Fundamentals of Acoustic Cavitation in Sonochemistry: Jia Luo, Zhen Fang, Richard L. Smith, Jr. and Xinhua Qi
Fundamentals of Acoustic Cavitation in Sonochemistry: Jia Luo, Zhen Fang, Richard L. Smith, Jr. and Xinhua Qi
Ultrasound is a longitudinal pressure wave that has a wave frequency higher than
normal human hearing range. In acoustic science, the classification of a sound
wave is according to its oscillating frequency. Waves that have oscillating fre-
quencies from 0.001 to 10 Hz are referred to as infrasonic sound, and waves with
oscillating frequencies from 16 kHz and higher are referred to as ultrasonic sound
or ultrasound [1]. Table 1.1 lists some key parameters including the wavelength,
speed, frequency, amplitude, power, pressure, intensity of acoustic wave, acous-
tic characteristic impedance of medium material and their typical values in sono-
chemistry [1].
Compared with audible sound and other wave phenomena, ultrasound causes
many special physical or chemical effects in materials. Because of the high oscil-
lating frequency of liquid molecules in an ultrasonic field and the high viscosity
of the liquid medium compared with a gas, molecules absorb a large number of
energy from the propagating ultrasonic wave. As shown in Table 1.1 for the acous-
tic pressure, liquid reactants in ultrasonic field can have a high pressure differ-
ence relative to the hydrostatic pressure of fluid [2]. The high energy absorption
and high acoustic pressure make it possible for ultrasonic auxiliary to overcome
intermolecular interactions in a solvent, and produce numerous cavities, which
is referred to as acoustic cavitation. The formed cavities drain and accumulate
ultrasonic energy, and explosively release their energy by the collapse of cavi-
ties. Therefore, there are two main routes that ultrasound provides energy to liquid
reactants: (i) ultrasonic microstreaming (mechanical force) and (ii) acoustic cavi-
tation. However, only acoustic cavitation is believed to activate chemical reactions
with high energy intensity.
As shown in Table 1.1, ultrasound is used in power applications at low fre-
quencies of 20–40 kHz and at high acoustic power (>50 W), or it can be used in
nondestructive detection or measurement at frequencies above 1 MHz and low
acoustic power as commonly used in diagnostics, clinical medicine and biological
processing.
In this chapter, the fundamentals in ultrasound and acoustic cavitation are intro-
duced. The dynamics and influencing factors of cavitation bubbles are summa-
rized. The secondary effects of acoustic cavitation are introduced and discussed.
The knowledge may be useful in studying the intensification mechanism of sono-
chemical reaction of biomass materials. The authors would like to emphasize the
relationship between energy intensity and energy efficiency with sonication. The
ability of cavitating bubbles to focus and concentrate energy, forces and stresses is
the basic of phenomena for ultrasonic auxiliary of biomass-related reactions [3],
while the development of bioenergy requires attentions to special energy intensity
that could activate chemical species and possibly lower energy cost.
1 Fundamentals of Acoustic Cavitation in Sonochemistry 5
Table 1.1 Key parameters in acoustic science and typical parameter values for studies in sono-
chemistry [1]
Parameters (symbol, unit) Physical definition Typical values
Acoustic wavelength (λ, m) Distance between two 3.0–7.5 cm for power ultrasound
neighboring points that have in liquids
the same phase position in an 0.015–0.075 cm for measurement
acoustic wave
Acoustic speed (c, m/s) Acoustic propagation distance Air, 340 m/s
per unit time Liquid, 1,000–2,000 m/s
Solids, 3,000–6,000 m/s
Acoustic frequency (f, Hz) Vibration number per unit 20–40 kHz for power ultrasound
time in liquids
1–20 MHz for measurement
Acoustic amplitude (PA, m) Maximum distance of mass
point at the ultrasonic field
from its balance position
Acoustic power (P, W) Acoustic energy that passes Extremely high power (>1,000 W)
through one surface perpen- for cavitational erosion of solids
dicular to the propagating and metal working
direction per unit time
Optimum acoustic power High power (50–1,000 W) for
depends on sample, reactor ultrasound-assisted conventional
volume and processing needs thermochemical and biochemical
reactions
Low power (1–10 W) for
stimulating biological cells or for
promoting particle aggregation
with low cavitation
Acoustic pressure (Pa, Pa) Difference of dynamic 0.1–1 MPa for ultrasonic cleaner
force per area from its
static value, as the result
of the compressed zone
and rarefacted zone of fluid
formed with sound transport
Pa = PA sin(2πft), where t
refers to the propagating time
Acoustic intensity (I, W/m2) Average acoustic energy that
passes through a unit surface
perpendicular to the propagat-
ing direction per unit time
Vector I = P2A /(2ρc), where ρ
is the medium density
Acoustic characteristic Ratio of acoustic pressure to Gases, 1.1–5.5 × 102 kg/(m2 s)
impedance (Z, kg/(m2 s)) acoustic velocity at one mass Liquids, 0.7–3.2 × 106 kg/(m2 s)
point in the medium Nonmetals, 0.1–1.5 × 107 kg/
(m2 s)
Characteristic of the Metals, 1.0–10.4 × 107 kg/(m2 s)
transmitting medium. Z = ρc,
where ρ is the medium density
6 J. Luo et al.
1.2 Acoustic Cavitation
The energy content of ultrasonic microstreaming is relatively low due to the small
amplitude of the oscillation of fluid elements and it is generally insufficient to
induce chemical reaction in the medium. The wavenumber of ultrasound in liq-
uid is much larger than ultraviolet light, and the excitation of electrons in com-
pound molecules is not possible with sonication. Sonication does not result in the
resonance of chemical molecules or atoms like other methods such as microwave
or infrared radiation, because the oscillating frequency of ultrasound is on the
order of 104–105 Hz. As a result, acoustic cavitation, which could concentrate and
explosively release acoustic energy, is believed to be the dominant mechanism for
process intensification and chemical activation in sonochemical systems.
