Vous êtes sur la page 1sur 8

SONOCHEYISTAY

Ultrasonics Sonochemistry 3 (1996) S83-S90

Sonolysis of trichloroethylene in aqueous solution: volatile organic intermediates


David Drijvers *, Robrecht De Baets, Alex De Visscher, Herman Van Langenhove
Department

of OrganicChemistry, Faculty of Agricultural & Applied Biological Sciences, University of Ghent. Coupure Links 653, B-9000 Gent, Belgium Received 15 February 1996;revised 30 April 1996

Abstract

The ultrasonic degradation of 3.34 mM trichloroethylene (TCE) in aqueous solution was measured at 20 and 520 kHz. As the degradation was energetically more efficient at 520 kHz, sonication at this frequency was further investigated. The effect of the saturating gas (air or argon) and the influence of the pH of the aqueous solution was studied. The degradation was fastest in basic solutions saturated with argon. TCE was not degraded in the bulk solution. During sonication volatile and non-volatile organic degradation products were formed. The most important volatile compounds were identified: C,HCl, C,Cl,, CqCl,, C2Cld, C,HCl, (2), C4Cl,, C,HCl, and C4Cl,. Those products are typical for pyrolysis of TCE and are an affirmation for the hot spot theory. The kinetics of five of those volatile intermediates were determined by headspace analysis for air-saturated as well as for argon-saturated solutions. The intermediates considered are formed in single cavitation events and disappeared from the aqueous

solution as well.
Keywords:

Sonolysis; Trichloroethylene;

Aqueous solutions -

1. Introduction Ultrasonic waves in water provoke the collapse of cavitation bubbles and induce the formation of chemical species such as H , OH , 0 and H,O, [l-3]. Two

theories have been suggested to explain their formation: the hot spot theory and the electrical discharge theory. According to the hot spot theory the collapse of a bubble is almost adiabatic and causes extremely high temperatures and pressures. By determining the first-order rate coefficients of sonochemical ligand substitution as a function of metal carbonyl vapour pressure, a temperature of 5000 K inside the collapsing cavitation and a temperature of 2000 K in the interfacial region between the bulk solution and the collapsing bubble was determined [4]. By using the temperature dependence of C-N bond pyrolysis of p-nitrophenol in oxygenated aqueous solutions, a temperature of 800 K was determined for the interfacial region [S]. According to the electrical discharge theory, the extreme conditions

* Corresponding f 32-9-264-6243.

author.

E-mail:

David.Drijvers@rug.ac.be;

fax:

leading to the formation of the radicals originate from intense electrical fields during collapse of the cavitation bubbles [6]. The combination of both theories has been used as well to explain all phenomena of ultrasonic irradiation [ 71. The existence of electrical discharge was argued by the analogies between the features of sonochemistry and corona discharge. However, data from sonochemical experiments are still mostly interpreted in terms of the hot spot theory. Ultrasonic irradiation causes the degradation of organic compounds in water. The OH radicals formed oxidize the organic compounds. Apolar and voltatile organic compounds, however, are also degraded thermally in and around the cavitations. Thus, both indirect and direct decomposition by ultrasonics occur. The solution can be divided into three different zones during sonication [S]. The collapsing microbubbles, filled with dissolved gas and solvent vapour, are the reaction zones in which solvent vapour is pyrolysed, producing radicals. Volatile organics too are pyrolysed in the collapsing bubbles. The interfacial region, a thin shell of liquid, is exposed to temperatures and pressures exceeding the critical ones and the resulting supercritical

1350-4177/96/$15.00 Copyright PII S1350-4177(96)00012-O

0 1996 Elsevier Science B.V. All rights reserved

S84

D. Drijvers.

et al. f Ultrasonics

Sonochemistry

3 (1996)

