Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Chemical Engineering & Processing: Process Intensification 179 (2022) 109080

Contents lists available at ScienceDirect

Chemical Engineering and Processing - Process


Intensification
journal homepage: www.elsevier.com/locate/cep

Influence of processing conditions on hydrogen Sonoproduction from


methanol sono-conversion: A numerical investigation with a
validated model
Aissa Dehane a, Slimane Merouani a, *, Atef Chibani a, Oualid Hamdaoui b,
Muthupandian Ashokkumar c
a
Laboratory of Environmental Process Engineering, Department of Chemical Engineering, Faculty of Process Engineering, University Salah Boubnider-Constantine 3, P.O.
Box 72, 25000, Constantine, Algeria
b
Chemical Engineering Department, College of Engineering, King Saud University, P.O. Box 800, 11421, Riyadh, Saudi Arabia
c
The University of Melbourne, Department of Chemistry, Parkville, Victoria, 3010, Australia

A R T I C L E I N F O A B S T R A C T

Keywords: A comprehensive numerical investigation was conducted to analyze the effect of ultrasound frequency (213-1000
Methanol sonolysis kHz), ultrasonic intensity (0.7-1.5 W/cm2) and medium temperature (293.15-333.15 K) on the production of
Hydrogen production hydrogen from methanol decomposition inside acoustic bubbles (methanol sono-conversion). The adopted model
Ultrasound frequency
was firstly compared with some reference data. It was found that in the absence of methanol (in the presence of
Acoustic intensity
Liquid temperature
methanol), the bubble temperature, hydrogen production (and methanol conversion) are decreased monoto­
nously with an increase in frequency. The maximal bubble temperature is slightly impacted by the presence of
methanol under ultrasonic frequencies that are equal to or greater than 515 kHz. In addition, the yield of H2 is
larger in the presence of methanol, regardless of the utilized frequency. Between 213 to 355 kHz, methanol
conversion and hydrogen production are most efficient. Both H2 production and CH3OH degradation are
accelerated under an ideal acoustic intensity of 1 W/cm2. At a liquid temperature of 303.15 K, a turning point in
bubble temperature may be noticed in the absence of methanol, and at a liquid temperature of 323.15 K, the
highest hydrogen production can be achieved. Hydrogen production and methanol conversion are most efficient
at liquid temperatures between 293.15 and 303.15 K.

mixtures have to be used to avoid the formation of nitrogen oxides [10].


The hydrogen as a fuel is a great source of energy that could be used in
1. Introduction different applications including space industry, transportation, fuel
cells, rockets, pumping and heating and more [10,11].
In comparison to other traditional energy sources (methane, diesel The sonochemical production of hydrogen has gained the interest of
and gasoline), hydrogen has numerous advantages, including higher many researchers as a clean, non-toxic and low-cost technique [12–22].
heating value (141.9 kJ g− 1) [1], ability of storage in different forms However, lower hydrogen yields has been reported from pure water
(liquid, gas, or in combination with metal hybrid), availability, and the sonolysis (<40 µM/min) [12]. To overcome this inconvenient, many
fact that it is a clean fuel with really no CO2 emissions, as well as being different additives, either in the saturating gas or in the sonolyzed water,
able to be utilized in fuel cells to generate electricity [2–4]. Fossil fuels can be used to increase the efficiency of the sonolytic yielding of
are now the primary source of hydrogen production, which can be hydrogen. For example, the inclusion of hydrocarbons such as methane
accomplished by a variety of techniques like gasification, vapor and ethane results in a weighty increase in the production of hydrogen
reforming and partial oxidation [1,5,6]. Alternative methods for [23]. In the hot gas phase of the bubble, methane and ethane can
hydrogen production, such as water electrolysis, biological photosyn­ penetrate, and their pyrolysis results in excessive amounts of hydrogen
thesis and photocatalytic processes have been developed [7–9]. Once [23]. Alcohols such as conventional and bio-methanol, on the other
produced, hydrogen is a clean synthetic fuel: when burnt with oxygen, hand, serve the same role as CH4 within the bubble, following the same
the only exhaust gas is water vapor, but when burnt with air, lean

* Corresponding author.
E-mail address: s.merouani03@gmail.com (S. Merouani).

https://doi.org/10.1016/j.cep.2022.109080
Received 3 April 2022; Received in revised form 18 July 2022; Accepted 28 July 2022
Available online 30 July 2022
0255-2701/© 2022 Elsevier B.V. All rights reserved.
A. Dehane et al. Chemical Engineering and Processing - Process Intensification 179 (2022) 109080

Nomenclature Tmax Maximum temperature inside a bubble, (K).


T∞ Bulk liquid temperature, (K).
Af (Ar) Pre-exponential factor of the forward (reverse) reaction, x Thermal diffusivity inside the bubble, (m2 s− 1)
[(cm3 mol− 1 s− 1) for two body reaction and (cm6 mol− 2 Cp Heat capacity concentration inside the bubble, (J m− 3 K− 1)
s− 1) for three body reaction]. MH2O Molar mass of water (kg mol− 1).
bf (br) Temperature exponent of the forward (reverse) reaction. ṁ Evaporation-condensation rate of water, (kg m− 2 s− 1).
c Speed of sound in the liquid medium, (m s− 1). Q Energy transferred by heat exchange, (J s− 1)
Eaf (Ear) Activation energy of the forward (reverse) reaction, (cal n Molar amount, (mol).
mol− 1). PB Liquid pressure on the external side of the bubble wall,
f Frequency of ultrasonic wave, (Hz). (Pa).
Ia Acoustic intensity of ultrasonic irradiation, (W m− 2). V Volume of the bubble, (m3).
kf (kr) Forward (reverse) reaction constant, [(cm3 mol− 1 s− 1) for
two body reaction and (cm6 mol− 2 s− 1) for three body Greek letters
reaction]. σ Surface tension of liquid water, (N m− 1).
p Pressure inside a bubble, (Pa). ρl Density of liquid water, (kg m− 3).
pmax Maximum pressure inside a bubble (Pa). ρg Density inside the bubble (kg m− 3).
p∞ Ambient static pressure, (Pa). λmix Thermal conductivity of the mixture (W m− 1 K− 1).
PA Amplitude of the acoustic pressure, (Pa). λi Thermal conductivity of species i (W m− 1 K− 1).
Pv Vapor pressure of water, (Pa). µ Dynamic viscosity (Pa s).
R Radius of the bubble, (m). α Accommodation coefficient.
Rmax Maximum radius of the bubble, (m). ρH2O Density of water vapor inside the bubble (kg m− 3).
R0 Ambient bubble radius, (m). ρsat,H2O Saturated vapor density (kg m− 3).
Rg Ideal gas constant (J mol− 1 K− 1). υki Stoichiometric coefficient of the kth chemical species in the
t Time, (s). ith reaction.
T Temperature inside a bubble, (K). U̇k Production rate of the kth species (mol s− 1 m− 3).

pyrolytic pathway [24,25]. The concentration of these additives, on the 40% of CH3OH. Kerboua et al. [19] also reported that a 40% methanol
other hand, should be carefully controlled because an excessive con­ concentration and a 40% molar fraction for argon provide the maximum
centration could reduce (or suppress) the yield of hydrogen production. molar output. Kerboua et al. [29] determined that the optimal combi­
On the other hand, hybrid technologies, such as sonocatalysis, photo­ nation for H2 generation and CH3OH conversion is achieved for a so­
sonolysis and sonophotocatalysis are being developed to improve the lution with 20% methanol (v/v) saturated with 70% molar of argon,
efficiency of the sono-generated hydrogen. An interesting review tracing thereby 99.9% of the methanol is removed and about 60% of the
the different studies on hydrogen production by ultrasound has been methanol is converted to hydrogen. Dehane et al. [31] used a compre­
provided by Merouani and Hamdaoui [12]. hensive mathematical model for single bubble sono-pyrolysis of meth­
The intensification of hydrogen sono-production (from water anol to investigate the effect of methanol dosage (in the bulk phase) on
sonolysis) using alcohols dilute aqueous solution has been proven the maximum sonochemical performance for hydrogen production,
experimentally. Buettner et al. [26] have studied the sonolysis of CH3OH consumption and the range of active bubbles. They discovered
water-methanol mixtures at 1 MHz under argon atmosphere. In deion­ that the aqueous phase methanol dosage (0-100 % (v/v)) has little
ized water, H2 was generated at 22 µM/min. When methanol is added to impact on the spectrum of active bubble sizes for CH3OH consumption.
the solution, the H2 rate increased to 100 µM/min at 5% (v/v) of In the case of hydrogen generation, however, for methanol concentra­
methanol and 150 µM/min at 10% (v/v) of methanol, but as methanol tions more than 20%, the active bubble ranges gradually diminish. At
higher concentrations rise, the molar yield of H2 progressively goes 80% argon (in the initial gas composition) and a methanol content of 7
down until no H2 was detected at roughly 80% (v/v) of methanol. When to 20%, the maximum efficacy for CH3OH conversion, hydrogen pro­
using 38 kHz ultrasound in water and water/ethanol (20% v/v) under duction and the width of active bubble sizes (for H2 generation and
argon saturation, Penconi et al. [27] observed a 1.4-fold increase in the CH3OH decomposition) is reached. Nevertheless, the Dehan’s investi­
rate of H2 generation. Rassokhin et al. [24] have noted that the gener­ gation [31] is not complete as all results were done for single matrix of
ation rates of the sonolysis products (e.g. H2, CH4 and CO) of methanol operational conditions (frequency: 355 kHz, In = 1 W/cm2 and Tliq =
increase (H2: from 33.3 to 167.0 µM/min) with the increase of its con­ 20◦ C). According to the available theoretical studies [19,29,31] focusing
centration from 0.001 to 0.50 M in the sonicated solutions, where the on methanol sono-decomposition, the effect of operational conditions
molar yield of these species is as: H2>CO>CH4. Rassokhin et al. [24] (ultrasound frequency, acoustic intensity and the bulk temperature)
have shown that depending on the mole fraction of methanol in the bulk have not covered yet.
solution, an optimum medium temperature is retrieved for the maximal In the present paper, the sono-production of hydrogen from meth­
hydrogen production ([H2]max = 200 µM/min at ~30◦ C and xMeOH= anol sono-conversion in a single bubble is analyzed at varying ultra­
0.018), where the increase of methanol concentration decreases the sound frequencies (213-1000 kHz), acoustic intensities (1 and 2 W/cm2)
temperature at which the maximal H2 formation rate is reached. and liquid temperatures (293.15-333.15 K). The model developed early
Several numerical studies have focused on hydrogen synthesis using [31] by our research group is adopted for this task. To the best of our
sonochemistry [12,14,28–30]. These studies have shown promising knowledge, the effects of these operational conditions have not been
outcomes of clean and efficient hydrogen generation. Merouani et al. treated previously. For this purpose, a reaction mechanism for methanol
[12] and Kerboua et al. [17] have recently published fascinating reviews combustion within the cavity has been adopted. The model was firstly
on hydrogen generation using sonochemistry. Kerboua et al. [19] stud­ validated by some available experimental data before assessing the pa­
ied numerically the influence of the saturating atmosphere (O2 and Ar) rameters’ influence. It should be stressed here that a deep analysis of the
on the sono-pyrolysis of CH3OH, finding that regardless of the quantity effects of these operating conditions (wave frequency, acoustic power,
of argon, a maximum molar yield was obtained for a solution comprising and liquid temperature) enables us to better understand their impacts on

