Вы находитесь на странице: 1из 9

Ultrasonics Sonochemistry 18 (2011) 864–872

Contents lists available at ScienceDirect

Ultrasonics Sonochemistry
journal homepage: www.elsevier.com/locate/ultsonch

The characterization of acoustic cavitation bubbles – An overview


Muthupandian Ashokkumar
School of Chemistry, University of Melbourne, VIC 3010, Australia

a r t i c l e i n f o a b s t r a c t

Article history: Acoustic cavitation, in simple terms, is the growth and collapse of preexisting microbubbles under the
Received 7 October 2010 influence of an ultrasonic field in liquids. The cavitation bubbles can be characterized by the dynamics
Accepted 26 November 2010 of oscillations and the maximum temperatures and pressures reached when they collapse. These aspects
Available online 4 December 2010
can be studied both experimentally and theoretically for a single bubble system. However, in a multibub-
ble system, the formation of bubble streamers and clusters makes it difficult to characterize the cumu-
Keywords: lative properties of these bubbles. In this overview, some recently developed experimental procedures
Cavitation bubbles
for the characterization of acoustic cavitation bubbles have been discussed.
Bubble size
Bubble temperature
Ó 2010 Elsevier B.V. All rights reserved.
Transient cavitation
Stable cavitation

1. Introduction bubbles. An overview of these experimental procedures and their


significance in understanding and controlling of acoustic cavitation
Ultrasound, when passing through a liquid medium causes bubbles has been provided. In particular, the experimental tech-
mechanical vibration of the liquid. In addition to this effect, ultra- niques for the characterization of the growth, temperature, lifetime
sound also generates acoustic streaming within the liquid. If the li- and capability (chemistry vs. light emission) of cavitation bubbles
quid medium contains dissolved gas nuclei, which will be the case have been briefly discussed in this overview. For detailed informa-
under normal conditions, they can be grown and collapsed by the tion on these techniques, appropriate references that have been ci-
action of the ultrasound. The phenomenon of growth and collapse ted can be referred.
of microbubbles under an ultrasonic field is known as ‘‘acoustic
cavitation’’ [1]. When cavitation bubbles oscillate and collapse, 2. Bubble growth
several physical effects are generated, namely, shock waves, micro-
jets, turbulence, shear forces, etc. The physical effects of ultrasound Unlike in a single bubble system, the bubble growth in a multi-
have been used for a number of applications that include emulsifi- bubble system involves two processes, namely, rectified diffusion
cation, extraction, cleaning, etc. [2]. The collapse of the acoustic and bubble coalescence [13]. While it is possible to use experimen-
cavitation bubbles is also near adiabatic and generates tempera- tal techniques such as light scattering, stroboscopic or fast video
tures of thousands of degrees within the bubbles for a short period recording to monitor the rectified diffusion growth of the bubbles
of time [3]. Under this extreme temperature conditions, highly in a single bubble system [3,14], these techniques cannot be easily
reactive radicals are generated. For example, if water is the med- adapted for multibubble systems. Added to this complexity is the
ium, H and OH radicals are generated by the homolysis of water. contribution from bubble coalescence to the growth of the bubbles.
These radicals have been used to achieve chemical reactions that Despite the complexities involved in the bubble growth process, a
include the synthesis of nanomaterials, polymers, degradation of simple experimental technique based on the initial growth of mul-
organic pollutants, etc. [1,2,4–7]. In addition, acoustic cavitation tibubble sonoluminescence (MBSL) has been developed by us to
is found to be useful in diagnostic and therapeutic medicine [8]. qualitatively monitor the bubble growth by rectified diffusion
Despite the use of ultrasonics and sonochemistry in a variety of and bubble coalescence.
applications [4–12], the studies devoted to the fundamental under- It has been shown that several acoustic pulses are required to
standing of the characteristics of cavitation bubbles are limited. attain a steady-state MBSL in aqueous solutions under pulsed son-
We have been, for the past 15 years, developing a number of exper- ication conditions, in particular at high ultrasound frequencies
imental procedures for the characterization of acoustic cavitation [15,16]. The reason for this induction period to reach a steady-state
MBSL is the absence of resonance sized bubbles at the initial stages
of sonication. The bubble nuclei present in the solution need to
E-mail address: masho@unimelb.edu.au grow to reach the resonance size range (see Fig. 1 [17]).

1350-4177/$ - see front matter Ó 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.ultsonch.2010.11.016
M. Ashokkumar / Ultrasonics Sonochemistry 18 (2011) 864–872 865

Fig. 1. Schematic representation of the growth of cavitation bubbles in an acoustic field and the corresponding experimental data (515 kHz, pulse on = 4 ms, pulse
off = 12 ms) on the growth of MBSL. Adapted from Ref. [17].

