18
Fluid Phase Equilibria, 13 (1983) 91-108 Ekvier Science Publishers B.V., Amsterdam -Printed in The Netherlands 91 EXTENSION OF THE PENG-ROBINSON EQUATION OF STATE TO COMPLEX MIXTURES: EVALUATION OF THE VARIOUS FORMS OF THE LOCAL COMPOSITION CONCEPT Paul M. Mathias and Thomas W. Copeman Air Products & Chemicals, Inc. Box 538, Allentown, PA 18105 ABSTRACT -- Density-dependent local composition (DDLC) mixing rules (Whiting and Prausnitz, 1981, 1982 and Mollerup, 1981) offer great promise to extend equations of state to highly nonideal mixtures. In this work we investigate various aspects of the DOLC concept and offer two useful forms. All development is done with the Peng-Robinson (1976) equation of state, but the results have general applicability. INTRODUCTION Practical models to describe phase equilibrium can be divided into two broad categories: equations of state and activity-coefficient models. Equations of state have been successfully applied to mixtures containing ,nonpolar and slightly polar components at all conditions of engineering interest, including the critical and retrograde regions. On the other hand, the mathematical flexibility of activity coefficient models has been considered necessary to model systems which exhibit high liquid-phase nonideality. The activity-coefficient approach works well at low reduced temperatures where the liquid phase is relatively incompressible and up to moderate pressures. Serious problems arise when these bounds are violated. ,Further, the use of different models for the various phases precludes the correct description of mixture critical points. Additional problems arise for supercritical components (Abrams, et al., 1975). The equation-of-state approach does not inherently suffer from these limitations and thus it would be very valuable to develop equations of state with the mathematical flexibility to describe complex mixture behavior. Recent research in equations of state has focussed on the mixing rules as a promising way to achieve mathematical flexibility through a semi-theoretical development. Vidal (1978) has shown that the standard mixing rules used with the Soave (1972) and Peng-Robinson (1976) equations of state are very similar to regular solution theory and thus inadequate for "chemical" systems. Further, Huron and Vidal (1979) have suggested improvements to the standard mixing rules by applying the ideas of NRTL model (Renon and Prausnltz, 1968) 0378-3812/83/$03.00 0 1983 Elsevier Science Publishers B.V.

extension of peng-robinson for complex mixtures

Embed Size (px)

DESCRIPTION

copeman

Citation preview

  • Fluid Phase Equilibria, 13 (1983) 91-108 Ekvier Science Publishers B.V., Amsterdam -Printed in The Netherlands

    91

    EXTENSION OF THE PENG-ROBINSON EQUATION OF STATE TO COMPLEX MIXTURES:

    EVALUATION OF THE VARIOUS FORMS OF THE LOCAL COMPOSITION CONCEPT

    Paul M. Mathias and Thomas W. Copeman Air Products & Chemicals, Inc. Box 538, Allentown, PA 18105

    ABSTRACT --

    Density-dependent local composition (DDLC) mixing rules (Whiting and

    Prausnitz, 1981, 1982 and Mollerup, 1981) offer great promise to extend

    equations of state to highly nonideal mixtures. In this work we investigate

    various aspects of the DOLC concept and offer two useful forms. All

    development is done with the Peng-Robinson (1976) equation of state, but the

    results have general applicability.

    INTRODUCTION

    Practical models to describe phase equilibrium can be divided into two

    broad categories: equations of state and activity-coefficient models.

    Equations of state have been successfully applied to mixtures containing

    ,nonpolar and slightly polar components at all conditions of engineering

    interest, including the critical and retrograde regions. On the other hand,

    the mathematical flexibility of activity coefficient models has been

    considered necessary to model systems which exhibit high liquid-phase

    nonideality. The activity-coefficient approach works well at low reduced

    temperatures where the liquid phase is relatively incompressible and up to

    moderate pressures. Serious problems arise when these bounds are violated.

