+ All Categories
Home > Documents > Mutual diffusion of binary liquid mixtures · NRTL and UNIQUAC. It is found that the Wilson model...

Mutual diffusion of binary liquid mixtures · NRTL and UNIQUAC. It is found that the Wilson model...

Date post: 02-Feb-2021
Category:
Upload: others
View: 5 times
Download: 0 times
Share this document with a friend
112
Mutual diffusion of binary liquid mixtures Mutual diffusion of binary liquid mixtures containing methanol, ethanol, acetone, benzene, cyclohexane, toluene and carbon tetrachloride a) Gabriela Guevara-Carrion, 1 Tatjana Janzen, 1 Y. Mauricio Muñoz-Muñoz, 1 and Jadran Vrabec 1, b) Thermodynamics and Energy Technology, University of Paderborn, 33098 Paderborn, Germany (Dated: 6 March 2016) Mutual diffusion coefficients of all 20 binary liquid mixtures that can be formed out of methanol, ethanol, acetone, benzene, cyclohexane, toluene and carbon tetrachlo- ride without a miscibility gap are studied at ambient conditions of temperature and pressure in the entire composition range. The considered mixtures show a varying mixing behavior from almost ideal to strongly non-ideal. Predictive molecular dy- namics simulations employing the Green-Kubo formalism are carried out. Radial distribution functions are analyzed to gain an understanding of the liquid structure influencing the diffusion processes. It is shown that cluster formation in mixtures containing one alcoholic component has a significant impact on the diffusion process. The estimation of the thermodynamic factor from experimental vapor-liquid equilib- rium data is investigated, considering three excess Gibbs energy models, i.e. Wilson, NRTL and UNIQUAC. It is found that the Wilson model yields the thermodynamic factor that best suits the simulation results for the prediction of the Fick diffusion coefficient. Four semi-empirical methods for the prediction of the self-diffusion co- efficients and nine predictive equations for the Fick diffusion coefficient are assessed and it is found that methods based on local composition models are more reliable. Finally, the shear viscosity and thermal conductivity are predicted and in most cases favorably compared with experimental literature values. Keywords: Fick diffusion coefficient, Maxwell-Stefan, Green-Kubo, thermodynamic factor, radial distribution function, shear viscosity, thermal conductivity, model pre- diction a) Supplementary Materials available. b) Electronic mail: [email protected] 1
Transcript
  • Mutual diffusion of binary liquid mixtures

    Mutual diffusion of binary liquid mixtures containing methanol, ethanol, acetone,

    benzene, cyclohexane, toluene and carbon tetrachloridea)

    Gabriela Guevara-Carrion,1 Tatjana Janzen,1 Y. Mauricio Muñoz-Muñoz,1 and Jadran

    Vrabec1, b)

    Thermodynamics and Energy Technology, University of Paderborn,

    33098 Paderborn, Germany

    (Dated: 6 March 2016)

    Mutual diffusion coefficients of all 20 binary liquid mixtures that can be formed out

    of methanol, ethanol, acetone, benzene, cyclohexane, toluene and carbon tetrachlo-

    ride without a miscibility gap are studied at ambient conditions of temperature and

    pressure in the entire composition range. The considered mixtures show a varying

    mixing behavior from almost ideal to strongly non-ideal. Predictive molecular dy-

    namics simulations employing the Green-Kubo formalism are carried out. Radial

    distribution functions are analyzed to gain an understanding of the liquid structure

    influencing the diffusion processes. It is shown that cluster formation in mixtures

    containing one alcoholic component has a significant impact on the diffusion process.

    The estimation of the thermodynamic factor from experimental vapor-liquid equilib-

    rium data is investigated, considering three excess Gibbs energy models, i.e. Wilson,

    NRTL and UNIQUAC. It is found that the Wilson model yields the thermodynamic

    factor that best suits the simulation results for the prediction of the Fick diffusion

    coefficient. Four semi-empirical methods for the prediction of the self-diffusion co-

    efficients and nine predictive equations for the Fick diffusion coefficient are assessed

    and it is found that methods based on local composition models are more reliable.

    Finally, the shear viscosity and thermal conductivity are predicted and in most cases

    favorably compared with experimental literature values.

    Keywords: Fick diffusion coefficient, Maxwell-Stefan, Green-Kubo, thermodynamic

    factor, radial distribution function, shear viscosity, thermal conductivity, model pre-

    diction

    a)Supplementary Materials available.b)Electronic mail: [email protected]

    1

  • Mutual diffusion of binary liquid mixtures

    I. INTRODUCTION

    Almost all separation processes in chemical engineering, such as distillation, absorption

    or extraction, are affected by diffusion in liquids. Traditionally, equilibrium stage methods

    predominate for modeling, design and control of these unit operations because they are

    comparably easy to implement and solve. The weakness of equilibrium approaches is that

    they often yield a solution that is far from the physical process, which is usually corrected for

    with an empirical efficiency factor1. In the last decades, advances in this field have been made

    with the continuous development of rate-based or non-equilibrium methods, which are much

    nearer to physical reality. Nowadays, rate-based methods are employed to solve complex

    modelling and simulation issues in steady and unsteady state operations including start-up

    and shut-down2,3. These non-equilibrium methods involve mass and energy transfer models,

    which require not only diffusion data, but also other transport coefficients like shear viscosity

    and thermal conductivity for pure components as well as mixtures4. Thus, there is a growing

    need for accurate transport properties, which experimental measurements alone are not able

    to satisfy5. Traditionally, transport data have played a lesser role than time independent

    properties, like vapor-liquid equilibria (VLE), so that the availability of experimental data on

    transport coefficients is still low5. Hence, there is an increasing interest in better methods for

    their prediction. Owing to the rapid development of computing power, molecular modelling

    and simulation has emerged as an alternative for such predictions, especially when dealing

    with hazardous substances or challenging thermodynamic conditions6,7.

    Molecular modeling and simulation comprises computational techniques derived from

    quantum chemistry and statistical mechanics to study macroscopic thermodynamic proper-

    ties by means of particle ensembles based on interaction potentials. These force fields provide

    a fundamental molecular-level physical description of the nature of matter. Thus, a detailed

    insight into the physics of equilibrium and non-equilibrium processes can be gained8. Be-

    cause of these microscopic considerations, force field-based simulation methods can be used

    for the understanding and interpretation of experimental results, to obtain predictive esti-

    mates and to inter- or extrapolate experimental data into regions that are difficult to access

    in the laboratory9.

    To describe diffusive mass transport in liquid mixtures, two approaches are commonly

    used: Fick’s law and Maxwell-Stefan (MS) theory1. Both relate a mass flux to a driving

    2

  • Mutual diffusion of binary liquid mixtures

    force1. In the case of Fick’s law, the driving force is expressed in terms of the mole fraction

    gradient ∇xj , which is a quantity that can be measured in the laboratory. The diffusive

    molar flux of component i is then

    Ji = −ρn−1∑

    j=1

    Dij∇xj , (1)

    where n is the number of components in the mixture, ρ is the molar density and Dij denotes

    the Fick diffusion coefficient coupling the flux of component i with the gradient of the mole

    fraction of component j.

    On the other hand, MS theory1 expresses the driving force in terms of the gradient of

    the chemical potential ∇µi, which is assumed to be balanced by a friction force that is

    proportional to the mutual velocity between the components ui − uj

    n∑

    j 6=i=1

    xj(ui − uj)Ðij

    = −∇µi

    kBT, (2)

    where kB is the Boltzmann constant and T the temperature. The MS diffusion coefficient

    Ðij thus plays the role of an inverse friction coefficient between components i and j.

    The MS approach accounts for thermodynamics and mass transfer separately such that

    only the latter contribution is characterized by the MS diffusion coefficient. Because Ðij is

    related to the chemical potential gradient, it cannot directly be measured in the laboratory.

    However, the MS diffusion coefficient can well be sampled by molecular dynamics (MD)

    simulation. It is usually calculated via equilibrium molecular dynamics (EMD) simulation

    from velocity correlation functions with the Green-Kubo formalism or, alternatively, from the

    mean square displacement with the Einstein formalism1,10. The thermodynamic contribution

    is considered by the so-called thermodynamic factor Γ.

    Eqs. (1) and (2) describe the same phenomenon so that a relation between both sets of

    diffusion coefficients exists1. For binary mixtures, because there is only a single independent

    MS and Fick diffusion coefficient, it is simply

    Dij = Ðij · Γ , (3)

    with

    3

  • Mutual diffusion of binary liquid mixtures

    Γ = 1 + x1

    (

    ∂ ln γ1∂x1

    )

    T,p

    = 1 + x2

    (

    ∂ ln γ2∂x2

    )

    T,p

    , (4)

    where γi stands for the activity coefficient of component i. The MS diffusion coefficient can

    thus be transformed to the Fick diffusion coefficient and vice versa, if the thermodynamic

    factor is known.