As mentioned above, the bubble growth in a multibubble filed in Fig. 4 also support the above discussion. In water, large coa-
can occur by either rectified diffusion or bubble coalescence or lesced bubbles are clearly visible whereas, they are completely ab-
by a combination of both processes. We have used the initial MBSL sent in 1 mM SDS solution.
growth data to qualitatively understand the bubble growth under Returning back to the discussion regarding the initial growth of
pulsed sonication conditions. Fig. 2 shows that about fifteen 4 ms MBSL (Fig. 2), it is possible to interpret the data by bubble coales-
pulses (duration between pulses is 12 ms) are required to reach a cence and rectified diffusion processes. In water, bubble coales-
steady-state bubble population of sonoluminescence bubbles in cence occurs as shown by the capillary technique data (Fig. 3).
water. Compared to water, the number of pulses required to reach Hence, the initial growth of the cavitation bubbles may occur by
steady-state increases to about 30 in aqueous solution containing bubble coalescence pathway. However, rectified diffusion also con-
1 mM of a surfactant (sodium dodecyl sulfate, SDS). Let us return tributes to this growth. The support for this argument comes from
to this observation after some discussion on the role of surfactants the SDS data in Fig. 3, which shows that SDS retards bubble coales-
in bubble coalescence. cence. Hence, in the presence of SDS, the growth of cavitation bub-
Surfactants are known to adsorb at bubble solution interface bles by bubble coalescence pathway is significantly reduced. The
and reduce the coalescence between bubbles [13]. We have used initial growth of MBSL observed in Fig. 2 must then be due to the
a capillary technique to monitor bubble coalescence among cavita- growth of cavitation bubbles primarily by rectified diffusion path-
tion bubbles in the absence and presence of surface active solutes way. Based on the number of acoustic pulses required for steady-
[18]. The capillary technique involves the attachment of a capillary state MBSL in water, it can be suggested that the contributions
that can measure the change in volume (DVT) that occurs due to for the bubble growth form bubble coalescence and rectified diffu-
the formation of large inactive bubbles formed by coalescence sion are approximately equal. In the absence of bubble coales-
among cavitation bubbles. This is schematically shown in Fig. 3 cence, almost double the number of acoustic pulses is required in
along with a photograph of a 500 kHz ultrasound generator and SDS solution where the bubbles primarily grow by rectified diffu-
capillary setup. Thus, a large DVT means more coalescence and a sion pathway. It should be clearly understood that the observed re-
small DVT means less coalescence between bubbles. sults are for a given ultrasound frequency under specific
Fig. 3 also shows that the relative coalescence between cavita- experimental conditions (power, duty cycle, etc.). The number of
tion bubbles in 1 mM SDS solution is significantly reduced to about acoustic pulses required may be different under different experi-
20% compared to that observed in water. The photographs shown mental conditions.

Fig. 2. The initial growth of MBSL as a function of acoustic pulse number. Frequency = 515 kHz, duty cycle = 4 ms on, 12 ms off.
866 M. Ashokkumar / Ultrasonics Sonochemistry 18 (2011) 864–872

Fig. 3. A 500 kHz ultrasound generator (Undatim Ultrasonics) along with the capillary setup. A schematic diagram showing the volume change in water and surfactant
solution and the experimental coalescence data are also shown. Adapted from Ref. [17].

