Sim ASPEN Biodiesel Cat MgO

download Sim ASPEN Biodiesel Cat MgO

of 6

Transcript of Sim ASPEN Biodiesel Cat MgO

  • 8/20/2019 Sim ASPEN Biodiesel Cat MgO

    1/13

    Simulation of heterogeneously MgO-catalyzed transesterificationfor fine-chemical and biodiesel industrial production

    Tanguy F. Dossin, Marie-Françoise Reyniers *, Rob J. Berger, Guy B. Marin

     Laboratorium Voor Petrochemische Techniek, UGent Krijgslaan 281 (S5), B-9000 Gent, Belgium

    Received 19 January 2006; received in revised form 5 April 2006; accepted 8 April 2006

    Available online 5 June 2006

    Abstract

    A heterogeneous magnesium oxide catalyst is a good alternative for homogeneous catalysts for the transesterification of alkyl esters for theproduction of fine-chemicals as well as for the production of biodiesel. The transesterification of ethyl acetate with methanol was used as a model

    reaction to simulate fine-chemical production in a batch slurry reactor at industrial conditions. The transesterification of triolein with methanol to

    methyl oleate was chosen to simulate continuous production of biodiesel from rapeseed oil. A kinetic model based on a three-step ‘Eley–Rideal’

    type of mechanism in the liquid phase was used in both process simulations. The transesterification reaction occurs between methanol adsorbed on

    a magnesium oxide free basic site and ethyl acetate or the glyceride from the liquid phase. Methanol adsorption is assumed to be rate-determining

    in both processes. Activity coefficients were required to account for the significant non-ideality of the reaction mixture in the simulations of both

    processes. The simulations indicate that a production of 500 tonnes methyl acetate per year can be reached at ambient temperature in a batch

    reactor of 10 m3 containing 5 kg of MgO catalyst, and that a continuous production of 100,000 tonnes of biodiesel per year can be achieved at

    323 K in a continuous stirred reactor of 25 m3 containing 5700 kg of MgO catalyst. Although various assumptions and simplifications were made

    in these explorative simulations the assumptions concerning the reaction kinetics used, the results indicate that for both processes a heterogeneous

    magnesium oxide catalyst shows promising potential as a viable industrial scale alternative.

    # 2006 Elsevier B.V. All rights reserved.

    Keywords:  Transesterification; Solid base catalyst; Magnesium oxide; Reactor modeling; Industrial simulation; Biodiesel; Rapeseed oil; Kinetics

    1. Introduction

    Transesterification of alkyl esters plays an important

    industrial role with numerous applications for large and small

    volume productions. Large-scale applications are, e.g. the

    production of biodiesel  [1], polyester or PET in the polymer

    industry   [2,3]   while small-scale fine-chemical production

    includes synthesis of intermediates for the pharmaceutical

    industry, the production of food additives or surfactants and thecuring of resins in the paint industry [2,4,5]. Nowadays, most

    industrial applications are performed in batch or continuous

    reactors depending on the production volume. Temperatures

    range from 333 to 523 K while pressure varies from 0.1 to

    10 MPa and alcohol to ester molar ratios from 5 to 15.

    Transesterification reactors are usually coupled with distillation

    columns to separate the alcohols from the reaction mixture

    [6–11]. Reactive distillation is also currently investigated as

    transesterification process   [12–15]. Transesterification reac-

    tions are mainly performed using acid or base homogeneous

    catalysts: sulfuric, sulfonic, phosphoric and hydrochloric acidsas acid catalysts   [2,14]   and alkaline hydroxides, metal

    alkoxides or acetates as base catalysts   [2,6]. Base catalysts

    are preferred over acid catalysts because of the higher reaction

    rates and the lower process temperatures [16]. However, the use

    of homogeneous catalysts requires neutralization and separa-

    tion from the reaction mixture, leading to a series of 

    environmental problems related to the use of large amounts

    of solvents and energy. Heterogeneous solid base catalysts, able

    to catalyze the transesterification of alkyl esters could solve

    these issues: they can be easily separated from the reaction

    www.elsevier.com/locate/apcatbApplied Catalysis B: Environmental 67 (2006) 136–148

     Abbreviations:   BuOAc, butyl acetate; BuOH, n-butanol; BuOK, potassium

    n-butanolate; D, diolein; EtOAc, ethyl acetate; EtOH, ethanol; FAME, fatty

    acid methyl ester; G, glycerine; M, monoolein; M/E, methanol to ethyl acetate

    molar ratio; M/T, methanol to triolein molar ratio; MeOAc, methyl acetate;

    MeOH, methanol; MgO, magnesium oxide; MeOl, methyl oleate; T, triolein

    * Corresponding author. Tel.: +32 9 264 45 17; fax: +32 9 264 49 99.

    E-mail address:  [email protected] (M.-F. Reyniers).

    0926-3373/$ – see front matter # 2006 Elsevier B.V. All rights reserved.

    doi:10.1016/j.apcatb.2006.04.008

    mailto:[email protected]://dx.doi.org/10.1016/j.apcatb.2006.04.008http://dx.doi.org/10.1016/j.apcatb.2006.04.008mailto:[email protected]

  • 8/20/2019 Sim ASPEN Biodiesel Cat MgO

    2/13

    mixture without requiring the use of a solvent, show easy

    regeneration, and have a less corrosive character, leading to

    safer, cheaper and more environment-friendly operation.

    Recently, numerous studies have been performed to investigate

    the use of heterogeneous base compounds as catalysts for

    transesterification reactions. This includes the use of catalysts

    T.F. Dossin et al. / Applied Catalysis B: Environmental 67 (2006) 136–148   137

    Nomenclature

    ai   activity of component i  (mol m3)

    a LS    liquid–solid interfacial area (m2 m3)

    amn   group interaction parameter (in calculation

    activity coefficient)

     A   pre-exponential factor (m3 kgcat

    1 s

    1)C i   concentration of component i  (mol m

    3)

    d 1,2   Flory–Huggins combinatorial term

    d  I    impeller diameter (m)

    d  p   solid particle diameter (m)

     Di   liquid diffusion coefficient of component   i

    (m2 s1)

     Di,eff    effective diffusion coefficient of component i  in

    the pellet pores (m2 s1)

     Dim   bulk diffusivity of component i  (m2 s1)

     D0 AB   binary diffusivity of component B  in  A  (m2 s1)

    E a   activation energy (J mol1)

    F   molar flow (mol s

    1

    )g   acceleration of the gravity (m2 s1)

    k BuOK    reaction rate coefficient for BuOK catalyzed

    reaction (m6 /mol2 s)

    k MeOH   reaction rate coefficient of methanol adsorption

    step (m3 kgcat1 s1)

    k  LS ,i   liquid–catalyst mass transfer coefficient for com-

    ponent  I  (m3 m2 s1)

    K eq   equilibrium constant of the overall reaction

    K  A   equilibrium constant of alcohol adsorption

    (m3 mol1)

    li   volume parameter (in calculation activity coeffi-

    cient)

    MWi   molecular weight of component i  (kg/mol)ni   number of moles of component  i  (mol)

     N    total number of components

     N  I    impeller revolution speed (s1)

     N P   impeller power number (=5) (s1)

     p   pressure (atm)

    P   yearly production (tonnes year1)

    q   induction parameter

    qi   molecular surface area parameter (in calculation

    activity coefficient)

    r BuOK    BuOK-catalyzed reaction rate (mol m3 s1)

    r i   van der Waals volume parameter (in calculation

    activity coefficient)r MgO   MgO-catalyzed reaction rate (mol kgcat1 s1)

     R   universal gas constant (J mol1 K 1)

     Re   Kolmogoroff Reynolds number = r3 L d 4 p N  pd 

    5 I  N 

    3i =

    ðm3 L V  L ÞSci   Schmidt number component  i =  k  LS ,i  d  p /  Dim  (s)

    t    time (s)

    T    temperature (K)

    V  L    liquid volume (m3)

    W    mass of catalyst (kgcat)

     xi   molar fraction of component i

     X i   conversion of component  i

    Y i   product yield of component  i

    Greek symbols

    a   acidity parameter

    b   basicity parameter

    g i   activity coefficient of component ig 1i   activity coefficient of component   i   at infinite

    dilution

    g C i   combinatorial contribution to activity coefficientof component  i

    g  Ri   residual contribution to activity coefficient of 

    component i

    Fi   segment fraction (in calculation activity coeffi-

    cient)

    u i   area fraction (in calculation activity coefficient)