    The Fick diffusion coefficient can either be determined directly with experimental meth-

    ods, because it is related to the gradient of a measurable quantity, or estimated from molec-

    ular simulation, theoretical, semi-empirical or empirical models. Predictive and empirical

    approaches relate the mutual diffusion coefficients to pure fluid properties or simplify the in-

    teraction between unlike molecules, which may lead to inaccurate approximations for many

    liquid mixtures1. Most methods for the prediction of the composition dependence of the

    Fick diffusion coefficient have been related theoretically and empirically to the diffusion co-

    efficients at infinite dilution or to the self-diffusion coefficients in the form of composition

    weighted averages. The classical interpolation methods by Darken11 and Vignes12 have been

    extended to better consider non-idealities. Leffler and Cullinan13 as well as Carman and

    Stein14 introduced the shear viscosity of the mixture and that of its neat components. The

    concept of local composition according to the Wilson model was employed by Li et al.15

    and Zhou et al.16 to improve Darken’s and Vignes’ models, respectively. Bosse and Bart17

    proposed an excess Gibbs energy (GE) correction term for Vignes’ equation. Several authors

    have introduced modifications of the thermodynamic factor18–23. Other predictive methods

    rely on the knowledge of a variety of properties of the involved pure fluids and/or binary

    adjustable parameters which are often not available24–32.

    The Fick diffusion coefficient can be also calculated directly with molecular simulation

    employing non-equilibrium MD methodologies7,33,34. However, these approaches are com-

    plex, time-consuming and usually require very high concentration gradients, which make

    them impractical to accurately sample the composition dependence of the Fick diffusion co-

    efficient35. Recently, Nichols and Wheeler36 reported a novel Fourier correlation method to

    calculate the Fick diffusion coefficient directly from systems in equilibrium. In the present

    work, EMD simulation and the Green-Kubo formalism were preferred. One of the advan-

    tages of this route is that all transport coefficients can be sampled simultaneously. On the

    other hand, the thermodynamic factor is needed to obtain the Fick diffusion coefficient.

    4

  • Mutual diffusion of binary liquid mixtures

    The thermodynamic factor is usually extracted from experimental VLE data or excess

    enthalpy measurements37,38. For this purpose equations of state34,39, excess Gibbs energy

    GE models or direct numerical integration20–22 can be employed. However, it can also be

    obtained from molecular simulation. E.g., it can be estimated from the integration of the

    radial distribution function (RDF) based on Kirkwood-Buff theory40–47. Another molecular

    simulation approach is to determine the composition dependence of the chemical potential

    using free energy perturbation methods, like Widom’s test particle insertion48,49, thermody-

    namic integration50,51 or gradual insertion52. However, these approaches are challenging in

    case of dense liquids consisting of strongly interacting molecules.

    The advantage of the classical approach to determine the thermodynamic factor is the

    good availability of experimental VLE data. However, this method should be employed

    carefully, taking two important issues into account. First, the thermodynamic factor is sen-

    sitive to the underlying thermodynamic model. In fact, different GE models may describe

    experimental VLE data equally well, but yield different values for the thermodynamic fac-

    tor38,53. Second, the thermodynamic factor determined with this approach corresponds to

    thermodynamic conditions under which the underlying VLE data were measured, i.e. for a

    given temperature the resulting thermodynamic factor may not correspond to the desired

    pressure, but to the varying phase equilibrium pressure. In order to investigate the uncer-

    tainty of the thermodynamic factor introduced by the GE model, the thermodynamic factor

    was calculated here for all studied mixtures with three different GE models, i.e. Wilson54,

    NRTL55 and UNIQUAC56.

    This work relies on the capability of molecular modeling and simulation to predict the

    diffusion coefficients of liquid mixtures. All binary systems that can be formed out of the

    seven components methanol, ethanol, acetone, benzene, cyclohexane, toluene and carbon

    tetrachloride (CCl4) were studied in a fully combinatorial manner, cf. Figure 1. The selection

    of these mixtures was driven by the unusually good availability of experimental transport

    data and by the presence, in many cases, of interesting thermodynamic and structural

    characteristics. Table I provides an overview of the 20 systems that were investigated at

    ambient conditions. The mixture methanol + cyclohexane was not considered because of its

    miscibility gap. Previous simulation results for the Fick diffusion coefficient of methanol +

    ethanol of our group52 were complemented here with more thorough simulations for the sake

    of consistency. Present simulation results were compared, wherever possible, to experimental

    5

  • Mutual diffusion of binary liquid mixtures

    data and to a set of predictive equations. Further, self-diffusion coefficients, shear viscosity

    and thermal conductivity were also predicted.

    FIG. 1. Graphical representation of the binary liquid mixtures studied in this work

    TABLE I. Overview of the binary liquid mixtures studied in this work.

    ethanol acetone benzene cyclo- toluene CCl4

    hexane

    methanol 1 2 3 a 4 5

    ethanol 6 7 8 9 10

    acetone 11 12 13 14

    benzene 15 16 17

    cyclohexane 18 19

    toluene 20

    Group I, Group II, Group III

    a Presence of a miscibility gap at ambient conditions.

    In the present work, rigid and non-polarizable molecular models were used, i.e. a simple

    model class that is suitable to predict thermodynamic and structural properties of fluids,

    including hydrogen-bonding, with a good accuracy, e.g. for water, methanol, ethanol, am-

    monia, dimethylamine and some of their binary mixtures57–59. Further, both alcohol models

    6

  • Mutual diffusion of binary liquid mixtures

    TABLE II. Present results for density, self-diffusion coefficient, shear viscosity and thermal con-

    ductivity of the pure liquids at 298.15 K and 0.1 MPa. a

    substance ρsim ρexp D0i,sim D

    0i,exp η

    0sim η

    0exp λ

    0sim λ

    0exp exp.

    mol L−1 mol L−1 10−9m2s−1 10−9m2s−1 10−4 Pa s 10−4 Pa s Wm−1K−1 Wm−1K−1 ref.

    methanol 24.541 (6) 24.552 2.410 (2) 2.415 5.3 (4) 5.44 0.21 (3) 0.201 64–67

    ethanol 17.129 (3) 17.046 0.974 (3) 1.075 10.87 (6) 10.82 0.190 (9) 0.160 68–71

    acetone 13.537 (3) 13.511 4.538 (8) 4.77 3.0 (2) 3.07 0.162 (6) 0.156 72–74

    benzene 10.283 (2) 10.300 2.239 (4) 2.204 6.1 (3) 6.14 0.123 (6) 0.141 75–78

    cyclohexane 9.220 (1) 9.200 1.524 (4) 1.424 7.7 (4) 8.83 0.144 (5) 0.117 74,79,80

    toluene 9.343 (1) 9.358 2.380 (4) 2.290 5.3 (3) 5.54 0.109 (7) 0.135 68,76,81,82

    CCl4 11.148 (1) 11.183 1.323 (3) 1.305 7.6 (4) 8.90 0.114 (3) 0.10465,75,83,84

    a The number in parentheses indicates the statistical uncertainty in the last given digit.

    have successfully been tested in preceding work for the prediction of transport properties

    in their ternary mixture with water and the according binary subsystems52. The mod-

    els for methanol, ethanol, acetone and cyclohexane were taken from previous work of our

    group60–63, whereas the molecular models for benzene, toluene and CCl4 were devised here.

    These three models were developed starting from quantum mechanical calculations and a

    subsequent optimization of the site-site distances and model parameters to experimental

    VLE and self-diffusion coefficient data following a recently published methodology63. The

    employed molecular models well reproduce the transport properties considered here under

    the relevant thermodynamic conditions. Table II lists the simulation results for self-diffusion

    coefficients, shear viscosity and thermal conductivity of the seven pure fluids. The exper-

    imental values were reproduced with an average relative deviation (ARD) of 3%, 6% and

    13%, respectively.

    Mutual diffusion of binary liquid mixtures by molecular simulation has been the subject of

    several publications. Not only simple mixtures of Lennard-Jones (LJ) spheres33,85–88, but also

    more complex systems of hydrocarbons89 or hydrogen-bonding components90,91 have been

    regarded in this sense using equilibrium and non-equilibrium MD simulations. Among the

    binary mixtures considered in this work, benzene + cyclohexane has been the one that was

    most widely studied by molecular simulation with respect to transport properties. Schaink

    7

  • Mutual diffusion of binary liquid mixtures

    et al.92 and Hoheisel and Würflinger93 calculated Fick and self-diffusion coefficients, shear

    viscosity and thermal conductivity using EMD and rigid molecular models. Zhang and

    Müller-Plathe94 calculated the thermal and transport diffusion coefficients of this mixture

    using reverse-NEMD and EMD methods with flexible all-atom molecular models95. Liu et

    al.35 investigated the MS diffusion coefficient of the binary mixtures methanol + acetone

    and acetone + CCl4 using EMD and rigid molecular models96. The thermodynamic factor

    reported by Liu et al. was determined with the molecular simulation approach proposed

    by Schnell et al.41,42. Wheeler and Rowley90 predicted the shear viscosity of methanol +

    acetone, employing rigid molecular models and non-equilibrium simulation methods. Per-

    era et al.97 also studied this mixture with molecular simulation techniques, however, only

    time independent thermodynamic and structural properties were considered. The diffusion

    coefficient at infinite dilution of the binary mixture cyclohexane + toluene as well as their

    self-diffusion coefficients in their ternary mixture with n-hexane were predicted by Liu et

    al.98,99. We are not aware of any other molecular simulation studies on transport properties

    covering the remaining 16 binary mixtures studied here.

