0st HTD Calculating Z and EOS
0st HTD Calculating Z and EOS
0st HTD Calculating Z and EOS
The following 3 effects should be taken into account for any valid model of the density behavior of pure liquids and their mixtures.
Temperature effect:
The expansion and contraction of the volume of a fixed mass of constant composition liquid is the most significant departure term that must be quantitatively
described. The change in specific volume with a change in temperature is defined by the thermodynamically fundamental term known as the coefficient of
thermal expansion.
1 dV
α=
V dT
The final form of any model relating α to easily obtainable measurements depends upon the integration of this basic definition. The integration, in turn, depends
upon the assumptions made and the sequence in the derivation at which the assumptions are invoked.
Typical temperature corrections for hydrocarbon oils in the neighborhood of 0.62 specific gravity are about 0.09$ per ℉ (0.2% per ℃). Data that have been
developed on butanes and lighter show much higher corrections, 0.2% per ℉ (0.4% per ℃).
Pressure effect:
Compressibility is a physical property causing an isothermal reduction in volume occupied by a constant composition fluid as pressure increases. The degree of
compressibility of liquids is defined by the instantaneous compressibility.
β= ( −1V ) dVdP
This equation states that the compressibility is the slope of the P-V curve at a given pressure and temperature divided by the total volume at that pressure.
For industrial applications the pressure effect is normally expressed by compressibility factor
F=
1
∆P (
1−
V
Vs )
Which can be written
F=
( V s−V
P−Ps )( )
1
Vs
In this equation ∆V/∆P represents the mean volume change from saturation pressure to system pressure.
For heavy hydrocarbons at pressures and temperatures well below their critical condition, the adjustment for pressure is small. It is on the order of 0.00005%
per psi (0.0003% per kPa). For light hydrocarbons at 100 ℉, the correction is in the magnitude of 0.004% psi (0.024% per kPa).
Mixing effect:
The volume of hydrocarbon mixtures will be less than the combined volume of the components existing at the same conditions.
The magnitude of this “volume shrinkage” is dependent on the composition and type of hydrocarbons being blended as well as the temperature and pressure of
the resulting mixture.
This chemical/physical phenomenon is caused primarily by 3 basic factors:
When combining components whose molecules differ considerably in size, the smaller molecules pack into voids between the larger molecules.
When the operating conditions approach the critical conditions of one of the components, there is more free space between the molecules of the near
critical component. In addition, the increased kinetic energy causes the molecules of the near critical component to become malleable. This permits the
more stable molecules of the heavier, and large, components to force their way between the lighter molecules. This effect is typical in NGL streams
where ethane is a near critical material. The excess volume in this case can excess -5% and increases with decreasing pressure and increasing ethane
mole fraction.
When the components exhibit a high degree of hydrogen bonding, the pseudochemical effect produces more compact volumes than can be attributed
to the purely physical mixing phenomenon previously described. The chemical shrinkage upon mixing is accompanied by discernable heat release.
The modeling effort usually attempts to account for the temperature and pressure effects by various integrations of Equation (α and β). Hence it is convenient to
define two states of liquids, one of which is pressure independent.
Saturated liquids
A saturated liquid is a pure liquid in equilibrium at its own vapor pressure, or a liquid mixture at its bubble point, either being at a specified temperature.
The bubble point pressure is that pressure at which the first bubble of vapor forms.
Densities of such liquids are sometimes called “orthobaric” or “bubble point” densities.
It is easily shown by the Phase Rule that in either case pressure is not an independent variable.
Compressed liquid.
A compressed liquid is defined as one for which the pressure at the prevailing temperature is greater than the vapor pressure for a pure compound or greater
than the bubble point pressure for a multicomponent liquid mixture.
The densities of many compressed liquids at nominal pressures are not much different from the saturated densities.
In fact, for many process applications this difference is considered to be insignificant. Thus, in these cases, the saturated liquid densities often adequately
represent compressed liquid densities for up to several atmospheres of excess pressure. On the other hand, at very high pressure, or even at moderate pressure
for liquid metering operations, changes in liquid density may be of considerable importance. This is especially true of liquid determinations for the purpose of
custody transfer of such liquids as LNG/LPG mixtures and NGL products. In these cases, the pressure effect is significant and high degree of accuracy of quantity
measurement is required.
The models used to represent these effects may be classified as Factor Models or Compositional Models.
Factor Models:
Factor models express the temperature and pressure dependence as differences in density from a known or base density.
