ARTICLE IN PRESS Chemical Engineering Science

(

)

--

Contents lists available at ScienceDirect

Chemical Engineering Science journal homepage: w w w . e l s e v i e r . c o m / l o c a t e / c e s

A new mechanistic model for liquid–liquid phase transfer catalysis: Reaction of benzyl chloride with aqueous ammonium sulfide Sunil K. Maity a , Sujit Sen b , Narayan C. Pradhan b, ∗ a b

Department of Chemical Engineering, National Institute of Technology Rourkela, Rourkela-769 008, Orissa, India Department of Chemical Engineering, Indian Institute of Technology Kharagpur, Kharagpur-721 302, West Bengal, India

A R T I C L E

I N F O

Article history: Received 10 December 2008 Received in revised form 3 July 2009 Accepted 6 July 2009 Keywords: Phase transfer catalysis Dynamic simulation Reaction engineering Multiphase reactions Dibenzyl sulfide Hydrogen sulfide

A B S T R A C T

A new mechanistic model for reactions involving two liquid phases and a homogeneous phase transfer catalyst (PTC) has been developed based on extraction mechanism considering the equilibrium of catalyst and active catalysts at the interface. The proposed kinetic model considers thermodynamic framework based aqueous phase ionic equilibrium and the separate contributions of non-PTC and PTC-enhanced reactions towards the overall reactions. The developed model was then applied to an industrially important reaction of benzyl chloride with aqueous ammonium sulfide for synthesis of dibenzyl sulfide and benzyl mercaptan. The kinetic parameters of the developed model were estimated at different temperatures using an indigenously developed non-linear regression technique based on modified Levenberg–Marquardt algorithm. Sensitivity analysis was then performed under various experimental conditions using the estimated parameters and the results were compared with experimental observations. A good agreement was observed between experimental and calculated values with proper trends of the results. © 2009 Elsevier Ltd. All rights reserved.

1. Introduction Phase transfer catalysis (PTC) is an attractive technique for synthesis of organic chemicals from two reactants existing in two different phases, which otherwise normally cannot react with each other due to their low mutual solubility in other phase and hence their low interaction (Dehmlow and Dehmlow, 1993; Starks et al., 1994). Now-a-days, this technique is widely practiced for large-scale production in industry as well because of its additional advantages of milder operating conditions, reduced consumption of solvents and raw materials, and enhanced selectivity over the conventional processes. The PTC technique has been applied successfully for a large number of reactions belonging to various categories such as alkylation, arylation, condensation, elimination, polymerization, etc. The liquid–liquid phase transfer catalysis (L-LPTC) is a process of favoring reactions between a lipophilic substrate dissolved in an organic phase with a hydrophilic reactant solubilized in water. The PTC facilitates the transfer of nucleophile of a hydrophilic reactant (via formation of a lipophilic ion-pair) from aqueous phase to organic phase, where the reaction with the lipophilic substrate takes place (Starks et al., 1994; Dehmlow and Dehmlow, 1983;

∗ Corresponding author. Tel.: +91 3222 283940; fax: +91 3222 255303. E-mail address: [email protected] (N.C. Pradhan).

Starks and Owens, 1973). This mechanism as suggested by Starks is known as extraction mechanism (Scheme 1a) and is applicable when the PTC is not highly lipophilic one and hence can distribute itself between the phases. When the PTC is a highly lipophilic one, interfacial mechanism (Scheme 1b) is generally used to explain the L-LPTC (Starks et al., 1994). According to this mechanism, the PTC remains entirely in the organic phase and exchanges nucleophile with hydrophilic reactant through aqueous-organic interface only. Extensive reviews have been published in the area of L-LPTC in the form of monographs and papers in peer reviewed journals (Dehmlow and Dehmlow, 1993; Starks et al., 1994; Sasson and Neumann, 1997; Halpern, 1997). Although much effort has been devoted to the kinetic study and synthesis of organic chemicals, very little attention has been paid in the past to the development of kinetic models for L-LPTC reactions of commercial importance. This is due to the fact that the modeling of the dynamic behavior of L-LPTC is quite complex as it involves the consideration of reactions in both aqueous and organic phases and transfer of catalyst and active catalysts between the phases. Moreover, the physical parameters of the catalyst (or active catalysts) such as distribution coefficients, mass transfer coefficients, and dissociation equilibrium constants are difficult to obtain experimentally as organic-soluble PTC (or active catalyst) cannot be isolated or identified during the reaction in most of the cases (Wang and Yang, 1991a; Wang and Chang, 1994).

0009-2509/$ - see front matter © 2009 Elsevier Ltd. All rights reserved.

Please cite this article as: S.Maity, K., et al., A new mechanistic model for liquid–liquid phase transfer catalysis: Reaction of benzyl chloride with aqueous ammonium sulfide. Chemical Engineering Science (2009), doi:10.1016/j.ces.2009.07.004

ARTICLE IN PRESS 2

S.K. Maity et al. / Chemical Engineering Science

RX + Q+Y-

RY + Q+X-

Organic phase

M+X - + Q+Y-

M+Y - + Q+X-

Aqueous phase

RX + Q+Y-

RY + Q+X-

Organic phase

M+X - + Q+Y-

M+Y - + Q+X-

M+X -

M+Y -

Interface

Aqueous phase

Scheme 1. (a) Extraction mechanism and (b) interfacial mechanism.

The consideration of mass transfer factor in the kinetic modeling is important when the phase transfer catalyzed reactions in the organic phase are very fast compared to the mass transfer of the active catalysts. Examples of mass transfer controlled reactions are however limited in literature (Lele et al., 1983) and in most of the cases the organic phase reactions are found to be rate-determining under intense agitation of the reaction mixture. The kinetic models developed in the past were mostly based on the two-film theory of mass transfer (of the catalysts and active catalysts) between two phases and the kinetics of the organic phase reactions (Wang and Wu, 1991; Wang and Yang, 1991a). Wang and Yang (1991b) later on analyzed the reaction system using a simplified model based on the pseudo-steady-state hypothesis. It is revealed from the analysis that the mass-transfer resistance of active catalysts from the aqueous phase to the organic phase needs no consideration in the kinetic model as the contribution is negligible compared to the chemical reaction rate either in the organic phase or in the aqueous phase. Wang and Wu (1991) assumed that the concentration of the active catalyst in the organic phase remains constant during the course of reaction and the organic phase reactions are represented by pseudo-first-order reaction kinetics. Recently, the modeling of a variety of liquid–liquid phase transfer catalyzed nucleophilic substitution reactions are reported based on this assumption (Wang and Tseng, 2003, 2002a, 2002b; Yadav and Jadhav, 2003; Yang and Lin, 2003). Wu (1993) analyzed the dynamic behavior of a liquid–liquid phase transfer catalyzed reactions system using a phase-plane model based on the two-film theory. The analysis reveals that the assumption of pseudo-first-order kinetics in the organic phase does not hold when mass transfer rates are comparable with the chemical reaction rates. Wu (1996) reworked on the pseudo-steady-state hypothesis to avoid the assumption of the pseudo-first-order reaction kinetics. It was demonstrated that the concentration of active catalyst in the organic phase decreased with the progress of the reaction. The model also suitably explained the concentration dependency of the apparent-first-order reaction rate coefficient (kapp = ko [QY]o ) as observed by previous researchers (Wang and Wu, 1991). In the past, the common approach was that the non-PTC reactions could either be ignored or incorporated in the PTC-enhanced reactions. With this approach, the correlation capability of the model will still remain unaffected as long as non-catalytic contributions relative to catalytic one remains constant. However, the conditions of reaction generally vary significantly with the progress of the reaction and the above assumption will therefore not remain valid throughout the course of the reaction. Separate consideration of these two contributions is therefore essential in order to develop a physically meaningful mechanistic model with a quite good parametric sensitivity. Recently, a kinetic model was developed considering separate contributions of both non-PTC and PTC-enhanced reactions and the effect of changes in electrolyte composition in the aqueous phase and the developed model was then validated for the industrially