S83-S90

fluid possesses physical properties other than the bulk solution. The dielectrical constant is lowered for polar solvents such as water and allows the accumulation of low-polarity solutes. The third region is the bulk solution. Here, the solutes react with radicals that have not yet recombined, disproportionated or that have not been scavenged. An important parameter for the efficiency of the degradation of organic compounds is the average specific heat ratio y (=C,/C,) of the gas and solvent vapour in the collapsing bubble. The temperature inside the collapsing bubble is closely related to this specific heat ratio. Riesz [9] calculated that the final collapse temperature for bubbles saturated with monoatomic gases, y = 1.67, is more than twice as high than for bubbles saturated with N,O, y = 1.30. Aspects of the 20 kHz ultrasonic degradation of most Cl and C2 chlorinated compounds and some chlorofluorocarbons in aqueous solutions have been investigated [ 10-141. CHCl, and CCL, supersaturated solutions have been sonicated. The observed organic products from the degradation of both CHCl, and Ccl, were hexachloroethane and tetrachloroethylene [lo]. The sonolysis of supersaturated and dilute aqueous solutions of l,l,l-trichloroethane was investigated as well [ 11,121. Organic degradation products such as chloroform, chloroethylene, l,Zdichloroethane, l,l-dichloroethane, chloroethane and 1,2-dichloroethylene were determined. However, these organic degradation products were not quantified. The ultrasonic degradation was more complete for the dilute solution than for the supersaturated solution. This was explained by the decrease of the cavitation intensity because of the higher vapour pressure of the supersaturated solution of l,l,l-trichloroethane [ 111. For the chlorofluorocarbons CFC 11 and CFC 113 [ 131 and eight chlorinated Cl and C2 volatile organic compounds [ 143, including trichloroethylene, dilute aqueous solutions have been sonicated and no organic byproducts were sought. Only the degradation rate, following first order, was measured. The Cl compounds seemed to degrade by pyrolysis because of the high vapour pressure of these compounds. As the vapour pressure of the Cl compounds increased, the first-order rate coefficient seemed to decrease. Unlike Cl compounds, there seemed to be no relation between the lower vapour pressure of the chlorinated C2 compound and its destruction rate. This led to the conclusion that C2 chlorinated compounds probably disintegrate by both mechanisms: pyrolysis-type reactions in the cavitation and free radical attack in the bulk liquid phase. At the higher frequency of 200 kHz, aspects of the ultrasonic decomposition of 10 ppm aqueous solutions of trichloroethylene, tetrachloroethylene, l,l,l-trichloroethane, chloroform and carbon tetrachloride were studied [ 151. After an irradiation time of 10 minutes more than 70% of the chlorinated hydrocarbon was

degraded and it was concluded that the main reactions are thermal decomposition or combustion in cavitation bubbles and not reactions by OH radicals or H atoms. The sonication of an argon saturated solution of 10 ppm or 0.11 mM trichloroethylene gave Cl-, CO and H, as major products and COZ, methane and ethylene as minor products. A trace amount of dichloroethylene was detected by GC-MS. The concentration versus time profile for the final products Cl-, H,, CO and COZ was also determined. The ultrasonic degradation at high frequency of less dilute solutions of chlorinated hydrocarbons has not yet been investigated. As the concentration of the volatile hydrocarbon will decrease the average specific heat ratio of the gas in the collapsing bubble, provoking lower temperatures and pressures, a higher amount of intermediate organics is expected than in the case of the sonication of the 10 ppm aqueous solutions [ 151. A higher extent of radical recombination reactions is also expected, leading to different degradation products. The sonication at 1 MHz of argon-saturated aqueous solutions of the volatile C2 compound acetylene in the millimolar range was studied [ 161. The following products were found: H,, CO, CH,, hydrocarbons containing two to about eight C atoms, formaldehyde and acetaldehyde, formic and acetic acid and insoluble soot. The products observed are similar to the ones observed in the pyrolysis and combustion of acetylene. Initially, all products appeared proportional to the irradiation time and it was concluded that all products are formed in single cavitation events and not by stepwise formation and subsequent sonolysis of intermediate compounds in different cavitation bubbles. The aim of this study is to investigate if the sonication of trichloroethylene (TCE) in the millimolar range leads to similar volatile organic intermediates as in the case of acetylene [ 161. Knowledge of the volatile organic intermediates formed will lead to a better understanding of the ultrasonic degradation of TCE. The effect of the saturating gas, air or argon, was investigated as well.

