Jaykumar B. Bhasarkara,
Mohit Singhb and
Vijayanand S. Moholkar*a
aDepartment of Chemical Engineering, Indian Institute of Technology Guwahati, Guwahati – 781039, Assam, India. E-mail: vmoholkar@iitg.ernet.in; Fax: +91 361 258 2291
bDepartment of Chemical Engineering, National Institute of Technology Tiruchirapalli, Tiruchirapalli – 620015, Tamil Nadu, India
First published on 19th November 2015
This paper attempts to gain physical insight into the phase transfer agent (PTA) assisted ultrasonic oxidative desulfurization process. Essentially, the synergistic links between mechanisms of PTA and ultrasound/cavitation have been identified by coupling experimental results with simulations of cavitation bubble dynamics and Arrhenius & thermodynamic analysis of reaction kinetics. It is revealed that ultrasonic oxidative desulfurization has a radical-based mechanism with a low activation energy. However, due to the high instability of radicals, the frequency factor is small leading to low dibenzothiophene (DBT) oxidation. PTA-assisted oxidative desulfurization has an ionic mechanism with a much higher activation energy. The synergistic effect of fine emulsification generated by micro-convection due to ultrasound and cavitation, and PTA-assisted interphase transport of oxidant results in almost complete oxidation of DBT. It is thus established that synergy between the mechanisms of ultrasound/cavitation and PTA is predominantly of a physical nature. Moreover, the effect of PTA is more marked for an ultrasonic system than a mechanically agitated system.
New technology for oxidative desulfurization has shown significant promise for achieving deep desulfurization. This technology is based on the conversion of non-polar aromatic hydrocarbons containing sulfur to the corresponding sulfones, which can subsequently be extracted from fuel. Oxidants used in this technique are organic or inorganic peroxyacids, which are immiscible with the fuel. Thus, the oxidative desulfurization reaction system is essentially a liquid–liquid heterogeneous system, which is limited by mass transfer. The kinetics as well as the yield of such a system depends on the interfacial area between the two phases. Ultrasound irradiation (or sonication) is a well-known technique for enhancing the interphase mass transfer in diverse physical and chemical processes through the generation of strong micro-convection.3–6 Transient cavitation also generates highly reactive radicals through the thermal dissociation of solvent vapor entrapped in the cavitation bubble at the moment of collapse. Another means of enhancing the interfacial mass transfer in a liquid–liquid heterogeneous system is the use of a phase transfer catalyst or phase transfer agent (PTA).7 A PTA enhances interphase mass transfer through the formation of a complex with the nucleophilic reagent in the aqueous phase (or the oxidant in the context of the present study), and the transport of the complex to the organic phase. Simultaneous application of ultrasound and PTA for oxidation desulfurization has been attempted by several authors.8–10 Hagenson et al.11 and Hagenson and Doraiswamy12 have analyzed the effect of ultrasound, a phase transfer agent and a microphase in the enhancement of the synthesis of benzyl sulfide from benzyl chloride and sodium sulfide. Ultrasound was revealed to enhance intrinsic mass transfer as well as the effective diffusivity of the organic reactant through the product layer. Ultrasound was revealed to boost the effects of the microphase and PTA on the enhancement of the reaction kinetics. This was attributed to enhancement of the interphase mass transfer due to strong micro-convection generated by ultrasound.
Most of the research in ultrasound-assisted oxidative desulfurization has focused on results rather than rationale. The physical mechanism of this process has remained largely unexplored. In a previous paper, we attempted to identify the physical mechanism of ultrasound assisted oxidative desulfurization by distinguishing between the individual effects of ultrasound and cavitation on the reaction system. We demonstrated that the generation of reducing species such as H2, CO and CH4 during transient cavitation in the organic phase hinders oxidation of sulfur compounds due to the competitive consumption of oxidant species.13,14 In another paper, we explored the effect of PTA on oxidative desulfurization using different peracid oxidant systems (performic acid and peracetic acid). In this study we show that the interfacial transport of oxidant in the form of an oxidant–PTA complex reduces the undesirable consumption of oxidant by reducing species, as observed in our earlier study.15 Although the paper demonstrated the beneficial effect of PTA on the ultrasonic oxidative desulfurization system, the underlying physical mechanism was not established. In the present study, we have attempted to identify the physical mechanism of the PTA-assisted ultrasonic oxidative desulfurization process. Our approach is based on the determination of the kinetic (Arrhenius) and thermodynamic parameters of oxidative desulfurization in different experimental categories, and its concurrent analysis with the simulation of cavitation bubble dynamics. For the kinetic analysis of PTA-assisted oxidative desulfurization, we have used the model reported by Zhao et al.16 The principal objectives of this study are two-fold: (1) assessment of the relative contributions of ultrasound and PTA to the enhancement of oxidative desulfurization, and (2) determination of the relative influence of interfacial mass transfer and intrinsic reaction kinetics in the process.