(

)

--

important reaction of synthesis of benzaldehyde from benzyl chloride (BC) (Satrio and Doraiswamy, 2002). The liquid–liquid phase transfer catalyzed reaction of BC with aqueous ammonium sulfide has enormous potential to utilize the hydrogen sulfide (H2 S) of various by-product gas streams and to produce value-added products, dibenzyl sulfide (DBS) and benzyl mercaptan (BM) (Maity et al., 2006). The aqueous ammonium sulfide based process has enormous commercial importance especially in the petroleum refineries dealing with large quantities of heavy crudes containing high amount of sulfur and nitrogen. During the process of reduction of the sulfur and nitrogen of the heavy crudes, large quantities of by-product gas streams containing H2 S and ammonia (NH3 ) are commonly produced. Conventionally, the gas streams containing these gases are first scrubbed with either water or dilute sulfuric acid to remove NH3 and then sent through amine treating unit to remove H2 S, which is further processed in the Claus unit to produce elemental sulfur (Kohl and Nielsen, 1997). One potential alternative of the conventional process is the simultaneous removal of both H2 S and NH3 from the gas streams in a single step using dilute ammonium hydroxide to produce aqueous ammonium sulfide, which is then utilized to produce commercially important chemicals like DBS and BM. Moreover, the costly regeneration of H2 S-rich solution is not required in the proposed alternative process. Considering enormous importance of the reaction system, the work was undertaken to develop a dynamic model following the actual reaction mechanism. The developed model considers the equilibrium of catalyst and active catalysts between the phases and the separation of non-PTC and PTC-enhanced contributions towards the overall reactions. Moreover, influence of complex aqueous phase ionic equilibrium on the reaction was considered for the first time in the developed kinetic model based on a thermodynamic framework. The reaction of BC with ammonium sulfide provides an ideal case study for development of kinetic model for multi-step L-LPTC systems as it involves more than one inorganic nucleophiles (HS− and S2− ) resulting from complex ionic equilibrium of NH3 –H2 S–H2 O in the aqueous phase, series–parallel reactions network in the organic phase, and significant non-PTC contribution (Scheme 2). The availability of a validated kinetic model for such kind of industrially important reaction with highly complex reaction mechanism is quite desirable one and is essential for the development, design, and scale-up of reactors. 2. Experimental The reactions of BC with aqueous ammonium sulfide were carried out batch wise in a 6.5 cm i.d. fully baffled mechanically agitated glass reactor (capacity 250 cm3 ). A 2.0 cm diameter six-bladed glass disk turbine impeller with the provision for stirring speed regulation located at a height of 1.5 cm from the reactor bottom was used for stirring the reaction mixture. The reactor was kept in a constant temperature water bath whose temperature could be controlled within ± 1 ◦ C. The H2 S absorption experiments were carried out in a 250 cm3 standard gas-bubbler using ammonium hydroxide solution of suitable ammonia concentration. In a typical experimental run, 50 cm3 of the aqueous ammonium sulfide containing a known concentration of sulfide and catalyst, tetra-n-butylammonium bromide (TBAB) was charged into the reactor and kept well stirred until steady-state temperature was reached. Then the organic phase containing measured volume of BC and solvent (toluene) kept separately at the reaction temperature was charged into the reactor. Samples were withdrawn from the organic layer at regular intervals after stopping the stirring and allowing the two phases to separate. All the samples from the organic phase were analyzed by gas–liquid chromatography using an FID detector and nitrogen as carrier gas. The initial sulfide

Please cite this article as: S.Maity, K., et al., A new mechanistic model for liquid–liquid phase transfer catalysis: Reaction of benzyl chloride with aqueous ammonium sulfide. Chemical Engineering Science (2009), doi:10.1016/j.ces.2009.07.004

ARTICLE IN PRESS S.K. Maity et al. / Chemical Engineering Science NH4+ + OH-

(I)

H2S

H+ +HS-

(II)

HS-

H+ + S2-

(III)

H2O

H+ + OH-

(IV)

NH4OH

Aqueous phase

NH4OH+HCl

+

Q + HS-

i

QBr/ QCl

Cl

Q

+

-

+ Br /Cl

-

QSQ (VIII)

(VI)

SNH4

Cl

NH4Cl +

+ NH4Cl

Cl

k3nc

+

S

+ HCl

mQSH

Cl

+QSH

k1

S

NH4Cl +

k'2

mQBr

SH

SQ

SNH4

mQSQ

Cl

c

+QCl

QCl+

Cl

+

S

+ HCl

S

QCl+

b

+QSQ

k2c

Cl

k3c

iii

+

nc

Organic phase SH

ii +(NH4)2S

k2nc

SH

a

d

3

k1nc

+NH4SH

iv

--

2Q++ S2-

SH

Cl

Interface

)

(V)

NH4Cl+H2O

(VII)

QSH

(

k'2c

SQ

c

+

concentration was determined by the standard iodometric titration method. The experimental details were reported in our earlier publication (Maity et al., 2006).