For example, the oil industry expresses the temperature dependence in terms of a volume correction factor defined by
ρ
VCF=
ρT 0
The factor approach is based on the requirement that the base density is known. This known value is usually measured but may be computed from a
compositional model in some cases.
The factor approach has the built in assumption that mixing effects are not functions of temperature or pressure. Hence the measurement of a 60 ℉ density of a
mixture of specified composition takes mixing effect into account. The factor model approach then corrects this single measurement for temperature and
pressure variation.
The approach is highly accurate when used for oils and petroleum fractions with densities greater than 620 kg/m³ at 60 ℉.
The compositional effects on light ends are, however, dependent on both temperature and pressure. Hence, the attempt to factor density from a base value for
the light ends results in a significant loss of absolute accuracy.
Compositional Models:
The compositional model depends on both the existence of a sufficient analysis of the compositions of the molecular species involved and on set of mixing rules
capable of blending model parameters using these compositions.
This common model is useful for process calculations involving mixture comprised of components of similar molecular size and structure.
Expected errors are, when the equation applies, on the order of 3 to 5%.
The challenge of the accurate compositional model is the adequate representation of the mixing effect. Hence, the typical set of mixing rules are far more
complicated than shown in Equation.
The most common Factor Model in worldwide use is embodied in the Petroleum Measurement Table (1980). These are represented by API Satndard 250, IP
Standard 200, ASTM Standard D1250 in addition to the standards of other national bodies and the International Standards Organization, ISO/R91.
The approach was developed in 1916 by Beurce and Peffer. Their data base consisted of but 18 crude oils and 60 products. All samples were of American origin.
These original tables, with few additions over the years, were the basis for oil and hydrocarbon product measurement until 1980.
In 1972, citing studies by Downer and Inkley, The Institute of Petroleum showed that thermal expansion coefficient data for crude oils of current economic
importance had thermal expansion coefficients averaging 5% greater than those represented in the 1940’s edition of Table 6 in API Standard 2540.
In 1974, the American Petroleum Institute (API) and the National Bureau of Standards (NBS) initiated a cooperative venture, funded by the API, to create a data
base of density measurements on both crude oils and refined petroleum products. The joint venture by API and NBS provided the scientific data base for the
development of the more accurate measurement tables.
The completion of this 5 years, $500,000 project in March, 1979 provided the necessary data to modernize the tables of API Standard 2540. Using the NBS
density data and related publications of outstanding technical authorities, the members of the Physical Properties Working Group of the AI/ASTM Joint
Committee on Static Petroleum Measurement (COSM) produced the modernized tables for Volume Correction Factors (VCF) presently available from API.
The new tables were developed from the NBS data by screening the data by the use of linear equations and computer generated plots. During the screening
process it was determined that there were five major identifiable groups of materials that had significantly different relationships between the coefficient of
thermal expansion and density. The coefficient of thermal expansion of crude and the 4 classes of products (gasolines, jet fuels, fuel oils and lube oils) follow
separate curves as a function of inverse density squared, or corresponding °API Gravity ranges. Thus the strict factor model was modified to incorporate the
compositional effect by grouping materials into classes.
A study was shown that it was not possible to represent the different classes of fluids with a single or table. Hence, separate tables were developed to represent
crude oils and products. The products table consists of segments representing each of the major products classes. The dashed region, called the transition zone,
was filled in to eliminate any discontinuities in the tables.
In addition to the crude and products tables, a third type of table was developed for special applications. This third table provides a mechanism for incorporating
those materials that do not have thermal expansion properties similar to those fluids tested by the NBS. It is highly probable that most materials of this class are
not naturally occurring petroleum oils or products obtained from such oils. It is anticipated that this third table will find maximum utilizations in the custody
transfer of alternative energy fluids. Such applications include liquids from coal, tar sands, shale oils and blends of petroleum with the biomass fluids or alcohol
from other sources. This, of course, does not prevent the special application tables from being used for crudes and products under the conditions described
below.
The new tables retained the format of the old Table 6, which Volume Correction Factors or densities tabulated as functions of temperature. The products Table
was computed in the segments tabled below:
Products: °API
Fuel Oil: 0 to 37 °API
Jet Fuel Type A & A1 37 to 50 °API
Gasoline 50 to 85 °API
Crude oil covers a range from 0 to °API Gravity
Table 6C, the Special Application Table, represents tabular entries of volume correction factor (VCF) against thermal expansion coefficient and temperature.