Experimental category | Solvent: toluene | ||
---|---|---|---|
ηb% | ka (min−1) | R2 | |
a Pseudo-1st order kinetic constant, k, obtained at a reaction temperature of 313 K for ultrasound assisted desulfurization experiments, and 333 K for experiments using mechanical stirring. Abbreviations: MS – mechanical stirring, US – ultrasound-assisted treatment (or under sonication), and ESP – elevated static pressure.b Percentage of DBT oxidation (η) calculated using: [(initial DBT concn − final DBT concn)/initial DBT concn] × 100. | |||
(A.1) MS + PFA | 28.30 ± 1.12 | 4.13 × 10−3 | 0.97 |
(A.2) US + PFA | 31.22 ± 0.98 | 4.43 × 10−3 | 0.95 |
(B.1) MS + PFA + TBAB | 63.50 ± 0.92 | 1.22 × 10−2 | 0.96 |
(B.2) US + PFA + TBAB | 96.65 ± 1.06 | 4.26 × 10−2 | 0.97 |
(C.1) US + PFA + TBAB + ESP (1.8 bar) | 77.63 ± 1.39 | 1.52 × 10−2 | 0.85 |
As stated earlier, we have determined the kinetic constants of PTA-assisted oxidative desulfurization in different experimental protocols using the model proposed by Zhao et al.16 This model is based on the model reported by Maw-Ling and Chen24 for the PTA-assisted synthesis of formaldehyde acetals from alcohols and dibromomethane in an alkaline solution of KOH and an organic solvent. This model was developed for a mechanically agitated system. We used this model for ultrasonic desulfurization in view of the predominant influence of micro-convection generated by ultrasound on the reaction system. Zhao et al.16 also proposed that complexation of the anion of the oxidant (HCOOO−) with the cation of the PTA reduces its polarity, which enables faster transport to the organic phase. Although this hypothesis is perceivable, it should be noted that the PTA–anion still has ionic character – as the net charge is not neutralized after complexation. For the convenience of the reader, we have reproduced this model below.
The oxidant in the present study, viz. performic acid, forms through the reaction between formic acid and H2O2 as follows:
HCOOH + H2O2 → HCOOOH + H2O |
Performic acid undergoes dissociation to yield the nucleophilic anionic oxidant species HCOOO− as follows:
HCOOOH ⇄ HCOOO− + H+ |
The nucleophile anion HCOOO− forms a complex with the cation of PTA (quaternary ammonium salt), represented as Q+Br−, as follows:
HCOOO− + Q+Br− → [HCOOO−–Q−–Br] | (1) |
The desulfurization reaction occurs in the organic phase in two steps: (1) formation of a PTA–oxidant complex in the aqueous phase (represented by the subscript aq), and (2) transfer of the complex across the interphase to the organic phase (represented by the subscript org). This process is reversible and is characterized by two rate constants (k1 and k−1, for the forward and backward reactions, respectively). Usually, the intrinsic reaction between PTA and the oxidant is very fast. Hence, the overall rate constants k1 and k−1 are essentially functions of the interphase mass transfer rate (which can be called the rate limiting step). In a liquid–liquid heterogeneous system, the rate of mass transfer across the interphase is a function of the interfacial area. The second step (reaction (3)) is the irreversible oxidation of the sulfur compound in the organic phase and this is characterized by rate constant k2.
