Pulgarin_Leon Innovaquimica

  • Upload
    landa

  • View
    218

  • Download
    0

Embed Size (px)

Citation preview

  • 8/12/2019 Pulgarin_Leon Innovaquimica

    1/10

    ANLISIS DE SENSIBILIDAD EN UN REACTOR DE CRAQUEO CATALTI-CO DE METANO PARA LA PRODUCCIN DE HIDRGENO

    J. Pulgarn-Len1, J. Saavedra-Rueda2, A. Molina1,*1 Bioprocesos y Flujos Reactivos, Facultad de Minas, Universidad Nacional de Colombia, Sede Medelln

    2 Instituto Colombiano del Petrleo, Ecopetrol, Bucaramanga

    Resumen:Se desarroll un modelo de un reactor de craqueo cataltico de metanopara la produccin de hidrgeno utilizando un mecanismo que simplifica las

    mltiples reacciones heterogneas que suceden en el reformado cataltico con tresreacciones que incluyen la adsorcin en sitios activos y el equilibrio qumico. El

    modelo considera al reactor como un lecho empacado y simplifica la transferencia

    de masa en el catalizador considerando las reacciones como pseudohomogneas.

    Para el desarrollo del modelo se utiliz el software comercial ASPEN Plus y se es-

    cribi un mdulo de usuario que utiliza una rutina en FORTRAN escrita especfi-camente para representar la cintica qumica. Las predicciones del modelo son si-

    milares a las tendencias reportadas en la literatura y en un reformador industrial. El

    artculo concluye utilizando el modelo para entender el efecto de la variacin de latemperatura, relacin vapor/carbono y tiempo de residencia sobre la conversin y

    el rendimiento fraccional global para el hidrgeno.

    Palabras Clave:Reformado de metano, Gas natural, Hidrgeno, ASPEN PLUS.

    SENSITIVITY ANALYSIS OF A METHANE STEAM REFORMER

    J. Pulgarn-Len1, J. Saavedra-Rueda2, A. Molina1,*1 Bioprocesos y Flujos Reactivos, Facultad de Minas, Universidad Nacional de Colombia, Sede Medelln

    2 Instituto Colombiano del Petrleo, Ecopetrol, Bucaramanga

    Abstract:A model that predicts the conversion in a methane steam reformer for

    hydrogen production was developed. The model makes use of a three-reactionmechanism that simplifies the multiple heterogeneous reactions involved in re-

    forming that include adsorption on catalyst active sites and chemical equilibrium.

    The model considers the reformer as a packed bed and simplifies mass transfer in

    the catalyst by considering reactions as pseudohomogeneous. The model was de-

    veloped using a FORTRAN subroutine specially written for the former that waslinked to the commercial software ASPEN Plus. Model predictions are similar to

    those found in the literature and with results from an industrial reformer. The pa-

    per ends with a sensitivity analysis that uses the model to understand the effect that

    the variation of temperature, steam to carbon ratio and residence time have on me-

    thane conversion and on fractional conversion of hydrogen.

    Keywords:Methane steam reforming, Natural Gas, Hydrogen, ASPEN PLUS.

    *Corresponding author: [email protected]

    1. IntroductionDespite the recent interest in different

    technologies for hydrogen production,

    such as biomass and coal gasification,

    fluidized bed membrane reactors, fermen-

    tative routes and water electrolysis; me-

    thane steam reforming remains the major

    production route of hydrogen. Although

    this interest is mainly motivated by the

    role that hydrogen can play as an energy

    resource, hydrogen production for the

    petrochemical, chemical and food industry

    will remain a major activity of chemical

    engineers.

    In methane steam reforming, hydrogen is

    produced in a furnace in which a mixture

    of methane and water reacts in tubes iso-

    lated from the combustion gases. With thehelp of a catalyst (typically with Ni as

    major component) a highly endothermic

  • 8/12/2019 Pulgarin_Leon Innovaquimica

    2/10

    Pulgarn et al., Innovaqumica, 2009

    2

    reaction converts methane and water into

    a mixture of hydrogen and other com-

    pounds, with carbon monoxide as major

    component.