    G k    group residual activity coefficient

    G ðiÞk    group residual activity coefficient in a reference

    solution containing only molecules of type   i

    e p   catalyst porosity

    es   solid fraction in the slurry reactor

    l   dispersion parameterLi, j   Wilson binary parameter of component   i   in

    solvent  j

    m L    liquid viscosity (kg m1 s1)

    n   liquid molar volume at 293 K (cm3 mol1)

    j    asymmetry parameter due to hydrogen bonding

    r L    liquid density (kg m3)

    r p   catalyst pellet density (kg m3)

    r p,wet   density of the catalyst particle filled with liquid

    (kg m3)

    t    polar parameter or catalyst tortuosity

    c   asymmetry parameter due to polarity difference

    cmn   group interaction parameter (in calculation

    activity coefficient)

    Subscripts and superscripts

    0 initial condition* basic site

    cat catalyst

    eq equilibrium

    i   component  i

    in inlet

     I    impeller

     j   parameter  j

    k k th experimental point

    out outlet p   catalyst pellet

     R   reactor

    s   solvent

    tot total

  • 8/20/2019 Sim ASPEN Biodiesel Cat MgO

    3/13

    such as alkaline-earth oxides, zeolites, and hydrotalcites

    [3,4,17–20]. Nowadays only one industrial application,

    developed by the IFP, uses heterogeneous base catalysts for

    the transesterification of vegetable oils to produce biodiesel

    [21].

    Although several authors investigated the kinetics of 

    transesterification catalyzed by homogeneous base catalysts

    [6,14,16,22–24], there is very little information available

    concerning the kinetics of the transesterification catalyzed by

    solid bases. Hattori et al. [18] have proposed a mechanism for

    the transesterification of ethyl acetate with different alcohols

    catalyzed by a variety of solid base catalysts, particularly

    alkaline-earth-metal oxides. However, this mechanistic study

    did neither provide values of activation energies nor rate

    constants. Recently, a kinetic model was proposed to describe

    the transesterification of ethyl acetate with methanol catalyzed

    by a heterogeneous magnesium oxide catalyst [25]. This model

    is based on a three-step ‘Eley–Rideal’ type of mechanism

    applied in liquid phase where methanol adsorption is assumed

    to be rate determining. The proposed value of the activationenergy is 20 kJ/mol and the value of the methanol adsorption

    equilibrium coefficient is 3.13 103 m3 /mol. From severalkinetic studies on homogeneous base catalysts followed a

    pseudo-second-order rate law with activation energies ranging

    from 22 to 83 kJ/mol, depending on the type of alcohol and

    ester used [6,14,16,22,23].

    The objective of this study was to evaluate the use of the

    heterogeneously MgO-catalyzed transesterification reaction in

    batch and continuous stirred tank reactors at industrially

    relevant conditions using the kinetic model based on the three-

    step ‘Eley–Rideal’ type mechanism assuming methanol

    adsorption as rate-determining step. Two processes have beensimulated: the transesterification of ethyl acetate with methanol

    in a batch slurry reactor for fine-chemical production and the

    transesterification of triolein with methanol to form methyl

    oleate to simulate biodiesel production from rapeseed oil in a

    continuous slurry reactor. The simulation results from both

    have been compared with those of homogeneously catalyzed

    transesterification processes.

    Biodiesel consists of alkyl esters of long chain fatty acids

    originating from a renewable resource such as vegetable oils.

    Biodiesel can be used in usual diesel engines and presents some

    advantages compared to traditional petroleum-based diesel.

    Biodiesel is an environment-friendly compound, non-toxic and

    emits less carbon monoxide, sulphur compounds, particulatematter and unburned hydrocarbons than traditional diesel.

    However, NO x   emissions are 10% higher compared to

    petroleum-based diesel. Moreover, since biodiesel is produced

    from a renewable source, the net production of carbon dioxide

    is lowered with reduced contribution to the greenhouse effect.

    Biodiesel is usually produced by transesterification of 

    vegetable oils, largely consisting of triglycerides, with

    methanol to produce fatty acid methyl esters (FAME), i.e.

    biodiesel, and glycerine.

    The catalysts used are typically alkali hydroxides such as

    KOH or NaOH and alkali alkoxides such as sodium methoxide

    or sodium butoxide. Non-ionic base catalysts such as TBD

    (1,5,7-triazabicyclo[4.4.0]dec-5-ene, C7H13N3) are currently

    studied [20,26]. The main purpose of transesterification is to

    reduce the viscosity of the oil. Vegetable oils cannot be used as

    such in common diesel engines because their viscosity is 10–20

    times that of conventional diesel, causing injector coking and

    engine deposits [1,26]. Biodiesel does not show these problems

    and offers the same performance and engine durability as

    petroleum-based diesel. Therefore, the importance of biodiesel

    increases due to environmental concerns and uncertainties

    concerning petroleum price and availabilities.

    The transesterification of triolein with methanol to methyl

    oleate (CH3(CH2)7CH=CH(CH2)7COOCH3) has been chosen

    as a model process for the transesterification of rapeseed oil to

    produce biodiesel. Rapeseed oil is the most abundant oil

    produced in Europe and triolein (C57H104O6) has been chosen

    to represent rapeseed oil as oleic acid is the major fatty acid in

    rapeseed oil [26].

    2. Modeling procedures

    2.1. Transesterification of ethyl acetate with methanol

    2.1.1. Thermodynamics

    Thermodynamic properties of the components and the

    transesterification reaction between 283 and 323 K were

    calculated using the ASPEN Engineering SuiteTM 11.1  [27].

    The reaction enthalpy varies from  2.8 at 273 K to  3.2 kJ/ mol at 328 K, showing a slightly exothermic reaction. The

    value of the equilibrium coefficient  K eq  ranges from 1.875 at

    293 K to 1.716 at 323 K, which is similar to values reported in

    previous studies of transesterification reactions   [14,22]. Theequilibrium conversion of ethyl acetate   X eq  varies from 9 to

    95% at molar ratios of methanol to ethyl acetate (M/E) ranging

    from 0.1 to 10.

    The non-ideal character of the reaction mixture was assessed

    and activity coefficients   g i   for each component   i   were

    calculated at temperatures ranging from 283 to 323 K based

    on the Wilson equation as follows  [28]:

    ln g i  ¼ ln

    X N  j¼1

     x jLi; j

    þ 1

    X N k ¼1

     xk Lk ;iP N  j¼1 x jLk ; j

    (1)

    where, N ,  xi and  L i, j  are the number of components, the molefraction of each component and the Wilson binary parameter of 

    component i diluted in solvent j. The Wilson binary parameters

    are calculated based on Eq.  (1)  assuming infinite dilution of 

    component   i   in solvent   j. The activity coefficient at infinite

    dilution is then calculated based on Eq.  (2)  by means of the

    MOSCED method [28]:

    ln g 1i   ¼  ni

     RT 

    ðls liÞ

    2 þq2sq

    2i ðt s t iÞ

    2

    cs

    þðas aiÞðbs biÞ

    j s

    þ d s;i   (2)

    T.F. Dossin et al. / Applied Catalysis B: Environmental 67 (2006) 136–148138

  • 8/20/2019 Sim ASPEN Biodiesel Cat MgO

    4/13

    where g 1i   is the activity coefficient of component i in solvent s

    at infinite dilution, and a, b, l, n, j , t , c, q and d s,i are tabulated

    parameters, specific for each component.

    Table 1 shows the influence of the M/E ratio at 303 K on the

    activity coefficients of each component at conditions corre-

    sponding to the start of the reaction and to the equilibrium

    conversion. Since the calculated activity coefficients were

    fairly different from 1.0, the non-ideality of the mixture was

    taken into account in the simulations.

    2.1.2. Kinetic model

    The kinetic model used to describe the transesterification of 

    ethyl acetate with methanol to form methyl acetate and ethanol

    is based on a three-step ‘Eley–Rideal’ type mechanism  [25].

    The elementary steps of the mechanism are given in Table 2.