![]() | (2) |
![]() | (3) |
The rate of oxidation of DBT in the organic phase (or the formation of the product DBT sulfone [C12H8SO]org) can be written as:
![]() | (4) |
In order to determine the rate of reaction in the organic phase, we need to know the concentration of the PTA–anion complex in the organic phase. The formation of the PTA–oxidant complex (Q+HCOOO−)org can be described by the following rate expression as per eqn (2) and (3) given above:
![]() | (5) |
At steady-state, the net concentration of the PTA–oxidant complex in the organic phase stays constant and this produces the condition:
![]() | (6) |
Other factors that also contribute to a constant concentration of the PTA–oxidant complex in the organic phase are fast ion exchange and mass transfer across the interphase, and a large excess of the oxidant and PTA, when compared to the concentration of the sulfur compound in the organic phase.19 It therefore follows that:
k1[Q+Br−]org[HCOOO−]aq = k−1[Q+HCOOO−]org[Br−]aq + k2[Q+HCOOO−]org[C12H8S]org | (7) |
The PTA exists in the organic phase of the reaction system in the pure form Q+Br− as well as in the form of the complex. Hence, the total concentration of the PTA cation [Q+]org, which forms a complex with the nucleophilic oxidant is written as:
[Q+]org = [Q+Br−]aq + [Q+HCOOO−]org | (8) |
Substituting for [Q+Br−]aq as [Q+]org − [Q+HCOOO−]org in eqn (7), we get:
![]() | (9) |
Then, substituting eqn (9) into eqn (4), the rate of the reaction of DBT oxidation in the organic phase is:
![]() | (10) |
Due to convection present in the system (either in the form of mechanical agitation or ultrasound irradiation), the rate of interphase mass transfer of (HCOOO−)aq can be assumed to be very fast. This results in the rapid accumulation of (HCOOO−)aq and (Q+)org in the organic phase. This favors the forward reaction of eqn (2). Moreover, the intrinsic kinetics of the DBT oxidation reaction in the organic phase is also likely to be much slower than the dissociation of performic acid. Under these conditions, the following inequality holds true:
k1[HCOOO−]aq ≫ k−1[Br−]aq + k2[C12H8S]org | (11) |
From this, the rate expression for product formation can be transformed into:
![]() | (12) |
The stoichiometry of the reaction between DBT (reactant) and DBT–sulfone (product) is essentially one. Hence, one can easily substitute the rate of formation of product in terms of the rate of disappearance of reactant (DBT):
![]() | (13) |
Therefore, eqn (12) becomes:
![]() | (14) |
Since the PTA concentration is usually in large excess, and the mass transfer rate across the interphase is also fast, the total concentration of the quaternary cation (in free and complex form) in the organic phase remains constant. Hence, we can club together the rate constant k2 and the concentration of PTC, [Q+]org, to give a new constant, k3.
![]() | (15) |
Integrating the above equation between the limits, at t = 0, [C12H8S]org = [C12H8S]org,0, and at t = t, [C12H8S]org = [C12H8S]org,t, we get:
![]() | (16) |
The magnitudes of the physical and chemical effects of cavitation bubbles have been determined using the diffusion limited ordinary differential equation model proposed by Toegel et al.25 It should be noted that cavitation phenomena occur in both the organic medium (toluene) and the aqueous medium (performic acid). However, since the DBT oxidation reaction occurs in the organic phase, we have considered the cavitation bubble dynamics phenomena in toluene only in our model. Cavitation phenomenon occurring in an aqueous medium mainly contributes to emulsification of the two phases, due to the convection induced by the radial motion of transient cavitation bubbles. The diffusion limited model reported by Toegel et al.25 is based on the comprehensive partial differential equation model reported by Storey and Szeri26 who showed that solvent vapor transport and entrapment in the cavitation bubble, leading to the formation of radicals, is essentially a diffusion limited process. This model has been extensively described in previous papers.27–29 For the convenience of readers, the essential equations and thermodynamic data of this model have been provided in the ESI (Tables S2 and S3†). The model is essentially a set of 4 ordinary differential equations: (1) the Keller–Miksis30 equation for the radial motion of the bubble, (2) an equation for the diffusive flux of water vapor and heat conduction through the bubble wall, and (3) the overall energy balance. The thermal conductivity of the bubble contents and the diffusion coefficient of the water vapor inside the bubble are determined using Chapman–Enskog theory using the Lennard–Jones 12-6 potential. An air bubble has been considered for the simulations. The condition for bubble collapse was taken as the first compression after an initial expansion. Various parameters which were used in the simulation of the bubble dynamics equation and their numerical values were as follows: ultrasound frequency (f) = 35 kHz; ultrasound (or acoustic) pressure amplitude (PA) = 150 kPa; equilibrium bubble radius (Ro) = 5 μm; the vapor pressure of toluene was calculated using an Antoine type correlation at the temperature of the reaction (298 K). Various physical properties of toluene are as follows: density (ρL) = 867 kg m−3, kinematic viscosity (ν) = 6.8 × 10−7 Pa s−1, surface tension (σ) = 0.0285 N m−1, sonic speed (c) = 1275 m s−1, and static pressure (Po) = 101.3 kPa (for experiments at atmospheric static pressure) or 162 kPa (for experiments with elevated static pressure).