The complete process of acoustic cavitation consists of three steps [4, 5]:
1. Formation of cavitation nuclei. Figure 1.1 shows bubble formation, growth and
collapse phenomena for transient and steady cavitation. Negative and posi-
tive acoustic pressure act alternately on one point in the liquid medium in an
ultrasonic cycle with the propagation of ultrasound wave (Fig. 1.1). The nega-
tive acoustic pressure stretches the liquid medium apart to make a relatively
rarefacted zone. If the acoustic pressure is increased to a value that is higher
than certain intensity, the microscopic distance between the liquid molecules
becomes far enough to initiate cavity formation that contains solvent vapors or
dissolved gases. The minimum requirement for acoustic pressure to form a cav-
ity is called the Blake threshold, PB. If the vapor pressure in a bubble can be
ignored, then Eq. 1.1 results [5]:
1/ 2
(2σ/R0 )3
2 (1.1)
PB = P0 +
3 3(P0 + 2σ/R0 )
where P0 is the static pressure on the liquid, R0 is the initial radius of the
formed bubble, and 2σ/R0 is the surface tension of the assumed bubble when
it forms. Low surface tension and low static pressure on a liquid favors the for-
mation and growth of cavitation nuclei. The existence of impurities and hetero-
geneous crevices in liquids decrease the actual pressure threshold of cavitation
by reducing the surface tension. The Blake threshold value for untreated sol-
vents is about 1–10 % of that in ultrapure solvents [4]. The influence of impuri-
ties and heterogeneous crevices on nucleation is kinetic. Similarly to a catalyst,
the presence of surface defects lowers the free energy barriers separating the
metastable liquid state from the vapor phase, and modifies the nucleation
1 Fundamentals of Acoustic Cavitation in Sonochemistry 7
Fig. 1.1 Calculation results of bubble dynamics: a transient cavitation and b steady cavitation:
acoustic pressure, bubble temperature, microturbulence velocity, bubble radius, bubble pressure,
and shockwave amplitude (adapted with permission from [8], Copyright © 2012 Elsevier)
mechanism [6]. As a result, the nucleation rate on surface defects with simple
geometries can be increased by up to five orders of magnitude compared with
that of a flat hydrophobic surface [6]. The presence of surface defects has also
been found to affect the nucleation rate for non-condensable gas as dissolved in
liquids [6]. For bubble nucleation in solutions containing solid substances, the
contact angle of gas bubbles on the solid surface is a critical index to evaluate
the possibility of bubble nucleation. A small contact angle means a low energy
barrier for the nucleation. In recent research by Zhang et al. [7], it was pointed
out that the contact angle of gas bubbles on spherical surfaces could be remark-
ably reduced by several methods such as decreasing the surface tension of the
solution, by reducing the size of solid particles, or by changing the hydrophilic
surface of solid particles to that of a hydrophobic surface.
2. Growth (radial motion) of cavitation bubbles. The formed microbubble contin-
ually grows to a maximum bubble radius of about 2–150 μm until the end of
the negative acoustic pressure phase. The oscillating motion of cavitation bub-
ble can be described with the Rayleigh-Plesset equation [5]:
d 2 R 3 dR 2
3γ
1 2σ R0 2σ 4µ dR
R 2 + = P0 + − − − P∞ (1.2)
dt 2 dt ρ R0 R R R dt
8 J. Luo et al.
where R is the radius of bubble in motion, dR/dt is the velocity item of bubble
wall away from the bubble center, R0 is the initial radius of bubble, ρ and μ
are the density and viscosity of the liquid, γ refers to the specific heat ratio of
gas in bubble and the P0 and P∞ are the static pressure nearby bubble and at
infinite distance in the liquid. The terms on the left-hand side of the equation
represent the radial motion (expansion or compression) of the bubble wall. On
the right-hand side of the equation, the first term shows the pressure variation
in the bubble, while the second and third terms are the surface tension and the
viscous stresses at the bubble surface, respectively. Equation 1.2 is based on
the assumption that the motion of a bubble is adiabatic and that the liquid is
incompressible [4]. The mass and heat transfer between the bubble and the sur-
roundings is not considered in Eq. 1.2. It is assumed that no chemical reaction
or physical changes occur and that there are no temperature, density or pressure
gradients in cavitating bubble.
3. Collapse or the next oscillation of bubbles. Positive acoustic pressure phase
comes after the negative pressure phase (Fig. 1.1), and under its influence, the
cavitation bubble undergoes contraction. The time required for the bubble to
contract is about 25–0.5 μs, which is half of the reciprocal of ultrasound fre-
quency. According to the literature [9], the oscillating frequency of a bubble in
a liquid, fb, increases with contraction of the bubble radius, R:
1/
2
1 3γ 2σ (1.3)
fb = P0 +
2πR ρ R
where ρ is liquid density, γ and 2σ/R are the gas specific heat ratio and sur-
face tension in the bubble and P0 is the static pressure. If the resonant fre-
quency of a bubble, fb, is smaller than the frequency of ultrasonic field, fa, at
the end of the positive acoustic pressure phase, the bubble survives and turns
to growth in a new ultrasonic cycle. The radial motion of the cavitation bub-
ble can be repeated for several acoustic cycles, and is called as steady cavita-
tion (Fig. 1.1b). However, if fb ≥ fa, the bubbles will collapse quickly in several
nanoseconds, and this is called as transient cavitation (Fig. 1.1a). Fragments
generated during the collapse of the parent bubble may become new nuclei for
subsequent cavitation phenomena. Steady and transient cavitation are both pre-
sent in most sonochemical systems, although the properties of the liquid may
favor one type of cavitation.
During the transient collapse of a bubble, work is done by the liquid such that fluid
elements impart energy to the bubble, thus raising its temperature and pressure.
The energy intensity of the cavitation bubble collapse is proportional to the com-
pression ratio of the bubble at the point of minimum radius during radial motion.
At the instance of bubble collapse, energy is released that does not have time to
be transferred to the surroundings and therefore local hotspots are produced [10,
11]. These local hotspots have extremely high temperatures (ca. 5,000 °C) and
1 Fundamentals of Acoustic Cavitation in Sonochemistry 9
Fig. 1.2 Range of duration time, pressure and energy for various energy chemistries (reprinted
with permission from [11], Copyright © 2010 WILEY-VCH Verlag GmbH & Co. KGaA,
Weinheim)
pressures (ca. 100 MPa) in its interior, and causes very high rates of local heat-
ing and cooling (>109 °C/s). The energy density of acoustic cavitation can be on
the order of 1015 W/cm3, while the rated acoustic power density used in common
sonochemical reactors is 10−2–10 W/cm3 [12]. Acoustic cavitation also produces
high intensity shockwave and violent flowing of localized liquid.
In Fig. 1.2, the average energy of acoustic cavitation (about 0.1–1 eV) is much
higher than that in conventional heating, so that breaking most of chemical bonds
is possible. However, this energy is not enough for plasma ionization of volatile
molecules and photon production in bubbles. Suslick et al. [13, 14] observed very
strong Ar atomic excitation in the sonication of 95 wt% sulfuric acid solution at
20 kHz at ambient temperature under an Ar atmosphere. The emission temperature
of Ar excitation in bubble was estimated at about 8,000 °C. However, the thermal
excitation of Ar atom needed very high energy of >13 eV, and thus Ar atom could
not be thermally excited at 8,000 °C, which shows that the distribution of tempera-
ture and pressure in a bubble is not uniform. Figure 1.3 shows the physical struc-
ture of a bubble and the active zones for sonochemistry. Optically opaque plasma
probably exists at the core of a collapsing bubble at extreme conditions, and the
collision of higher energy electrons in this plasma zone is believed to result in the
excitation of an Ar atom. Spectral lines of Ar excitation are emitted just on the
outer zone of the plasma ionization (emissive shell) and can be detected by ultravi-
olet-visible spectroscopy [14]. The emissive shell is optically transparent and has a
temperature that is much lower than the plasma zone.
On the other hand, the energy intensity of bubbles is also influenced by the
mass and heat transfer of the surrounding liquid. It is believed that the vapor or gas
10 J. Luo et al.
Fig. 1.3 Physical structure of collapsing bubble and active zones for sonochemistry (adapted
with permission from [14], Copyright © 2011 Elsevier)
in the cavity can not escape into the surroundings before bubble collapse occurs.
