Sei sulla pagina 1di 12

Ultrasonics Sonochemistry 18 (2011) 197208

Contents lists available at ScienceDirect

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

A calorimetric study of energy conversion efciency of a sonochemical reactor at 500 kHz for organic solvents
Maricela Toma a,*, Satoshi Fukutomi a, Yoshiyuki Asakura b, Shinobu Koda a
a b

Department of Molecular Design and Engineering, Graduate School of Engineering, Nagoya University, Furo-cho, Chikusa-ku, Nagoya, Aichi 464-8603, Japan Honda Electronics Co. Ltd., 20 Oyamazuka, Oiwa-cho, Toyohashi, Aichi 441-3193, Japan

a r t i c l e

i n f o

a b s t r a c t
It would seem that the economic viability is yet to be established for a great number of sonochemical processes, owning to their perfectible ultrasonic equipments. Industrial scale sonoreactors may become more important as a result of mastering the parameters with inuence on their energy balance. This work related the solvent type to the energy efciency as the rst step of a complex study aiming to assess the energy balance of sonochemical reactors at 500 kHz. Quantitative measurements of ultrasonic power for water and 10 pure organic solvents were performed by calorimetry for a cylindrically shaped sonochemical reactor with a bottom mounted vibrating plate. It was found that the ultrasonic power is strongly related to the solvent, the energy conversion for organic liquids is half from that of water and there is a drop in energy efciency for lling levels up to 250 mm organic solvents. Surface tension, viscosity and vapor pressure inuence the energy conversion for organic solvents, but it is difcult explain these ndings based on physical properties of solvents alone. The apparent intensity of the atomization process shows a good agreement with the experimentally determined values for energy conversion for water and the solvent group studied here. This study revealed that to attain the same ultrasonic power level, more electrical energy is need for organic solvents as compared to water. The energy balance equation has been dened based on these ndings by considering an energy term for atomization. 2010 Elsevier B.V. All rights reserved.

Article history: Received 18 December 2009 Received in revised form 14 May 2010 Accepted 17 May 2010 Available online 24 May 2010 Keywords: Sonochemical reactor Organic solvents Calorimetry Energy conversion Atomization Energy balance

1. Introduction It is important to recognize that a key issue to be addressed for every sonochemical process with potential industrial application is the efciency as compared with the classical pathway. Sonochemistry and its hybrid technologies are advanced techniques that rely upon the acoustic cavitation effects and yields. Cavitational eld intensity is known to be under the strong inuences of physical and chemical properties of the host solvent, treatment condition and ultrasonic irradiation characteristics [16]. It is the understanding of these parameters mode of interaction together with the optimization of the acoustic cavitation equipment that will nally authorize successful application of sonochemistry [7,8]. Given the concept of true and false sonochemical processes [9,10], there are two types of ultrasound applications: those based on the chemical effect (sonochemistry) and those based on the physical effects generated by bubble collapse (sonoprocessing). The entrenchment of power ultrasonic devices in industrial processing is an important gain but investigations that seek to scale-up for processes that take advantage on chemical effects of cavitation are always worthy of consideration in the endeavor to
* Corresponding author. Tel.: +81 52 789 3275; fax: +81 52 789 3273. E-mail address: maricela_toma@yahoo.com (M. Toma). 1350-4177/$ - see front matter 2010 Elsevier B.V. All rights reserved. doi:10.1016/j.ultsonch.2010.05.005

achieve a full recognition of sonochemistry in industry [11]. The eld of sonochemical equipment modeling has been active for many years but earlier efforts focused mostly on facilitating the chemistry rather than getting better acoustics parameters [8]. It is almost two decades ago that Bernard and Mason, assessing a large range of ultrasonic devices concluded that the equipment for industrial scale sonochemistry can easily be derived from the type of those existing at laboratory level [12]. Ultrasonic systems manufacturing was seen to follow the expansion of ultrasound applications by tailoring reactors to t new strategies [13] but the main goal: sonoreactors featuring high energy efciency, is not materialized yet. The energy conversion is known to be a critical factor in industrial applications. In order to lower the electric energy demands for the scaled-up version of sonochemical reactors, their energy balance is important to be throughly evaluated. To resolve the energy balance of the sonochemical reactors it is not a simple matter as there is a simultaneous input of mechanical and chemical energy into the reaction medium during sonication [12] and more rigorous and realistic characterization of the energy consumed to produce cavitations together with the thermal, viscous and radiation losses is needed. Instead, the sonochemical efciency (SE) concept [14] was introduced to determine the energy conversion toward the required effect in lieu of the calorimetrically measured acoustic

198

M. Toma et al. / Ultrasonics Sonochemistry 18 (2011) 197208

energy available in the sonicated medium. Beside being an elegant, reliable method to express the sonoreactors performance, SE (mol dm3)/(J dm3) is a basic parameter to compare different ultrasonic systems (the lack of a standard methodology in the past allowed many confusions when trying to correlate chemical with acoustical information). Whenever is the case of calibration or scale-up, the acoustic power (W) dissipated by the ultrasonic wave into the bulk medium is a key parameter required to express the efciency of a sonochemical reactor, also considered to have the main contribution on the sonochemical yield [15,16]. Therefore, a variety of investigative techniques has been proposed during the years to evaluate the output ultrasonic power. According to the type of cavitational outcome investigated, these methods can be categorized as following: (a) methods dealing merely with the physical effect of cavitation such as the thermal effect (calorimetry [1720]) and acoustic pressure (aluminium foils erosion [2123,40] or hydrophone probe [24]); (b) methods that are sensitive to radical species produced during solutions sonolysis (chemical dosimetry [2527] and electrochemical dosimetry [2830]), and (c) methods that combine the above mentioned effects [31]. However, comparative tests of calorimetry and chemical dosimetry carried out on the same sonoreactors demonstrated that these two methods provide similar prediction of the ultrasonic power [32,33]. The energy conversion in sonochemistry is strongly dependent upon the type and operation mode of the sonication equipment and sonication medium involved [10,31,34,35]. Nevertheless, another critical parameter to be assessed for optimizing or for designing reactors with adequate ultrasonic elds is the energy distribution on the sonication volumes [36,37]. There are indications that by controlling the geometric and operating conditions of the sonoreactor, the required intensity of cavitation can be achieved with maximum energy efciency [38,39]. The height of the sonoreactor is regarded to play an important role since it was reported that there is an axial and radial distribution of the cavitational eld intensity [4044]. Having achieved an increased understanding over many aspects of sonoreactors efciency [14,4549] our research group is aiming at developing industrial scale sonochemical reactors with high ultrasonic efciency. In pursuing this aim, a special attention was paid at easing the acoustics admittance in the benet of chemistry [48,49]. The efciency of energy conversion from electricity to ultrasound was 70% for the tower type sonoreactor that we developed in a previous work [47]. Therefore, we were interested to determine which changes in the reactor sonication medium might affect the energy conversion. Since the relative signicance of the parameters that inuence energy conversion depends upon the physical and chemical characteristics of the liquid being irradiated [34], we were interested to obtain informations about the energy conversion for organic solvents. Back in 1965, it was revealed by Weissler et al. [50] in their pioneering study that cavitation could occur in organic solvents at 800 kHz and 3 W/cm2. The non-aqueous sonochemistry is an active research eld related to polymer chemistry and solvent decomposition [1,51] but information concerning ultrasonic energy conversion in organic solvents is sparse. This study focused on the investigation regarding the ultrasonic energy dissipation into various organic solvents sonicated at 500 kHz in a cylindrically shaped sonoreactor with a bottom mounted transducer.