    There is a vast literature on modeling

    steam reformers. [1-22]. Most studies

    date to the late 80s. In the 90s and early

    21st century the number of papers in the

    open literature on modeling methane

    steam reformers decreased, probably be-

    cause of licensing agreements between

    production companies and research insti-

    tutions as well as an increased interest in

    alternative technologies for reforming

    such as Fluidized Bed Membrane Reac-tors [10, 13, 16-19], optimized control

    strategies [12, 14, 15] and novel catalysts

    formulations [11, 21].

    One of the most refereed studies on me-

    thane reformers was the work by Xu and

    Froment [5, 7] that describes a model for

    these reactors and carefully compares

    experimental data with predictions. In

    Colombia, as early as in 1981, Ros de-

    veloped a model that predicted hydrogen

    production and methane conversion in anindustrial methane steam reformer [1].

    More recently Acua [14], in order to

    improve the controllability of steam re-

    formers, developed a detailed model of

    the methane steam reforming process.

    This paper presents preliminary results on

    a model that predicts the conversion of

    steam and methane in a reformer. Model

    details, such as kinetic expressions and

    input parameters are defined. The resultsfrom the model are validated against pre-

    vious literature results, and data from an

    industrial reformer. The paper ends with a

    sensitivity analysis that allows an addi-

    tional test to the model and further under-

    standing of the methane steam reforming

    reaction. The results show that the model

    can correctly predict most of the trends

    observed in an industrial reformer.

    2. ModelTo simulate the conversion of methane to

    hydrogen in a catalytic reactor it is neces-

    sary to consider three elements: (1) the

    kinetic expressions that describe the rateof methane and water reaction on the cata-

    lyst surface to produce hydrogen and other

    byproducts, (2) the mathematical expres-

    sions that address the mole, momentum

    and energy balances in the reactor and (3)

    the process parameters (both operational

    and geometrical) that characterize the

    reactor.

    2.1 Kinetic data

    As mentioned in the introduction, bothRos [1] and Acua [14] modeled methane

    steam reformers. While Ros fitted expe-

    rimental data to obtain the kinetic parame-

    ters for his model, Acua [14] used the

    kinetic expressions described by Haldor

    [23], as quoted by Acua [14], that date

    from 1965. The present research aimed at

    making use of more contemporary kinetic

    data.

    A review of previous literature showed

    that the model proposed by Xu and Fro-ment [7] in 1989 is the one more quoted in

    the literature. Although it is an old model,

    it was recently successfully used by Za-

    maniyan et al. [20] and Alberton et al.

    [22] in the simulation of industrial me-

    thane reformers.

    Table 1shows the reactions considered by

    Xu and Fromment [7].

    These authors came to the conclusion that

    reactions R 6 to R 11 are not important

    because the ratios of forward to reverse

    reaction rates are always higher than 1 at

    typical steam reforming conditions. Reac-

    tions R 4 to R 5 are also considered of

    minor importance because they, although

    thermodynamically favored at typical

    reforming conditions, describe trends

    (lower CO2concentration as the length in

    the reactor increases) that do not agree

    with normal reactor performance. Al-

  • 8/12/2019 Pulgarin_Leon Innovaquimica

    3/10

    Pulgarn et al., Innovaqumica, 2009

    3

    though these are rather weak assumptions

    as local conditions inside the reformer

    may have concentrations that will thermo-

    dynamically favor forward reactions, as a

    first approximation in this paper, we

    adopted Xu and Fromment [7] recom-

    mendations and simply considered R 1 to

    R 3 as the most important reactions in the

    process.