However, the evaporation of liquid molecules at the interface between bubbles
and surroundings is possible and materials flowing into the bubbles can change
the vapor composition in a bubble. Dissociation and chemical reactions of volatile
molecules in the bubble can occur and this will consume the energy of the bubble
and reduce its energy release when it collapses. For example, sonication in alcohol
solution under Ar atmosphere at the frequencies of 20, 363 and 1,056 kHz gives
contradictory results [15]. Much less energy is released in a high frequency acous-
tic field of 1,056 kHz than at low frequencies, which is determined by the volume
change of bubbles when bubbles collapse. However, the calculated temperature of
the cavitation bubbles has the following order of high to low according to frequen-
cies of 1,056 > 363 > 20 kHz. It is highly possible that the evaporation of sol-
vent in the acoustic field of 20 kHz is more than that at 1,056 kHz, because of the
longer acoustic cycle and larger bubble size during the radial motion of bubbles
under sonication at 20 kHz.
Therefore, it is necessary to modify the Rayleigh-Plesset equation (Eq. 1.2),
because Eq. 1.2 deals with slow bubble motions, and many of the assumptions on
which Eq. 1.2 is based are invalid at the time of rapid collapse of bubbles [16].
Equation 1.2 does not take mass transfer, heat transfer, chemical reactions, non-
equilibrium phase changes and non-uniform pressures in the bubble interior into
account, so that modeling results have large deviations from their true values.
Boundary layer approximation on the basis of physical observations allows simpli-
fication of the theory.
1 Fundamentals of Acoustic Cavitation in Sonochemistry 11
For example, considering the compressibility of liquid and the acoustic attenu-
ation, the bubble motion can be described by the Keller-Miksis equation (Eq. 1.4)
[17, 18]:
dR/dt d 2 R 3 dR/dt dR 2
(1 − )R 2 + (1 − )
c dt 2 3c dt
1 dR/dt R dPb 2σ 4µ dR
= (1 + ) Pb − (Pam − PA sin 2πft) + − −
ρ c c dt R R dt
(1.4)
where R is the radius of bubble in motion, dR/dt is the velocity item of bubble
wall away from the bubble center, ρ is the liquid density, μ is the dynamic vis-
cosity of the liquid (μ = ρν, ν is kinematic viscosity), and the c, f and PA are the
speed, frequency and pressure amplitude of the sound wave in liquid, respectively.
In Eq. 1.4, 2σ/R is the surface tension of the bubble according to the liquid tem-
perature, Pam is the ambient pressure (1 atm), Pb is the pressure inside the bubble,
and its time derivative is given by a van der Waals type equation of state as Eq. 1.5
[17, 18]:
Ntot (t)kTb
Pg (t) =
(1.5)
3
4π
(R(t))3 − Req (t)/8.86
3
In Eq. 1.5, k is the Boltzmann constant, Ntot is the total number of vapor molecules
in the bubble due to the condensation and evaporation, Tb is the temperature of
bubble contents and Req is the equilibrium radius of the bubble.
Other important parameters for numerical simulation of bubble motion are
the mass diffusion and non-equilibrium phase change of water vapor during the
sonication in water. In 2000, Storey and Szeri [16, 19] proposed a two-step pro-
cess that consisted of the diffusion of vapor molecule to the bubble wall and the
condensation of vapor at the bubble wall, along with three important time scales,
namely, the time scale of bubble dynamics (tosc), the time scale of mass diffusion
(tdif) and the time scale of condensation (tcond) as follows:
R
tosc =
|dR/dt|
R2 1
tdif = ReSc M ≈ ReSc √ , Re = ρ0 R0 v0 /µ0 , Sc = µ0 /ρ0 Davg
DH2 O R Tb
(1.6)
R 2πMH2 O
tcond =
σ 9M0 T0
tosc ≫ tdif, tcond, so that bubble motion is slow enough to allow the completion of
For mass diffusion of water vapor, in the earlier phase of bubble collapse,
one cycle of mass transfer between liquid and bubble interior, which results in uni-
tosc ≪ tdif, the bubble motion is so rapid that the water vapor has insufficient time
form bubble composition [18, 19]. Then, with the acceleration of the bubble wall,
to diffuse to the bubble wall that results in nearly unchanged composition of vapor
equilibrium with respect to the bubble motion. If tosc ≪ tcond, the phase change
will be non-equilibrium, which results in the entrapment of water molecules in the
tosc ≪ tdif is reached well before tosc ≪ tcond during bubble collapse and so the
bubble [18, 19]. However, Storey and Szeri [19] demonstrated that the condition
slow mass diffusion is responsible for trapping water vapor in the bubble.
Actually, during bubble oscillation, the surface temperature of the bubble exceeds
the temperature of bulk water only for a very brief moment [17, 18]. Therefore, the
bubble is divided into two parts, a “cold” boundary layer in thermal equilibrium
with liquid and an eventually hot homogeneous core. Based on this assumption, the
instantaneous diffusive penetration depth ldif is taken to be Eq. 1.7 [17, 18]:
ldif = Dtosc (1.7)
where CH2 O and CH2 O,eq are the actual concentrations of water molecules in the
bubble core and the equilibrium concentration of water molecules at bubble wall,
respectively. The CH2 O,eq is calculated from the vapor pressure of water at bulk
temperature (Tl).
At the instant that the bubble stops growing, the velocity of the bubble wall is
zero, and the upper limit of the diffusion length can be calculated [17, 18]:
1 Fundamentals of Acoustic Cavitation in Sonochemistry 13
R
ldif = min( Dtosc , ) (1.9)
π
For diffusion of heat in analogy with diffusion of mass, in the earlier phases of
bubble collapse when the time scale of bubble oscillation is longer than that for
heat transfer, sufficient heat transfer helps to keep the bubble interior at ambient
liquid temperature, while during bubble collapse, the bubble motion is so rapid
that bubble collapse behaves nearly adiabatically with negligible outflow of bub-
ble energy [19]. Equation 1.10 provides a method to estimate the rate of heat loss
across a bubble wall during its motion [17, 18]:
dQ 2 T0 − T R Rκ
= 4πR tc , lth = min( , ),
dt lth π |dR/dt|
n (1.10)
tc
κ= , ρmix Cp, mix = ρi Cpi
ρmix Cp, mix
i=1
where λtc is the thermal conductivity of bubble contents and T0 is the temperature
of the cold bubble surface that is nearly the same with the temperature in the liq-
uid. In Eq. 1.10, lth is the thermal diffusion length, κ is the thermal diffusivity and
terms ρi and Cpi are the density and molecular specific heat of compound i, respec-
tively. The ρmix and Cp,mix in Eq. 1.10 are the density and molecular specific heat
of the vapor mixture in the bubble, respectively.
