Supplemental Derivation

download Supplemental Derivation

of 10

Transcript of Supplemental Derivation

  • 8/14/2019 Supplemental Derivation

    1/10

    Solvation Thermochemistry

    The treatment of solvent effects on the thermochemistry and kinetics of systems in

    condensed phases has been studied theoretically and empirically in a number of fields.

    However, often ambiguities present in the underlying assumptions and confusion with

    regard to standard states and reference states can cause difficulty when attempting to use

    the results correctly. An example of where confusion arrives could be in the typical

    concentration-based equilibrium constant CK , which is a physical property of the

    system. However, the free energy change of reaction that it is related to in

    phenomenological thermodynamics is actually the standard state change oRxnG , since

    the actual change in free energy at equilibrium must be zero. This means that there are an

    infinite number ofo

    RxnG values, since the standard state can be chosen arbitrarily. The

    same problem is seen for gas phase systems, but it is less problematic since the nearly

    universal standard state is an ideal gas at 1 atmosphere or bar of pressure. A simple

    example derivation for a reaction in a condensed phase is given in terms of the typical

    standard state thermodynamic quantities, as well as the pseudo-chemical potential

    * *i ior G of Ben-Naim. The concepts of Gibbs free energy and chemical potential will

    be used almost interchangeably here, with the knowledge that chemical potential is more

    specifically a partial-molar Gibbs free energy.

    The pseudo-chemical potential (PCP) as defined by Ben-Naim is the chemical potential

    (CP) of a solute confined to a fixed position, and is denote by the * superscript. The

    concentration dependence due to the translation is no longer present in a PCP; however,

    any concentration dependence of the chemical potential due to non-ideal effects would

    still be included. As will be seen below, one benefit of the PCP is that it will allow for a

    direct correspondence with experimental measurements, without the need for standard

    states. There are other subtleties to the PCP definition and use, and the reader is

    encouraged to seek the aforementioned references for more information. Equation

    relates the traditional CP to the PCP as defined by Ben-Naim. Equation is a useful

    relationship that will be used to convert between the standard state CP and the PCP,

  • 8/14/2019 Supplemental Derivation

    2/10

    where is the number density (or molar concentration) and 3 is the momentum partition

    function. It is derived by equating the traditional and PCP-based expressions for the

    chemical potential.

    # * # 3lni i i ikT

    #, #,

    *# 3 *# 3

    , ,ln ln ln lno o o oi ii i i i i i i io o

    i i

    kT kT or kT kT

    In all equations, an o superscript will be used to signify the standard state, a + will be

    used to signify an arbitrary reference state/behavior, a # will be used for the actual state

    of the mixture, an will be used for the dilute-limit behavior reference, and a p will

    be used for the pure i reference state. For example,*#

    i is the PCP at the actual state of

    the mixture, ando

    i is the CP at the standard state conditions. The symbolxi will be used

    for condensed phase mole fractions, and yi will used for gas phase mole fractions (orpi

    for the partial pressure). Unless otherwise noted, the condensed phase fugacity reference

    will be either dilute-limit behavior extrapolated to xi = 1, or Lewis-Randall behavior atxi

    = 1. The analysis is not requisite on this assumption, but simplifies the notation

    somewhat since we do not need a symbol for the mole fraction at which the reference was

    defined. The terms,

    i

    or,

    i

    represent the correction that must be made to the

    behavior at the state to achieve the behavior at state , and are known as fugacity

    and activity coefficients. All activity coefficients discussed here will be based on the

    mole fraction concentration scale, as is typical in chemical engineering thermodynamics

    using the fugacity formalism.

    Derivation of a Condensed-Phase Equilibrium ExpressionThe general method used to derive the equilibrium relationships shown in the paper will

    be demonstrated here through a simple solution-phase equilibrium constant example. It

    will be based on the use of the concept of fugacity as a means to relate the behavior of the

    chemical potential in the actual state to some arbitrary reference behavior. This reference

    behavior could theoretically be anything, but practically it is useful to define the

  • 8/14/2019 Supplemental Derivation

    3/10

    reference fugacity as a straight line as a function of mole fraction. Typical examples of

    this would be to assume the reference behavior as Henrys Law behavior or ideal solution

    behavior over the entire mole fraction range. These reference lines are illustrated by

    dashed lines in the figure below, which was adapted from the 3rd edition of

    Thermodynamics and Its Applications by Tester and Modell. Henrys Law is the

    extrapolation of the fugacity slope at low concentrations, and ideal solution behavior is a

    straight line connecting zero and the pure component fugacity pif at a mole fraction of

    one. First, the concentration-based equilibrium constant (KC) will be derived in terms of

    the standard state free energy change of reaction. Then, an expression forKC in terms of

    the PCP difference will be derived.