temperatures in the range of 5000 K – >50,000 K [3,19] have been


estimated. However, experimental estimates of the temperature
within the collapsing bubbles based on multibubble sonochemis-
try and sonoluminescence are reported to be between 750 K and
6000 K. Misík et al. [20], using the kinetic isotope effect in an EPR
spin-trapping study of the sonolysis of H2O/D2O mixtures found
that the cavitation temperature determined was dependent on
the specific spin trap used and fell in the range of 1000–4600 K.
They attributed the difference in the values obtained by the spin
traps as being due to the sampling of different regions of the cav-
itation ‘‘hot spot’’. Misík and Riesz [21] made use of the kinetic
isotope effect in the ultrasound induced production of radicals
Fig. 4. Photographs taken during sonication of water and 1 mM SDS solution at in organic liquids to estimate the temperatures during cavitation.
515 kHz. In water the formation of large bubbles due to coalescence is clearly seen.
The temperature region where hydrogen radical abstraction oc-
In 1 mM SDS solution, no large bubbles are observed.
curred in n-dodecane was estimated to be about 750 K, whereas
the region where benzyl radical formation occurred in toluene
3. Bubble temperature was about 6000 K. An interesting aspect of this work, as in the
H2O/D2O study described above, was the determination of a mean
The bubble temperature is another important characteristic of temperature in different regions of a ‘‘hot spot’’ depending on
cavitation bubbles. Based on a simple thermodynamic model for which radicals were detected. Using comparative rate thermome-
bubble collapse and assuming adiabatic compression takes place, try in alkane solutions, Suslick et al. [22] postulated that there are
the maximum theoretical temperature within the bubble (Tmax) two regions of sonochemical reactivity: a gas phase zone within
can be calculated using Eq. (1) [19]: the collapsing cavity with an estimated temperature and pressure
of 5200 ± 650 K and 500 atm, respectively, and a thin liquid layer
  immediately surrounding the collapsing cavity with an estimated
Pm ðc  1Þ
T max ¼ T 0 ð1Þ temperature of 1900 K.
Pv
Sonolysis of methane in argon saturated water has been used by
T0 is the ambient solution temperature, Pm is the pressure in the Henglein and coworkers [23] to estimate the bubble core temper-
liquid (a sum of the hydrostatic and acoustic pressures), c ¼ ccPv is ature. Depending upon the percentage of methane and argon pres-
the specific heat ratio of the gas/vapor mixture. Pv is the pressure ent in water, the temperature was estimated to be in the range of
in the bubble at its maximum size; usually assumed to be equal 1930–2720 K (the temperature decreased with an increase in the
to the vapor pressure of the liquid, although it is known that gas percentage of the methane). Tauber et al. [24] used a similar tech-
molecules originally dissolved in the fluid will also be present. nique in their study of the sonolysis of t-butanol in water, and esti-
For example, if we assume that pm is 2 atm, the gas inside the bub- mated the temperature to be in the range 2300 K and 3600 K. The
ble is equal to the equilibrium water vapor pressure (Pv = 0.031 lower temperature was obtained for solution with the highest con-
atm), the ambient solution temperature is 298 K and c for water centration of t-butanol (0.5 M). Both these studies show that the
vapor is 1.32, then the theoretical estimate of Tmax will be core temperature of the bubble is strongly dependent upon the
6150 K. It should be noted that Eq. (1) over estimates the Tmax, amount of organic solute in the system.
because it does not take into account the heat leaking from the Ashokkumar and Grieser [25] have shown that there exists two
bubble or thermal conductivity of the gases or the energy con- bubble temperatures: a time and volume averaged temperature
sumed in the decomposition of the vapor/gas within the bubble. that is responsible for the chemical reactions and the peak temper-
Experimental determinations of the temperature within a cav- ature that is responsible for the SL. In Fig. 5, it can be seen that the
itation bubble have been made by a number of research groups. bubble temperatures measured by the methyl radical recombina-
By fitting the experimentally recorded single bubble sonolumi- tion (MRR) method [23–26] is not significantly affected in the pres-
nescence spectra using the spectrum radiated by a blackbody, ence of up to 50 mM of ethanol whereas the SL intensity is almost
M. Ashokkumar / Ultrasonics Sonochemistry 18 (2011) 864–872 867

Fig. 5. It is shown in this figure that the relative MBSL intensity decreases to more than 90% of that observed in water in the presence of ethanol whereas the bubble
temperature measured by the MRR method is not significantly affected by the same amount of ethanol. Frequency: 356 kHz. Adapted from Ref. [25].

90% quenched. It has been discussed that the volatile ethanol and discussed later. It can also be noticed that the average bubble tem-
its decomposition products can decrease the peak temperature peratures (chemical temperature discussed above) do decrease at
reached upon bubble collapse without affecting the average tem- higher alcohol concentrations at higher frequencies: the cavitation
perature, measured by the MRR method. bubble temperatures measured decrease with an increase in the
We have also measured the cavitation bubble temperatures concentrations of the alcohols and for a given concentration of
using the MRR method in argon saturated alcohol solutions at the alcohol, the longer chain alcohol shows a greater effect. The
20 kHz, 363 kHz, and 1056 kHz and the results are shown in reason for the lowering of the bubble temperatures at high fre-
Fig. 6 [27]. It can be seen at 20 kHz that the bubble temperatures quencies is the hydrocarbon product accumulation as a result of
did not significantly change in the presence of the alcohols. The the evaporation and decomposition of volatile solutes under the
reason for almost no change in the bubble temperature may be high temperature conditions generated during bubble collapse
due to the transient nature of the cavitation bubbles, which is [28]. The higher surface activity of butanol compared to that of

Fig. 6. Cavitation bubble temperatures measured using the MRR [23–26] method at 20 kHz, 363 kHz and 1056 kHz as a function of the concentrations of ethanol and butanol.
Adapted from Ref. [27].
868 M. Ashokkumar / Ultrasonics Sonochemistry 18 (2011) 864–872