It is suggested that Table C be used when:
1.Table A & B do not adequately represent the thermal expansion properties of the fluids of interest and;
2.Precise thermal expansion coefficients may be obtained directly or indirectly by experiments, and;
3.If buyers and sellers of petroleum products or crude agree that for their use, a greater degree of equity can be obtained.
Correlation development:
The 2,278 point NBS/API data base was reduced to a predictive correlation using an integrated form of Equation of α. The following equation was premised:
α =α T + β ∆ T
0
where
∆ T =T −T 0
These equations were statistically validated by computer studies of the NBS data base. The precise value of k was selected from a consideration of:
(a) the computer studies,
(b) the theoretical curvature of density with temperature, and
(c) high temperature literature data on crudes, petroleum fractions, and C6 through C32 alkanes.
These literature data were obtained from work of Jessup and Orwall and Flory.
The value of k best expressing these criteria is 1.6.
Where
It was determined that the coefficients of thermal expansion at the base temperature for each group are related to the densities at the base temperature by:
k 0+ k 1 ρ T
αT = 2
0
0
ρT 0
The values of k1 and k2 were established for each major group from a simultaneous nonlinear regression of all data points within that group.
Constants for English units (per ℉). Obtained from Global regression of NBS data.
Because of the economic impact of the new table, two major independent tests were commissioned. The first, a study of Prudhoe Bay crude oil, supplied by
SOHIO, was performed in the laboratories of Phillips Petroleum Company. A detailed description of these results is given in Volume X of the new table. These
data both confirmed the accuracy of the new table and showed errors of up to 0.3 percent in the old tables.
The second set of tests were performed on twenty of the original samples at temperatures up to 300 ℉ and down to - 50 ℉. The results of these tests, which
were performed under API contract at the University of Missouri-Rolla, have not yet been publically released. The evaluation definitely shows a confirmation of
the new tables to 300 ℉ and to low temperature of -50 ℉.
Expansions to the Factor Model have been developed for lubricating oils and gasohol. The lubricating oils publication became the source document for the
fourth in the series of standard tables and was given the designation “D” series.
In summary, the new tables are based on the largest and most representative set of density data on oils and fractions ever collected. The data were measured by
the most modern techniques in one of the most highly respected laboratories in the world; that of the U.S. National Bureau of Standards.
The crude oils in the data base were obtained from over 40% of the world’s crude production in 1974. The refined products were representative of the major
classes of fluids transported in bulk. When compared to limited amount of data on primarily domestic stock collected in 1915 and the late 1920’s, which formed
the basis of the old measurement tables, the new data base must be considered orders of magnitude more representative of modern crudes and products.
These new measurement table will significantly improve the pipeline industry’s ability to measure petroleum more accurately in custody transfer transactions
and improve operating overages and shortages caused by these transactions.
Liquid density correlations may be divided into classes in another way. They may be empirical, theoretical, or semitheoretical in origin. Most of the published
correlations have been empirical; that is, they are not based on particular physical model. The developer of the correlation has found some mathematical
expression describing the relationship between one or more independent variables usually temperature, pressure, and composition, and the dependent
variable, density. Theoretical correlations are derived from the behavior or well-defined models. Two examples are hard sphere model in which the atoms or
molecules of the liquid are assumed to behave as rigid spheres, and the cell model in which each atom or molecule of the liquid is assumed to be contained in a
cell or cage made up of the other liquid particles.
There are also some theoretical models describing the behavior of mixtures. Some of these models are based on the assumption of random mixing while others
treat non-random mixing. Corresponding states correlations may be considered to be semitheoretical in that they adhere to the Principle of Corresponding
States but the analytical forms are of empirical origin.
The principal of corresponding states was developed from the observation of experimental data. It was discovered that many properties of pure compounds
change by about the same percent when compared on a common basis. The classical common basis is the same reduced temperature and pressure. The
reducing parameters are the critical temperature and critical pressure. Enhancements to the principal have been common during the past twenty-five years. The
most successful enhancement is the addition of a “third parameter” by Curl and Pitzer. They called their third parameter the acentric factor.
The use of the principal of corresponding states for mixtures presents a significant problem. A suitable set of reducing parameters must be selected to represent
mixture behavior as a function of composition. The reducing parameters for mixtures are called pseudocritical properties and are defined by a variety of mixing
rules.