2
A. Dehane et al. Chemical Engineering and Processing - Process Intensification 179 (2022) 109080

the formation of H2 from CH3OH sono-conversion as well as to deter­ between the bubble and the neighboring liquid. Table 1 outlines the
mine the appropriate conditions (of ultrasound frequency, acoustic in­ equations that govern the model. All numerical calculations were done
tensity and liquid temperature) for the maximal production for a bubble in Ar-O2-saturated water with 20 (vol.)% methanol. The
(intensification) of H2 (from methanol sono-combustion). interactions between bubbles are disregarded. Tables 2 and 3 indicate
the reaction pathways utilized to analyze the interior bubble chemistry
2. Model for the cases of Ar-O2-bubble (0% of methanol, Table 2, 38 reactions)
and Ar-O2-CH3OH bubble (Table 3, 32 reactions). The physical proper­
In the presence of CH3OH, the single-bubble model we developed ties of the surrounding liquid in presence of methanol (20%, v/v) are
early [31] is adopted herein. The essential points of this model are given calculated according to the molar fraction contribution of water and
in this section. Based on ordinary differential equations, our model takes methanol (see Table 4). Table 1 provides the following main equations:
into consideration chemical processes (of both water and methanol
pyrolysis), methanol and water molecules’ non-equilibrium condensa­ 1 Eq. 1 (the Keller-Miksis modified model [32]) describes the
tion and vaporization at the bubble wall, as well as heat transfer bubble-radius dynamics, R(t), in an aqueous methanol solution.

Table 1
Principal equations of the model (see detail in Refs. [31,35])*.
Bubble dynamics:
( ) ( ) ( →)[ ( ( )) ]
R̂ ṁ 3 2 R 2ṁ 1 R R
1− + RR̈ + Ṙ 1 − + = 1+ PB (t) − PA sin 2πf t + − P∞
c CpL 2 3C 3cL ρL C C
(Eq. 1)
( ) ( )
m̈R Ṙ ṁ ṁ ṁ Ṙṁ R dPB
+ 1− + + Ṙ + + +
ρL C CρL ρL 2ρL 2CρL CρL dt
- Pressure at the external bubble wall:
2σ 4μṘ
PB (t) = P(t) − − (Eq. 2)
R R
- Bubble pressure and Temperature:
nRg T an2
P(t) = + (Eq. 3)
V − nb) V2
2
( an
E+
V
T= (Eq. 4)
Cv nt
Mass transfer (water vapor and methanol condensation and evaporation):
{Psat,i [R] − Pi
ṁ = α √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ i = H2 O or CH3 OH (Eq. 5)
2πT[R] Rg
Mi
Heat transfer (thermal conduction):
(Tliq− T)
Q̇ = 4πR2 λmix (Eq. 6)
Lth
}
{ R √̅̅̅̅̅̅
RX
Lth = min ′ (Eq. 7)
π R
(n ) ( ) ( ) ( )
nAr nCH3OH nO2
(Eq. 8)
H2 O
λmix = λH2 O (T) + λAr (T) + λCH3OH (T) + λO2 (T)
nt nt nt nt
Internal bubble energy:
ṁ,H2O (TIiq − T) 4 3 ∑n ṁ,CH3OH
ΔE = − P(t)ΔV(t) + 4πR2 Δt eH2O + 4πR2 Δtλ − πR Δt i=1 ΔHi ri + 4πR2 Δt eCH3OH (Eq. 9)
MH2O Lth 3 MCH3OH
Change in species quantities (mol)
- For H2O and CH3OH:
ṁ,i
ni (t +Δt) = ni (t) + 4πR2 Δt + VΔtU̇i i = H2 O or CH3 OH (Eq. 10)
Mi
- For other species k (except Ar):
nk (T + Δt) = nk (T) + VΔtUk (Eq. 11)
Where:
1 dnk ∑I
U̇k = (υ′′ − υ )ri (k = 1, …, K) (Eq. 12)

v dt i=1
∏K ∏ x
ri = kfi k=1 [Xk ]υ ki − kri Kk=1 [Xk ]υki (Eq. 13)

( Ea )
kfi = Afi Tbfi exp − fi
(Eq. 14)
Rg T
( E )
a
kri = Ari Tbri exp − ri
(Eq. 15)
Rg T

*Variables description: dots denoted here time derivative (d/dt), R is the bubble radius, C is the sound speed in the medium (water), ρL is the liquid density, ṁ is the net
rate of evaporation per unit area and unit time and P∞ is the ambient static pressure. PA is the acoustic amplitude (linked to the acoustic intensity Ia by: PA=(2IaρLC)1/2),
PB(t) is the liquid pressure at the liquid side of the bubble, P(t) is the pressure inside the bubble. σ is the surface tension, μ is the liquid viscosity), f is the sound
frequency, Pv is the vapor pressure within the bubble, a and b (in Eqs. 3and 4) are the Van de Waals constants (given in [36]), Rg is the universal gas constant. V is the
volume of the bubble [V=4/3(πR3)], T is the temperature inside the bubble, E is the internal bubble energy, Psat[R] is the saturated vapor pressure (calculated by using
Antoine’s equation) at the interface temperature T[R] = Tliq, MH2O or CH3OH is the molecular weight of water vapor or methanol. ‘α’ is the accommodation coefficient
(given in [29]). λmix, χ and Lth are the heat conductivity, thermal diffusivity of the gas mixture and the thickness of the thermal boundary layer, respectively. [In­
dividual λi of gases [36–38]: λH2O(T) = 9.967213 × 10− 5T - 1.1705 × 10− 2, λAr(T) = 3.5887 × 10− 5T + 6.81277 × 10− 3, λCH3OH(T) = -0.0161+1.02 × 10− 4T, λO2(T)=
7.2211 × 10− 3+6.4478 × 10− 5T, χ =[λmix/Cp]. Cp is the heat capacity concentration (J m− 3 K− 1) for H2O, Ar and CH3OH mixture, Cv is the molar heat of gases and
vapor in the bubble [Cv = (3/2)Rg for monoatomic gases (Ar, H…), (5/2)Rg for diatomic gases (O2, N2, …) and (6/2)Rg for triatomic gases]. ΔHi and ri are the enthalpy
change and the rate of the ith reaction, respectively, and eH2O (eCH3OH) is the energy transported by 1 mole of an evaporating or condensing water vapor (methanol) [e,i
= Cv,iT, i=H2O or CH3OH], U̇i (U̇k ) is the production rate of H2O or CH3OH (kth species) within the bubble. It should be noted that the physical properties of liquid in
presence of methanol are calculated according to the molar fraction contribution of water and methanol (20%, v/v) (see Table 4).

