ballard 1

8
The Next Generation of Hydrate Prediction: An Overview A. L. Ballard and E. D. Sloan, Jr. * Colorado School of Mines, Golden, CO 80401 Abstract The van der Waals and Platteeuw hydrate equation of state, coupled with the classical thermodynamic equation for hydrates, has been used in the prediction of hydrate formation for over forty years. The standard state used in these equations is a hypothetical empty hydrate lattice. In part I of this series, we proposed an alternative derivation of these equations using a different standard state. The new hydrate equations were shown to be simpler to use. In part II of this series, we proposed an aqueous phase model tailored specifically for the presence of hydrate inhibitors such as salts and methanol in the aqueous phase. Part III provides a prescription for the incorporation of the new hydrate and aqueous phase models into a multi-phase Gibbs energy minimization program (CSMGem). Part IV compares predictions from the CSMGem program with four other commercially available programs. In this paper, we give a brief overview of each of the papers in the series, discussing the non- ideal solid solution hydrate model, incorporation of all fugacity models using the Gibbs energy minimization technique, and overall results of the CSMGem program. * Corresponding Author. E-mail address: [email protected] Introduction Clathrate hydrates are non-stoichiometric crystalline compounds that consist of a hydrogen bonded network of water molecules and enclathrated molecules. Davy (1811) first observed clathrate hydrates in the chlorine+water system. It wasnt until 1934, however, that clathrate hydrates were extensively studied. Hammerschmidt (1934) found that natural gas transport lines could be blocked by the formation of clathrate hydrates. This raised substantial attention in the oil and gas industry, prompting more hydrate research. Natural gas hydrates are formed when natural gas is brought into contact with water, generally at low temperatures and high pressures. The guest molecules most common in natural gas systems are hydrocarbons ranging from methane to i-pentane. These gases make up greater than 98 mole percent of a typical natural gas in United States pipelines and comprise the majority of the experimental work in the last 70 years. For all topics of hydrate research, a thermodynamic model for hydrate stability is needed. Hydrates are non-stoichiometric in that the composition of the hydrate is not constant. The filling of the cavities in the hydrate depends on the composition of the gas mixture as well as the temperature and pressure of the system, posing a modeling challenge. In the early 1950s, von Stackelberg and co- workers determined, via X-ray diffraction, that there were two types of hydrates: sI and sII. With knowledge of the crystal structure of hydrates, a statistical thermodynamic model of hydrate phase equilibria was conceived by van der Waals and Platteeuw (1959). In their work, an expression for the chemical potential of water in any hydrate structure was developed using an approach analogous to Langmuir adsorption. The van der Waals and Platteeuw model was difficult to use compared to earlier methods. While the gas gravity (Katz 1945) and K-value (Wilcox et al. 1941) methods could be performed with a hand calculation, the method of van der Waals and Platteeuw could not. This posed problems for the van der Waals and Platteeuw models applicability in industrial settings. With the advent of computers, however, this method became the method of choice. In 1964, Saito et al. (1964) developed expressions for calculating hydrate phase equilibria using the chemical potential of water in hydrates developed by van der Waals and Platteeuw. Their approach allowed the hydrate formation pressures and temperatures to be determined by equating the chemical potential of water in the hydrate with that in the aqueous or ice phase. This method was not widely used until Parrish and Prausnitz (1972) developed an iterative scheme, using the equations developed by van der Waals and Platteeuw and Saito et al., for use on a computer. The Parrish and Prausnitz scheme does not explicitly incorporate the hydrate phase. In their method, the thermodynamic equilibrium of the fluid phases is determined and then compared to the hydrate phases. The temperature or pressure is then found such that the chemical potential of water is equal in all phases (including the hydrate phase). The Parrish and Prausnitz work was a landmark in that it allowed, for one of the first times, the application of statistical thermodynamics in an industrial setting. The van der Waals and Platteeuw model,

Upload: ediabc

Post on 16-Sep-2015

224 views

Category:

Documents


2 download

DESCRIPTION

ggghhhhh

TRANSCRIPT

  • The Next Generation of Hydrate Prediction: An Overview

    A. L. Ballard and E. D. Sloan, Jr.* Colorado School of Mines, Golden, CO 80401

    Abstract

    The van der Waals and Platteeuw hydrate equation of state, coupled with the classical thermodynamic equation for hydrates, has been used in the prediction of hydrate formation for over forty years. The standard state used in these equations is a hypothetical empty hydrate lattice. In part I of this series, we proposed an alternative derivation of these equations using a different standard state. The new hydrate equations were shown to be simpler to use. In part II of this series, we proposed an aqueous phase model tailored specifically for the presence of hydrate inhibitors such as salts and methanol in the aqueous phase. Part III provides a prescription for the incorporation of the new hydrate and aqueous phase models into a multi-phase Gibbs energy minimization program (CSMGem). Part IV compares predictions from the CSMGem program with four other commercially available programs.