2. Materials and methods Sonication experiments were performed with a 20 kHz Branson 250 sonicator and a 520 kHz Undatim Ortho Reactor. The high frequency reactor was equipped with an extra voltmeter to allow sonication with a constant power transfer to the liquid. At low frequency, 120 ml solution in a 160 ml reaction cell, and at high frequency 150 ml solution in a 200 ml reaction cell, was sonicated. The steady state reaction temperature was 32 + 1C in the 20 kHz reactor and 29.5 t 0.5C in the 520 kHz reactor.

D. Drijvers,

et al. / Ultrasonics

Sonochemistry

3 (1996)

S83490

S85

2.1. Chemicals

All organic chemicals commercially available were of a purity 399%. Only two of the eight major organic intermediates formed during sonication are commercially available, namely tetrachloroethylene and hexachlorobutadiene. Three intermediate products, dichloroacetylene, tetrachlorobutenyne and pentachlorobutadiene were synthesized. Tetrachlorobutenyne and pentachlorobutadiene were synthesized out of hexachlorobutene. Hexachlorobutene was made by dimerisation of trichloroethylene in the presence of dibenzoylperoxide [ 171. Dehydrochlorination of hexachlorobutene by OH - yielded pentachlorobutadiene [ 181. By using the stronger base NH; for further dehydrochlorination of hexachlorobutene tetrachlorobutenyne was obtained [ 191. Both products, pentachlorobutadiene and tetrachlorobutenyne, were purified by preparative GC. Purity was verified by H- or 13C-NMR and mass spectrometry. Dichloroacetylene is a very reactive compound which forms explosive mixtures with air. The presence of diethylether in the system retards the violent autooxidation of dichloroacetylene by formation of a molecular complex [20]. This complex was obtained from a mixture of TCE and diethylether in the presence of an aqueous solution of sodium hydroxide at 70C and with an ammonium salt as a phase-transfer catalyst [20]. A 13C-NMR and a mass spectrum of the obtained mixture of dichloroacetylene and ether was taken. Traces of TCE were also present. It was not possible to separate the mixture by distillation or preparative GC because of the reactivity of dichloroacetylene with air. 2.2. Determination of the kinetics of sonochemical degradation of TCE The initial TCE concentration in all experiments was 3.34 mM. The solution was buffered at pH 4.7 (40 mM acetate buffer), pH 7 (40 mM phosphate buffer) and pH 10 (40 mM borate buffer). In the argon experiments the solution was saturated with argon before adding TCE and the reactor was flushed with argon before adding the solution. A syringe needle was pierced through the septum of the screw cap of both reaction cells for sampling. Using a glass syringe 2 ml samples were taken from the 520 kHz reactor at various time intervals during sonication while the sonication was restarted in the case of the 20 kHz reactor for every analysis. This was done because the low frequency reactor did not allow sampling during sonication. The 2 ml samples were transferred to 2.5 ml bottles, and sealed with Mininert stoppers (Alltech Ass.). 50 ul of a 1 ~01% 2-hexanone aqueous solution was added as an internal standard. In the case of sonication at 20 kHz, 100 ul of toluene in dichloromethane (volume ratio l/5) was added to the 120 ml sonicated solution as

an internal standard. 1 ul of the mixture was then analysed on a Varian 3700 gas chromatograph (FID-detector) with a DB-5 fused silica column (15 m, film thickness 1.5 urn, ID 0.53 mm). The initial GC oven temperature was 40C and a 3C min- temperature rise was used.