The overall energy balance for the bubble as an open system can be written as
Eq. 1.11 [18]:
dE dQ dW dNH2 O
= − + h H2 O (1.11)
dt dt dt dt
where dW/dt refers to the work done by the bubble. The enthalpy per water mol-
ecule that condenses at the cold bubble surface is given by hH2 O = 4kT0, where k
is the Boltzmann constant. Equation 1.11 can be rewritten as [17, 18]:
dT dQ dV dNH2 O
Cv,mix = − Pi + hH2 O − UH2 O
dt dt dt dt
∂E θi T
UH2 O = = NH2 O kT 3 +
∂NH2 O exp (θi /T ) − 1 (1.12)
n
Cv,mix = Cv,i Ni
i=1
where UH2 O is the internal energy, namely, the specific energy of water molecules
in the bubble, Pi is the pressure in the bubble interior, θi is the characteristic vibra-
tional temperature of species i, Ni is the number of molecules of component i and
Cv,i and Cv,mix are the specific heats of individual component i and the mixture,
respectively.
14 J. Luo et al.
Equation 1.12 can be used to estimate the temperature change of a bubble dur-
ing its motion, and it gives better results in the numerical simulation of the sono-
chemical phenomena [17].
In summary, cavitation behavior of a single bubble under a high intensity ultra-
sonic field has been discussed along with fundamental relationships. Information
on the formation, growth, collapse and energy change of single bubble as the
result of certain ultrasonic field parameters in a sonochemical reactor has been
highlighted. By solving related mathematical equations, the motion and energy
behavior of a single bubble can be estimated theoretically. By trapping a single
bubble of gas in partially degassed liquid by a standing acoustic wave and then
driving it into highly nonlinear oscillation, single bubble cavitation can be realized
and studied in the laboratory [20], which demonstrates that cavitation takes place
when the acoustic intensity is higher than the Blake threshold, and it provides
much higher energy intensity than conventional acoustic streaming. The energy
intensity of cavitation bubble depends on the compression ratio of bubble radius
and also the net heat capacity of the bubble content. However, the entrapment of
solvent vapor in the bubble at the moment of collapse consumes the heat capac-
ity of the bubble through evaporation, which lowers the temperature attained at
the moment of minimum radius during collapse. Moreover, dissociation of solvent
vapor mainly involves endothermic reactions that further consume energy and con-
tribute to the lowering of the peak temperature attained at collapse.
1.2.2 Multibubble Cavitation
In an actual sonochemical reaction, there are numerous active bubbles (ca. 104 bub-
bles/1 mL water) for sonication at 20 kHz [2]. Bubbles at different positions of
acoustic field follow different cavitation behavior and have different cavitation
energy due to spatial diversity of the acoustic pressure intensity. Bubble behavior
is influenced not only by the acoustic field, but also by the cavitation behavior of
neighboring bubbles. As a result, multibubble cavitation shows statistical results of
cavitation behavior and gives rise to the interactions between bubbles.
In multibubble cavitation, the intensity and efficiency of acoustic cavitation
is determined by the number density of effective bubbles in the liquid. Rectified
diffusion, bubble coalescence and concerted collapse influence the behavior of
bubbles.
Rectified diffusion. In acoustic cavitation, two possibilities exist for the newly
formed bubble nuclei—dissolution in the liquid phase if the ultrasonic intensity
is below a threshold value, or growth to a larger size under the action of nega-
tive acoustic pressure. If the size of the bubble expands to over a critical (reso-
nance) size, the bubbles will quickly collapse in the following positive acoustic
pressure stage, which initiates sonochemical reactions. However, several or tens of
acoustic cycles may be necessary for a significant number of bubbles to reach the
critical size (Fig. 1.4) [21]. Rectified diffusion is the main way to enlarge the sizes
1 Fundamentals of Acoustic Cavitation in Sonochemistry 15
Fig. 1.4 Bubble growth in surfactant solution under sonication: a bubble growth as the function
of time in water and sodium dodecyl sulfonate (SDS) solution at acoustic pressure of ~0.022 MPa
and frequency of ~22 kHz; b sonoluminescence (SL) intensity as the function of acoustic pulse
number in water under pulse ultrasonic field at 515 kHz; c SL intensity as the function of acous-
tic pulse number in 1 M aqueous methanol solution under pulse ultrasonic field at 515 kHz; d SL
intensity as the function of acoustic pulse number in 0.75 mM aqueous SDS solution under pulse
ultrasonic field at 515 kHz (adapted with permission from [21], Copyright © 2007 Elsevier)
hinders coalescence of small bubbles to larger ones. As a result, the growth of bub-
bles through bubble coalescence requires more sonication time or more acoustic
pulses in surfactant solutions. On the other hand, for high concentration of SDS,
the adsorption of SDS at the bubble/liquid interface is thought to restrict outflow
of chemical matters in the bubbles during bubble shrinkage. Therefore, the growth
of bubbles via rectified diffusion route is promoted and accelerated with the addi-
tion of surfactants.
Concerted collapse. In a bubble cloud, the collapse of a bubble emits strong
shockwaves. The produced shockwave elevates the ambient pressure or external
energy intensity on neighboring bubbles, and accelerates the collapse of neigh-
boring bubbles and other bubbles in the cloud. This assumption is referred to as
1 Fundamentals of Acoustic Cavitation in Sonochemistry 17
concerted collapse [2, 25]. Concerted collapse starts at the boundary of a bub-
ble cloud and proceeds until collapse of bubbles at the cloud center occurs. As
a result, shockwaves from multiple bubbles converge together and form a single
shockwave directing towards the solid surface with much higher shockwave inten-
sity (several hundreds of kPa) than the collapse of multiple single bubbles.
In a practical sonochemical reactor, the behavior of bubbles is influenced by
the pattern of ultrasonic streaming. Some sonochemical systems show interfer-
ence characteristics in ultrasonic wave propagating in the whole reactor space,
which means numerous alternate or fixed regions with strong or weak ultrasonic
intensity. An example is the common ultrasonic bath with special depths (integral
multiples of quarter-wavelength of ultrasound wave) from the bath bottom to the
liquid surface. For this type of sonochemical reactor, the collapse of bubbles are
generally concentrated at points that have high ultrasonic intensity (antinodes) in
the reactor space, which gives high cavitation activities as shown in Fig. 1.5 [26,
27]. Regions with long-term low or even zero ultrasonic intensity (nodes) may
favor the growth of bubbles via rectified diffusion and coalescence. At these nodes,
the size of the bubbles may increase beyond the critical size. The bubbles pos-
sibly survive even for long time after the ultrasound is shut off [21]. The behavior
of bubbles at antinodes or nodes is also used in distinguishing the intensification
mechanism of cavitation from that of ultrasonic streaming [28].