The principal physical effect of cavitation is the generation of strong micro-convection in the bulk medium through two phenomena, viz. micro-turbulence, and shock or acoustic waves. The magnitudes of these two parameters can be calculated using the bubble dynamics model as follows:33–35
![]() | (17) |
![]() | (18) |
![]() | (19) |
ΔH = Ea − RT | (20) |
ΔG = ΔH − TΔS | (21) |
(1) Comparing the extent of DBT oxidation for categories A.2 and B.2, a 3.1× (or 3.1-fold) rise in oxidation is seen when PTA is used in the presence of ultrasound. On the other hand, comparing the results for categories A.1 and B.1, a relatively lower rise of 2.25× (or 2.25-fold) in oxidation is seen for PTA when applied in the mechanically stirred systems. Thus, the enhancement in DBT oxidation in the presence of PTA is more pronounced for the ultrasonic systems than the mechanically stirred systems. This result is in complete concurrence with the conclusion reported by Hagenson et al.11 that the enhancement effect of PTA on reaction kinetics acquires higher significance in the presence of ultrasound.
(2) Comparison of the results of categories B.1 and B.2 reveals that the effect of PTA is more pronounced for the ultrasound-assisted system than the mechanically agitated one. This is yet more proof of the synergism between the mechanism of ultrasound and PTA – which is corroborated by the conclusions reported by Hagenson et al.11
(3) Reduction in DBT oxidation at elevated static pressure, which is evident from the comparison of the results of categories B.2 and C.1, is attributed to suppression of the transient cavitation phenomenon at elevated static pressure, as subsequently explained in greater detail.
(1) Use of PTA in the mechanically stirred system reduces the activation energy of oxidative desulfurization, as can be observed from comparison of the activation energies in categories A.1 and B.1. This result is in concurrence with the observations of Zhao et al.16
(2) The ultrasound-assisted systems (categories A.2 and B.2) have significantly lower activation energies compared to the mechanically stirred systems (categories A.1 and B.1, respectively). Moreover, it is interesting to note that the use of PTA in the ultrasound assisted desulfurization systems leads to an increase in activation energy, as can be seen from the Arrhenius parameters for categories A.2 and B.2.
(3) The frequency factors for the ultrasound assisted systems (categories A.2 and B.2) are one to two orders of magnitude smaller than for the mechanically stirred system (categories A.1 and B.1). The use of PTA with ultrasound leads to a rise in the frequency factor, as is evident from comparison of the frequency factors for categories A.2 and B.2.
(4) Scatter of the data points in the Arrhenius plots and the low regression coefficients in Fig. S4B.2 and S4D.2† (corresponding to the systems with ultrasound) are attributed to the inverse effect of temperature on the intensity of transient cavitation. Both the physical and chemical effects of cavitation diminish with increasing temperature. The reason underlying this effect is large evaporation of solvent vapor (due to high vapor pressure) and subsequent entrapment in the bubble at elevated temperature. The entrapped vapor “cushions” the transient collapse of the cavitation bubble. This causes reduction in the intensity of the collapse (the peak temperature and pressure reached in the bubble) and also the physical/chemical effects associated with it. Consequently, the rates of the reactions that are accelerated by the physical/chemical effects of cavitation also reduce with temperature. This phenomenon violates the postulation of an increase in the rate of reaction with temperature in the Arrhenius theory. Therefore, the kinetic data of the ultrasound-assisted reactions (in the present context experimental categories A.2 and B.2) show scatter when fitted to the Arrhenius model leading to low regression coefficients as can be seen in Fig. S4B.2 and S.4D.2 in the ESI.† However, despite this limitation, the trends in the Arrhenius parameters obtained for the different experimental categories reveal a physically meaningful picture of the mechanism of PTA-assisted ultrasonic oxidative desulfurization.