ethanol causes more butanol molecules to evaporate into the bub- water from the surface of the bubble. Thus, the relatively lower
ble during the expansion phase of the bubble oscillation leading to amount of water vapor present inside the collapsing bubbles at
the production of a relatively larger amount of hydrocarbon prod- higher frequencies results in a lower heat energy consumption that
ucts. A similar trend in MBSL quenching has been reported and dis- leads to relatively higher temperatures on bubble collapse. It
cussed based on the surface activity of the alcohols [28]. should also be noted that the amount of primary radicals generated
The temperature data shown in Fig. 6 can also be used to obtain per collapsing bubble is larger at lower frequencies. However, the
the cavitation bubble temperature in the absence of any solute by overall radical yield, which is the product of the radicals per bubble
extrapolating the temperature data to zero concentration of the and the total number of bubbles, is higher at higher frequencies.
solutes. By extrapolating to zero, both alcohols data yield a ‘‘zero’’ This is due to the increase in the number of bubbles at higher fre-
concentration temperature in the range 3700–4200 K at 20 kHz. quencies. In fact, the overall radical yield seems to be the maxi-
This variation in the cavitation bubble temperature might be due mum in the frequency range 200–600 kHz [1]. The radical yield
to experimental variations. The bubble temperature obtained at decreases at very high frequencies (for example, 1 MHz) due to
363 kHz (Fig. 6) by extrapolating to zero solute concentrations is the lack of time available for evaporation of water as discussed
in the range of 4200–4700 K. A similar treatment at 1056 kHz above.
yields a bubble temperature in the range of 5000–6200 K. The tem-
perature trend thus observed for water at different frequencies is, 4. Bubble size
T1056 kHz > T363 kHz > T20 kHz. This trend is opposite to what can be
expected theoretically. At 20 kHz, the theoretical resonance size While it is relatively convenient to monitor the cavitation bub-
of the bubble is about 150 lm; this is very large compared to the ble size in a single bubble system using either light scattering or
bubble size at 1056 kHz, which is about 3 lm. A 20 kHz bubble stroboscopic imaging, it is difficult to monitor the bubble size in
would have a large dV (change in volume) term on collapse leading a multibubble environment due to bubble clustering and interfer-
to a higher collapse temperature compared to that of a 1056 kHz ence caused by bubble movements. A few experimental techniques
bubble. However, the observed trend can be explained, if the that are available to measure the bubble size are laser light diffrac-
amount of water vapor that evaporated within one acoustic cycle tion [30], active cavitation detection [31] and phase-Doppler [32].
was taken into account. We have introduced a relatively simple technique based on pulsed
A 20 kHz bubble has a relatively longer expansion cycle time MBSL [33]. The method is based on the dissolution of the cavitation
than a bubble at a higher frequency; this would result in the evap- bubbles during the ‘‘off’’ time of the ultrasound between pulses as
oration of relatively more water molecules at 20 kHz than that at a shown in Fig. 7. During the pulse ‘‘on’’ time, cavitation bubbles
higher frequency. These molecules can effectively consume more grow as discussed earlier. During the pulse ‘‘off’’ time, bubbles dis-
heat during the bubble collapse. Owing to a very short expansion solve. For a given duty cycle, the system continues to maintain a
cycle time, a 1056 kHz bubble would not have as many water mol- steady-state population of active bubbles and approximately con-
ecules within the bubble as at a lower frequency to consume the stant MBSL intensity as shown in Fig. 2 [15,16].
heat energy. It can be seen in Fig. 7 that the average MBSL intensity does not
Tronson [29] has calculated the total mass of water vapor that change significantly when the ‘‘off’’ time is increased from 12 ms to
could evaporate into an expanding bubble by using Eq. (2). 200 ms. However, beyond this value, the MBSL intensity decreases
Z t¼t and almost reaches zero beyond 290 ms. What this indicates is that
mvap ¼ 4pJ obs R2 ðtÞdt ð2Þ the active cavitation bubble population has reached a size range
t¼0
that is below a critical value to grow the bubbles to that required
where mvap is the upper bound for the total mass of water vapor for SL to occur during the next pulse on period. This critical size
evaporated into the bubble, Jobs is the rate of efflux of water vapor can be considered as equal to the resonance size of the bubbles:
per unit area; =178 g s1 m2, R is the bubble radius and t is the bubbles that reach the resonance size range grow to a maximum
expansion time (approximately equal to the expansion cycle time and collapse to emit light. It can be assumed that the bubbles are
of the sound wave). This quantity can be compared to the mass just below the resonance size when the average MBSL intensity
(in grams) of water present in a monolayer (mmon) at the surface reaches zero at this critical pulse off time. Using the Epstein–Ples-
of the bubble at Rmax [29] using Eq. (3). set equation shown in Fig. 7 Caption and the procedure discussion
by Lee et al. [33], the resonance size of the cavitation bubbles can
4pR2max Mw be estimated. The data provided in Table 2 shows that the bubble
mmon ¼ ð3Þ
N A Aw size measured using the pulsed SL technique is in agreement with
where NA is the Avagadro number, Mw is the molar mass of water that measured by other techniques.
(=18 g mol1) and Aw is the cross sectional area of a molecule of
water at a surface, taken to be 11 Å2 in this case. Table 1 summa- 5. Bubble lifetime (transient vs. stable)
rises the mvap and mmon data calculated using Eqs. (2) and (3) for
the three frequencies used in this investigation. There is always confusion with the terminology referring to
The data in Table 1 clearly shows that the amount of water va- transient and stable bubbles. Stable bubbles are considered to be
por evaporated into the bubble is equivalent to several monolayers weakly and symmetrically oscillating bubbles, whereas transient
of water for 20 kHz, whereas for the higher frequencies the expan- bubbles are considered to be ‘‘active cavitation’’ bubbles. Leighton
sion time is not sufficient for the evaporation of one monolayer of [19] has provided an appropriate discussion to clarify this

Table 1
mvap and mmon data along with Rmax and Ro data used in the calculations [27].