    ,Further, the use of different models for the various phases precludes the

    correct description of mixture critical points. Additional problems arise for

    supercritical components (Abrams, et al., 1975). The equation-of-state

    approach does not inherently suffer from these limitations and thus it would

    be very valuable to develop equations of state with the mathematical

    flexibility to describe complex mixture behavior.

    Recent research in equations of state has focussed on the mixing rules as a

    promising way to achieve mathematical flexibility through a semi-theoretical

    development. Vidal (1978) has shown that the standard mixing rules used with

    the Soave (1972) and Peng-Robinson (1976) equations of state are very similar

    to regular solution theory and thus inadequate for "chemical" systems.

    Further, Huron and Vidal (1979) have suggested improvements to the standard

    mixing rules by applying the ideas of NRTL model (Renon and Prausnltz, 1968)

    0378-3812/83/$03.00 0 1983 Elsevier Science Publishers B.V.

  • 92

    to the excess Gibbs energy of the Soave equation. The results are promising

    but the Huron and Vidal model does not meet the important low-density 1 imit.

    This has been pointed out by Mollerup (1981) and Whiting and Prausnitz (1981,

    1982); these investigators have separately proposed a physically appealing

    model to account for mixture nonideality: density-dependent local composition

    (DDLC) theory.

    In this work we have investigated density-dependent local-composition

    theory as applied to the popular Peng-Robinson (PR) equation of state. The

    standard Peng-Robinson equation provides a very good description of

    asymmetric, nonpolar systems but the DDLC-PR model proposed by Mollerup (1981)

    .can be inaccurate by up to 3-4 orders of magnitude for highly asymmetric

    systems. Noting this, we have adopted a modlficatlon of an idea proposed by

    Dimitrelis and Prausnitz (1982). The new model enables improved correlations

    of a wide variety of complex mixtures. Our proposed DOLC model (along with

    the other recently proposed forms) could prove to be too expensive in computer

    time for systems with a large number of components. Thus we propose a

    truncated version of the theory which, however, retains several attractive

    qualities.

    STANDARD PENG-ROBINSON EQUATION OF STATE

    According to the Peng-Robinson

    pure fluid is given as a function

    follows:

    P=RT_ a

    v-b v(v+b) + b(v-b)

    (1976) equation of state the pressure of a

    of the temperature and molar volume as

    (1)

    a and b are component dependent parameters. b is independent of temperature

    and a is temperature dependent such that the vapor pressure of the pure fluid

    i s accurately carrel ated. Thus,

    b = 0.07780 R Tc/Pc

    a = ace(T)

    a C

    = 0.45724 R=T,=/P,

    B = [ 1 + CI(1 - fiR) + c2(1 - FR)Z + c3(l - j-YR)l ] 2

    Equation (5) Is slightly different from that origlnally proposed by Peng

    and Robinson (1976). The original PR model is obtained when c2=c3=0. The

    additional parameters are necessary to correlate the vapor pressure of highly

    polar substances 11 ke water and methanol. In this work we wish to focus on

    the development of mixing rules. Inaccurate pure component vapor pressure

  • 93

    representation artificially distorts the excess Helmholtz free energy and

    obscures the analysis of mixture effects (Van Ness, et al., 1973).

    For mixtures, Peng and Robinson assumed that eqn. (1) remains unchanged

    but the parameters a and b are calculated as the mole fraction averages shown

    below:

    a = I E xixjaji

    ij

    (6)

    b = Z xibi

    i

    (7)

    a.. is given by the usual combining rules which include the pair dependent

    biiary parameter kji

    aji = (1 - kji) (aiaj)'fl

    Equations (1) and (6-8) deflne the "standard" Peng-Robinson model. We

    refer to this form as the van der Waals l-Fluid (VOW-l) or "conformal" model.

    It should be noted that the mixing rule defined by eqns. (6) and (7) do not

    imply that the mixture is random (all pair radial distribution functions are

    identical) but rather that the various pair radial distribution functions can

    be superimposed on a reduced basis (Mansoori and Leland, 1972).

    ANALYSIS OF THE MOLLERUP DDLC MODEL

    Mollerup (1981) derived his version of the density-dependent local

    composition Peng-Robinson equation of state by assuming that local

    compositions arise from differences fn the attractive parameter "a" per unit

    surface area (q).