    In this paper, we give a brief overview of each of the papers in the series, discussing the non-ideal solid solution hydrate model, incorporation of all fugacity models using the Gibbs energy minimization technique, and overall results of the CSMGem program.

    * Corresponding Author. E-mail address: [email protected]

    Introduction Clathrate hydrates are non-stoichiometric

    crystalline compounds that consist of a hydrogen bonded network of water molecules and enclathrated molecules. Davy (1811) first observed clathrate hydrates in the chlorine+water system. It wasnt until 1934, however, that clathrate hydrates were extensively studied. Hammerschmidt (1934) found that natural gas transport lines could be blocked by the formation of clathrate hydrates. This raised substantial attention in the oil and gas industry, prompting more hydrate research.

    Natural gas hydrates are formed when natural gas is brought into contact with water, generally at low temperatures and high pressures. The guest molecules most common in natural gas systems are hydrocarbons ranging from methane to i-pentane. These gases make up greater than 98 mole percent of a typical natural gas in United States pipelines and comprise the majority of the experimental work in the last 70 years.

    For all topics of hydrate research, a thermodynamic model for hydrate stability is needed. Hydrates are non-stoichiometric in that the composition of the hydrate is not constant. The filling of the cavities in the hydrate depends on the composition of the gas mixture as well as the temperature and pressure of the system, posing a modeling challenge. In the early 1950s, von Stackelberg and co-workers determined, via X-ray diffraction, that there were two types of hydrates: sI and sII. With knowledge of the crystal structure of hydrates, a statistical thermodynamic model of hydrate phase equilibria was conceived by van der Waals and Platteeuw (1959). In

    their work, an expression for the chemical potential of water in any hydrate structure was developed using an approach analogous to Langmuir adsorption.

    The van der Waals and Platteeuw model was difficult to use compared to earlier methods. While the gas gravity (Katz 1945) and K-value (Wilcox et al. 1941) methods could be performed with a hand calculation, the method of van der Waals and Platteeuw could not. This posed problems for the van der Waals and Platteeuw models applicability in industrial settings. With the advent of computers, however, this method became the method of choice.

    In 1964, Saito et al. (1964) developed expressions for calculating hydrate phase equilibria using the chemical potential of water in hydrates developed by van der Waals and Platteeuw. Their approach allowed the hydrate formation pressures and temperatures to be determined by equating the chemical potential of water in the hydrate with that in the aqueous or ice phase. This method was not widely used until Parrish and Prausnitz (1972) developed an iterative scheme, using the equations developed by van der Waals and Platteeuw and Saito et al., for use on a computer. The Parrish and Prausnitz scheme does not explicitly incorporate the hydrate phase. In their method, the thermodynamic equilibrium of the fluid phases is determined and then compared to the hydrate phases. The temperature or pressure is then found such that the chemical potential of water is equal in all phases (including the hydrate phase).

    The Parrish and Prausnitz work was a landmark in that it allowed, for one of the first times, the application of statistical thermodynamics in an industrial setting. The van der Waals and Platteeuw model,

  • coupled with the Parrish and Prausnitz algorithm to solve for hydrate prediction have remained relatively unchanged in the last 30 years. Motivation for a New Approach to Hydrate

    Modeling For industrial use, the method of Parrish and

    Prausnitz for determining the formation pressures and temperature of hydrates has its purpose. The main use of it is to prevent hydrate formation in pipelines. For example, if the temperature and pressure profiles are known for a given pipeline, the regions of possible hydrate formation can be determined using the hydrate model. With that determined, chemicals can be injected into the pipelines to prevent hydrate formation. For this reason, there has been no need in developing another method for hydrate phase equilibria. However, current hydrate related problems dictate that another approach to hydrate modeling is needed. Hydrates at High Pressure. As the demand for oil and natural gas becomes greater, there is a need to go to deeper depths of water to obtain it. Therefore, the pressure and temperature ranges of thermodynamic models that are the basis of flow assurance strategy need to be extended. Predictions from the current hydrate model are relatively good at moderate pressures and temperatures. However, at high pressures (P>200 bar) and for natural gas hydrates the model predictions deviate from experimental data. This is indication that the properties of the hydrate are not properly treated in the current model. Hydrates in Natural Environments. Significant amounts of hydrate are found in deep ocean sediment. Along a vertical plane of ocean floor, acoustic imaging is used to determine the depths at which these hydrates are present. It has been proven that the bottom simulating reflector (BSR), represents the bottom of a hydrate accumulation. That is, hydrates are present at depths less than the BSR and not present below the BSR. The pressure and temperature conditions at the BSR are such that hydrates are not stable below it (i.e. the temperature is too high). Therefore, the BSR represents the pressure and temperature at which hydrates first form.