pipe (4 mm thick and 58 mm inner diameter) vertically mounted on a 0.1 mm stainless steel (SUS304) vibrating plate attached to a PZT disk type transducer with a resonance frequency of 490 kHz and 50 mm diameter, manufactured by Honda Electronics Co., Ltd. A fan was incorporated into the transducer housing to avoid the over-heating during the sonication. The reactor was congured to accept a 500 mm liquid column, and equipped with a lid in order to minimize the solvent evaporation during sonication. The measurements at 20 kHz frequency were performed by employing a Langevin type transducer in connection with a 2 mm vibration plate. The transducer was driven by a continuous sinusoidal wave produced by a signal generator (W1942, NF Corp.) and delivered through a high-frequency power amplier (L400BM-H, Honda Electronics Co., Ltd.). The effective electric power input to the transducer was accurately calculated from the voltage at both ends of the transducer and the current as measured with an oscilloscope (TDS3012B, Tektronix Inc.) connected to a current probe (TCP202, Tektronix Inc.). The temperature measurements were performed using two kinds of thermocouples (Sheath T-type with a diameter of 2.3 mm, Takahashi Thermo Sensor Ltd. or copperconstantan with a diameter of 0.2 mm, Omega Engineering Inc.) and a thermometer (NR-500, NR-TH08 Keyence Corp.) connected to a personal computer equipped with an in-house developed software for post processing. Precise movements between two successive measurement positions were provided by a SGSP33-200 Sigma Koki motorized 3D stage. The scheme shown in Fig. 1a illustrates the experimental set-up details. 2.2. Reagents The chemicals were purchased as follows: ethanol (99.5%), toluene (99.5%), cyclohexane (99.5%), propanol (98.5%) methanol (99.5%), benzene (99.5%), xylene (99.5%) and 1,2,3,4-tetrahydronaphthalene (tetralin) (99.5%) from Chameleon Reagent; hexane; n-butanol (99.5%) from Nacalai Tesque and decane from Wako Pure Chemical Industries Ltd. All solvents used here were analytical grade and were used without extra purication. Distillate water (conductivity less that 0.1 mS/m) was prepared on the laboratory site employing a Yamato Autostill WG 25 system. Prior to sonication, samples were air-saturated by bubbling the air for a period of 30 min at 25 1 C to provide uniform distribution for the gas in the bulk liquid as a precaution raised from the known inuence of bubble dynamics on the liquid media [52]. Moreover, in order to facilitate the gas transfer into the relatively large sample volumes handled here (ranging from 100 to 1500 ml), solutions were also slowly stirred during the purging for reproducibility concerns. Precautions were taken to avoid solvents temperature variation prior the measurement starting point. 2.3. Quantication of dissipated ultrasonic power The procedure we adopted here for acoustic power measurements was the calorimetric method. Based on the assumption that the mechanical energy generated by the ultrasonic waves is reduced to heat, the dissipated ultrasonic power Up was calculated from the rate of temperature increase as:

Up C p M

dT dt

2. Method and materials 2.1. The sonochemical reactor system The ultrasonic equipment used for this study was a custombuild sonoreactor whose design incorporates a cylindrical glass

where Cp is the heat capacity of the solvent at constant pressure (J kg1 K1), M is the mass of solvent (kg) and dT/dt is temperature rise per second [53]. The choice of the quantication method was based upon the need to make a correct comparison on acoustic power for the solvents involved in this work. We found it desirable to avoid the experimental errors arising from using solvent specic chemical

M. Toma et al. / Ultrasonics Sonochemistry 18 (2011) 197208

199

A
Cylindrical glass vessel

58

Vibration plate

Signal generator

Transducer

Cooling fan

Current probe Oscilloscope Power amplifier

B
Motorized 3D stage Thermometer
PC

readings were taken within the bulk liquid for the rst 100 s sonication with a thermocouple reproducibly positioned at the middle of the liquid sample and 15 mm apart from the central axis as to avoid the larger intensity reported from the reactor centerline [39]. The measurements were repeated three times and their means were used in calculations. For an accurate picture of the liquid height inuence on the ultrasonic power dissipation, the temperature measurements have been carried out stepwise, while the lling levels were gradually varied with a constant increment of 10 mm in a range from 10 mm up to 500 mm. It is worthwhile noting that the reactor cell was relled with a fresh portion of air-saturated solvent prior to each measurement to assure reproducibility. To simplify the ultrasonic waves environment, the measurements for this study where carried out without a reector. It is known that this can affect the topology of sound and bubble eld in the reactor [8,56]. Previous studies have shown that the calorimetric method is rather independent of the bulk liquid temperature (below 50 C), reactor shape and insulation, and concluded that the barely discernible heat gains or losses to the experimental environment are negligible for the short period of measurement [18,32,35,55]. Accordingly, thermal insulation was omitted for the sonoreactor glass cell. All measurements were made at ambient temperature and pressure. The details of temperature measurement set-up are depicted in Fig. 1b. 3. Results and discussion High-frequency ultrasound (200600 kHz) generates chemically active bubbles but their applications came much later as compared to the low frequency ultrasound, owning to their inefciency in processing applications. However, this ultrasonic region became intensively approached in scale-up studies soon after their effectiveness for pollutants decomposition was revealed [57]. The energy conversion is known to be a critical factor in industrial applications. It should be emphasized, however, that the energy terms for ultrasonic systems are not always evident and the models reported in literature are likely to be applicable to a specic set of operating parameters. This work focused on the effect of liquid, electrical power and liquid height on energy efciency at 500 kHz. 3.1. The inuence of solvent type on ultrasonic power Previous studies on the effect of changes in reactor shape and solvent characteristics on the energy conversion and power density suggest that carefully investigation are required to understand the system in order to evaluate its energy balance [12,42]. Thus, the dependence of energy conversion on the type of solvent will be addressed herein. Although the liquid properties inuencing cavitational events are well studied, the sonochemical literature has little to tell about the liquid parameters that inuence the power dissipation, especially for the non-aqueous systems. The liquid selection for this study included distilled water and pure organic solvents, namely: ethanol, butanol, propanol, hexane, cyclohexane, toluene, benzene, tetralin, xylene and pentadecane. Seeking for accurate informations on the energy conversion for organic solvents, comparative tests were conducted at 500 and 20 kHz. The ultrasonic power has been estimated calorimetrically from the initial temperature rise (dT/dt) that gives a reasonable indication of the quantity of energy effectively dissipated into the sonicated liquid. At this point, the measurements were carried out for 100 ml liquid as the electrical power supplied to the transducer was set constant at 30 W. The temperature rise dependence on the type of liquid under sonication at 500 and 20 kHz is depicted in Fig. 2. The plots clearly show that for both frequencies the slope of temperature rise depends strongly on the liquid