    The first step describes the adsorption of methanol on a

    catalytically active site. Then, adsorbed methanol reacts with

    ethyl acetate from the liquid phase to form methyl acetate and

    adsorbed ethanol. Ethanol finally desorbs in the third step. The

    proposed model assumes methanol adsorption as rate-deter-

    mining step, leading to the following rate equation:

    r MgO  ¼k MeOHðaMeOH ð1=K eqÞðaMeOAcaEtOH=aEtOAcÞÞ

    1 þ ðK  A=K eqÞðaMeOAcaEtOH=aEtOAcÞ þ K  AaEtOH(3)

    where k MeOH  is the methanol adsorption rate constant,  ai, the

    activities of reaction component  i,  K eq the equilibrium coeffi-

    cient of the overall transesterification reaction and   K  A   is the

    equilibrium constant of the methanol adsorption step. The value

    of the adsorption equilibrium coefficient is assumed to be the

    same for both alcohols. The values of the activation energy  E aand the pre-exponential factor  A  of the rate constant  k MeOH  as

    well as the adsorption equilibrium coefficient   K  A   have been

    estimated and their values are presented in   Table 3. Theparameter estimation procedure and its results are described

    elsewhere [25]. The value of  K eq has been calculated based on

    thermodynamic data for all compounds.

    According to Madon and Iglesia  [29], it is appropriate in

    non-ideal systems to use the activities instead of the

    concentrations in the rate expression if the rate-determining

    step is an adsorption or desorption step, because the

    corresponding activated complexes may be solvated by the

    surrounding fluid phase. However, if the solvation effect of the

    activated complex is similar to that of the reactants (products),

    the concentrations instead of the activities should be applied

    because a cancellation of activity coefficients leads to the

    apparent dependence of rates on concentration rather than

    activities. For the current reaction, it is unclear which situation

    applies and therefore the effect of this was investigated, thus

    using the following rate equation:

    r MgO  ¼k MeOHðC MeOH ð1=K eqÞðaMeOAcaEtOH=aEtOAcÞÞ

    1þ ðK  A=K eqÞðaMeOAcaEtOH=aEtOAcÞ þ K  AC EtOH(3a)

    Since the location of the thermodynamic equilibrium in non-

    ideal mixtures is determined by the equilibrium constants in

    combination with the component activities rather than the

    component concentrations, the activities are maintained in

    the terms correcting for the equilibrium. This thus results in

    a rather unusual rate expression containing activities as well as

    concentration terms. A similar form of rate equation, however,

    has been proposed by Madon and Iglesia  [29]   for the cyclo-

    hexene hydrogenation over Pd.

    Simulations results using rate Eq.  (3a)   are compared to

    simulation results using Eq. (3)  to study the possible influence

    of solvation in the non-ideal liquid phase.

    2.1.3. Reactor model

    The slurry reactor, operated batch wise or in a continuous

    way, was chosen as industrial reactor because of its easy

    operability and frequent use in fine-chemical industry. The

    T.F. Dossin et al. / Applied Catalysis B: Environmental 67 (2006) 136–148   139

    Table 1

    Influence of the M/E ratio on the activity coefficients of MeOH, EtOH, MeOAc,

    EtOAc at 303 K, at the start of the reaction and at equilibrium

    M/E ratio 0.1 1 5 10

    Conversion (mol%) 0 9 0 58 0 89 0 95

    MeOH 3.08 3.11 1.29 1.30 1.03 1.03 1.01 1.01

    EtOH – 2.79 – 1.26 – 1.09 – 1.09MeOAc – 1.00 – 1.30 – 1.90 – 2.10

    EtOAc 1.02 1.02 1.42 1.42 2.22 2.24 2.51 2.53

    Table 2

    Elementary reactions of Eley–Rideal type reaction mechanism

    CH3OH + * CH3OH*

    CH3OH* + CH3COOCH2CH3   CH3CH2OH

    * + CH3COOCH3CH3CH2OH

    * CH3CH2OH + *

    Table 3

    Parameter values of the kinetic model used

     A  (m3 /kgcat s) 0.148

    E a  (103 J/mol) 20.1

    K  A  (103 m3 /mol) 5.29

    Table 4

    Range of process parameters

    Conditions

    T  (K) 293–323

     p  (atm) 1

    M/E () 0.1–30W  (kg) 5–10000

    F inEtOAc(mol s1)   1–100

    Reactor

    V  L  (m3) 10–40

    d  R  (m) 2.3–3.7

    h R  (m) 2.3–3.7

     N  I  (s1) 1

    d  I  (m) 1.2–1.9

    Catalyst

    d  p  (106 m) 25

    e p   0.45

    t  p   5.0

    r p  (kg/m3) 2460

  • 8/20/2019 Sim ASPEN Biodiesel Cat MgO

    5/13

    reactor was considered to be cylindrical with equal reactor

    diameter and height and with a volume between 10 and 40 m 3.

    The range of process parameters considered in the simulations

    is given in   Table 4. Since transesterification reactions are

    equilibrium-controlled reactions, an excess of methanol is used

    to shift the equilibrium towards the product side. The methanol

    to ethyl acetate molar ratio (M/E) varies from 5 to 30  [16]. The

    reactor geometry follows the standard tank geometry for

    turbulent mixing with a radial flow impeller proposed by

    Tatterson   [30]. The impeller is a traditional Rushton turbine

    with a diameter from 1.2 to 1.9 m. The impeller speed is

    calculated to fulfill perfect mixing and complete suspension of 

    solid particles, following Eq.  (4) [31]:

     N  I  >

    12

     g

    d  I 

    0:5 r p;wet r L r L 

    0:5d  p

    d  I 

    0:165e

    0:25S    (4)

    where N  I  is the impeller speed,  d  I , the impeller diameter, g, the

    gravitational constant,  r  p;wet, the density of the catalyst pellet

    filled with liquid,r L , the liquid density, d  p, the particle diameterand   es  is the catalyst fraction in the slurry. Since the criterion

    was only validated in the range 0.01  <   eS  

  • 8/20/2019 Sim ASPEN Biodiesel Cat MgO

    6/13

    values are close to 1.0 since the reactions are almost

    thermoneutral and the entropy change is likely to be very

    small as well. For simplicity, all the equilibrium constants were

    taken equal to 1.0 in the simulations, independent of 

    temperature.

    Since it was observed that the reaction mixture consisting of 

    ethyl acetate, methanol, methyl acetate and ethanol was not

    ideal (see Section 2.1.1), the need for taking non-ideality into

    account was also checked for the biodiesel simulations.

    However, unlike the transesterification of ethyl acetate with

    methanol, the MOSCED method could not be applied to solve

    the Wilson equation since data for glyceride molecules are nottabulated. Therefore, the UNIFAC contribution method was

    used to calculate activity coefficients   [28]. The activity

    coefficient is separated into two parts: one part provides the

    contribution due to molecular size and shape and is defined as

    the combinatorial contribution (ln g C i  ), and the other provides

    the contribution due to molecular interactions, known as the

    residual contribution (ln g  Ri ):

    ln g i  ¼ ln g C i   þ ln g 

     Ri   (10)

    where

    ln g C i   ¼ lnFi

     xiþ 5qiln  u i

    Fiþ li Fi

     xi

    Xallcomp: j

     x jl j   and

    ln g  R ¼Xallgroupsk 

    nðiÞðlnG k  lnG ðiÞk   Þ;

    u i  ¼  xiqiP

     j x jq j;   Fi  ¼

      xir iP j x jr  j

    ;

    li  ¼ 5ðr i qiÞ ðr i 1Þ;   r i  ¼Xk 

    nðiÞk   Rk ;

    qi  ¼Xk 

    nðiÞk   Qk 

    and

    lnG k  ¼ Qk 

    1 ln

     Xallcomp:m

    u mC mk 

    Xallcomp:m

    u mC kmPallcomp:n   u nC nm

    ;

    C mn  ¼ exp

    amn

    :

     Rk   and   Qk   are group volumes and area parameters and aretabulated as a function of the type of group.  T is the temperature

    expressed in Kelvin, nðiÞk    is the number of groups k  present in the

    molecule i and xi is eitherthe molar fraction of component i in the

    mixture or the mole fraction of group i in the component. Table 5

    shows the influence of the M/T ratio at 323 K on the activity

    coefficientsof each component andat thestart of thereactionand

    at equilibrium in terms of mole fraction. Since the calculated

    activity coefficients were fairly different from 1.0, the non-

    ideality of the mixture was taken into account in the simulations.

