Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Chemical Engineering Journal 433 (2022) 133272

Contents lists available at ScienceDirect

Chemical Engineering Journal


journal homepage: www.elsevier.com/locate/cej

Methanol sono-pyrolysis for hydrogen recovery: Effect of methanol


concentration under an argon atmosphere
Aissa Dehane a, Slimane Merouani b, *, Oualid Hamdaoui c
a
Laboratory of Environmental Engineering, Department of Process Engineering, Faculty of Engineering, Badji Mokhtar – Annaba University, P.O. Box 12, 23000 Annaba,
Algeria
b
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
c
Chemical Engineering Department, College of Engineering, King Saud University, P.O. Box 800, 11421 Riyadh, Saudi Arabia

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

Keywords: A number of theoretical and experimental works showed the possibility of increasing the sonochemical pro­
Methanol sonolysis duction of hydrogen through the pyrolysis of methanol within the acoustic cavitation (i.e. sonolysis of aqueous
Hydrogen production methanol solution). In this study, numerical simulations have been conducted in order to reveal, for the first
Active bubbles range
time, the effect of methanol concentration (in the bulk liquid) on the maximal sonochemical efficiency for
Methanol conversion
hydrogen yielding, methanol conversion and the broadness of active bubbles range. The current adopted model is
Sonochemical efficiency
built on a set of ordinary differential equations that account for non-equilibrium evaporation and condensation of
water vapor and methanol at the bubble wall, thermal conduction both within and outside the bubble and heat of
chemical reactions. All numerical simulations were carried out for a single bubble oscillating in O2 and/or argon-
saturated water with varying concentration of methanol. It was found that the variation of methanol concen­
tration (0–100% (v/v)) affects slightly the broadness of active bubbles ranges for methanol consumption.
Conversely, for hydrogen production, a gradual decrease of the active bubbles ranges is observed for methanol
concentration greater than 20%. The maximal sonochemical efficiency for hydrogen yielding, methanol con­
version and the broadness of active bubbles ranges (for H2 production and CH3OH conversion) is obtained at
80% of argon and for methanol concentration between 7% and 20%. The yield of hydrogen increased by 26.16
and 22.4 times at 10 and 20% (v/v) of methanol in bulk solution, respectively, whereas CH3OH conversion
decreased by 13.84 and 22.3%, respectively.

spot theory considers the bubble at the last stage of collapse as a


microreactor within which high energy combustion reactions and
1. Introduction sonoluminescence occur in very short time (scale of nanoseconds). After
final collapse, the bubble content is diffused into the thin layer imme­
Sonochemistry is considered as one of the cleaner ways for hydrogen diately surrounding the collapsing bubble, where oxidants can drive
production [1], particularly if the sonicated solution (water principally) other reaction chain in the liquid phase [6].
contains alcohols at low concentrations [2]. The ultrasonic (US) irradi­ Methanol is an important chemical product. It is mostly utilized in
ation of water, provoke the generation of strong oxidizing agents (i.e. the manufacture of other compounds such as formaldehyde, acetic acid,

OH, HO2● and O) and molecular species (i.e. H2 and H2O2) through the methyl and vinyl acetates, methyl methacrylate, methylamines, methyl
acoustic cavitation, which is the central event of this technique [3]. tert-butyl ether (MTBE), fuel additives, and other chemicals. Every year,
After formation, acoustic cavities can grow and collapse violently. approximately 98 million tonnes are produced, almost entirely from
During the fast compression of the entrapped vapor and gases, hot fossil fuels (either natural gas or coal) [7,8]. The use of methanol in
extreme conditions of temperature and pressure (~5000 K and ~ 500 sonochemistry field has found a place through several experimental and
bar [4]) are formed, leading to the decomposition of bubble content (i.e., theoretical works. In the early days of sonochemistry, Henglein and
water vapor, gases and entrapped volatile solutes) and the generation of Schulz [9] have indicated a decrease of iodide oxidation yield with the
different substances (radicals and molecules) [5]. The sonochemical hot

* Corresponding author.
E-mail address: s.merouani@yahoo.fr (S. Merouani).

https://doi.org/10.1016/j.cej.2021.133272
Received 14 August 2021; Received in revised form 30 September 2021; Accepted 27 October 2021
Available online 6 November 2021
1385-8947/© 2021 Elsevier B.V. All rights reserved.
A. Dehane et al. Chemical Engineering Journal 433 (2022) 133272

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

Uk Production rate of the kth species (mol s− 1
m− 3)
T Temperature inside a bubble, (K).

increase of methanol amount added to sonicated potassium iodide have observed the low consumption of DPPH (1,1-diphenyl-2-pic­
aerated-solutions at 500 kHz. This was ascribed to the high vapor rylhydrazyl) compared to the formation of hydrogen, which indicated
pressure of methanol, which causes the suppress of bubble sonoactivity. the scavenging of hydroxyl radicals by methanol molecules inside
Similarly, Son and Zoh [10] have found that the increase of methanol bubbles. In addition, Okitsu et al. [14] and Mizukoshi et al. [15] have
quantity into the bulk liquid causes the formation of hydrogen peroxide attributed the different degradation rates of DPPH to the evaporation
and the oxidation rate of 1,4-dioxane to be reduced. The decrease of rates of the organic solutes rather than their vapor pressure. Moreover,
H2O2 generation at low temperatures was attributed to the scavenging the effect of methanol (in addition to other organic species) on bubbles
effect of methanol towards hydroxyl radicals at the bubble-solution coalescence and its relation with sonoluminescence has been extensively
interface, whereas, the increase of liquid temperature transfer this analyzed in several works [16–18]. According to the experimental works
mechanism into the acoustic cavitation. Additionally, Son and Zoh [10] of Rassokhin et al. [12] and Buttner et al. [11] an important amount of
indicated that the scavenging effect of ●OH radicals is time dependent hydrogen can be produced in aqueous solution compared to that of pure
(accelerated at the beginning of sonolysis), where, they observed that water.
the degradation of 1,4-dioxane was accelerated after some time Recently, numerous works have been concentrated on the sono­
(depending on CH3OH concentration) in the presence of methanol. chemical production of hydrogen [2,19–22]. Promising results have
Buttner et al. [11] showed that typical pyrolysis and combustion prod­ been achieved through these studies for an efficient and clean hydrogen
ucts are obtained via the sonolysis of a methanol–water mixture under production. Merouani et al. [2] and Kerboua et al. [23] have lately
an argon and oxygen atmospheres, respectively, where the composition presented interesting reviews on the sonochemical production of
of the sonolysis products changes proportionally with the concentration hydrogen. Kerboua et al. [24] have also investigated theoretically the
of methanol in the bulk liquid. Additionally, Buttner et al. [11] observed effect of saturating gases (Ar and O2) on the sonolysis of methanol,
clearly the increase of the chemical production in 10% (v/v) methanol where a maximal molar yield was retrieved for a solution containing
solution compared to the case of pure water, whereas the sonoactivity is 40% of methanol in spite of the amount of argon. Additionally, Kerboua
completely suppressed for the solutions containing more than 80% of et al. [24] showed that the highest molar production is obtained through
methanol. Similarly, under an argon atmosphere, Rassokhin et al. [12] a concentration of 40% (v/v) for methanol and molar fraction of 40% for
demonstrated the increased formation of the sonolysis products (e.g. argon. In another theoretical study, Kerboua et al. [21] found that the
hydrogen, methane and carbon monoxide) of methanol proportionally best combination for hydrogen production and methanol conversion is
with the increase of its concentration in the solutions (diluted solutions, obtained for a solution containing 20% (v/v) of methanol saturated with
0.001–0.5 M). However, it is found that the yield of hydrogen peroxide 70% molar of argon, where 99.9% of methyl alcohol is eliminated, and
goes down with the increase of methanol concentration due to the 59% of methanol is converted to hydrogen.
thermal dissociation of methanol, which promotes the formation of CO, In general, the theoretical studies performed for methanol sono­
whereas in very diluted solution of methanol, the reaction of hydroxyl chemistry adopt a single ambient bubble radius (mean radius) for rep­
radicals with methanol favors the formation of H2O2 and CO. Krishna et resenting the bubbles population. However, in reality this is not the true
al. [13] have indicated the formation of CH3● (in a relatively important situation, because, bubbles population in acoustic field is an interval
quantity) and CH2OH● radicals. The generation of methyl radicals was rather than a single value; the statement which is well proven experi­
explained by the pyrolysis of methanol in the hot spot. However, the mentally through several technics [25–30] and theoretically through
formation of CH2OH● radicals was attributed either to the pyrolysis of different models [25–33]. Besides, because of the dual role of methanol
methanol or to the H abstraction through the reaction of methanol with within the bubble (i.e. generator of methyl radicals and quencher of
H●, ●OH and CH3● radicals. Okitsu et al. [14] and Mizukoshi et al. [15] hydroxyl radicals), the active bubbles population in the sonochemical

2
A. Dehane et al. Chemical Engineering Journal 433 (2022) 133272

reactor may be greatly modified due to the change of the sono-activity of unit area and unit time and P∞ is the static pressure (1 atm). PA is
individual cavities. In another word, methanol sonochemistry can af­ connected to the acoustic intensity Ia by: PA= (2Ia ρL C)1/2. The liquid
fects one of the most significant features of the cavitation field, which is pressure PB(t) exerted on the external side of the bubble wall is related to
the size of active bubbles. This problem was never addressed previously, the internal bubble pressure P(t) as [42]:
neither experimentally nor theoretically. The determination of the
2σ 4μṘ
population of active bubbles has a great importance, because if this PB (t) = P(t) − − (2)
range is properly determined, it will be facile to study the bubbles effects R R
and their interactions with the acoustic irradiation [34]. In fact, the σ and µ are is the surface tension and viscosity of the liquid,
bubble temperature, and so the bubble sono-activity, are strongly respectively. The van der Waals equation (Eq. (3)) is used for calculating
affected by the ambient bubble size (i.e., R0) [32,33,35]. the internal bubble pressure P(t) :
Therefore, based on the chemical mechanisms adopted by Kerboua
nRg T an2
et al. [21,24] for methanol pyrolysis and combustion inside acoustic P(t) = + (3)
V − nb V 2
cavitation, the present study attempts to reveal, for the first time, the
effect of methanol concentration [% (v/v)] on the range of active bub­ a and b are the Van de Waals constants (calculated as functions of O2,
bles for hydrogen production and methyl alcohol conversion. However, Ar, H2O and CH3OH as in [43,44]), T is the temperature inside the
in our work, some important chemical reactions such as water dissoci­ bubble, Rg is the universal ideal gas constant and V is the bubble volume
ation and methane formation (i.e. ignored in the Kerboua et al.’s [4/3(πR3)].
mechanism) were added for an accuracy approach. The mass flux of evaporation and condensation of H2O and CH3OH at
Model the interface is expressed using Hertz-Knudsen formula derived from
The mathematic model we used combined in one part the bubble kinetic theory of gases [41]:
dynamics and the energy balance equations developed by Yasui [36–38]
{Psat i [R] − Pvi }
and, in another part, the Toegel model [39,40] for heat transfer, as well ṁi = a √̅̅̅̅̅̅̅̅̅̅̅̅ (i = H2 O, CH3 OH) (4)
2π T[R] Rg
as the Hertz-Knudsen model for mass transport [41]. Additionally, an Mi
improved chemical kinetics mechanism for methanol pyrolysis within
Psat_i [R] is the saturation pressure of water or methanol (calculated
the bubble is adopted, where many essential chemical reactions for
by using Antoine’s equation as in [45,46]) at the interface temperature
water sonolysis and methane formation were added to the Kerboua
T[R] = Tliq, PV is the vapor pressure (H2O or CH3OH) within the bubble.
et al.’s [21,24] reaction scheme. These combinations provide an excel­
Mi is the molecular weight of H2O or CH3OH and α is the accommoda­
lent prediction of the experimental bubble temperature and the yields of
tion coefficient (for both H2O and CH3OH), given as [21]:
H2, CH2O and CO over a wide range of methanol concentration, as it will
be seen in Section 3.1. In addition, our model fitted perfectly the Yasui’s