    While it is important to know the properties of the hydrate at the BSR depth, the majority of the hydrate accumulation is not at these temperature and pressure conditions. Therefore, the properties need to be known at conditions higher than at the formation conditions. Natural hydrates have been around for many millennia and have most likely reached an equilibrium state with their surrounding phases (i.e. aqueous, vapor). The method of Parrish and Prausnitz can determine only the properties of the hydrate at the formation conditions (at the BSR). However, the properties need to be known at pressures well above this condition.

    A New Approach to Hydrate Modeling In order to give a complete description of the

    hydrate phase at any temperature and pressure, the hydrate phase must be explicitly part of a flash algorithm. The Parrish and Prausnitz method does not allow this. Therefore, another method is needed which essentially treats the hydrate phase as just another phase in a flash calculation.

    Multi-phase flashes are commonly performed for V-L-L equilibria and have been around for many years (Baker et al. 1982). These types of flashes are based on the criterion of minimizing the Gibbs energy of a closed system. A multi-phase flash calculation is just an extension of the common two-phase flash, if care is taken. In this sense, it should be a simple task to add the hydrate phase into a flash calculation. In Bishnois lab, Gupta (1990) was the first to accomplish this feat. In his breakthrough work, Gupta developed the equations and solution procedure necessary to calculate hydrate phase equilibria at conditions other than hydrate formation. However, the method of Gupta has been overlooked in the last 10 years.

    The work of Gupta was certainly ground-breaking in that the hydrate phase was, for the first time, incorporated in flash calculations. However, Gupta did not report some key aspects of his method. Hydrates are highly non-ideal from a thermodynamic perspective in that the composition of water in the hydrate is bounded and the fugacity of water in the hydrate is a strong function of this bounded composition. These two aspects of hydrates require a special treatment in a flash calculation, which are not reported by Gupta. In this work, we give the complete procedure needed for incorporating hydrates in a multi-phase flash program (CSMGem). We have also developed a new initial guess procedure based on ideal K-values.

    The CSMGem program is the culmination of several separate components of this laboratory. Figure 1 shows the schematic used for the development of the CSMGem program. Each box and circle represents

    hydrate phasemeasurements

    next generation hydrate program

    (CSMGem)

    RamanNMRXRD

    Gibbs energyminimization

    new hydratefugacitymodel

    aqueousfugacitymodel

    hydrocarbonfugacitymodel

    Windowspackaging

    pure phasefugacitymodels

    Figure 1. Schematic of the development procedure for the

    CSMGem program a different aspect of the development while the lines attaching them represent the process of incorporating

  • them to each other. The diagram is read in a left-to-right, top-to-bottom procedure.

    The proposed modifications to the hydrate fugacity model stem from direct measurements of the hydrate phase using X-ray diffraction (XRD), Raman spectroscopy, and nuclear magnetic resonance (NMR) spectroscopy. The proposed hydrate fugacity model incorporates a substantial amount of hydrate phase measurement data via optimization of the parameters. The hydrate, aqueous, hydrocarbon, and pure phase fugacity models are all linked using a Gibbs energy minimization (GEM) flash algorithm. The GEM flash algorithm is coupled with a Windows interface, using Visual Basic, to make the next generation hydrate program (CSMGem).

    Gibbs Energy Minimization

    To calculate thermodynamic equilibrium for a

    closed system, three fundamental conditions must be met: 1) temperature equilibrium of all phases, 2) pressure equilibrium of all phases, and 3) equality of fugacity in each phase present,

    all resulting from the Gibbs energy being at a minimum (Gibbs 1875). These conditions are commonly used in developing procedures for solving for thermodynamic equilibrium. For a system of known phases, meeting the first three conditions will ensure that the Gibbs energy is at a minimum. The most common implementation of these conditions is for the two-phase system, vapor and liquid hydrocarbon, known as the VLE Flash.