    xi

    #i

    f

    p

    if

    if

    #,

    i

    #,p

    i

    ##,

    #

    i

    i

    i i

    f

    x f

    +

    +=

    #i

    f

    0 1

    Standard State Methodology

    The derivation begins with a typical expression for the chemical potential in terms of

    fugacity in the condensed phase. A straightforward extension of this example to the gas

    phase or liquid-vapor equilibrium is possible.

    #

    #

    , ,

    ln

    o i

    i liq i liq o

    i liq

    fRT

    f

  • 8/14/2019 Supplemental Derivation

    4/10

    In this expression,#

    ,i liq is the chemical potential in the actual state, ,o

    i liq is the standard

    state chemical potential,#

    if is the fugacity in the actual state, and o

    if is the fugacity in

    the standard state. Since the goal is to derive an equilibrium relationship for the example

    reaction A B , one must realize that the actual state chemical potentials ofA and B

    must be equal at equilibrium. Writing equation for each species and setting the CPs

    equal allows one to arrive at equation , dropping the explicit condensed phase notation. A

    simple rearrangement of the terms allows one to extract the well-known activity-based

    equilibrium constant,KA, which is defined in terms of the standard state CP difference or

    the standard state Gibbs free energy change of reaction as shown in equation .

    # #

    ln ln

    o oA BA Bo o

    A B

    f fRT RT f f

    ##

    exp exp

    o o o oB A Rxn B A

    A o

    B A

    G f fK

    RT RT f f

    In order to simplify this relationship, one has to redefine the fugacity as a product of a

    simple reference fugacity and a term accounting for the non-linear behavior as a function

    of composition. The reference fugacity behavior will take the form of a line connecting

    the zero fugacity point atxi = 0 with some arbitrary point at xi = 1. As alluded to earlier,

    typical examples of this arbitrary point are the Henrys Law extrapolation or the pure

    component fugacity. The common name for the second term is the activity coefficient,

    which corrects the reference state fugacity at a given composition to the actual state

    fugacity at that composition. In general, the activity coefficient is very complicated and

    depends on the composition of the mixture, the reference behavior chosen, temperature,

    and pressure. A general expression for the fugacity of a species in solution is givenbelow in equation .

    # # #,

    ii i i

    i

    ff x

    x

  • 8/14/2019 Supplemental Derivation

    5/10

    Here,#

    ix is the mole fraction in the actual state,

    if is the reference fugacity defined at

    the mole fraction ix

    on the arbitrary reference line betweenxi = 0 andxi = 1, and#,

    i

    is

    the activity coefficient needed to convert from the reference fugacity to the actual state

    fugacity. Since if

    lies on the reference line between mole fractions of 0 and 1, it is

    customary and logical to take 1ix , which simplifies the expression and results in

    equation .

    # # #,i i i if x f

    Now, if

    is the reference state fugacity atxi = 1, which means that the reference fugacity

    at any mole fraction can be defined as#

    i ix f . Essentially, equation says that the fugacity

    in the actual state is equal to the reference fugacity at the same composition as the actual

    state multiplied by the activity coefficient that corrects for deviations from the reference

    state behavior. It is often helpful to see this graphically, and the reader is urged to

    supplement this text with graphical representations of the fugacity coefficient that can be

    found in most chemical engineering thermodynamics and physical chemistry textbooks.

    Now that the fugacity has been defined, it can be used in the previously derived

    equilibrium constant relationship. This definition replaces the actual state fugacity and

    the standard state fugacity, and the result is shown in equation . To be explicitly clear,

    reference state fugacities typically make an assumption about linear behavior across the

    entire composition range; however, actual state and standard state fugacities and chemical

    potentials should include all non-idealities.