    2.2.2. Kinetic model

    Several kinetic studies of methyl ester production from

    different vegetable oils have been performed using homo-geneous base catalysts [6,16,20,23]; however, to the best of our

    knowledge, no kinetic data of vegetable oil transesterification

    involving heterogeneous catalysts have been reported. All

    kinetic studies performed with homogenous catalysts show that

    all three consecutive transesterification reaction rates follow a

    pseudo-second-order law with respect to the glyceride and the

    alcohol. As can be seen in Freedman’s work   [16], the three

    consecutive transesterification reactions have significantly

    different activation energies.

    Since the reaction mechanism of the transesterification

    reactions over the MgO catalyst is assumed to be of the

    Eley–Rideal type with the adsorption of methanol being the

    T.F. Dossin et al. / Applied Catalysis B: Environmental 67 (2006) 136–148   141

    Table 5

    Influence of the M/T ratio on the activity coefficients (g i) of all components at 323 K, at the start of the reaction and at equilibrium ( xi = mole fraction)

    Triolein Diolein Monoolein Methyl oleate Glycerine MeOH

    M/T = 5

     xi   0.16648 0.00004 0.00006 0.00029 0.00003 0.83311

    g  I    1.34479 0.65299 0.83860 2.15770 28.27353 1.23042

     xi  (equilibrium) 0.01987 0.06435 0.07885 0.23293 0.00368 0.60033

    g  I    1.23714 0.60129 0.77292 2.06543 26.08368 1.26260

    M/T = 10

     xi   0.09086 0.00000 0.00002 0.00012 0.00003 0.90897

    g  I    2.58711 0.71287 0.51951 3.67922 9.93927 1.12314

     xi  (equilibrium) 0.00432 0.02538 0.05645 0.15259 0.00478 0.75648

    g  I    2.35993 0.64583 0.46744 3.50896 8.88203 1.13984

    M/T = 20

     xi   0.04759 0.00000 0.00001 0.00007 0.00002 0.95231

    g  I    6.56754 0.97234 0.38074 6.66597 3.91396 1.05075

     xi  (equilibrium) 0.00077 0.00826 0.03344 0.09059 0.00516 0.86177

    g  I    6.22731 0.90549 0.34822 6.45176 3.51569 1.05628

  • 8/20/2019 Sim ASPEN Biodiesel Cat MgO

    7/13

    rate-determining step, the kinetic model used for simulation of 

    the three heterogeneous transesterification reactions over the

    MgO catalyst is assumed to be the same as that for the

    transesterification of ethylacetate, as discussed in Section 2.1.2.

    As this mechanism is of the Eley–Rideal type with the

    adsorption of methanol being the rate-determining step, it can

    be assumed that the values of the activation energy  E a and the

    pre-exponential factor   A   for each of the three transesterifica-

    tions are equal to the values for the transesterification of ethyl

    acetate with methanol, as presented in   Table 3. The rate

    equations for the three glyceride transesterifications are thus

    assumed to be, analogous to Eq.  (3):

    r T  ¼

    k MeOH

    aMeOH

      1K eq

    aMeOla DaT 

    1þ   K  AK eq

    aMeOla DaT 

    þ K  Aa D þ K  Aa M þ K  AaG(11a)

    r  D ¼

    k MeOH

    aMeOH

      1K eq

    aMeOla M a D

    1 þ  K  AK eq

    aMeOla M a D

    þ K  Aa D þ K  Aa M þ K  AaG (11b)

    r  M  ¼

    k MeOH

    aMeOH

      1K eq

    aMeOlaGa M 

    1 þ  K  AK eq

    aMeOlaGa M 

    þ K  Aa D þ K  Aa M þ K  AaG(11c)

    where k MeOH is the methanol adsorption rate constant equal for

    all three reactions, ai, the activities of each reaction component,

    K eq the equilibrium coefficient of the overall transesterification

    reaction and  K  A  is the adsorption equilibrium constant of the

    alcohols present, assumed to be the same for all alcohols

    present. The elementary steps and the corresponding reactionmechanism for each transesterification reaction are shown in

    Table 6.

    It should be mentioned that the assumption of exactly the

    same reaction kinetics for the transesterification of large

    glyceride molecules as that for the transesterification of ethyl

    acetate is rather speculative. The long organic tails of the

    glyceride molecules will probably have some influence on the

    kinetics. Additionally, the assumption of equal rate coefficients

    for the three consecutive reactions seems rather speculative in

    view of the significant differences between the activation

    energies of the three consecutive transesterification reactions

    observed with homogeneous catalysts, as mentioned above.

    Similarly as for batch slurry reactor simulations (see Section

    2.1.2), the results obtained using the rate equations based on

    the component activities   (11a)–(11c)   were compared to the

    results using the rates based on the component concentrations

    (Eqs. (12a)–(12c)):

    r T  ¼k MeOH

    C MeOH

      1

    K eq;1

    aMeOla D

    aT 

    1 þ   K  AK eq;1

    aMeOla DaT 

    þ K  AC  D þ K  AC  M þ K  AC G(12a)

    r  D ¼

    k MeOH

    C MeOH

      1K eq;2

    aMeOla M a D

    1 þ   K  AK eq;2

    aMeOla M a D

    þ K  AC  D þ K  AC  M þ K  AC G(12b)

    r  M  ¼

    k MeOH

    C MeOH

      1K eq;3

    aMeOlaGa M 

    1 þ   K  AK eq;3

    aMeOlaGa M 

    þ K  AC  D þ K  AC  M þ K  AC G(12c)

    2.2.3. Reactor model

    As for the simulation of fine-chemical production, a slurry

    reactor has been chosen to simulate continuous biodiesel

    production. All assumptions mentioned for the simulations in

    the EtOAc/MeOH case are also valid for the biodiesel

    production process (see Section   2.1.3). Additionally, the

    continuous slurry reactor was assumed to be at steady state,

    start-up time was not taken into account and the catalyst was

    retained in the reactor by a microporous filter.

    For simplicity reasons, the possibility of phase separation

    between alcohols and glycerides was not included in the

    explorative simulations discussed in this paper. According toNoureddini and Zhou   [23], this phase separation disappears

    rapidly if a high impeller speed of at least 5 s1 is applied since

    methyl oleate functions as mutual solvent to form a single-

    phase system as reaction advances. However, it seems more

    likely that the good mixing leads to a micro-emulsion of a polar

    and a non-polar liquid phase [1]. Particularly the high activity

    coefficients of methyl oleate and glycerol (Table 5), which were

    calculated assuming one homogenous liquid phase, indicate the

    possibility of such a phase separation.

    The existence of a phase separation may strongly influence

    the reactor performance. On the one hand it may negatively

    influence the performance if mass transport limitations occur

    between both phases. On the other hand it may influence theoverall equilibrium composition, since a phase separation

    generally leads to a significant decrease of the values of the

    activity coefficients.   Table 5   shows the composition at

    equilibrium as a function of the M/T if no phase separation

    is taken into account. For instance, at a M/T ratio of 10, the

    maximum achievable mole fraction of methyl oleate amounts to

    0.1526, corresponding to a yield of only 0.56. According to

    these calculations, the reaction product will thus always contain

    significant amounts of diolein and monoolein besides methyl

    oleate, as shown in Table 5.

    Experimental studies using homogeneous alkali catalysts,

    however, have shown that methyl ester yields of over 0.9 can be

    T.F. Dossin et al. / Applied Catalysis B: Environmental 67 (2006) 136–148142

    Table 6

    Elementary steps and reaction mechanisms for the biodiesel transesterification

    reactions

    Elementary reactions Reaction mechanism

    T ! D D ! M M ! G

    CH3OH + * CH3OH* 1 1 1

    CH3OH* + T D* + MeOl 1

    CH3OH* + D M* + MeOl 1

    CH3OH* + M G* + MeOl 1

    D* D + * 1

    M* M + * 1

    G* G + * 1

  • 8/20/2019 Sim ASPEN Biodiesel Cat MgO

    8/13

    reachedatanM/Tabove5 [23,24,46,39]. Besides the existence of 

    a phase separation, this discrepancy may also originate from the

    estimation of the equilibrium constants or from the UNIFAC

    contribution method used to estimate the activity coefficients.