    Table 1. Reactions involved in methane steam reforming as reportedby Xu and Fromment [7]

    Reaction H298(kJ/mol)

    R 1224 H3COOHCH 206.1

    R 2222 HCOOHCO 41.15

    R 3224 H4COOH2CH 165.0

    R 4224 H2CO2COCH 247.3

    R 5 OH2CO4CO3CH 224

    330.0R 6

    24 H2CCH 74.82R 7

    2COCCO2 173.3

    R 8 OHCHCO 22 131.3R 9 OH2CH2CO 222 90.13

    R 10 OH2C3CO2CH 24 187.6R 11 OH2C2COCH 224 15.3

    Xu and Fromment [7] carried out a bal-

    ance of active sites on the catalysts for aheterogeneous mechanisms they devel-

    oped to simulate R 1 to R 3. From this

    balance these authors derived three rate

    expressions shown in E 1 toE 3 that con-

    sider the partial pressure of hydrogen,

    methane, water and carbon monoxide, (Pi)

    as well as kinetic (ki), equilibrium (Ki) and

    adsorption constants (Kj, with j represents

    CO, H2, CH4, H2O). E 5 toE 7present the

    mathematical expressions used to

    represent the variation of the constant withtemperature. Table 2 to Table 4 show the

    kinetic constants finally used in the model.

    Equilibrium constants were calculated

    using the free-domain code for gaseous

    equilibrium Gas-Eq [24].

    2

    1

    CO

    3

    OHCH5.2

    2H

    1

    1DEN

    K

    pppp

    p

    k3

    r

    2H

    24

    E 1

    2

    2

    COH

    OHCO2H

    2

    2DEN

    K

    pp

    ppp

    k

    r

    22

    2

    E 2

    2

    3

    CO

    4

    4

    4CH5.3

    2H

    3

    3DEN

    K

    pppp

    p

    k4

    r

    22H

    O2H

    E 3

    2

    2

    244

    22

    H

    OH

    OHCHCH

    HHCOCO

    p

    pKpK

    pKpK1

    DEN E 4

    ir

    T,iiT

    1

    T

    1

    R

    Eaexpkk

    E 5

    ir

    j

    T,iiT

    1

    T

    1

    R

    HexpKK

    E 6

    i

    ibi

    ii,eqT

    EaexpTAK

    E 7

  • 8/12/2019 Pulgarin_Leon Innovaquimica

    4/10

    Pulgarn et al., Innovaqumica, 2009

    4

    Table 2. Kinetic constants used in the model(adapted from [7])

    ki, 648K Eai

    (kmol bar1/2

    / kg cat h) (kJ/mol)

    k1 1.84 10-4 240.1

    k2 7.558 67.13

    k3 2.19 10-5

    243.9

    Table 3. Adsorption constants used in the model(adapted from [7])

    Kj TrH

    (bar-1

    ) (K) (kJ/mol)

    CH4 0.1791 823 -38.28

    H2O 0.4152* 823 88.68CO 40.91 648 -70.65

    H2 0.02960 648 -82.90

    *Dimensionless

    Table 4. Equilibrium constants used in the mod-el (calculated from[24])

    Ai (bar2) bi Eai (K

    -1)

    Keq1 4.94106 1.89 -25216

    Keq2 1.2010-6

    * 1.24 5425

    Keq3 5.93 3.13 -19791

    *Dimensionless

    2.2 Balance equations in the reactorAs one of the objectives of the paper was

    to perform a sensitivity analysis of the

    methane steam reformer and given the

    extended use of the commercial software

    ASPEN in the Chemical Engineering in-

    dustry, this research employed ASPEN

    PLUS V 7.0, as simulation tool. Unfortu-

    nately, as it happens with most commer-

    cial software, the presence of non-conventional kinetic expressions, (e.g.E 1

    to E 3), impose the use of user-defined

    models in ASPEN.

    There are typically two possibilities when

    modeling non-conventional rate expres-

    sions in ASPEN. The first one, as illu-

    strated by Barrera et al. [25], consists in

    linking ASPEN to Excel and developing

    the kinetic expressions through Excel by a

    spreadsheet. The second one, adopted in

    this research, consist on defining FOR-

    TRAN-based subroutines. We consider the

    second approach more suited for our re-

    search needs as FORTRAN allows more

    complex calculations that Excel.