2.3. Identification
intermediate

and determination products

of the kinetics of

To identify the volatile intermediate products the mass spectra were determined with a Varian 2700 gas chromatograph with a RSL 150 fused silica column (60 m, film thickness 1.5 urn, ID 0.53 mm) coupled to a MAT 112 mass spectrometer. A constant temperature increase of 3C min- from 30C to 200C was used as a temperature program. The intermediates were stripped from the diluted sonicated solution with helium, absorbed on Tenax GC and injected by the cold trap GC method. The Kovats indices of the intermediate products were determined with the Varian 3700 gas chromatograph with a RSL 150 fused silica column (30 m, film thickness 5 urn, ID 0.53 mm) and with a FID-detector. The same procedure was followed as for the mass spectra but 0.5 ul of the alkanes C5-Cl6 (1 ~01%) in CS2 was spiked on the Tenax tube before desorption. The Kovats index of the standard solution of dichloroacetylene and tetrachlorobutenyne was determined by direct liquid injection. The kinetics of the intermediate volatile products were determined by headspace analysis: after various sonication times, 100 ml of the solution was transferred to a 118 ml bottle. The bottle was closed immediately with a Mininert valve (Alltech Ass.) and 50 ul of cyclohexane in methanol (0.06 ~01%) was added as an internal standard. The bottle was then incubated overnight in a thermostatic water bath at 25 f O.lC. Earlier studies indicated that this time is sufficient for attaining equilibrium partitioning of the organic compounds between the water phase and the headspace [21,22]. After incubation 1 ml headspace (Syringe Pressure-Lok Series A, 1 ml) was injected in the Varian 3700 gas chromatograph with the column used for the Kovats index determination. The GC oven was kept at 35C for three minutes and then heated to 200C at a rate of 10C min- . A standard curve for this headspace analysis was determined for dichloroacetylene, tetrachloroethene, tetrachlorobutenyne, pentachlorobutadiene and hexachlorobutadiene by injection of the headspace above standard solutions. As dichloroacetylene was only available as a mixture with ether and TCE, the standard curve for ether and TCE was also determined by headspace injection. So the quantity of dichloroacetylene injected could be calculated by subtraction of the quantity of ether and TCE from the quantity of the mixture injected.

S86

D. Drijvers, et al. / Ultrasonics Sonochemistry 3 (1996) S83-S90 Table 2 First-order reaction rate coefficients k, with standard deviations (sd.) and correlation coefficients r (n = 7) for the sonochemical degradation of TCE at different pH values PH 4.1 7 7 10 10 k, (mini ) 0.03053 0.03728 0.03944 0.04269 0.04044 sd. (min-r) 0.00335 0.00246 0.00295 0.00407 0.00409 r 0.995 0.997 0.998 0.997 0.996

3. Results and discussion 3.1. General characteristics

The effective ultrasonic power Qes of the sonicators was determined calorimetrically [23] in triplicate. For the 20 kHz Branson 250 sonicator the power transferred to the liquid was 51.80 f 1.28 W or 0.43 W ml- and for the 520 kHz Undatim Ortho Reactor 14.23 & 0.73 or 0.095 W ml-i. In Fig. 1 the concentration of trichloroethylene (TCE) versus sonication time for non-buffered, air-saturated solutions using 20 kHz and 520 kHz ultrasonic waves is shown. As was found previously [ 141, the concentration decreased exponentially with sonication time, strongly indicating first-order or pseudo-first-order kinetics. The experiments were repeated twice at each frequency and the corresponding first-order rate coefficients k, are given in Table 1. After 20 minutes of sonication the initially neutral pH decreased to 2.9 _t 0.1 at 20 kHz and to 2.4 f 0.1 at 520 kHz. It is clear from Fig. 1 that the ultrasonic degradation of TCE is energetically more efficient at 520 kHz. This is in agreement with the results found by comparing the ultrasonic degradation of benzene, chlorobenzene, phenol and chlorophenol at 20 and 500 kHz [ 241. The sonochemical degradation at 520 kHz was carried out at different pH values. Table 2 shows the k,-values obtained. During sonication the buffers limited the pH-decrease to less than 0.4 pH-units. Obviously, there
Table 1 First-order reaction rate coefficients k, with standard deviations (s.d.), correlation coefficients r and n, the number of data points on which the regression was calculated, for the sonochemical degradation of TCE at two frequencies Frequency 20 20 520 520 (kHz) k, (mini ) 0.02758 0.02625 0.03383 0.03299 s.d. (min- ) 0.00507 0.00411 0.00364 0.00476 r 0.991 0.994 0.996 0.992 n 6 6 7 7