    Nevertheless, to the best of our knowledge, the UNIFAC

    contribution method is currently the best method to estimate

    activity coefficients of complex and non-tabulated molecules.

    The occurrence of phase separation therefore seems to be the

    most likely explanation of the discrepancy since the activity

    coefficients of the final twoproducts, methyl oleate and glycerol,

    are the highest of all components, thus limiting the equilibrium

    conversion to a relatively low value, which can be understood

    based on Eq. (11c); a decrease of the activity coefficients as a

    result of a phase separation will thus increase the conversion to

    the desired product methyl oleate at equilibrium. Due to the

    significant deviation of the equilibrium conversion, the simula-

    tion results are to be interpreted in a qualitative way only.

    Another important aspect in this process is the presence of 

    mass transport limitations around and inside the catalyst

    particles as a result of the relatively low diffusivity, which iscaused by the high viscosity of the reaction mixture. It may give

    rise to significant concentrations gradients. In contrast, heat

    transport limitations do not have to be taken into account as a

    result of the thermoneutrality of the reaction. In order to take

    the effect of mass transfer limitation into account, the reactor

    model equation in the bulk phase for a continuous slurry reactor

    considering steady-state operation and mass transport limita-

    tions is as follows:

    F ini   F outi   ¼ V  L k  LS ;ia LS ðC  L ;i C S ;iÞ   (13)

    where F ini   and F outi   are inlet and outlet molar flows of compound

    i, k  LS ,i  the liquid–solid mass transfer coefficient of the compo-nent i  in methanol, assumed to be in excess in the reactor, and

    a LS   the liquid–catalyst interface area, and  V  L  is the total liquid

    volume. C  L ,i and  C S ,i are the concentrations in the liquid phase

    and at the external pellet surface, respectively.

    The mass transfer coefficient is calculated from Sano et al.

    [40]:

    k  LS ;i ¼ 2þ 0:4 Re0:25Sc0:333i   (14)

    The required liquid density and viscosity were calculated as the

    weight-weighted average value of the values of the individual

    components, which were takenfrom various sources as indicated

    in Table 7.

    Since uniform spherical catalyst particles are assumed, the

    mass balance inside the pellet is, applying spherical coordi-

    nates:

    4

    e pd 2 p

    1

    j 2@

    @j 

    j 2 Deff ;i

    @C i

    @j 

    ¼ r 

    r p

    e p

    (15)

     Deff,i   and   C i   are the effective diffusion coefficient and theconcentration of component   i   in methanol inside the pellet;

    e p and  r p are the porosity and the density of the catalyst pellet,

    respectively. j is the spatial coordinate inside de pellet, equal to

    1 at the surface and 0 at the centre of the pellet. r  is the net rate

    of formation of this specific component. The effective diffu-

    sivity Deff,i  depends on the concentrations through the depen-

    dence of the activity coefficients on the mole fractions. The

    effective diffusion coefficient was estimated from  [41]:

     Deff ;i  ¼  e p

    t  p Dim   (16)

    The bulk diffusivity Dim was calculated from the binary diffu-

    sion coefficient in methanol ( A =  i,   B = methanol)   [42]   and

    corrected for the effect of non-ideality [43] using:

     D AB ¼ ð x A D0 BA þ x B D

    0 ABÞ

    1 þ

    @ ln g  A@ ln x A

      (17)

    Since  x A D0 BA x B D

    0 AB, this equation was simplified to:

     D AB ¼  x B D0 AB

    1 þ

    @ ln g  A@ ln x A

      (18)

    resulting in:

     Dim  ¼  xMeOH D0

    im

    1 þ

    @ ln g i@ ln xi

      (19)

    At the pellet surface, the effective diffusion of the component  i

    inside the pellet is equal to the diffusion of this component from

    the bulk phase to the pellet surface. This is expressed by the

    following equation:

    2 Deff ;i

    k  LS ;id  p

    @C i

    @j 

    j ¼1

    ¼ C  L ;i C S ;i   (20)

    Eq. (13) was applied to all reaction compounds and the reaction

    rate r  in this equation is the sum of the three transesterification

    reaction rates (Eqs. (11a)–(11c)) taking the reaction stoichio-

    metry into account. Simulations were performed with the

    T.F. Dossin et al. / Applied Catalysis B: Environmental 67 (2006) 136–148   143

    Table 7

    Pure component densities and viscosities used in the biodiesel production process simulations

    Component Densitya (kg/m3) Source Viscosityb (mPa s)   T 0  (K) Source

    Triolein 1113–0.684T    [27,47]   45.1 313   [47,48]

    Diolein 1042–0.42T    [27]   68 313   [49]

    Monoolein 1108–0.543T    [27]   91.5 313   [49]

    Methyl oleate 1082–0.712T    [27,47]   5.2 313   [48,47,50]

    Glycerine 1453–0.656T    [27,51]   1500 293   [51]

    Methanol 1079–0.971T    [27,51]   0.55 298   [51]

    a T : temperature (K).b The Lewis and Squires liquid viscosity–temperature correlation [28] wasusedto obtain theviscosity at therequiredtemperature:m0:2661T    ¼ m

    0:2661T 0

      þ T T 0233

     , with

    mT  = liquid viscosity at the required temperature  T  (K), and  mT 0  = known viscosity at temperature  T 0  (with viscosities expressed in mPa s).

  • 8/20/2019 Sim ASPEN Biodiesel Cat MgO

    9/13

    Athena Visual Workbench software developed by Stewart &

    Associates Engineering Inc., using the PDAPlus solver. This

    software allows the solution of boundary value problems

    involving partial differential equations. The integration of 

    the catalyst particles was performed using the method of finite

    differences with 150 discretization points.

    The biodiesel production in a continuous slurry reactor

    typically amounts to 100,000 tonnes year1 [44,45]. The

    applicable temperature range is narrow due to the high

    viscosity of the glycerides and the low boiling point of 

    methanol. Since the melting point of the monoglyceride is

    about 310 K and the methanol boiling point is about 340 K, the

    operating temperature window only ranges from 320 to 330 K.

    Nowadays, the methanol to oil molar ratio is typically set to 6,

    although a recent patent showed that higher molar ratios can

    improve the yield sufficiently to make it economically

    attractive   [46]. Therefore, the operating molar methanol/ 

    triolein ratio will be in the range between 5 and 30.

    The conversions and the yields were calculated according to:

     xT  ¼ 1  F T ;out

    F T ;feed;   xMeOH  ¼ 1

     F MeOH;out

    F MeOH;feed(21)

    Y MeOl ¼F MeOl;out

    3F T ;feed;   Y  D ¼

     F  D;out

    F T ;feed;   Y  M  ¼

     F  M ;out

    F T ;feed;

    Y G ¼ F G;out

    F T ;feed

    (22)

    The yearly production of methyl oleate and glycerine were

    calculated as follows:

    PMeOl  ¼ 3Y 

    MeOl W 

    MWMeOl

    60

    60

    24

    365

    1000ðW =F T ;feedÞ

    ðtonnes=yearÞ   (23)

    PG ¼Y G W MWG 60 60 24 365

    1000ðW =F T ;feedÞ  ðtonnes=yearÞ

    (24)

    3. Results and discussion

    3.1. Batch slurry reactor for fine-chemical production

    Simulations of a batch slurry reactor for the heterogeneouslyMgO-catalyzed transesterification of ethyl acetate with

    methanol at industrial process conditions have been performed

    using Eq. (5). Influences of temperature, amount of catalyst and

    molar ratio have been investigated to achieve a typical

    production of 500 tonnes MeOAc year1. Traditionally, batch

    times do not exceed 2 days.