Lid
Liquid height: 30-500 mm

Thermocouples

x
Fig. 1. Experimental set-up for calorimetric measurements of ultrasonic power dissipated: general assemble (A) with details on the reactor unit (B).

dosimetry methods. At the moment, the literature offers few data about the dosimetry in non-aqueous solvents [43,54,55] and a general method for capturing and quantifying the cavitational effects on organic solvents has not yet been validated. Despite the reported drawbacks (the convective cooling measurements, the heating of transducer and the sensor that may disturb the measurements), calorimetry is an universal, yet precise method for quantifying the acoustic power [18]. This method, however, does provide a way of determining the power dissipation based on the dependence of the temperature rise (dT/dt) upon the solvent acoustic and thermal properties. This sonoreactor had previously been characterized by alternative measurement techniques, such as KI dosimetry or hydrophone probe and the details can be seen in Ref. [48]. 2.3.1. Experimental procedures 2.3.1.1. Electrical power input measurements. By using the above experimental set-up, the electric power change could be held within 0.5 W by continually tuning the voltage at both ends of the transducer as observed on the oscilloscope display reading. 2.3.1.2. Temperature measurements. A set of experiments was carried out by recording the temperature evolution under sonication for every organic solvent investigated and for water. Temperature

200

M. Toma et al. / Ultrasonics Sonochemistry 18 (2011) 197208

38 36 34 32 30 28
To

Temperature [oC]

Water T etralin Ethanol Butanol Hexane Benzene Propanol Cyclohexane Pentadecane T oluene Xylene

To

Temperature [ C]

36 32 28 24

Water Ethanol P ropanol Butanol Cyclohexane

26 24 0 50

50 Time [s]

100

150

100

150

200

Time [s]
Fig. 2. Experimental curves of temperature evolution vs. time for different solvents at 500 kHz. Depicted in the graph inlet are the experimental curves at 20 kHz.

investigated and there is a signicant difference between the water and the non-aqueous medium. Immediately after the ultrasonic source was turned on, a dynamic linear rise in temperature (DTo) was observed for the rst ms sonication at high-frequency. This behavior is attributed to the solvent shear viscosity stress [59]. During the experimental work carried out for this study, we noticed that DNo depends on the viscosity of the liquid being sonicated, ultrasonic frequency and liquid height. In examining the inuence of liquid on the ultrasonic power dissipated into de sonication medium, we found that water accommodates more ultrasonic power than the organic solvents as shown in Fig. 3 and this difference is more evident at 500 kHz as compared to 20 kHz. 3.1.1. Inuence of physical properties Contrary to what one might intuitively expect, our experimental data show little dependence on the acoustic impedance. Viscosity, surface tension and vapor pressure are among the most

important variables of the sonicated medium to inuence the ultrasound effect [35,58,60]. As shown in Fig. 4, the plots of ultrasonic power at 500 kHz against the liquids physical properties (calculated from data in Ref. [61]) are in keeping with the previous ones. The data was distributed into two distinct groups: organic liquids and distillate water indicating again a signicant difference between water and the organic solvents investigated here. The ultrasonic power for organic solvents decreases as following: hexane > ethanol > benzene > xylene > toluene > propanol > butanol > tetralin > cyclohexane. The difference in viscosity cannot explain the difference seen between water and organic solvents ultrasonic power values since water and cyclohexane have similar viscosity. On the other hand, it would seem that the ultrasonic power in the organic solvents group decreases for solvents with elevated viscosity such as propanol, butanol and tetralin. Vapor pressure is an important property that controls the cavitational activity in solvents [63]. Hexane and cyclohexane show

500kHz
30 25
30 25 20 15 10 5 0

20kHz

Ultrasonic power [W]

20 15 10 5 0

at e Et r ha no H l ex a Be ne nz en e X yl en To e lu e Pr ne op Pe an nt ad ol ec an Bu e ta no Te l Cy tra l cl oh in ex an e

er

no l

no l

no ta Cy

op a

Bu

Et

Pr

Fig. 3. Dependence of ultrasonic power on solvent at 20 and 500 kHz.

cl

oh

ex an

at

ha

M. Toma et al. / Ultrasonics Sonochemistry 18 (2011) 197208

201

30

Water Xylene Ethanol Butanol Propanol Hexane Cyclohexane Benzene Toluene Tetralin

Ultrasonic Power [W]

25 20 15 10 5 0 0 5 10 15 20 25 0.5 1.0 1.5 2.0 2.5 20 30 40 50 60 70 80

Vapor Pressure [kPa at 298K]

Viscosity [mPas]

Surface Tension [mNm-1]

Fig. 4. The inuence of vapor pressure, viscosity and surface tension on the ultrasonic power.