3. Modeling 3.1. Role of mass transfer resistance For discerning the true kinetics, the mass transfer resistance should be negligible during the reaction. To determine the role of mass transfer resistance, the effect of stirring speed on the rate of reaction of BC was studied in the range of 1000–3000 rev/min under otherwise identical experimental conditions in presence and absence of catalyst as shown in Fig. 1 . The variation of reaction rate of BC with the stirring speed was found to be negligibly small for both in the absence and presence of catalyst as observed from the figure. The reaction may, therefore, be considered as free from mass transfer resistance and controlled by kinetics only. Accordingly, effect of mass transfer was not considered in the developed model for this reaction. Moreover, the aqueous phase reactions and extraction of catalysts and active catalysts at the interface are assumed to remain in equilibrium. The overall rate of reaction in presence of a catalyst has two contributions: non-catalytic and catalytic as given by following equations:    d[RX]  d[RX]  d[RX]  = + dt overall dt non-catalytic dt catalytic

(1)

   d[RSH]  d[RSH]  d[RSH]  = + dt overall dt non-catalytic dt catalytic

(2)

   d[RSR]  d[RSR]  d[RSR]  = + dt overall dt non-catalytic dt catalytic

(3)

Reaction rate of benzyl chloride × 104, kmol/m3s

Scheme 2.

18 non-catalytic catalytic (TBAB = 9.3×10-2 kmol/m3 of org. phase)

16

14

12 6

4

2 1000

1500 2000 2500 Speed of agitation, rpm

3000

Fig. 1. Effect of speed of agitation on rate of reaction of benzyl chloride. Volume of organic phase = 5.0×10−5 m3 ; concentration of benzyl chloride = 2.6 kmol/m3 ; volume of aqueous phase = 5.0×10−5 m3 ; concentration of sulfide = 1.20 kmol/m3 ; NH3 concentration = 6.05 kmol/m3 ; temperature = 333 K; matching reaction time = 5 min.

3.2. Non-catalytic contribution Inorganic reagents are generally insoluble in the organic phase. It is therefore reasonable to assume that the non-catalytic reactions proceed at the aqueous-organic interface. Based on the product distribution, the mechanism of the non-catalytic reactions was represented by the reactions (i)–(iv) of Scheme 2. The ammonium hydrosulfide (NH4 SH) and sulfide ((NH4 )2 S) in the aqueous phase non-catalytically react with organic phase benzyl chloride (RX) at the

Please cite this article as: S.Maity, K., et al., A new mechanistic model for liquid–liquid phase transfer catalysis: Reaction of benzyl chloride with aqueous ammonium sulfide. Chemical Engineering Science (2009), doi:10.1016/j.ces.2009.07.004

ARTICLE IN PRESS 4

S.K. Maity et al. / Chemical Engineering Science

interface to yield benzyl mercaptan (RSH) and dibenzyl sulfide (RSR) respectively as represented by the reactions (i)–(iii) of Scheme 2. The nucleophilic substitution reactions are in general second order overall. The direct reaction of BC with (NH4 )2 S was therefore assumed to proceed through an intermediate, RSNH4 , which further reacts with RX to produce RSR. Since RSNH4 was not identified in GLC analysis, the net rate of its formation was approximated as zero based on pseudo-steady-state hypothesis. The RSH reacts further with RX to produce RSR (reaction (iv) of Scheme 2). This conclusion was reached from the concentration profile of a typical batch (Maity et al., 2006). The net rate of formation of RX, RSH, and RSR are represented by the following equations:  d[RX]  nc = − knc 1 [RX][NH4 SH] − 2k2 [RX][(NH4 )2 S] dt non-catalytic (4) − knc 3 [RX][RSH]  d[RSH]  nc = knc (5) 1 [RX][NH4 SH] − k3 [RX][RSH] dt non-catalytic  d[RSR]  nc = knc (6) 2 [RX][(NH4 )2 S] + k3 [RX][RSH] dt non-catalytic

(

)

--

where f=

volume of organic phase, VO volume of aqueous phase, Va

The total initial amount of PTC added in the reaction mixture, Q0, expressed as kmol/m3 of organic phase, is distributed between the aqueous and organic phases. The PTC in the aqueous phase remains in the form of ionic, molecular, and active catalysts, whereas it is present in molecular and active catalysts forms in the organic phase. 1 Q0 =[QX]o +[QSH]o +2[QSQ]o + ([Q + ]+[QX]a +[QSH]a +2[QSQ]a ) (14) f Combining Eqs. (10)–(12) with Eq. (14), the following expression is obtained:     1 1 Q0 = [QX]a mQX + + [QSH]a mQSH + f f   1 + 1 + [Q ] (15) + 2[QSQ]a mQSQ + f f The concentration of X− ions generated from the ionization of the PTC (reaction VI of Scheme 2) in the aqueous phase is given by following equation: [X − ] = [Q + ] + [QSH]a + 2[QSQ]a + f ([QSH]o + [QSQ]o )

3.3. Catalytic contribution In presence of a PTC, the reactions proceed through both noncatalytic and catalytic pathways and so contributions of both the reaction to overall rate need to be considered. The contribution of non-catalytic reactions will be governed by Eqs. (4)–(6). The rate expression of the catalytic contribution is however quite complex and depends on the type of mechanism involved, which in turn depends on the lipophilicity of the PTC used in the reaction. The reaction of BC with aqueous ammonium sulfide was studied in presence of a PTC, TBAB, which is not a highly lipophilic one and can distribute itself between aqueous and organic phases. Therefore, the catalytic reactions could be best represented by extraction mechanism of the L-LPTC as shown by the reactions a–d of Scheme 2. The dissociation equilibrium constants of the PTC, QX and active catalysts, QSH and QSQ, in the aqueous phase are defined as follows (reactions VI–VIII of Scheme 2):

= [Q + ] + [QSH]a (1 + mQSH f )

+ 2[QSQ]a (1 + mQSQ f ) (Using Eq. 11 & 12) [Q+ ],

(16)

[X− ],

KVI =

[Q + ][X − ] [QX]a

(7)

The concentration of [QX]a , [QSH]a , and [QSQ]a are obtained by solving the set of five non-linear equations, Eqs. (7)–(9), (15) and (16) provided the sulfide and hydrosulfide ions concentration in the aqueous phase are known. The calculation steps followed are (1) assume a value of [Q+ ], (2) calculate [QSH]a and [QSQ]a using Eqs. (8 and 9), (3) calculate [X− ] using Eq. (16), (4) calculate [QX]a using Eq. (7), (5) calculate Q0 using Eq. (15), (6) if the Q0 calculated from Eq. (15) is not equal to the Q0 added in the reaction mixture, repeat the steps from (1) to (5). (7) Calculate the concentration of [QSH]o , and [QSQ]o using Eqs. (11) and (12). The sulfide and hydrosulfide ions concentration in the aqueous phase are calculated based on a thermodynamic framework developed by Edwards et al. (1975) as described in the subsequent section. The reaction of BC with QSQ was assumed to proceed through an intermediate of RSQ. The rate of formation of RSQ is given by

KVII =

[Q + ][HS− ] [QSH]a

(8)

d[RSQ] = kc2 [RX][QSQ]o − k2 c[RX][RSQ] dt

(9)

The RSQ was not identified in GLC analysis. The concentration of RSQ was therefore expressed assuming pseudo-steady-state hypothesis as given by the following equation.