(A) Kinetic and Arrhenius analysis | |||||||
---|---|---|---|---|---|---|---|
Experimental category | Parameter | Temperature (K) | Ea (kJ mol−1) | A (mol L−1 min−1) | |||
303 | 313 | 323 | 333 | ||||
a Notation: k – pseudo-1st order kinetic constant (min−1), R2 – regression coefficient, Ea – activation energy (kJ mol−1), and A – frequency factor or pre-exponential factor (mol L−1 min−1). | |||||||
MS + PFA | k (min−1) | 1.12 × 10−3 | 2.63 × 10−3 | 3.42 × 10−3 | 4.20 × 10−3 | 38.74 | 5805.14 |
R2 | 0.92 | 0.97 | 0.94 | 0.95 | |||
US + PFA | k (min−1) | 4.40 × 10−3 | 6.10 × 10−3 | 5.80 × 10−3 | 5.59 × 10−3 | 5.96 | 0.052 |
R2 | 0.95 | 0.94 | 0.96 | 0.96 | |||
MS + PFA + TBAB | k (min−1) | 6.70 × 10−3 | 9.80 × 10−3 | 1.22 × 10−2 | 2.02 × 10−2 | 29.52 | 805.93 |
R2 | 0.95 | 0.96 | 0.97 | 0.97 | |||
US + PFA + TBAB | k (min−1) | 1.80 × 10−2 | 4.26 × 10−2 | 3.67 × 10−2 | 3.59 × 10−2 | 16.43 | 15.96 |
R2 | 0.92 | 0.97 | 0.98 | 0.95 |
(B) Thermodynamic analysis | ||||||||||||
---|---|---|---|---|---|---|---|---|---|---|---|---|
Experimental category | Thermodynamic property | |||||||||||
ΔH (kJ mol−1) | −ΔS (kJ mol−1 K−1) | ΔG (kJ mol−1) | ||||||||||
Temperature | ||||||||||||
303 K | 313 K | 323 K | 333 K | 303 K | 313 K | 323 K | 333 K | 303 K | 313 K | 323 K | 333 K | |
(A.1) MS + PFA | 36.22 | 36.13 | 36.05 | 35.97 | 0.22 | 0.21 | 0.22 | 0.22 | 101.96 | 102.92 | 105.57 | 108.34 |
(A.2) US + PFA | 3.44 | 3.36 | 3.28 | 3.91 | 0.31 | 0.31 | 0.31 | 0.31 | 98.23 | 100.72 | 104.14 | 107.55 |
(B.1) MS + PFA + TBAB | 27.00 | 26.91 | 26.83 | 26.75 | 0.23 | 0.23 | 0.23 | 0.23 | 97.12 | 99.47 | 102.15 | 104.00 |
(B.2) US + PFA + TBAB | 13.91 | 13.82 | 13.74 | 13.67 | 0.27 | 0.26 | 0.27 | 0.27 | 94.68 | 95.65 | 99.23 | 102.40 |
The trends in the thermodynamic parameters that can be identified from the results presented in Table 2B are as follows: (1) the use of PTA in mechanically stirred systems leads to a reduction in ΔH and ΔG and an increase in −ΔS. (2) The use of PTA in ultrasound assisted systems shows rather the opposite trend in that ΔH increases, while −ΔS and ΔG reduce in the presence of PTA. (3) Compared to mechanically stirred systems, the ΔH and ΔG values for ultrasound-assisted systems are significantly smaller, while −ΔS values are higher.