    This paper is organized as follows: First, the simulation methodology is described. Sec-

    ond, the results for the thermodynamic factor, the MS and the Fick diffusion coefficients

    are presented for the 20 studied binary mixtures. The calculated Fick diffusion coefficient is

    compared with experimental data and different predictive methods for mutual diffusion. The

    observed behavior of the diffusion coefficients is analyzed based on the microscopic physical

    structure of the mixtures as provided by radial distribution functions (RDF). Subsequently,

    the predictions for self-diffusion coefficients, shear viscosity and thermal conductivity of the

    mixtures are compared with the available experimental data. Finally, conclusions are drawn.

    A detailed description of the new molecular models for benzene, toluene and CCl4 and the

    technical simulation details are given in the Supporting Information.

    II. MOLECULAR MODELS

    Throughout this work, rigid and non-polarizable molecular models of united-atom type

    were used. The models account for the intermolecular interactions, including hydrogen-bon-

    ding, by a set of LJ sites and superimposed point charges, point dipoles or point quadrupoles

    which may or may not coincide with the LJ site positions. The molecular models for

    8

  • Mutual diffusion of binary liquid mixtures

    methanol, ethanol, acetone and cyclohexane were taken from prior work60–63, whereas the

    models for benzene, toluene and CCl4 were developed here. For detailed information on

    the molecular models the interested reader is referred to the Supplementary Material of this

    work and to the original publications60–63.

    To define a molecular model for a binary mixture on the basis of pairwise additive pure

    substance models, only the unlike interactions have to be specified. In case of polar in-

    teraction sites, this can straightforwardly be done by following the laws of electrostatics.

    However, for the unlike LJ parameters there is no physically sound approach100 and combin-

    ing rules have to be employed for predictions. Vrabec et al.101 have shown in a systematic

    study on 267 binary mixtures that in many cases an adjustable binary parameter is nec-

    essary to describe the VLE with a high accuracy. Thus, the use of such parameters may

    be important to accurately cover the phase behavior of binary mixtures near liquid-liquid

    phase separation, i.e. when Γ → 0. However, in this work, a strictly predictive route was

    followed that exclusively relies on the pure fluid models. Thus, the interactions between LJ

    sites of unlike molecules were specified by the Lorentz-Berthelot combining rules.

    III. METHODOLOGY

    A. Transport properties

    Transport data were sampled by EMD simulation and the Green-Kubo formalism102,103.

    This formalism establishes a direct relationship between a transport coefficient and the time

    integral of the correlation function of the corresponding microscopic flux in a system in

    equilibrium. The general Green-Kubo expression for an arbitrary transport coefficient γ is

    given by

    γ =1G

    ∫ ∞

    0dt〈

    Ȧ(t) · Ȧ(0)〉

    , (5)

    Therein, G is a transport property specific factor, A the related perturbation and Ȧ its time

    derivative. The brackets denote the ensemble average. In case of the self-diffusion

    coefficient, A(t) is the position vector of a given molecule at some time t and Ȧ(t) is its center

    of mass velocity vector. In this way, the self-diffusion coefficient is related to the velocity

    autocorrelation function. On the other hand, the shear viscosity is associated with the time

    9

  • Mutual diffusion of binary liquid mixtures

    autocorrelation function of the off-diagonal elements of the stress tensor and the thermal

    conductivity to the autocorrelation functions for the energy. The detailed expressions can

    be found in the Supporting Information and in previous publications58,104.

    The partial molar enthalpy, which is necessary to calculate the heat flow in a mixture105,

    was determined in two steps. First, the molar enthalpy of the binary mixture h was calcu-

    lated in the isobaric-isothermal (NpT ) ensemble over the entire composition range. Second,

    a third order polynomial h = h(x1) was fitted by a least squares optimization to these data.

    Finally, the partial molar enthalpy was calculated analytically by

    hi = h + xj

    (

    ∂h

    ∂xi

    )

    T,p

    , (6)

    where h represents the molar enthalpy of the mixture at the desired composition.

    B. Thermodynamic factor

    For the calculation of the thermodynamic factor, the composition dependence of the

    activity coefficients is required. Following the Gibbs-Duhem equation, the excess Gibbs

    energy GE of the binary mixture is related to the individual activity coefficients by106

    kBT lnγ1 =

    (

    ∂GE

    ∂n1

    )

    T,p

    and kBT lnγ2 =

    (

    ∂GE

    ∂n2

    )

    T,p

    , (7)

    where ni is the number of moles of component i in the mixture. Therefore, a mathematical

    expression for GE as a function of composition is required to calculate the thermodynamic

    factor from Eq. (3). In this work, three different well-established local composition models

    were considered to estimate the uncertainty of the thermodynamic factor, i.e. Wilson54,

    NRTL55 and UNIQUAC56. These classical GE models require adjustable binary parameters,

    which were regressed to experimental VLE data assuming that the vapor phase is an ideal gas

    and that the Poynting correction factor is negligible. The according regressions were carried

    out with the RecPar tool from the Dortmund Data Bank107. The composition derivatives of

    the activity coefficient models were evaluated analytically to determine the thermodynamic

    factor as described by Taylor and Kooijman38.

    The quality of the underlying experimental VLE data is crucial for the calculation of

    the thermodynamic factor, therefore, only thermodynamically consistent VLE data should

    10

  • Mutual diffusion of binary liquid mixtures

    be employed17. For this purpose both the point to point test108,109 and the integral or

    area consistency test110,111 were applied. The VLE data sets with the largest number of

    measured data points were chosen at or close to the target temperature of 298.15 K. For the

    regarded mixtures no significant changes of the thermodynamic factor were observed for a

    temperature interval of ±10 K around the target temperature for a given GE model. Further,

    the thermodynamic factor was calculated for other VLE data sets, if available, in order to

    confirm the consistency of the data. The parameters of the Wilson, NRTL and UNIQUAC

    models determined for all studied mixtures and their ARD in terms of the vapor pressure

    together with the respective experimental VLE data sources are given in the Supporting

    Information.

    C. Predictive equations

    There are numerous correlations in the literature to determine the composition depen-

    dence of the self-diffusion coefficients in mixtures. Four of these were assessed in this work:

    The correlation by Carman and Stein14, which relates the self-diffusion coefficient in the

    mixture Di with that of the pure liquid D0i , its shear viscosity η and that of the pure com-

    ponent η0i , its correction proposed by Li et al.15, the correlations by Krishna and van Baten89

    and by Liu et al.99, which relate the self-diffusion coefficients with their values at infinite

    dilution and its mass fraction wi or molar fraction xi. These predictive equations are listed

    in Table III.

    Because the present simulations provide MS and self-diffusion coefficients simultaneously,

    a comparison with the classical interpolation approach for the determination of the mutual

    diffusion coefficients suggested by Darken11 is straightforward. Darken’s model takes only

    self-correlations into account, resulting for binary mixtures to11

    Ðij = xi · Dj + xj · Di· (8)

    Hence, Darken’s model is applicable for ideally diffusing mixtures, where the contribution

    of the velocity cross-correlations to the net velocity correlation function is negligible112.

    Further, the limiting values of the mutual diffusion coefficient are given by the self-

    diffusion coefficients

    11

  • Mutual diffusion of binary liquid mixtures

    D∞ij = Dxj→1i and D

    ∞ji = D

    xi→1j · (9)

    Dxj→1i is the self-diffusion coefficient of component i when it is infinitely diluted in compo-

    nent j. It is thus possible to obtain the mutual diffusion coefficient at infinite dilution by

    extrapolation of the self-diffusion coefficients. The logarithmic average proposed by Vignes12

    to predict the composition dependence of the MS diffusion coefficient

    Ðij =(

    D∞ij

    )xj·

    (

    D∞ji

    )xi, (10)

    requires the knowledge of the diffusivities in the infinite dilution limit. Eq. (10) has a simple

    form and it is easy to use so that it is widely applied. However, it may lead to large deviations

    for mixtures containing associating components1.

    There is a variety of methods attempting to improve Darken’s and Vignes’ interpolation

    methods. Among the most widely used are those that relate mutual diffusion with the shear

    viscosity, e.g. by Leffler and Cullinan13 and by Carman and Stein14. Li et al.15 modified

    Darken’s equation using local volume fractions according to the Wilson model, whereas Zhou

    et al.16 applied the same concept to Vignes’ equation. Bosse and Bart17 expanded Vignes’

    equation by an excess Gibbs energy term. On the other hand, D’Agostino et al.20 and Zhu

    et al.23 proposed to rise the thermodynamic factor by an empirical factor α = 0.64. Table III

    gives an overview on the nine predictive equations that were assessed here.

    The shear viscosity of liquid mixtures is in many cases very sensitive to association

    effects among its components. Thus, the shape of the mole fraction dependence of the shear

    viscosity can be linear for ideal mixtures or exhibit a maximum, a minimum or both for

    highly non-ideal mixtures106. Predictive equations are usually interpolative and relate the

    shear viscosity of the mixture with that of its neat components. Because a comprehensive

    discussion of the different predictive shear viscosity equations is out of scope here, present

    simulation results were only compared with experimental data and the widely used relation

    by Grunberg and Nissan113

    ln η = xi · ln η0i + xj · ln η0j + xi · xj · Gij , (11)

    where the binary interaction parameter was set to zero, i.e. Gij = 0, which is termed as

    behavior of the ideal mixture.

    12

  • Mutual diffusion of binary liquid mixtures

    TABLE III. Predictive equations for the mole fraction dependence of self-diffusion, Maxwell-Stefan

    and Fick diffusion coefficients of binary mixtures.

    Self-Diffusion Coefficients ref.