    Lx a /q E

    P=RT_ 1 j j ji ji ji

    zxq v-b v(v+b) + b(v-b) i i i 1 x. E..

    J J' j

    where, Eji = exp ~~~~~

    (I 1 = exp - aji/qji ~ en

    RT 2-&b

    (9)

    (10)

    (11)

    UserHighlight

    UserSticky Notewhy assume?

  • 94

    and, aji/qj, = (1 - kji) (aiaj/qiqj)lB (12)

    Actually, Mollerup did not employ the universal "degree-of-randomness"

    parameter a (Whiting and Prausnitz, 1981), however, it is included here for

    generality.

    The attractive Helmholtz energy corresponding to eqn. (9) is given by

    Aa = RT

    - - Z xiqian

    [I

    BXjEji

    II. i j

    (13)

    Useful insight into the nature of the DDLC theory is obtained by an

    expansion of the logarithm and exponential terms in eqn. (13) as follows:

    -a A = - Zxiqi Exjaji/qji

    I i j I

    r

    FV

    1 Q - 2xiqi Exjafji/qZji - (Exjaj

    i \ j j

    In eqn. (14) we have used the simp

    1 v + (1 + 2)b $ Fv e - en

    2$b I 1 v + (I -4)b

    (14) /qji)=

    r'

    - Fv2 - .._._ 2RT

    ifying notation,

    (IS)

    L J

    It is interesting to note that the first term of eqn. (14) has exactly the

    same functional dependence as the standard or VDW-1 Peng-Robinson equation of

    state. (A minor difference arises in the i-j combining rule when qi # q., but

    this could be absorbed by an appropriately chosen binary parameter, kji. ; To

    fix ideas, we define the first term of eqn. (14) as the "equivalent" VDW-1

    contribution and lump all the remaining terms into the "asymmetric" or

    "nonconformal" contribution.

    It is illustrative to compare the fugacity coefficients of eqn. (13) with

    those of its equivalent VDW-1 form. Due to the complexity of the equations,

    we only show the comparison for a binary with component 1 at infinite

    dilution.

    01-

    - 1 = -9 2

    _!I

    -VDW-1

    1) - (16)

  • 95

    a12 a22 1 where, X = - - - - Fv

    [ d 912 qz RT (17) The right side of eqn. (16) is always negative and thus the fugacity

    coefficient is reduced from its VDW-1 value. (The opposite effect is obtained

    for negative values of 0~. However, only positive values of (I are

    physically meaningful.)

    In general, it can be shown that the local composition effect reduces the

    Helmholtz energy of a system at fixed T, V, g. This is a qualitatively

    correct result since the DDLC configuration should be more favorable than its

    equivalent VDW-1 form.

    EVALUATION OF THE MOLLERUP MODEL FOR ASYMMETRIC NONPOLAR SYSTEMS

    The VOW-1 PR-EOS provides a good description of most nonpolar systems,

    including those which are highly asymmetric. For example, the methane-decane

    binary is accurately correlated over a temperature range of 311-583 K, even

    with the default interaction parameter, k2I=O; the average absolute errors in

    the K-values of methane and decane are 3.5% and 8.8%, respectively. It is

    extremely important that the DDLC mixing rules retain this desirable

    capability. Therefore we have conducted an investigation of the model

    proposed by Mollerup (eqn. 9) by using the methane-decane binary as a test

    case.

    Figures 1, 2, and 3 present the results of the investigation. We have

    somewhat arbitrarily chosen to show the liquid-phase fugacity coefficient of

    methane as a function of composition at 310 K and 270 atm. In general, the

    DDLC mixing rules cause a decrease in the fugacity coefficient. Figure 1

    shows that this decreasecan be enormous, giving fugacity coefficients too low

    by as much as 3-4 orders of magnitude when the q of decane is unity. Figure 1

    also presents the fugacity coefficients with the size parameter from Bondi

    (1968), q=6, and two higher values. The local composition effect is still

    large. A non-randomness parameter (o) of 0.2 along with the same decane

    size parameters is shown in figure 2. Figure 3 illustrates the additional

    effect of setting the binary parameter to 0.5. These comparisons indicate

    that the Mollerup model requires large and unreasonable parameters to

    correlate this relatively simple system.