    Table 8  shows the evolution of the annual production as a

    function of temperature ranging from 293 to 323 K, at M/E = 10,

    W  = 2 kg,   V  L  = 10 m3.   Fig. 1   shows the corresponding ethyl

    acetate conversion as a function of the batch time. The minimum

    required batch time is defined as the time to reach 95% of the

    ethyl acetate equilibrium conversion. At the conditions stated

    above, this corresponds to an EtOAc equilibrium conversion of 

    90%. As expected, the minimum required batch time decreases

    with increasing temperature. For the considered process

    parameters, a yearly production of almost 500 tonnes year1

    is already achieved at 323 K (see  Table 8). Nevertheless, it is

    useful to study the influence of the other process parameters such

    as the amount of catalyst and the M/E molar ratio on the yearly

    production. The influence of the catalyst mass on the yearly

    production is presented in Table9. From these results follows that

    a similar yearly production (524 tonnes year1) can be achieved

    at 293 K if the amount of MgO catalyst is increased from 2 to4.5 kg (M/E = 10,  V  L  = 10 m

    3). Running the process at 293 K 

    could eliminate theneed forreactor heating, whichmay make the

    process easier and less expensive, assuming the cost of the

    catalyst is much smaller than the cost of the process equipment

    combined with the energycost. Theinfluence of the M/E ratio on

    the minimum required batch time at 293 K,   W  = 5 kg,

    V  L  = 10 m3 is given in Table 10 while Fig. 2 shows the evolution

    of ethyl acetate conversion as a function of time for differentM/E

    ratios at the same reaction conditions. The minimum required

    batch time needed to reach 95% of the equilibrium conversion is

    reduced when M/E is increased. The yearly production is also

    increased. However, the absolute amount of product obtained perbatch is also reduced since the initial amount of ethyl acetate is

    lowerat higherM/E molarratios. This results in an increase of the

    number of batches and thus an increase of the time needed for

    emptying and reloading the reactor, thus resulting in an overall

    productivity loss. A more in-depth productivity analysis is

    thereforeneeded to determinethe optimal M/Eratio to maximize

    the profit.

    The effect of using concentrations instead of activities in the

    rate expressions, as discussed in Section 2.1.1, was found to be

    small around the optimal process conditions, i.e. 293 K and

    M/E = 10, as shown in  Table 10 (compare the first entry with

    the second entry). At M/E > 1, the activity coefficients of 

    methanol and ethanol are close to 1.0 and only a small influencecan thus be expected.

    A comparison has been made with homogeneously

    catalyzed transesterification. Schmidt et al. have proposed a

    T.F. Dossin et al. / Applied Catalysis B: Environmental 67 (2006) 136–148144

    Table 8

    Influence of the temperature on the annual production (simulation performed using Eqs. (3) and (5) at M/E = 10,  W  = 2 kg,  V  L  = 10 m3,  X EtOAc = 90%)

    Temperature (K) Minimum required batch time (h) MeOAc production (tonnes year1) batches year1

    293 51.1 233.5 171.3

    303 39.6 301.3 221

    313 31.8 375.1 275.2

    323 25.6 465.3 341.3

  • 8/20/2019 Sim ASPEN Biodiesel Cat MgO

    10/13

    kinetic model for the transesterification of butyl acetate withethanol catalyzed by potassium butanolate to form ethyl acetate

    and butanol [22]. The activation energy and the pre-exponential

    factor of the rate coefficient were derived from experimental

    data, based on a second-order rate law (Eq.   (25)) and

    accounting for the polarity of the reaction mixture:

    r BuOK  ¼ k BuOK 

    C BuOAcC EtOH

    C BuOHC EtOAc

    K eq

      (25)

    This model has been applied to the homogeneous catalyzed

    transesterification of ethyl acetate with methanol using the

    kinetic parameters reported by Schmidt et al.: activation energyand pre-exponential factor amount to 51 kJ/mol and 24.4 m6 / 

    mol2 s, respectively   [22]. Note that this activation energy is

    twice that for the heterogeneously catalyzed reaction used in

    this paper. A simulation has been performed at the same process

    conditions as the heterogeneous catalyzed industrial transes-

    terification, i.e. 293 K, M/E = 10,   V  L  = 10 m3 integrating

    Eq.   (5)   using Eq.   (25)  as rate equation. In order to achieve

    the same yearly production of 500 tonnes year1, 0.150 mol/m3

    catalyst concentration is needed, i.e. 0.170 kg of potassium

    butanolate catalyst for a comparable batch time and a 10 m 3

    reactor volume. The required mass of homogeneous catalyst is

    thus almost 30 times smaller than the mass of solid magnesium

    oxide catalyst to achieve the same production. However,

    the lower cost of the solid magnesium oxide catalyst, the

    elimination of the reaction mixture neutralization step and

    the easy catalyst separation from the reaction mixture make

    the heterogeneous process a promising alternative to the

    homogeneous process.

    3.2. Continuous slurry reactor for biodiesel production

    Simulations were performed of the industrial scale biodiesel

    production in a continuously perfectly mixed slurry reactor by

    transesterification of triolein with methanol using an hetero-geneous MgO catalyst. The influence of the temperature, the

    space time and the MeOH/triolein feed molar ratio (M/T ratio)

    on the production has been investigated to achieve about

    100,000 tonnes of methyl oleate per year. As discussed in

    Section 2.2.3, the results are to be interpreted in a qualitative way

    only, since the equilibrium conversion to the desired product

    methyl oleate is much lower than the ones found experimentally.

    In order to obtain a high productivity, the amount of catalyst

    was chosen quite high: 5700 kg. In a total liquid volume of 

    25 m3, the volume fraction of magnesium oxide catalyst

    amounts to about 0.091. According to Eq. (4), with this catalyst

    fraction, the impeller revolution speed should be at least0.76 s1 at the standard conditions: 323 K, a M/T molar ratio of 

    10, a space time of 1000 kg-cat s/mol-triolein, and a catalyst

    particle size of 25 mm. Therefore, an impellor speed of 1.0 s1

    seems sufficient. However, impeller revolutions of 5 and 10 s1

    T.F. Dossin et al. / Applied Catalysis B: Environmental 67 (2006) 136–148   145

    Table 9

    Influence of the catalyst mass on the annual production (simulation performed using Eqs.  (3) and (5) at  T  = 293 K, M/E = 10,  V  L  = 10 m3,  X EtOAc = 90%)

    W  (kg) Minimum required batch time (h) MeOAc production (tonnes year1) Batches year1

    2 51.1 233.5 171.3

    4.5 22.7 525.5 385.4

    5 20.4 583.9 428.2

    10 10.2 1167.8 856.515 6.8 1751.6 1284.7

    Table 10

    Influence of the M/E molar ratio on the annual production (simulation performed using Eqs.  (3) and (5) at  T  = 293 K, W  = 5 kg, V  L  = 10 m3, X EtOAc = 95% of  X eq)

    M/E Minimum required batch time (h) MeOAc production (tonnes year1) Batches year1

    10 20.4 583.9 428.2

    10a 19.9a 601.4a 441.1a

    15 9.9 873.8 883.3

    20 5.7 1163.3 1545.8

    30 2.8 1637.9 3122.1

    a

    Simulations using rate Eq.  (3a) with concentrations instead of activities.

    Fig. 1. Conversion of ethyl acetate vs. batch time at different temperatures, at

    M/E = 10,   W  = 2 kg,   V  L  = 10 m3: (—) 293 K; ( ) 303 K; (– – –) 313 K;

    (– - – -) 323 K. The simulated values were obtained from Eqs.  (3) and (5)

    using the parameter values of  Table 3.

  • 8/20/2019 Sim ASPEN Biodiesel Cat MgO

    11/13

    were chosen in order to achieve a homogeneous mixture (or amicro-emulsion) of the methanol and glycerides.

    The reaction rates were calculated using Eqs.  (12a)–(12c).

    The conversion of the triolein, together with the yearly

    production of methyl oleate, is plotted in Fig. 3 as a function of 

    space time at the standard set of reactor conditions: M/T feed

    ratio = 10,   T  = 323 K,   d  p = 25 mm,   V  L  = 25 m3, triolein feed

    molar flow = 5.0 mol s1,   W cat = 5700 kg,   N  I  = 5 s1. As

    expected, the conversion increases with space time until

    equilibrium conversion is reached. The yearly production of 

    methyl oleate decreases strongly with increasing space time

    due to the decrease of the reaction rate upon approaching the

    equilibrium composition. Working at short space times willthus result in a high production; however, the quality of the

    biodiesel will be worse due to the larger amounts of unreacted

    triolein and diolein. This can be seen in  Fig. 4   in which the

    yields of the products are plotted as a function of space time.