    As this was the first time that FORTRAN

    was linked with ASPEN in our research

    group, we developed our own code to

    model a PFR reactor using MATLAB to

    validate the ASPEN/FORTRAN approach.

    Both programs (MATLAB and AS-

    PEN/FORTRAN) yielded identical results.

    Although the methane steam reformer

    simulations by Ros [1] and Acua [14]

    considered energy and momentum bal-ances inside the reactor, this research used

    predefined pressure and temperature pro-

    files as inputs to the model. This was done

    because emphasis was placed in the kinet-

    ic approach and, as a first step; it was con-

    sidered more important to address the

    chemistry. Current work conducted in our

    research group, is incorporating energy

    and momentum expressions to the model.

    In the treatment of the reactor in ASPENwe adopted a PFR formulation and consi-

    dered did not consider heterogeneous limi-

    tations (pseudo-homogeneous approach).

    This is clearly an oversimplification, al-

    though frequently adopted in the literature

    (e.g. [1, 14]). Future developments of the

    FORTRAN subroutine will include mass

    transfer to the particle with approaches

    similar to those described by Xu and

    Fromment [5], Zamaniyan et al. [20] and

    Alberton et al. [22]

    2.3 Process and reactor parametersAlthough, as explained above, the AS-

    PEN/FORTRAN reactor model was vali-

    dated with an independent program de-

    signed by our research group, additional

    validation of the kinetic expressions was

    mandatory before simulation of an indus-

    trial furnace could be undertaken. This

    kinetic validation was carried out using the

    data presented by Zamaniyan et al. [20]. A

  • 8/12/2019 Pulgarin_Leon Innovaquimica

    5/10

    Pulgarn et al., Innovaqumica, 2009

    5

    comparison of our results with those of

    Zamaniyan et al. [20] is presented in the

    Results section.

    In the simulations used to validate our

    model against that of Zamaniyan et al.

    [20], we employed the same input parame-

    ters (temperature and pressure profile,

    geometry, input stream gas composition,

    temperature and pressure profiles) that

    Zamaniyan describe in their paper [20].

    For the simulation of the industrial me-

    thane steam reformer a 15-m length, 12.6-

    cm internal diameter reactor was consi-

    dered. Table 5 shows the input-stream

    composition as supplied by the reformeroperator. O2 and N2 were considered as

    inert gases, given their low concentrations.

    The concentrations of ethane, propane and

    butane were added to that of methane fol-

    lowing the equal-carbon concept described

    by Zamaniyan et al. [20].

    Table 5. Input-stream composition

    Species CH4 H2O O2 N2Molar

    fraction0.1384 0.8424 0.0003 0.0017

    Species C2H6 C3H8 C4H10Molar

    fraction0.0143 0.0027 0.0002

    As only input and output data were availa-

    ble for temperature and pressure, the tem-

    perature and pressure profiles reported for

    an industrial methane steam reformer in

    the literature [20] were adjusted to obtain

    the same input and output values as those

    reported by the reformer operator. Figure1 compares the temperature and pressure

    profiles used in this paper for the industri-

    al reformer with those of Zaminiyan et al.

    [20]. It is evident that the use of these

    profiles is a rather crude simplification.

    Nevertheless, as it will be shown below, it

    turned out to be an acceptable approxima-

    tion.

    Because, mass-transfer limitations were

    not modeled, no data on the catalyst is

    given different from the fact that it was a

    Ni/Al2O3 catalyst and the density was as-

    sumed to be 2355.5 kg m-3, as reported in

    [20].

    3.Results and discussion3.1 Model validationFigure 2 compares the profiles of major

    species (CH4, H2O, H2and CO) along the

    reactor predicted by Zamaniyan et al. [20]

    with those obtained using our model. Al-

    though there are some clear differences in

    trends, exit concentrations are very similar

    for all species. Particularly the two more

    important species, H2and CH4, show simi-

    lar concentrations at the reactor exit. This

    gives confidence on the model ability tocorrectly predict the final result for the

    reactor.