a Data reproduced from ref. 14. | ||
---|---|---|
Species | Toluene | |
Air bubble | Air bubble | |
Ro = 5 μm | Ro = 5 μm | |
Po = 101.3 kPa | Po = 162 kPa | |
Tmax = 4835 K | Tmax = 1076 K | |
Pmax = 961 MPa | Pmax = 80.25 bar | |
Vturb = 0.15 m s−1 | Vturb = 0.014 m s−1 | |
PAW = 30.20 MPa | PAW = 0.69 bar | |
xN2 = 0.79 | xN2 = 0.79 | |
xO2 = 0.21 | xO2 = 0.21 | |
xTOL = 2.77 × 10−6 | xTOL = 4.91 × 10−7 | |
![]() |
||
Equilibrium mole fraction | ||
N2 | 7.11 × 10−1 | 7.90 × 10−1 |
O2 | 1.25 × 10−1 | 2.10 × 10−1 |
O | 2.05 × 10−2 | — |
O3 | 3.80 × 10−5 | — |
N | 1.91 × 10−4 | — |
N3 | 3.82 × 10−6 | — |
NO | 1.41 × 10−1 | 7.60 × 10−5 |
NO2 | 2.27 × 10−3 | 2.18 × 10−5 |
NO3 | 1.64 × 10−6 | — |
N2O3 | 1.14 × 10−6 | — |
N2O | 7.41 × 10−4 | — |
CO | 1.23 × 10−6 | — |
CO2 | 2.18 × 10−6 | — |
OH | 3.37 × 10−6 | — |
H2O | — | — |
HO2 | — | — |
H2O2 | — | — |
HNO2 | — | — |
Raising of the static pressure in the reaction mixture results in a drastic reduction of the physical and chemical effects of transient cavitation. Not only do the temperature and pressure peaks generated at transient bubble collapse drop sharply, but the magnitudes of shock waves (or acoustic pressure waves) and micro-turbulence generated by the bubble also show marked reduction. Moreover, no formation of any radical species is seen from the thermal dissociation of the bubble contents at the moments of transient collapse at elevated static pressure.
(1) Mechanical agitation of the reaction mixture at 300 rpm generates a limited interfacial area, which results in low (∼28.3%) DBT oxidation. The PTA cation forms a complex with the oxidant anion, which is transported across the interphase. After completion of the oxidation reaction in the organic phase, the PTA–cation returns to the aqueous phase. The chemical mechanism of PTA-assisted oxidative desulfurization with mechanical agitation is depicted in Scheme S1 in the ESI.† It could be perceived from Scheme S1† that the process has purely ionic character. The addition of PTA to this system enhances DBT oxidation by more than 2× as a result of faster and more effective transfer of oxidant across the interface. This is reflected in the reduction of the activation energy of the oxidative desulfurization, as pointed out by Zhao et al.16 For the same reason, the net enthalpy change (ΔH) for oxidative desulfurization also reduces with the use of PTA.
(2) Despite very low activation energy and intense emulsification, the extent of DBT oxidation in the ultrasound assisted system (category A.2) is almost same as that for the mechanically agitated system (category A.1). This result is attributed to a radical based mechanism of ultrasound-assisted oxidative desulfurization (as shown in Scheme S2 in the ESI†). The oxidative radicals (O˙) generated by transient bubble collapse are highly unstable and do not diffuse or penetrate much in the reaction medium from the point of bubble collapse. Hence, the probability of their interaction with DBT molecules is rather limited, which is reflected in the low value of the frequency factor (A) compared to the mechanically agitated system (category A.1), and low DBT oxidation. Nonetheless, since O˙ radicals are extremely energetic species (with an oxidation potential of 2.42 eV), the activation energy for the DBT oxidation induced by these species is quite small. For the same reasons, ΔH for category A.2 is much smaller than that for category A.1.