    For the explorative simulations discussed in this paper, it was

    assumed that a production of 100,000 tonnes methyl oleate per

    reactor per year is typical. Additionally, it was assumed that theminimum required triolein conversion and methyl oleate yields

    are close to the values shown in Table 11. Of course, a thorough

    economic evaluation is needed to determine the real optimum

    operation conditions, which includes the maximum allowed

    contents of remaining glycerides in the biodiesel according to

    the ASTM or DIN standards, and also the value of the

    byproduct glycerine. Currently glycerine is still a valuable

    product that can be used in several industrial applications;

    however, its value will decrease once the biodiesel production

    takes off on a large scale.

    The influence of temperature, M/T ratio, particle diameter,

    and impeller revolution speed on conversions and yields wereinvestigated and some of these results are presented in Table 11.

    According to the simulations, a production of 100,000 tonnes

    methyl oleate is achieved at a space time of 871 kg s/mol-

    triolein at 323 K, M/T feed ratio = 10,   d  p = 25 mm and

     N  I  = 5 s1. A triolein conversion of 93% is achieved while

    methyl oleate and glycerine yields amount to 0.54 and 0.05,

    respectively. At the same conditions, according to Eq.   (24),

    944 tonnes of glycerine per year are produced. Separation from

    methyl oleate is easy since glycerine is practically immiscible

    with methyl oleate and therefore can be recovered by settling

    (decantation) [7]  or centrifugation [46].

    Increasing theM/T ratio from 10 to 20 results in a significantly

    higher triolein conversion and also higher yields of methyl oleateand glycerine (compare first entry with the second entry in

    Table 11) as was also pointed out by Boocock  [46]. These yields

    rise from 0.54 and 0.05 to 0.62 and 0.10, respectively. However,

    this increase of M/Tratio also increasesmethanol flow, leading to

    a larger effort to recycle the methanol, as pointed out by

    Connemann et al. [7] and Zhang et al.  [8]. The extra cost for

    separating the methanol from the reaction mixture by means of 

    distillation will thus compensate the advantage of the higher

    yields of methyl oleate and glycerine. An economic evaluation

    will thus be needed to estimate the optimal M/T ratio.

    It appears that the influence of the temperature, as well as

    that of the catalyst particle size on the reactor performance is

    T.F. Dossin et al. / Applied Catalysis B: Environmental 67 (2006) 136–148146

    Fig. 2. Conversion of ethyl acetate vs. batch time at different M/E ratios, at

    T  = 293 K, W  = 5 kg, V  L  = 10 m3: (—) M/E = 10; ( ) 15; (– – –) 20; (– - – -)

    30. The simulated values have been obtained using Eqs.  (3) and (5). The

    simulated values were obtained from Eqs. (3)and (5) using the parameter values

    of  Table 3.

    Fig. 3. Conversion of triolein (—; left axis) and the yearly production (kilo-

    tonnes year1) of methyl oleate (– – –; right axis) and glycerine ( ; right axis)as a function of space time (kgcat   s/mol-triolein) at 323 K, M/T = 10,

    W  = 5700 kg,  V  L  = 25 m3,  N  I  = 5 s

    1,  d  p = 25 mm. The simulated values were

    obtained using Eqs.  (11a)–(11c),   (13),   (15), and  (20), and using the kinetic

    parameter values from Table 3.

    Fig. 4. Yields of diolein (– – –), monoolein (– - – -), glycerine ( ), andmethyl oleate (—) as a function of space time (kgcat   s/mol-triolein) at

    323 K, M/T = 10, W  = 5700 kg, V  L  = 25 m3, N  I  = 5 s

    1, d  p = 25 mm. The simu-

    lated values were obtained using Eqs.  (11a)–(11c),  (13),   (15), and  (20), and

    using the kinetic parameter values from  Table 3.

  • 8/20/2019 Sim ASPEN Biodiesel Cat MgO

    12/13

    very small. This is due to the rather close approach to

    equilibrium, which is neither influenced by the temperature,

    nor by the particle size. However, since a rather strong diffusion

    limitation of triolein occurs at low conversion, the use of smaller

    catalyst particles will significantly increase the initial conversion

    rate whereas the temperature has only a minor effect. The strong

    diffusion limitation of triolein at low conversion is due to the

    combination of the relatively fast reaction rate and the fact that

    the reaction rate is zeroth order with respect to triolein. This

    zeroth order causes the rate to remain high if the triolein

    concentration already decreases due to the diffusion limitation.Obviously, both external and internal transport limitations

    disappear when equilibrium is approached.

    Also the impeller revolution speed has very little influence

    on the reactor performance since the conversion rate is not

    significantly influenced by external mass transfer limitation

    (i.e. transfer from the liquid bulk towards the catalyst particles).

    Nevertheless, it remains crucial to check whether the stirring

    rate is really sufficient to achieve a single-phase system or

    micro-emulsion system and a good suspension of the catalyst

    particles in the reactor.

    The effect of using concentrations instead of activities in the

    rate expressions, as discussed in Section 2.2.2 was found to besmall at all conditions considered. At the standard conditions the

    calculated effect is shown in  Table 11 (compare the first entry

    with the third entry). Apparently, the net effect of the activity

    coefficients on thevalue of thedenominators of therate equations

    is small. The numerator of the rate equations hardly changes

    since the activity coefficient of methanol is always close to 1.0.

    As for the EtOAc/MeOH-case, the biodiesel production

    catalyzed by homogeneous catalysts, based on data from Zhang

    et al. [8] and Connemann et al. [45], has been compared to the

    heterogeneously catalyzed transesterification. According to

    Zhang et al., a stream of 10 kg/h of sodium hydroxide is needed

    to reach a production of 8000 tonnes biodiesel year1 at 333 K 

    with M/E = 6. This leads to a yearly catalyst consumption of 88tonnes. Based on Connemann’s data on existing biodiesel

    plants, which typically operate at 333–348 K, 3–4 kg catalyst

    per tonne of biodiesel is needed to reach a yearly production of 

    80,000–125,000 tonnes of biodiesel. This thus requires more

    than 300 tonnes of homogeneous base catalyst. These amounts

    are much larger than the estimated amount of 5.7 tonnes of 

    heterogeneous MgO catalyst that follows from the simulations

    in this paper. In contrast to the batch reactor simulation for the

    EtOAc/MeOH-case, the heterogeneous catalyst thus seems to

    be much more efficient for biodiesel production than the

    homogeneous catalysts. Although this is an encouragement to

    intensify the research on the heterogeneously catalyzed system,

    it must be interpreted with care in view of the assumptions and

    simplifications made in the explorative simulations presented

    here. Particularly the assumption of the reaction kinetics to be

    exactly the same as for the transesterification of ethyl acetate

    may have a major impact. On the other hand it might be that the

    catalytic activity of the KOH catalyst in the homogeneous

    process is rather low. Despite the uncertainties in the

    explorative simulation results of the biodiesel production

    process, they offer a useful starting point towards further

    investigation of the industrial biodiesel production using MgO

    catalysts. Moreover, these simulation results indicate thatheterogeneous catalysts may allow good production levels and

    present an interesting alternative to homogeneous catalysts

    towards a more environment-friendly process.

    4. Conclusions

    Magnesium oxide offers a viable heterogeneous solid base

    catalyst for the standard transesterification of ethyl acetate or

    triolein with methanol at industrial conditions for the industrial

    production of fine-chemicals or biodiesel. Simulation of batch

    and continuous slurry reactors, based on an Eley–Rideal kinetic

    model, show that industrial production of fine-chemicals can bereached at ambient temperature and atmospheric pressure using

    finely dispersed MgO catalyst. This allows the development of 

    a more environmentally friendly chemical process for the

    transesterification reaction of low and high methyl acetate

    production volumes since reaction neutralization is avoided and

    product separation is made much easier. Also, the use of MgO

    makes the batch process cost-effective since reactor heating is

    not required.

    Explorative process simulations for the continuous produc-

    tion of biodiesel indicate that commercially attractive produc-

    tion rates can be achieved. However, further experimental

    investigation is needed to quantitatively determine the potential

    of MgO as an alternative towards a more sustainabledevelopment of the industrial biodiesel production process.

    Acknowledgement

    This research has been funded in the frame of the STWW

    project ‘‘HYPCAT’’ (STWW 141) of the IWT of the Flemish

    Government.