    # #, ,

    , # #,exp

    o o o

    Rxn B B B A A AA o o

    B B B A A A

    G x f x f K

    RT x f x f

    Up to this point, very few assumptions have been made regarding the behavior or state of

    the system, and it is desired that the assumptions be kept to a minimum to allow the

    reader to see where all terms arise in the equilibrium expression. However, we will now

    make one simplifying assumption, which is that the reference behavior of a given species

  • 8/14/2019 Supplemental Derivation

    6/10

    is taken to be the same for the actual state and the standard state. This is a very logical

    assumption that was implicitly made when equation was written with if

    identical for

    the actual state and standard state. This means that ratio of if

    s in the equation will be

    equal to one, resulting in the simplified equation . However, let it be clear that A Bf f

    was notrequired to make the simplification, only that the standard state and actual state

    fugacities for a given species were based upon the same reference behavior, such as

    Henrys Law or ideal solution behavior. Different reference behaviors may be chosen for

    each molecule as desired, and the choice will not affect the resulting equilibrium

    constant, assuming any activity coefficient can be calculated to the same degree of

    accuracy.

    # #, ,

    , # #,exp

    o o o

    Rxn B B A AA o o

    B B A A

    G x xK

    RT x x

    Equation is a general expression for the equilibrium constant and shows several

    important aspects ofKA. First, it is dependent on the standard state, which should be

    obvious given that it is defined by the standard state free energy change, and there are

    standard state mole fractions in the definition as well. It can also be seen that the

    reference behavior chosen in defining the fugacity is completely arbitrary, but this

    arbitrariness is compensated for by the activity coefficient. Therefore, it does not affect

    the equilibrium constant. For example, if you define your reference far from the behavior

    of the actual system, then the actual state activity coefficient will be very large, and if you

    define it close to the actual state behavior the opposite will be true. This brings to light

    another important aspect, which is that the activity coefficients for the actual state and

    standard state are completely independent because the standard state is arbitrary. In fact,

    it is usually the case that the standard state and the actual state compositions are different

    (otherwise 0o

    RxnG ), which necessarily makes the activity coefficients different. It is

    often possible to define the reference behavior beneficially, such that the activity

    coefficient of the actual state or standard state is close to one. In many cases and

    especially with computational chemistry calculations, the standard state is defined by the

  • 8/14/2019 Supplemental Derivation

    7/10

    calculation or experiment that one can perform, and the reference state follows logically

    from that. In a computational chemistry calculation, the state of the molecule is usually

    an isolated molecule in a continuum, which essentially mimics Henrys Law behavior

    where solute-solute interactions are negligible. If the standard state concentration is also

    taken to be low, then the Henrys Law reference would be a logical choice because it

    would allow one to assume that, 1oi

    , simplifying the expression for the equilibrium

    constant.

    If one is attempting to construct detailed kinetics models, then it is necessary to have the

    concentration-based equilibrium constant (KC) to calculate reverse rate constants. The

    activity-based equilibrium constant can be related to KC with relatively little effort. In

    order to do this, the solution phase mole fractions must be converted into concentrations,

    by defining the mole fraction as the concentration of i iC divided by the total

    concentration TC . Applying this definition, one can arrive at equation .

    # #, # ,

    # , # #,

    #, # ,

    # , #,

    expo o o o

    Rxn B B T T A AA o o o

    T B B A A T

    o o o

    B T T A A

    C o o oT B B A T

    G C C C C K

    RT C C C C

    C C C

    K C C C

    One can quickly notice that the total solution concentration is not necessarily the same at

    the actual state composition and the standard state composition. In this equimolar

    reaction example, the total concentration factors would cancel as long as the standard

    state was chosen to be a mixture of A and B in a solvent witho

    AC and

    o

    BC being the

    concentrations in solution. Note that one does not necessarily need the standard state

    concentrations of A and B to be equivalent, though they usually are taken to be so.

    However, for a non-equimolar reaction, one would always be left with a term of the form

    #n

    o

    T TC C

    . Therefore, a more useful condition would be when the total concentration

    under the standard state conditions and actual state conditions are equal. If this is the

    case, then all total concentration terms will cancel out and equation can be further

  • 8/14/2019 Supplemental Derivation

    8/10

    simplified. Generally, this condition is usually not satisfied rigorously, but at relatively

    low concentrations in aqueous solution, the error introduced by making this assumption

    will be small. Using this assumption, one can arrive at equation , which relates the

    concentration-based equilibrium constant to the standard state free energy change.

    , #,

    #, ,exp

    o o o

    Rxn B B AC o o

    A B A

    G CK

    RT C

    Equation can then be evaluated if one is able to estimate a standard state free energy

    change, the activity coefficient of A and B in the actual system state, and the activity

    coefficient of A and B in the standard state mixture. These are not necessarily easy tasks

    to accomplish, and often further assumptions can be made based upon the system

    conditions and the choice of standard and reference states to simplify the process.