3
A. Dehane et al. Chemical Engineering and Processing - Process Intensification 179 (2022) 109080

Table 2 Table 3
Scheme of the possible chemical reactions inside a collapsing Ar-O2-bubble [39, Scheme of the possible chemical reactions inside a collapsing Ar-O2-CH3OH
40]. M is the third Body. A is in (m3 mol− 1 s− 1) for two body reaction [(m6 mol− 2 bubble [19,29,41,42]. M is the third Body. A is in (m3 mol− 1 s− 1) for two body
s− 1) for a three body reaction], and Ea is in (kJ mol− 1) and ΔH in (kJ mol− 1). reaction [(m6 mol− 2 s− 1) for a three body reaction], and Ea is in (kJ mol− 1) and
N◦ Reaction A n Ea ΔH
ΔH in (kJ mol− 1).
7 N◦ Reaction A n Ea ΔH
1 H2O+M → H + OH+M
• •
1.912 × 10 -1.83 28.35 508.82
2 O2+M → O+O+M 4.515 × -0.64 28.44 505.4 1 H2O+M → H•+•OH+M 1.912 × -1.83 28.35 508.82
1011 107
3 •
OH+M →O+H•+M 9.88 × 1011 -0.74 24.43 436.23 2 CH3OH + O2 → CH2OH 2.0 × 107 0.0 1.88 × 105 196.492
4 H•+O2 →O+•OH 1.915 × 108 0.0 3.93 69.17 + H2O2
5 H•+O2 +M →HO•2 +M 1.475 0.6 0.0 -204.80 3 CH3OH + HO2→ 8.0 × 107 0.0 8.11 × 104 -35.508
6 O+H2O → •OH+•OH 2.97 2.02 3.21 72.59 CH2OH• + H2O2
7 HO•2+H• → H2+O2 1.66 × 107 0.0 1.97 × 10− 1
-239.67 4 CH3OH + H• → CH2OH• 1.35 × 3.2 1.46 × 104 -34.057
8 HO•2+H• →•OH+•OH 7.079 × 107 0.0 7.06 × 10− 2
-162.26 + H2 10− 3
9 HO•2+O → •OH+O2 3.25 × 107 0.0 0.0 -231.85 5 •
OH + •OH → HO•2 + H• 1.08 × 0.61 1.51 × 154.957
10 HO•2+•OH → H2O+O2 2.89 × 107 0.0 - 1.19 × -304.44 1005 1005
10− 1 6 HO•2 + •OH → H2O+ O2 2.89 × 0.00 − 2.08 × − 291.451
11 H2+M → H•+H•+M 4.577 × -1.4 24.98 444.47 1007 1003
1013 7 HO•2 + O → •OH + O2 2.00 × 0.00 0.00 − 224.131
12 O+H2 → H•+•OH 3.82 × 106 0.0 1.9 8.23 1007
13 •
OH+H2 → H•+H2O 2.16 × 102 1.52 8.25 × 10− 1
-64.35 8 HO•2 + H → H2 O+ O

3.10 × 0.00 7.20 × –22.957
14 H2O2+O2 → HO•2+HO•2 4.634 × -0.35 12.12 175.35 1007 1003
1010 9 HO•2 + H → H2 + O2

1.66 × 0.00 3.44 × − 230.293
15 H2O2+M → 2.951 × 108 0.0 11.59 217.89 1007 1003

OH+•OH+M 10 HO•2 + H• → 2 •OH 7.08 × 0.00 1.23 × − 155.613
16 H2O2+H• → H2O+•OH 2.410 × 107 0.0 9.5 × 10− 1
-290.93 1007 1003
17 H2O2+H• → H2+HO•2 6.025 × 107 0.0 1.9 -64.32 11 CH2OH• + O2 → CH2O+ 5.00 × 0.00 0.00 − 79.404
1
18 H2O2+O → •OH+HO•2 9.550 2.0 9.5 × 10− -56.08 HO•2 1006
19 H2O2+•OH → 1.0 × 106 0.0 0.0 -128.67 12 CH3OH + •OH → 4.4 × 1000 2.0 6.3 × 10 3
-54.215
H2O+HO•2 CH3O•+ H2O
10
20 H•+•OH+M → H2O+M 2.2 × 10 -2.0 0.0 - 508.82 13 CH3OH + •OH → 1.44 × 2.0 -3.51 × -95.215
21 O+O+M → O2+M 6.165 × 103 -0.5 0.0 -505.4 CH2OH• + H2O 1000 103
22 O+H•+M → •OH+M 4.714 × 106 -1.0 0.0 -436.23 14 CH3O•+ M → CH2OH• + 1.00 × 0.00 8.00 × − 38.000
23 O+•OH → H•+O2 5.481 × 105 0.39 -7.01 × -69.17 M 1008 1004
10− 2 15 CH3O•+ M → CH2O+ 7.78 × 0.00 5.65 × 126.297
24 HO•2 +M → H•+O2 +M 3.09 × 106 0.53 11.7 204.80 H•+ M 1007 1004
25 •
OH+•OH → O+H2O 1.465 × 2.11 -6.94 × -72.59 16 CH3O•+ O2 → CH2O+ 4.28 × 7.60 − 1.48 × − 117.404
10− 1 10− 1 HO•2 10− 19 1004
26 H2+O2 → HO2• +H •
3.164 × 106 0.35 13.3 239.67 17 CH2O+ HO•2 → HCO• + 4.11 × 2.50 4.27 × − 48.976
27 •
OH+•OH → HO•2+H• 2.027 × 104 0.72 8.8 162.26 H2O2 10− 02 1004
28 •
OH+O2 → HO•2+O 3.252 × 106 0.33 12.75 231.85 18 CH2O+ •OH → HCO• + 3.90 × 0.89 1.70 × − 128.159
29 H2O+O2 → HO•2+•OH 5.861 × 107 0.24 16.53 304.44 H2O 1004 1003
30 H•+H•+M → H2+M 1.146 × 108 -1.68 1.96 × 10− 1
- 444.47 19 CH2O+ O → HCO• + 3.50 × 0.00 1.47 × − 60.839
31 H•+•OH → O+H2 2.667 × 2.65 1.17 - 8.23 •
OH 1007 1004
10− 2 20 CH2O+ H• → HCO •+ H2 5.74 × 1.90 1.15 × − 67.001
32 H•+H2O → •OH+H2 2.298 × 103 1.40 4.38 64.35 1001 1004
33 HO•2+HO•2 → H2O2+O2 4.2 × 108 0.0 2.87 - 175.35 21 HCO• + O2 → CO + HO•2 7.58 × 0.00 1.72 × − 140.530
34 •
OH+•OH+M → 1.0 × 102 -0.37 0.0 - 217.89 1006 1003
H2O2+M 22 HCO + M → CO + H +
• •
1.86 × − 1.00 7.11 × − 65.171
35 H2O+•OH → H2O2+H• 1.269 × 102 1.31 17.08 290.93 M 1011 1004
36 H2+HO•2 → H2O2+H• 1.041 × 105 0.70 5.74 64.32 23 CO + •OH → CO + H•+ 4.40 × 1.50 − 3.10 × − 212.853
37 •
OH+HO•2 → H2O2+O 8.66 × 10− 3 2.68 4.45 56.08 CO2 1000 1003
38 H2O+HO•2 → 1.838 × 104 0.59 7.4 128.67 24 2 HO•2 → H2O2 + O2 3.02 × 0.00 5.80 × − 212.268
H2O2+•OH 1006 1003
25 H2O2 + M → 2 •OH + M 8.15 × − 1.92 2.08 × 262.356
Third body efficiency factors: R1: αH2 = 0.73, αH2O = 12, αAr = 0.38; R2: αH2 =
k0 1017 − 1.39 1005
2.5, αH2O = 12, αAr = 0.75; R3: αH2 = 2.5, αH2O = 12, αAr = 0.83; R5: αH2 = 1.3, k∞ 2.62 × 2.15 ×
αH2O = 14, αAr = 0.67; R11: αH2 =2.5, αH2O=12. R15: αH2 = 2.5, αH2O = 12, αAr = 1013 1005
0.64; R20: αH2 = 0.73, αH2O = 12, αAr = 0.38; R21: αH2 = 2.5, αH2O = 12, αAr = 26 H•+ O2 + M → HO•2 + M 5.75 × − 1.40 0.00 − 205.701
0.75; R22: αH2 = 2.5, αH2O = 12, αAr = 0.83; R24: αH2 = 1.3, αH2O = 14, αAr = 0.67; k0 1007 0.44 0.00
R30: αH2 =2.5, αH2O=12; R34: αH2 = 2.5, αH2O = 12, αAr = 0.64. k∞ 4.65 ×
1000
27 H2 + O → •OH + H• 5.06 × 2.67 2.63 × 6.162
2 Eqs. 3 and 4 provide the internal bubble pressure and temperature 10− 02 1004
during oscillation. 28 H + O2 → OH + O
• •
3.52 × − 0.70 7.14 × 68.518
3 Eq. 5 describes the Hertz-Knudsen model [33] for water vapor and 1010 1004
29 CH2OH•+ H• → CH2O+ 3 × 107 0.0 0.0 -309.578
methanol mass transport.
H2
4 Eqs. 6-8 (heat dissipation by conduction [34]) describe the heat ex­ 30 2 •OH → HO•2 1.08 × 0.61 1.51 × 105 154.957
change dQ/dt inside and outside the bubble during oscillation. 105
5 Eq. 9 describes the change vs. time of the internal bubble energy. 31 CH2OH + OH →
• •
2.4 × 107 0.0 0.0 -370.7
6 Eqs. 10-15 describe the change, with time, in quantities of H2O and CH2O+ H2O
32 •
CH3+ HO•2 → CH4+O2 3.61 × 0.0 0.0 -233.852
all other species ‘k’ within the bubble during oscillation. Through
106
applying Eqs. (12)-(15), the temporal production rates (Uk) of all
species (10 species) within the bubble can be established as function Third body efficiency factors: R1:αH2=2.5, αH2O =16, αAr =1, αCO =1.9, αCO2
of concentrations and kinetic parameters (rate constants, k) as (ac­ =3.8, αCH4 =16, αCH3OH =5; R7: αH2= 2, αH2O =6, αAr =0.7, αCO =1.5, αCO2 =2,
αCH4 =, others 1; R15:αH2=2, αH2O =6, αAr =0.7, αCO =1.5, αCO2 =2, αCH4 =2,
cording to Table 2):