    Di = D0i · η

    0i /η Carman-Stein

    14

    Di = D0i · (η

    0i /η) ·

    (

    n0i /(

    1 +(

    n0i − 1)

    xi))0.5 (a) Li et al. 15

    Di =∑n

    j=1 wj · Dxj→1i Krishna-van Baten

    89

    Di = 1/(∑n

    j=1 xj/Dxj→1i ) Liu et al.

    99

    Mutual Diffusion Coefficients ref.

    Ðij = xi · Dj + xj · Di Darken11

    Ðij =(

    D∞ij

    )xj·

    (

    D∞ji

    )xiVignes 12

    Ðij =(

    D∞ij · η0j

    )xj·

    (

    D∞ji · η0i

    )xi/η Leffler-Cullinan 13

    Ðij =(

    xj · D∞ij · η

    0j + xi · D

    ∞ji · η

    0i

    )

    /η Carman-Stein 14

    Ðij = Di · φjj · v/vj + Dj · φii · v/vi(b) Li et al. 15

    Ðij =(

    D∞ij

    )φjj ·v/vj·

    (

    D∞ji

    )φii·v/vi (b) Zhou et al. 16

    Ðij =(

    D∞ij

    )xj·

    (

    D∞ji

    )xi· exp

    (

    −gE/(RT ))

    Bosse-Bart 17

    Dij = (xj · Di + xi · Dj) · Γ0.64 D’Agostino et al. 20

    Dij = (xjj · Di + xii · Dj) · Γ0.64 (c) Zhu et al. 23

    (a) n0i

    = (η0j

    · D∞ij

    )/(η0i

    · D0i)2

    (b) φii = xi/(xi + xjΛij) and Λij = (vi/vj) exp (−∆λij/(RT )), where ∆λij is the Wilson parameter, vi is the partial

    molar volume of component i and v = xivi + xjvj

    (c) xii = xi/(xi + xjGji) and Gij = exp(−αij ∆gij/(RT )), where αij and ∆gij are NRTL parameters

    Predictive equations for the thermal conductivity of mixtures are usually also based on

    the interpolation of pure component data. In this work, the simple expression by Filippov114

    was chosen for comparison

    λ = wi · λ0i + wj · λ0j + 0.72 wi · wj · (λ

    0j − λ

    0i ), (12)

    13

  • Mutual diffusion of binary liquid mixtures

    where λ and λ0i stand for the thermal conductivity of the mixture and of the pure component

    i, respectively. This equation is expected to predict the thermal conductivity of binary

    mixtures within 5%106.

    IV. RESULTS AND DISCUSSION

    In order to facilitate the discussion of the results obtained in this work, the 20 studied

    mixtures were divided according to their mixing behavior into three groups. This catego-

    rization was based on the maximum deviation of the thermodynamic factor from that of

    the ideal mixture, i.e. less than 10% (group I), up to 45% (group II), and greater than 60%

    (group III). Five mixtures constitute group I: methanol + ethanol, benzene + toluene, ben-

    zene + CCl4, cyclohexane + CCl4 and toluene + CCl4. Group II contains seven mixtures,

    i.e. methanol + acetone, ethanol + acetone, acetone + benzene, acetone + toluene, acetone

    + CCl4, benzene + cyclohexane and cyclohexane + toluene. The remaining eight mixtures,

    methanol + benzene, methanol + toluene, methanol + CCl4, ethanol + benzene, ethanol

    + cyclohexane, acetone + cyclohexane, ethanol + toluene and ethanol + CCl4, form group

    III, cf. Table I.

    A. Density

    The density specified in the simulations to determine the transport properties was ob-

    tained from MD simulations in the NpT ensemble under ambient conditions and compared

    with experimental data. For all studied mixtures, a very good agreement was found between

    simulation and experiment with an ARD of 0.35%. The largest relative deviation of 1.5%

    was found for the strongly non-ideal mixtures containing CCl4. The simulation results of

    six selected mixtures are exemplarily shown in Figure 2 together with experimental values.

    Among these mixtures a diverse density behavior is present, i.e. from “ideal” mixing to large

    positive or negative excess volume. Tabulated numerical simulation data and a graphical

    representation for all 20 mixtures are given in the Supporting Information.

    14

  • Mutual diffusion of binary liquid mixtures

    ���� �������

    ��

    ���� �������

    ��

    ��

    ������������

    ��

    ��� �� �� ��� ��� ���

    ���� �������

    ��

    ��

    ������������

    ��

    ��� �� �� ��� ��� ���

    ��

    ��� ���

    ��� ���

    ��� ���

    FIG. 2. Mole fraction dependence of the density of (a) methanol (1) + ethanol, (b) benzene (1) +

    CCl4, (c) methanol (1) + acetone, (d) benzene (1) + cyclohexane, (e) ethanol (1) + toluene, (f)

    acetone (1) + cyclohexane at 298.15 K and 0.1 MPa. Present simulation results (◦) are compared

    with experimental data (+) as referred to in Table IV. The simulation results by Liu et al.35 (△)

    are also shown.

    B. Structure

    To gain an insight into the underlying microscopic structure RDF were sampled for all

    studied mixtures. The RDF gA−B(r) between like and unlike sites were calculated for the

    pure liquids and mixtures at different compositions. Further, the running coordination

    number NA−B(r) between the sites A and B, was determined from the integral of the RDF

    NA−B(r) = 4π ̺∫ r

    0r2gA−B(r) dr, (13)

    where r is the distance from the reference site and ̺ is the bulk number density of site B.

    The RDF of all studied mixtures are given in the Supporting Information. Exemplarily,

    the RDF of six selected mixtures are shown in Figures 3 to 5. Relevant structural aspects

    15

  • Mutual diffusion of binary liquid mixtures

    TABLE IV. Sources of experimental binary liquid mixture data on density ρ, self-diffusion coeffi-

    cients Di, Fick diffusion coefficient Dij , shear viscosity η and thermal conductivity λ.

    system ρ Di Dij η λ

    no.

    1 72,115–117 118 52 117,119 120

    2 72,117,121 122 117,121 123

    3 64,124 118,125 118,126 127–131

    4 129,130,132,133 134,135 129–131,136

    5 130,131,137,138 65,139 65,140,141 130,131,137,138,142–144 145,146

    6 117,147 117,147 120

    7 64,148,149 150 16,150–152 148,153–155

    8 117,156 157,158 117,143,155,159 120

    9 68 158 160,161 162

    10 163 164 140,141,158,165,166 144,165 120

    11 167,168 73,169 73,151,170 171–173

    12 117,174,175 176 117,175

    13 133 177 133,173,178,179 145

    14 117 164 16,151,170,180 117,181 120

    15 79,138,182–185 125,186,187 79,183,188–190 155,182,184,185,189,191–197 198

    16 79,183,199 79,183 199–202 198

    17 75,203 150,164 118,170,204 144,191,201,204,205

    18 79,183,206 207 79,135,183 197,202,206 198

    19 79,117,183,208,209 79,165,183,210 117,138,144,165,191,194,196,197,211 120

    20 76,212 76 76 76,202,212,213 198

    provided by the RDF analysis are discussed together with the results for mutual diffusion.

    C. Thermodynamic factor

    The thermodynamic factor was calculated with the Wilson, NRTL and UNIQUAC models

    for all mixtures. The model parameters, determined from adjustments to carefully selected

    VLE data, yield relative deviations in terms of the vapor pressure of below 2% in all cases. It

    is well known that GE models that fit the same VLE data set equally well may lead to quite

    different values for the thermodynamic factor38. In the present work, the thermodynamic

    factor calculated with the different models differs by less than 2% for the mixtures in group

    16

  • Mutual diffusion of binary liquid mixtures

    ����

    �����

    � �� � �� �

    ����

    ����

    ������ � � �

    ����

    ������ � � �

    ����

    ������ � � �

    ����

    ������

    ������ � � �

    ������

    ������

    ������ � � �

    ������

    �����

    � �� � �� �

    ������

    ������ � � �

    ������

    ���

    ���

    ���

    ���

    ���

    ���

    FIG. 3. Selected radial distribution functions and the corresponding running coordination numbers

    (inset) of methanol (1) + ethanol (left) and toluene (1) + CCl4 (right) at 298.15 K and 0.1 MPa

    between (a) the oxygen and hydroxyl hydrogen sites of methanol gO−H, (b) the hydroxyl hydrogen

    sites of methanol and ethanol gH−H, (c) the oxygen and hydroxyl hydrogen sites of ethanol gO−H,

    (d) the methine sites of toluene gCH−CH, (e) the methine and chlorine sites of toluene and CCl4

    gCH−Cl, (f) the chlorine sites of CCl4 gCl−Cl. Data for pure methanol, ethanol, toluene and CCl4

    (· · · ) as well as the mixtures with x1 = 0.1 (–), 0.5 (–) and 0.9 mol mol−1 (–) are depicted.