    It is useful to say a bit about what equations and actually mean from a physical

    perspective. One starts with the free energy change under the standard state conditions.

    The standard state activity coefficients in the equation serve to remove the non-idealities

    at the standard state, essentially leaving a free energy change under the reference state

    behavior at the standard state concentration. The ratio of standard state concentrations

    addresses the simple concentration dependence of the free energy. The actual stateactivity coefficients then take the free energy change under the reference state behavior to

    the behavior at the actual system conditions. It is very much like a thermochemical cycle

    with three steps: (i) real behavior to reference behavior at the standard state composition,

    (ii) correction for the composition difference between the standard state and actual state,

    and (iii) reference behavior to real behavior at the actual state composition. It is not

    exactly a cycle as written, but looks very much like one if the equation is solved for

    o

    RxnG

    . This brings up an interesting point about what one needs to do when defining thestandard state free energy change. Although the work here assumes that

    o

    RxnG includes

    all non-idealities, making it a true free energy change under the standard state

    conditions; one should realize that this is not a necessity. As in the computational

    chemistry calculations, it may be much easier to obtain energies under very dilute

  • 8/14/2019 Supplemental Derivation

    9/10

    conditions, ignoring non-ideal effects. However, it is also possible to evaluate the

    partition functions at a concentration of far from the dilute limit. If this is done, one is

    left with a free energy estimate that ignores non-idealities, but is not evaluated under

    conditions where non-idealities can be ignored. The result is an estimate that lies on the

    reference behavior line, which would be the Henrys Law extrapolation in this dilute-

    referenced example. If this type of standard state energy data is used, one must simply

    throw out the standard state activity coefficients in equations and because you already

    lie on the reference behavior line. Then steps (ii) and (iii) of the cycle are the same.

    The point to be taken away here is that the standard state is arbitrary and can be defined

    in a variety of ways; however, one needs to be sure of exactly what assumptions are

    implicitly accounted for in the equations and data.

    Pseudo-Chemical Potential Methodology

    In this part, extensive use will be made of the previously-derived standard state

    expressions. First, if one recalls equation for the condensed phase and writes the

    standard state CP difference for the reaction, equation is the result. Equation was

    derived by equating and , using equation to define the fugacity, and assuming that the

    total concentration of the actual state and the standard state are equivalent # 0T TC C .

    This assumption could be relaxed, but would result in more complicated notation andmore intimidating equations.

    3 #, ,

    *# *#

    3 , #,

    3 #, ,

    3 , #,

    ln ln

    ln ln

    o oo o B B B A

    B A B A o o

    A A B A

    o oo B B B A

    Rxn Rxn o o

    A A B A

    kT kT or

    G G kT kT

    Once equation has been established, the expressions in terms of the PCP difference

    RxnG come rather naturally from the previous equations in terms of oRxnG . The

    relationship between KA and the PCP difference is shown in equation and is found by

    simply inserting equation into equation . The expression forKC can be found by

    combining equations and , realizing that i and Ci are the same, and canceling terms.

  • 8/14/2019 Supplemental Derivation

    10/10

    The simple result is shown in equation . The momentum partition function terms still

    remain because the PCP is the CP without the momentum partition function. In a

    unimolecular reaction or when examining phase equilibria for a single molecule, the

    lambda ratio will be unity because of conservation of mass and the classical definition

    used by Ben-Naim that is only a function of mass and temperature. One can also see that

    equation can be derived directly by equating for A and B at equilibrium and solving for

    the concentration ratio,KC.

    3 #, ,

    3 , #,exp

    o o

    Rxn A A B AA o o

    B B B A

    GK

    RT

    3

    3expRxn A

    C

    B

    GKRT

    Although this example was for the specific case of unimolecular reaction equilibrium in a

    condensed phase, it should be straightforward to extend this to non-equimolar reactions,

    gaseous processes, and liquid-vapor equilibria. Many other equilibrium relationships can

    be derived in a similar manner. In this derivation, an attempt was made to keep

    assumptions to a minimum and to pedagogically explain them when implemented. It is

    hoped that it was completed with enough clarity and rigor to allow others to evaluate

    what assumptions hold for their application and to modify the given expressions as

    necessary.