    DEVELOPMENT OF A NEW DDLC MODEL: DEPARTURE FROM THE GEOMETRIC MEAN

    The results of the last section show that the "asymmetric" contribution

    greatly over-predicted by the Mollerup version of the DDLC-PR EOS for

    too

    iS

    asymmetric, nonpolar systems. A similar conclusion has already been reached

  • All u: P - 1.0 kz,- 0.0

    q-+-y-- o.b 0.0 0.2

    x(t&aM) OS6

    Fig 1. Effect of LC model (eqn. 9). Liquid-phase of methane-decane binary at 310 K and 270 atm.

    10

    1 VOW-l. 9'1 (OATA wlnuw 10x) k

    $; U.H -/- -- _---,/H -- s 0.1: -xiq=ro___--- ,/-- - __--- #H _I All lx%

    :__/-/- u. 40 t*;%l 0.01 , I 1 I I

    0.0 0.2 0.4 0.8 0.8

    %(Methano)

    Fig 2. Effect of LC model (eqn. 9). 310 K and 270 atm.

    Liquid-phase of methane-decane binary at

  • 91

    i VOW-~. q=l (DATA WITNIN 10%)

    -1 ei LC.qrlO -__- _____c _--- u. q=s __-- --- __-- 9 I l.c. q*o All u

    ,,i _

    0.0 0:2 I

    x(&a"*)

    0:s 0.k

    Fig 3. Effect of LC model 310 K and 270 atm.

    (eqn. 9). Liquid-phase of methane-decane binary at

    Fig 4. Computer time for calculation of fugacity coeffkients for n-component system.

  • 98

    by Dimitrelis and Prausnitz (1982) who further point out that models analogous

    to eqn. (9) make the rather stringent assumption that local compositions

    Inevitably occur In the j-l pair interaction whenever ai/qi # aj/qj. They

    have proposed that local composition effects are not caused when ji

    interactions are different from ii but rather when ji are different from some

    "ideal" combination of ii and jj interactions, referred to as ji".

    Dimitrelis and Prausnitz define jiO as an arithmetic mean. Since the standard

    Peng-Robinson generally produces satisfactory results for nonpolar systems

    with kji = 0. the geometric mean should be a very good first approximation for

    jiO.

    We assume that the total molar internal energy of the ji pair in a

    multi-component fluid is divideb into two parts: ideal and excess. The

    "ideal" part is entirely conformal as in the standard Peng-Robinson model.

    However, the "excess" contribution causes non-conformal local compositions

    described by the DDLC theory. \

    -a -a0 -aEX U ji

    U = ji + u..

    31

    -a0 a FV U ji =_._._

    (1 - kji) -

    al/T T I VP!.!

    -aEX a FV

    "ji = ~ (-dji) -

    al/T T I V,N

    (19)

    (20)

    Equation (19) corresponds to the standard Peng-Robinson model. The usual

    binary parameter kji is retained. However, in the data analysis we have

    generally set kji = 0. The parameter dji determines that part of the internal

    energy that produces local composition effects. Note that 4 is not symmetric

    (i.e., d jf # dfj). The total mixture molar internal energy is obtained by

    summing the ideal and excess ji contributions using the VDW-1 and DDLC mixing

    rules, respectively.