    I and by less than 4.5% for the mixtures in group II. Therefore, the related error is not

    expected to exceed 5% for these groups. The consistency of the thermodynamic factor

    calculated with the three different models breaks down for the mixtures in group III, which

    can be considered as thermodynamically more challenging. For these mixtures, the difference

    among the calculated thermodynamic factor data can even achieve one order of magnitude if

    the thermodynamic factor approaches zero, being near liquid-liquid phase separation, which

    may translate to significant errors when the Fick diffusion coefficient is determined. Figure 6

    exemplarly shows the uncertainty of the thermodynamic factor introduced by the GE model,

    17

  • Mutual diffusion of binary liquid mixtures

    ����

    ����

    � �� � �� �

    ������

    �����

    ����� � � � �

    ����

    ����� � � � �

    �����

    ����� � � � �

    ������

    ����

    ��

    ��

    ��

    ����� � � � �

    ����

    ����

    ����� � � � �

    ����

    ����

    � �� � �� �

    ����

    ����� � � � �

    ����

    ���

    ���

    ���

    ���

    ���

    ���

    FIG. 4. Selected radial distribution functions and the corresponding running coordination numbers

    (inset) of methanol (1) + acetone (left) and acetone (1) + benzene (right) at 298.15 K and 0.1

    MPa between (a) the oxygen and hydroxyl hydrogen sites of methanol gO−H, (b) the oxygen sites

    of methanol and acetone gO−O, (c) and (d) the oxygen sites of acetone gO−O, (e) the oxygen and

    methine sites of acetone and benzene gO−CH, (f) the methine sites of benzene gCH−CH. Data for

    pure methanol, acetone and benzene (· · · ) as well as for the mixtures with x1 = 0.1 (–), 0.5 (–)

    and 0.9 mol mol−1 (–) are depicted.

    indicated as a shaded area, for six selected mixtures belonging to the three introduced groups.

    As can be seen, the thermodynamic factor may change significantly, not only in magnitude

    but also in shape when different GE models are used.

    Further uncertainties of the thermodynamic factor may arise when different methodolo-

    gies are used to obtain it. To illustrate this point, the thermodynamic factor sampled with

    molecular simulation by Liu et al.35 is compared in Figure 6 with the present values for

    methanol + acetone. Accordingly, a larger uncertainty of the thermodynamic factor could

    18

  • Mutual diffusion of binary liquid mixtures

    ����

    ��

    ��

    ��

    ��

    ����

    � � �� �� �� ��

    ��������

    �������

    ����� � � � �

    ����

    ����� � � � �

    �������

    ����� � � � �

    ��������

    ����

    ��

    ��

    ��

    ��

    ����� � � � �

    ����

    ������

    ����� � � � �

    ������

    ����

    � � �� �� �� ��

    ������

    ����� � � � �

    ������

    ���

    ���

    ���

    ���

    ���

    ���

    FIG. 5. Selected radial distribution functions and the corresponding running coordination numbers

    (inset) of methanol (1) + benzene (left) and ethanol (1) + cyclohexane (right) at 298.15 K and

    0.1 MPa between (a) the oxygen and hydroxyl hydrogen sites of methanol gO−H, (b) the methyl

    and methine sites of methanol and benzene gCH3−CH, (c) methine sites of benzene gCH−CH, (d)

    the oxygen and hydroxyl hydrogen sites of ethanol gO−H, (e) the methyl and methylene sites of

    ethanol and cyclohexane gCH3−CH2, (f) the methylene sites of cyclohexane gCH2−CH2. Data for

    pure methanol, ethanol, benzene and cyclohexane (· · · ) as well as for the mixtures with x1 = 0.1

    (–), 0.3 (–), 0.5 (–) and 0.9 mol mol−1 (–) are depicted.

    be inferred, however, the calculations by Liu et al.35 are based on molecular models and

    not on experimental data. On the other hand, Moggridge22 determined the thermodynamic

    factor via a piecewise fit of experimental vapor pressure data in three regions followed by

    numerical integration. The difference between the thermodynamic factor calculated here

    and that by Moggridge22 suggests an increase of approximately 5% in terms of the thermo-

    dynamic factor uncertainty for mixtures in group III, cf. Figure 6.

    To select a model for the thermodynamic factor that is most suitable for the present

    19

  • Mutual diffusion of binary liquid mixtures

    FIG. 6. Mole fraction dependence of the thermodynamic factor of (a) benzene (1) + CCl4, (b)

    cyclohexane (1) + CCl4, (c) methanol (1) + acetone, (d) acetone (1) + CCl4, (e) methanol (1)

    + CCl4, (f) ethanol (1) + benzene. The shaded area represents the range of the results of the

    three considered GE models. The thermodynamic factor obtained via Kirkwood-Buff integrals35

    (△) and via direct numerical integration of experimental VLE data22 (�) is also shown.

    simulation results, the Fick diffusion coefficient was determined with the data from the

    three GE models and compared with the available experimental data. It was found that the

    Fick diffusion coefficient calculated on the basis of the Wilson model yields, on average, the

    smallest deviation from the experiment data for all regarded mixtures, i.e. 16%. Therefore,

    the Wilson thermodynamic factor was chosen to be applied in the following.

    D. Mutual diffusion coefficients

    A set of nine predictive equations for the mutual diffusion coefficients based on Vignes’

    and Darken’s models was assessed. For a fair comparison, all predictive equations were eval-

    uated using the present thermodynamic factor from the Wilson model and its parameters.

    20

  • Mutual diffusion of binary liquid mixtures

    In case of the Darken based equations, the values of the self-diffusion coefficients in the

    mixture were taken from the best polynomial fit of the experimental values, if available, or

    from present molecular simulation values otherwise. For the assessment of the predictive

    diffusivity equations that include a shear viscosity correction13,14, a polynomial function of

    third order was fitted to the available experimental data sets in order to minimize data

    scatter inaccuracies. In case of Vignes’ based equations, which require the mutual diffu-

    sion coefficients at infinite dilution D∞ij and D∞ji as an input, the average of the available

    experimental data was employed.

    The diffusion coefficients at infinite dilution were estimated from molecular simulation

    results by extrapolation of the corresponding self-diffusion coefficients in the dilute region to

    the limiting value at vanishing concentration. Note that infinite dilution diffusion coefficients

    may also be obtained from the extrapolation of the MS diffusion coefficient, but self-diffusion

    coefficient data were preferred because of their inherently lower statistical uncertainty. On

    the other hand, semi-empirical relationships like the Wilke-Chang equation214 can also be

    applied for such predictions. The diffusion coefficients at infinite dilution predicted by

    simulation yield in general a better agreement with experimental data than those predicted

    by the Wilke-Chang equation. The overall ARD from experimental data is 8% for the

    simulation results compared with 19% for the Wilke-Chang equation, which is recommended

    by Poling and Prausnitz106 for non-aqueous mixtures. Present numerical values together with

    experimental data from the literature and predictions from the Wilke-Chang equation214 are

    listed in the Supporting Information.

    Among the nine tested predictive models, those by Zhu et al.23, Zhou et al.16 and Li et

    al.15 were found to be the best three with overall ARD from polynomial fits to experimental

    data for 19 binary mixtures of 12%, 13% and 15%, respectively. The model by D’Agostino et

    al.20 follows with an overall ARD of slightly above 15%, whereas the remaining five models

    yield an overall ARD between 19 and 25%.

    For all 20 mixtures, the Fick diffusion coefficient was determined by molecular simulation

    in combination with the Wilson thermodynamic factor for 11 different mole fractions, cov-

    ering the entire composition range, and was compared with the available experimental data.

    In general, a good agreement was found between simulative predictions and experimental

    data, having an overall ARD of 16%. Figure 7 depicts the ARD of the present simulation

    results for the Fick diffusion coefficient. All data are plotted and the numerical values are

    21

  • Mutual diffusion of binary liquid mixtures

    listed in the Supporting Information. In the following, the results for the MS and Fick

    diffusion coefficients are analyzed in detail.

    �������

    � �� � � �� ��

    �����������������

    �����������������

    �����������������

    �����������������

    ��������������

    �����������������

    �����������������

    ���������������������

    �����������������

    ��������������

    �����������������

    ���������������������

    �����������������

    ��������������

    ���������������������

    �����������������

    ��������������

    ���������������������

    ������������������

    ��������������

    ����� �!!�"�������!!������

    "���#�$�"��"���

    ���#������ ����$���

    FIG. 7. Average relative deviation (ARD) of present simulation results for Fick diffusion coefficient,

    shear viscosity and thermal conductivity from the best polynomial fit of the available experimental

    data.

    1. Group I

    The mixtures methanol + ethanol and benzene + toluene exhibit an almost ideal behavior

    because their components have very similar molecular structures. Thus, the MS diffusion

    coefficient is almost a linear function of the mole fraction and corresponds to the Fick

    diffusion coefficient because Γ ≃ 1, cf. Figure 8. In these simple cases, the interpolation

    equations by Darken11 and Vignes12 are able to accurately predict the mutual diffusion

    coefficient with deviations below 1%. The RDF of nearly ideal mixtures are expected to

    be similar215, and indeed, the RDF sampled here for both mixtures show solvation shells

    located at practically the same distances with similar magnitudes that undergo only small

    changes when the composition is varied, explaining the linear behavior of the MS, Fick and

    self-diffusion coefficients, cf. Figure 3.