[Q + ]2 [S2− ] KVIII = [QSQ]a

The equilibrium distribution coefficients of the PTC and active catalysts are defined as follows: mQX =

[QX]o [QX]a

(10)

mQSH =

[QSH]o [QSH]a

(11)

mQSQ =

[QSQ]o [QSQ]a

(12)

The total sulfide concentration (TS ) expressed as kmol/m3 of aqueous phase at any reaction time is the sum of the sulfides present in the aqueous phase in the form of molecular and ionic sulfur species and active catalysts and organic phase in the form of active catalysts as given by the following equation: TS = [H2 S]+[HS− ]+[S2− ]+[QSH]a +[QSQ]a +f ([QSH]o +[QSQ]o )

(13)

[RSQ] =

kc2 [QSQ]o k2 c

(17)

(18)

The rates of formation of benzyl chloride (RX), benzyl mercaptan (RSH), and dibenzyl sulfide (RSR) by catalytic pathway are expressed by the following equations. The concentration of RSQ as appeared in the rate expression was replaced by using Eq. (18).  d[RX]  = −kc1 [RX][QSH]o − 2kc2 [RX][QSQ]o − kc3 [RX][RSH] (19) dt catalytic  d[RSH]  = kc1 [RX][QSH]o − kc3 [RX][RSH] (20) dt catalytic  d[RSR]  = kc2 [RX][(NH4 )2 S] + kc3 [RX][RSH] (21) dx catalytic The total sulfide and ammonia concentrations in the aqueous phase decrease as the reaction proceeds. Therefore, the total sulfide

Please cite this article as: S.Maity, K., et al., A new mechanistic model for liquid–liquid phase transfer catalysis: Reaction of benzyl chloride with aqueous ammonium sulfide. Chemical Engineering Science (2009), doi:10.1016/j.ces.2009.07.004

ARTICLE IN PRESS S.K. Maity et al. / Chemical Engineering Science

(

)

--

5

 +  2− aH+ aS2− mH+ mS2− m  = × H S = KIII KIII aHS− mHS− HS−

(TS ) and ammonia (TNH3 ) concentrations in the aqueous phase at any reaction time were obtained from overall mass balance as given by the following equations:

KIII =

TS = TS0 − f ([RSH] + [RSR])

(22)

TNH3 = TNH3 0 − f ([RX]0 − [RX])

(23)

where `ai ' (ai = mi i ) is the activity of the species i. The dissociation equilibrium constants of reactions I–IV at a given temperature are calculated using the correlation reported by Beutier and Renon (1978). As observed from Eqs. (E1)–(E4), the calculation of concentrations of seven species in the aqueous phase involves fourteen variables (concentration and activity coefficients of each of the seven species) and fourteen independent equations are therefore needed to solve these equations. In addition to above four equations, three more balance (mass and charge) equations can be written as given by the following equations: Mass balance for NH3 :



m KIV = aH+ aOH− = mH+ mOH− × H+ OH− = KIV KIV

where [RX]0 is the initial concentration of BC. 4. Modeling of aqueous phase ionic equilibrium Edwards et al. (1975) developed a thermodynamic framework for calculation of vapor–liquid equilibrium for aqueous solutions containing one or more volatile weak electrolytes: NH3 , CO2 , H2 S, SO2 , and HCN and successfully applied the developed model for two ternary systems: NH3 –H2 S–H2 O and NH3 –CO2 –H2 O. In the present study, this framework was used to represent the ionic equilibrium of NH3 –H2 S–H2 O (reactions I–IV of Scheme 2) in the aqueous phase. The ionic equilibrium of NH3 –H2 S–H2 O results seven (molecular and ionic) species (in addition to water) in the aqueous phase, namely NH3, H2 S, NH4 + , H+ , HS− , S2− , and OH− as represented by Scheme 2. The ionic equilibria of reactions I–IV in Scheme 2 can be represented by the dissociation equilibrium constants (K) in terms of molar concentration (m) and activity coefficients () of the species as represented by Eqs. (E1)–(E4). 4

aNH3

=

mNH+ mOH− 4

mNH3

×

NH4+ OH−  = KIm KI NH3

 + − aH+ aHS− mH+ mHS−  = × H HS = KIIm KII aH2 S mH2 S H2 S

Hydrosulfide ions, mol/lit

KII =

aNH+ aOH−

1.2 1 0.8 0.6 0.4 0.2 0 7 Am 6 5 mo nia 4 ,m ol/l it

3

0.2

0.4

1

0.8

0.6

lfide,

su Total

(E4)

(E5)

4

Mass balance for sulfide: TS = mH2 S + mHS− + mS2−

(E6)

Electroneutrality/charge balance: mNH+ + mH+ = mHS− + 2mS2− + mOH−

(E7)

4

(E1)

The activity coefficients (i ) for each of seven species were expressed by the following equations (Edwards et al., 1975):

(E2)

ln i =

sulfide ions concentration, mol/lit

KI =

TNH3 = mNH3 + mNH+

(E3)

√ 7  −Zi2 I ik mk √ +2 1+ I k=1

(E8–E14)

x 10-3 4 3.5 3 2.5 2 1.5 1 0.5 0

A mm

t

mol/li

7 6 5 onia c onc ent ra 4 tion,

3 mol/ lit

1.0 0.8 ol/ lit m , n tio 0.2 c entra n o c e s ulfid Total 0.4

0.6

Molecular hydrogen sulfide, mol/lit

x 10-3 6 5 4 3 2 1 0 7 6 5 mo nia ,m o

Am

4

l/lit

3

0.2

0.4

0.6

0.8

1

ol/lit

lfide, m

u Total s

Fig. 2. Concentration of different sulfur species in the aqueous phase. Simulation results are at 333 K.