    References

    [1] F. Ma, M.A. Hanna, Biodiesel, Bioresour. Technol. 70 (1999) 1.

    [2] J. Otera, Chem. Rev. 93 (1993) 1449.

    T.F. Dossin et al. / Applied Catalysis B: Environmental 67 (2006) 136–148   147

    Table 11

    Space time,yields, and conversions obtained at a production rate of 100,000 tonnes methyl oleate per year as a function of the methanol/triolein feed ratio and the use

    of concentrations instead of activities in the rate equations (Eqs.  (12a)–(12c)) instead of Eqs.  (11a)–(11c))

    T  (K) M/T   d  p  (mm)   N  I  (s1)   W  / F tot  (kg s/mol)   Y MeOl   Y G    Y  D   Y  M    X T    X MeOH

    323 10 25 5 870.6 0.54270 0.05002 0.28953 0.59090 0.92938 0.16215

    323 20 25 5 1013.5 0.62279 0.10319 0.18789 0.68123 0.97137 0.09300

    323 10 25 5 851.6 0.52718a 0.04366a 0.30708a 0.56878a 0.91843a 0.15757a

    a Simulations using Eqs. (12a)–(12c) with concentrations instead of activities.

  • 8/20/2019 Sim ASPEN Biodiesel Cat MgO

    13/13

    [3] U. Meyer, W.F. Hoelderich, Appl. Catal. A 178 (1999) 159.

    [4] A. Corma, S. Iborra, S. Miquel, J. Primo, J. Catal. 173 (1998) 315.

    [5] J. Barrault, Y. Pouilloux, J.M. Clacens, C. Vanhove, S. Bancquart, Catal.

    Today 75 (2002) 177.

    [6] D. Darnoko, M. Cheryan, J. Am. Oil Chem. Soc. 77 (12) (2000) 1269.

    [7] J. Connemann, A. Krallmann, E. Fischer, Process for the continuous

    production of lower alkyl esters or higher fatty acids, US Patent 5354878,

    1993.

    [8] Y. Zhang, M.A. Dubé, D.D. McLean, M. Kates, Bioresour. Technol. 89(2003) 1.

    [9] N. Ergün, P. Panning, Method for producing fatty acid methyl esters and

    equipment for realizing the same, US Patent 6440057, 2002.

    [10] M. Canonge, J.-C. Joly, Process for the continuous preparation of acetates,

    Patent EP0280471A1, 1988.

    [11] A.F. Eastgate Sims, J. Southwood, Production of acrylic and methacrylic

    esters, Patent GB960005, 1961.

    [12] Y.-J. Fang, W.-D. Xiao, Sep. Purif. Technol. 34 (2004) 255.

    [13] L. Jimenez, A. Garvin, J. Costa-Lopez, Ind. Eng. Chem. Res. 41 (2002)

    6663.

    [14] B. Davies, G.V. Jeffreys, Trans. Inst. Chem. Eng. 51 (1973) 271.

    [15] B. Davies, G.V. Jeffreys, Trans. Inst. Chem. Eng. 51 (1973) 275.

    [16] B. Freedman, R.O. Butterfield, E.H. Pryde, J. Am. Oil Chem. 63 (10)

    (1986) 1375.

    [17] S. Gryglewicz, Appl. Catal. A 192 (2000) 23–28.[18] H. Hattori, M. Shima, H. Kabashima, Stud. Surf. Sci. Catal. 130 (2000)

    3507.

    [19] M. Di Serio, R. Tesser, A. Ferrara, E. Santacesaria, J. Mol. Catal. A (2004)

    251.

    [20] U. Schuchardt, R. Sercheli, R.M. Vargas, J. Braz. Chem. Soc. 9 (1) (1998)

    199.

    [21] L. Bournay, D. Casanave, B. Delfort, G. Hillion, J.A. Chodorge, Catal.

    Today 106 (2005) 190.

    [22] J. Schmidt, D. Reusch, K. Elgeti, R. Schomäcker, Chem. Ing. Tech. 71

    (1999) 704.

    [23] H. Noureddini, D. Zhu, J. Am. Oil Chem. Soc. 74 (11) (1997) 1457.

    [24] D. Darnoko, M. Cheryan, J. Am. Oil Chem. Soc. 77 (12) (2000) 1263.

    [25] T.F. Dossin, M.-F. Reyniers, G.B.M. Marin, Appl. Catal. B Environ. 61

    (2006) 35.

    [26] A. Demirbas, Energy Conversion Manage. 44 (2003) 2093.[27] ASPEN Engineering SuiteTM 11.1 version, Aspen Technology Inc.

    [28] C. Reid, J.M. Prausnitz, D.E. Poling, The Properties of Gases and Liquids,

    third and fourth ed., McGraw-Hill, New York, 1987.

    [29] R.J. Madon, I. Iglesia, J. Mol. Catal. A Chem. 163 (2000) 189.

    [30] G.B. Tatterson, Fluid Mixing and Gas Dispersions in Agitated Tanks,

    McGraw-Hill, New York, 1991.

    [31] D. Thoenes, Chemical Reactor Development, Kluwer Academic Publish-

    ers, Dordrecht, The Netherlands, 1994, p. 91.

    [32] F. Kapteijn,G.B. Marin, J.A. Moulijn, Catalytic Reaction Engineering: An

    Integrated Approach of Homogeneous, Heterogeneous and IndustrialCatalysis, Elsevier, 1993 (Chapter 7 in Catalysis).

    [33] G.F. Froment, K. Bischoff, Chemical Reactor Analysis and Design, J.

    Wiley & sons, 1990.

    [34] D.E. Mears, J. Catal. 20 (1971) 127.

    [35] H. Duński, W.K. Jóźwiak, H. Sugier, J. Catal. 146 (1994) 166.1.

    [36] B.C. Gates, Catalytic Chemistry, Wiley, New York, 1992.

    [37] L.R. Petzold, SIAM J. Stat. Comput. 4 (1983) 136.

    [38] Netlib,  http://www.netlib.org.

    [39] J.M. Encinar, J.F. Gonzalez, A. Rodriguez-Reinares, Ind. Eng. Chem. Res.

    44 (2005) 5491.

    [40] Y. Sano, N. Yamaguchi, T. Adachi, J. Chem. Eng. Jpn. 7 (1974) 255.

    [41] F. Kapteijn, J.A. Moulijn, in: G. Ertl, H. Knözinger, J. Weitkamp (Eds.),

    Laboratory Reactors: Handbook of Heterogeneous Catalysis, vol. 3, VCH

    Verlagsgesellschaft mbH, Weinheim, Germany, 1997, p. 1359 (Chapter 9).

    [42] C.S. Caldwell, A.L. Babb, J. Phys. Chem. 60 (1956) 51.[43] R.H. Perry, D.W. Green, J.O. Maloney, Perry’s Chemical Engineers’

    Handbook, seventh ed.,McGraw-Hill,NewYork,1997(Chapter5, pp.5–52).

    [44] D. Bockey, Situation and Development Potential for the Production of 

    Biodiesel—An International Study, Union zur Förderung von Oel und

    Proteinpflanzen e.V., 2002.

    [45] J. Conneman, J. Fischer, International Liquid Biofuels Congress, Curitiba,

    Brazil, July 19–22, 1998.

    [46] D.G.B. Boocock, Process for productivity of fatty acid methyl esters from

    fatty acid triglycerides, US Patent No. 6,712,867 B1, March 30, 2004.

    [47] X. Lang, A.K. Dalai, N.N. Bakhshi, M.J. Reaney, P.B. Hertz, Bioresour.

    Technol. 80 (2001) 53.

    [48] G. Knothe, K.R. Steidley, Fuel 84 (2005) 1059.

    [49]   http://www.patechfc.com.tw/p2e.htm, November 2005.

    [50] G. Knothe, Fuel Proc. Technol. 86 (2005) 1059.

    [51] C.W. Weast, CRC Handbook of Chemistry and Physics, 63rd ed., CRCPress, Boca Raton, USA, 1982–1983.

    T.F. Dossin et al. / Applied Catalysis B: Environmental 67 (2006) 136–148148

    http://www.netlib.org/http://www.patechfc.com.tw/p2e.htmhttp://www.patechfc.com.tw/p2e.htmhttp://www.netlib.org/