is an influence of pH on the degradation rate of TCE. The k,-values obtained at pH 7 and 10 are significantly higher than those obtained at pH 4.7 and without buffering. The effect of the pH on the ionization of the CO, formed (pK,, = 6.38 and pK,, = 10.25) is thought to be the reason for the higher degradation rate in basic solutions. CO, will be ionized to carbonate and bicarbonate in basic solutions, while in acidic solutions CO2 is present as dissolved gas. The specific heat ratio y is 1.304 [25] at 15C for CO, and 1.403 [25] at 17C for air. As CO2 has a lower y than air, it will decrease the efficiency of the adiabatic bubble collapse and thus of the ultrasonic degradation of TCE. Further experiments were carried out at pH 7. In order to evaluate the influence of the dissolved gas, the experiment was also carried out with an argon saturated solution. The y-value of argon is 1.668 [25] at 15C. A first-order rate coefficient k, = 0.06155 ) 0.00253 min- (Y= 0.998; n = 13) was obtained. The degradation of 3.34 mM TCE occurs significantly faster when the solution is saturated with the monoatomic argon instead of air. This is again explained by the higher value of y. The hydroxyl radical production rate in air-saturated water was determined in the absence of TCE and buffers. The rate of oxidation of 1 M potassium iodide to iodine and 30 mM ferrous sulphate to ferric sulphate can be used as respectively a lower and upper estimate of the hydroxyl radical production rate in the bulk solution [23]. The concentration of I; ~~~~~ = 26400 M-l cm- ) and Fe3+

3.5 4 1 . 520 kHz (0.095 W/ml) A 20 kHz (0.43 W/ml)

I .

. .

o0

. . i
60

10

20

30

40

50

time (min)
Fig. 1. Concentration versus time profiles of TCE for two ultrasonic frequencies with corresponding effective power Q.n,

D. Drijvers, et al. / Ultrasonics Sonochemistry 3 (1996) S83-S90

SE37

(%5= 2197 M- cm-)

was measured spectrophotometrically. The rate of oxidation was obtained from the linear regression of these concentrations as a function of irradiation time (YZ 0.993; n = 5). The lower estimate to the hydroxyl radical yield was 0.0040 mM min- and the upper estimate 0.0176 mM min- . So, during the 60 minutes of sonication the hydroxyl radical production in the bulk solution lies between 0.24 and 1.06 mM or between 7 and 32% of the initial TCE concentration. This quantity of hydroxyl radicals makes a considerable degradation of TCE in the bulk solution possible. To measure the importance of this degradation of TCE in the bulk solution, the radical scavenger NaBr was added to the TCE solution buffered at pH 7. The reaction rate coefficient [2] kOH of Br- with OH is 10 M-' s-i. Makino showed, by spin trapping, that 1 M formate, with the same kOH, was enough to scavenge all OH in the bulk solution when air-saturated water was sonicated [2]. For the experiment in which 1 M NaBr was added to the solution a rate coefficient kl = 0.03909 + 0.00369 min- (r = 0.993; n = 11) was obtained. As the degration rate of TCE was not influenced by the addition of NaBr, TCE is not degraded in the bulk solution at 520 kHz. The apolar volatile compound TCE is degraded in or at least near the collapsing microbubbles. 3.2. Intermediate
volatile organic products

from the headspace analysis of an aqueous solution of the purchased or synthesized product (Istand). The peak of tetrachloroethylene coincided in both cases with the peak of octane. The Kovats indices confirmed the structure of the five intermediates proposed by the mass spectra. Taylor studied the high-temperature, oxygen-free pyrolysis of TCE in the 573-1273 K temperature range [26]. Initial decomposition was observed at 1000 K with formation of HCl and C,Cl,. Pronounced molecular growth was observed at higher temperatures with the formation of C,C&, C,Cl, and C,Cl, (cyclic (cy)) as major (25 mole%) products and C,Cl,, CqCls, CgHCls (cy), CgC& (cY), C&~S (cY), C10Cls (CY)and C12C18 (CY)as minor (< 5 mole%) products. The mole% of the products depended strongly on the temperature of pyrolysis. A pyrolysis mechanism of TCE was proposed and tested using computer simulations based on the rate coefficients and the activation energies E, for all possible reactions. Generally, there was good agreement between predicted and experimental species profiles. The unimolecular C-Cl bond scission is the dominant initiation step and the Cl produced then abstracts the H atom of TCE yielding C,Cl; and HCl (Eqs. (1) and (2)) CzHC1, + C,HCl; + Cl , C2HCl, + Cl + &Cl; + HCl. (1) (2)