⎪ a = 0.35, for T ≤ 350K

⎪ ( ) ( )( )( )
⎨ T T T T
a = 0.35 − 0.05 − 7 2 + 0.025 − 7 − 8 − 9 , for 350 ≤ T ≤ 500K (5)

⎪ 50 50 50 50



a = 0.05, for T ≥ 500K

bubble dynamics at 140 kHz (please see Fig. S1 of the Supplementary Heat transfer through the bubble wall is calculated as [40]:
Material). The model’s detail is presented below. ( )
Tliq − T
The model is built on a set of ordinary differential equations that Q̇ = 4π2 Rλmix (6)
account for non-equilibrium evaporation and condensation of water ξ
vapor and methanol at the bubble wall, thermal conduction (i.e. heat { √̅̅̅̅̅̅ }
transfer) both within and outside a bubble and chemical reactions (i.e. R Rχ
ξ = min , (7)
heats of reactions are incorporated). Interactions between bubbles are π Ṙ
ignored. All numerical simulations were carried out for a single bubble λmix, χ and Lth are the heat conductivity, thermal diffusivity of the gas
oscillating in O2 and/or argon-saturated water with varying amounts of mixture and the thickness of the thermal boundary layer respectively.
methanol. Bubble temperature, bubble pressure, bubble radius and λmix is estimated by [47]:
bubble wall velocity can be calculated at any time during the bubble ( )

oscillation. The radial dynamics of the bubble is described by the λmix = λi (T)
ni
(8)
following form of the Keller-Miksis equation [36]: nt

( ) ( ) ( )[ ( ( )) ] ( ) ( )
Ṙ ṁ 3 2 Ṙ 2ṁ 1 Ṙ R m̈R Ṙ ṁ ṁ ṁ Ṙṁ R dPB
1− + RR̈ + Ṙ 1 − + = 1+ PB (t) − PA sin 2πf t + − P∞ + 1− + + Ṙ + + +
C C ρL 2 3C 3CρL ρL C C ρL C C ρL ρL 2ρL 2CρL CρL dt
(1)

where t is time, dots (●) signifies time derivative (d/dt), R is the radius of Where λi(T) is the thermal conductivity of water vapor, methanol
the bubble, ρL is the density of the liquid, C is the speed of sound in the and argon and/or oxygen at temperature T (in K), respectively. λi (in W/
liquid, ṁ is the net rate of evaporation of water vapor and methanol per m K) of H2O, Ar, O2 and CH3OH are estimated by [42,48,49]:

3
A. Dehane et al. Chemical Engineering Journal 433 (2022) 133272

λH2O (T) = 9.967213 × 10− 5 T − 1.1705 × 10− 2


(9) The variation of the internal bubble energy (ΔE) in time (Δt) as a
result of the pressure work, heat of chemical reactions within the bub­
λAr (T) = 3.5887 × 10− 5 T + 6.81277 × 10− 3
(10) ble, energy transported by water and methanol molecules by evapora­
tion and condensation and heat exchange by thermal conduction at the
λCH3OH (T) = 1.02 × 10− 4 T − 0.0161 (11) bubble interface, is given by [38]:

λO2 (T) = 6.4478 × 10− 5 T + 7.2211 × 10− 3


(12)

[ ] ( )
ṁ ṁ Tliq − T 4 ∑n
ΔE = − P(t)ΔV(t) + 4πR2 Δt e H2 O + eCH3 OH + 4πR2 Δtλ − πR3 Δt ΔHi ri (14)
MH2 O MCH3 OH ξ 3 i=1

The temperature in the bubble is calculated as [50]: In Eq. (14), ΔHi and ri are the enthalpy change and the rate of the ith
( ) reaction, respectively, and eH2O and eCH3OH are the energy transported
by 1 mol of an evaporating or condensing water vapor and methanol,
2
E + anV
T= (13) which are expressed by:
Cv nt
ei = Cv,i T (i = H2 O, CH3 OH) (15)
where E is the internal energy of the bubble, nt is the total number of
moles inside the bubble, V is the bubble volume and Cv is the molar heat
1.1. Internal bubble chemistry
of the gases and vapor within the bubble [Cv = (3/2)Rg for monoatomic
gases (Ar, He…), (5/2)Rg for diatomic gases (H2,O2…) and (6/2)Rg for
The reaction scheme given in Table 1 is considered in the absence of
the other gases].
methanol. It is composed of 38 reversible chemical reactions involving

Table 1
Scheme of the possible chemical reactions inside a collapsing Ar-O2-bubble [70,71]. M is the third Body. A is in (m3 mol− 1 − 1
s ) for two body reaction [(m6 mol− 2 − 1
s )
for a three body reaction], Ea is in (kJ mol− 1) and ΔH is in (kJ mol− 1).
N◦ Reaction A n Ea ΔH

1 H2O + M → H•+•OH + M 1.912 × 107 − 1.83 28.35 508.82


2 O2 + M → O + O + M 4.515 × 1011 − 0.64 28.44 505.4
3 •
OH + M → O + H•+M 9.88 × 1011 − 0.74 24.43 436.23
4 H•+O2 → O+•OH 1.915 × 108 0.0 3.93 69.17
5 H•+O2 + M → HO2• +M 1.475 0.6 0.0 − 204.80
6 O + H2O → •OH+•OH 2.97 2.02 3.21 72.59
7 HO2•+H• → H2 + O2 1.66 × 107 0.0 1.97 × 10-1 − 239.67
8 HO2•+H• →•OH+•OH 7.079 × 107 0.0 7.06 × 10-2 − 162.26
9 HO2•+O → •OH + O2 3.25 × 107 0.0 0.0 − 231.85
10 HO2•+•OH → H2O + O2 2.89 × 107 0.0 − 1.19 × 10-1 − 304.44
11 H2 + M → H•+H•+M 4.577 × 1013 − 1.4 24.98 444.47
12 O + H2 → H•+•OH 3.82 × 106 0.0 1.9 8.23
13 •
OH + H2 → H•+H2O 2.16 × 102 1.52 8.25 × 10-1 − 64.35
14 H2O2 + O2 → HO2•+HO2• 4.634 × 1010 − 0.35 12.12 175.35
15 H2O2 + M → •OH+•OH + M 2.951 × 108 0.0 11.59 217.89
16 H2O2 + H• → H2O+•OH 2.410 × 107 0.0 9.5 × 10-1 − 290.93
17 H2O2 + H• → H2 + HO2• 6.025 × 107 0.0 1.9 − 64.32
18 H2O2 + O → •OH + HO2• 9.550 2.0 9.5 × 10-1 − 56.08
19 H2O2+•OH → H2O + HO2• 1.0 × 106 0.0 0.0 − 128.67
20 H•+•OH + M → H2O + M 2.2 × 1010 − 2.0 0.0 − 508.82
21 O + O + M → O2 + M 6.165 × 103 − 0.5 0.0 − 505.4
22 O + H•+M → •OH + M 4.714 × 106 − 1.0 0.0 − 436.23
23 O+•OH → H•+O2 5.481 × 105 0.39 − 7.01 × 10-2 − 69.17
24 HO2• +M → H•+O2 + M 3.09 × 106 0.53 11.7 204.80
25 •
OH+•OH → O + H2O 1.465 × 10-1 2.11 − 6.94 × 10-1 − 72.59
26 H2 + O2 → HO2•+H• 3.164 × 106 0.35 13.3 239.67
27 •
OH+•OH → HO2•+H• 2.027 × 104 0.72 8.8 162.26
28 •
OH + O2 → HO2•+O 3.252 × 106 0.33 12.75 231.85
29 H2O + O2 → HO2•+•OH 5.861 × 107 0.24 16.53 304.44
30 H•+H•+M → H2 + M 1.146 × 108 − 1.68 1.96 × 10-1 − 444.47
31 H•+•OH → O + H2 2.667 × 10-2 2.65 1.17 − 8.23
32 H•+H2O → •OH + H2 2.298 × 103 1.40 4.38 64.35
33 HO2•+HO2• → H2O2 + O2 4.2 × 108 0.0 2.87 − 175.35
34 •
OH+•OH + M → H2O2 + M 1.0 × 102 − 0.37 0.0 − 217.89
35 H2O+•OH → H2O2 + H• 1.269 × 102 1.31 17.08 290.93
36 H2 + HO2• → H2O2 + H• 1.041 × 105 0.70 5.74 64.32
37 •
OH + HO2• → H2O2 + O 8.66 × 10-3 2.68 4.45 56.08
38 H2O + HO2• → H2O2+•OH 1.838 × 104 0.59 7.4 128.67

Third body efficiency factors: R1: αH2 = 0.73, αH2O = 12, αAr = 0.38; R2: αH2 = 2.5, αH2O = 12, αAr = 0.75; R3: αH2 = 2.5, αH2O = 12, αAr = 0.83; R5: αH2 = 1.3, αH2O = 14,
αAr = 0.67; R11: αH2 = 2.5, αH2O = 12. R15: αH2 = 2.5, αH2O = 12, αAr = 0.64; R20: αH2 = 0.73, αH2O = 12, αAr = 0.38; R21: αH2 = 2.5, αH2O = 12, αAr = 0.75; R22: αH2 = 2.5,
αH2O = 12, αAr = 0.83; R24: αH2 = 1.3, αH2O = 14, αAr = 0.67; R30: αH2 = 2.5, αH2O = 12; R34: αH2 = 2.5, αH2O = 12, αAr = 0.64.