    22

  • Mutual diffusion of binary liquid mixtures

    ��������������������

    �������

    ����������������

    ����������������������

    ������������

    ��

    �� �� �� �� �� ��

    ����������������������

    ������������

    ��

    �� �� �� �� �� ��

    ���

    ���

    ���

    ���

    FIG. 8. Mole fraction dependence of the Maxwell-Stefan (left) and Fick (right) diffusion coefficients

    of (a) benzene (1) + toluene, (b) benzene (1) + CCl4, (c) cyclohexane (1) + CCl4, (d) toluene (1) +

    CCl4 at 298.15 K and 0.1 MPa. The simulation results for the Maxwell-Stefan diffusion coefficient

    (•) are compared with the models by Darken11 (◦), Vignes12 (−−), Li et al.15 (− ⋄ −) and Zhou et

    al.16 (–) based on present simulation data. The simulation results for the Fick diffusion coefficient

    (•) are compared with experimental data (+) as referred to in Table IV. The models by Li et al.15

    (−⋄−), Zhou et al.16 (–) and Zhu et al.23 (−⋄−) based on present simulation data are also shown.

    The components of the remaining mixtures of this group, i.e. benzene + CCl4, cyclohex-

    ane + CCl4 and toluene + CCl4, do not have similar molecular structures. However, they

    behave nearly ideal mainly because their components have similar sizes and interactions,

    which is reflected by their RDF. Exemplarily, Figure 3 shows the RDF between toluene-

    toluene methine sites gCH−CH, CCl4-CCl4 chlorine sites gCl−Cl and the unlike sites gCH−Cl

    23

  • Mutual diffusion of binary liquid mixtures

    at three different toluene mole fractions of 0.1, 0.5 and 0.9 mol mol−1. Here, the peaks of

    the first and second solvation shells are found at similar distances, i.e. around 4 and 6 Å.

    The running coordination numbers indicate a rather homogeneous structure and the small

    changes of the RDF for different compositions suggest an insensitivity of the structure upon

    mixing because of the similarity of the interaction sites. The homogeneous distribution of

    the molecules in the mixture can clearly be seen in the snapshots shown in Figure 9. For

    these mixtures, the mole fraction dependence of the MS diffusion coefficient is only slightly

    above a linear course. The equations by Darken and Vignes are thus still adequate with

    deviations of 3 and 4%, respectively. However, other predictive equations, e.g. by Li et al.15

    or Zhou et al.16, achieve an even better accuracy. Consequently, the Fick diffusion coeffi-

    cient shows only slight deviations from a linear behavior, cf. Figure 8. The ARD between

    simulation and experiment is only 6% for this group of mixtures.

    FIG. 9. Snapshots of (a) toluene (1) + CCl4, (b) acetone (1) + benzene, (c) ethanol (1) +

    cyclohexane at 298.15 K, 0.1 MPa and three mole fractions x1 = 0.1 (left), 0.5 (center) and 0.9 mol

    mol−1 (right). At mole fractions of 0.1 and 0.9 mol mol−1 the solvent molecules are not depicted

    to improve visibility. The methyl and methylene groups are shown in orange, the methine groups

    in brown, the oxygen atoms in red and the chlorine atoms in green.

    24

  • Mutual diffusion of binary liquid mixtures

    2. Group II

    Mixtures in this group show a moderate deviation from ideality in terms of Γ. In general,

    the mole fraction dependence of the MS diffusion coefficient is not linear, but exhibits

    a convex curvature which is centered near the equimolar composition in most cases, cf.

    Figure 10. This curvature can be explained with the presence of some grade of association

    between the molecules, like solvation. In case of the mixtures of acetone with benzene and

    toluene, the RDF between the methyl and the oxygen sites of acetone gCH3−O shows a sharp

    main peak, which indicates the presence of an ordered nearest-neighbor structure related

    to the tendency of acetone molecules to associate into dimers216. Further, the RDF remain

    alike when the composition is changed. Thus, the packing structure of the solvents changes

    little with composition, indicating that acetone tends to stay in segregated pockets216, cf.

    Figure 4. This fact can also clearly be observed in the simulation snapshots for acetone +

    benzene shown in Figure 9 and explains the moderate decrease of the self-, Fick and MS

    diffusion coefficients.

    For mixtures of acetone with methanol and ethanol, the composition dependence of the

    MS diffusion coefficient has a maximum at an alcohol mole fraction of around 0.2 mol

    mol−1. This is related to the pronounced presence of alcohol association at this composition

    as revealed by the corresponding RDF. The main peak of the RDF between the oxygen

    and hydroxyl hydrogen sites of the alcohol gO−H, which is related to the hydrogen-bonding

    structure, becomes sharper and higher when acetone is added, cf. Figure 4. The relatively

    low main peak observed for the mixtures with a high alcohol content is a result from sta-

    tistical standardization, i.e. more methanol molecules can be found in the far range of the

    simulation volume217. However, the increase of the first neighbors peak with decreasing

    methanol concentration is important, suggesting the enhancement of the self-associating

    alcohol structure97. Further, shape and amplitude of the peaks do not change with compo-

    sition, indicating that at low concentrations the alcohol molecules form clusters which are

    surrounded by acetone218. These findings lead to the observed maximum of the MS diffusion

    coefficient. Therefore, deviations of the MS diffusion coefficient from Vignes’ and Darken’s

    interpolation methods become important.

    The mole fraction dependence of the Fick diffusion coefficient exhibits a concave cur-

    vature, showing a decrease of up to 30% from ideal diffusion behavior, depending on the

    25

  • Mutual diffusion of binary liquid mixtures

    �������

    ��������������

    ������

    ���������������

    ������������

    ��

    ��� ��� �� ��� ��� ���

    ���������������������

    ������������

    ��

    ��� ��� �� ��� ��� ���

    ��������������������

    ���

    ���

    ���

    ���

    FIG. 10. Mole fraction dependence of the Maxwell-Stefan (left) and Fick (right) diffusion coeffi-

    cients of (a) methanol (1) + acetone, (b) ethanol (1) + acetone, (c) acetone (1) + benzene, (d)

    benzene (1) + cyclohexane at 298.15 K and 0.1 MPa . The simulation results for the Maxwell-

    Stefan diffusion coefficient (•) are compared with the models by Darken11 (◦), Vignes12 (−−), Li

    et al.15 (− ⋄ −) and Zhou et al.16 (–) based on present simulation data. The simulation results

    for the Fick diffusion coefficient (•) are compared with experimental data (+) as referred to in

    Table IV and to the simulation results by Liu et al.35 (△). The models by Li et al.15 (− ⋄−), Zhou

    et al.16 (–) and Zhu et al.23 (− ⋄ −) based on present simulation data are also shown.

    thermodynamic factor. Here, the Darken based methods by Li et al.15 and D’Agostino et

    al.20 achieve the best agreement with experimental values, the ARD is approximately 5%.

    Present simulation results are in good agreement with experimental data for the mixtures

    acetone + benzene, acetone + toluene, benzene + cyclohexane and cyclohexane + toluene

    with ARD between 4 and 10%. Unfortunately, the present simulation results poorly predict

    26

  • Mutual diffusion of binary liquid mixtures

    the Fick diffusion coefficient of the mixture acetone + CCl4 in the acetone-rich composition

    range, resulting in an ARD of approximately 45% for this mixture. Here, the displacement

    of the peaks of the unlike RDF gCH3−Cl towards larger distances suggests that intermolecular

    interactions debilitate when CCl4 is added. This observation may explain the present over-

    estimation of the Fick diffusion coefficient that increases the overall ARD of the simulation

    data for this group to 15%.

    To the best of our knowledge, there are no experimental Fick diffusion coefficient data

    of the mixture ethanol + acetone. As mentioned above, the shape of the mole fraction

    dependence of the MS diffusion coefficient indicates the presence of self-association at low

    ethanol concentrations, similar to the one observed from experimental methanol + acetone

    data, cf. Figure 10.

    3. Group III

    Most systems containing one alcoholic component exhibit a well pronounced peak in the

    mole fraction dependence of the MS diffusion coefficient with a maximum located between

    0.2 and 0.3 mol mol−1 of alcohol content, cf. Figure 11. This sharp increase of the MS

    diffusion coefficient at low alcohol concentration is typical for this group of mixtures and is

    related to cluster formation due to solute self-association219. The MS diffusion coefficient of

    acetone + cyclohexane exhibits a pronounced peak centered at an acetone mole fraction of

    0.4 mol mol−1, indicating significant self-association. The presence of clusters due to self-

    association can also be inferred from the RDF. The sharpness and magnitude of the main

    double peak, corresponding to the hydrogen-bonding structure of the alcohol gO−H, increases

    significantly as the alcohol is depleted, cf. Figure 5. This fact suggests that the nearest

    neighbor hydrogen-bonding structure at low alcohol concentration is more stable than that

    of the pure alcohol. Moreover, the insensitivity of the location of the RDF peaks to a

    change in composition indicates that the alcohol molecules conserve their local environment

    of nearest neighbors of the neat liquid, supporting the thesis of the presence of strong alcohol

    self-association in clusters, causing the maximum of the MS diffusion coefficient. The running

    number of nearest oxygen neighbors around the hydrogen site has a well defined step-like

    form with a plateau at approximately unity. It is noteworthy that upon mixing, the value at

    which the plateau is reached remains almost constant at least up to an alcohol mole fraction

    27

  • Mutual diffusion of binary liquid mixtures

    �������

    ��������������

    ��

    ��

    ��������������������

    ��������������������

    ��

    ��

    ������������

    ��

    ��� ��� �� ��� ��� ���

    ����������������������

    ������������

    ��

    ��� ��� �� ��� ��� ���

    ���

    ���

    ���

    ���

    FIG. 11. Mole fraction dependence of the Maxwell-Stefan (left) and Fick (right) diffusion coeffi-

    cients of (a) methanol (1) + benzene, (b) methanol (1) + toluene, (c) ethanol (1) + benzene, (d)

    acetone (1) + cyclohexane at 298.15 K and 0.1 MPa. The simulation results for the Maxwell-Stefan

    diffusion coefficient (•) are compared with the models by Darken11 (◦), Vignes12 (−−), Li et al.15