    The requirement that the Gibbs energy of the system must be at a minimum, at a given temperature and pressure, is a statement of the second law of thermodynamics. Meeting conditions 1 through 3 is necessary for thermodynamic equilibrium but is not sufficient for the minimization of the Gibbs energy. Baker et al. (1982) discuss this further in their development of a solution procedure for multi-phase equilibrium calculations. For simple systems in which the phases present at equilibrium are known (i.e. vapor and liquid hydrocarbon), however, conditions 1 through 3 are commonly used with no problem. When solving for thermodynamic equilibrium in a more complex system in which several phases could form, a fourth criterion, the minimum of Gibbs energy, may be used. That is, if the amount of phases present at the solution is not known a priori, the Gibbs energy must be used to determine what phases are present.

    Implementation of conditions 1 to 3 is done following a similar procedure as that of a two-phase system (Rachford and Rice 1952). For a system with C components and possible phases, satisfying a simple mass balance for each component in each phase results in the following objective function:

    1

    1

    10

    1 1

    ikiC

    irk

    i ijj

    j irj r

    xzx

    Exx

    =

    =

    = = +

    k = 1,, 1

    where zi is the mole fraction of component i in the feed, j is the molar phase fraction of phase j, and xik is the mole fraction of component i in phase k. The subscript r refers to a reference phase, which is present at thermodynamic equilibrium. Implementation of conditions 1 to 3 results in the following expression for the mole fraction ratio: ik ir

    ikir ik

    x Kx

    = =

    i = 1,,C k = 1,,. 2

    where Kik is the distribution coefficient. Note that the definition of the distribution coefficient (Equation 2) is valid only for phases present at thermodynamic equilibrium. Therefore, this approach can only be used if all phases (k = 1,,) are present. This is quite a drawback in that, for natural gases and water, the phases present at equilibrium are usually not known.

    Gupta (1990) showed that, by defining the mole fraction ratio in terms of the distribution coefficient as k kik ir

    ikir ik

    x e K ex

    = =

    i = 1,,C k = 1,, 3

    where k is defined as

    exp ln ikkir

    ff

    = k = 1,,, 4

    is equivalent to minimizing the Gibbs energy of the system conditional to 0k kk

    k k

    S

    = =

    + k = 1,,. 5

    That is, if k > 0 then phase k is present and k = 0 and likewise, if k = 0 then phase k is not present and k 0. Note that Equation 6 is always satisfied for the reference phase by definition of k.

    Replacing the mole fraction ratio in Equation 1 with Equation 3, we obtain

    ( )( )1

    1

    10

    1 1

    k

    j

    Ci ik

    ki

    j ijjj r

    z K eE

    K e

    =

    =

    = =

    +

    k = 1,, 6

    which is valid for all phases (present or not) at a solution. Equation 6 is used to determine the set of phase amounts, , and stability variables, , that satisfy thermodynamic equilibrium.

    The solution procedure is split into two parts: 1. minimizing Gibbs energy by updating phase amounts and stability variables at a given set of K-values and 2. updating K-values at a given set of phase amounts and stability variables. This approach is typical in phase equilibria problems such as this. The Newton procedure is used to minimize the Gibbs energy of the system at a given set of K-values (via Equation 6). This approach is convergent as long as the fugacity coefficients are not

  • strong functions of composition. This is certainly the case for the fluid phases such as vapor and liquid hydrocarbon but is not necessarily true for the aqueous and hydrate phases. With proper care, which will be discussed later, this approach is convergent for all systems.

    The Newton method gives a correction to the phase amounts and stability variables based on the gradient of the Gibbs energy at the given set of K-values. Due to the highly non-ideal behavior of the hydrate phases, all corrections are scaled such that no phase amount or stability variable is changed more than 25% of its original value. That is, the entire correction vector is scaled so as to keep the proper direction of convergence.

    In the development of Equation 6, the composition of each phase is given as

    ( )1

    1 1

    k

    j

    i ikik

    j ijjj r

    z K exK e

    =

    =

    + i = 1,,C k = 1,,. 7

    Equation 7 is used to update the composition of each phase. With the new composition, the fugacity coefficients are calculated to get the new set of K-values. As is the case for minimizing the Gibbs energy, this approach is convergent as long as the fugacity coefficients are not strong functions of composition. Successive substitution of the composition gives linear convergence whereas Newtons method gives quadratic convergence.

    Due to the highly non-ideal behavior of the hydrate phases, Equation 7 is not used to update the composition. The hydrate equation of state gives an explicit formula for calculating the successive composition. This formula is given later in this paper. However, the compositions of all non-hydrate phases are determined via Equation 7. All composition corrections of a given phase are scaled such that no composition in that phase is changed by more than 50% of its original value.