Equation of state, relationships between pressure, temperature, and composition, are some times used to calculate both liquid as well as vapor densities. They
have one advantage over most liquid density correlations: the calculated properties change smoothly as the critical region is crossed. This is not generally true of
liquid density correlations. However, equations of state such as the Soave-Redlich-Kwong and the Peng-Robinson equations are generally not suitable for
accurate calculation using equations of state are seldom smaller than five percent and are often as large as ten to fifteen percent.
In the discussion below, the application of saturated liquid density correlations to LNG’s will be emphasized.
Klosek-McKinley Correlation
This correlation has been called a “totally empirical recipe for calculating the density of an LNG-like mixture.”
The calculated density is a function of temperature and composition, but not pressure. The specific volume of the mixture is calculated from the equation
V m =∑ x i V i−k x CH 4
i
Where the xi and the Vi are the mole fractions and the specific volumes of the components, xCH4 is the mole fraction of methane, and k is a constant which
depends on both temperature and the molecular weight of the mixture. The Vi vary with temperature also, of course.
McCarty-Klosek-McKinley
Fitting very accurate density data for “LNG-like mixture” measured recently at the U.S. National Bureau of Standards laboratories in Boulder, Colorado, McCarty
reported that the values of Klosek and McKinley’s k fell on one of two curves. The curve depended on whether the mixture contained no nitrogen or about 4.5%
nitrogen. He then developed the equation:
V m =∑ x i V i−
i
[ k 1+(k 2−k 1 )x N 2
0.0425 ]x CH 4
In which all terms have the same meaning as those in Klosek-McKinley’s equation except that k1 and k2 have different numerical values and xN2 is the mole
fraction of nitrogen. Of the forty densities calculated for the “LNG-like” mixtures in the NBS data base, thirty-one were within 0.1% of the measured values and
eight of the other nine were within 0.2%.
The advantages of McCarty’s modification of the Klosek-McKinley method are its accuracy and its simplicity. The disadvantages are its severely limited
composition and temperature ranges. It is valid only for LNG-like mixture containing at least 60% methane, less than 4% nitrogen, less than 4% each of iso and
normal butane, and less than 2% total of normal and isopentane, at temperature between 105 and 115 K.
Rackett equation
The Rackett equation is an empirical correlation which is said to reproduce the densities of pure, saturated liquids from their triple points to their critical points.
The equation as originally written by Rackett is:
( )
2
Vf
log =(1−T R )7 log Z c
Vc
Where Vf is the saturated liquid specific volume. The term Vf/Vc is the reduced volume. The equation is a relating reduced volume to reduced temperature. The
Rackett equation is, then, an equation relating reduced volume to reduced temperature, which makes it a corresponding states type of equation. The equation is
empirical in the sense that, as Rackett wrote, “There is no known theoretical basis for (the equation) or for the specific value selected for this exponent.” Rackett
also points out that several liquid density corresponding correlations have used one or more terms containing the quantity (1-TR) for their temperature
functions. In summarizing his comparisons, Rackett wrote “… the equation here proposed is more precise than any of the others.” It is interesting to note,
however, that his equation gave the smallest standard errors for only five out of twenty-nine data sets for sixteen compounds, while each of the other
correlations gave the lowest errors in eight cases. Rackett also points out that several polar liquids did not “conform” to his equation. That is, their density-
temperature curves did not have the correct shape.
Rackett devised rather complex mixing rules for his equation and made rather limited comparisons of their reliability. He compared the results calculated using
his method to experimental values and to the values calculated using four other methods for only six different methane-pentane mixture at 100 ℉. In addition,
he compared his own results with experimental measurements on twenty-two other mixtures.
In testing the Rackett equation and its closet competitor, the Francis equations, Spencer and Danner found that the former reproduced densities for
hydrocarbons considerably better, but that the latter was better for “organics” (non-hydrocarbons) and inorganics. Although they conclude that “… the modified
Rackett equation is far the best for the hydrocarbons and quite good for the other organic and inorganic materials,” a rough adjustment for the number of data
points in each class (1948 for hydrocarbons), 652 for “organics”, and 148 for inorganics) shows that the Francis equations are significantly better.