    -a0 -a0 U = E Z xjxj uji (21)

    ij

    -aEX -aEX U = B Xrqi x ji j< (22)

    i j

  • Eji where 'ji

    = x. - J

    X 'k Eki k

    and Eji = exp (-dji)Fv 1

    99

    (23)

    (24)

    The Helmholtz energy may be obtained by integrating eqns. (18-24) with

    respect to l/T. Thus,

    _a RT A = - L E xixj (aiaj)'f2 (1 - kji)Fv - - X xiqi rn I: xj Eji

    [ 1 (25) ij cri j It should be noted that the integration to obtain Helmholtz energy requires

    an expression for the athermal limit. Here we assume that the repulsive term

    of the standard Peng-Robinson equation of state yields a reasonable athermal

    contribution to the total Helmholtz free energy. Whiting and Prausnitr (1982)

    have shown that the Peng-Robinson repulsive term resembles the Flory-Huggins

    athermal term. Also we note that highly asymmetric nonpolar systems (e.g.,

    methane-decane) are well described by the standard Peng-Robinson model.

    Expressions for all configurational properties can easily be obtained from

    eqn. (25). However no derived properties are shown in order to save space.

    The above derivation is only valid from a phenomenological viewpoint. It

    does however state the assumptions that have been made. The new model meets a

    very important limit in addition to the ones at high temperature and low

    density noted by Whiting and Prausnitz (1981) and Mollerup (1981): that is,

    the nonconformal effect is not necessarily large (or even nonzero) for an

    asymmetric system. This feature allows the model to retain the good results

    of the standard Peng-Robinson model and facilitates accurate correlation of

    the phase equilibria of highly nonideal polar systems.

    DEVELOPMENT OF AN APPROXIMATE OR TRUNCATED DDLC MODEL

    The local-composition Peng-Robinson model defined by eqn. (25) produces

    greatly improved correlations over the standard Peng-Robinson model, as shown

    in the next section. However, the computational load might prove to be

    unacceptably high for practical applications with a large number of

    components. This arises from the bulk of computer time being consumed in mole

    fraction summation calculations. (Also see Boston and Mathias, 1980.) In the

    model defined by eqn. (25), these summations are density dependent and thus

    must be redone at each iteration for the mixture molar volume.

  • 100

    Figure 4 shows a comparison of the average computer time required by the

    standard Peng-Robinson model and DDLC version (eqn. 25) for a single

    calculation of the fugacity coefficients in an n-component system. In our

    opinion, a reasonable increase in computer time (say, a factor of two) is

    justified by a superior model, but an order of magnitude increase is

    impractical (with current computer capabilities for process simulation) when

    the number of components in the mixture is greater than about 15. Figure 4

    also includes a comparison for the approximate or truncated model we

    propose. We conclude that the truncated model may be used for large systems

    without incurring a prohibitive computational burden.

    According to the truncated DDLC model, the attractive Helmholtz energy of a

    mixture is given by

    -a aNC

    A = - a Fv - - Fvz

    2RT

    (26)

    NC where a is given by the standard Peng-Robinson form (eqns. 6 and 8) and a ,

    the NonConformal term, is given by

    aNc = H xi+ aci E 1 j

    The parameters

    xj tji (27)

    tii are the lumped contribution of all the nonconformal

    effects. Note thatt is not symmetric (tji # tij) and thus the model has 3

    parameters per binary pair; the third parameter is kji.

    The truncated model defined by eqns. (6, 8, 26, and 27) may be derived by

    expanding eqn. (25) analogously to eqn. (14) and retaining only the first

    nonconformal term. -aNC

    Denoting this term as A we obtain,

    -aNC 1 2 P z

    1 I

    A =-- Zxiaci xj dji - (z xj dji) F,

    2RT i j

    (28)

    Notice that 01 and the qs and as have been absorbed in the ds, This does

    not affect the composition dependence since only the first nonconformal term

    Is used. Further, the temperature dependence of a (eqns. 3 and 5) has been

    removed. This has facilitated good representation of data over a wide

    temperature range with temperature-independent parameters (e.g.,

    water-hydrocarbon systems).

  • 101

    The temperature dependence of eqn. (28) may also be justified on a

    theoretical basis. Mansoori and Leland (1972) have shown that the first

    nonconformal contribution to the compressibility factor is of order 1/T2.

    This is consistent with eqn. (28).