Because of the uncertainties in the spatial distribution of bubble size and bub-
ble density as well as the complex interaction between bubbles, the determina-
tion of cavitation in a multibubble system is more involved than that for a single
bubble system [15, 20]. Several effective methods have been developed for ana-
lyzing cavitation in multibubble systems [14, 15, 29, 30]. These methods are
summarized in Table 1.2. Acoustic frequency spectroscopy (AFS, Fig. 1.6) is
Horizontal plane
8 cm
Vertical direction from the
bottom of ultrasonic bath
6 cm
4 cm
2 cm
0 cm
Fig. 1.5 Cavitation erosion pattern on aluminum foil vertically positioned in ultrasonic bath
(water depth of 10 cm) at 20 kHz (adapted with permission from [27], Copyright © 2009
Elsevier)
18 J. Luo et al.
Table 1.2 Qualitative and quantitative approaches for measuring ultrasonic cavitation and
secondary effects
Method Sensor Principle of measurement Reference
Spectroscopic
Acoustic frequency Hydrophone Steady and transient bubbles [27, 31]
oscillating at subharmonics,
harmonics and ultraharmonics
frequencies produce variation of
local acoustic pressure
Sonoluminescence Light intensity Light intensity is detected and [14, 15, 29]
plotted as the function of scanning
wavelength from 300–900 nm
(ultraviolet-visible range)
Chemical
Iodine method ·OH radical H2O → ·OH + ·H, [30]
I− + ·OH → I2− → I3−
I3− concentration determined by
ultraviolet spectroscopy
Fricke method H2O → ·OH + ·H,
Fe2+ + ·OH → Fe3+ + OH−
Fe3+ concentration determined by
ultraviolet spectroscopy
Salicylic acid method Hydroxylation of salicylic acid by
·OH radical
Salicylic acid concentration deter-
mined by high performance liquid
chromatography
Terephthalic acid p-C6H4(COOH)2 + ·OH → C6H3
method OH(COOH)2
Hydroxyterephthalic acid concen-
tration determined by fluorescence
spectroscopy
Sonochemiluminescence 3-Aminophthalhydrazide (lumi-
nol) + ·OH → 3-aminophthalate,
in alkaline solution
Blue light emission of 3-ami-
nophthalate proportional to
concentration
Physical
Aluminum foil erosion Al foil Indentations on foil and weight [27, 30]
loss
Laser scattering Light Volumetric concentration of bub- [32, 33]
attenuation bles from Beer-Lambert law
Laser phase-Doppler Frequency Velocity of bubbles from [33, 34]
variation frequency variation between
scattering and incident waves
while bubble size by phase
displacement between scattering
and reflected waves
1 Fundamentals of Acoustic Cavitation in Sonochemistry 19
(d)
f
a direct but approximate way for measuring the occurrence of acoustic cavita-
tion. Hydrophones are placed in liquid under sonication, and record the inten-
sity of pressure pulse at local place as the function of scanning frequency. AFS
results reflect the change of acoustic parameters as the result of acoustic activa-
tion. Since it uses hydrophones for detection of sound waves, AFS needs com-
plex mathematical transformation and processing to obtain useful information
from the as-received spectra [27]. AFS spectrum with high intensity peaks at the
harmonic, higher harmonic, or even subharmonic frequencies proved the occur-
rence of cavitation [27, 31]. Chemical measurements are normally used in the
quantitative determination of cavitation yield and cavitation efficiency [12, 30].
In common horn- and bath-type reactors, the cavitation yields are determined as
3.5 × 10−9 and 5.8 × 10−7 g/J, respectively, which give relatively low energy effi-
ciency (5–20 %) of acoustic cavitation [12]. Because of less energy requirement,
the results from chemical measurement represent the less violent energy character-
istics in bubble dynamics than that from sonoluminescence [15]. Erosion analysis
of aluminum foil of several micrometer thickness is used to evaluate the cavitation
number (punching number on foil) and cavitation intensity (weight loss of foil).
20 J. Luo et al.
Results from chemical methods and erosion analysis describe the intensity of ·OH
radical and shockwave, secondary effects of cavitation, respectively.
Advanced technologies such as pulsed multibubble sonoluminescence (MBSL),
laser diffraction [32, 33], laser phase-Doppler [33, 34], methyl radical recombina-
tion and others are widely used in analyzing multibubble cavitation in recent years,
and these methods give the scale or distribution of energy, number, size, velocity
of bubble clusters. For MBSL, light emitted at certain wavelength relates to the
formation of special plasma. Light intensity is proportional to the energy intensity
of bubbles. By fitting the emission spectra of special plasma, the emission tem-
perature of plasma in bubbles can be estimated, which indicates the energy level of
bubble cavitation. However, the analysis of multibubble cavitation is still challeng-
ing. In fact, as the time for bubble collapse is short, cavitation bubbles analyzed by
various methods are mainly in growth or in oscillating, which does not apply to the
bubbles before collapse. For laser phase-Doppler, the number and volume factions
of bubbles are plotted against to the motion velocity and size of bubbles at certain
plane in acoustic field. Velocity of bubbles is determined by frequency variation
between scattering and incident waves, while bubble size is determined by phase
displacement between scattering and reflected waves. Laser scattering and laser
phase-Doppler are both direct methods to determine the cavitation number.
In multibubble cavitation, the production of cavitation bubbles means the for-
mation of another immiscible phase besides the liquid medium. The intensity of
ultrasonic energy is scattered and weakened for a heterogeneous liquid containing
high concentrations of insoluble solid particles or gas microbubbles. High acous-
tic intensity produces high number density of bubbles (bubble cloud) in the liq-
uid phase, while high number density of bubbles greatly influence the delivery of
acoustic intensity to the entire reactor space. These two opposing effects have con-
tradictory consequences [26]. For example, in a horn type sonochemical reactor, a
high density of bubble cloud gathers at the region near the tip of the transducer, and
results in inefficient cavitation in sonochemical processing [35]. Thus, the system
with a controlled concentration of cavitation nuclei gives much higher cavitation
intensity than the system with too many nuclei and the system suffering completely
denucleation [26]. Appropriate number of cavitation nuclei uses energy efficiently.
The formation and distribution of cavitating bubbles in the whole space of a
sonochemical reactor is not uniform and is greatly influenced by ultrasonic
streaming. Ultrasonic streaming occurs according to the geometry of the reactor
and the variation of its ultrasonic parameters. By using high-speed photography
[36] and chemical methods such as the iodine method [37] and sonochemilumi-
nescence [38], it is possible to observe and record the spatial distribution of active
bubbles. Bubbles are forced to migrate directionally in the acoustic field due to
primary Bjerknes forces. The primary Bjerknes force is defined as the translational
force on a cavitating bubble in a liquid when the nonlinear oscillation of the cavi-
tating bubble is interacting with the acoustic pressure field [39]. Parlitz et al. [40]
demonstrated that when the amplitude of the acoustic pressure was not high, cavi-
tating bubbles tended to move along the direction of pressure rise (pressure anti-
node). This deduction is consistent with the phenomenon that is described in the
1 Fundamentals of Acoustic Cavitation in Sonochemistry 21
section dealing with concerted collapse (Fig. 1.5). However, when the amplitude
is further increased, the primary Bjerknes force may change, and the high pres-
sure amplitude region may become repulsive for the bubbles because of nonlinear
oscillation of bubbles [36, 40]. As a result, in an acoustic field near a sonotrode
with the pressure amplitude of greater than 190 kPa [36], cavitation bubbles tend
to move along the direction of pressure drop. This creates different spatial distribu-
tions and migration features of bubbles in a reactor that is commonly referred to as
acoustic cavitation structure.