4
A. Dehane et al. Chemical Engineering and Processing - Process Intensification 179 (2022) 109080

others 1. R22:αH2=1.9, αH2O =12, αAr =, αCO =2.5, αCO2 =2.5, αCH4 =, others 1; kHz [44], 3 µm for 515 kHz [45] and 2 µm for 1000 kHz [44] were used
R25:αH2=2, αH2O =6, αAr =0.4, αCO =1.5, αCO2 =2, αCH4 =, others 1; R26:αH2=2.5, in the numerical simulation of this study (see also Refs. [21,46,47]).
αH2O =16, αAr =0.7, αCO =1.2, αCO2 =2.4, αCH4 =, others 1. These values are the mean ambient bubble radii of active bubbles at the
given frequencies (determined experimentally). Several numerical in­
vestigations [22,48–50] based on single bubble sonochemistry models
Table 4
supported the R0 values adopted in this study. The variation of the
Physical properties of methanol and water at 293.15 K and 1 atm [43].
bubble temperature (Eq. 4), pressure (Eq. 3), radius and the wall ve­
Compound Surface tension Viscosity Sound velocity Density (kg/
locity (Eq. 1), as well as the mole numbers of all species (except argon,
(N/m) (Pa.s) (m/s) m3)
which is chemically inert) in the bubble (Eqs. 10, 11) are all outputs of
2
Methanol 2.256 × 10− 5.7548 × 1116.2 791.02 Table 1 equations, which are applied throughout the bubble lifetime.
10− 4
− 2 3 The initial pressures of Ar, O2, H2O and CH3OH inside the bubble fulfill
Water 7.245 × 10 1.0 × 10− 1482.0 998.12
the equilibrium relation: P0 = P∞ +2σ /R0, where P0 is the internal
bubble pressure (Pv,0+PCH3OH,0+PAr,0+ PO2,0) at equilibrium, R=R0 (PAr,
• U̇H2O = dn
V dt = -{kf1[H2O][M]- kf20[H][OH][M]} –{kf6[H2O][O]-
H2O
0+ PO2,0: initial gas pressure “Pg,0”). The initial pressure of CH3OH
kf25[OH][OH]}+ {kf10[HO2][OH]- kf29[H2O][O2]}+ {kf13[OH] within the bubble is governed by Raoult’s law: PCH3OH,0=XCH3OH,0 × Psat,
[H2]- kf32[H][H2O]} +{kf16[H2O2][H]-kf35[H2O][OH]} + CH3OH,0 (Psat,CH3OH,0=0.13 bar at 293.15 K), where X CH3OH,0 is the
{kf19[H2O2][OH]-kf38[H2O][HO2]}. CH3OH mole fraction in the aqueous phase (from 0 to 1). The physical
• U̇O2 = dn properties of methanol and water are given in Table 4. The water vapor
V dt = -{kf2[O2][M]- kf21[O][O][M]} – {kf4[H][O2]-kf23[O]
O2

pressure (Pv,0) is calculated through Antoine’s equation (2.33 kPa at


[OH]}- {kf5[H][O2][M]-kf24[HO2][M]}+ {kf7[HO2][H]-kf26[H2]
293.15 K [51]). The initial gas pressure Pg,0 is calculated as Pg,0 = P0-(Pv,
[O2]}+ {kf9[HO2][O]-kf28[OH][O2]}+ {kf10[HO2][OH]-kf29[H2O]
0+P CH3OH,0). Initial mole fractions of CH3OH, H2O, Ar and O2 were
[O2]}- {kf14[H2O2][O2]-kf33[HO2][HO2]}
calculated as Pi,0/P0, where i = CH3OH, H2O, Ar and O2. More details are
• U̇OH = dn OH
V dt = {kf1[H2O][M]-kf20[H][OH][M]}- {kf3[OH][M]-
given in Ref. [31].
kf22[O][H][M]}+ {kf4[H][O2]-kf23[O][OH]}+2 × {kf6[O][H2O]-
kf25[OH][OH]}+ 2 × {kf8[HO2][H]-kf26[OH][OH]}+ {kf9[HO2] 3. Results and discussions
[O]-kf28[OH][O2]}- {kf10[HO2][OH]-kf29[H2O][O2]}+{kf12[O]
[H2]-kf31[H][OH]}- {kf13[OH][H2]-kf32[H][H2O]}+ 2 × According to our previous work [31], the maximal H2 yield and
{kf15[H2O2][M]-kf34[OH][OH][M]}+ {kf16[H2O2][H]-kf35[H2O] CH3OH conversion could be achieved at a specific methanol aqueous
[OH]} + {kf18[H2O2][O]-kf37[OH][HO2]}- {kf19[H2O2][OH]- dosage of 20 (vol)% with 80% (molar %) of argon inside the bubble.
kf38[H2O][HO2]} Consequently, to facilitate our task (analyzing of frequency, temperature
• U̇H = Vdndt
H
= {kf1[H2O][M]-kf20[H][OH][M]}+{kf3[OH][M]- and sound intensity impacts), throughout the present study, the aqueous
kf22[O][H][M]} – {kf4[H][O2]-kf23[O][OH]} – {kf5[H][O2][M]- concentration of methanol is fixed at 20% with a bubble initial gas
kf24[HO2][M]}- {kf7[HO2][H]-kf26[H2][O2]} –{kf8[HO2][H]- content of 80% argon and 20% oxygen (the rest is for methanol and
kf27[OH][OH]}+2 × {kf11[H2][M]-kf30[H][H][M]}+{kf12[O] water vapor). The effects of the operating conditions were investigated
[H2]-kf31[H][OH]}+ {kf13[OH][H2]-kf32[H][H2O]} – {kf16[H2O2] by the variation of the ultrasound frequency (213, 355, 515 and 100
[H]-kf35[H2O][OH]}-{kf17[H2O2][H]-kf36[H2][HO2]}. kHz), acoustic intensity (0.7, 1 and 1.5 W/cm2) and liquid temperature
• U̇O = Vdndt
O
= 2 × {kf2[O2][M] –kf21[O][O][M]}+ {kf3[OH][M]- (293.15 to 333.15 K). Accordingly, their values are injected in the
kf22[O][H][M]}+ {kf4[H][O2]-kf23[O][OH]} - {kf6[O][H2O]- mathematical model resumed in Table 1.
kf25[OH][OH]]}- {kf9[HO2][O]-kf28[OH][O2]} – {kf12[O][H2]-
kf31[H][OH]} – {kf18[H2O2][O]-kf37[OH][HO2]}. 3.1. Model validation
• U̇HO2 = dn
V dt = {kf5[H][O2][M]-kf24[HO2][M]} - {kf7[HO2][H]-
HO2

kf26[H2][O2]} - {kf8[HO2][H]- kf27[OH][OH]} –{kf9[HO2][O]- Before going into the various findings in this paper, we would want
kf28[OH][O2]}- {kf10[HO2][OH]-kf29[H2O][O2]} + 2 × to compare the predicted bubble temperatures achieved by our model to
{kf14[H2O2][O2]-kf33[HO2][HO2]}+ {kf17[H2O2][H]-kf36[H2] those estimated empirically by Rae et al. [52] (by the methyl radical
[HO2]}+ {kf18[H2O2][O]-kf37[OH][HO2]}+ {kf19[H2O2][OH]- recombination technique) for an argon-methanol aqueous solution at
kf38[H2O][HO2]}. varying dosages of the alcohol (0-500 mM), where the frequency, the
• U̇H2 = dn intensity and the bulk solution temperature were 355 kHz, 2W/cm2 and
V dt = {kf7[HO2][H]-kf26[H2][O2]} – {kf11[H2][M]-kf30[H]
H2

~293.15 K. According to Fig. 1(a), our model goodly captures the data
[H][M]}- {kf12[O][H2]-kf31[H][OH]} – {kf13[OH][H2]-kf32[H]
of methanol effect on the bubble temperature, particularly for lower
[H2O]}+ {kf17[H2O2][H]-kf36[H2][HO2]}.
dnH2O2
alcohol dosages. For 0 mM methanol, the model provided 4570.2 K
• U̇H2O2 = V dt = - {kf14[H2O2][O2]-kf33[HO2][HO2]}-
against 4600 K for Rae et al. data, showing a deviation of <1%. Simi­
{kf15[H2O2][M]-kf34[OH][OH][M]}- {kf16[H2O2][H]-kf35[H2O] larly, lower deviations of ~1.8% and 2.5% are also marked for 60 and
[OH]}- {kf17[H2O2][H]-kf36[H2][HO2]}- {kf18[H2O2][O]-kf37[OH] 150 mM of methanol, respectively. However, a relatively higher devi­
[HO2]}- {kf19[H2O2][OH]-kf38[H2O][HO2]}. ation is recorded for greater CH3OH concentration (> 200 mM).
• U̇Ar = VdnArdt = 0. Nevertheless, the global tendency of the bubble temperature variation
with the alcohol dosage was excellent followed by the model.
The above kinetical equations system (derived from Eq. 12) is It should be noted that there is no peak temperature at 100 mM of
coupled with the two differential equation derived from Eq. 1 (i.e., MeOH, because above 50 mM of MeOH in solution, the experimental
providing dR/dt and d2R/dt2) and Eqs. (4)-(6), (10) and (11), and the bubble temperature fluctuated around a near plateau value (3500 K).
system of all these differential equations are solved using finite differ­ This phenomenon (trend) is confirmed experimentally by several re­
ence method (using FORTRAN 95 software). The initial conditions are: searchers where the bubble temperature was estimated using different
at t = 0, R = R0, dR/dt = d2R/dt2 = 0, T=T∞, dPdtB = 0, dm alcohols (methanol, ethanol, n-propanol, n-butanol and tert-butanol) at
2
d m
dt = 0, dt2 = 0. A
time step of 10− 4 µs is used for the resolution of the system of the dif­ different frequencies [52–54]. On the other hand, the predicted bubble
ferential equation. On the other hand, fixed experimental values of temperature (using our model) is that achieved at the end of the bubble
ambient bubble radii of R0 =3.9 µm for 213 kHz [44], 3.2 µm for 355 collapse (the maximum bubble temperature), without considering the
interactions between bubbles in the sonicated medium (isolated bubble,

5
A. Dehane et al. Chemical Engineering and Processing - Process Intensification 179 (2022) 109080

confronting the results of H2 production obtained theoretically using our


model to those retrieved experimentally by Buttner et al. [26] at 1000
kHz and an acoustic intensity of 2 W/cm2. In Fig. 2(b), the normalized
yields of H2 are given over a range of 0 % to 80 % (v/v) for an aqueous
solution of methanol. As it can be seen, the outputs of our chemical
mechanism (Ar-CH3OH-bubble, Table 2) are in good concordance with
those obtained empirically by Buttner et al. [26] for the whole range of
methanol volume concentration (0-80%).