Frequency (kHz) Expansion cycle time (s) Rmax (m) Ro (m) mvap (g) mmon (g)
6 6 6 10
20 25  10 150  10 15  10 6.2  10 7.7  1011
360 1.4  106 8  106 0.8  106 9.8  1014 2.2  1013
1060 0.47  106 2.8  106 0.28  106 4.1  1015 2.7  1014
M. Ashokkumar / Ultrasonics Sonochemistry 18 (2011) 864–872 869

Fig. 7. The growth and dissolution of cavitation bubbles under pulsed sonication conditions (left); MBSL observed during the 4 ms pulse on time at various pulse off times
(right top); Epstein and Plesset equation (right bottom) [33] for a dissolving bubble: D – diffusion coefficient, C – gas concentration, qg – gas density, Ro – initial bubble radius
(assumed to be equal to the resonance radius), M – molar mass of the gas, c – surface tension of the liquid, R – gas constant, T – temperature. Using the pulse off time for ‘‘t’’, Ro
is calculated.

confusion. He mentions, quote: ‘‘the terms ‘‘stable’’ and ‘‘transient’’ Table 2


are perhaps unfortunate, since they suggest a temporal differenti- Experimentally measured resonance bubble radii at various frequencies [33].
ation, when in fact they are most useful in describing the energy
Experimental technique Frequency (kHz) Experimental Ro (lm)
and effects of the collapse’’. For example, single bubble cavitation
Pulsed MBSL 515 2.8–3.7
can be considered as ‘‘stable’’ cavitation in a sense that the same
Active cavitation 1100 0.9, 1.38
bubble undergoes continuous oscillations emitting light on col- Laser diffraction 20 3.8
lapse during each cycle. The energy of the collapse is significant Phase-doppler 20 5.0
in that it can also be classified as ‘‘transient collapse’’. In order to
overcome this confusion, Leighton further classified the high-en-
ergy collapse (transient cavitation) into ‘‘Fragmentary Transient
Cavitation’’ and ‘‘Repetitive Transient Cavitation (High-energy sta- acoustic cycles whereas stable bubbles existed for hundreds of
ble cavitation)’’. When we first investigated [28,34] cavitation at acoustic cycles [34]. We observed that the sonoluminescence
low and high ultrasound frequencies, we have referred to fragmen- intensity is decreased significantly in the presence of volatile sol-
tary transient cavitation as simply transient cavitation and the utes at higher ultrasound frequencies (e.g., 515 kHz) whereas SL
high-energy stable cavitation as just stable cavitation. From the intensity is not affected by the same solute at 20 kHz (Fig. 8). We
collapse energy point of view, they both refer to high-energy col- proposed that the accumulation of hydrocarbon products within
lapse. The transient cavitation bubbles then refer to high-energy cavitation bubbles is important. At 515 kHz, the solution volume
collapse followed by fragmentation and generation of new cavita- (and height) used was sufficient to establish standing waves in
tion nuclei and the stable cavitation bubbles refer to high-energy the reaction cell. Under standing wave conditions, bubbles undergo
collapse where same bubbles repetitively collapse and grow sev- ‘‘stable’’ cavitation. At 20 kHz, the solution in the reaction cell was
eral times. not sufficient for the establishment of standing waves and under
We have provided experimental evidence based on MBSL these circumstances, bubbles underwent transient cavitation. Sta-
quenching where transient bubbles existed only for one or a few ble bubbles are considered to undergo a large number of oscilla-
tions during their existence and in doing so exist for a sufficient
time for surface active solutes to adsorb to the bubble/solution
interface, accumulate gaseous decomposition products, and to al-
low a significant amount of the gases solubilised in solution to en-
ter the bubble. In contrast, for transient bubbles these events can
be significantly reduced.
Ashokkumar et al.’s recent work [35] on MBSL at low and high
frequencies introduced another complexity to the behavior of the
cavitation bubbles. It was thought that the reason for the genera-
tion of transient bubbles might be the absence of a standing wave
in a horn system and the stable bubbles can only be generated un-
der standing wave conditions. In order to clarify this aspect, MBSL
quenching experiments were carried out at four ultrasound fre-
quencies keeping the sonication cell configuration the same. The
solution height and volume were chosen to make sure that stand-
ing waves could be formed at all frequencies. In the first system, a
20 kHz horn was positioned closer to the water/air interface (see
Fig. 9). When the MBSL was photographed using this arrangement,
active cavitation was observed closer to the tip of the horn only.
Despite the possibility for a standing wave, there were no active
cavitations observed in any other region of the reactor. However,
Fig. 8. MBSL quenching at 20 kHz and 515 kHz as a function of alcohol concentra- when a 25 kHz plate transducer was attached at the bottom of
tion. Adapted from Ref. [34]. the sonication cell (reactor), standing waves were visible in the
870 M. Ashokkumar / Ultrasonics Sonochemistry 18 (2011) 864–872

Fig. 9. (a) Photographs showing the sonication cell used at 20 kHz (horn; left), MBSL at 20 kHz (horn) and MBSL at 37 kHz (right); (b) MBSL quenching data at different
frequencies. Adapted from Ref. [35].