Please cite this article as: S.Maity, K., et al., A new mechanistic model for liquid–liquid phase transfer catalysis: Reaction of benzyl chloride with aqueous ammonium sulfide. Chemical Engineering Science (2009), doi:10.1016/j.ces.2009.07.004

ARTICLE IN PRESS 6

S.K. Maity et al. / Chemical Engineering Science

¨ where  is the Debye–Huckel proportionality factor defined by ¨  = 2.303A , where A is the Debye–Huckel constant for activity coefficient = 0.511 (kg/mol) 1/2 at 25 ◦ C (Bromley, 1973), zi is the ionic charge of species i, I is the ionic strength of the solution defined  by I = 12 7i=1 Zi2 mi , and ik is the specific interaction parameter for species i and k and obtained from the works of Edwards et al. (1975). These 14 equations (Eqs. (E1)–(E14)) constitute the model of Edwards et al. (1975) for calculation of concentrations and activity coefficients of each of the seven species in the aqueous phase. The concentrations of different sulfur species in the aqueous phase at 333 K were calculated from this model for varying total ammonia and sulfide concentration in the aqueous phase as shown in Fig. 2. It is clear from this figure that hydrosulfide ions are the dominating sulfur species and its concentration increases with total sulfide concentration in the aqueous phase (Fig. 2a). It is also observed from the figure that the concentration of hydrosulfide ions is in the same order of magnitude as that of total sulfide. The concentration of sulfide ions (Fig. 2b), on the other hand, was only significant at high total ammonia concentration and the concentration of molecular H2 S (Fig. 2c) was found to be significant at high total sulfide concentration and low total ammonia concentration. From this result, it may be concluded that at high ammonia concentration the dissociation equilibria of reactions I–IV of Scheme 2 shift towards more ionization leading to low molecular H2 S concentration and high sulfide ion concentration in the aqueous phase. 5. Regression analysis and parameters estimation It is worthy to mention here that the reaction of BC with BM apparently seems to be non-catalytic one. If the above reaction was assumed to proceed non-catalytically only, even in presence of a PTC, the appropriate trend of conversion of BC and selectivity of DBS was not observed during the regression analysis. This observation clearly indicates that the above reaction proceeds catalytically as well. However, exact catalytic mechanism for this reaction is still not clear. The required model parameters, distribution coefficient and dissociation constants of TBAB were taken from the work of Wu et al. (1998) measured in chlorobenzene solvent at different temperatures in the range of 283–323 K as it closely resembles the present reaction system. However, the values of these constants for the active catalysts, QSH and QSQ, could not be obtained experimentally as these catalysts could not be isolated during the reaction. Therefore, the values of these parameters for the active catalysts were assumed to be the same as that of the catalyst. 5.1. Non-catalytic contribution The kinetic constants for non-catalytic contribution were estimated using the experimental data of independent experiments in absence of any catalyst at four different temperatures in the range of 303–333 K. The non-catalytic model consisting of three ordinary differential equations, Eqs. (4)–(6), algebraic equations, Eqs. (22) and (23), and model of aqueous phase ionic equilibrium, Eqs. (E1)–(E14), were regressed using an indigenously developed non-linear regression algorithm based on modified Levenberg–Marquardt algorithm. The optimum values of the parameters were estimated by minimizing the objective function as given by the following equation: E=

n  [{([RX]exp t )i − ([RX]pred )i }2 + {([RSH]exp t )i i=1

− ([RSH]pred )i }2 + {([RSR]exp t )i − ([RSR]pred )i }2 ]

(24)

The optimized non-catalytic rate constants at different temperatures were reported in Table 1. It is observed from the table that with

(

)

--

Table 1 Variation of rate constants with temperature in the absence of catalyst. Temperature (K)

3 −1 −1 knc 1 (kmol/m ) s

3 −1 −1 knc 2 (kmol/m ) s

3 −1 −1 knc 3 (kmol/m ) s

303 313 323 333

3.30×10−05 9.63×10−05 2.45×10−04 3.17×10−04

1.61×10−04 5.63×10−04 6.03×10−04 1.26×10−03

4.88×10−06 7.89×10−06 1.60×10−05 2.27×10−05

Table 2 Frequency factor and activation energy of various rate constants in the absence of catalyst. Rate constants

Frequency factor, k0 ((kmol/m3 )−1 s−1 )

Activation energy, E (kJ/mol)

Correlation factor, R2

knc 1 knc 2 nc k3 Experimental

6.53×1006 2.21×1005 2.38×1002 −

65.20 52.54 44.66 62.17

0.95 0.90 0.99 1.0

increase in temperature, the rate constants increase as expected. The Arrhenius plots of ln(k) against 1/T were made and the activation energies and pre-exponential factors were obtained from the slopes and intercepts of the Arrhenius plots as shown in Table 2. The initial rate of reaction at different temperatures were calculated from experimental data and the experimental activation energy (apparent) was calculated from the slope of the Arrhenius plot of ln(initial rate) versus 1/T as shown in Table 2. As it is observed from the table, the activation energy calculated from the estimated rate constants was comparable with the experimental activation energy. The conversion of BC and selectivity of DBS at various reaction times were calculated using the estimated rate constants at different temperatures and compared with those of experimental values as shown in Fig. 3. A good agreement was observed with average absolute deviation of about 5.1%, 4.2%, 4.3%, and 5.2% at 303, 313, 323, and 333 K, respectively. 5.2. Catalytic contribution In order to estimate the kinetic constants corresponding to the catalytic contribution, experiments were conducted in presence of catalyst, TBAB, under otherwise identical experimental conditions (as that of non-catalytic reactions) as shown in Fig. 4. The entire kinetic model, with both catalytic and non-catalytic contributions, was regressed using previously estimated non-catalytic kinetic constants. The optimized catalytic kinetic constants at different temperatures are shown in Table 3. It is evident from Tables 1 and 3 that the estimated rate constants corresponding to catalytic contribution are higher compared to the rate constants corresponding to non-catalytic contribution. This indicates the importance of PTC in enhancing the rate of reaction. The activation energy and pre-exponential factor of the rate constants were obtained from the slope and intercept of the Arrhenius plot of ln(k) against 1/T as shown in Table 4. The conversion of BC and selectivity of DBS at various reaction times were calculated using the estimated rate constants at different temperatures and were compared with those of experimental values as shown in Fig. 4. A good agreement was observed in this case as well with average absolute deviation of about 5.8%, 4.7%, 4.4%, and 4.1% at 303, 313, 323, and 333 K, respectively. The experimental activation energy in presence of catalyst has been found to be smaller than that calculated from the estimated rate constants. This is due to the fact that the experimental activation energy in presence of catalyst reflects contributions of both catalytic and non-catalytic reactions. The activation energies calculated from the rate constants of catalytic reactions are found to be significantly

Please cite this article as: S.Maity, K., et al., A new mechanistic model for liquid–liquid phase transfer catalysis: Reaction of benzyl chloride with aqueous ammonium sulfide. Chemical Engineering Science (2009), doi:10.1016/j.ces.2009.07.004

ARTICLE IN PRESS

70

70

60

60

50 40 30 Expt. Model 303 K 313 K 323 K 333 K

20 10

% Selectivity of DBS

% Conversion of benzyl chloride

S.K. Maity et al. / Chemical Engineering Science

(

)

--

7

Expt. Model 303 K 313 K 323 K 333 K

50 40 30 20 10 0

0 0

0

50 100 150 200 250 300 350 400 450 500

50 100 150 200 250 300 350 400 450 500

Reaction time, min

Reaction time, min

100

100

90

90

80

80

70 60 50 40 Expt. Model 303 K 313 K 323 K 333 K

30 20 10 0

0

50 100 150 200 250 300 350 400 450 500 Reaction time, min

%Selectivity of DBS

% Conversion of benzyl chloride

Fig. 3. Comparison of calculated and experimental results in the absence of catalyst. Volume of organic phase = 5.0×10−5 m3 ; concentration of benzyl chloride = 2.6 kmol/m3 ; volume of aqueous phase = 5.0×10−5 m3 ; concentration of sulfide = 1.20 kmol/m3 ; NH3 concentration = 6.05 kmol/m3 ; stirring speed = 1500 rev/min.