3.2. Effect of ultrasound frequency on hydrogen production and methanol


conversion

In Figs. 2(a)-(d), the yield of H2 production and CH3OH conversion


are shown as a function of ultrasound frequency (from 213 to 1000 kHz)
at 1 W cm− 2 in the presence (20%, v/v) and absence of methanol in the
liquid. First, as it was expected, the increase of the wave frequency af­
fects negatively (decreases) the molar production of hydrogen and the
peak temperature of the bubble (and methanol conversion when it is
present in solution, Figs. 2(b,c)) either in the presence or absence of
methanol, Fig. 2(a). This trend was largely reported in different exper­
imental and theoretical inquiries [44,47,55,56]. This behavior is logi­
cally ascribed to the reduction of the collapse intensity as well as the
molar production of the acoustic cavitation with the rise of frequency
[47,55,57]. In other words, as the wave frequency increases (the
acoustic period shortens), less water vapor (and methanol molecules, if
present in solution) enters the acoustic cavitation during the expansion
phase, resulting in a milder collapse. Consequently, a lower bubble
temperature will be obtained at the end of bubble compression, which
means that a lesser amount of hydrogen (and methanol conversion)
molecules will be obtained via water vapor dissociation (and methanol
combustion). Additionally, the presence of methanol (20% in solution)
decreases more the maximal temperature of the bubble (especially at
213 and 355 kHz) through the increase of its heat capacity and the
Fig. 1. Comparison between (a) our simulated bubble temperature and those
retrieved experimentally by Rea et al. [52] at 355 kHz, In = 2 W/cm2, Tliq = endothermic dissociation of methanol [31,58]. However, as it can be
293.15 K, P∞= 1 atm (the R0, ambient radius, used for this simulation is 3.2 µm seen in Fig 2(a), for wave frequencies that are equal to or greater than
[44]), and (b) our normalized simulated production rate of H2 and those ob­ 515 kHz, the peak temperature of the bubble is slightly affected by the
tained experiemebntally by Butner et al. [26] at 1000 kHz and 2 W/cm2 under presence of methanol, where at 515 and 1000 kHz, the maximal tem­
argon atmosphere (the R0, ambient radius, used for this simulation is 2 peratures are 2593 K (2741 K in absence of methanol) and 1481 K (1489
µm [44]). K in absence of methanol), respectively. This may be explained by the
rise of ultrasound frequency, where less amount of methanol is encap­
as indicate in the model section). Adding alcohol into the sonicated sulated within the bubble at the end of the collapse due to the short
solution could increase the methanol portion in the bubble, which could bubble lifetime. Therefore, the intensity of bubble collapse (its maximal
decrease the maximum bubble temperature at certain level of MeOH (i. temperature) will be marginally affected by the presence of small
e., 50 mM), above which no further bubble reduction is obtained. amounts of methanol, which is responsible for the increase in bubble
Therefore, the numerical intergradation could not yield any peak tem­ heat capacity and the number of endothermal reactions within it
perature at 100 mM, because, simply, there is no effective peak tem­ (Table 3). On the other hand, in the presence of methanol, the yield of H2
perature (even experimentally) in the presence of methanol. Therefore, is greater than that retrieved in the absence of methanol independently
the best trend of temperature variation with methanol additions was of the applied frequency. A similar tendency was obtained experimen­
that obtained numerically through our model. tally by Rassokhin et al. [24] at 724 kHz and by Buttner et al. [26] at
Finally, taking into account the maximum difference (280 K at 500 1000 kHz. This tendency (increase of H2 yielding) is mainly ascribed to
mM) observed between our findings and those of Rae et al. [52], in the scavenging effect of methyl alcohol toward hydroxyl radicals (re­
addition to the experimental errors observed in the experimental actions 12 and 13, Table 3) and its reaction with hydrogen atoms (re­
method of Rae et al. (±250 K), it can be concluded that our results are action 4, Table 3). As a result, the recombination reaction of •OH
very acceptable. Therefore, our model is reliable in forecasting the radicals with hydrogen atoms is reduced proportionally with the
bubble temperature, and as a result, it may be trusted to be used to enhancement of hydrogen production. At 213 kHz, a close production of
investigate the impact of the diverse acoustical and medium parameters H2 is obtained for both cases [absence (1.22 × 10− 17 mol) and presence
(i.e., for instance: frequency, intensity and liquid temperature) on the of CH3OH (4.6 × 10− 17 mol)]. This is due to the large decrease in bubble
acoustical generation of hydrogen from methanol sono-conversion temperature (4000 K) in the presence of methanol (which causes the
within the cavitation bubble. Note that with 100 mM CH3OH, the increase in bubble heat capacity and the number of endothermal re­
temperature of the cavity drops from 4610 to 3940 K due to the actions, Table 3). Therefore, lower amount of H2 is produced (at 213
CH3OH-induced augmentation in the heat capacity of the gas mixture in kHz). However, in the presence of methanol and for ultrasound fre­
the bubble, as well as its endothermal dissociation. However, the in­ quencies greater than 355 kHz, a rapid decrease in hydrogen production
crease in alcohol dosage above 100 mM has relatively little influence on is observed: 1.16 × 10− 18 at 355 kHz, 1.7 × 10− 21 at 515 kHz and 1.8 ×
the maximum bubble temperature (Tmax ~3910 K). 10− 31 mol at 1000 kHz.
Further evidence of the precision of our model was done through In Fig 2(b), the maximal conversion of methanol (per collapse) is
calculated as a function of the maximal amount of methanol found at the

6
A. Dehane et al. Chemical Engineering and Processing - Process Intensification 179 (2022) 109080

Fig. 2. Effect of ultrasound frequency (from 213 to 1000 kHz) on (a) the bubble temperature and hydrogen production per collapse, (b) methanol conversion
(calculated according to its maximal amount found at Rmax) and bubble temperature and (c) the quantity of converted methanol, the total yield of the bubble and
methanol conversion percentage (Calculated with respect to the total production of the bubble). Conditions: In= 1 W/cm2, Tliq= 293.15 K and P∞= 1 atm.

end of bubble expansion (Rmax). At first sight, it seems that these con­ amount of water vapor and volatile species (methanol) found at the end
versions (<16.7%) are very low compared to the registered spectrum of of the rarefaction phase. In other words, in spite of the same observed
bubble temperatures (4000-1481 K) over the whole range of frequency temperatures for both cases (in the presence and absence of methanol),
(213-1000 kHz). However, according to Fig 2(c), it is clearly concluded the positive impact of methanol (limitation of H• and •OH recombina­
that important quantities of methanol (compared to the total yield of tion and enhancing the yield of H2 molecules, Table 3) inside the bubble
bubble [from 2.03 × 10− 16 to 1.52 × 10− 22 mol]) are converted espe­ outweighs the decrease in its temperature (resulting from the increase in
cially in the range of ultrasound frequency from 213 (1.73 × 10− 16 mol, its heat capacity and the combustion of methanol). The same trend was
85.22%) to 355 kHz (7.22 × 10− 18 mol, 71.2%). The comparison of Fig 2 shown previously in our study [62] in the presence of volatile CCl4. The
(a) and 2(b) shows that the higher decomposition of methanol and rise of acoustic intensity from 1 to 1.5 W/cm2, increases the yield of H2
hydrogen production may be obtained in the range of ultrasound fre­ (and bubble temperature) from 6.92 × 10− 21 mol (4205 K) to 7.36 ×
quency from 213 to 355 kHz. This conclusion is sustained especially by 10− 18 mol (6646 K), respectively, in the absence of methanol. However,
the increase of the number density proportionally with the increase of in the presence of methanol, it is clearly observed that the increase in the
ultrasound frequency [59–61], which may increase the molar yield of H2 molar production of hydrogen (and the peak temperature of the bubble)
and the conversion of methanol. is amortized despite the rise of acoustic intensity from 1 (H2: 1.16 ×
10− 18 mol, Tmax: 3456 K) to 1.5 W/cm2 (H2: 8.97 × 10− 18 mol, Tmax:
3801 K). This is obviously owing to the negative effect of evaporating
3.3. Effect of acoustic intensity on hydrogen production and methanol methanol (via increasing the bubble heat capacity and the combustion of
conversion CH3OH) within the bubble at 1.5W/cm2 on both the bubble temperature
and the molar yield of H2. This is because at high acoustic intensity,
In Fig 3(a), the molar yield of hydrogen production and bubble higher maximal radii (Rmax) will be attained, which allows the pene­
temperature are shown as functions of acoustic intensity (0.7, 1 and 1.5 tration of huge amounts of water vapor and methanol (more volatile
W/cm2) at 355 kHz. At 0.7 W/cm2, it is observed that approximately the than water) inside the bubble. As a result, a milder implosion will result
same bubble temperatures are retrieved for both cases, in the presence thanks to the existence of these large quantities of water vapor and
and absence of CH3OH, whereas the production of hydrogen in the methanol. A similar tendency was obtained by Ferkous et al. [63] for the
presence of methanol is greater than that obtained without methanol. yield of ●OH radicals at 585, 860 and 1140 kHz and by Yasui et al. [64]
This indicates clearly that the molar production of the acoustic bubble is for the maximal bubble temperature at 1 MHz. This result demonstrates
not only dependent on its peak temperature but also on the maximal