    600

    800

    1000

    1200

    0 2 4 6 8 10 12 14distance along reactor (m)

    T(K)

    Zamaniyan et al.

    this paper

    a.

    11

    12

    13

    14

    15

    16

    17

    0 2 4 6 8 10 12 14distance along reactor (m)

    P(bar)

    Zamaniyan et al.

    this paper

    b.

    Figure 1. Comparison of (a) temperature and(b) pressure profiles along the reactor. Datafrom Zamaniyan et al. [20] and this paper

    Contrary to the relative success in predict-

    ing the final concentrations,Figure 2 a. to

    d. show that our model tends to overesti-

    mate methane conversion. This is evident

    in Figure 2 because the predictions by

    Zamaniyan et al. [20], show that methane

    concentration is always higher and hydro-

    gen concentration lower than our model

  • 8/12/2019 Pulgarin_Leon Innovaquimica

    6/10

    Pulgarn et al., Innovaqumica, 2009

    6

    0.0

    0.1

    0.2

    0.3

    0 2 4 6 8 10 12

    distance along reactor (m)

    CH4

    (molfraction)

    a.

    Zamaniyan et al.

    this paper

    0.2

    0.3

    0.4

    0.5

    0 2 4 6 8 10 12

    distance along reactor (m)

    H2O

    (molfraction)

    b.

    Zamaniyan et al.

    this paper

    0.0

    0.1

    0.2

    0.3

    0 2 4 6 8 10 12

    distance along reactor (m)

    H

    2(molfraction)

    Zamaniyan et al.

    this paper

    c.

    0.0

    0.1

    0.2

    0.3

    0 2 4 6 8 10 12

    distance along reactor (m)

    CO(molfraction)

    d.

    Zamaniyan et al.

    this paper

    Figure 2. Comparison of concentration profiles as predicted by the model described by Zamaniyan et al. [20] and by this paper. (a)CH4, (b), H2O, (c) H2and (d) CO.

  • 8/12/2019 Pulgarin_Leon Innovaquimica

    7/10

    Pulgarn et al., Innovaqumica, 2009

    7

    results, at short distances inside the reac-

    tor. This result is not surprising as Zama-

    niyan et al. [20] considered mass-transfer

    limitations inside the catalyst (using an

    effectiveness factor) and our model consi-dered a pseudo-homogeneous approach.

    Neglecting mass transfer limitations caus-

    es the observed overestimation of methane

    conversion at short distances in the reac-

    tor. As stated above, future model ver-

    sions should consider mass transfer limita-

    tion in the catalyst.

    3.2 Simulation of industrial methanesteam reformer.

    Once the model was validated againstliterature data, the ability of the model to

    predict the behavior in an industrial re-

    former was addressed. Table 6 compares

    the exit concentrations (in mole fractions)

    as reported by the reformer operator with

    model predictions. Although there are

    some differences, particularly in CO, the

    agreement between predictions and the

    data for the industrial reformer is deemed

    as excellent, considering some of the

    crude approximations taken with the pres-sure and temperature profiles.

    Table 6. Comparison of exit concentrations(mole fraction) in an industrial reformer with

    model predictions

    Industrial

    reformer

    Model

    CH4 0.0110 0.0145

    H2O 0.4536 0.4878

    H2 0.4121 0.3885

    CO 0.0600 0.0419

    CO2 0.0620 0.0657

    N2 0.0013 0.0015

    3.3 Sensitivity analysisTo make an additional validation of the

    model and get a better understanding of

    the reforming reaction, a sensitivity analy-

    sis of the model response to variations in

    temperature profile, steam-to-carbon ratio

    in

    CS (in molar units) and residence

    time was carried out. There are different

    variables that can be used to evaluate the

    performance of a methane steam reformer

    (e.g. mass flow hydrogen produced, con-

    version of different species, rate of coke

    deposition, rate of H2to CO at the reactor

    exit). Among those and for clarity of thediscussion we selected only two: conver-

    sion of methane 4CHX defined as theratio of methane moles converted in the

    reactor to those entering the reformer and

    the fractional conversion of hydrogen

    42 CHH defined inE 8.

    (moles)convertedCH

    molesreactor

    theleavingH

    CHH

    4

    2

    42

    E 8

    Figure 3 shows the variation of methane

    conversion 4CHX and hydrogen fraction-al conversion 42 CHH with changes innominal reactor temperature. Because

    there is a temperature profile, the abscissa

    inFigure 3 represents a nominal tempera-

    ture that reflects the mean temperature in

    the profile.