4
A. Dehane et al. Chemical Engineering Journal 433 (2022) 133272

Table 2
Scheme of the possible chemical reactions inside a collapsing Ar-CH3OH-bubble [24,51,52]. M is the third Body. A is in (m3 mol− 1
s− 1) for two body reaction [(m6
mol− 2 s− 1) for a three body reaction], Ea is in (kJ mol− 1) and ΔH is in (kJ mol− 1).
N◦ Reaction A n Ea ΔH

1 H2O + M ⇌ H•+•OH + M 1.912 × 107 − 1.83 28.35 508.82


2 CH3OH ⇌ CH3• + •OH 1.9 × 1010 0.0 3.84 × 105 374.114
3 CH3OH ⇌ CH2OH• + H• 1.54 × 1010 0.0 4.05 × 105 390.114
4 CH3OH + H• ⇌ CH2OH• + H2 1.35 × 10-3 3.2 1.46 × 104 − 34.057
5 CH3OH + •OH ⇌ CH3O•+ H2O 4.4 × 1000 2.0 6.3 × 103 − 54.215
6 CH3OH + •OH ⇌ CH2OH• + H2O 1.44 × 1000 2.0 − 3.51 × 103 − 95.215
7 CH3O•+ M ⇌ CH2OH• + M 1.00 × 1008 0.00 8.00 × 1004 − 38.000
8 •
OH + •OH ⇌ HO2• + H 1.08 × 1005 0.61 1.51 × 1005 154.957
9 HO2• + •OH ⇌ H2O + O2 2.89 × 1007 0.00 − 2.08 × 1003 − 291.451
10 HO2 •+ O ⇌ •OH + O2 2.00 × 1007 0.00 0.00 − 224.131
11 HO2 + H ⇌ H2 O + O 3.10 × 1007 0.00 7.20 × 1003 –22.957
12 HO2• + H• ⇌ H2 + O2 1.66 × 1007 0.00 3.44 × 1003 − 230.293
13 HO2• + H• ⇌ 2 •OH 7.08 × 1007 0.00 1.23 × 1003 − 155.613
14 CH2OH• + O2 ⇌ CH2O + HO2• 5.00 × 1006 0.00 0.00 − 79.404
15 CH3O•+ M ⇌ CH2O + H•+ M 7.78 × 1007 0.00 5.65 × 1004 126.297
16 CH3O•+ O2 ⇌ CH2O + HO2• 4.28 × 10− 19 7.60 − 1.48 × 1004 − 117.404
17 CH2O + HO2 •⇌ HCO• + H2O2 4.11 × 10− 02 2.50 4.27 × 1004 − 48.976
18 CH2O+ •OH ⇌ HCO• + H2O 3.90 × 1004 0.89 1.70 × 1003 − 128.159
19 CH2O + O ⇌ HCO• + •OH 3.50 × 1007 0.00 1.47 × 1004 − 60.839
20 CH2O + H• ⇌ HCO• + H2 5.74 × 1001 1.90 1.15 × 1004 − 67.001
21 HCO• + O2 ⇌ CO + HO2• 7.58 × 1006 0.00 1.72 × 1003 − 140.530
22 HCO• + M ⇌ CO + H•+ M 1.86 × 1011 − 1.00 7.11 × 1004 − 65.171
23 CO +• OH ⇌ CO + H•+ CO2 4.40 × 1000 1.50 − 3.10 × 1003 − 212.853
24 2 HO2 •⇌ H2O2 + O2 3.02 × 1006 0.00 5.80 × 1003 − 212.268
25 H2O2 + M ⇌ 2 •OH + M k0 8.15 × 1017 − 1.92 2.08 × 1005 262.356
k∞ 2.62 × 1013 − 1.39 2.15 × 1005
26 H•+ O2 + M ⇌ HO2• + M k0 5.75 × 1007 − 1.40 0.00 − 205.701
k∞ 4.65 × 1000 0.44 0.00
27 H2 + O ⇌ •OH + H • 5.06 × 10− 02 2.67 2.63 × 1004 6.162
28 H•+ O2 ⇌ •OH + O 3.52 × 1010 − 0.70 7.14 × 1004 68.518
29 CH2OH•+ H• ⇌ CH2O + H2 3 × 107 0.0 0.0 − 309.578
30 2 •OH ⇌ HO2• 1.08 × 105 0.61 1.51 × 105 154.957
31 CH2OH• + •OH ⇌ CH2O + H2O 2.4 × 107 0.0 0.0 − 370.7
32 •
CH3 + HO2• ⇌ CH4 + O2 3.61 × 106 0.0 0.0 –233.852

Third body efficiency factors: R1:αH2 = 2.5, αH2O = 16, αAr = 1, αCO = 1.9, αCO2 = 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, others 1. R22:αH2 = 1.9, αH2O = 12, αAr =, αCO = 2.5, αCO2 = 2.5, αCH4 =, others 1; R25:αH2 =
2, αH2O = 6, αAr = 0.4, αCO = 1.5, αCO2 = 2, αCH4 =, others 1; R26:αH2 = 2.5, αH2O = 16, αAr = 0.7, αCO = 1.2, αCO2 = 2.4, αCH4 =, others 1.

H2O, H•, O, •OH, HO2•, H2O2, H2 and O2 chemical species. For a bubble et al. [15]. According to the experimental study of Buttner et al. [11], the
initially containing methanol in addition to argon and water vapor, a typical pyrolysis (sonolysis) products of methanol detected under argon
chemical kinetics consisting in 32 chemical reactions (Table 2) is taken atmosphere are: H2, CH2O, CO, CH4 and traces of C2H4 and C2H6 (very
into account, involving H2O, H•, O, •OH, HO2•, H2O2, H2, O2, CH3OH, low yielding), whereas, under oxygen atmosphere, the typical products
CH3O•, CH3•, CH2OH•, CH2O, HCO•, CO, CO2 and CH4 molecules. For a of methanol combustion inside the bubbles are: CO2, CO, HCOOH,
bubble initially containing methanol in addition to argon, oxygen and CH2O, H2O2 and traces of H2. On the other hand, Mizukoshi et al. [15],
water vapor, a chemical kinetics consisting in 32 chemical reactions shows that in presence of argon as a saturating gas, the detected prod­
(Table 3) is taken into account, involving the same species as in Table 2. ucts of methanol pyrolysis are: H2 (main product), CO, CH4 and alde­
It should be noted that the chemical reactions involved in Tables 2 and 3 hydes. More interestingly, schemes of Table 2 and 3 are validated by the
are validated from static, flow and shock-tube reactors [51,52]. comparison of our results for the production of H2, CH2O and CO to
From a theoretical viewpoint many chemical reactions are expected those obtained experimentally by Buttner et al. [11] and theoretically by
through the pyrolysis and combustion of methanol. This has been Kerboua et al. [24] at 1 MHz (see Fig. 2). Additionally, the efficacy of our
demonstrated by various works [51–53]. However, in sonochemistry chemical mechanism was asserted through the confrontation of our
and inside the bubble (an effective micro-reactor), the space available bubble temperatures to those obtained experimentally by Rae et al. [54]
for chemical reactions is quite different (small and varying volume), in for a range of methanol concentration from 0 to 500 mM, where a good
addition the time available for chemical reactions is very short (a few accordance is observed between our results and the Rae et al.’s results
nanoseconds, as in Fig. 1). Moreover, other phenomena are expected (Fig. 3).
such as heat exchange, variation of reactor volume, segregation of vapor The chemical kinetics model by which the reaction schemes of Ta­
and gas content of bubble, light emission … [43]. Consequently, it is bles 1-3 are involved in the bubble dynamics is provided in detail in our
deduced that the acoustic cavitation is a different environment for previous work [55]. The water vapor and methanol change with time
chemical reactions; therefore, depending on the initial composition of inside the bubble, due to the evaporation and condensation at the bubble
bubble, the kinetical mechanism is very limited [43]. The present wall and due to chemical reactions is given by:
chemical mechanisms used in our study (Tables 2 and 3) are principally
dnH2 O ṁH O 4
inspired from the reduced mechanisms of Seiser et al. [51] and Norton = 4πR2 2 + πR3 U̇ H2 O (16)
dt MH2 O 3
and Dryer [52], which are in good agreement with the experimental
results validated by static, flow and shock tube reactors. It should be dnCH3 OH ṁCH3 OH 4 3
noted that methane formation (from CH3) is considered in Tables 2 and = 4πR2 + πR U̇ CH3 OH (17)
dt MCH3 OH 3
3. On the other hand, the mechanisms of Tables 2 and 3 are supported by
the experimental sonolytic results of Buttner et al. [11] and Mizukoshi For other species k (except Ar):

5
A. Dehane et al. Chemical Engineering Journal 433 (2022) 133272

Table 3
Scheme of the possible chemical reactions inside a collapsing Ar-O2-CH3OH bubble [21,24,51,52]. M is the third Body. A is in (m3 mol− 1 − 1
s ) for two body reaction
[(m6 mol− 2 s− 1) for a three body reaction], Ea is in (kJ mol− 1) and ΔH is in (kJ mol− 1).
N◦ Reaction A n Ea ΔH