    The solution procedures discussed above are implemented into the algorithm shown in Figure 2.

    estimate xik, Kikk, k

    input feedconditions

    update k and k

    update xik and Kik

    Ek = 0 ?

    xik = 0 ?

    stop

    no

    no

    yes

    yes

    estimate xik, Kikk, k

    input feedconditions

    update k and k

    update xik and Kik

    Ek = 0 ?

    xik = 0 ?

    stop

    no

    no

    yes

    yes

    Figure 2. Algorithm to solve for thermodynamic

    equilibrium

    As can be seen in Figure 2, thermodynamic equilibrium is achieved when the Gibbs energy is at a minimum (Ek = 0) and the difference in the updated mole fractions and previous mole fractions (xik) for each phase is zero. On a computer, it is difficult for a variable to become exactly zero. Therefore, in this work, convergence is met when Ek and xik are less than 1E-6. One of the crucial steps in obtaining a solution is creating a good initial estimate for the unknown variables.

    Several authors have determined ideal K-values for component distribution between V-Lhc (DePriester 1953, Hadden and Grayson 1961) and V-Hydrate (Wilcox et al. 1941) phases. However, Ballard (2002) has developed composition independent sets of K-values to determine the component distribution between all possible phases: vapor, liquid hydrocarbon, aqueous, sI hydrate, sII hydrate, sH hydrate, ice, solid NaCl, solid KCl, and solid CaCl2.

    The Ideal Solid Solution Hydrate Fugacity Model

    In 1959, van der Waals and Platteeuw (1959)

    used statistical thermodynamics to derive several thermodynamic properties of hydrates, most importantly the chemical potential. Their approach paralleled the work of Langmuir for adsorption. That is, van der Waals and Platteeuw statistically treated hydrate cages as adsorption sites in which species become adsorbed or encaged. This section provides a brief outline of the statistical model derived by van der Waals and Platteeuw. This is the hydrate fugacity model that has been used for the last 40 years. The next section will discuss the derivation and modifications made to the van der Waals and Platteeuw model for this work.

    The chemical potential of water in the hydrate is given by the following equation

    ln 1wwH m imm i

    gRT RT

    = + 8 where, gw is the Gibbs energy of water in the standard (empty) hydrate lattice at a given volume and the system temperature and pressure, m is the number of cages of type m divided by the number of water molecules in the hydrate lattice, and im is the fractional occupancy of guest i in hydrate cage m. Notice that the fractional occupancy of a given hydrate cage will always sum to a value less than one. Therefore, the natural log of the term in parenthisis will always be a number less than zero. In other words, the summation term in Equation 8 lowers the total energy of the hydrate as a result of the insertion of the guest molecules into the empty hydrate lattice. A complete definition of the fractional occupany is given by Ballard (2002). In order to perform the Gibbs energy minimization, the fugacity of each component in each phase is needed to perform thermodynamic equilibrium calculations. Therefore, we still need to determine the fugacity of each component in the hydrate. The standard approach to this was to relate the chemical potential of

  • water in the hydrate to that in the aqueous or ice phase (Parrish and Prausnitz 1972). This was done by considering the energy change associated with forming a hydrate and an aqueous or ice phase, both from the empty hydrate lattice. That is,

    ln 1wH wwH m imm i

    gRT RT

    = = 9 and

    0

    0 0

    / /

    /20

    ln

    wAq Ice wAq Ice w

    T Pw w w

    wAq IceT P

    gRT RT RT

    g h vdT dP aRT RT RT

    =

    = + +

    10

    where gw, hw, and vw are the molar Gibbs energy, enthalpy, and volume of pure liquid water or ice, respectively, and the term refers to the difference of that property between the pure liquid water or ice and empty hydrate phases. Note that Equation 10, developed by Saito et al. (1964), and simplified by Holder et al. (1980) and Menten et al. (1981), utilizes the classical expression for the chemical potential of water in the aqueous or ice phase and empty hydrate.