7
A. Dehane et al. Chemical Engineering and Processing - Process Intensification 179 (2022) 109080

Fig. 3. Effect of acoustic intensity on (a) the bubble temperature and hydrogen Fig. 4. Variation of (a) the bubble temperature and hydrogen production and
production and (b) the quantity of converted methanol, the total yield of bubble (b) the quantity of converted methanol, total yield of bubble and the percentage
and the methanol conversion parentage (calculated with respect to the total of converted methanol (calculated according to the total production of bubble).
production of bubble). Conditions: f = 355 kHz, Tliq= 293.15 K and P∞= 1 atm. Conditions: In= 1 W/cm2, f = 355 kHz, and Psta= 1 atm.

the existence of an ideal acoustic intensity (1 W/cm2) for the efficient temperature (4136 K) is observed in the absence of methanol at the
production of hydrogen by ultrasound. liquid temperature of 303.15 K, then this maximal bubble temperature
In Fig 3(b), it is shown that the conversion of methanol is promoted gradually decreased with the rise of the bulk temperature [16.8% of
in the range from 0.7 W/cm2 (1.1 × 10− 21 mol, 0.33% of the total bubble decrease from 303 K (Tmax=4136 K) to 333.15 K (Tmax=3441 K)]. On the
yield) to 1 W/cm2 (7.22 × 10− 18 mol, 71.24% of the total yield of other hand, optimal production of hydrogen (4.26 × 10− 17 mol) has
bubble). However, a further increase of In to 1.5 W/cm2 seems to be been obtained at a liquid temperature of 323.15 K in the absence of
inefficient for methanol decomposition (3.72 × 10− 17 mol, 75.09% of methanol in the solution. This optimum of H2 production results from
the total production of the bubble), mainly due to the lower increase of the competition between the maximal bubble temperature and the
the maximum bubble temperature (3801 K, Fig 2(a)) at this acoustic maximal amount of water vapor found at the end of the expansion phase
intensity (1.5 W/cm2). This agrees with the Alippi et al. [65] and [22,62]. This means that, over a bulk temperature of 323.15 K, the
Weissler et al. [66] findings for the yield of iodine in a sonicated solution decrease in bubble temperature (due to the evaporation of large
of CCl4 for electrical power in the range from 200 to 450 W and from 0 to amounts of water vapor) is unable to maintain the molar increase of H2
600 W, respectively. According to Figs. 2(a) and 3(b), it can be observed production. As a result, an optimum balance between the decrease of
that the acoustic intensity should be accurately controlled for maximal bubble temperature and the increase of H2 formation is retrieved at the
and efficient production of H2 and CH3OH conversion with respect to the liquid temperature of 323.15 K. In contrast, a rapid decrease of the
delivered energy into the sonoreactor. bubble temperature {from 3456 K (at 293.15 K) to 2613.42 K (at 313.15
K)} is observed in the presence of methanol (20% v/v). Nevertheless, the
sono-activity of the bubble is totally suppressed at the liquid tempera­
3.4. Effect of liquid temperature on hydrogen production and methanol
ture of 323.15 and 333.15 K. Accordingly, a maximal hydrogen pro­
conversion
duction of 1.48 × 10− 18 mol is obtained at the liquid temperature of
303.15 K. Our results are consistent with those obtained elsewhere by
In this section, the molar production of hydrogen and methanol
Rassokhin et al. [24] at 724 kHz, where an optimum of H2 production
decomposition are visualized on a range of liquid temperatures (293.15
rate (165 µM/min) is observed at the liquid temperature of ~300 K in
to 333.15 K) at 1 W/cm2 and 355 kHz, Figs. 4(a) and 4(b). It should be
the presence of 22.26% of methanol in the bulk liquid. The suppression
noted that according to the different experimental studies, the highest
of the bubble activity at the liquid temperatures of 323.15 and 333.15 K
sono-activity of the micro-bubble is observed around this frequency
(Fig 4(a) and 3(b)), is owing to the huge amounts of water vapor and
(355 kHz) [67–70]. In Fig 4(a), a turning point of the bubble

8
A. Dehane et al. Chemical Engineering and Processing - Process Intensification 179 (2022) 109080

CH3OH evaporated inside the bubble during the rarefaction phase (data Data Availability Statements
not shown). Therefore, the bubble lifetime at the liquid temperatures of
323.15 and 333.15 K exceeds the acoustic period (2.81 µs at 355 kHz). All data generated or analyzed during this study are included in this
This negative effect of methanol is observed through the sonolysis of an manuscript itself.
appropriate amount of other volatile substrates such as CO2, N2O and
CH4 [58,71–73]. In Fig 4(b), it is seen that the maximal conversion of Declaration of Competing Interest
methanol is obtained at the liquid temperatures of 293.15 K (7.22 ×
10− 18 mol, 71.2% of the total yield) and 303.15 K (4.89 × 10− 18 mol, The authors declare no competing interests.
41.27% of the total yield); however, a critical decrease of methanol
decomposition is observed at the liquid temperature of 313.15 K (6.9 × Data Availability
10− 23 mol). The analysis of Fig 4(a) and 4(b) indicates that the maximal
conversion of methyl alcohol and hydrogen production are obtained in a No data was used for the research described in the article.
narrow range of liquid temperature from 293.15 to 303.15 K, which
means that the sonoactivity of the bubble is very sensitive to the increase
of liquid temperature in the presence of CH3OH in the liquid phase. Acknowledgements

4. Conclusion We acknowledge The Ministry of Higher Education and Scientific


Research of Algeria and the General Directorate of Scientific Research
Methanol sono-conversion and H2 production were investigated on a and Technological Development (GD-SRTD) for their support in con­
range of ultrasound frequency (213-1000 kHz), acoustic intensity (1 and ducting the RFU project A16N01UN250320220002, named
2 W/cm2) and liquid temperature (293.15-333.15 K). Either in the “Développement, scale-up et intensification de procédés innovants
presence or absence of methanol, the peak temperature of the bubble, d’oxydation et de réduction (POA/PRA) pour la destruction rapide des
the hydrogen production and the methanol conversion are decreased micropolluants émergents des effluents aqueux industriels.”
proportionally with the rise of the wave frequency. For ultrasound fre­
quencies that are equal or greater than 515 kHz, the maximal temper­
References
ature of the bubble is slightly affected by the presence of methanol.
Additionally, in the presence of methanol, the yield of H2 is greater than [1] I. Dincer, C. Acar, Review and evaluation of hydrogen production methods for
that obtained in the absence of methanol, independently of the used better sustainability, Int. J. Hydrogen Energy. 40 (2015) 11094–11111, https://
frequency. The highest efficacy of methanol combustion and hydrogen doi.org/10.1016/j.ijhydene.2014.12.035.
[2] S. Merouani, O. Hamdaoui, Correlations between the sonochemical production rate
production is obtained for frequencies ranging from 213 to 355 kHz. On of hydrogen and the maximum temperature and pressure reached in acoustic
the other hand, accurate control of acoustic intensity is needed for bubbles, Arab. J. Sci. Eng. 43 (2018) 6109–6117, https://doi.org/10.1007/s13369-
maximal production of H2 and conversion of CH3OH with respect to the 018-3266-3.
[3] A. Chibani, S. Merouani, C. Bougriou, L. Hamadi, Heat and mass transfer during the
delivered energy in the sonicated solution. Accordingly, an ideal storage of hydrogen in LaNi5-based metal hydride: 2D simulation results for a large
acoustic intensity of 1 W/cm2 was observed for H2 yielding and CH3OH scale, multi-pipes fixed-bed reactor, Int. J. Heat Mass Transf. 147 (2020), 118939,
conversion. https://doi.org/10.1016/j.ijheatmasstransfer.2019.118939.
[4] A. Chibani, C. Bougriou, S. Merouani, Simulation of hydrogen absorption/
In the absence of methanol in the solution, a turning point of bubble desorption on metal hydride LaNi5-H2: Mass and heat transfer, Appl. Therm. Eng.
temperature (4136 K) is observed at the liquid temperature of 303.15 K. 142 (2018) 110–117, https://doi.org/10.1016/j.applthermaleng.2018.06.078.
Moreover, optimal production of hydrogen is obtained at the liquid [5] A. Haryanto, S. Fernando, N. Murali, S. Adhikari, Current Status of Hydrogen
Production Techniques by Steam Reforming of Ethanol : a Review, Energy Fuels 19
temperature of 323.15 K. Conversely, in the presence of methanol, the
(2005) 2098–2106, https://doi.org/10.1021/ef0500538.
peak temperature of the bubble is decreased proportionally with the rise [6] I. Dincer, Green methods for hydrogen production, Int. J. Hydrogen Energy. 37
of liquid temperature from 293.15 to 313.15 K. However, the bubble (2012) 1954–1971, https://doi.org/10.1016/j.ijhydene.2011.03.173.
[7] F. ezzahra Chakik, M. Kaddami, M. Mikou, Effect of operating parameters on
sono-activity is completely suppressed at the liquid temperature of
hydrogen production by electrolysis of water, Int. J. Hydrogen Energy. 42 (2017)
323.15 and 333.15 K. Consequently, the maximal production of 2–9, https://doi.org/10.1016/j.ijhydene.2017.07.015.
hydrogen and methanol conversion are obtained in a narrow range of [8] D. Das, T.N. Veziroglu, Advances in biological hydrogen production processes, Int.
liquid temperature from 293.15 to 303.15 K, which indicates the J. Hydrogen Energy. 33 (2008) 6046–6057, https://doi.org/10.1016/j.
ijhydene.2008.07.098.
sensitivity of acoustic cavitation to the change of the bulk liquid tem­ [9] M. Ni, M.K.H. Leung, D.Y.C. Leung, K. Sumathy, A review and recent developments
perature in the presence of methanol. in photocatalytic water-splitting using TiO2 for hydrogen production, Renew.
Sustain. Energy Rev. 11 (2007) 401–425, https://doi.org/10.1016/j.
rser.2005.01.009.
CRediT authorship contribution statement [10] L. Schlapbach, A. Züttel, Hydrogen-storage materials for mobile applications,
Nature 414 (2001) 353–358, https://doi.org/10.1038/35104634.
Aissa Dehane: Conceptualization, Methodology, Software, Acqui­ [11] T.N. Veziro⋗lu, F. Barbir, Hydrogen: the wonder fuel, Int. J. Hydrogen Energy. 17
(1992) 391–404, https://doi.org/10.1016/0360-3199(92)90183-W.
sition of data, Formal analysis, Writing - original draft, Writing - review [12] S. Merouani, O. Hamdaoui, The sonochemical approach for hydrogen production,
& editing. Slimane Merouani: Project administration, Conceptualiza­ in: A.A. Inamuddin (Ed.), The sonochemical approach for hydrogen production,
tion, Supervision, Visualization, Writing - review & editing Methodol­ Sustain. Green Chem. Process. Their Allied Appl. Nanotechnol. Life Sci. (2020)
1–29, https://doi.org/10.1007/978-3-030-42284-4_1.
ogy, Formal analysis, Writing-review & editing. Atef Chibani:
[13] H. Harada, Sonochemical reduction of carbon dioxide, Ultrason. Sonochem. 5
Acquisition of data, Revision. Oualid Hamdaoui: Visualization, Vali­ (1998) 73–77, https://doi.org/10.1016/S1350-4177(98)00015-7.
dation, Writing - review & editing. Muthupandian Ashokkumar: [14] H. Islam, O.S. Burheim, B.G. Pollet, Sonochemical and Sonoelectrochemical
Production of Hydrogen, Ultrason. Sonochem. 51 (2019) 533–555, https://doi.
Visualization, Formal analysis, Writing - review & editing.
org/10.1016/j.ultsonch.2018.08.024.
[15] S.S. Rashwan, I. Dincer, A. Mohany, An investigation of ultrasonic based hydrogen
Funding production, Energy 205 (2020), 118006, https://doi.org/10.1016/j.
energy.2020.118006.
[16] S.S. Rashwan, I. Dincer, A. Mohany, A unique study on the effect of dissolved gases
This work received financial support from The Ministry of Higher and bubble temperatures on the ultrasonic hydrogen (sonohydrogen) production,
Education and Scientific Research of Algeria (project code: Int. J. Hydrogen Energy. 45 (2020) 20808–20819, https://doi.org/10.1016/j.
A16N01UN250320220002) and the General Directorate of Scientific ijhydene.2020.05.022.
[17] K. Kerboua, O. Hamdaoui, Energetic challenges and sonochemistry: a new
Research and Technological Development (GD-SRTD). alternative for hydrogen production? Curr. Opin. Green Sustain. Chem. 18 (2019)
84–89, https://doi.org/10.1016/j.cogsc.2019.03.005.