    (− ⋄−) and Zhou et al.16 (–) based on present simulation data. The simulation results for the Fick

    diffusion coefficient (•) are compared with experimental data (+) as referred to in Table IV. The

    models by Li et al.15 (−⋄−), Zhou et al.16 (–) and Zhu et al.23 (−⋄−) based on present simulation

    data are also shown.

    of 0.3 mol mol−1. For the lowest alcohol concentration, the coordination number is usually

    lower, suggesting a smaller size of the alcohol clusters. This conclusion is supported by the

    analysis of the RDF between the like sites of the solvents. Again, it can be observed that the

    structure of the pure liquid remains almost unchanged for alcohol mole fractions of 0.1, 0.3

    28

  • Mutual diffusion of binary liquid mixtures

    and 0.5 mol mol−1, but looses its long range behavior for 0.9 mol mol−1. Here, the structure

    beyond the first solvation shell exhibits less stable long range structures. These findings are

    supported by the analysis of the simulation snapshots of this type of mixtures, which back

    up the theory of Pozar et al.216 that species are microsegregated and explain the low values

    of the Fick diffusion coefficient, cf. Figure 9.

    Usually, the maximum of the mole fraction dependence of the MS diffusion coefficient

    coincides with the presence of a dominating minimum in the mole fraction dependence of

    the Fick diffusion coefficient, which implies slower diffusion due to molecular association.

    The differences in the location of the minimum can be traced back to the contribution

    of the thermodynamic factor to the Fick diffusion coefficient. In general, the presence of

    strong non-idealities in these mixtures is a challenge for predictive equations and molecular

    simulation techniques because of the proximity of liquid-liquid phase separation at certain

    compositions. The predictive equations by Zhu et al.23 and Zhou et al.16 yield the best results

    for this group, with ARD of 19 and 22% from the experimental values, respectively. The

    other predictive equations yield ARD above 30%. Present molecular simulation results for

    the individual mixtures in this group deviate between 15 and 20% from experimental data,

    except for the mixtures containing CCl4. The Fick diffusion coefficient in the CCl4-poor

    region was strongly overpredicted, leading to an overall ARD of 29% for this group.

    E. Self-diffusion coefficients

    Self-diffusion coefficients of the individual species in their binary mixture were predicted

    in this work with an estimated statistical uncertainty between 1 and 2%. Figure 12 shows

    present simulation values for six selected mixtures together with experimental data and

    the two predictive equations that were found to be the best in the present assessment.

    According figures for all mixtures can be found in the Supporting Information together with

    the numerical simulation data.

    Four semi-empirical equations were assessed for the prediction of the self-diffusion coeffi-

    cients. However, none of the them was found to be satisfactory for all considered mixtures.

    The model by Liu et al.99, based on a development of the linear response theory, yields the

    best overall agreement with the available experimental data for eleven mixtures. Its overall

    ARD is 14% for Di and 5% for Dj, where i is the component with the lower molecular

    29

  • Mutual diffusion of binary liquid mixtures

    �����������

    ��

    �����������

    ��

    �����������

    ��

    �� �������

    ��

    ��� �� ��� ��� ��� ���

    ��� ���

    ��� ���

    ��� ���

    �� �������

    ��

    ��� �� ��� ��� ��� ���

    ����������

    � �

    ���

    FIG. 12. Mole fraction dependence of the self-diffusion coefficients of (a) benzene (1) + CCl4, (b)

    toluene (1) + CCl4, (c) acetone (1) + benzene, (d) benzene (1) + cyclohexane, (e) methanol (1)

    + benzene, (f) ethanol (1) + benzene at 298.15 K and 0.1 MPa. Present simulation results for

    components (1) (•) and (2) (•) are compared with experimental data (+) as referred to in Table IV.

    mass. Comparatively, present predictions by molecular simulation yield deviations from

    experiment of 9% and 7%, respectively.

    1. Group I

    The self-diffusion coefficients of these mixtures vary almost linearly with the mole fraction

    and the lighter component always shows a higher mobility, cf. Figure 12. As expected, the

    simple interpolative equations by Krishna and van Baten89 and Liu et al.35 yield on average

    predictions within 3% of the experimental values. The ARD of present simulation results

    from experimental data are 6 and 3% for Di and Dj, respectively.

    30

  • Mutual diffusion of binary liquid mixtures

    2. Group II

    For mixtures belonging to this group, a departure from linearity for the self-diffusion

    coefficients can be observed. Therefore, the ARD increases up to 6% for the interpolative

    models by Krishna and van Baten89 and Liu et al.99, similarly to present molecular simulation

    values. In case of acetone + benzene, acetone + toluene, acetone + CCl4 and benzene +

    cyclohexane, the self-diffusion coefficient of the lighter component is higher in the whole

    composition range, cf. Figure 12. Here, the influence of molecular size on molecular mobility

    prevails over any association effect. In case of methanol + acetone, ethanol + acetone and

    cyclohexane + toluene, molecular association overcomes the size effect such that the lighter

    component propagates slower than the heavier one at least in one part of the composition

    range.

    3. Group III

    As expected, the largest deviations between simulation and experiment were found for

    mixtures in this group. For the alcoholic mixtures, the self-diffusion coefficient of the lower

    mass component, i.e. the alcohol, is in a large composition range significantly smaller than

    that of the heavier component. The slower movement of the lighter molecules in the mixture

    can be explained by the presence of association among the alcohol molecules, which increases

    their effective diameter and hinders their mobility. The strong self-association of the alcohol

    molecules into clusters explains the sharp decrease on the alcohol self-diffusion coefficients

    and may be inferred from the RDF, cf. Figure 5. Only molecular simulation is able to predict

    this complex molecular behavior, while all four tested predictive equations fail. Among

    the predictive equations, the relation by Li et al.15 yields the best agreement with the

    experimental data. This model, which is highly dependent on shear viscosity data, predicts

    deviations from linearity of self-diffusion in non-ideal mixtures, however, in many cases the

    self-diffusion coefficients exhibit a strong overcorrection, yielding an ARD of 25% for Di.

    The inadequacy of the model by Liu et al.99 for this kind of mixtures exposes its major

    weakness, i.e., the assumption of negligible velocity cross-correlations.

    Figure 12 shows the simulation results for the mole fraction dependence of the self-

    diffusion coefficients of six selected mixtures compared with experimental values and the

    31

  • Mutual diffusion of binary liquid mixtures

    results from the relations proposed by Li et al.15 and Liu et al.99. In case of the self-diffusion

    coefficients of methanol + benzene and ethanol + benzene, only molecular simulation is able

    to qualitatively correctly predict the composition dependence of the alcohol self-diffusion

    coefficient.

    F. Shear viscosity

    Although the simulation results for the Fick diffusion coefficient may deviate strongly

    from experimental data in particular cases, the predicted shear viscosity does usually agree

    qualitatively and quantitatively with the experimental data, having an overall ARD of only

    8% from the best polynomial fit of the experimental data given in Table IV. The ARD

    for each mixture are represented graphically in Figure 7. Figure 13 exemplarily shows the

    predicted shear viscosity for six selected mixtures together with experimental data and the

    ideal mixture model by Grunberg and Nissan113, cf. Eq. (11). Analogous Figures for all

    mixtures are given in the Supporting Information together with the numerical data.

    1. Group I

    As expected, the shear viscosity of these mixtures shows a nearly linear mole fraction

    dependence so that it can be accurately predicted by the ideal mixture model by Grunberg

    and Nissan191, cf. Figure 13. Present molecular simulation results deviate with an overall

    ARD of 10% for this group, mainly because of the underestimation of the shear viscosity

    of pure CCl4. If this offset would be corrected, a significant improvement of the agreement

    between simulation and experiment is expected.

    2. Group II

    The shear viscosity of these mixtures shows some negative deviation from ideal behavior.

    Especially for methanol + acetone, ethanol + acetone, benzene + cyclohexane and cyclo-

    hexane + toluene, a significant negative deviation from the ideal behavior is observed. Here,

    the shear viscosity of the mixture is lower than that implied by their individual components,

    which is related to their differences in size and shape and to weak unlike intermolecular

    32

  • Mutual diffusion of binary liquid mixtures

    interactions220. Present simulation work was able to predict the shear viscosity composition

    dependence with an overall ARD of 7% for this group.

    3. Group III

    In this group of mixtures, positive and negative deviations from the ideal shear viscosity

    behavior were found. The binary mixtures of methanol with benzene, toluene and CCl4

    exhibit a positive deviation from ideality, cf. Figure 13. This is most pronounced for

    methanol + benzene, where the shear viscosity is 50% higher than its ideal value around

    equimolar composition, indicating strong intermolecular interactions, which is related to

    the presence of alcohol self-association leading to clusters. The negative deviation from

    ideal behavior found for the mixtures ethanol + benzene, ethanol + cyclohexane, ethanol +

    toluene and acetone + cyclohexane is a consequence of the combination of the interacting and

    non-interacting forces221 between unlike molecules. With the exception of methanol + CCl4,

    present simulations are able to predict the composition dependence of the shear viscosity

    well also for mixtures belonging to this group, where the overall ARD from experimental

    values is 7%.

