, Z., & Liang, Y. (2020b). Modeling acoustic cavitation in homogeneous mixture framework. International Journal of Multiphase Flow, 122, 103142. https://doi.org/10.1016/j.ijmultiphaseflow.2019.103142
  • Ye, Y., Liang, Y., Dong, C., Bu, Z., Li, G., & Zheng, Y. (2021a). Numerical modeling of ultrasonic cavitation by dividing coated microbubbles into groups. Ultrasonics Sonochemistry, 78(6), 105736. https://doi.org/10.1016/j.ultsonch.2021.105736
  • Ye, Y., Liang, Y., Dong, C., Xu, Y., & Zhang, Z. (2021b). Treating the phase change of cavitation as the source of vapor inside bubbles. Modern Physics Letters B, 35(5), 2150093. https://doi.org/10.1142/S0217984921500937
  • Yusefi, H., & Helfield, B. (2022). Ultrasound contrast imaging: Fundamentals and emerging technology. Frontiers in Physics, 10, 791145. https://doi.org/10.3389/fphy.2022.791145
  • Zhang, Y. (2013). A generalized equation for scattering cross section of spherical Gas bubbles oscillating in liquids under acoustic excitation. Journal of Fluids Engineering, 135(9), 091301. https://doi.org/10.1115/1.4024128