close values for vapor pressure and surface tension, but the viscosity of hexane is three time lower than those of cyclohexane and this can be a reason for their difference in the ultrasonic power. The ultrasonic power dissipated decreased for solvents with high volatility such as tetralin and butanol but again, it is hard to explain the difference between water and the organic solvents. A variation in the dissipated ultrasonic power with their surface tension values can be seen for water and organic solvents. These results suggested that it is difcult to explain the energy dissipation in organic solvents by looking at a particular physical property. Instead, it would be logical to consider a complex interference of liquid viscosity, surface pressure and vapor pressure together with their variation during the sonication responsible for the energy conversion. Since the surface tension (another colligative property) shows a good t with the ultrasonic power variation among the investigated liquids, we looked at the atomization effect for a better explanation of our experimental results. 3.1.2. Inuence of ultrasonic atomization The uctuation of the atomization intensity with the liquid and sonication parameters detected by visual observations made during the calorimetric measurements also suggested to us a certain inuence of atomization on the dissipated ultrasonic power. Therefore, a detailed description of the solvent atomization that takes place during the calorimetric measurements for the ultrasonic power will be presented here. Ultrasonic excitation of the solvent results in a break of capillary surface waves followed by a liquid protuberance and a subsequent expulsion of drops (atomization fountains). It is known that the instability of capillary surface waves increases at high-frequency owing to the difference in ultrasonic beam directivity [52] and that the drop size is small at high frequencies where the capillary wavelength is short and vice versa [6264]. This can explain the fact that the discrepancy between the ultrasonic power for water and organic solvents is more reduced at 20 kHz as compared to 500 kHz (Fig. 4). An illustration of the capillary wave disintegration at 500 kHz during a typical experimental run for calorimetric measurements is shown in Fig. 5. The sonication resulted in large volumes of ultrasonic mist from cyclohexane, ethanol and toluene where liquid splashes were also observed for the last two (Fig. 5A). Decane and tetralin displayed vigorous splash fountains while a gently splashing fountain was observed for butanol but the capillary waves for water never teared up to form drops under the experimental conditions employed here. It is worth mentioning that the difference in the drop size can be clearly attributed to the dif-

ference in vapor pressure and surface tension as a heavy mist layer was formed for solvents with high vapor pressure and low surface tension. However, it is unclear why decane, a solvent with low values of surface tension, viscosity and density developed the higher liquid crest with large sized drops but no mist. We assumed that not only the mist formation but also the liquid splashes have a sure contribution to the temperature decrease during the measurement period. These experimental results made obvious the dependence of the ultrasonic power dissipated into the solvents at high-frequency on the colligative properties (viscosity and surface pressure) and vapor pressure rather than on acoustic impedance. This nding is in good agreement with the dependence shown by other authors [35,58]. The fact that the power transferred into the sonicated medium decreased at high volatilities and high viscosities of the liquid, cleared the link between the energy efciency and the atomization. 3.2. The inuence of electrical power delivered to the transducer on the ultrasonic power The electrical power supplied at transducer was varied in a range from 10 to 50 W. As it was expected, the ultrasonic power increases linearly with the electric power but there is a signicant difference in the curve slope for water and organic solvents as shown in Fig. 6. The slope is 0.82 for water and about 0.30.4 for organic solvents indicating that the energy base efciency for solvents is half compared to the efciency of water. This is in agreement with a low energetic yield for toluene sonicated at 40 W electrical power reported previously [55]. The energy based efciency has been calculated from ultrasonic power (Pu) and electrical power (Pe) delivered at transducer as following:

ge

Pu Pe

The relative energy efciency has been also calculated from:

gr

gs gw

where gw is the energy efciency for water and gs is energy efciency for organic solvents. The experimental results of the inuence of liquid on energy conversion at 500 kHz and for comparison, at 20 kHz are shown in Table 1. The calculated energy efciency is half of that for water under the same sonication parameters or less than half as for the rest of solvents measured here.

202

M. Toma et al. / Ultrasonics Sonochemistry 18 (2011) 197208

Fig. 5. An illustration of liquid atomization of the solvents during the calorimetric measurements: (A) the dependence on solvent; (B) the inuence of electrical power supplied at transducer on atomization of tetralin and distillate water and (C) the inuence of liquid height on vaporization.

It is important to emphasize at this stage that an increase in electrical power is followed by an increase in atomization intensity, as well (Fig. 5B). The apparent intensity of the atomization process is in good agreement with the experimentally determined values for ultrasonic power dissipated in water and the solvent group. Thus, the energy conversion is not automatically a pure electrical power input effect since the ultrasonic atomization is also involved and consequently, a part of the input energy is lost due to atomization. Low cavitation based efciencies were previously reported for organic solvents as shown in Table 1 but it is difcult to compare the results for solvents individually since the ultrasonic frequency is different. These low energy efciencies correspond well with those found in the sonolysis of organic compounds [51,63]. It is known that more energy is needed to achieve a certain cavitational intensity in organic solvents than in water, owning

to the difference in cavitation threshold [65]. Here it was found that to attain the same ultrasonic power level, more electrical energy is needed at transducer in the case of organic solvents than for water. 3.3. The inuence of liquid height on the ultrasonic power Webber and Chon [39] were rst to observe variation of the acoustic power density along the sonication volumes with a pattern of the standing waves that were later conrmed in many papers. For more information on the atomization inuence on energy conversion we nd it vital to investigate on the inuence of liquid height on dissipated acoustic power. Quantitative measurements of absorbed acoustic energy were carried out with different levels of liquid in reactor cell for water and organic solvents. The electrical power input to the transducer

M. Toma et al. / Ultrasonics Sonochemistry 18 (2011) 197208

203

50 45 40
Water Ethanol Propanol Butanol Hexane Cyclohexane Pentadecane Benzene Toluene Xylene Tetralin

35 30 25 20 15 10 5 0 0

10

20

30

40

50

Electrical Power [W]


Fig. 6. The inuence of electrical power on the ultrasonic power dissipated into the sonicated medium for water and organic solvents.

was set constant at 30 W over the measurements period and the amount of acoustic power that a particular solvent can accommodate was measured for water, tetralin, butanol, hexane, decane, ethanol and cyclohexane.

3.3.1. Local vs. overall temperature investigation Calorimetry is a method based on bulk liquid temperature and assumes uniformity in the liquid temperature. However, it has been shown that the calorimetric data are compromised by high viscous media and insufcient convection [58]. A characteristic of sonoreactors operating with bottom mounted transducers is the location of the active zone in a cylindrically shaped hollow dened by the directivity of the ultrasonic beam [44,49]. The sonochemical reactor employed here has the advantage of hosting the bulk liquid in this region since the difference between the diameter of transducer and the diameter of glass cell is only 8 mm. Moreover, our previous measurements on the velocity prole of the ultrasonic streaming [49] are indicating high liquid velocities: 11.8 cm s1 for this zone. Boldo et al. [44] have also reported a well mixing effect for the bulk solvent during the sonication in a similar sonoreactor as a result of the convective currents in association with the cavitational hydrodynamic phenomena. In order to get a valid temperature value of the bulk liquid, the measurement location was set intuitively at the sample height median point (h/2) and 10 mm apart from the reactor centerline were the liquid velocity is optimal. However, a concern about temperature uniformity for the relatively large volumes and heights of liquid employed led us to investigate the local temperature variations for this point. We also suspected that the signicant acoustic pressure variations at positions hosted in the near-eld and close to the transducer can affect the temperature data along the liquid