Fig. 10. (a) Sonoluminescence from water and (b) sonochemiluminescence from an aqueous solution containing luminol; frequency = 170 kHz; power = 12 W [39].

MBSL photograph. A similar pattern was observed at 37 kHz (Fig. 9) bubble lifetimes at different ultrasound frequencies, which range
and at 440 kHz. The MBSL quenching data observed in the presence from 100 to 350 ls that correspond to 70–100 acoustic cycles.
of various concentrations of propanol is also shown in Fig. 9.
One interesting aspect that can be seen immediately is a rela-
tively small SL quenching (about 20%) observed with 20 kHz horn 6. SL bubbles vs. SC bubbles
system compared to almost no quenching reported by Tronson
et al. [34]. It may be speculated that there might be some active The data shown in Fig. 5 indicates that the lower MBSL does not
stable bubbles present in the 20 kHz system used by Ashokkumar directly correlate with the cavitation activity of a system. When
et al. [35] that were not visible in the photograph. However, com- MBSL intensity is decreased by more than 90% in the presence of
pared to the SL quenching observed with the horn system, about small amounts of ethanol, there are still hydrocarbon products pro-
80% SL quenching was observed at 25 kHz and 100 mM propanol duced in the bubbles indicating the existence of cavitation bubbles
when a plate transducer was used. A similar SL quenching was also generating sonochemical reactions. As indicated previously, the
observed at 37 kHz. Considering that the SL quenching observed at cavitation bubbles may not attain the temperatures required for
25 kHz and 37 kHz were almost similar to that observed at SL to occur, however, the temperatures may be sufficient to
447 kHz, it was suggested that these systems consisted of primar- decompose molecules. In following reports, we have used sonolu-
ily stable bubbles. The reason for the generation of primarily tran- minescence and sonochemiluminescence photographs to show the
sient bubbles by the 20 kHz horn is the high energy density existence of two groups of cavitation bubbles in a given reactor
delivered at the tip of the horn compared to the more diffused en- [38,39]. One group of bubbles reaches higher temperatures for SL
ergy density delivered in plate type transducers. It has been esti- to occur and the 2nd group causing chemical reactions. When
mated that the stable bubbles may have lifetimes of up to 5000 water was sonicated at 170 kHz, sonoluminescence was observed
acoustic cycles depending upon the acoustic frequency [36]. Using primarily closer to the water/air interface as can be seen by the
bubble coalescence data, Sunartio et al. [37] have estimated the weak emission in Fig. 10a. However, when luminol was present
M. Ashokkumar / Ultrasonics Sonochemistry 18 (2011) 864–872 871