70 60 50 40 Expt. Model 303 K 313 K 323 K 333 K

30 20 10 0

0

50 100 150 200 250 300 350 400 450 500 Reaction time, min

Fig. 4. Comparison of calculated and experimental results in the presence of catalyst (TBAB). Concentration of TBAB = 9.3×10−2 kmol/m3 of org. phase; all other conditions are same as in Fig. 3.

Table 3 Variation of rate constants with temperature in the presence of catalyst (TBAB). Temperature (K)

kc1 (kmol/m3 )−1 s−1

kc2 (kmol/m3 )−1 s−1

kc3 (kmol/m3 )−1 s−1

303 313 323 333

7.75×10−03 4.21×10−02 6.84×10−02 3.44×10−01

11.75 44.54 126.98 241.40

6.00×10−06 1.78×10−05 3.55×10−05 5.39×10−05

higher than those calculated for non-catalytic reactions (Tables 2 and 4). Therefore, the reaction in the presence of catalyst is much more temperature sensitive than the reactions in the absence of catalyst. Moreover, the high values of activation energies (Tables 2 and 4) confirm the fact that the reaction is kinetically controlled and it provides further justification for neglecting the mass transfer resistance in the developed model. 6. Sensitivity study The parametric sensitivity of the validated mechanistic model was further tested using the above estimated rate constants and determining the effect of various process parameters on conversion of BC and selectivity of DBS. The results obtained from the above analysis were then compared with that of experimental observations under different experimental conditions as shown in the Figs. 5–8. The discussions of the results thus obtained are presented below.

Table 4 The frequency factor and activation energy of various rate constants in the presence of catalyst (TBAB). Rate constants

Frequency factor, k0 ((kmol/m3 )−1 s−1 )

Activation energy, E (kJ/mol)

Correlation factor, R2

kc1 kc2 kc3 Experimental

1.22×1015 6.46×1015 2.61×1005 −

99.40 85.09 61.28 68.02

0.96 0.98 0.96 0.99

6.1. Effect of PTC concentration The effect of catalyst loading on conversion of BC and selectivity of DBS was studied at four different catalyst concentrations in the range of 0.0–0.093 kmol/m3 and compared with the model based calculations as shown in Fig. 5. Both conversion of BC and selectivity of DBS at a fixed reaction time were found to increase with increasing PTC concentration as observed from the figures. A reasonably good correlation was obtained with proper trends as observed from the figure. It is worthy to mention here that the developed model is capable of correlating the data even in the absence of any catalyst, which is otherwise not possible if the non-PTC reactions are incorporated in the PTC-enhanced reaction.

Please cite this article as: S.Maity, K., et al., A new mechanistic model for liquid–liquid phase transfer catalysis: Reaction of benzyl chloride with aqueous ammonium sulfide. Chemical Engineering Science (2009), doi:10.1016/j.ces.2009.07.004

ARTICLE IN PRESS S.K. Maity et al. / Chemical Engineering Science

90

100

80

90

70 60 50 40 30

kmol/m3 org. phase Expt. Model 0.0 0.031 0.062 0.093

20 10

% Selectivity of DBS

% conversion of benzyl chloride

8

(

)

--

80 70 60 50 40 kmol/m3 org. phase Expt. Model 0.0 0.031 0.062 0.093

30 20 10 0

0 50 100 150 200 250 300 350 400 450 500

0

0

50 100 150 200 250 300 350 400 450 500

Reaction time, min

Reaction time, min

100 90 80 70 60 50 40 30

0.78 kmol/m3 3 1.44 kmol/m 3 2.0 kmol/m

20 10

Expt. Model

% Selectivity of DBS

% Conversion of benzyl chloride

Fig. 5. Effects of PTC concentration on conversion of benzyl chloride and selectivity of DBS. Temperature = 333 K; all other conditions are same as in Fig. 3.

0 0

120 110 100 90 80 70 60 50 40 30 20 10 0

50 100 150 200 250 300 350 400 450

0.78 kmol/m3 1.44 kmol/m3 2.0 kmol/m3

0

Expt. Model

50 100 150 200 250 300 350 400 450

Reaction time, min

Reaction time, min

100

100

90

90

80 70 60 50 40 30 3

Expt. Model

1.06 kmol/m 3 1.32 kmol/m 3 2.29 kmol/m

20 10

% Selectivity of DBS

% Conversion of benzyl chloride

Fig. 6. Effects of benzyl chloride concentration on conversion of benzyl chloride and selectivity of DBS. The experimental results are taken from Maity et al. (2006).

1.06 kmol/m3 3 1.32 kmol/m 2.29 kmol/m3

80

Expt. Model

70 60 50 40 30 20 10

0

0 0

50

100

150

200

250

Reaction time, min

0

50

100

150

200

250

Reaction time, min

Fig. 7. Effects of sulfide concentration on conversion of benzyl chloride and selectivity of DBS. The experimental results are taken from Maity et al. (2006).

6.2. Effect of benzyl chloride concentration

6.3. Effect of sulfide concentration

The effect of BC concentration on conversion of BC and selectivity of DBS is shown in Fig. 6. With increase in concentration of BC the conversion of BC decreases because of the limited quantity of sulfide in the aqueous phase as observed from the figure. However, the selectivity of DBS increases with increase in concentration of BC. It is also observed from the figure that the model prediction is reasonably close to that of experimental observations.

Fig. 7 shows a comparison of the model-based predictions and experimental results of the effect of sulfide concentration on the conversion of BC and selectivity of DBS. The conversion of BC increases with increase in sulfide concentration in the aqueous phase as expected (Fig. 7a). However, the selectivity of DBS was found to be higher for lower sulfide concentration in the aqueous phase as observed from the Fig. 7b. Also, as evident from the figures, quite a

Please cite this article as: S.Maity, K., et al., A new mechanistic model for liquid–liquid phase transfer catalysis: Reaction of benzyl chloride with aqueous ammonium sulfide. Chemical Engineering Science (2009), doi:10.1016/j.ces.2009.07.004

ARTICLE IN PRESS

100

70

90

60

80 70 60 50 40 30

3.93 kmol/m3 3 6.73 kmol/m 8.14 kmol/m3

20 10

Expt. Model

% Selectivity of DBS

% Convesion of benzyl chloride

S.K. Maity et al. / Chemical Engineering Science

(

)

--

9

50 40 30 20

3.93 kmol/m3 3 6.73 kmol/m 8.14 kmol/m3

10

Expt. Model

0

0 0

50

100

150

200

250

0

50

100

150

200

250

Reaction time, min

Reaction time, min

Fig. 8. Effects of ammonia concentration on conversion of benzyl chloride and selectivity of DBS. The experimental results are taken from Maity et al. (2006).

good agreement was observed between experimental and calculated results with proper trends of the parameters.