    The composition dependence chosen for eqn. (27) is somewhat arbitrary. It

    is suggested by writing eqn. (28) for a binary,

    -aNC 1 rzz 221

    A = - - x1 x2 x1 acl d21 + x2 ac2 d12

    2RT 1 1

    (29)

    Equations (28) and (29) always give a negative contribution to the total

    Helmholtz energy. This is physically correct but it could restrict the mathematical flexibility of the model. Therefore, we have chosen the

    relatively empirical form of eqn. (27). However, we should note that if the

    regressed value of t is large and negative, it suggests that the physical

    meaning is lost.

    In this work only binary data were analyzed and thus it was not possible to

    evaluate whether the more correct composition dependence of eqn. (29) is

    necessary for multicomponent mixtures. Further, for the systems we have

    analyzed the t's have been positive and thus eqns. (27) and (28) are

    identical.

    An extremely appealing aspect of the model is that it reduces exactly to

    the standard Peng-Robinson model when all the t's are equal to zero. Thus it

    is especially applicable to multicomponent systems whfch are largely nonpolar

    but have a few important polar components (e.g., water in petroleum or coal

    conversion processes).

    CORRELATION OF PHASE EQUILIBRIUM

    Vapor-liquid equilibrium was correlated for various systems using van der

    Waals one-fluid and local-composition mixing rules, eqn. (25). One parameter

    (k12) was fit with the van der Waals mfxing rules and two parameters (d12 and

    d21) were fit with the local composition rules. The non-randomness parameter

    (a) was set to unity and size parameters (q) were derived from Bondi (1968).

    Table 1 presents maximum and average K-value deviations for these two methods.

    In each system, local composition mixing rules enabled an improved fit to the

    data. The methanol-benzene system was correlated to a much higher accuracy,

    while improvement was less, but St.111 significant, for the other systems.

    Figures 5 and 6 present plots of K-value deviations from data for the

    methanol-benzene system. We should note the symmetric form of eqn. 25

    (d12 = d21) did not show significant improvement over the VDW-1 form.

  • Methanol(l)-

    Benzene(2)

    Methanol(l)-

    Carbon

    Dioxide(P)

    Acetone(l)-

    Water(2)

    Isobutylene(l)-

    Methanol(2)

    Ref

    .

    Nag

    a ta

    1969

    Ohgak

    i 1976

    Griswold

    1952

    Churkin

    1976

    Temp.

    (OC)

    60.8-76.7

    25-40

    100

    50

    TABLE

    1

    Correlation

    of Vapor-Liquid

    Equilibrium

    Data

    Pre

    ssu

    re

    (Atml

    1.0

    5.7-80.0

    1.1-3.6

    3.1-6.1

    Mix Rules

    vow-

    1 0.1

    0 -57.0

    62.1

    LC

    0.61

    -0.17

    -7.0

    -14.3

    VDW-1

    0.06

    14.2

    27.0

    LC

    0.23

    -0.37

    -4.1

    -22.5

    VDW-1

    -0.23

    LC

    -0.35

    -0.02

    ::::

    -49.6

    22.6

    -29.7

    10.6

    VDW-1

    0.04

    33.0

    LC

    0.0

    0.59

    -1.6

    Parameters

    %Kl max*

    XK2 max

    %Kl avg

    %K2 a"g

    k

    d

    d

    12

    21

    12

    2"::

    16.9

    2.5

    --

    17.1

    3.0

    14.8

    8.0

    23.9

    8.5

    * Average

    absolute

    deviation

    in K-value

    of component

    1.

  • 103

    15

    1,

    A

    X

    x (Methanol)

    Fig 5. Comparison of new LC model (eqn. 25) with experimental data (Nagata, 1969).

    Fig 6. Comparison of new LC model (eqn. 25) and VI@1 form with experimental data (Nagata, 1969).