    This approach takes advantage of the fact that, at equilibrium, the chemical potential of water is equal in all phases. Using this as a basis, it can easily be shown that Equation 9 is equal to Equation 10 at equilibrium and that the fugacity of water in the hydrate is

    // exp

    wH wAq IcewH wAq Icef f RT

    =

    . 11

    Since the fugacity of water in the aqueous or ice phase must be known in Equation 11, an aqueous or ice phase must be present to correctly calculate the fugacity of water in the hydrate. This is a major drawback of this approach in that an aqueous or ice phase is not present in all hydrate scenarios. The next section will discuss a new derivation of the fugacity of water in the hydrate that eliminates this constraint. The fugacity of hydrate guests is assumed to be equal to that in the reference phase during the Gibbs energy minimization procedure, which will be true at equilibrium. This approach causes the stability of hydrates to be solely a function of the energy or fugacity of water in the hydrate. For components that are not present in the hydrate, the fugacity is assumed to be infinite. As discussed previously, Equation 7 is not used to update the composition of hydrates in the Gibbs energy minimization procedure. Depending on the type of hydrate that forms (sI, sII, sH), the composition of water is constrained due to the finite number of cages. The lower limit on the composition of water in the hydrates, assuming that all cages are filled, is 0.852, 0.850, and 0.850 for sI, sII, and sH hydrates, respectively. The upper limit is not as defined, as it depends on the temperature and pressure of the system as well as the hydrate guests. However, the composition of water in the hydrate must be less than 1.0 because, if it were equal to

    1.0, the phase would not be a hydrate, rather some new type of ice phase. The expression for the composition of species in the hydrate follows from a simple mass balance is

    1

    m imm

    iHm jm

    m j

    x

    =

    +

    12

    where m is the number of hydrate cages of type m per water molecule in the hydrate. This equation is used in place of Equation 7 while determining the solution. Improvements to the of van der Waals and

    Platteeuw Model: Hydrates as the Non-Ideal Solid Solution

    As the demand for oil and natural gas becomes

    greater, there is a need to go to deeper depths of water to obtain them. Therefore, the pressure and temperature ranges of thermodynamic models that are the basis of flow assurance strategy need to be extended.

    The current hydrate fugacity model described in the previous section applies statistical thermodynamics and classical thermodynamics. Predictions from this model are relatively good at moderate pressures and temperatures. However, at high pressures (P>200 bar) and for natural gas hydrates the model predictions deviate from experimental data. This suggests that the hydrate standard state (empty lattice) properties are not well defined. As discussed in the previous section, hydrate predictions must be made with an aqueous phase present. Ballard and Sloan (2002) proposed two modifications to the existing theory for hydrate fugacity: 1) an activity coefficient based on exact knowledge of the hydrate volume and 2) the first description of the hydrate standard state itself.

    Activity Coefficient for Water in the Hydrate. In the development of the statistical thermodynamic hydrate model (Equation 8), the free energy of water in the standard hydrate (empty hydrate lattice), gw, is assumed to be known at a given T and v. Since the model was developed at constant volume, the assumption requires that the volume of the empty hydrate lattice, v, be equal to the volume of the equilibrium hydrate, vH, so that the only energy change is due to occupation of the hydrate cavities, as shown in Figure 3.

    . . v (T,P) vH (T,P,x)

    . . . . . . . . . .

    . .

    . . 1

    Figure 3. Current model not allowing for distortion of

    hydrate due to guests (v = vH)

  • Process (1) in Figure 3 is given by the summation term in Equation 8.

    Traditionally, the chemical potential of the standard hydrate is assumed to be at a given volume, independent of the hydrate guests. If the standard hydrate volume is not the volume of the equilibrium hydrate, there should be an energy change proportional to the difference in volume (vH = vH-v ). Note that, in the development of Equation 8, vH is assumed to be equal to zero (i.e. all hydrates of a given structure are at the same volume). However, Ballard and Sloan (2002) accounted for the distortion via an activity coefficient.

    Equation 8 is considered to be an ideal solid solution model. Ballard and Sloan defined the chemical potential of water in hydrates as

    ln 1 lnwH w m im wHm i

    g RT RT

    = + + 13 where the activity coefficient of water in the hydrate, wH, accounts for non-idealities due to the inclusion of the hydrate guests. Note that Equation 13 can be visualized in a similar manner as in Figure 3. Figure 4 illustrates the processes needed to determine the chemical potential of water in the hydrate (as given by Equation 13).

    vH (T,P,x) v (T,P) v (T,P)

    1 2

    . . . . . . . . . . . . . . . .

    . . . . . . . .

    . . . . . . . .

    Figure 4. Corrected model allowing for distortion of hydrate due to guests (v vH)

    Process (1) in Figure 4 is done at constant volume and therefore, the van der Waals and Platteeuw statistical model can be used. Process (2) in Figure 4, the volume change of the hydrate from its standard state volume, is done at constant composition and is described by the activity coefficient in Equation 13.

    Ballard (2002) noted that chemical potential is a state function and, therefore Equation 13 can be visualized in another manner (as shown in Figure 5). vH (T,P,x) vH (T,P) v (T,P)

    1 2

    . . . . . . . .

    . . . . . . . .