1 H2O + M ⇌ H•+•OH + M 1.912 × 107 − 1.83 28.35 508.82


2 CH3OH ⇌ •CH3 + •OH 1.9 × 1010 0.0 3.84 × 105 374.114
3 CH3OH ⇌ CH2OH• + H• 1.54 × 1010 0.0 4.05 × 105 390.114
4 CH3OH + H• ⇌ CH2OH• + H2 1.35 × 10-3 3.2 1.46 × 104 − 34.057
5 •
OH + •OH ⇌ HO2• + H• 1.08 × 1005 0.61 1.51 × 1005 154.957
6 HO2• + •OH ⇌ H2O + O2 2.89 × 1007 0.00 − 2.08 × 1003 − 291.451
7 HO2• + O ⇌ •OH + O2 2.00 × 1007 0.00 0.00 − 224.131
8 HO2• + H• ⇌ H2 O + O 3.10 × 1007 0.00 7.20 × 1003 –22.957
9 HO2• + H• ⇌ H2 + O2 1.66 × 1007 0.00 3.44 × 1003 − 230.293
10 HO2• + H• ⇌ 2 •OH 7.08 × 1007 0.00 1.23 × 1003 − 155.613
11 CH2OH• + O2 ⇌ CH2O + HO2• 5.00 × 1006 0.00 0.00 − 79.404
12 CH3OH + •OH ⇌ CH3O•+ H2O 4.4 × 1000 2.0 6.3 × 103 − 54.215
13 CH3OH + •OH ⇌ CH2OH• + H2O 1.44 × 1000 2.0 − 3.51 × 103 − 95.215
14 CH3O•+ M ⇌ CH2OH• + M 1.00 × 1008 0.00 8.00 × 1004 − 38.000
15 CH3O•+ M ⇌ CH2O + H•+ M 7.78 × 1007 0.00 5.65 × 1004 126.297
16 CH3O•+ O2 ⇌ CH2O + HO2• 4.28 × 10− 19 7.60 − 1.48 × 1004 − 117.404
17 CH2O + HO2• ⇌ HCO• + H2O2 4.11 × 10− 02 2.50 4.27 × 1004 − 48.976
18 CH2O+ •OH ⇌ HCO• + H2O 3.90 × 1004 0.89 1.70 × 1003 − 128.159
19 CH2O + O ⇌ HCO• + •OH 3.50 × 1007 0.00 1.47 × 1004 − 60.839
20 CH2O + H• ⇌ HCO •+ H2 5.74 × 1001 1.90 1.15 × 1004 − 67.001
21 HCO• + O2 ⇌ CO + HO2• 7.58 × 1006 0.00 1.72 × 1003 − 140.530
22 HCO• + M ⇌ CO + H•+ M 1.86 × 1011 − 1.00 7.11 × 1004 − 65.171
23 CO + •OH ⇌ CO + H•+ CO2 4.40 × 1000 1.50 − 3.10 × 1003 − 212.853
24 2 HO2• ⇌ H2O2 + O2 3.02 × 1006 0.00 5.80 × 1003 − 212.268
25 H2O2 + M ⇌ 2 •OH + M k0 8.15 × 1017 − 1.92 2.08 × 1005 262.356
k∞ 2.62 × 1013 − 1.39 2.15 × 1005
26 H•+ O2 + M ⇌ HO2• + M k0 5.75 × 1007 − 1.40 0.00 − 205.701
k∞ 4.65 × 1000 0.44 0.00
27 H2 + O ⇌ •OH + H• 5.06 × 10− 02 2.67 2.63 × 1004 6.162
28 H•+ O2 ⇌ •OH + O 3.52 × 1010 − 0.70 7.14 × 1004 68.518
29 CH2OH•+ H• ⇌ CH2O + H2 3 × 107 0.0 0.0 − 309.578
30 2 •OH ⇌ HO2• 1.08 × 105 0.61 1.51 × 105 154.957
31 CH2OH• + •OH ⇌ CH2O + H2O 2.4 × 107 0.0 0.0 − 370.7
32 •
CH3 + HO2• ⇌ CH4 + O2 3.61 × 106 0.0 0.0 –233.852

Third body efficiency factors: R1:αH2 = 2.5, αH2O = 16, αAr = 1, αCO = 1.9, αCO2 = 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, others 1. R22:αH2 = 1.9, αH2O = 12, αAr =, αCO = 2.5, αCO2 = 2.5, αCH4 =, others 1; R25:αH2 =
2, αH2O = 6, αAr = 0.4, αCO = 1.5, αCO2 = 2, αCH4 =, others 1; R26:αH2 = 2.5, αH2O = 16, αAr = 0.7, αCO = 1.2, αCO2 = 2.4, αCH4 =, others 1.

dnk 4 3 of reaction energy. kfi and kri are the forward and reverse rate constants
= πR U̇ k (18)
dt 3 of the ith reaction, respectively. kfi and kri for the ith reactions are given
by the modified Arrhenius law [56]:
where, U̇H2 O , U̇CH3 OH andU̇k are the variation rates of H2O, CH3OH, and ( )
Eafi
all other species inside the bubble (except Ar, i.e. inert species) due to kfi = Afi T bfi exp − (21)
Rg T
chemical reactions involved in Tables 1, 2 and 3. U̇k is calculated as:
( )
1 dnk ∑
I Eari
U̇k = (υ˝ − υ’)ri (k = 1, ⋯, K) (19) kri = Ari T bri exp − (22)
V dt i=1 Rg T

Arrhenius parameters (A, b and E) of each chemical reaction are


where nk is the mole number of the kth species (including H2O and given in Tables 1-2.
methanol). The rate ri for the ith reaction is given as:
∏K ∏K 1.2. Resolution methodology
ri = kfi k=1
[Xk ]υ’ki − kri k=1
[Xk ]υ˝ki (20)

[Xk] is the concentration of the kth species. If a third body (M) is The nonlinear second-order differential equation of the bubble dy­

applicable, Eq. (20) should be multiplied by [ (αki)(Xk)], (k = 1,…K), namics (Eq. (1)) has been converted into a system of two differential
where αki is the third body effeciency factor of the kth species in the ith first-order equations as
reaction. If it is not stated under Table 1-3, αki = 1 for all substances. The dR
role M is stabilizing chemical products through eliminating the excessive = Ṙ (23)
dt

( )[ ( ( )) ]
1 Ṙ R
1+ PB (t) − PA sin 2π f t + − P∞ +
ρL C C
( ) ( ) ( )
m̈R Ṙ ṁ ṁ ṁ Ṙṁ R dPB 3 2 Ṙ 2ṁ
1− + + Ṙ + + + − Ṙ 1 − +
dṘ ρL C C ρL ρL 2ρL 2CρL CρL dt 2 3C 3CρL
= ( ) (24)
dt
1 − CṘ + Cṁρ R
L

6
A. Dehane et al. Chemical Engineering Journal 433 (2022) 133272

Equations (13), (14), (16), (17), (18), (23) and (24) are solved using P0 is the internal bubble pressure (Pv,0 + PCH3OH,0 + Pi,0) at equilibrium,
finite difference method. The outputs of these equations are the varia­ R = R0 (Pi,0: initial gas pressure, i = Ar, O2). The initial pressure of
tion of bubble temperature, pressure, radius and wall velocity as well as CH3OH within the bubble is governed by Raoult’s law: PCH3OH,0 =
the quantities (mole numbers) of all species (except argon, which is XCH3OH,0 . Psat,CH3OH,0 (Psat,CH3OH,0 = 0.13 bar at 20 ◦ C), where X CH3OH,0
chemically inert) in the bubble, all along the bubble oscillation (with is the CH3OH mole fraction in the aqueous phase (i.e. from 0 to 1). The
respect to time). The initial pressures of Ar and/or O2, H2O and CH3OH water vapor pressure (Pv,0) is calculated through Antoine’s equation
inside the bubble fulfill the equilibrium relation: P0 = P∞ +2σ /R0, where (2.33 kPa at 20 ◦ C [57]). The initial gas pressure Pi,0 is calculated as Pi,0

Fig. 1. Evolution of the different chemical species, versus time, inside an argon bubble at around the end of the bubble collapse, in the absence (a) and in the
presence (b) of methanol at 10% in the bulk solution (simulation conditions: frequency = 355 kHz, Acoustic intensity In = 1 w/cm2, liquid temp. Tliq = 20 ◦ C, static
pressure P∞ = 1 atm, initial bubble radius R0 = 3.2 µm [typical value at 355 kHz], saturation gas: Argon).

7
A. Dehane et al. Chemical Engineering Journal 433 (2022) 133272

Fig. 2. The normalized yields of H2, CH2O and CO as function of methanol concentration (v/v) in solution for the present paper compared to those obtained by
Kerboua et al. [24] and Buttner et al., [11] at 1000 kHz under argon atmosphere.

8
A. Dehane et al. Chemical Engineering Journal 433 (2022) 133272

typical ambient radius of the bubbles is 3.2 µM [29]. For the same
conditions, the simulation results for methanol/water solution (10% v/v
of CH3OH) is depicted in Fig. 1(b). According to Fig. 1(b), it is clearly
observed that the decomposition of methanol inside an argon-saturated
bubble (Table 2) modifies completely the chemistry of the bubble
(especially around the end of collapse). The important substances are:
CH2OH• (2.107 × 10-19 mol), CH2O (1.159 × 10-21 mol), CH3O● (1.717
× 10-20 mol), CH3● (5.1 × 10-21 mol), ●OH (7.03 × 10-20 mol), H●
(1.617 × 10-20 mol) and H2 (1.4 × 10-19 mol), as calculated at the
maximum bubble temperature (Tmax = 3876.4 K). This behavior
(modification of bubble chemistry) has been corroborated experimen­
tally by Henglein and Schulz [9], Rassokhin et al. [12] and through other
volatile species as it was demonstrated by Henglein [58] in presence of
CO2, N2O and CH4. Moreover, a similar trend has been obtained in our
previous works [43,44] in the presence of carbon tetrachloride within
the acoustic bubble. Compared to the argon bubble, the CH3OH (at 10%
v/v) decreases the Tmax by 15.73% [from 4600 K to 3876 K, Fig. 1(a)-
(b)]. Note that the time scale in Fig. 1(a) is only 0.006 µs (from 0.155 to
2.161 µs) and in Fig. 1(b) is 0.0014 µs (from 2.2585 to 2.2599 µs), at
around the end of the bubble collapse. As a result, the temperature
profile is always superior to 1500 K. However, if we reveal the complete
view of the bubble temperature profile we get the plots of Figs. S2(a) and
S2(b) of the Supplementary Material. These figures show clearly that
bubble temperature goes up gradually from 293.15 K and comes back to
this lower temperature (293.15 K) after attaining its maximum
depending on the chemical composition of the acoustic bubble.
Fig. 1(b) shows clearly the increase (by 4.7 times) of hydrogen
production compared to the case where methanol is absent in the bulk
solution (Fig. 1(a)). In addition, the yield of hydroxyl radicals decreased
by 8.85 times due to the scavenging effect of methyl alcohol within the
acoustic cavitation (Reactions (4) and (5) in Table 2). The results
Fig. 3. Bubble temperature (Tmax) as function of (a) the molar and (b) the depicted in Fig. 1(a) and (b) indicate the plausible modification of the
volumetric concentration of methanol in the bulk solution (frequency: 355 kHz). range of active bubbles for hydrogen yield (targeted species) because of
In Figure (a) a comparison has been made between our simulated temperature the dissociation of methyl alcohol inside the bubble. This effect will be
and those retrieved experimentally by Rea et al. [54]. analyzed with respect to the optimum amount of methanol (in solution)
and argon and oxygen in the bubble.
= P0-(Pv,0 + P CH3OH,0). Initial mole fractions of CH3OH, H2O and Ar Before discussing the different results obtained in our work, we
and/or O2 were calculated as Pi,0/P0, where i = CH3OH, H2O, Ar and O2. would like to firstly compare the normalized production of H2, CH2O
Accordingly, the initial bubble composition varies with the CH3OH and CO for an Ar-CH3OH-bubble (Table 2) to those found theoretically
liquid concentration due to the change in PCH3OH,0. For a methanol by Kerboua et al. [24] and experimentally by Buttner et al. [11] at 1000
concentration of 50% (v/v) in the bulk liquid and 50% (mole fraction, kHz (In: 2 W/cm2). In Fig. 2(a)-(d) the normalized yields of H2, CH2O
inside the bubble) of argon (at 20 ◦ C), initial mole fractions of CH3OH, and CO are given over a range of 0 to 80 % (v/v) for an aqueous solution
H2O, Ar and O2 were 0.029 (2.9%),0.017 (1.7%), 0.477 (47.7%) and of methanol. As it can be seen, the outputs of our chemical mechanism
0.477 (47.7%), respectively. (Ar-CH3OH-bubble, Table 2) are in good concordance with those ob­
tained by Kerboua et al. [24] and Buttner et al. [11] for the whole range of
2. Results and discussions methanol volume concentration (0–80%). For example, in the case of
our study and that of Kerboua et al.’s work [24], the maximal produc­
Throughout this work, the word “efficiency” is reserved for the tion of hydrogen is obtained at ~ 15% of methanol in solution, whereas
comparison of H2 production and CH3OH conversion (in mol) [and the this optimum is shifted to ~ 5% for the experimental work of Buttner
wideness of R0’s range] at different molar concentrations of argon (or et al. [11] (Fig. 2(a)). On the other hand, the maximal yield of formal­
oxygen) inside the bubble and volumetric concentrations of CH3OH in dehyde and monoxide of carbon are observed at ~ 30% and 10%,
the solution. Therefore, the maximal efficiency is observed when the respectively, as it is observed from the experimental results of Buttner’s
maximal molar production of H2 and CH3OH conversion (and the largest group [11] (Fig. 2(b) and (c)). However, the optimums of these species
of R0’s range) is obtained at a determined molar fraction of argon within (CH2O and CO, respectively) are obtained at around ~ 20% for Kerboua
the bubble and methanol concentration in the bulk liquid. et al. [24] and ~ 15% according to our results. These optimums [Fig. 2
(a)-(d)] are due to the fact that when methanol concentration increased,
2.1. Methanol effect on the bubble kinetics bubble temperature goes down, however, the yield of the different
species continues to be increased. Nevertheless, after certain concen­
Fig. 1(a) shows the evolution of the different species at around the tration of methanol, the lower temperature inside the bubble is unable to
end of the bubble collapse for an argon bubble (Table 1), where the main ensure the increase of the different substances production, therefore,
substances created at this level are H● (5.937 x10-19 mol), ●OH (6.225 optimum productions (of the different species) are created. On the other
x10-19 mol), O (1.539 × 10-20 mol) and H2 (2.98 × 10-20 mol), where the hand, Fig. 3(a) shows a good agreement between the bubble tempera­
maximum bubble temperature is Tmax = 4600 K. The same trend was tures found in our study (Table 2) and those obtained experimentally by
obtained by several studies [24,47,55]. The numerical simulations of Rae et al. [54] at 355 kHz in the range of methanol concentration from
Fig. 1(a) were given for a bubble oscillates in Ar-saturated water (20 ◦ C) 0 to 500 mM. These results indicate that our model is accurate for pre­
at frequency of 355 kHz and an acoustic intensity of 1 W/cm2. The dicting the bubble temperature, and therefore, it can be used with