6.4. Effect of ammonia concentration The concentration of sulfide ions relative to hydrosulfide ions vary significantly with ammonia concentration in the aqueous phase which in turn affect the conversion of BC and selectivity of DBS. Therefore, the effect of ammonia concentration on this reaction can be used to test the suitability of the model of Edwards et al. (1975) for predicting the concentration of sulfide and hydrosulfide ions in the aqueous phase. The experimental result of the effect of ammonia concentration on the conversion of BC and selectivity of DBS was compared with that of model based prediction as shown in Fig. 8. The appropriate trend of results was obtained from the model. However, the calculated results based on developed mechanistic model were found to deviate slightly from the experimental observations because of the inaccuracy of the model of Edwards et al. (1975). The more appropriate model is therefore needed to represent the aqueous phase ionic equilibrium.

7. Conclusions A kinetic model based on extraction mechanism was developed for the liquid-liquid phase transfer catalyzed reaction of BC with aqueous ammonium sulfide considering separate contribution of non-PTC and PTC-enhanced reactions, the complex ionic equilibrium of NH3 –H2 S–H2 O in the aqueous phase, and equilibrium of catalyst and active catalysts between the phases. A thermodynamic framework developed by Edwards et al. (1975) was used to represent the aqueous phase ionic equilibrium. The rate constants and equilibrium constants of the developed kinetic model were estimated at different temperatures using a modified Levenberg–Marquardt algorithm. The frequency factors and activation energies of the rate constants were determined from the intercept and slope of the Arrhenius plot of ln(rate constants) versus 1/T. Sensitivity analysis was then performed using the estimated parameters under various experimental conditions and then the results were compared with experimental ones. A reasonably good agreement was observed between the calculated results obtained from the developed model and the experimental observations.

Acknowledgment Senior Research Fellowship awarded to Sujit Sen by the Council of Scientific and Industrial Research (CSIR), New Delhi, India is gratefully acknowledged. References Beutier, D., Renon, H., 1978. Representation of NH3 –H2 S–H2 O, NH3 –CO2 –H2 O, and NH3 –SO2 –H2 O vapor–liquid equilibria. Industrial & Engineering Chemistry Process Design and Development 17 (3), 220–230. Bromley, L.A., 1973. Thermodynamic properties of strong electrolytes in aqueous solutions. AIChE Journal 19 (2), 313–320. Dehmlow, E.V., Dehmlow, S.S., 1993. Phase Transfer Catalysis. third ed. Verlag Chemie, Weinheim, Germany. Dehmlow, E.V., Dehmlow, S.S., 1983. Phase Transfer Catalysis, Monographs in Modern Chemistry, vol. II. Verlag Chemie, Weinheim. Edwards, T.J., Newman, J., Prausnitz, J.M., 1975. Thermodynamics of aqueous solutions containing volatile weak electrolytes. AIChE Journal 21 (2), 248–259. 1997. In: Halpern, M.E. (Ed.), ACS Symposium Series 659: Phase-transfer Catalysis, Mechanisms and Syntheses. American Chemical Society, Washington DC, USA. Kohl, A.L., Nielsen, R.B., 1997. Gas Purification. Gulf Publishing Company Houston, Texas. Lele, S.S., Bhave, R.R., Sharma, M.M., 1983. Fast and very low liquid–liquid reaction: phase transfer catalysis. Chemical Engineering Science 22, 765–773. Maity, S.K., Pradhan, N.C., Patwardhan, A.V., 2006. Reaction of benzyl chloride with ammonium sulfide under liquid–liquid phase transfer catalysis: reaction mechanism and kinetics. Journal of Molecular Catalysis A: Chemical 250, 114–121. 1997. In: Sasson, Y., Neumann, R. (Eds.), Handbook of Phase Transfer Catalysis. Blackie Academic & Professional, an imprint of Chapman & Hall, London, UK. Satrio, J.A.B., Doraiswamy, L.K., 2002. Phase-transfer catalysis: a new rigorous mechanistic model for liquid–liquid systems. Chemical Engineering Science 57, 1355–1377. Starks, C.M., Liotta, C.L., Halpern, M., 1994. Phase Transfer Catalysis. Chapman & Hall, New York, USA. Starks, C.M., Owens, R.M., 1973. Phase transfer catalysis. II. Kinetic details of cyanide displacement on 1-haloctanes. Journal of American Chemical Society 95, 3613– 3617. Wang, M.L., Chang, S.W., 1994. Model of the kinetics of synthesizing formaldehyde acetals by phase-transfer catalysis. Industrial & Engineering Chemistry Research 33, 1606–1611. Wang, M.L., Tseng, Y.H., 2003. Phase transfer catalytic reaction of n-bromobutane and sodium sulfide in a two-phase solution and its kinetics. Journal of Molecular Catalysis A: Chemical 203, 79–93. Wang, M.L., Tseng, Y.H., 2002a. Phase-transfer catalytic reaction of dibromo-o-xylene and 1-butanol in two-phase solution. Journal of Molecular Catalysis A: Chemical 179, 17–26.  Wang, M.L., Tseng, Y.H., 2002b. Kinetic model for synthesizing 4,4 -diethanoxy biphenyl by phase transfer catalysis. Journal of Molecular Catalysis A: Chemical 188, 51–61. Wang, M.L., Wu, H.S., 1991. Kinetic and mass transfer studies of a sequential reaction by phase transfer catalysis. Chemical Engineering Science 46 (2), 509–517. Wang, M.L., Yang, H.M., 1991a. Dynamics of phase transfer catalyzed reaction for the allylation of 2,4,6-tribromophenol. Chemical Engineering Science 46 (2), 619–627.