    Figure 5. Alternative expression for corrected model allowing for distortion of hydrate (v vH)

    Equation 14 is the analogous expression for the processes shown in Figure 5.

    ln 1wH w w m imm i

    g g RT

    = + + 14

    where gw is a perturbed Gibbs energy of the empty hydrate lattice, equal to the natural log of the activity coefficient in Equation 13, defined as

    00 0

    20

    lnT P

    w w Hw wH

    T P

    g h vg dT dPRT RT RT

    = = + . 15 Process (1) is given by Equation 15 and process (2) by the van der Waals and Platteeuw statistical model, since it is done at constant volume. Note that, since chemical potential is a state function, Figures 4 and 5 are equivalent processes.

    In his work, Ballard (2002) arbitrarily defined the perturbed Gibbs energy and enthalpy of formation be linear in vH;

    0 0w Hg a v = and 0 0w Hh b v = . Note

    that the subscript 0 refers to the volume difference at T0 and P0, the temperature and pressure at which the formation properties are known. A more complete discussion of the non-ideal solid solution hydrate model is given in Ballard (2002). New Development of Fugacity. In order to solve for thermodynamic equilibrium, the fugacity of water in the hydrate must be known. Recall that the current approach to this is to relate the chemical potential of water in the hydrate to that in the aqueous phase. This approach can be burdensome; if an aqueous phase is not present, comparing chemical potentials will not be consistent with thermodynamic equilibrium. Ballard and Sloan (2002) suggested that the fugacity of water in the hydrate be defined as

    exp wH wowH wogf f

    RT

    = 16

    where fwo is 1 bar, gwo is the Gibbs energy of pure water in the ideal gas state at 1 bar, and wH is given by Equation 13. Note that the fugacity of water in the hydrate, as determined by Equation 16, does not require that an aqueous phase be present. This approach is much simpler to use and understand since the properties of the empty hydrate lattice are no longer referenced to pure liquid water and allows the hydrate phase to be calculated in standard thermodynamic sense. That is, calculation of the hydrate phase fugacity is no different than any other solid phase, with the exception of the activity, due to the fact that hydrates are not pure. Results

    A comparison of predictions from CSMGem,

    the program developed in this work, and four other commercial hydrate prediction programs is given for all hydrate data reported in literature. The four commercial programs compared in this work are:

    CSMHYD - CSM (1998 version), DBRHydrate - DBRobinson (version 5.0), Multiflash - Infochem (version 3.0), and PVTsim - Calsep A/S (version 11).

  • A total of 1685 hydrate formation data points were used in the comparison.

    The most common type of hydrate data taken is the formation temperatures and pressures. This type of data is most important to natural gas pipeline situations. We have compiled a list of all experimental data of this type and compared it with predictions from CSMGem, CSMHYD, DBRHydrate, Multiflash, and PVTsim. It should be noted that DBRHydrate has a prediction limit to pressures below 680 bar and so comparisons could not be made for all data points.

    The results are given as the average error either in temperature or pressure. The average error in temperature and pressure are calculated as:

    ( )predicted experimental

    # data pointsError

    T -TT =

    # data points

    17

    ( )predicted experimental# data points experimental

    Error

    P -PP

    P = 100%# data points

    , 18

    and are given in Kelvin and bar, respectively. Figures 6 and 7 show the comparisons for all five hydrate prediction programs.

    -0.2

    -0.1

    0

    0.1

    0.2

    Ave

    rage

    Err

    or in

    Tem

    pera

    ture

    (K)

    -4%

    -2%

    0%

    2%

    4%

    Ave

    rage

    Err

    or in

    Pre

    ssur

    e (%

    )

    CSMGemCSMHYDDBRHydrateMultiflashPVTsim

    6%

    T Error P Error

    0.24

    Figure 6. Hydrate formation T and P errors (relative) for

    all hydrates (1685 points)

    0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    Ave

    rage

    Abs

    olut

    e Er

    ror i

    n Te

    mpe

    ratu

    re (K

    )

    0%

    2%

    4%

    6%

    8%

    10%

    12%

    Ave

    rage

    Abs

    olut

    e Er

    ror i

    n Pr

    essu

    re (%

    )

    CSMGemCSMHYDDBRHydrateMultiflashPVTsim

    T Error P Error

    Figure 7. Hydrate formation T and P errors (absolute) for

    all hydrates (1685 points)

    Note that Figure 6 is the relative error (defined by Equations 17 and 18) and Figure 7 is the absolute error (taking the absolute differences in Equations 17 and 18). There is certainly cancellation of errors for all programs given the larger absolute errors in Figure 6. However, CSMGem gives the smallest prediction errors for both cases (relative and absolute). Conclusions