9
A. Dehane et al. Chemical Engineering Journal 433 (2022) 133272

Fig. 4. Variation of (a) the bubble temperature (Tmax), (b) hydrogen production, (c) methanol conversion and (d) the total bubble yield as function of the molar
quantity of argon (inside the bubble) and methanol concentration in the bulk liquid [% (v/v)]. Simulations were conducted for the same conditions as in Fig. 1 (f =
355 kHz, In = 1 w/cm2, Tliq = 20 ◦ C, P∞ = 1 atm, R0 = 3.2 µm [typical value at 355 kHz], sat. gas: Argon).

confidence for studying hydrogen recovery from methanol sono- ~ 3900 K). This is probably owing to the lower concentrations adopted
pyrolysis within acoustic bubbles. by Rae et al. [54]. For this reason, we extended the evolution of the
As it can be seen from Fig. 3(a), the presence of 100 mM of methanol bubble temperature over the whole range of methanol concentration in
(in the bulk solution) reduces bubble temperature from 4600 to 3930 K solution [0–100 % (v/v)] (Fig. 3(b)). As it was expected, the increase of
(14.56 % of decrease) as a result for the increase of bubble heat capacity methanol concentration in the bulk liquid has a negative impact on the
(presence of CH3OH) and the endothermal dissociation of methanol. peak temperature of our acoustic cavitation, where the maximal tem­
However, the increase of methyl alcohol concentration (from 100 to 500 perature goes down gradually from 4600 (0% of CH3OH) to 2941 K
mM) has practically no effect on the maximal bubble temperature (Tmax (100% of CH3OH).

10
A. Dehane et al. Chemical Engineering Journal 433 (2022) 133272

Fig. 4. (continued).

2.2. Optimal conditions for the maximal hydrogen yielding and methanol K for the same interval of argon quantity (0 to 100%). On the other hand,
conversion for hydrogen production, methanol conversion and the total yield of the
bubble (Fig. 4(b)-(d)), optimum points are observed at 80%, 60–80%
In order to reveal the effect of methanol on the range of active and 80% of argon mole fraction (inside the bubble), respectively, for the
bubbles (in term of R0) for hydrogen production, the ideal conditions of majority of methanol concentrations. The maximal conversion of
bubble composition (argon amount) should be determined for the pur­ methanol (1.38 × 10-17 mol) is retrieved at 3% (v/v) of CH3OH and 70%
pose to get a clear view of methanol impact on the range of active of argon mole fraction (Fig. 4(c)), which corresponds to 9.46 × 10-19 mol
bubbles, and obtain the initial composition of liquid and acoustic cavi­ for the maximal production of hydrogen (Fig. 4(b)).
tation giving the maximum efficiency. According to several experi­ On the other hand, the maximal yield of hydrogen is obtained at 7%
mental works [59–62], the maximal activity of the bubble is obtained at (v/v) of methanol and 80% of argon (mole fraction) with a value of 1.13
around 355 kHz. Therefore, the ultrasound frequency is kept fixed at × 10-18 mol (19.45% of increase compared to the previous case, Fig. 4
355 kHz (In = 1 W/cm2). Fig. 4(a)-(d) show the evolution of the bubble (b)). This production (1.13 × 10-18 mol) corresponds to 1.3 × 10-17 mol
temperature (Tmax), hydrogen production, maximal methanol conver­ of methanol conversion, giving a decrease of 5.79% compared to the
sion and the total yield of the bubble, respectively. previous case (1.38 × 10-17 mol) (Fig. 4(c)). For the total yield of
From Fig. 4(a), it can be observed that the bubble temperature is acoustic cavitation (Fig. 4(d)), it is observed that the majority of
increased proportionally with the rise of argon amount inside the bubble maximum productions are obtained at 80 % (mole fraction) of argon
or through the reduction of methanol volume fraction in solution inside the bubble over the whole range of methanol concentration
(especially over 10% (v/v) of methanol in the bulk liquid). For example, (0–100% (v/v)). It should be noted that according to Fig. 4(c)-(d), in
in the absence of methanol, the increase of the initial argon percentage general, the total yield of the bubble is slightly lower than the maximal
in the bubble from 0 to 100% causes the maximal cavitation temperature conversion of methanol. This is ascribed to the conversion of a small
to be increased from 2806 to 4540 K. Nevertheless, for a methanol so­ amount of methanol to water (during the compression phase) which in
lution (100% CH3OH), bubble temperature increases from 2009 to 2942 turn is condensed during the strong collapse (lost). The analyzation of

11
A. Dehane et al. Chemical Engineering Journal 433 (2022) 133272

Fig. 5. Effect of methanol concentration [% (v/v)] in the bulk solution on the range of active bubbles for (a) hydrogen yielding, (b) methanol conversion and (c) the
normalized wideness of active bubbles range (for H2 production and methanol conversion) at 80 % of argon within the bubbles, for the same conditions as in Fig. 1 (f
= 355 kHz, In = 1 w/cm2, Tliq = 20 ◦ C, P∞ = 1 atm, R0 = 3.2 µm [typical value at 355 kHz], sat. gas: Argon).

Fig. 4(b)-(d) indicates that the maximal efficiency of the bubble is ob­ It is worth mentioning that the obtained optimal composition of bubble
tained at 80% of argon (20% oxygen), where the best combination for in saturating gases (80% Ar and 20% O2) is in line with various exper­
the maximal hydrogen production (1.13 × 10-18 mol) and methanol imental studies [61,63–65] investigating the sonoactivity of acuostic
conversion (1.3 × 10-17 mol) is observed at 7% (v/v) of methanol and bubble.
80% of argon. At this level, hydrogen production (1.13 × 10-18 mol) is Fig. 4(b)-(d) confirm also that the concentration of methanol (in
increased by 32.7 times compared to the maximal value obtained in solution) can be increased until 20%, where the yield of hydrogen is
absence of methanol (and 100% Ar), which is equal to 3.46 × 10-20 mol. increased by 26.16 (9.05 × 10-19 mol) and 22.4 (7.74 × 10-19 mol) times