  • TABLE 2

    Correlation

    of Liquid-Liquid

    Equilibrium

    Data

    1

    2

    3

    System

    REP

    Temp

    Mix Rules

    Parameter

    %l

    avg

    w

    avg

    0 c

    k

    k

    d

    d

    d

    12a

    12b

    21a

    12a

    12b

    1-Butanol(l)-

    Water(e)

    Hill

    5-80

    VDW-1

    -0.18

    49.1

    26.7

    1926

    LC

    -0.1

    -0.05

    -0.05

    0.14

    1.0

    0.7

    LC(tr)4

    0.07

    0.12

    0.45

    1.2

    2.6

    Benzene

    (l)-

    Water(e)

    Tsonopoulos

    10-200

    ZW-'

    -0.08

    132.1

    2691.1

    1982

    8:::

    -0.15

    1.08

    0.06

    LC(tr)

    0.04

    0.74

    :::

    :::

    Hexane(1)

    Water(Z)

    @clpoulos

    10-200

    VDW-1

    -0.24

    >10000

    LC

    0.15

    -0.24

    2.39

    0.15

    "X

    12.0

    LCItr)

    0.48

    0.0

    1.35

    16:7

    12.9

    I-Hethylnaph-

    thalene(l)-

    Water(P)

    Brady

    37-277

    1982

    ZW-'

    -:z

    1036.0

    -0.16

    1.16

    0.21

    23.2

    '"Z

    i_C(tr)

    0:31

    0.01

    1.36

    3.5

    618

    1

    d

    =d

    +

    d

    12

    12a

    T

    12b

    2 Average

    absolute

    deviation

    in calculated

    mole fraction

    of component

    1 in liquid phase 2.

    3 Average

    absolute

    deviation

    in calculated

    mole fraction

    of component

    2 in liquid phase

    1.

    4 For LC(tr),

    read "de as et".

  • 106

    Fig 7. Correlation of benzene-water mutual solubility data (Tsonopoulos and Wilson, 1982) with truncated LC model.

    1

    0.1

    0.01

    0.001

    0.0001-

    o.oooo1

    Fig 8. z.Zrap;lation of I-methylnapthalene-water mutual solubility data (Brady ., 1982) with truncated LC model.

  • 106

    Liquid-liquid equilibrium was correlated for a few selected systems using

    van der Waals one-fluid and local composition (including the truncated) mixing

    rules. Table 2 presents average absolute deviations in calculated mole

    fractions for these methods. One parameter was fit to the van der Waals

    mixing rule, four parameters (to include temperature dependence) were fit

    the local composition and three to the truncated local composition mixing

    rules. Although the overall improvement is substantial with four local

    composition parameters, it should be noted that temperature dependence in

    k12 and d12 parameters must be correlated for an accurate fit. In this

    to

    the

    regard, the local composition equation of state is analogous to the current

    local composition activity coefficient models. It should be noted that even

    with four binary parameters (also typically used with the UNIQUAC activity

    coefficient model) the local composition method provides a significant

    improvement beyond methods like Peng and Robinson (1980) for correlation of

    liquid-liquid equilibria. These methods use separate interactlon parameters

    (temperature dependent) for each phase. No ambiguity or discontinuity arises

    with the local composition method at a critical end point where (any) two

    phases merge.

    The results with the truncated local composition model are particularly

    encouraging since the (three) interaction parameters are not temperature

    dependent. Figures 7 and 8 present a comparison of model calculations against

    experimental data. Note that tZII for water in the hydrocarbons are nearly

    zero. A two-parameter fit would yield about the same accuracy.

    CONCLUSION

    In this work we have investigated the density-dependent local composition

    theory as applied to the Peng-Robinson equation of state. We propose two

    useful models to account for nonconformal or asymmetric effects. The approach

    is also applicable to any other equation of state derived from the generalized

    van der Waals partition function.

    Finally, the results of this work demonstrate that an equation of state

    model which employs DDLC mixing rules and thus meets the required limits at

    high temperature and low density is promising as a substitute for activity

    coefficient models in practical applications.

    ACKNOWLEDGEMENT

    The authors wish to thank Professors J. M. Prausnitz and F. P. Stein and

    D. Dimitrelis for helpful discussions.

    The work reported in this paper was sponsored in part by the International

    Coal Refining Company and the United States Department of Energy under

    Contract DE-AC05-780R03054.