    4. Data discrimination

    One of the powerful applications of molecular simulation is data discrimination. There-

    fore, inconsistencies found for several mixtures among experimental literature data are dis-

    cussed in the following.

    In case of methanol + benzene, five experimental data sets were found at ambient condi-

    tions, three sets127–129 disagree with the other two130,131. While the former three suggest a

    nearly ideal behavior, present simulation results predict a strongly non-linear shear viscosity

    mole fraction dependence, which is consistent with the data sets by Rathore at al.130 and

    Goyal et al.131, thus backing up their results. A similar observation was made for methanol

    + toluene, where four different experimental data sets are available. The composition de-

    pendence of the shear viscosity of three of them has a convex shape129,131,136, whereas the

    fourth one130 exhibits a concave shape, cf. Figure 13. In this case, the simulation results do

    not show a strong deviation from ideality and agree well with the data sets by Han et al.129,

    33

  • Mutual diffusion of binary liquid mixtures

    �� ����������

    ��

    �� ���������

    ��� �� ��� ��� ��� ���

    �������������

    �� �������

    ��

    ��� �� ��� ��� ��� ���

    �� ����������

    ��� ���

    ��� ���

    ��� ���

    FIG. 13. Mole fraction dependence of the shear viscosity of (a) methanol (1) + ethanol, (b)

    benzene (1) + toluene, (c) acetone (1) + toluene, (d) benzene (1) + cyclohexane, (e) methanol

    (1) + benzene, (f) methanol (1) + toluene at 298.15 K and 0.1 MPa. Simulation results for the

    shear viscosity (•) are shown together with the viscosity of the ideal mixture (−−). Experimental

    data: (+)117,119 for methanol + ethanol, (+)200,201, (+)202 for benzene + toluene; (+)178, (+)179,

    (+)133 for acetone + toluene; (+)138,155,185,194,197, (+)202 for benzene + cyclohexane; (+)130, (+)131,

    (+)129 for methanol + benzene; (+)130, (+)131, (+)129, (+)136 for methanol + toluene.

    Goyal et al.131 and Wanchoo et al.136, reducing the credibility of the data set by Rathore

    et al.130. For acetone + toluene, the experimental data set by Rajagopal et al.133 indicates

    a quite significant deviation from ideality. Molecular simulation results and other experi-

    mental data sets173,178,179 show a fairly linear behavior of the shear viscosity. For benzene

    + cyclohexane, benzene + toluene, cyclohexane + toluene and toluene + CCl4, the experi-

    mental data by Pandey et al.202 are always higher than other data sets from the literature

    and than the present simulation results. In case of benzene + cyclohexane and cyclohexane

    + toluene, again the data by Pandey et al.202 suggest a rather ideal behavior, where present

    results and other experimental data indicate a negative deviation from ideality. In case of

    34

  • Mutual diffusion of binary liquid mixtures

    toluene + CCl4, further inconsistencies among experimental data sets were found, present

    results show a good agreement with the experimental data by Reddy et al.212.

    G. Thermal conductivity

    All present simulation results for the thermal conductivity are listed and plotted in the

    Supporting Information. Figure 14 shows simulation results for six selected mixtures to-

    gether with experimental data and the predictions from the Filippov relation114. The ARD

    from experimental data for each mixture are shown in Figure 7.

    ���� ���

    ������

    ���

    ���

    ��

    ���� ���

    ������

    ���

    ���

    ���� ���

    ������

    ���

    ���

    ��

    ���� ���

    ������

    ���

    ���

    ��

    �� ��������

    ��

    ��� �� �� ��� ��� ���

    ���� ���

    ������

    ���

    ���

    ��

    �����������

    ��

    ��� �� �� ��� ��� ���

    ���

    ���

    ���

    ���

    ���

    ���

    ���

    ���

    FIG. 14. Mole fraction dependence of the thermal conductivity of (a) benzene (1) + toluene, (b)

    toluene (1) + CCl4, (c) methanol (1) + acetone, (d) acetone (1) + CCl4, (e) methanol (1) + CCl4,

    (f) ethanol (1) + toluene at 298.15 K and 0.1 MPa. Present simulation results (•) are compared

    with the predictions from the Filippov relation114 (−−) and experimental data (+) as referred to

    in Table IV.

    As expected, for all studied mixtures, the mole fraction dependence of the thermal con-

    ductivity does not exhibit significant deviations from the simple interpolative relation by

    35

  • Mutual diffusion of binary liquid mixtures

    Filippov114. This behavior is usually correctly predicted by the present simulations, cf. Fig-

    ure 14. Nonetheless, in the case of methanolic mixtures with ethanol, acetone and benzene,

    the predicted thermal conductivity may show a strong scatter and overestimation in the

    methanol-rich composition range that is related to the high noise to signal ratio of the ther-

    mal conductivity autocorrelation function of both alcohols and to the offset for the pure

    fluids. Nonetheless, the overall ARD is only 11% for 14 mixtures for which experimental

    data are available.

    V. CONCLUSIONS

    This work is aimed at the understanding of transport property behavior in liquid mixtures,

    considering diffusion coefficients, shear viscosity and thermal conductivity. 20 binary liquid

    mixtures were studied on the basis of classical rigid force fields by molecular dynamics in

    a strictly predictive way. It was shown that these properties can be predicted from little

    sophisticated molecular models with a good accuracy.

    Values for the MS diffusion coefficient were sampled directly from equilibrium simulations

    with the Green-Kubo formalism, whereas the thermodynamic factor was obtained from

    selected experimental VLE data. The uncertainty of the thermodynamic factor was assessed

    employing three different GE models: Wilson, NRTL and UNIQUAC. In this way, it was

    estimated that the corresponding uncertainty does not to exceed 5% for mixtures with up to

    moderate deviations from the ideal behavior. For strongly non-ideal mixtures, it was found

    that the thermodynamic factor determined with different GE models may vary even by an

    order of magnitude, if the value of the thermodynamic factor approaches zero, i.e. near

    phase separation. The best results for the Fick diffusion coefficient were obtained with the

    thermodynamic factor based on the Wilson model with an overall ARD from experimental

    data of 16%. For the mixtures not deviating significantly from the ideal mixture behavior,

    the achieved overall ARD was usually below 10%, whereas for the most challenging strongly

    non-ideal mixtures, ARD between 15 and 20% were found. Further, an excellent qualitative

    agreement for the composition dependence of the Fick diffusion coefficient was obtained for

    the majority of the studied mixtures. Exceptions are mixtures of an alcohol and CCl4, where

    the Fick diffusion coefficient was strongly overestimated in the alcohol-rich region.

    In addition, nine different predictive relations for the mutual diffusion coefficients were

    36

  • Mutual diffusion of binary liquid mixtures

    tested against the available experimental data. The predictive equations based on local

    composition models, i.e. those by Li et al.15 and Zhou et al.16, as well as those including a

    correction of the thermodynamic factor, i.e. by D’Agostino et al.20 and Zhu et al.23, yield the

    best results with overall ARD from 12 to 15%. Further, four different predictive expressions

    for the self-diffusion coefficients in the mixture were assessed. The method by Liu et al.99

    was found to be the one with the lowest overall ARD from the available experimental values.

    For the highly non-ideal mixtures considered here, it was found that the equation by Liu

    et al.99 is not able to predict the strong decrease of the self-diffusion coefficient of the most

    polar substance in the mixture. The only predictive method, based on physical arguments,

    that was able to predict this extreme behavior of the self-diffusion coefficients is molecular

    simulation.

    The EMD method employed here yields the MS and self-diffusion coefficients as well as

    the shear viscosity and thermal conductivity directly from one simulation run. Therefore,

    the simulation results for shear viscosity and thermal conductivity were also compared with

    the available experimental values. An overall ARD of only 8% was found for the shear

    viscosity, whereas the ARD for the thermal conductivity is 11%.

    The microscopic structure of the studied mixtures was analyzed thoroughly. The RDF

    of mixtures that exhibit a nearly ideal behavior, albeit their components may be quite

    different, have similar distances and characteristics of the nearest-neighbors solvation shells.

    The similarity of their intramolecular interactions explains their almost ideal behavior. The

    RDF of the mixtures containing an alcohol and a less polar component suggest strong alcohol

    self-association, the presence of clusters and, in many cases, microheterogeneity. These are

    related to a sharp decrease of the self-diffusion coefficient of the alcohol in the mixture and

    to low values of the Fick diffusion coefficient.

    SUPPLEMENTARY MATERIALS

    A detailed description and parameters of the new molecular models for benzene, toluene

    and CCl4 are presented together with the calculated VLE and transport properties in com-

    parison to the corresponding reference equations of state or experimental data. The Simu-

    lation methodology is explained and details of the carried out simulations are reported. GE

    model parameters for the Wilson, NRTL and UNIQUAC models for all mixtures are given.

    37

  • Mutual diffusion of binary liquid mixtures

    The obtained diffusion coefficients at infinite dilution are listed for all mixtures compared

    to the Wilke-Chang equation an experimental data. Tabulated numerical simulation data

    and graphical representations for the density, Fick and MS diffusion coefficients, self-diffusion

    coefficients, shear viscosity and thermal conductivity for all mixtures are given. The RDF

    of all studied mixtures are


Recommended