12
A. Dehane et al. Chemical Engineering Journal 433 (2022) 133272

at 10% and 20% of methanol, respectively, compared to the case in concentrations (0–100%), the mean radii (for the maximal yield of
absence of methanol (and 100% Ar). At 10% and 20% (v/v) of methanol hydrogen and methanol conversion) are observed at 3.2 µm, which is in
in the bulk solution, its conversion inside the bubble decreased by good agreement with that obtained experimentally by Brotchie et al.
13.84% (1.12 × 10-17 mol) and 22.3 % (1.01 × 10-17 mol), respectively, [29] for sonochemiluminescing bubbles at 355 kHz. In order to reveal
compared to the maximal conversion registered at 7% (1.3 × 10-17 mol). the evolution of active bubbles ranges (for H2 production and CH3OH
More information could be obtained by the analyzation of the whole conversion) as function of the amount of methanol in the bulk liquid,
range of active bubbles (detailed in the next section). Fig. 5(c) shows the normalized wideness of R0 (of active bubbles) ranges
for hydrogen yield and methanol conversion as function of methanol
2.3. Effect of methanol on the range of active bubbles for hydrogen concentrations (0–100%). As it can be observed in Fig. 5(c), the varia­
production tion of methanol amount (volume fraction) in solution seems to have a
little impact on the wideness of R0 (active bubbles radii) ranges for
From Fig. 5(a) and (b) it can be observed that for methanol con­ methanol conversion, where the normalized wideness of R0 ranges are in
centrations (in the bulk liquid) lower than 10%, some fluctuation is the range of 0.9 to 1. The increase of methanol concentration from 1% to
observed either for the maximal conversion of methanol or for the 70–80% increases hardly (10% of increase) the range of active bubbles
maximal yield of hydrogen. As it was pointed out in the previous section, (normalized values vary from 0.9 to 1). At 90% of methanol (in solution)
where the best combination (maximal production of hydrogen and this normalized value goes down to 0.94 (for 100% methanol). At the
conversion of methanol) is obtained at 7% of methanol (and 80% of opposite, hydrogen yielding shows a different behavior compared to
argon) in the aqueous solution. Fig. 5(a) and (b) indicate clearly the that of methanol conversion (Fig. 5 (c)), where the normalized wideness
possibility of extension of this concentration to 20% (v/v) of methanol of active bubbles (for H2 production) ranges increases rapidly from 0.5
(and 80% of Ar). Over this amount (20% of CH3OH), a relatively rapid (0% of CH3OH) to 1 by the introduction of only 1% (v/v) of methanol.
decrease of the maximal yield of hydrogen and methanol conversion is This normalized value is slightly decreased (by 4%) when methanol
recorded. For example, at 30% (v/v) of methanol, the maximal pro­ concentration is increased to 20% in the bulk liquid. From this con­
duction of hydrogen (5.72 × 10-19 mol) and methanol conversion (8.32 centration (20%), the normalized wideness of active bubbles ranges (for
× 10-18 mol) are reduced by 49.4% and 36%, respectively, compared to H2 production) starts to be gradually decreased attaining 0.56 (44% of
those values obtained at 7% of methanol (1.13 × 10-18 and 1.3 × 10-17 decrease) for a pure solution of methanol (100% CH3OH). As a reca­
mol for H2 production and CH3OH conversion, respectively). This pitulation, Table 4 gives the evolution of R0 ranges for hydrogen
decrease goes up to 89.4% and 77.92% respectively for H2 production yielding (Fig. 5(a)) and methanol conversion (Fig. 5 (b)). The pro­
(1.2 × 10-19 mol) and CH3OH conversion (2.87 × 10-18 mol) for meth­ nounced effect of methanol concentration (in liquid) on the ranges of
anol concentration of 80% (v/v) in the bulk liquid. This behavior is hydrogen production (decrease with the increase of methanol concen­
owing to the fact that when the amount of methanol is increased within tration in liquid) compared to that of methanol conversion (weak
the bubble, heat capacity of this cavitation is increased which means impact) is considered as a logical result. Because as it was observed
that the maximal temperature is reduced (Fig. 4 (a)), consequently, the previously for a single bubble (Fig. 4 a) and (c)), in spite of the decrease
conversion of CH3OH and H2 production go down. This negative impact of bubble temperature with the increase of methanol concentration in
of methanol was obtained previously in the case of other volatile com­ solution (for a fixed amount of argon), the molar production of the
pounds such as carbon dioxide [66–68] and carbon tetrachloride [44] at bubble and methanol conversion persist to be clearly greater than the
different amounts depending on their physical properties. yielding threshold (4.67 × 10-22 mol). At the opposite, the molar pro­
Fig. 5(a) and (b) show that approximately for all methanol duction of hydrogen seems to be more affected by the decrease of bubble
temperature as one of the methanol conversion species (Fig. 4(a) and
Table 4
(b)). Therefore, when methanol concentration is increased in the bulk
Variation of active bubbles ranges (µm) for hydrogen production and methanol liquid, more bubbles (at the range limits) are supposed to be activated
conversion as function of methanol concentration (% vol.) and argon molar (for methanol conversion) due to the additional amount of methanol
fraction within the bubble. (Conditions: frequency = 355 kHz, Intensity In = 1 W/ present within these bubbles [43], however, the decrease of tempera­
cm2, liquid temp. TLiq = 20 ◦ C, static pressure P∞= 1 atm). tures in these bubbles makes it difficult to keep the conversion of
/ Range of active bubbles for Range of active bubbles for methanol over the threshold (4.67 × 10-22 mol). Consequently, the
80% of argon (inside the 100% of argon (inside the range of active bubbles for methanol consumption is hardly increased
bubble) bubble) (Table 4). On the other hand, the increase of methanol concentration in
Methanol H2 CH3OH H2 CH3OH the bulk liquid (and inside bubbles) affects negatively the range of active
concentration (vol, Production Conversion Production Conversion bubbles for H2 production, where more bubbles are deactivated
%) (smallest and largest bubbles), because with the decrease of these bub­
0 1.5–3.92 – 1.35–4.6 – bles temperatures it is not possible to keep the quantity of hydrogen
1 1.3–6.23 0.79–7.85 1.7–5.28 1.1–5.86 produced inside bubbles greater than that of production threshold (4.67
2 1.33–6.22 0.79–7.85 1.6–5.28 1–5.87 × 10-22 mol).
3 1.4–6.22 0.79–7.86 1.6–5.27 1.05–5.86
4 1.42–6.21 0.79–7.86 1.64–5.26 1.05–5.86
For more comparison, from Fig. 6(a) and (b), it can be clearly
5 1.42–6.2 0.79–7.87 1.6–5.25 1.06–5.86 observed that the sonochemical activity is reduced by the increase of
6 1.43–6.2 0.79–7.88 1.62–5.25 1.06–5.86 argon mole fraction inside bubbles to 100%. For example, for hydrogen
7 1.45–6.2 0.78–7.88 1.63–5.25 1.06–5.84 production, it is shown in Fig. 6(a) that the yield of hydrogen is
8 1.5–6.2 0.78–7.89 1.64–5.25 1.05–5.84
decreased (compared to the previous case Fig. 5(a)) over the range of
9 1.52–6.2 0.78–7.9 1.6–5.24 1.05–5.85
10 1.52–6,2 0.78–7.9 1.61–5.23 1.05–5.85 active bubbles for all methanol concentrations, where this production is
20 1.44–6.16 0.77–7.97 1.65–5.18 1.17–5.79 lower than 8.2 × 10-20 mol (for all R0′ s). Similarly, the consumption of
30 1.45–6.11 0.76–8.05 1.75–5.12 1.14–5.73 methanol goes down over the whole range of R0 and for all methanol
40 1.47–6 0.75–8.15 1.74–5 1.1–5.67 concentrations (Fig. 6(b)). The maximal conversion is equal or lower
50 1.57–6 0.74–8.26 1.79–4.95 1.2–5.59
60 1.62–5.9 0.72–8.39 1.89–4.83 1.29–5.49
than 3.67 × 10-19 mol (for all R0′ s). In addition, it is observed in Fig. 6(c)
70 1.65–5.75 0.7–8.56 1.86–4.67 1.27–5.34 that the wideness of active bubbles either for hydrogen production or for
80 1.75–5.61 0.67–8.52 2.05–4.42 1.5–5.15 methanol conversion is negatively affected by the increase of methanol
90 1.84–5.35 0.63–8.43 2.2–4.05 1.55–4.85 concentration (0–100 %). Some exception is registered for methanol
100 2–4.8 0.69–8.1 2.22–4.35 1.7–5.02
concentration equal or lower than 2% (for H2 production and CH3OH

13
A. Dehane et al. Chemical Engineering Journal 433 (2022) 133272

Fig. 6. Effect of methanol concentration [% (v/v)] in the bulk solution on the range of active bubbles for (a) hydrogen yielding, (b) methanol conversion and (c) the
normalized wideness of active bubbles range (for H2 production and methanol conversion) at 100 % of argon within the bubbles, for the same conditions as in Fig. 5.

conversion) and greater than 90% (for hydrogen production). By conversion) is obtained at 80% of argon (inside bubble) and for meth­
comparing Fig. 5 and Fig. 6(a)-(b), it can be concluded that the maximal anol concentration between 7 and 20%. This conclusion indicates also
efficiency in terms of hydrogen production, methanol conversion and that not only the maximal temperature (inside the bubble) is the con­
the broadness of active bubbles ranges (for H2 yielding and CH3OH trolling key for the sonoactivity of the bubble but also the amounts of

14
A. Dehane et al. Chemical Engineering Journal 433 (2022) 133272

oxygen, argon and volatile solutes encapsulated within the bubble (see Declaration of Competing Interest
section 3.2).
The authors declare that they have no known competing financial
interests or personal relationships that could have appeared to influence
2.4. Energetic cost
the work reported in this paper.