Stripping of the sonicated solutions saturated with air or with argon and analysis by GCMS of the Tenax tubes resulted in nine major peaks of which two had the same mass spectrum. Based on the mass spectra, the following volatile products were tentatively identified: chloroacetylene (C,HCl), dichloroacetylene (CzC1,), dichlorodiacetylene (C,Cl,), tetrachloroethylene (CzC14), two isomers of trichlorobutenyne (CqHC13), tetrachlorobutenyne (&CL,), pentachlorobutadiene (CqHC15) and hexachlorobutadiene (C,Cl,). To confirm the results of interpretation of the mass spectra of these products, the Kovats index for five products was determined. Table 3 shows the Kovats indices obtained from the headspace analysis of the sonicated solution of TCE (Isonic) and
Table 3 The Kovatz indices Isonic and Istand, respectively determined by injecting headspace above the sonicated solution and the standard solution, of the intermediate products Product
GC ,

The &Cl; radical then either loses a Cl radical to form CzClz or reacts with Cl2 to give CzCld and a Cl radical. Although C,HCl was not detected, model results indicate that C,HCl is formed as a trace product (< 0.5% of the initial C,HCl, concentration). In a study of the thermal or CO,-laser induced decomposition of TCE, &Cl,, C,CL, C,HCl,, CqC16, C,HCl,, C&l,, CsHCl,, C8HC17 and C8C1, were detected as degradation products [27]. The reaction was proposed to proceed by elimination of HCl (Eq. (3)) from TCE, followed by radical formation and oligomerization C,HCl, -+ &Cl, + HCl. (3)

~aonic 509 800* 1033 1160 1219 with octane.

I stand 509 800* 1033 1161 1218

CzC14 C4C4 C,HCI, C l CL * Coincided

The similarity of products found in the sonolysis experiment with the most volatile ones from the two aforecited studies indicates that TCE is at least partially degraded thermally. In the study of the OH radical initiated oxidation of TCE [ 281 no similar products were detected by GC-analysis. In a recent study on the plasma-assisted decomposition of TCE in a pulsed corona discharge reactor [29] the volatile organic intermediates were not looked for. It would be interesting to do so. The absence of the above mentioned nine volatile organic intermediates would confirm the thermal degradation of TCE during sonication. In both pyrolysis studies [26,27] the formation of the higher molecular weight products is found to proceed via CzC1, and neither C,HCl nor C,HCl, was detected.

S88

D. Drijvers,

et al. / Ultrasonics

Sonochemistry

3 (1996)

S83-S90

The formation of for example C,Cl, was explained by the following reactions [27] (Eqs. (4) and (5)): clc~ccl-+clc~c +c1 , ClC~C +C1C~CCl+C1C~C-CCl=CCl (4) (A). (5) By picking up a free chlorine atom or by abstracting one from another molecule C,Cl, is formed out of the resulting radical A. By applying the same mechanism, CqHC13 must be formed out of &Cl, and C,HCl. This means that both CZHC1 as CzClz are formed during sonolysis of TCE. Formation of &Cl4 on the other hand confirms that the reaction mechanism proposed by Taylor proceeds [ 261. Recently, some studies were done on the mechanism of unimolecular dissociation of trichloroethylene [ 30,3 11. The enthalpy diagram of the possible channels for the TCE dissociation process is given in Fig. 2. The Cl, elimination reactions, which are expected to be very high in energy, are not considered [30]. TSa and TSb are the transition states of the three- and four-centered elimination of HCl respectively. As can be seen from Fig. 2, the difference in enthalpy for reactions la, lb and 2a is small and those reactions will compete. It can be concluded that both C-Cl bond scission (Eq. (1)) and HCl elimination (Eq. (3)) are thermodynamically possible as initial reactions in the degradation of TCE. The formation of C,HCl can be explained as the result of two stepwise Cl atom eliminations from TCE [31] (Eqs. (1) and (6)). C,HCl; + C,HCl+ Cl . (6)

are so extreme that even the Cl, elimination of TCE could compete with the Cl abstraction and the HCl elimination. The fact that the same volatile organic intermediates are found for the solutions saturated with air as with argon is explained by the low oxygen concentration in the aqueous solution. Normally combustion of TCE [ 321 gives partially different degradation products than oxygen-free pyrolysis. But in the aqueous air-saturated solution, initially only 0.25 mM oxygen is present and during the sonication the oxygen content decreases: 0.15 mM, 0.1 mM and 0.05 mM after 10, 30 and 60 minutes sonication respectively. This decrease is explained by degassing of the solution and by Eq. (7): 02+H +HO,. (7)