Typical acoustic cavitation structures can be conical bubble structure (CBS),
smoker bubble structure, tailing bubble structure (TBS), jet-induced bubble struc-
ture (JBS) and acoustic Lichtenberg figure (ALF). Acoustic cavitation structure
CBS (Fig. 1.7a) [36] and smoker (Fig. 1.7b) [41] are two common forms that
Fig. 1.7 Acoustic cavitation structures: a conical bubble structure; b smoker bubble structure;
c tailing bubble structure; d jet-induced bubble structure; e acoustic Lichtenberg figure (a, b, d
and e are reprinted with permission from [36], Copyright © 2014 Elsevier. c is reprinted with
permission from [42], Copyright © 2014 Elsevier)
22 J. Luo et al.
The oscillation and transient collapse of cavitation bubbles produce many special
secondary physicochemical phenomena, including shockwave, microturbulence,
microjet, radical effects and sonoluminescence, which can have great influence on
the dynamics and process equilibrium in a system. Mass and heat transfer coef-
ficients of a reacting system can be multiplied by sonication with the predomi-
nant reason for enhancement being related to acoustic cavitation and its secondary
effects [28, 44]. The collapse of transient bubbles produces local hot spots with
temperature and pressure that is much higher than that of the surrounding liquid.
The huge temperature and pressure drop generates strong shockwaves towards
bubble outsides [45]. The shrinkage and transient collapse of bubbles creates
large voids in the liquid, while the quick influx of liquid stream to the voids forms
1 Fundamentals of Acoustic Cavitation in Sonochemistry 23
and emulsion separation under low intensity ultrasound by standing wave prin-
ciple. The action ranges of acoustic cavitation and relevant secondary effects
are much shorter. The surrounding liquid absorbs much of the cavitation energy,
while the reactive radicals generated by cavitation are easily quenched and cannot
migrate far from the location of the collapsed bubble. The attenuation of the pres-
sure pulse for the bubbles is greater than 5 μm, which makes the pressure ampli-
tude measured at 1 mm from the bubble center only 10−3 of its actual value [31].
In a heterogeneous system, such as in systems containing a high concentration
of small-size solid particles, the absorption of cavitation energy by the surround-
ings become significant due to the scattering effects. Near the surface of the solid
materials, the shockwave that has high temperature and high pressure from bub-
ble collapse, impacts the solid surface and changes the morphology and properties
of solid surface. The high energy shockwave lasts for short time of about 200 ns
and this happens within a short distance of only tens of nanometers from the solid
surface [2]. However, the energy from many shockwaves can be concentrated and
enhanced due to concerted collapse. Therefore, it is possibly for ultrasound wave
to make indentations of 10–100 μm on solid surface even if the solid surface is
exposed under acoustic cavitation for a short period of time [2].
As shown in Fig. 1.3, the active zones for acoustic cavitation and most of the
derived physicochemical effects can be divided into four parts—bubble interior,
bubble/liquid interface, nearby zone and far-distance zone. The bubble interior
allows the access of volatile chemicals and dissolved gases. The bubble generates
high temperatures and pressures in its interior during the nearly adiabatic com-
pression process, and initiates ionization, thermolysis and violent radical reactions
of vapor molecules. Radical reactions not only occur in the bubble interior, but
also at the bubble/liquid interface and in the nearby liquid, especially for nonvola-
tile compounds. In the sonochemical degradation of phenol in water, the increase
of relative concentration of nonvolatile phenol at the bubble/liquid interface pro-
motes the degradation efficiency of phenol multiply even though the concentration
of phenol in the bulk liquid remains almost unchanged [49]. This demonstrates
that the bubble/liquid interface plays a vital role in radical-induced chemical reac-
tions of nonvolatile compounds under sonication. However, it does not mean that
the chemical reactions of nonvolatile compounds cannot take place in the bubble
interior. In the study of sonoluminescence in concentrated H3PO4 solution, peaks
assigned to radicals ·OH and PO· are both significantly visible in the ultraviolet
spectrum [29]. In the sonoluminescence of H3PO4, radical PO· must be produced
from nonvolatile H3PO4 in the plasma zone, namely the deepest core in the bubble
structure. There are probably two different types of cavitating bubbles that exist
in sonicated H3PO4 solution [29]: (1) stationary bubbles that undergo highly sym-
metrical collapse and produce radical ·OH and (2) rapidly moving bubbles that
have less symmetrical collapse. In the second type of bubbles, nanodroplets of
liquid containing H3PO4 might be injected into the bubble interior due to capil-
lary wave action, microjetting, or bubble coalescence in the asymmetry collapse
of bubbles [29]. However, the ratio of chemical reactions of nonvolatile species in
bubbles is not large. As the collapse of transient bubble occurs, numerous active
1 Fundamentals of Acoustic Cavitation in Sonochemistry 25
radicals are released into the nearby liquid which initiate radical reactions such as
the oxidative degradation of chemical materials in wastewater or are captured by
chemical reagent in determining cavitation activity. For the nearby zone which is
within 100 μm from the bubble boundary, the dominant ultrasound-intensification
mechanism is physical, such as shockwave and microturbulence. High impact and
high temperatures that result from shockwave and high-speed microjet are respon-
sible for the modification of solid surfaces under high-energy ultrasound, such
as surface erosion, particle breakage, metal melting and peeling of surface oxide
coating [2, 50]. Microturbulence promotes dispersion and fine emulsification in
immiscible liquid mixtures and helps to dislodge substances that adhere onto solid
surfaces (ultrasonic cleaning). In the far-distance zone, cavitation and shockwaves
lose their strength, while microturbulence may be the main mechanism for ultra-
sonic intensification of chemical processing such as emulsification [48].
Most sonochemical reactions do not occur in the interior of active bubbles.
The intensification of these sonochemical reactions is through the action of sec-
ondary effects, but not the cavitation behavior of bubbles themselves. However,
the energy intensity of secondary effects greatly depends on the behavior of the
bubbles, as well as the liquid properties and heterogeneous characteristics of the
reaction system. For chemical reactions that are related to biomass conversion, the
effective intensification/activation zones are mainly the zones that are nearby the
bubbles and the zones that have distance within one wavelength from the bubbles.
Therefore, the physical mechanism should be applicable to the discussion of ultra-
sound-enhanced biomass processing, and intensification of the mass transfer.
For most practical sonochemical reactions, it is necessary to know the factors that
could influence acoustic cavitation, and these factors include the physicochemical
properties of liquids, acoustic operation and heterogeneous characteristics of reac-
tion system [4, 5, 51, 52]. The selection and optimization of those factors allows
proper energy intensity to be applied and gives higher energy efficiency. Table 1.3
describes some important factors and their effects on cavitation. Energy efficiency
for cavitation is related to the proper design of sonochemical reactors that provide
a uniform acoustic energy field for bubble cavitation. Generally, the influence of
the factors shown in Table 1.3 can be divided into four categories:
1. Influence on acoustic environment for the occurrence of cavitation. Factors
such as heterogeneous characteristics, acoustic frequency and liquid tempera-
ture influence the propagation and attenuation of acoustic energy in the liquid
and thus affect the intensity and uniformity of the acoustic field. The charac-
teristics of acoustic streaming in different reactors also influences the spatial
distribution of acoustic energy [54].