  • 107

    List of Symbols

    Helmholtz Energy Peng-Robinson parameter, eqn. 1 Second Virial Coefficient Peng-Robinson parameter, eqn. I Binary Interaction Parameter Binary Interaction Parameter Surface Area Size Parameter Number of Moles Pressure Gas Constant Temperature Binary Interaction Parameter Volume Molar Volume Mole Fraction Non Randomness Parameter Temperature Dependent Pure Component Parameter, eqn. 5 Fugacity Coefficient

    Superscipts

    a Attractive Contribution Infinite Dilution

    Subscripts

    i Component i i Comoonent 5 i ComPonent k C Critical

    REFERENCES Abrams, 0. S. and Prausnitz, J. M., 1975, AIChE J. 21:116-128.

    Abrams, 0. S., Seneci, F., Chueh, P. L., and Prausnitz, J. M., 1975, Ind. Eng. Chem., Fundam., 14:52-54.

    Bondi, A., 1968. Physical Properties of Molecular Crystals, Liquids, and Glasses, John Waley and Sons, New York.

    Boston, J. F. and Mathias, P. M., 1980, Phase Equiilibria and Fluid Properties in the Chemical Industry - 2nd Internation Conference, Berlin. March 17-21:823-848.

    Brady, C. J., Cunningham, J. R., and Wilson, G. M., 1982, Gas Processors Association, RR-62.

    Churkin, V. N., Gonshov, V. A., and Pavlov, S. Y., 1978. Zh Fiz Khim. 52:488.

    Dimitrelis, D. and Prausnitz, 3. M., 1982. "Thermodynamic Properties of Strongly Nonideal Fluid Mixtures from an Extended Quasi-Chemical Theory," paper presented at Fall Annual Meeting. American Institute of Chemical Engineering, Los Angeles, CA, Nov 14-19, 1982.

    Gmehling, J. and Onken, U., 1981. Vapor-Liquid Equilibria Data Collection, Dechema, Frankfurt.

  • Griswold, J. and Wong, S. Y.. 1952, Chem. Eng. Prog. Symp. Ser. 48:18.

    Hill, A. E., and Malisoff, W. M. 1926, J. Am. Chem. Sot. 48:918.

    Huron, M. J. and Vidal, J., 1979, Fluid Phase Equilibria, 3:255-271.

    Mansoori, G. A. and Leland, Jr., T. W., 1972, J. Chem. Sot., Faraday Trans. II, 68:320-344.

    Mollerup, J., 1981, Fluid Phase Equilibria, 7:121-138.

    Nagata, I., 1969, J. Chem. Eng. Data, 14:418.

    Ohgaki , K. and Takashi, K., 1976, J Chem. Eng. Data 21:53.

    Peng, D. Y., and Robinson, D. 8.. 1976. Ind. Eng. Fundam., 15:59-64.

    Peng, 0. Y., and Robinson, D. B., 1980, in ACS Symposium Series 133, Ed: Newman, S. A., Washington, DC, 393-414.

    Reamer, H. H., Olds, R. H.. Sage, 8. H. and Lacey, W. N.. 1942, Ind. Eng. Chem., 36:1526.

    Soave. G., 1972, Chem. Eng. Sci., 27:1197.

    Tsonopoulos, C. and Wilson, G. M., 1982. High Temperature Mutual Solubilities of Hydrocarbons and Water. I. Benzene, Cyclohexane, and n-Hexane. accepted for publication in AIChE Journal.

    Van Ness, H. C., Byer, and S. M., Gibbs, R. E., 1973. AIChE Journal 19:238-244.

    Vidal, J., 1978, Chem. Eng. Sci., 33:787-791.

    Whiting, W. B. and Prausnitz, J. M., 1981. Equations of State for Strongly Nonideal Fluid Mixtures: Application of The Local-Composition Concept, paper presented at Spring National Meeting. American Institute Chemical Englneering, Houston, Texas, 5-g April 1981.

    Whiting, W. B. and Prausnitz, J. M., 1982, Fluid Phase Equilibria, 9:119-147.