Please cite this article as: S.Maity, K., et al., A new mechanistic model for liquid–liquid phase transfer catalysis: Reaction of benzyl chloride with aqueous ammonium sulfide. Chemical Engineering Science (2009), doi:10.1016/j.ces.2009.07.004

ARTICLE IN PRESS 10

S.K. Maity et al. / Chemical Engineering Science

Wang, M.L., Yang, H.M., 1991b. Simplified dynamic model for the allylation of 2,4,6tribromophenol by phase-transfer catalysis. Industrial & Engineering Chemistry Research 30, 631–635. Wu, H.S., Fang, T.R., Meng, S.S., Hu, K.H., 1998. Equilibrium and extraction of quaternary salt in an organic solvent/alkaline solution: effect of NaOH concentration. Journal of Molecular Catalysis A: Chemical 136, 135–146. Wu, H.S., 1993. Phase-plane modeling of a liquid–liquid phase transfer catalyzed reaction. Industrial & Engineering Chemistry Research 32, 1323–1327. Wu, H.S., 1996. Mathematical modeling for liquid–liquid phase-transfer catalysis. Chemical Engineering Science 51 (5), 827–830.

(

)

--

Yadav, G.D., Jadhav, Y.B., 2003. Kinetics and modeling of liquid–liquid phase transfer catalysed synthesis of p-chlorophenyl acetonitrile: role of co-catalyst in intensification of rates and selectivity. Journal of Molecular Catalysis A: Chemical 192, 41–52. Yang, H.M., Lin, C.L., 2003. Phase-transfer catalyzed benzylation of sodium benzoate using aliquat 336 as catalyst in liquid–liquid system. Journal of Molecular Catalysis A: Chemical 206, 67–76.

Please cite this article as: S.Maity, K., et al., A new mechanistic model for liquid–liquid phase transfer catalysis: Reaction of benzyl chloride with aqueous ammonium sulfide. Chemical Engineering Science (2009), doi:10.1016/j.ces.2009.07.004

A new mechanistic model for liquid–liquid phase ...

tion, arylation, condensation, elimination, polymerization, etc. ..... sion algorithm based on modified Levenberg–Marquardt algorithm. The optimum values of the ...

620KB Sizes 1 Downloads 76 Views

Recommend Documents

Mechanistic Force Model of an Arbitrary Drill Geometry
May 4, 2009 - Drilling is a machining process used to make a straight hole in wood, metal, ceramic and a variety of other materials. Commonly the process is performed by using a twist drill bit. The flutes are responsible for transferring material fr

Multicomponent phase-field model for extremely large ...
Jan 31, 2014 - We develop a multicomponent phase-field model specially formulated to robustly ... The phase-field method introduces a new thermodynamic.

A Thermal Lattice Boltzmann Two-Phase Flow Model ...
simulation by using the method given in Sec. 4 the so-called shifting equilibrium velocity method or by adding an additional term after the collision process.

Nonequilibrium phase transitions in a model with social ...
science. The Axelrod's model of social interaction is proposed to understand the .... phase transitions also does not belong to any known universality class of.

Strategic downsizing-a new model for implementing ...
and overhead. As the pressure for cutting jobs ... ruption to business continuity. THE WRONG WAYS ... ing the strategic needs of the business. Fortunately, there ...

Chromosome numbers and a new model for karyotype ...
cool regions of northern Europe and Canada, while tetraploids inhabit warm or temperate areas from ..... University of California Press, Berkeley,. CA, 356 pp.

Real Candidacy Games: A New Model for ... - Zinovi Rabinovich
May 8, 2017 - breaking schemes, a stable candidate positioning exists. While for several voting rule ... Appears in: Proc. of the 16th International Conference on. Autonomous ..... We shall call the same rule Super Condorcet-consistent (SCC), if, whe

on a new framework for anisotropic damage model
Based on the hypothesis of strain equivalence, the stress-driven damage model proposed by Lemaitre et al.[5] obtained symmetric stiffness tensor but the shear- bulk effects experimentally evidenced in those quasi-brittle materials. Introducing the en

Link-PLSA-LDA: A new unsupervised model for ... - Semantic Scholar
The output of the new model on blog data reveals very inter- ... modeling topics and topic specific influence of blogs. Introduction ..... 07/04/2005 and 07/24/2005.

Chromosome numbers and a new model for karyotype evolution in ...
Ca strovie jo (1983). (su b. R . cirrhosa). 20. Spain. Ciruja no (1986). 20. Italy, Sard inia. M a rc .... Brock, M.A., 1982. Biology of the salinity tolerant genus Ruppia L. in saline lakes in South ... I. Chromosome idiograms and a list of chro- mo

FISHERIES RESEARCH Developing a Mechanistic ...
found limited evidence to support our second prediction regarding premature preparation .... tudes, training, and institutions, to achieve extensive ..... groups for 1 week in either 0 ppt (fresh),. 14 ppt ... tory to possess the range of skills need

Strategic downsizing-a new model for implementing ...
Fortunately, there is an alternative to these ill-advised ... Employment Relations Today DOI 10.1002/ert. 42 ..... energy to replace the modest value he or. John W.

Redefining a New-Formed Average Model for Three ...
ca. i i i. + + = Figure 2 The derived model in literature [3] for boost rectifiers with ...... [1] Rusong Wu, S. B. Dewan and G. R. Slemon, “Analysis of an ac-to-dc.

A New Tractable Model for Cellular Coverage
with the mobile users scattered around the network either as a Poisson point ... be achieved for a fixed user with a small number of interfering base stations, for.

Hemicraniectomy: A New Model for Human ...
Berger improved his EEG recording technology, he was able to acquire EEG from .... approximate underlying Brodmannʼs area (BA; Koessler et al., 2009; Okamoto ...... (Canolty et al., 2006) that information obtained through interfrequency ...

Hemicraniectomy: A New Model for Human ...
rhythms, we hypothesized that that data from the elec- trodes placed over the ... patient with at least 4 weeks recovery time following their ... was 3 in the field.

Study of a low Mach model for two-phase flows with ...
May 4, 2016 - In the low Mach number approach, one possible modelling (system of PDEs .... (Section 2.3) in pure phases based on tabulated values. ...... At time t3 = 20 s the security pumps are turned on, thus the inflow is re-established. ... which

A New Entorhinal Cortex Model
predictive qualities of the compressed representation provided by the .... build a strong relationship between the phasic stimulus and the entorhinal node it.

A New Entorhinal Cortex Model
Abstracts of the Society for Neuroscience Annual Meeting (Miami, FL) 25 ... For the source code, please visit http://alessandro.gagliardi.name/HPC/Stim.ml.html ...

Study of a low Mach nuclear core model for single-phase flows
study is based on a monophasic low Mach number model (Lmnc model) coupled to the .... in the bounded domain Ω = (0,L). ..... The computational cost is about.

Study of a low Mach nuclear core model for single-phase flows
Conditions (4) mean that the fluid is injected at the bottom of the core .... Indeed, as β0 does not depend on h anymore, Eq. (5a) is decoupled from (5b-5c) and can ...... Dealing with unstructured grids requires an efficient localization procedure.

A Behavioural Model for Client Reputation - A client reputation model ...
The problem: unauthorised or malicious activities performed by clients on servers while clients consume services (e.g. email spam) without behavioural history ...