26 J. Luo et al.
Table 1.3 (continued)
Factor Direct effect Remarks
6. Acoustic intensity
1. Bubble dynamics 1. High acoustic intensity promotes cavitation
2. Number of cavitation events 2. Capacity for indefinite increase of acous-
tic intensity depends on material stability
of transducer and the properties of liquid
medium. Acoustic intensity is determined
according to reaction requirement
3. Bubble cloud effect seriously decreases
cavitation efficiency at high acoustic inten-
sity. Additional auxiliaries such as mechani-
cal agitation avoid bubble clouds gathering
in zones that have high acoustic intensity
7. Acoustic frequency
1. Acoustic cycle 1. Cavitation threshold increases with acous-
2. Bubble dynamics tic frequency
3. Cavitation threshold 2. High acoustic frequency of >1 MHz
4. Sound attenuation means short acoustic cycle and insufficient
time for the generation, growth and collapse
of bubbles
3. High frequency increases attenuation of
ultrasonic energy in liquid
4. Number of useful cavitation events per
unit time increases for acoustic frequency
of <200 kHz. This increases cavitation
efficiency
5. Longer duration of bubble motion and
larger bubble size exist with low frequency.
This increases the evaporation of liquid at
the bubble surface and chemical reaction of
vapor content in bubble, which consumes
energy in cavitation bubble, and decreases
the energy intensity of transient collapse
acoustic emission intensity is observed in the liquid that is degassed but not denu-
cleated which contains a controlled concentration of noncondensable air [26]. In
most cases, degassing before sonochemical treatment enhances the intensity of
cavitation and dramatically improves cavitation distribution. Cavitation is con-
centrated more closely to transducers when sonication is performed in tap water,
while it is highly dispersed and evenly in the whole reactor space when sonication
is performed in boiled water [60]. Nuclei concentration effects seem to artificially
regulate cavitation intensity in homogeneous liquids.
For solid-liquid heterogeneous systems, especially patterned geometry of
defects on solid surfaces, the size and concentration of the solid micro-/nanoparti-
cles are important for selective production or control of cavitation intensity [6, 7].
As mentioned in Sect. 1.2.1, smart design of surface defects can enhance the
nucleation rate by several orders of magnitude [6]. In biomass reactions handing
lignocellulose particles, appropriate particle size (several tens to several hundreds
of micrometers) [61, 62] and solid concentration (<5 wt%) [51] seems to promote
efficient reaction.
Another interesting heterogeneous system is the hot-pressed solvents such as
subcritical CO2 [63]. Acoustic cavitation only occurs below the critical point of
liquid, because no phase boundary exists above the critical point. Compared to
ambient water, acoustic cavitation in subcritical CO2 requires a lower threshold
pressure, as the high vapor pressure of CO2 counteracts the hydrostatic pressure.
The Blake threshold for subcritical CO2 at 20 °C and 5.82 MPa is only 0.1 MPa,
while water needs 5.9 MPa for the same conditions [63]. Theoretical calcula-
tions show that with stimulation of ultrasound at 20 kHz, liquid CO2 at 20 °C and
5.82 MPa gives high cavitation intensity, with the maximum bubble radius compa-
rable to that in water at 20 °C and ambient pressure [63]. The acoustic cavitation
in subcritical CO2 is already used in radical-induced chemical reactions, such as
polymer synthesis.
References
7. Zhang L, Belova V, Wang HQ, Dong WF, Möhwald H (2014) Controlled cavitation at nano/
microparticle surfaces. Chem Mater 26(7):2244–2248
8. Parkar PA, Choudhary HA, Moholkar VS (2012) Mechanistic and kinetic investigations in
ultrasound assisted acid catalyzed biodiesel synthesis. Chem Eng J 187:248–260
9. Minnaert M (1933) On musical air bubbles and the sounds of running water. Philos Mag
16(104):235–248
10. Suslick KS (1990) Sonochemistry. Science 247(4949):1439–1445
11. Bang JH, Suslick KS (2010) Applications of ultrasound to the synthesis of nanostructured
materials. Adv Mater 22(10):1039–1059
12. Gogate PR, Tayal RK, Pandit AB (2006) Cavitation: a technology on the horizon. Curr Sci
91(1):35–46
13. Flannigan DJ, Suslick KS (2005) Plasma formation and temperature measurement during
single-bubble cavitation. Nature 434(7029):52–55
14. Suslick KS, Eddingsaas NC, Flannigan DJ, Hopkins SD, Xu H (2011) Extreme condi-
tions during multibubble cavitation: sonoluminescence as a spectroscopic probe. Ultrason
Sonochem 18(4):842–846
15. Ashokkumar M (2011) The characterization of acoustic cavitation bubbles—An overview.
Ultrason Sonochem 18(4):864–872
16. Storey BD, Szeri AJ (2001) A reduced model of cavitation physics for use in sonochemistry.
Proc R Soc Lond A 457(2011):1685–1700
17. Toegel R, Gompf B, Pecha R, Lohse D (2000) Does water vapor prevent upscaling sonolumi-
nescence? Phys Rev Lett 85(15):3165–3168
18. Sivasankar T, Paunikar AW, Moholkar VS (2007) Mechanistic approach to enhancement of
the yield of a sonochemical reaction. AlChE J 53(5):1132–1143
19. Storey BD, Szeri AJ (2000) Water vapour, sonoluminescence and sonochemistry. Proc R Soc
Lond A 456(1999):1685–1709
20. Didenko YT, Suslick KS (2002) The energy efficiency of formation of photons, radicals and
ions during single-bubble cavitation. Nature 418(6896):394–397
21. Ashokkumar M, Lee J, Kentish S, Grieser F (2007) Bubbles in an acoustic field: an overview.
Ultrason Sonochem 14(4):470–475
22. Bremond N, Arora M, Dammer SM, Lohse D (2006) Interaction of cavitation bubbles on a
wall. Phys Fluids 18(12):121505
23. Jiao JJ, He Y, Leong T, Kentish SE, Ashokkumar M, Manasseh R, Lee J (2013) Experimental
and theoretical studies on the movements of two bubbles in an acoustic standing wave field.