In order to evaluate the energetic cost of our sono-system, it is


Acknowledgements
necessary to know the number of bubbles generated within the sono-
reactor. According to the semi-empirical technique of Merouani et al.
This study was supported by The Ministry of Higher Education and
[69], the number of active bubbles under a wave frequency of 355 kHz
Scientific Research of Algeria (project No. A16N01UN250320180001)
is ~ 2.8428 × 107 bubble/L s. On the other hand, according to Fig. 1(b),
and the General Directorate of Scientific Research and Technological
the amount of methanol converted per one acoustic bubble (at 355 kHz,
Development (GD-SRTD) of Algeria.
In = 1 W/cm2) during one cycle is 2.137 × 10-14 mol. Therefore, the
amount of methanol converted per unit volume per unit time is
Appendix A. Supplementary data
6.075 × 10-7 mol/L s, if it supposed that bubbles are fragmented at
the first collapse. Now, if we suppose that we have a reactor with the
Supplementary data to this article can be found online at https://doi.
following dimensions: volume V = 10-2 m3 and section S = 5 × 10-3 m2
org/10.1016/j.cej.2021.133272.
(i.e. a Lab-scale reactor), this means that our power density (Pd) in the
sono-reactor is 290.9 W/L [Pd= (Pa2. S)/(2 × ρL × C × V)], where Pa = References
√̅̅̅̅̅̅̅̅̅̅̅̅
2ICρL (ρL is the mass density of the liquid and C is the sound velocity in
the liquid). As a result, we have a power density of 290.9 J/s L and a [1] K. Kerboua, O. Hamdaoui, M.H. Islam, A. Alghyamah, H.E. Hansen, B.G. Pollet,
Low carbon ultrasonic production of alternate fuel: Operational and mechanistic
molar conversion of methanol of 6.075 × 10-7 mol/ L s in the whole concerns of the sonochemical process of hydrogen generation under various
reactor. This means that our energetical consumption is 7.54 × 10-3 scenarios, Int. J. Hydrogen Energy. (2021) in press. https://doi.org/10.1016/j.
mol/kWh for an acoustic intensity of 1 W/cm2 under a frequency of 355 ijhydene.2021.05.191.
[2] S. Merouani, O. Hamdaoui, The sonochemical approach for hydrogen production,
kHz.
in: Inamuddin, A. Asiri (Eds.), Sustain. Green Chem. Process. Their Allied Appl.
Nanotechnol. Life Sci., Springer, 2020: pp. 1–29. https://doi.org/10.1007/978-3-
3. Conclusion 030-42284-4_1.
[3] K. Yasui, Fundamentals of acoustic cavitation and sonochemistry, in:
M. Ashokkumar (Ed.), Pankaj, Springer ScienceþBusiness Media, New York, Theor.
This paper gives a special focus on the impact of methanol concen­ Exp. Sonochemistry Involv. Inorg. Syst., 2011, pp. 1–29.
tration (in the bulk liquid) on the range of active bubbles for hydrogen [4] T. Sivasankar, A.W. Paunikar, V.S. Moholka, Mechanistic approach to
enhancement of the yield of a sonochemical reaction, AIChE J. 53 (2007)
production and methanol conversion by sonolysis. It was demonstrated 1132–1143, https://doi.org/10.1002/aic.
that the dissociation of methanol within bubbles is able to completely [5] S. Merouani, O. Hamdaoui, Sonochemical treatment of textile wastewater, in: M.
modify the chemistry of these cavities (at the end of collapse) as well as P. Inamuddin, A. Asiri (Eds.), Water Pollut, Remediat. Photocatal, Springer-Nature
Switzerland, 2021 https://doi.org/https://doi.org/10.1007/978-3-030-54723-3_5.
the range of active bubbles for the yield of hydrogen and methanol
[6] K.S. Suslick, Sonoluminescence and Sonochemistry, Encycl. Phys. Sci. Technol. 3rd
conversion. It has been observed that the maximal sonochemical effi­ Ed.R. A. Meyers (Ed.); Acad. Press. Inc. San Diego, 2001. (2001).
ciency of bubbles is obtained at 80% of argon (inside the bubble), where [7] M.S. Peralta, J. Glassmire, M. Lazopoulou, K. Sumner, X. Vallvé, P. Lilienthal,
M. Ayuso, F. Boshell, Innovation Outlook: Renewable Mini-Grids, International
the best combination for the maximal H2 production and CH3OH con­
Renewable Energy Agency (IRENA) (2016). https://www.irena.org/publications/
version is obtained at 7% (v/v) of methanol and 80% of argon. In 2016/Sep/Innovation-Outlook-Renewable-mini-grids.
addition, it has been shown that it is possible to increase the concen­ [8] F. Dalena, A. Senatore, A. Marino, A. Gordano, M. Basile, A. Basile, Methanol
tration of methanol (in solution) until 20% without highly affecting the production and applications: An overview, in: A. Basile, F. Dalena (Eds.), Methanol
Sci. Eng., Elsevier B.V., 2018: pp. 3–28. https://doi.org/10.1016/B978-0-444-
bubble activity. On the other hand, it was obtained that the variation of 63903-5.00001-7.
methanol concentration [0–100% (v/v)] has a little effect on the wide­ [9] A. Henglein, R. Schulz, Der Einfluß organischer Verbindungen auf einige chemische
ness of active bubbles ranges for methanol conversion. In contrast, a Wirkungen des Ultraschalls, Zeitschrift Für Naturforsch. B. 8 (1953) 77–284.
https://doi.org/10.1515.
different behavior is observed for the production of hydrogen, where a [10] H.-S. Son, K.-D. Zoh, Effects of methanol and carbon tetrachloride on sonolysis of
gradual decrease of the active bubbles ranges is obtained for methanol 1,4-Dioxane in relation to temperature, Ind. Eng. Chem. Res. 51 (26) (2012)
concentrations greater than 20%. It has been concluded that the 8939–8944, https://doi.org/10.1021/ie201766h.
[11] J. Buettner, M. Gutierrez, A. Henglein, Sonolysis of water-methanol mixtures,
maximal efficiency for the production of hydrogen, wideness of active J. Phys. Chem. 95 (4) (1991) 1528–1530, https://doi.org/10.1021/j100157a004.
bubbles ranges (for H2 production and CH3OH conversion) and meth­ [12] D.N. Rassokhin, G.V. Kovalev, L.T. Bugaenko, Temperature effect on the sonolysis
anol conversion is obtained at 80% of argon and for methanol concen­ of methanol/water mixtures, J. Am. Chem. Soc. 117 (1) (1995) 344–347, https://
doi.org/10.1021/ja00106a037.
tration between 7 and 20%. Additionally, this work shows clearly that
[13] C.M. Krishna, Y. Lion, T. Kondo, P. Riesz, Thermal decomposition of methanol in
the sonoactivity of bubbles is not only controlled by the maximal tem­ the sonolysis of methanol-water mixtures. Spin-trapping evidence for isotope
peratures attained inside these cavities, but also by the amounts of exchange reactions, J. Phys. Chem. 91 (23) (1987) 5847–5850, https://doi.org/
saturating gases (e.g. O2, Ar…) and volatile solutes housed within these 10.1021/j100307a007.
[14] K. Okitsu, H. Nakamura, N. Takenaka, H. Bandow, Y. Maeda, Y. Nagata,
bubbles. Sonochemical reactions occurring in organic solvents: reaction kinetics and
Finally, the simulations performed in the present paper are aiming reaction site of radical trapping with 1, 1-diphenyl-2-picrylhydrazyl, Res. Chem.
principally to reveal the effect of CH3OH on the range of active bubbles Intermed. 30 (7-8) (2004) 763–774, https://doi.org/10.1163/
1568567041856864.
for H2 production and methanol conversion. Therefore, as it is clarified, [15] Y. Mizukoshi, H. Nakamura, H. Bandow, Y. Maeda, Y. Nagata, Sonolysis of organic
the best combinations of bubble composition (80% Ar and 20% O2) and liquid: effect of vapour pressure and evaporation rate, Ultrason. Sonochem. 6
liquid concentration in methanol (7–20 % (v/v)) are determined. This (1999) 203–209, https://doi.org/10.1016/S1350-4177(99)00012-7.
[16] M. Ashokkumar, R. Hall, P. Mulvaney, F. Grieser, Sonoluminescence from aqueous
means that our theoretical simulations give the first contribution to the alcohol and surfactant solutions, J. Phys. Chem. B. 101 (50) (1997) 10845–10850,
practical side of sonochemistry in terms of hydrogen production and https://doi.org/10.1021/jp972477b.
methanol conversion by ultrasound. Additionally, the discussed results [17] M. Ashokkumar, L. Crum, C.A. Frensley, F. Grieser, T.J. Matula, W.
B. Mcnamara III, K.S. Suslick, Effect of solutes on single-bubble sonoluminescence
(variation of R0 range for H2 production and CH3OH conversion) in the in water, J. Phys. Chem. A. 104 (2000) 8462–8465, https://doi.org/10.1021/
present paper are treated for the first time, consequently, our work is jp000463r.
considered as a guideline for increasing the efficacy of hydrogen pro­
duction in a sonochemical reactor.