    As expected, inFigure 3 methane conver-

    sion increases with temperature. However,

    42 CHH decreases, with temperature,as thermodynamic restrictions on the for-

    ward reaction in R 3 augment with tem-

    perature. Figure 3 suggests that a temper-

    ature of the order of 1100 K guarantees

    acceptable methane conversion, while the

    fractional hydrogen conversion remains at

    an acceptable value of 2.8. It is important

    to note that accordingR 1 toR 3,the max-

    imum possible value of 42 CHH is 4.A value of 2.8, such as that obtainedaround 1100 K, seems a good compromise

    between methane conversion and hydro-

    gen production.

    As Figure 4 shows, a higher value in the

    steam to carbon ratio in

    CS favors both,

    methane conversion and fractional con-

    version to hydrogen. This happens be-

    cause a higher steam concentration favors

  • 8/12/2019 Pulgarin_Leon Innovaquimica

    8/10

    Pulgarn et al., Innovaqumica, 2009

    8

    equilibrium toward hydrogen production

    in E 3. Favoring hydrogen production

    increases as well methane conversion.

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    700 900 1100 1300

    temperature (K)

    XCH4

    2.4

    2.6

    2.8

    3.0

    3.2

    (H2/CH4)

    Figure 3. Predicted variation of methane con-

    version 4CHX and fractional conversion ofhydrogen 42 CHH with changes in no-minal reactor temperature

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    4 5 6 7 8(S/C)in(mol/mol)

    XCH4

    2.4

    2.6

    2.8

    3.0

    3.2

    (H2/CH4)

    Figure 4. Predicted variation of methane con-

    version 4CHX and hydrogen fractional con-version 42 CHH with steam-to-carbonratio

    inCS

    Figure 4 suggests that the use of a higher

    value of in

    CS would be advisable, as

    higher methane conversion and fractional

    conversion of hydrogen are desirable out-

    comes. However, as Figure 5 shows, the

    total mass flow of hydrogen2H

    m de-

    creases as the steam to carbon ratio in-

    creases. Therefore, a balance between a

    reduction on hydrogen production and an

    increase in methane conversion and frac-

    tional conversion of hydrogen needs to be

    performed during reactor design. A final

    sensitivity analysis was conducted for

    residence time (Figure 6). The model

    predicts a reduction on methane conver-

    sion and fractional selectivity of hydrogen

    as the residence time in the reactor de-

    creases. This is an expected result. How-

    ever, Figure 6 gives very important infor-mation for model validation as shows that

    for typical conditions (residence time of

    0.11 s) methane conversion and the frac-

    tional conversion of hydrogen present a

    plateau that favors stable operation.

    900

    1000

    1100

    1200

    1300

    3.5 4.5 5.5 6.5 7.5(S/C)in(mol/mol)

    h/

    kg

    m

    2

    H

    Figure 5. Predicted variation of hydrogen mass

    flow2H

    m with changes in the steam-to-carbon

    ratio in

    CS

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    0.00 0.02 0.04 0.06 0.08 0.10 0.12

    residence time (s)

    XCH4

    2.4

    2.6

    2.8

    3.0

    3.2

    (H2/CH4)

    Figure 6. Predicted variation of methane con-

    version 4CHX and hydrogen fractional con-version 42 CHH with residence time in thereactor

    4. ConclusionsA model that integrates complex chemical

    kinetics into ASPEN using a FORTRAN

    subroutine correctly predicts the behavior

    of an industrial methane steam reformer.