Table 1 Calculated values for the energy based efciency. Solvent

Ultrasonic Power [W]

ge
500 kHz 20 kHz 0.27 0.21 0.23 0.20 0.19

gr
500 kHz 0.39 0.41 0.5 0.46 0.41 0.41 0.40 0.36 0.41 0.51 20 kHz

ge
500 kHz 10300 mm 0.74 0.56 0.57 0.59 0.52 0.59 0.49 300500 mm 0.68 0.59 0.62 0.81 0.70 0.77 0.61

Cavitation efciency* % from water

Water Tetralin Butanol Decane Ethanol Toluene Cyclohexane Propanol Xylene Benzene Hexane
*

0.82 0.32 0.34 0.41 0.38 0.34 0.34 0.33 0.30 0.34 0.42

0.77 0.85 0.75 [55] 0.74 0.70

70 43 46 71 42 64 43

Cavitational based efciency at 46 kHz (water = 100) when controlled at k/2 from Ref. [60].

0.045 0.04 0.035 0.03

bottom middle top average value

dT/dt

0.025 0.02 0.015 0.01 0.005 0 0


100 200 300 400 500

Liquid height [mm]


Fig. 7. Experimental data of calorimetric measurements at three different points of the sample and their variation along the liquid height.

204

M. Toma et al. / Ultrasonics Sonochemistry 18 (2011) 197208

33 32 31

50mm 100mm 150mm 200mm 250mm 300mm 350mm 400mm 450mm

30 29 28 27 26 25 24 0 20 40 60 80 100

120

Time [s]
Fig. 8. The temperature radiated by the transducer body during measurements for different liquid height.

column. Thus, the temperature variation along the cylinder central axis has been measured.

Temperature readings were made simultaneously at different locations within the bulk liquid with three sensors located as following: (a) at 15 mm from vibrating plate, (b) in the middle of liquid column and (c) at 15 mm beneath the liquid surface. The temperature raise (dT/dt) for these locations as a function of liquid height can be seen in Fig. 7 together with the average value. As we expected, the albeit small difference in the temperature distribution along the liquid height increases with increasing the liquid volume. The temperature at the sonochemical reactor bottom stays relatively distant from the averaged value, compared with the top and middle positions, thus being less inuenced by the liquid height. A rst explanation for this behavior is an additional input of heat from the transducer body and to conrm this hypothesis, the temperature radiated from the transducer body was been measured. The results indicate that the temperature change upon the liquid height are in the range of 0 to 1 C and can be considered negligible (Fig. 8). As this cannot explain the bottom temperature variation with the liquid height especially after 250 mm, we assumed that the temperature at this location is controlled mainly by the frictional heat between the thermocouple or the liquids frictional heat of vibrating particles in the thermocouple. In an effort to better understand the mixing effect, the local temperature variations considered as Ds = s(n+1) sn where sn is

Temperature [ oC]

A
0.375 0.325 0.275 0.225 0.175 0.125 0.075 0.025 -0.025 0 -0.075 20

NEAR-FIELD

50mm 100mm 150mm 200mm 250mm

40 Time [s] FAR-FIELD

60

80

100

B
0.375 0.325 0.275 0.225 0.175 0.125 0.075 0.025 -0.025 -0.075 0 20

300mm 350mm 400mm 450mm 500mm

40 Time [s]

60

80

100

Fig. 9. Experimental data of local variation of temperature at the calorimetric measurements points for: A near-eld locations and B far-eld locations.

M. Toma et al. / Ultrasonics Sonochemistry 18 (2011) 197208

205

the liquid temperature and s(n+1) is the consecutive temperature recorded after one second at h/2 location were plotted against the time. The near-eld for this reactor was calculated to be located in rst 300 mm near the transducer [48] but there is no signicant difference in the temperature variation recorded in the near-eld locations (Fig. 9A) as compared to the far-eld (Fig. 9B), except for the rst 10 s, when they are signicantly larger. The results show that the temperature variation around thermocouple position is hold within 0.03 C under the experimental conditions employed here and support our presumption that an efcient mixing process takes place in the reactor cell as a result of strong convection currents made possible by the reactor design. The signal recorded over the rst 10 s exhibited turbulent variations also attributed to viscosity stress associated phenomena. Therefore, the temperature rise for the rst 10 s of sonication was ignored in the calculations of the dissipated acoustical power. Considering these experimental evidences in support of the above assumption, all the correlations developed in the present

work are based on the fairly uniform distribution of the temperature in the available sample volume. The sample height in the sonochemical reactor cell was varied from 30 to 500 mm and an average value of dissipated ultrasonic power for different levels along the sonoreactor cell was calculated. The inuence of liquid height on the dissipated acoustic power, together with their average value is illustrated individually in Fig. 10. The results indicate that the average values do not vary much between the organic solvents and water and this is a departure from what was observed for 100 ml samples (shown in Section 3.1). It has been reported that the cavitation intensity in organic solvents at 46 kHz is strongly inuenced by the liquid height and it is possible to induce in some organic liquids cavitation of intensity almost equal to that achieved in water by modeling the thickness of the irradiated liquid layer in terms of k/2 [60]. This pattern of acoustic energy density involving the standing waves has been reported by many authors in investigations concerning with the

30 25 20 15 10 5 0

Water

average value

100

200 300 400 Liquid height [mm] Tetralin

500

30 25 20 15 10 5 0

Ethanol

Up [W]

Up [W]

average value

100

200 300 400 Liquid height [mm] Toluene

500

Up [W]

15 10 5 0

Up [W]

30 25 20

average value

100

200 300 400 Liquid height [mm] Cyclohexane

500

30 25 20 15 10 5 0

average value

100

200 300 400 Liquid height [mm] Decane

500

30 25 20 15 10 5 0

average value

100

200 300 400 Liquid height [mm]

500

30 25 20 15 10 5 0

Up [W]

Up [W]

average value

100

200 300 400 Liquid height [mm]


water butanol tetralin ethanol

500

30 25 20 15 10 5 0

Butanol Up [W]

50 40 30 20 10

toluene cyclohexene decane

Up [W]

atomization

average value

100

200 300 400 Liquid height [mm]

500

100

200 300 Liquid height [mm]

400

500

Fig. 10. The inuence of liquid height on the ultrasonic power dissipated in organic solvents.