(3) With the simultaneous use of PTA and ultrasound (as in categories B.2 and C.1), the chemical mechanism of the oxidative desulfurization has dual character, viz. ionic and radical. The cation of PTA forms a complex with the anion of the oxidant. This reduces the polarity of the oxidant anion, which assists its effective transfer across the interface to the non-polar organic phase. It should however be noted that complexation of the oxidant–anion with PTA–cation does neutralize the polarity of the latter, and hence, the overall mechanism of oxidative desulfurization still has ionic character. On a comparative basis, the process of complex formation between PTA–cation and oxidant–anion, the transfer of this complex across the interface and the reaction between the oxidant and DBT molecules, requires higher activation energy than the radical-induced DBT oxidation within the organic phase. However, as the amount of oxidant and PTA present in the system is in large excess compared to DBT (more specifically 1.24 mM of PTA, 25 mM of PFA oxidant, and 0.54 mM of DBT), the contribution of PTA-based oxidation to the overall DBT oxidation is much higher than radical-induced oxidation. Thus, the chemical mechanism of PTA-assisted ultrasonic oxidative desulfurization is predominantly ionic, which is manifested in terms of a higher activation energy compared to ultrasonic oxidative desulfurization. An explanation for the ∼4× higher ΔH value for category B.2 compared to category A.2 can also be given along similar lines. Due to the contribution of radical-induced reactions to the overall DBT oxidation in category B.2, the activation energy for this category is ∼2× less than that for category B.1, in which PTA is applied with mechanical stirring. The role of ultrasound and cavitation in experimental categories B.2 and C.1 is more or less of a physical nature, i.e. emulsification between the organic and aqueous phases.
(4) The reduction in the intensity of transient cavitation at elevated static pressure in category C.1, as indicated by the results of the simulation of cavitation bubble dynamics, is manifested in terms of less emulsification and interfacial area, compared to category B.2. The physical and chemical effects of transient cavitation are practically eliminated at elevated static pressure. This causes significant reduction in the intensity of micro-convection in the reaction system, which in turn, adversely affects emulsification and interfacial area. In this case, the ultrasonic system resembles a mechanically stirred system (as in category B.1). This essentially results in lower DBT oxidation in category C.1. However, the micro-streaming induced by ultrasound (which remains unaffected by the elevated static pressure) generates finer emulsion between the phases, compared to (macroscopic) mechanical stirring. The addition of PTA to the reaction system further boosts the extent of DBT oxidation. The beneficial effect of these two factors is manifested in significantly higher (>2×) DBT oxidation, compared to categories A.1 and A.2.
(5) The entropy change values (−ΔS) in the different experimental categories are again manifestations of the predominant chemical mechanism of oxidative desulfurization. The highest −ΔS values are seen for category A.2, in which the chemical mechanism for oxidative desulfurization is purely radical-based and reactions occur almost instantly. The −ΔS values for mechanically stirred systems increase with the addition of PTA, indicating faster reactions. The −ΔS values for category B.2 are intermediate between those of categories B.1 and A.2, indicating the dual (ionic and radical induced) nature of the chemical mechanism of the PTA-assisted oxidative desulfurization. The dual nature of the chemical mechanism in category B.2 also results in pseudo-1st order kinetics constants of intermediate value between the constants for categories A.2 and B.1.
(6) Despite significantly dissimilar chemical mechanisms that lead to dissimilar values of ΔH and −ΔS, the Gibbs energy change (ΔG) is similar (±5%) for experimental categories B.1 and B.2, in which PTA was simultaneously applied with ultrasound. This essentially is an indication of the physical role played by ultrasound in PTA assisted oxidative desulfurization. Moreover, lower ΔG values for categories B.1 and B.2, compared to categories A.1 and A.2, respectively, point to a supportive effect of the PTA in the enhancement of the oxidative desulfurization through effective interphase transport of the oxidant anion.
Footnote |
† Electronic supplementary information (ESI) available: The following supplementary materials have been provided with this manuscript: (1) schematic diagram of the experimental setup, (2) preliminary experiments for determination of the composition of the reaction mixture, (3) model equations and thermodynamic data for cavitation bubble dynamics, (4) experimental categories with the exact composition of the reaction mixture, (5) kinetic analysis of oxidative desulfurization under ultrasound treatment and mechanical stirring using performic acid as oxidant coupled with TBAB as a phase transfer agent, (6) simulations of radial motion of a 5 micron air bubble in toluene at elevated static pressure, (7) simulations of radial motion of a 5 micron air bubble in toluene at atmospheric static pressure, (8) reaction mechanism for only ultrasound assisted oxidative desulfurization and PTA-assisted ultrasonic oxidative desulfurization. See DOI: 10.1039/c5ra12178g |
This journal is © The Royal Society of Chemistry 2015 |