15
A. Dehane et al. Chemical Engineering Journal 433 (2022) 133272

[18] R. Tögel, S. Hilgenfeldt, D. Lohse, Squeezing alcohols into sonoluminescing improves the sonolytic degradation of nonvolatile organic contaminants, Sep.
bubbles: The universal role of surfactants, Phys. Rev. Lett. 84 (11) (2000) Purif. Technol. 275 (2021) 118614, https://doi.org/10.1016/j.
2509–2512, https://doi.org/10.1103/PhysRevLett.84.2509. seppur.2021.118614.
[19] M.H. Islam, J.J. Lamb, K.M. Lien, O.S. Burheim, J.-Y. Hihn, B.G. Pollet, Novel fuel [45] D.W. Green, R.H. Perry, Perry’s chemical engineers’ handbook, Eighth Edition,
production based on sonochemistry and sonoelectrochemistry, ECS Trans. 92 McGraw-Hill, New York, USA, 2008.
(2019) 1–16, https://doi.org/10.1149/09210.0001ecst. [46] W. Chen, X.i. Chen, M. Lu, G. Miao, R. Wei, Single bubble sonoluminescence driven
[20] Md.H. Islam, O.S. Burheim, B.G. Pollet, Sonochemical and Sonoelectrochemical by non-simple-harmonic ultrasounds, J Acoust Soc Am. 111 (6) (2002) 2632–2637,
Production of Hydrogen, Ultrason. Sonochem. 51 (2019) 533–555, https://doi. https://doi.org/10.1121/1.1480417.
org/10.1016/j.ultsonch.2018.08.024. [47] K. Yasui, Effect of liquid temperature on sonoluminescence, Phys. Rev. E. 64
[21] K. Kerboua, O. Hamdaoui, S. Al-Zahrani, Sonochemical production of hydrogen: A (2001), 016310, https://doi.org/10.1103/PhysRevE.64.016310.
numerical model applied to the recovery of aqueous methanol waste under [48] Y.S. Touloukian, P.E. Liley, S.C. Saxena, Thermal Conductivity: Nonmetallic
Oxygen-Argon atmosphere, Environ. Prog. Sustain. Energy. in press (2020), Liquids and Gases, IFI/Plenum (1990).
https://doi.org/10.1002/ep.13511. [49] K. Yasui, Effect of volatile solutes on sonoluminescence, J. Chem. Phys. 116 (7)
[22] K. Kerboua, S. Merouani, O. Hamdaoui, A. Alghyamah, M.H. Islam, H.E. Hansen, B. (2002) 2945–2954, https://doi.org/10.1063/1.1436122.
G. Pollet, How do dissolved gases affect the sonochemical process of hydrogen [50] K. Yasui, T. Tuziuti, Y. Iida, Optimum bubble temperature for the sonochemical
production? An overview of thermodynamic and mechanistic effects – On the “ hot production of oxidants, Ultrasonics. 42 (1-9) (2004) 579–584, https://doi.org/
spot theory”, Ultrason. Sonochem. 72 (2021) 105422, https://doi.org/10.1016/j. 10.1016/j.ultras.2003.12.005.
ultsonch.2020.105422. [51] R. Seiser, K. Seshadri, F.A. Williams, Detailed and reduced chemistry for methanol
[23] K. Kerboua, O. Hamdaoui, Energetic challenges and sonochemistry: A new ignition, Combust. Flame. 158 (9) (2011) 1667–1672, https://doi.org/10.1016/j.
alternative for hydrogen production? Curr. Opin. Green Sustain. Chem. 18 (2019) combustflame.2011.02.008.
84–89, https://doi.org/10.1016/j.cogsc.2019.03.005. [52] T.S. Norton, F.L. Dryer, Toward a comprehensive mechanism for methanol
[24] K. Kerboua, O. Hamdaoui, Oxygen-argon acoustic cavitation bubble in a water- pyrolysis, Int. J. Chem. Kinet. 22 (3) (1990) 219–241, https://doi.org/10.1002/
methanol mixture: Effects of medium composition on sonochemical activity, (ISSN)1097-460110.1002/kin.v22:310.1002/kin.550220303.
Ultrason. - Sonochemistry. 61 (2020) 104811, https://doi.org/10.1016/j. [53] W. Tsang, Chemical kinetic data base for combustion chemistry. Part 2. Methanol,
ultsonch.2019.104811. J. Phys. Chem. Ref. Data. 16 (3) (1987) 471–508, https://doi.org/10.1063/
[25] A. Thiemann, T. Nowak, R. Mettin, F. Holsteyns, A. Lippert, Characterization of an 1.555802.
acoustic cavitation bubble structure at 230 kHz, Ultrason. Sonochem. 18 (2) (2011) [54] J. Rae, M. Ashokkumar, O. Eulaerts, C. von Sonntag, J. Reisse, F. Grieser,
595–600, https://doi.org/10.1016/j.ultsonch.2010.10.004. Estimation of ultrasound induced cavitation bubble temperatures in aqueous
[26] S. Labouret, J. Frohly, Bubble size distribution estimation via void rate dissipation solutions, Ultrason. Sonochem. 12 (5) (2005) 325–329, https://doi.org/10.1016/j.
in gas saturated liquid. Application to ultrasonic cavitation bubble field, Eur. Phys. ultsonch.2004.06.007.
J. Appl. Phys. 19 (2002) 39–54, https://doi.org/10.1051/epjap:2002047. [55] A. Dehane, S. Merouani, O. Hamdaoui, A. Alghyamah, Insight into the impact of
[27] F. Burdin, N.A. Tsochatzidis, P. Guiraud, A.M. Wilhelm, H. Delmas, excluding mass transport, heat exchange and chemical reactions heat on the
Characterisation of the acoustic cavitation cloud by two laser techniques, Ultrason. sonochemical bubble yield: Bubble size-dependency, Ultrason. Sonochem. (2021),
Sonochem. 6 (1-2) (1999) 43–51, https://doi.org/10.1016/S1350-4177(98)00035- 105511, https://doi.org/10.1016/j.ultsonch.2021.105511.
2. [56] S. Merouani, O. Hamdaoui, Y. Rezgui, M. Guemini, Sensitivity of free radicals
[28] N.A. Tsochatzidis, P. Guiraud, A.M. Wilhelm, H. Delmas, Determination of production in acoustically driven bubble to the ultrasonic frequency and nature of
velocity, size and concentration of ultrasonic cavitation bubbles by the phase- dissolved gases, Ultrason. Sonochem. 22 (2014) 41–50, https://doi.org/10.1016/j.
Doppler technique, Chem. Eng. Sci. 56 (5) (2001) 1831–1840, https://doi.org/ ultsonch.2014.07.011.
10.1016/S0009-2509(00)00460-7. [57] S. Merouani, O. Hamdaoui, Y. Rezgui, M. Guemini, Computer simulation of
[29] A. Brotchie, F. Grieser, M. Ashokkumar, Effect of power and frequency on bubble- chemical reactions occurring in collapsing acoustical bubble: dependence of free
size distributions in acoustic cavitation, Phys. Rev. Lett. 102 (084302) (2009) 1–4, radicals production on operational conditions, Res. Chem. Intermed. 41 (2) (2015)
https://doi.org/10.1103/PhysRevLett.102.084302. 881–897, https://doi.org/10.1007/s11164-013-1240-y.
[30] A. Brotchie, T. Statham, M. Zhou, L. Dharmarathne, F. Grieser, M. Ashokkumar, [58] A. Henglein, Sonolysis of carbon dioxide, nitrous oxide and methane in aqueous
Acoustic bubble sizes, coalescence, and sonochemical activity in aqueous solution, Zeitschrift F{ü}r Naturforsch, B. 40 (1985) 100–107, https://doi.org/
electrolyte solutions saturated with different gases, Langmuir. 26 (15) (2010) 10.1515/znb-1985-0119.
12690–12695, https://doi.org/10.1021/la1017104. [59] M.A. Beckett, I. Hua, Elucidation of the 1,4-dioxane decomposition pathway at
[31] S. Merouani, O. Hamdaoui, Y. Rezgui, M. Guemini, Computational engineering discrete ultrasonic frequencies, Environ. Sci. Technol. 34 (18) (2000) 3944–3953,
study of hydrogen production via ultrasonic cavitation in water, Int. J. Hydrogen https://doi.org/10.1021/es000928r.
Energy. 41 (2) (2016) 832–844, https://doi.org/10.1016/j.ijhydene.2015.11.058. [60] J.-W. Kang, H.-M. Hung, A. Lin, M.R. Hoffmann, Sonolytic destruction of methyl
[32] K. Yasui, T. Tuziuti, J. Lee, T. Kozuka, A. Towata, Y. Iida, The range of ambient tert-Butyl ether by ultrasonic irradiation: The role of O3, H2O2, frequency, and
radius for an active bubble in sonoluminescence and sonochemical reactions, power density, Environ. Sci. Technol. 33 (1999) 3199–3205, https://doi.org/
J. Chem. Phys. 128 (18) (2008) 184705, https://doi.org/10.1063/1.2919119. 10.1021/es9810383.
[33] S. Merouani, O. Hamdaoui, Y. Rezgui, M. Guemini, Effects of ultrasound frequency [61] M.A. Beckett, I. Hua, Impact of ultrasonic frequency on aqueous sonoluminescence
and acoustic amplitude on the size of sonochemically active bubbles-Theoretical and sonochemistry, J. Phys. Chem. A. 105 (15) (2001) 3796–3802, https://doi.
study, Ultrason. Sonochem. 20 (3) (2013) 815–819, https://doi.org/10.1016/j. org/10.1021/jp003226x.
ultsonch.2012.10.015. [62] S. Koda, T. Kimura, T. Kondo, H. Mitome, A standard method to calibrate
[34] S. Labouret, J. Frohly, Distribution en tailles des bulles d’un champ de cavitation sonochemical efficiency of an individual reaction system, Ultrason. Sonochem. 10
ultrasonore, 10ème Congrès Français d’Acoustique. (2010) Lyon, 12–16 April. (3) (2003) 149–156, https://doi.org/10.1016/S1350-4177(03)00084-1.
[35] S. Merouani, O. Hamdaoui, The size of active bubbles for the production of [63] E.J. Hart, A. Henglein, Free radical and free atom reactions in the sonolysis of
hydrogen in sonochemical reaction field, Ultrason. Sonochem. 32 (2016) 320–327, aqueous iodide and formate solutions, J. Phys. Chem. 89 (20) (1985) 4342–4347,
https://doi.org/10.1016/j.ultsonch.2016.03.026. https://doi.org/10.1021/j100266a038.
[36] K. Yasui, Effects of thermal conduction on bubble dynamics near the [64] A. Henglein, Chemical effects of continuous and pulsed ultrasound in aqueous
sonoluminescence threshold, J. Acoust. Soc. Am. 98 (5) (1995) 2772–2782, solutions, Ultrason. Sonochem. 2 (2) (1995) S115–S121, https://doi.org/10.1016/
https://doi.org/10.1121/1.413242. 1350-4177(95)00022-X.
[37] K. Yasui, A new formulation of bubble dynamics for sonoluminescence, PhD thesis, [65] J.D. Schramm, I. Hua, Ultrasonic irradiation of dichlorvos: Decomposition
Waseda University, 1996. mechanism, Water Res. 35 (3) (2001) 665–674, https://doi.org/10.1016/S0043-
[38] Y. Shen, K. Yasui, T. Zhu, M. Ashokkumar, A model for the effect of bulk liquid 1354(00)00304-3.
viscosity on cavitation bubble dynamics, Phys. Chem. Chem. Phys. 19 (31) (2017) [66] O. Authier, H. Ouhabaz, S. Bedogni, Modeling of sonochemistry in water in the
20635–20640, https://doi.org/10.1039/C7CP03194G. presence of dissolved carbon dioxide, Ultrason. Sonochem. 45 (2018) 17–28,
[39] R. Togel, R. Tögel, Reaction-Diffusion Kinetics of a Single Sonoluminescing Bubble, https://doi.org/10.1016/j.ultsonch.2018.02.044.
PhD Thesis, University of Twente, UK, 2002. [67] S. Gireesan, A.B. Pandit, Modeling the effect of carbon-dioxide gas on cavitation,
[40] R. Toegel, D. Lohse, Phase diagrams for sonoluminescing bubbles: A comparison Ultrason. Sonochem. 34 (2017) 721–728, https://doi.org/10.1016/j.
between experiment and theory, J. Chem. Phys. 118 (4) (2003) 1863–1875, ultsonch.2016.07.005.
https://doi.org/10.1063/1.1531610. [68] S. Merouani, O. Hamdaoui, S. Al-Zahrani, Toward understanding the mechanism of
[41] S. Sochard, A.M. Wilhelm, H. Delmas, Modelling of free radicals production in a pure CO2-quenching sonochemical processes, J. Chem. Technol. Biotechnol. 95
collapsing gas-vapour bubble, Ultrason. Sonochem. 4 (1997) 77–84, https://doi. (2020) 553–566, https://doi.org/10.1002/jctb.6227.
org/10.1016/S1350-4177(97)00021-7. [69] S. Merouani, H. Ferkous, O. Hamdaoui, Y. Rezgui, M. Guemini, A method for
[42] K. Yasui, T. Tuziuti, W. Kanematsu, Extreme conditions in a dissolving air predicting the number of active bubbles in sonochemical reactors, Ultrason.
nanobubble, Phys. Rev. E. 94 (2016) 0131061_013106–13. https://doi.org/ Sonochem. 22 (2014) 51–58, https://doi.org/10.1016/j.ultsonch.2014.07.015.
10.1103/PhysRevE.94.013106. [70] S. Merouani, O. Hamdaoui, Y. Rezgui, M. Guemini, Mechanism of the
[43] A. Dehane, S. Merouani, O. Hamdaoui, Effect of carbon tetrachloride (CCl4) sonochemical production of hydrogen, Int. J. Hydrogen Energy. 40 (11) (2015)
sonochemistry on the size of active bubbles for the production of reactive oxygen 4056–4064, https://doi.org/10.1016/j.ijhydene.2015.01.150.
and chlorine species in acoustic cavitation field, Chem. Eng. J. 426 (2021) 130251, [71] K. Yasui, Chemical reactions in a sonoluminescing bubble, J. Phys. Soc. Japan. 66
https://doi.org/10.1016/j.cej.2021.130251. (9) (1997) 2911–2920, https://doi.org/10.1143/JPSJ.66.2911.
[44] A. Dehane, S. Merouani, O. Hamdaoui, Carbon tetrachloride (CCl4) sonochemistry:
A comprehensive mechanistic and kinetics analysis elucidating how CCl4 pyrolysis

16

You might also like