    In this work, we have given an overview of the

    hydrate fugacity model proposed by Ballard and Sloan (2002) and the implementation of the model into a Gibbs energy minimization program, CSMGem (Ballard 2002). The CSMGem program compares favorably to other commercially available programs. Nomenclature Latin Letters

    E objective function, error f fugacity g molar Gibbs energy (J/mol) h molar enthalpy (J/mol) K distribution coefficient (K-value) P pressure (bar) R universal gas constant S objective function T temperature (K) v molar volume (cm3/mol) x mole fraction of phase z mole fraction of feed

    Greek Letters molar phase fraction fugacity coefficient activity coefficient chemical potential (J/mol) number of phases, pi number of cavities per water molecule fractional occupancy, stability variable

    property difference between ice/Lw and phases (current model)

    property difference between real H and phases (proposed model)

    Subscripts Aq aqueous phase H hydrate phase i arbitrary component k arbitrary phase m hydrate cavity o ideal gas at 1 bar r reference phase w water (as a component)

    0 property at reference conditions (298.15 K, 1 bar) standard empty hydrate lattice

  • References Baker, L.E., Pierce, A.C., & Luks, K.D. (1982). Gibbs

    energy analysis of phase equilibria. Soc. Petr. Eng. J., October, 731-742.

    Ballard, A.L. Ph.D. Thesis (2002). A non-ideal hydrate solid solution model for a multi-phase equilibria program. Colorado School of Mines, Golden, CO, USA.

    Ballard, A.L., & Sloan, E.D., Jr. (2002). The next generation of hydrate prediction: I. Hydrate standard states and incorporation of spectroscopy. Fluid Phase Equilibria, in press.

    Davy, H. (1811). The Bakerian lecture. On some of the combinations of oxymuriatic gas and oxygen, and on the chemical relations of these principles to inflammable bodies. Phil. Trans. Roy. Soc. (London), 101, 1-35.

    DePriester, C.L. (1953). Light-hydrocarbon vapor-liquid distribution coefficients: Pressure-temperature-composition charts and pressure-temperature nomographs. Chem. Eng. Prog. Symp. Ser., 49(7), 1-43.

    Gibbs, J.W. (1875). On the equilibrium of heterogeneous substances. Reprinted in The Scientific Papers of J. Willard Gibbs, Vol. I, Dover Publications, Inc., N.Y., 196.

    Gupta, A.K., Ph.D. Thesis (1990). Steady state simulation of chemical processes. University of Calgary, Calgary, Canada.

    Hadden, S.T., & Grayson, H.G. (1961). New charts for hydrocarbon vapor-liquid equilibria. Hydrocarbon Proc. and Petrol. Refiner, 40(9), 207-218.

    Hammerschmidt, E.G. (1934). Formation of gas hydrates in natural gas transmission lines. Ind. Eng. Chem., 26, 851-855.

    Holder, G.D., Corbin, G., & Papadopoulos, K.D. (1980). Thermodynamic and molecular properties of gas hydrates from mixtures containing methane, argon and krypton. Ind. Eng. Chem. Fund., 19(3), 282-286.

    Katz, D.L. (1945). Prediction of conditions for hydrate formation in natural gases. Trans. AIME, 160, 140-149.

    Menten, P.D., Parrish, W.R., & Sloan, E.D., Jr. (1981). Effect of inhibitors on hydrate formation. Ind. Eng. Chem. Proc. Des. Dev., 20(2), 399-401.

    Parrish, W.R., & Prausnitz, J.M. (1972). Dissociation pressures of gas hydrates formed by gas mixtures. Ind. Eng. Chem. Proc. Des. Dev., 11(1), 26-35.

    Rachford, H.H., & Rice, J.D. (1952). Procedure for use of electronic digital computers in calculating flash vaporization hydrocarbon equilibrium. J. Petrol. Tech., 4(10), 10-19.

    Saito, S., Marshall, D.R., & Kobayashi, R. (1964). Hydrates at high pressures: Part II. Application of statistical mechanics to the study of hydrates of methane, argon, and nitrogen. AIChE J., 10(5), 734-740.

    Waals, J.H. van der, & Platteeuw, J.C. (1959). Clathrate solutions. Adv. Phys. Chem., 2, 1-57.

    Wilcox, W.I., Carson, D.B., & Katz, D.L. (1941). Natural gas hydrates. Ind. Eng. Chem., 33, 662.

    AbstractNomenclatureReferences