    The model, validated against literature and

    industrial data, was used to conduct a sen-

  • 8/12/2019 Pulgarin_Leon Innovaquimica

    9/10

    Pulgarn et al., Innovaqumica, 2009

    9

    sitivity analysis of methane conversion

    and fractional conversion to hydrogen

    with temperature, steam to carbon ratio

    and residence time. Although the higher

    the temperature, the higher the methaneconversion, thermodynamic restrictions

    decrease methane and steam conversion to

    hydrogen as temperature increases. A

    nominal temperature of 1100 K seems to

    give a reasonable balance between me-

    thane conversion and factional conversion

    of hydrogen. Higher steam to carbon ratio

    favors both methane conversion and fac-

    tional conversion of hydrogen, however it

    decreases the mass flow of hydrogen. It is

    necessary to establish a balance on prod-uctivity and conversion to assure optimal

    reformer behavior. The model gives in-

    formation on the variation of methane

    conversion and fractional conversion of

    hydrogen with residence time. The results

    suggest that for the industrial reformer

    analyzed, a residence time close to 0.12 s

    is optimal because it guarantees that varia-

    tions in residence time do not have a ma-

    jor impact in conversion.

    Acknowledgment

    We acknowledge Aspen Technology for

    having granted special permission for the

    use of the ASPEN. PLUS under the condi-

    tion of the academic licensing agreement.

    References

    1. Ros, J.A., Modelo matemtico y simulacin de

    un reformador de metano, 1981, M.Sc. Ingeniera

    Qumica. Universidad Industrial de Santander,

    Ingeniera Qumica: Bucaramanga. p. 242.

    2. Murray, A.P. and Snyder, T.S., Steam-methane

    reformer kinetic computer model with heat trans-

    fer and geometry options. Ind. Eng. Chem. Proc.

    Des. Dev., 1985. 24(2): p. 286-294.3. Soliman, M.A., El-Nashaie, S.S.E.H., Al-Ubaid,

    A.S., and Adris, A., Simulation of steam reformers

    for methane. Chem. Eng. Sci., 1988. 43(8): p.1801-1806.4. Murty, C.V.S. and Murthy, M.V.K., Modeling

    and simulation of a top-fired reformer. Ind. Eng.

    Chem. Res., 1988. 27(10): p. 1832-1840.

    5. Xu, J. and Froment, G.F., Methane steam re-

    forming: Ii. Diffusional limitations and reactor

    simulation.AIChE J, 1989. 35(1): p. 97-103.6. Alatiqi, I.M., Meziou, A.M., and Gasmelseed,

    G.A., Modelling, simulation and sensitivity analy-

    sis of steam-methane reformers.Int. J. Hyd. Ener-gy, 1989. 14(4): p. 241-256.7. Xu, J. and Froment, G.F., Methane steam re-

    forming, methanation and water-gas shift: I. In-

    trinsic kinetics.AIChE J, 1989. 35(1): p. 88-96.8. Rajesh, J.K., Gupta, S.K., Rangaiah, G.P., and

    Ray, A.K., Multiobjective optimization of steam

    reformer performance using genetic algorithm.

    Ind. Eng. Chem. Res., 2000. 39(3): p. 706-717.9. Pina, J., Schbib, N.S., Bucala, V., and Borio,

    D.O., Influence of the heat-flux profiles on the

    operation of primary steam reformers. Ind. Eng.

    Chem. Res., 2001. 40(23): p. 5215-5221.

    10. Prasad, P. and Elnashaie, S.S.E.H., Novelcirculating fluidized-bed membrane reformer for

    the efficient production of ultraclean fuels from

    hydrocarbons. Ind. Eng. Chem. Res., 2002.

    41(25): p. 6518-6527.11. Snoeck, J.W., Froment, G.F., and Fowles, M.,

    Steam/co2 reforming of methane. Carbon forma-tion and gasification on catalysts with various

    potassium contents. Ind. Eng. Chem. Res., 2002.