In a second phase the kinetics of the intermediate volatile products were determined. The standard curves for the five products considered are given in Table 4. The standard deviation on the concentration of these products was quite high. The analysis of the concentrations after 30 minutes sonication was performed twice and standard deviations as high as 20% were found. Figs. 3 and 4 show the kinetics for those intermediate products. There is no basic difference between the kinetics under air and under argon. This means that the temperature near and inside the collapsing bubble is, in both cases, so high that the difference has no importance. Although no smooth curves are obtained, a linear concentration increase during the first 10 minutes of sonication under air and argon can be deduced for most
Table 4 The standard products Product

The dissociation energy [ 311 for Eq. (6) is only 16.1 kcal mol- . The reason why C,HCl and C,HCl, were not found by Taylor and by Earl as pyrolysis products, could be the difference in conditions [26,27]. Under less extreme conditions the small differences in enthalpy between reactions la, lb and 2a from Fig. 2 could make the reaction thermodynamically exclusive. The HCl elimination is then energetically preferred. Another explanation for the formation of C,HCl could be that the conditions during bubble collapse

curves

for the headspace

analysis

of the intermediate

Slope (relative 15.99 23.81 17.25 11.19 39.93

area count, mM_ )

Intercept (relative area count) -0.1040 0.0036 0.0072 0.0052 0.0010

C*Cl, C,C& GCl, C, HCI, C,Cl,

(2b) CH=CC12 101.9

+ Cl \

\CHCl=CC12 Fig. 2. Enthalpy diagram for the TCE dissociation reaction.

/ are written in units of kcal mol- (after Yokoyama et al. [26])

Energies

D. Drijvers,

et al. / Ultrasonics

Sonochemistry

3 (1996)

S83-S90

S89

0.16 0.12 s c 5 0.08 0.1 0.06 0.04 0.02 0

--1-

C2Cl2 - air

50

100

150

200

time (min)
Fig. 3. Concentration versus time profiles of dichloroacetylene (C,C12) and tetrachloroethylene (C,C&)

0.006

, ---t
-m*------t --*---

C4Cl4 - air

, ~-*-- C4C14 - argon


C4HCl5 - air C4HCl5 - argon C4Cl6 - air C4Cl6 - argon

50

100

150

200

time (min)
Fig. 4. Concentration versus time profiles of tetrachlorobutenyne (C,C&), pentachlorobutadiene (C,HCl,) and hexachlorobutadiene (C,Cl,).

compounds. Furthermore, all maxima in Figs. 3 and 4 lie around 30 minutes. These two observations indicate that the five volatile intermediates are formed in single cavitation events, as was found previously for the products formed during sonication of acetylene [ 161. The concentration of C,CI,, a primary product, is at least 10 times higher than the concentration of the four other products. C,Cl, is formed during the first minutes at a rate of approximately 0.01 mM min- . At that moment, the rate at which the 3.34mM TCE breaks down is approximately 0.12 mM min- . So, the formation of &Cl, represents about 10% of the total degradation, Considering that during those first minutes at least C,Cl,, C,HCl, and C,Cl, are formed out of C,Cl, in single cavitation events, this percentage is higher. Since the response factor of CPHCl is not known, the amount of CZHCl could not be quantified. But the areas obtained for CLHCl were in the same order of magnitude as for C,Cl,. So the thermal degradation of TCE during sonication at 520 kHz represents a significant degradation pathway. Also, non-volatile products are formed during the sonication. Work is currently being done to identify and quantify those non-volatile products.

Acknowledgements

DD wishes to thank the Vlaams Instituut voor de Bevordering van het Wetenschappelijk-Technologisch Onderzoek in de Industrie (IWT) for financial support and ADV wishes to thank the Vlaams Impulsprogramma Milieutechnologie (VLIM) for financial support.