9
A. Dehane et al. Chemical Engineering and Processing - Process Intensification 179 (2022) 109080

[18] K. Kerboua, O. Hamdaoui, Sonochemical production of hydrogen: enhancement by [45] J. Lee, M. Ashokkumar, S. Kentish, F. Grieser, Determination of the size
summed harmonics, Chem. Phys. 519 (2019) 27–37, https://doi.org/10.1016/j. distribution of sonoluminescence bubbles in a pulsed acoustic field, J. Am. Chem.
chemphys.2018.11.019. Soc. 127 (2005) 16810–16811, https://doi.org/10.1021/ja0566432.
[19] K. Kerboua, O. Hamdaoui, Oxygen-argon acoustic cavitation bubble in a water- [46] S. Merouani, O. Hamdaoui, N. Kerabchi, Liquid compressibility effect on the
methanol mixture: effects of medium composition on sonochemical activity, acoustic generation of free radicals, J. Appl. Water Eng. Res. 0 (2020) 1–16,
Ultrason. Sonochem. 61 (2020), 104811, https://doi.org/10.1016/j. https://doi.org/10.1080/23249676.2020.1787245.
ultsonch.2019.104811. [47] S. Merouani, O. Hamdaoui, Y. Rezgui, M. Guemini, Sensitivity of free radicals
[20] K. Kerboua, O. Hamdaoui, Numerical estimation of ultrasonic production of production in acoustically driven bubble to the ultrasonic frequency and nature of
hydrogen: effect of ideal and real gas based models, Ultrason. Sonochem. 40 (2018) dissolved gases, Ultrason. Sonochem. 22 (2014) 41–50, https://doi.org/10.1016/j.
194–200, https://doi.org/10.1016/j.ultsonch.2017.07.005. ultsonch.2014.07.011.
[21] S. Merouani, O. Hamdaoui, Y. Rezgui, M. Guemini, Computational engineering [48] K. Yasui, T. Tuziuti, J. Lee, T. Kozuka, A. Towata, Y. Iida, The range of ambient
study of hydrogen production via ultrasonic cavitation in water, Int. J. Hydrogen radius for an active bubble in sonoluminescence and sonochemical reactions,
Energy. 41 (2016) 832–844, https://doi.org/10.1016/j.ijhydene.2015.11.058. J. Chem. Phys. 128 (2008), 184705, https://doi.org/10.1063/1.2919119.
[22] S. Merouani, O. Hamdaoui, The size of active bubbles for the production of [49] S. Merouani, O. Hamdaoui, Y. Rezgui, M. Guemini, Effects of ultrasound frequency
hydrogen in sonochemical reaction field, Ultrason. Sonochem. 32 (2016) 320–327, and acoustic amplitude on the size of sonochemically active bubbles-Theoretical
https://doi.org/10.1016/j.ultsonch.2016.03.026. study, Ultrason. Sonochem. 20 (2013) 815–819, https://doi.org/10.1016/j.
[23] E.J. Hart, C.H. Fischer, A. Henglein, Sonolysis of hydrocarbons in aqueous solution, ultsonch.2012.10.015.
Int. J. Radiat. Appl. Instrumentation. Part C. Radiat. Phys. Chem. 36 (1990) [50] A. Dehane, S. Merouani, O. Hamdaoui, A. Alghyamah, Insight into the impact of
511–516, https://doi.org/10.1016/1359-0197(90)90198-Q. excluding mass transport, heat exchange and chemical reactions heat on the
[24] D.N. Rassokhin, G.V. Kovalev, L.T. Bugaenko, Temperature effect on the sonolysis sonochemical bubble yield: Bubble size-dependency, Ultrason. Sonochem. 73
of methanol/water mixtures, J. Am. Chem. Soc. 117 (1995) 344–347, https://doi. (2021), 105511, https://doi.org/10.1016/j.ultsonch.2021.105511.
org/10.1021/ja00106a037. [51] S. Merouani, O. Hamdaoui, Computer simulation of chemical reactions occurring
[25] C.M. Krishna, Y. Lion, T. Kondo, P. Riesz, Thermal decomposition of methanol in in collapsing acoustical bubble: dependence of free radicals production on
the sonolysis of methanol-water mixtures. Spin-trapping evidence for isotope operational conditions, Res. Chem. Intermed. (2015) 881–897, https://doi.org/
exchange reactions, J. Phys. Chem. 91 (1987) 5847–5850, https://doi.org/ 10.1007/s11164-013-1240-y.
10.1021/j100307a007. [52] J. Rae, M. Ashokkumar, O. Eulaerts, C. Von Sonntag, J. Reisse, F. Grieser,
[26] J. Buettner, M. Gutierrez, a. Henglein, Sonolysis of water-methanol mixtures, Estimation of ultrasound induced cavitation bubble temperatures in aqueous
J. Phys. Chem. 95 (1991) 1528–1530, https://doi.org/10.1021/j100157a004. solutions, Ultrason. Sonochem. 12 (2005) 325–329, https://doi.org/10.1016/j.
[27] M. Penconi, F. Rossi, F. Ortica, F. Elisei, P.L. Gentili, Hydrogen production from ultsonch.2004.06.007.
water by photolysis, sonolysis and sonophotolysis with solid solutions of rare earth, [53] E. Ciawi, J. Rae, M. Ashokkumar, F. Grieser, Determination of temperatures within
gallium and indium oxides as heterogeneous catalysts, Sustain 7 (2015) acoustically generated bubbles in aqueous solutions at different ultrasound
9310–9325, https://doi.org/10.3390/su7079310. frequencies, J. Phys. Chem. B. 110 (2006) 13656–13660, https://doi.org/10.1021/
[28] M.H. Islam, J.J. Lamb, K.M. Lien, O.S. Burheim, J.-Y. Hihn, B.G. Pollet, Novel fuel jp061441t.
production based on sonochemistry and sonoelectrochemistry, ECS Trans. 92 [54] M.P. Kanthale, M. Ashokkumar, F. Grieser, Estimation of Cavitation Bubble
(2019) 1–16, https://doi.org/10.1149/09210.0001ecst. Temperatures in an Ionic Liquid, J. Phys. Chem. C. 111 (2007) 18461–18463,
[29] K. Kerboua, O. Hamdaoui, S. Al-Zahrani, Sonochemical production of hydrogen: a https://doi.org/10.1021/jp710148k.
numerical model applied to the recovery of aqueous methanol waste under [55] A. Dehane, S. Merouani, O. Hamdaoui, A. Alghyamah, A complete analysis of the
Oxygen-Argon atmosphere, Environ. Prog. Sustain. Energy. 40 (2021), e13511, effects of transfer phenomenons and reaction heats on sono-hydrogen production
https://doi.org/10.1002/ep.13511. from reacting bubbles: Impact of ambient bubble size, Int. J. Hydrogen Energy. 46
[30] K. Kerboua, S. Merouani, O. Hamdaoui, A. Alghyamah, M.H. Islam, H.E. Hansen, B. (2021) 18767–18779, https://doi.org/10.1016/j.ijhydene.2021.03.069.
G. Pollet, How do dissolved gases affect the sonochemical process of hydrogen [56] P. Kanthale, M. Ashokkumar, F. Grieser, Sonoluminescence, sonochemistry (H2O2
production? An overview of thermodynamic and mechanistic effects – On the “hot yield) and bubble dynamics: Frequency and power effects, Ultrason. Sonochem. 15
spot theory, Ultrason. Sonochem. 72 (2021), 105422, https://doi.org/10.1016/j. (2008) 143–150, https://doi.org/10.1016/j.ultsonch.2007.03.003.
ultsonch.2020.105422. [57] K. Yasui, T. Tuziuti, T. Kozuka, A. Towata, Y. Iida, Relationship between the bubble