206

M. Toma et al. / Ultrasonics Sonochemistry 18 (2011) 197208

effect of liquid height on the ultrasonic systems efciency [28,41]. Also, there are evident uctuations around their average values along the reactor cell lling levels, but our experimental data are in disagreement with the standing wave pattern of this system. An explanation for this can be found in the different measurement methods or in the difference of the sonication parameters. Most of the reported data focused on cavitational energy or were recorded by screening the energy distribution in the system after the equilibrium was established but our investigation focused on with the energy conversion based on the rate of initial temperature rise. It is know that heating is a bulk effect, whereas cavitation is regarded as a local effect [58]. Furthermore, the variations we

recorded cannot be clearly associated with a power density distribution in a standing wave pattern under the experimental condition employed here since the k/2 is only 1.52 mm and the ultrasonic power is relatively high. Thus, the experimental data variation along the liquid column can be only attributed to the perfectible experimental condition. It is expected that for a certain input of electric power, the ultrasonic power dissipated into the liquid medium would diminish as the sample height increases. However, this behavior can be seen only for water as there is an opposite trend for the organic solvents. It is also observed that there is a clear difference in the ultrasonic power dissipated in water and organic liquids for lling levels

Water

Liquid height: 30-250mm


1.4

Liquid height: 250-500mm


1.4 1.35 1.3 1.25 1.2 1.15
Toluene Decane Butanol Tetralin Ethanol Cyclohexane

Log ultrasonic power[W

1.35 1.3 1.25 1.2 1.15 0 1 2


o

2
o

Viscosity [mPa s]
Liquid height 30-250mm
1.4
1.4

Viscosity [mPa s]
Liquid height 250-500mm
Water Toluene

Log ultrasonic power [W

1.35 1.3 1.25 1.2 1.15 0 2 4 6 8 10 12 14 16

1.35 1.3 1.25 1.2 1.15 0 2 4 6 8 10 12 14

Decane Ethanol Butanol Tetralin Cyclohexane

16

Vapor pressure[kPa at 298K]


Liquid height 30-250mm
1.4

Vapor pressure[kPa at 298K]


Liquid height 250-500mm
1.4
Toluene Water

Log ultrasonic power [W

1.35 1.3 1.25 1.2 1.15 0 20 40


o

1.35
Decane

1.3 1.25 1.2 1.15


60 80

Ethanol Butanol Tetralin Cyclohexane

Surface tension [mN m-1]

20 40 60 Surface tension [mN m-1]


o

80

Fig. 11. The averaged ultrasonic power from two regions of liquid height: 30250 mm and 250500 mm as a function of viscosity, vapor pressure and surface tension.

M. Toma et al. / Ultrasonics Sonochemistry 18 (2011) 197208

207

located in the rst 250 mm of the reactor cell. Fig. 10 contains also a comparative plot emphasizing on the difference between the solvents and water. A plausible explanation for this difference is the atomization phenomena for water and organic solvents as the atomization intensity is strongly inuenced by solvents height [61]. Fig. 5C shows the inuence of liquid height on atomization observed in the case of ethanol. It can be seen that as the liquid level in column increases, the liquid protuberance gradually diminishes and it is no longer visible for levels that exceed 300 mm. However, it is worthwhile to notice that a disturbance of the upper surface is visible at every lling level of the reactor for all the particular liquids investigated here. We observe that there is a different variation in the energy conversion for organic liquid and water along the reactor cell. As shown in Table 1, the conversion values for a particular liquid are different in the 30250 mm segment, where the atomization is an active process as compared to the 250500 mm segment of reactor cell. Furthermore, the energy data plotted against the viscosity, vapor pressure and surface tension (Fig. 11) individually for these two column segments also suggest an inuence of atomization on power conversion. Good correlations are found between energy conversion for toluene and cyclohexane and their viscosity. Their low viscosity can be responsible for the decrease of ultrasonic energy due to atomization at low lling levels as compared with tetralin, a viscous solvent that show little variation with the sample height. The high volatile cyclohexane and ethanol show a bigger difference in the energy absorbed for the upper and lower part of the cell as compared to tetralin. It is the difference in surface tension of water and organic solvents that can be seen as a reasonable explanation for the energy conversion variation from the lower to the upper part of the reactor cell due to atomization. Based on these evidences, we consider that the assumable decline in the ultrasonic power with increasing the sample height has been reversed by a drop in the dissipated ultrasonic power due to solvents atomization at low lling levels of the reactor cell. As we expected, this study also conrmed the inuence of atomization on the energy conversion of organic liquids. Therefore, it would appear reasonable to consider a certain inuence of ultrasonic atomization on the energy balance for organic solvent and high-frequency systems. In dening the energy balance for a sonochemical reactor running on organic solvents at high-frequency, an energy term for atomization should be considered. Based on the observations of this work, the energy balance is considered as:

ric method, it was possible to study the effects of the type of sonicated liquid, electrical power and liquid height on the ultrasonic power. This experimental work revealed a difference in the energy efciency between water and the organic solvents. The ultrasonic power conversion for organic liquids depends on the viscosity, surface pressure and vapor pressure rather than on acoustic impedance but the energy conversion for organic solvents cannot be attributed to a particular physical property as there is a combined inuence from these parameters. Under the experimental conditions employed here, the energy conversion for organic solvents at 500 kHz is only half or less from the energy conversion calculated for water. Contrary to what may be expected, there is a drop in the energy conversion efciency in organic liquids for low lling levels as compared to the 250500 mm segment of the reactor cell. All these ndings appear to indicate a strong impact of ultrasonic atomization on the energy conversion for organic solvents. Therefore, an energy term for atomization was considered for the energy balance of a sonochemical reactor running on organic liquids at high-frequency. The interest in this topic arises from our previous work on sonochemical reactors efciency and the ndings of these experiments have inspired the development of a research project to quantify the inuence of atomization on the energy balance at high-frequency. References
[1] A.A. Atchley, L. Crum, Acoustic cavitation and bubble dynamics, in: K.S. Suslick (Ed.), Ultrasound: Its Chemical, Physical and Biological Effects, VCH Publishers, Inc., 1988, pp. 162. [2] L.H. Thompson, L.K. Doraiswamy, Sonochemistry: science and engineering, Ind. Eng. Chem. Res. 38 (1999) 12151249. [3] Y.G. Adewuyi, A. Babatunde, A. Oyenekan, Ind. Eng. Chem. Res. 46 (2007) 411 420. [4] E.A. Neppiras, Acoustic cavitation, Phys. Rev. Lett. 61 (1980) 159251. [5] O. Behrend, H. Shubert, Inuence of hydrostatic pressure and gas content on continuous ultrasound emulsication, Ultrason. Sonochem. 8 (2001) 271276. [6] I. Hua, M.R. Hoffmann, Optimization of ultrasonic irradiation as an advanced oxidation technology, Environ. Sci. Technol. 31 (1997) 2237. [7] V. Renaudin, N. Gondrexon, P. Boldo, C. Ptrier, A. Bernis, Y. Gonthier, Method for determining the chemically active zones in a high-frequency ultrasonic reactor, Ultrason. Sonochem. 1 (1994) S81S85. [8] C. Horst, Y.S. Chen, U. Kunz, U. Hoffmann, Design, modeling and performance of a novel sonochemical reactor for heterogeneous reactions, Chem. Eng. Sci. 51 (1996) 18371846. [9] M. Chanon, J.-L. Luche, Sonochemistry: quo vadis, in: J.-L. Luche (Ed.), Synthetic Organic Chemistry, Plenum, 1998, pp. 376392. [10] L.A. Crum, Comments on the evolving eld of sonochemistry by a cavitation physicist, Ultrason. Sonochem. 2 (1995) S147S152. [11] T.J. Mason, Recent trends in sonochemistry and ultrasonic processing, in: 11th Meeting of the European Society of Sonochemistry, Program and Book of Abstracts, 2008, p. 37. [12] J. Berlan, T.J. Mason, Sonochemistry from research laboratory to industrial plant, Ultrasonics 30 (1992) 203212. [13] G. Cravotto, E.C. Gaudino, L. Boffa, J.-M. Lvque, J. Estager, W. Bonrath, Preparation of second generation ionic liquids by efcient solvent-free alkylation of N-heterocycles with chloroalkanes, Molecules 13 (2008) 149 156. [14] S. Koda, T. Kimura, T. Kondo, H. Mitome, A standard method to calibrate sonochemical efciency of an individual reaction system, Ultrason. Sonochem. 10 (2003) 149156. [15] A.G. Chakinala, P.R. Gogate, A.E. Burgess, D.H. Bremner, Intensication of hydroxyl radical production in sonochemical reactors, Ultrason. Sonochem. 14 (2007) 509514. [16] S.I. Nikitenko, C. Le Naour, P. Moisy, Comparative study of sonochemical reactors with different geometry using thermal and chemical probes, Ultrason. Sonochem. 14 (2007) 330336. [17] J. Zieniuk, R.C. Chivers, Measurement of ultrasonic exposure with radiation force and thermal methods, Ultrasonics 14 (1976) 161172. [18] C.J. Martin, A.N.R. Law, The use of thermistor probes to measure energy distribution in ultrasound elds, Ultrasonics 18 (1980) 127133. [19] M.A. Margulis, I.M. Margulis, Calorimetric method for measurement of acoustic power absorbed in a volume of a liquid, Ultrason. Sonochem. 10 (2003) 343345. [20] I.G. Mikhailov, Methods of measuring the absolute intensity of ultrasonic waves in liquids and solids, Ultrasonics 2 (1964) 129133. [21] V. Saez, A. Fras-Ferrer, J. Iniesta, J. Gonzalez-Garca, A. Aldaz, E. Riera, Characterization of a 20 kHz sonoreactor. Part I: analysis of chemical effects by classical and electrochemical methods, Ultrason. Sonochem. 12 (2005) 5962.

Ielectrical Idissipated Iatomisation Itransducer Ilost

where Ielectrical is the electrical power applied at transducer, Idissipated is the ultrasonic power accommodated into the sonicated liquid, Iatomisation is the power lost through the liquid atomization, Itransducer is the heat stored at transducer and Ilost stand for the missing part that is more difcult to quantify such as the heat lost to the environment, liquid vaporization (apart from atomization) or ultrasonic wave attenuation. Our research group is now developing a separate experimental work to quantify atomization term from energy balance. 4. Conclusions The intensity of the cavitation is obviously linked to the acoustic power and any increase in ultrasonic power entering the reaction will be associated with an increase in cavitational effect within the system [34]. The purpose of this work was to obtain information about the energy conversion for organic solvents at 500 kHz as a step of a complex study aiming to ascertain the energy balance for the sonochemical reactors. By using the calorimet-

208

M. Toma et al. / Ultrasonics Sonochemistry 18 (2011) 197208 [43] V. Renaudin, N. Gondrexon, P. Boldo, C. Petrier, A. Bernis, Y. Gonthier, Method for determining the chemically active zones in a high-frequency ultrasonic reactor, Ultrason. Sonochem. 1 (1994) S81S85. [44] P. Boldo, V. Renaudin, N. Gondrexon, M. Chouvellon, Enhancement of the knowledge on the ultrasonic reactor by an interdisciplinary approach, Ultrason. Sonochem. 11 (2004) 2732. [45] Y. Kojima, S. Koda, H. Nomura, Effects of sample volume and frequency on power in solutions on sonication, Jpn. J. Appl. Phys. 37 (1998) 2992 2995. [46] Y. Asakura, M. Maebayashi, S. Koda, Study on efciency and characterization in a cylindrical sonochemical reactor, J. Chem. Eng. Jpn. 38 (2005) 10081014. [47] Y. Asakura, K. Yasuda, D. Kato, Y. Kojima, S. Koda, Development of a large sonochemical reactor at a high frequency, Chem. Eng. J. 139 (2008) 339 343. [48] Y. Asakura, T. Nishida, T. Matsuoka, S. Koda, Effects of ultrasonic frequency and liquid height on sonochemical efciency of large scale sonochemical reactor, Ultrason. Sonochem. 15 (2008) 244250. [49] Y. Kojima, Y. Asakura, G. Sugiyama, S. Koda, The effects of acoustic ow and mechanical ow on the sonochemical efciency in a rectangular sonochemical reactor, Ultrason. Sonochem. 17 (2010) 978984. [50] A. Weissler, I. Pecht, M. Anbar, Ultrasound chemical effects on pure organic liquids, Science 150 (1965) 12881289. [51] Y. Mizukoshi, H. Nakamura, H. Bandow, Y. Maeda, Y. Nagata, Sonolysis of organic liquid: effect of vapor pressure and evaporation rate, Ultrason. Sonochem. 6 (1999) 203209. [52] M.G. Boucher, Sonochemistry at low and high ultrasonic frequencies, Br. Chem. Eng. 15 (1970) 363367. [53] T.J. Mason, J.P. Lorimer, D.M. Bates, Quantifying sonochemistry: casting some light on a black art, Ultrasonics 30 (1992) 4042. [54] A. Gaplovsky, M. Gaplovsky, T. Kimura, S. Toma, J. Donovalova, T. Vencel, Method for comparing the efciency of ultrasound irradiation independent of the shape and the volume of the reaction vessel in sonochemical experiments, Ultrason. Sonochem. 14 (2007) 695698. [55] Ratoarinoro F. Contamine, A.M. Wilhelm, J. Berlan, H. Delmas, Power measurement in sonochemistry, Ultrason. Sonochem. 2 (1995) S43S47. [56] Y. Asakura, Doctoral Thesis, Nagoya University, 2007. [57] A. Francony, C. Petrier, Sonochemical degradation of carbon tetrachloride in aqueous solution at two frequencies: 20 kHz and 500 kHz, Ultrason. Sonochem. 3 (1996) S77S82. [58] J.M. Loning, C. Horst, U. Hoffmann, Investigations on the energy conversion in sonochemical processes, Ultrason. Sonochem. 9 (2002) 169179. [59] C.J. Martin, A.N.R. Law, Design of thermistor probes for measurements of ultrasound intensity distributions, Ultrasonics 21 (1983) 8590. [60] B. Niemczewski, A comparison of ultrasonic cavitation intensity in liquids, Ultrasonics 18 (1980) 107110. [61] J.A. Riddick, W.B. Bunger, T.K. Sakano, Organic Solvents-Physical Properties and Methods of Purication, Fourth Edition in Techniques of Chemistry, vol. 2, John Wiley and Sons, 1986. [62] M.N. Topp, Ultrasonic atomization a photographic study of the mechanism of disintegration, Aerosol Sci. 4 (1973) 1725. [63] K.S. Suslick, J.J. Gawienowski, P.E. Schubert, H.H. Wang, Sonochemistry in nonaqueous liquids, Ultrasonics 22 (1984) 2233. [64] B. Avvaru, M.N. Patil, P.R. Gogate, A.B. Pandit, Ultrasonic atomization: effect of liquid phase properties, Ultrasonics 44 (2006) 146158. [65] K.S. Suslick, Homogeneous sonochemistry, in: K.S. Suslick (Ed.), Ultrasound: Its Chemical, Physical and Biological Effects, VCH Publishers, Inc., 1988, pp. 123 159.