References
M. Anbar and I. Pecht, J. Phys. Chem. 68 (1963) 352. K. Makino, M.M. Mossoba and P. Riesz, J. Phys. Chem. 87 (1983) 1369. [3] C. P&trier, M. Micolle, G. Merlin, J.-L. Luche and G. Reverdy, Environ, Sci. Technol. 26 (1992) 1639. [4] K.S. Suslick, D.A. Hammerton and R.E. Cline, Jr., J. Am. Chem. Sot. 108 (1986) 5641. [S] A. Kotronarou, G. Mills and R. Hoffmann, J. Phys. Chem. 95 (1991) 3630. [6] M.A. Margulis, Ultrasonics 23 (1985) 157. [7] T. Lepoint, N. Voglet, L. Faille and F. Mullie, Bubble Dynamics and Interface Phenomena, J.R. Blake et al. (Eds.) (Kluwer Academic Publishers, The Netherlands, 1994) 321. [ 83 V. MiSik, N. Miyoshi and P. Riesz, J. Phys. Chem. 99 (1995) 3605. [9] P. Riesz and F. Takashi, Free Rad. Biol. Med. 13 (1992) 247. [lo] B.H. Jennings and S.N. Townsend, J. Phys. Chem. 65 (1961) 1574. [l] [2]

s90

D. Drijvers,

et al. / Ultrasonics

Sonochemistry

3 (1996)

S83490

Cl11 MS. Toy, M.K. Carter and T.O. Passell, Environ. Technol. 11 (1990) 837. Cl21 MS. Toy, R.S. Stringham and T.O. Passell, Pollution Prevention in Industrial Processes, J.J. Breen and M.J. Dellarco (Eds.) (Am. Chem. Sot. Sym. Series No. 508, Washington, DC, 1992) 284. and S. Kurup, Environ. Sci. Technol. 28 Cl31 H.M. Cheung (1994) 1619. and H.M. Cheung, Environ. Sci. Technol. 28 Cl41 A. Bhatnagar (1994) 1481. Cl51 K. Inazu, Y. Nagata and Y. Maeda, Chem. Lett. 57 (1993). Cl61 E.J. Hart, C.-H. Fischer and A. Henglein, J. Phys. Chem. 94 (1990) 284. and A.U. Blackham, J. Am. Chem. Sot. 72 Cl71 C.E. Frank (1950) 3283. Cl81 A. Roedig and R. Kloss, Chem. Ber. 90 (1957) 2902. Chem. Ber. Cl91 A. Roedig, G. Bonse, R. Helm and R. Kohlhaupt, 104 (1971) 3378. c201 J. Pielichowski and R. Popielarz, Synthesis (1984) 433. c211 J.M. Gossett, Environ. Sci. Technol. 21 (1987) 202. c221 A.H. Lincoff and J.M. Gossett, Gas Transfer at Water Surfaces, W. Brutsaert and G.H. Jirka (Eds.) (Reidel, Dordrecht, 1984) 17.

~231 A. Kotronarou,

G. Mills and R. Hoffmann, Environ. Sci. Technol. 26 (1992) 2420. of the ~241M.F. Lamy, C. Petrier and G. Reverdy, Proceedings Third Meeting of the European Society of Sonochemistry, Figuera da Foz, Portugal (1993) 87. s Chemical Engineers ~251P.E. Liley, R.C. Reid and E. Buck, Perry Handbook, R.H. Perry and D. Green (Eds.) (McGraw-Hill, New York, 1984). C261 P.H. Taylor, D.A. Tirey, W.A. Rubey and B. Dellinger, Combust. Sci. and Tech. 101 (1994) 75. ~271B.L. Earl and R.L. Titus, Collect. Czech. Chem. Commun. 60 (1995) 104. C281N. Itoh, S. Kutsuna and T. Ibusuki, Chemosphere 28 (1994) 2029. ~291M.C. Hsiao, B.T. Merritt, B.M. Penetrante and G.E. Vogtlin, J. Appl. Phys. 78 (1995) 3451. J. Chem. Phys. c301J.-F. Riehl, D.G. Musaev and K. Morokuma, 101 (1994) 5942. G. Fujisawa and A. Yokoyama, J. Chem. Phys. c311 K. Yokoyama, 102 (1995) 7902. ~321W.-D. Chang and S.M. Senkan, Environ. Sci. Technol. 23 (1989) 442.

Vous aimerez peut-être aussi