    41(15): p. 3548-3556.12. Nandasana, A.D., Ray, A.K., and Gupta, S.K.,

    Dynamic model of an industrial steam reformer

    and its use for multiobjective optimization. Ind.

    Eng. Chem. Res., 2003. 42(17): p. 4028-4042.13. Prasad, P. and Elnashaie, S.S.E.H., Coupled

    steam and oxidative reforming for hydrogen pro-

    duction in a novel membrane circulating fluidized-

    bed reformer.Ind. Eng. Chem. Res., 2003. 42(20):p. 4715-4722.

    14. Acua, A., Simulacin de un reactor para

    generar hidrgeno y desarrollo de un sistema de

    control avanzado, 2004, Ph.D. Thesis. Universidad

    Industrial de Santander, Ingeniera Qumica: Buca-

    ramanga. p. 205.

    15. Posada, A. and Manousiouthakis, V., Heat and

    power integration of methane reforming based

    hydrogen productionInd. Eng. Chem. Res.,2005. 44(24): p. 9113-9119.16. Rakib, M.A. and Alhumaizi, K.I., Modeling of

    a fluidized bed membrane reactor for the steam

    reforming of methane: Advantages of oxygen addi-

    tion for favorable hydrogen production. Energy

    Fuels, 2005. 19(5): p. 2129-2139.17. Patil, C.S., van Sint Annaland, M., and Kui-

    pers, J.A.M., Design of a novel autothermal mem-brane-assisted fluidized-bed reactor for the pro-

    duction of ultrapure hydrogen from methane. Ind.

    Eng. Chem. Res., 2005. 44(25): p. 9502-9512.18. Sarvar-Amini, A., Sotudeh-Gharebagh, R.,

    Bashiri, H., Mostoufi, N., and Haghtalab, A., Se-

  • 8/12/2019 Pulgarin_Leon Innovaquimica

    10/10

    Pulgarn et al., Innovaqumica, 2009

    10

    quential simulation of a fluidized bed membrane

    reactor for the steam methane reforming using

    aspen plus. Energy Fuels, 2007. 21(6): p. 3593-3598.

    19. Abashar, M.E.E., Alhabdan, F.M., and Elna-

    shaie, S.S.E.H., Staging distribution of oxygen incirculating fast fluidized-bed membrane reactors

    for the production of hydrogen. Ind. Eng. Chem.Res., 2007. 46(17): p. 5493-5502.20. Zamaniyan, A., Ebrahimi, H., and Moham-

    madzadeh, J.S.S., A unified model for top fired

    methane steam reformers using three-dimensional

    zonal analysis.Chem. Eng. Proc., 2008. 47(5): p.946-956.

    21. Schwaab, M., Alberton, A.L., Fontes, C.E.,

    Bittencourt, R.C., and Pinto, J.C.,Hybrid modeling

    of methane reformers. 2. Modeling of the industri-

    al reactors.Ind. Eng. Chem. Res., 2009.

    22. Alberton, A.L., Schwaab, M., Fontes, C.E.,Bittencourt, R.C., and Pinto, J.C.,Hybrid modeling

    of methane reformers. 1. A metamodel for the

    effectiveness factor of a catalyst pellet with com-

    plex geometry.Ind. Eng. Chem. Res., 2009.

    23. Haldor, T., Some aspects of catalytic cracking.

    1965, Insittute of gas engineers.24. Morley, C. Gas-eq: A chemical equilibrium

    program for windows. 2003 [cited 2003 August];

    Available from:

    http://www.c.morley.ukgateway.net/.

    25. Barrera-Zapata, R., Villa de P, A.L., and Mon-

    tes de Correa, C., Modelado y simulacin de la

    epoxidacin de limoleno con aspen plus, in XXIVConferencia Internacional Ingeniera Qumica,

    A.C.d.I. Qumica, Editor. 2007: Cali.

    http://www.c.morley.ukgateway.net/http://www.c.morley.ukgateway.net/http://www.c.morley.ukgateway.net/