[31] A. Dehane, S. Merouani, O. Hamdaoui, Methanol sono-pyrolysis for hydrogen temperature and main oxidant created inside an air bubble under ultrasound,
recovery: effect of methanol concentration under an argon atmosphere, Chem. Eng. J. Chem. Phys. 127 (2007), 154502, https://doi.org/10.1063/1.2790420.
J. 433 (2021), 133272, https://doi.org/10.1016/j.cej.2021.133272. [58] S. Gireesan, A.B. Pandit, Modeling the effect of carbon-dioxide gas on cavitation,
[32] K. Yasui, Effects of thermal conduction on bubble dynamics near the Ultrason. Sonochem. 34 (2017) 721–728, https://doi.org/10.1016/j.
sonoluminescence threshold, J. Acoust. Soc. Am. 98 (1995) 2772–2782, https:// ultsonch.2016.07.005.
doi.org/10.1121/1.413242. [59] S. Merouani, H. Ferkous, O. Hamdaoui, Y. Rezgui, M. Guemini, A method for
[33] S. Sochard, A.M. Wilhelm, H. Delmas, Modelling of free radicals production in a predicting the number of active bubbles in sonochemical reactors, Ultrason.
collapsing gas-vapour bubble, Ultrason. Sonochem. 4 (1997) 77–84, https://doi. Sonochem. 22 (2014) 51–58, https://doi.org/10.1016/j.ultsonch.2014.07.015.
org/10.1016/S1350-4177(97)00021-7. [60] K. Kerboua, O. Hamdaoui, Void fraction, number density of acoustic cavitation
[34] R. Toegel, D. Lohse, Phase diagrams for sonoluminescing bubbles: a comparison bubbles, and acoustic frequency: a numerical investigation, J. Acoust. Soc. Am. 146
between experiment and theory, J. Chem. Phys. 118 (2003) 1863–1875, https:// (2019) 2240, https://doi.org/10.1121/1.5126865.
doi.org/10.1063/1.1531610. [61] B. Avvaru, A.B. Pandit, Oscillating bubble concentration and its size distribution
[35] A. Dehane, S. Merouani, O. Hamdaoui, Carbon tetrachloride (CCl4) sonochemistry: using acoustic emission spectra, Ultrason. Sonochem. 16 (2009) 105–115, https://
A comprehensive mechanistic and kinetics analysis elucidating how CCl4 pyrolysis doi.org/10.1016/j.ultsonch.2008.07.003.
improves the sonolytic degradation of nonvolatile organic contaminants, Sep. [62] A. Dehane, S. Merouani, O. Hamdaoui, Effect of carbon tetrachloride (CCl4)
Purif. Technol. 275 (2021), 118614, https://doi.org/10.1016/j. sonochemistry on the size of active bubbles for the production of reactive oxygen
seppur.2021.118614. and chlorine species in acoustic cavitation field, Chem. Eng. J. 426 (2021),
[36] K. Yasui, T. Tuziuti, W. Kanematsu, Extreme conditions in a dissolving air 130251, https://doi.org/10.1016/j.cej.2021.130251.
nanobubble, Phys. Rev. E. 94 (2016) 013106–013113, https://doi.org/10.1103/ [63] H. Ferkous, S. Merouani, O. Hamdaoui, Y. Rezgui, M. Guemini, Comprehensive
PhysRevE.94.013106. experimental and numerical investigations of the effect of frequency and acoustic
[37] Y.S. Touloukian, P.E. Liley, S.C. Saxena, Thermal Conductivity: Nonmetallic intensity on the sonolytic degradation of naphthol blue black in water, Ultrason.
Liquids and Gases, IFI/Plenum, 1990. Sonochem. 26 (2015) 30–39, https://doi.org/10.1016/j.ultsonch.2015.02.004.
[38] K. Yasui, Effect of liquid temperature on sonoluminescence, Phys. Rev. E. 64 [64] K. Yasui, T. Tuziuti, Y. Iida, H. Mitome, Theoretical study of the ambient-pressure
(2001), 016310, https://doi.org/10.1103/PhysRevE.64.016310. dependence of sonochemical reactions, J. Chem. Phys. 119 (2003) 346, https://
[39] S. Merouani, O. Hamdaoui, Y. Rezgui, M. Guemini, Mechanism of the doi.org/10.1063/1.1576375.
sonochemical production of hydrogen, Int. J. Hydrogen Energy. 40 (2015) [65] A. Alippi, F. Cataldo, A. Galbato, Ultrasound cavitation in sonochemistry:
4056–4064, https://doi.org/10.1016/j.ijhydene.2015.01.150. decomposition of carbon tetrachloride in aqueous solutions of potassium iodide,
[40] K. Yasui, Chemical reactions in a sonoluminescing bubble, J. Phys. Soc. Japan. 66 Ultrasonics 30 (1992) 148–151, https://doi.org/10.1016/0041-624X(92)90064-S.
(1997) 2911–2920, https://doi.org/10.1143/JPSJ.66.2911. [66] A. Weissler, W. cooper Herbert, S. Snyder, Chemical effect of ultrasonic waves:
[41] R. Seiser, K. Seshadri, F.A. Williams, Detailed and reduced chemistry for methanol oxidation of potassium iodide solution by carbon tetrachloride, J. Am. Chem. Soc.
ignition, Combust. Flame. 158 (2011) 1667–1672, https://doi.org/10.1016/j. 72 (1950) 1769–1775, https://doi.org/10.1021/ja01160a102.
combustflame.2011.02.008. [67] M.A. Beckett, I. Hua, Elucidation of the 1,4-dioxane decomposition pathway at
[42] T.S. Norton, F.L. Dryer, Toward a comprehensive mechanism for methanol discrete ultrasonic frequencies, Environ. Sci. Technol. 34 (2000) 3944–3953,
pyrolysis, Int. J. Chem. Kinet. 22 (1990) 219–241, https://doi.org/10.1002/ https://doi.org/10.1021/es000928r.
kin.550220303. [68] M.A. Beckett, I. Hua, Impact of ultrasonic frequency on aqueous sonoluminescence
[43] D.W. Green, R.H. Perry, Perry’s chemical engineers’ handbook, McGraw-Hill, New and sonochemistry, J. Phys. Chem. A. 105 (2001) 3796–3802, https://doi.org/
York, USA, 2008. 10.1021/jp003226x.
[44] A. Brotchie, F. Grieser, M. Ashokkumar, Effect of power and frequency on bubble- [69] J.-W. Kang, H.-M. Hung, A. Lin, M.R. Hoffmann, Sonolytic Destruction of Methyl
size distributions in acoustic cavitation, Phys. Rev. Lett. 102 (2009), 084302, tert -Butyl Ether by Ultrasonic, Environ. Sci. Technol. 33 (1999) 3199–3205.
https://doi.org/10.1103/PhysRevLett.102.084302, 1–4. [70] R.A. Torres, C. Pétrier, E. Combet, M. Carrier, C. Pulgarin, Ultrasonic cavitation
applied to the treatment of bisphenol A. Effect of sonochemical parameters and

10
A. Dehane et al. Chemical Engineering and Processing - Process Intensification 179 (2022) 109080

analysis of BPA by-products, Ultrason. Sonochem. 15 (2008) 605–611, https://doi. [72] O. Authier, H. Ouhabaz, S. Bedogni, Modeling of sonochemistry in water in the
org/10.1016/j.ultsonch.2007.07.003. presence of dissolved carbon dioxide, Ultrason. Sonochem. 45 (2018) 17–28,
[71] S. Merouani, O. Hamdaoui, Toward understanding the mechanism of pure CO2 https://doi.org/10.1016/j.ultsonch.2018.02.044.
-quenching sonochemical processes, J. Chem. Technol. Biotechnol. 95 (2019) [73] A. Henglein, Sonolysis of carbon dioxide, nitrous oxide and methane in aqueous
553–566, https://doi.org/10.1002/jctb.6227. solution, Zeitschrift F{ü}r Naturforsch. B. 40 (1985) 100–107, https://doi.org/
10.1515/znb-1985-0119.

11

You might also like