[4] K.S. Suslick, Ultrasound: Its Chemical, Physical and Biological Effects, VCH
Publishers, New York, 1988.
[5] G. Price (Ed.), Current Trends in Sonochemistry, RSC, Cambridge, UK, 1992.
[6] M. Ashokkumar, Sonochemical synthesis of inorganic nanoparticles, in: P.D.
Cozzoli (Ed.), Advanced Wet-Chemical Synthetic Approaches to Inorganic
Nanostructures, Transworld Research Network, 2008, pp. 107–131 (Chapter
4).
[7] S. Muthukumaran, S.E. Kentish, M. Ashokkumar, G.W. Stevens, Application of
ultrasound in membrane separation processes: a review, Rev. Chem. Eng. 22
(2006) 155–194.
[8] F. Cavalieri, M. Zhou, M. Ashokkumar, The design of multifunctional
microbubbles for ultrasound image-guided cancer therapy, Curr. Top. Med.
Chem. 10 (2010) 1198–1220.
[9] R. Bhaskaracharya, S.E. Kentish, M. Ashokkumar, Selected applications of
ultrasonics in food processing, Food Eng. Rev. 1 (2009) 31–49.
[10] T.J. Mason, C. Petrier, in: S. Parsons (Ed.), Advanced Oxidation Processes for
Water and Wastewater Treatment, IWA Publishing, 2004, pp. 185–208.
[11] J.-L. Luche (Ed.), Synthetic Organic Chemistry, Plenum Press, New York, 1998.
[12] M.J. Povey, T.J. Mason (Eds.), Ultrasound in Food Processing, Springer, New
York, 1998.
[13] M. Ashokkumar, J. Lee, S. Kentish, F. Grieser, Bubbles in an ultrasonic field: an
overview, Ultrason. Sonochem. 14 (2007) 470–475.
Fig. 11. Bubble radius distribution of SL and SC bubbles. Frequency: 575 kHz. [14] J. Lee, S. Kentish, M. Ashokkumar, Effect of surfactants on the rate of growth
Adapted form Ref. [40]. of an air bubble by rectified diffusion, J. Phys. Chem. B 109 (2005) 14595–
14598.
in the same reactor, chemiluminescence was observed from more [15] A. Henglein, R. Ulrich, J. Lilie, Luminescence and chemical action by pulsed
ultrasound, J. Am. Chem. Soc. 111 (1989) 1974–1979.
than 90% of the reactor under the same experimental conditions [16] M. Ashokkumar, R. Hall, P. Mulvaney, F. Grieser, Sonoluminescence from
(Fig. 10b). It is well known that the luminol chemiluminescence aqueous alcohol and surfactant solutions, J. Phys. Chem. B 101 (1997) 10845–
is initiated by the reaction between cavitation generated OH radi- 10850.
[17] J. Lee, The Behavior of Ultrasound Generated Bubbles in the Presence of
cals and luminol molecules. Surface Active Solutes, Ph.D. Thesis, University of Melbourne, 2005.
It was also shown that the SC bubbles appeared at relatively [18] J. Lee, S. Kentish, M. Ashokkumar, The effect of surface-active solutes on bubble
lower acoustic power levels compared to the SL bubbles at a given coalescence in the presence of ultrasound, J. Phys. Chem. B 109 (2005) 5095–
5099.
frequency and the power required to observe SL bubbles increased
[19] T.G. Leighton, The Acoustic Bubble, Academic Press, London, 1994.
with an increase in the ultrasound frequency [39]. In this study, it [20] V. Misík, N. Miyoshi, P. Reisz, EPR spin-trapping of the sonolysis of H2O/D2O
was speculated that the sonoluminescence bubbles may reach a mixtures: probing the temperatures of cavitation regions, J. Phys. Chem. 99
relatively larger size and generate higher Tmax compared to those (1995) 3605–3611.
[21] V. Misík, P. Reisz, EPR study of free radicals induced by ultrasound in organic
bubbles that are sonochemically active. Further experimental sup- liquids II. Probing the temperatures of cavitation regions, Ultrason. Sonochem.
port to this speculation was later obtained using the bubble size 3 (1996) 25–37.
measurement procedure discussed earlier. The bubble size distri- [22] K.S. Suslick, D.A. Hammerton, R.E. Cline, Sonochemical hotspot, J. Am. Chem.
Soc. 108 (1986) 5641–5642.
butions were measured using pulsed SL and pulsed SCL and the re- [23] E.J. Hart, C.-H. Fischer, A. Henglein, Sonolysis of hydrocarbons in aqueous
sults are shown in Fig. 11 [40]. solution, Radiat. Phys. Chem. 36 (1990) 511–516.
It is clear from this data that the size distribution of SL bubbles [24] A. Tauber, G. Mark, H.-P. Schuchmann, C. Von Sonntag, Sonolysis of tert-
butyl alcohol in aqueous solution, J. Chem. Soc., Perkin Trans. 2 (1999)
is relatively larger than that of SC bubbles and supports the spec- 1129–1135.
ulation that two groups of bubbles exist in a reactor. Those bubbles [25] M. Ashokkumar, F. Grieser, A comparison between multibubble
located in high energy regions can grow to relatively larger sizes sonoluminescence intensity and the temperature within cavitation bubbles,
J. Am. Chem. Soc. 127 (2005) 5326–5327.
and attain relatively higher temperatures to generate SL. These [26] J. Rae, M. Ashokkumar, O. Eulaerts, C. Von Sonntag, J. Reisse, F. Grieser,
bubbles are relatively lower in numbers compared to the sono- Estimation of ultrasound induced cavitation bubble temperatures in aqueous
chemically active bubbles as can be seen in Fig. 10. solutions, Ultrason. Sonochem. 12 (2005) 325–329.
[27] T. Vu, Effect of Organic and Inorganic Solutes on Multibubble
Sonoluminescence in Aqueous Solutions, M.Sc. Thesis, University of
7. Final remarks Melbourne, 2004.
[28] M. Ashokkumar, P. Mulvaney, F. Grieser, The effect of pH on multibubble
sonoluminescence from aqueous solutions containing simple organic weak
This overview is focused on only a few aspects of acoustic cav- acids and bases, J. Am. Chem. Soc. 121 (1999) 7355–7359.
itation bubbles. It is known that the presence of solutes [41] affect [29] R. Tronson, The Effect of Surface Active Solutes and Ultrasound Frequency on
Sonoluminescence in Aqueous Solutions, University of Melbourne, Ph.D.
the cavitation bubbles in a number of ways. One important issue
Thesis, 2004.
that still remains is the estimation of bubble populations. Iida [30] F. Burdin, N.A. Tsochatzidis, P. Guiraud, A.M. Wilhelm, H. Delmas,
and coworkers [42] have recently used pulsed laser diffraction Characterization of the acoustic cavitation cloud by two laser techniques,
Ultrason. Sonochem. 6 (1999) 43–51.
method to estimate the bubble population. A number of other
[31] W.S. Chen, T. Matula, L.A. Crum, The disappearance of ultrasound cavitation
groups [43–50] have studied the cavitation structures and other bubbles: observation of bubble dissolution and cavitation nuclei, Ultrasound
aspects that have not been discussed in this overview. However, Med. Biol. 28 (2002) 793–803.
the aspects that are covered in this article expose some simple [32] N.A. Tsochatzidis, P. Guiraud, A.M. Wilhelm, H. Delmas, Determination of
velocity, size and concentration of ultrasonic cavitation bubbles by phase-
experimental techniques that have been developed to characterize Doppler technique, Chem. Eng. Sci. 56 (2001) 1831–1840.
the cavitation bubbles, which will be highly beneficial to the son- [33] J. Lee, M. Ashokkumar, S. Kentish, F. Grieser, Determination of the size
ochemistry community. distribution of sonoluminescence bubbles in a pulsed acoustic field, J. Am.
Chem. Soc. 127 (2005) 16810–16811.
[34] R. Tronson, M. Ashokkumar, F. Grieser, Comparison of the effects of water
References soluble solutes on multibubble sonoluminescence generated in aqueous
solutions by 20- and 515-kHz pulsed ultrasound, J. Phys. Chem. B 106
(2002) 11064–11068.
[1] M. Ashokkumar, T. Mason, ‘‘Sonochemistry’’, Kirk-Othmer Encylcopedia of
[35] M. Ashokkumar, J. Lee, Y. Iida, K. Yasui, T. Kozuka, T. Tuziuti, A. Towata, The
Chemical Technology, John Wiley and Sons, 2007, doi:10.1002/
detection and control of stable and transient cavitation bubbles, Phys. Chem.
0471238961.1915141519211912.a01.pub2, Article Online Posting Date:
Chem. Phys. 11 (2009) 10118–10121.
October 19, 2007.
[36] R. Tronson, M. Ashokkumar, F. Grieser, Multibubble sonoluminescence from
[2] T.J. Mason (Ed.), Advances in Sonochemistry, vol. 1–6, Elsevier, Amsterdam,
aqueous solutions containing mixtures of surface active solutes, J. Phys. Chem.
The Netherlands, 1990–2001.
B 107 (2003) 7307–7311.
[3] F.R. Young, Cavitation, McGraw-Hill, London, 1994.
872 M. Ashokkumar / Ultrasonics Sonochemistry 18 (2011) 864–872