J Phys Chem B 117(41):12549–12555
24. Stricker L, Dollet B, Fernández Rivas D, Lohse D (2013) Interacting bubble clouds and their
sonochemical production. J Acoust Soc Am 134(3):1854–1862
25. Hansson I, Mörch KA (1980) The dynamics of cavity clusters in ultrasonic (vibratory) cavi-
tation erosion. J Appl Phys 51(9):4651–4658
26. Moholkar VS, Warmoeskerken MMCG (2003) Integrated approach to optimization of an
ultrasonic processor. AlChE J 49(11):2918–2932
27. Avvaru B, Pandit AB (2009) Oscillating bubble concentration and its size distribution using
acoustic emission spectra. Ultrason Sonochem 16(1):105–115
28. Moholkar VS, Warmoeskerken MMCG, Ohl CD, Prosperetti A (2004) Mechanism of mass-
transfer enhancement in textiles by ultrasound. AlChE J 50(1):58–64
29. Xu HX, Glumac NG, Suslick KS (2010) Temperature inhomogeneity during multibubble
sonoluminescence. Angew Chem Int Ed 49(6):1079–1082
30. Sutkar VS, Gogate PR (2009) Design aspects of sonochemical reactors: techniques for under-
standing cavitational activity distribution and effect of operating parameters. Chem Eng J
155(1–2):26–36
31. Moholkar VS, Sable SP, Pandit AB (2000) Mapping the cavitation intensity in an ultrasonic
bath using the acoustic emission. AlChE J 46(4):684–694
32. Tuziuti T, Yasui K, Iida Y (2005) Spatial study on a multibubble system for sonochemistry by
laser-light scattering. Ultrason Sonochem 12(1–2):73–77
32 J. Luo et al.
33. Burdin F, Tsochatzidis NA, Guiraud P, Wilhelm AM, Delmas H (1999) Characterisation of
the acoustic cavitation cloud by two laser techniques. Ultrason Sonochem 6(1–2):43–51
34. Tsochatzidis NA, Guiraud P, Wilhelm AM, Delmas H (2001) Determination of velocity, size
and concentration of ultrasonic cavitation bubbles by the phase-Doppler technique. Chem
Eng Sci 56(5):1831–1840
35. Kadkhodaee R, Povey MJW (2008) Ultrasonic inactivation of Bacillus α-amylase. I. Effect
of gas content and emitting face of probe. Ultrason Sonochem 15(2):133–142
36. Bai LX, Xu WL, Deng JJ, Li C, Xu DL, Gao YD (2014) Generation and control of acoustic
cavitation structure. Ultrason Sonochem 21(5):1696–1706
37. Gogate PR, Tatake PA, Kanthale PM, Pandit AB (2002) Mapping of sonochemical reactors:
review, analysis, and experimental verification. AlChE J 48(7):1542–1560
38. Yin H, Qiao YZ, Cao H, Li ZP, Wan MX (2014) Cavitation mapping by sonochemilumines-
cence with less bubble displacement induced by acoustic radiation force in a 1.2 MHz HIFU.
Ultrason Sonochem 21(2):559–565
39. Louisnard O (2012) A simple model of ultrasound propagation in a cavitating liquid. Part II:
primary Bjerknes force and bubble structures. Ultrason Sonochem 19(1):66–76
40. Parlitz U, Mettin R, Luther S, Akhatov I, Voss M, Lauterborn W (1999) Spatio-temporal
dynamics of acoustic cavitation bubble clouds. Philos Trans R Soc London, Ser A
357(1751):313–334
41. Bai LX, Ying CF, Li C, Deng JJ (2012) The structures and evolution of Smoker in an ultra-
sonic field. Ultrason Sonochem 19(4):762–766
42. Bai LX, Deng JJ, Li C, Xu DL, Xu WL (2014) Acoustic cavitation structures produced by
artificial implants of nuclei. Ultrason Sonochem 21(1):121–128
43. Mettin R, Luther S, Ohl CD, Lauterborn W (1999) Acoustic cavitation structures and simula-
tions by a particle model. Ultrason Sonochem 6(1–2):25–29
44. Legay M, Gondrexon N, Le Person S, Boldo P, Bontemps A (2011) Enhancement of heat
transfer by ultrasound: review and recent advances. Int J Chem Eng 2011:17
45. Ohl CD, Kurz T, Geisler R, Lindau O, Lauterborn W (1999) Bubble dynamics, shock waves
and sonoluminescence. Philos Trans R Soc London Ser A 357(1751):269–294
46. Kalva A, Sivasankar T, Moholkar VS (2009) Physical mechanism of ultrasound-assisted syn-
thesis of biodiesel. Ind Eng Chem Res 48(1):534–544
47. Sutkar VS, Gogate PR, Csoka L (2010) Theoretical prediction of cavitational activity distri-
bution in sonochemical reactors. Chem Eng J 158(2):290–295
48. Cucheval A, Chow RCY (2008) A study on the emulsification of oil by power ultrasound.
Ultrason Sonochem 15(5):916–920
49. Sivasankar T, Moholkar VS (2008) Mechanistic features of the sonochemical degradation of
organic pollutants. AlChE J 54(8):2206–2219
50. Suslick KS, Skrabalak SE (2008) Sonocatalysis. In: Ertl G, Knözinger H, Schüth F,
Weitkamp J (eds) Handbook of heterogeneous catalysis, 2nd edn. Wiley-VCH Verlag GmbH
& Co. KGaA, Weinheim, pp 2007–2017
51. Luo J, Fang Z, Smith RL Jr (2014) Ultrasound-enhanced conversion of biomass to biofuels.
Prog Energy Combust Sci 41:56–93
52. Vichare NP, Senthilkumar P, Moholkar VS, Gogate PR, Pandit AB (2000) Energy analysis in
acoustic cavitation. Ind Eng Chem Res 39(5):1480–1486
53. Rooze J, Rebrov EV, Schouten JC, Keurentjes JTF (2013) Dissolved gas and ultrasonic cavi-
tation—A review. Ultrason Sonochem 20(1):1–11
54. Gogate PR, Sutkar VS, Pandit AB (2011) Sonochemical reactors: important design and
scale up considerations with a special emphasis on heterogeneous systems. Chem Eng J
166(3):1066–1082
55. Tran KVB, Kimura T, Kondo T, Koda S (2014) Quantification of frequency dependence of
mechanical effects induced by ultrasound. Ultrason Sonochem 21(2):716–721
56. Hua I, Hoffmann MR (1997) Optimization of ultrasonic irradiation as an advanced oxidation
technology. Environ Sci Technol 31(8):2237–2243
1 Fundamentals of Acoustic Cavitation in Sonochemistry 33
57. Eddingsaas NC, Suslick KS (2007) Evidence for a plasma core during multibubble sonolu-
minescence in sulfuric acid. J Am Chem Soc 129(13):3838–3839
58. Oxley JD, Prozorov T, Suslick KS (2003) Sonochemistry and sonoluminescence of room-
temperature ionic liquids. J Am Chem Soc 125(37):11138–11139
59. Bradley M, Ashokkumar M, Grieser F (2013) Multibubble sonoluminescence in ethylene
glycol/water mixtures. J Phys Chem B 118(1):337–343
60. Liu LY, Yang Y, Liu PH, Tan W (2014) The influence of air content in water on ultrasonic
cavitation field. Ultrason Sonochem 21(2):566–571
61. Bussemaker MJ, Xu F, Zhang DK (2013) Manipulation of ultrasonic effects on lignocel-
lulose by varying the frequency, particle size, loading and stirring. Bioresour Technol
148:15–23
62. Rezania S, Ye ZL, Berson RE (2009) Enzymatic saccharification and viscosity of saw-
dust slurries following ultrasonic particle size reduction. Appl Biochem Biotechnol
153(1–3):103–115
63. Kuijpers MWA, van Eck D, Kemmere MF, Keurentjes JTF (2002) Cavitation-induced reac-
tions in high-pressure carbon dioxide. Science 298(5600):1969–1971