[22] K.-V. Jenderka, C. Koch, Investigation of spatial distribution of sound eld parameters in ultrasound cleaning baths under the inuence of cavitation, Ultrasonics 44 (2006) 401406. [23] B. Zeqiri, M. Hodnett, A.J. Carroll, Studies of a novel sensor for assessing the spatial distribution of cavitation activity within ultrasonic cleaning vessels, Ultrasonics 44 (2006) 7382. [24] P.A. Lewin, Miniature piezoelectric polymer ultrasonic hydrophone probes, Ultrasonics 19 (1981) 213216. [25] K.S. Suslick, P.F. Schubert, J.W. Goodale, Chemical dosimetry of ultrasonic cavitation, IEEE Ultrason. Symp. Proc. (1981) 612615. [26] Y. Iida, K. Yasui, T. Tiutzi, M. Sivakumar, Sonochemistry and its dosimetry, Microchem. J. 80 (2005) 159164. [27] V. Saez, A. Fras-Ferrer, J. Iniesta, J. Gonzalez-Garca, A. Aldaz, E. Riera, Characterization of a 20 kHz sonoreactor. Part II: analysis of chemical effects by classical and electrochemical methods, Ultrason. Sonochem. 12 (2005) 67 72. [28] F. Trabelsi, H. Ait-lyazidi, J. Berlan, P.L. Fabre, H. Delmas, A.M. Wilhelm, Electrochemical determination of the active zones in a high-frequency ultrasonic reactor, Ultrason. Sonochem. 3 (1996) S125S130. [29] F. Faid, F. Contamine, A.M. Wilhelm, H. Delmas, Comparison of ultrasound effects in different reactors at 20 kHz, Ultrason. Sonochem. 5 (1998) 119124. [30] F. Contamine, F. Faid, A.M. Wilhelm, J. Berlan, H. Delmas, Chemical reactions under ultrasound: discrimination of chemical and physical effects, Chem. Eng. Sci. 49 (1994) 58655873. [31] M.W.A. Kuijpers, M.F. Kemmere, J.T.F. Keurentjes, Calorimetric study of the energy efciency for ultrasound-induced radical formation, Ultrasonics 40 (2002) 675678. [32] T. Kimura, T. Sakamoto, J.-M. Leveque, H. Sohmiya, M. Fujita, S. Ikeda, T. Ando, Standardization of ultrasonic power for sonochemical reaction, Ultrason. Sonochem. 3 (1996) 6772. [33] P.M. Kanthale, P.R. Gogate, A.B. Pandit, A.M. Wilhelm, Mapping of an ultrasonic horn: link primary and secondary effects of ultrasound, Ultrason. Sonochem. 10 (2003) 331335. [34] T.J. Mason, J.P. Lorimer, D.M. Bates, Y. Zhao, Dosimetry in sonochemistry: the use of aqueous terephthalate ion as a uorescence monitor, Ultrason. Sonochem. 1 (1994) S91S95. [35] J. Raso, P. Manas, R. Pagan, F.J. Sala, Inuence of different factors on the output power transferred into medium by ultrasound, Ultrason. Sonochem. (1999) 157162. [36] P.R. Gogate, A.B. Pandit, Engineering design methods for cavitation reactors I: sonochemical reactors, AIChE J. 46 (2000) 372379. [37] G. Servant, J.-L. Laborde, A. Hita, J.-P. Caltagirone, A. Grard, Spatio-temporal dynamics of cavitation bubble clouds in a low frequency reactor: comparison between theoretical and experimental results, Ultrason. Sonochem. 8 (2001) 163174. [38] T.J. Mason, J.P. Lorimer, Applied Sonochemistry: The Uses of Power Ultrasound in Chemistry and Processing, Wiley-VCH, Weinheim, Germany, 2002. pp. 36 45. [39] M.E. Webber, W.Y. Chon, Distribution of ultrasonic cavitation intensities in a liquid system, Can. J. Chem. Eng. 45 (1967) 238240. [40] B. Pugin, Qualitative characterization of ultrasound reactors for heterogeneous sonochemistry, Ultrasonics 25 (1987) 4955. [41] M. Romdhane, G. Gourdon, G. Casamatta, Local investigation of some ultrasonic devices by means of a thermal sensor, Ultrasonics 33 (1995) 221 227. [42] M. Romdhane, A. Gadri, F. Contamine, C. Gourdon, G. Casamatta, Experimental study of the ultrasound attenuation in chemical reactors, Ultrason. Sonochem. 4 (1997) 235243.

Potrebbero piacerti anche