[37] D. Sunartio, M. Ashokkumar, F. Grieser, Study of the coalescence of acoustic [44] W. Lauterborn, T. Kurz, Physics of bubble oscillations, Rep. Prog. Phys. 73
bubbles as a function of frequency, power and water soluble additives, J. Am. (2010) 1065011–1065088.
Chem. Soc. 129 (2007) 6031–6036. [45] J. Lee, K. Yasui, T. Tuziuti, T. Kozuka, A. Towata, Y. Iida, Spatial distribution
[38] D. Sunartio, K. Yasui, T. Tuziuti, T. Kozuka, Y. Iida, M. Ashokkumar, F. Grieser, enhancement of sonoluminescence activity by altering sonication and solution
Correlation between Na⁄ emission and chemically active acoustic cavitation conditions, J. Phys. Chem. B 112 (2008) 15333–15341.
bubbles, ChemPhysChem 8 (2007) 2331–2335. [46] J.F. Guan, T.J. Matula, Understanding the pulse-to-pulse evolution of
[39] M. Ashokkumar, J. Lee, Y. Iida, K. Yasui, T. Kozuka, T. Tuziuti, A. Towata, Spatial microbubble clouds using light scattering, IEEE Ultrason. Symp. 1–3 (2004)
distribution of acoustic cavitation bubbles at different ultrasound frequencies, 218–221.
ChemPhysChem 11 (2010) 1680–1684. [47] N. Segebarth, O. Eulaerts, J. Reisse, L.A. Crum, T.J. Matula, Correlation between
[40] A. Brotchie, F. Grieser, M. Ashokkumar, Effect of power and frequency on acoustic cavitation noise, bubble population and sonochemistry, J. Phys. Chem.
bubble size distributions in acoustic cavitation, Phys. Rev. Lett. 102 (2009) B 106 (2002) 9181–9190.
0843021–0843024. [48] A. Otta, T. Nowak, R. Mettin, F. Holsteyns, A. Lippert, Characterization of a
[41] M. Ashokkumar, F. Grieser, The effect of surface active solutes on bubbles in an cavitation bubble structure at 230 kHz: bubble population, sonolumines-
acoustic field, Phys. Chem. Chem. Phys. 9 (2007) 5631–5643. cence and cleaning potential, Solid State Phenomena 145–146 (2009) 11–13.
[42] Y. Iida, M. Ashokkumar, T. Tuziuti, T. Kozuka, K. Yasui, A. Towata, J. Lee, Bubble [49] R. Mettin, A.A. Doinikov, Translational instability of a spherical bubble in a
population phenomena in sonochemical reactor: I estimation of bubble size standing ultrasound wave, Appl. Acoustics 70 (2009) 1330–1339.
distribution and its number density with pulsed sonication – laser diffraction [50] M. Arora, O.D. Ohl, D. Lohse, Effect of nuclei concentration on cavitation cluster
method, Ultrason. Sonochem. 17 (2010) 473–479. dynamics, J. Acoust. Soc. Am. 121 (2007) 3432–3436.
[43] W. Lauterborn, Cavitation bubble dynamics-new tools for an intricate
problem, Appl. Sci. Res. 38 (1982) 165–178.

Вам также может понравиться