In their modification of the Rackett equation for mixture Spencer and Danner give these mixing rules:
Z RA =∑ x i Z RA
m i
i
V cm =∑ x i V ci
i
T cm=∑ ∑ φi φ j T cij
i j
x i V ci
φ i=
∑ xi V ci
i
T cij =( T ci T cj )1/ 2 ( 1−k ij )
k ij=1−[ ]
Spencer and Danner found “… there is no apparent criterion for recommending Harmen’s method of the modified Rackett”
The advantages of Spencer and Danner’s modification of the Rackett equation are its applicability to a large number of compounds and mixtures over wide range
of temperatures, its accuracy, and the simplicity of the pure-liquid equation. The drawback are its lower accuracy for polar compounds and relatively small
number of compounds for which values of ZRA are available. The Spencer -Danner-Rackett equation has not, to our knowledge, been evaluated explicitly for
LNG’s.
Hankinson and Thomson tested several different sets of mixing rules and recommended the following:
1 /2
V ij T cij=( V i T ci V j T cj )
¿ ¿ ¿
N N
∑ ∑ ( xi x j √V ¿ i T ci V ¿ j T cj )
i=1 j=1
T c ,m= ¿
V m
(∑ (∑ )(∑ ))
N N 2 N 1
¿ 1 ¿ ¿3 ¿ 3
V = x i V +3
i xi V i xi V i
4 i =1 i=1 i=1
N
ω SRK ,m =∑ x i ω SRK ,i
i=1
Comparisons of COSTALD with the Yen-Woods correlation and the Spencer-Danner-Rackett method over a fairly large quantity of mixture density data showed
that COSTALD gave significantly lower errors.
Interaction parameters have been developed for some of the light paraffins, carbon dioxide, and nitrogen to permit the use of COSTALD for custody transfer of
LNG’s. In comparison COSTALD with McCarty’s version of the Klosek-McKinley equation, Hankinson, Coker, and Thomson found that after inserting interaction
parameters, COSTALD gave an average absolute error for the densities of the whole 285 points of National Bureau of Standards mixture data of 0.199%
compared with 0.227% for McCarty’s equation.
Advantages of the COSTALD correlation are its high accuracy for many types of liquids and their mixtures over wide temperature ranges, its overall accuracy, and
the availability of the required parameters for many liquids. The COSTALD correlation has several unique features. One of these is its accuracy in the prediction
of LNG densities. It has also been shown that COSTALD fits densities of crude oils and petroleum fractions quite well, and it has been suggested that it can fit
densities of aqueous solutions well. Finally, Hankinson and Thomson showed that, given good values of Tc and ωSRK, COSTALD can be used to give good
estimates of the critical volumes of nonpolar compounds.
There are a few disadvantages. The COSTALD equations are not as simple as one might like, although this presents no problem on a digital computer. Its
predictions close to critical point are less accurate than we would like, and it does not predict densities of polar liquids as well as would be desired. COSTALD is
still being applied to new systems, however, and we may see some interesting new results soon.
The existing factor model as contained in the present API 1101 is generally accepted as being inadequate. The tables 1101 were constructed graphically and
basic mathematical model to represent the values does not exist. In 1979 Hankinson and Phillips regressed these tables to a complex function. This model was
given to the API and it in general domestic use.
In 1979 Donner of British Petroleum presented and alternative model to the ISO/TC 28/SC3. This model is given by:
Where
T: ℃
Ρ : 15 ℃ density, kg/L
He claims a calculational error of 0.05% in volume for temperatures up to 76 ℃ and pressure of 34 bar. The American Petroleum Institute has a project under
the technical direction of Dr. Mark Plummer (Marathon Oil) to incorporate new data into the existing base and develop a replacement for API 1101. This work is
complete and is presently in COPM ballot. Equations and results cannot be released until the ballot procedure is complete.
Previous Work:
Thomson, Brobst, Hankinson and Hankinson-Thomson have carefully examined the previous available compressed liquid density correlations. Each of these
correlations was found to be deficient for general for general use in one respect or another. The most common problem was a lack of mixing rules. Other
problems included narrow ranges of temperature, pressure, and composition; restricted component list; and difficulty in translating the method to computer
applications
The rationale behind the development of COSTALD was to combine as many desirable features as possible into a single unified, accurate liquid density
correlation. The following attributes were considered to be important in the formulation of a new compressed liquid density correlation:
1.Applicable to many types of liquids including nonpolar, slightly polar, polar, and quantum.
2.Contain generalized mixing rules which are consistent for all types of liquid mixtures.
3.Contain only generalized parameters of readily available, or easily determined, pure component data.
4.High degree of accuracy, particularly for refrigerated liquid mixtures such as NGL products and LNG/LPG mixtures.
5.Predictive, so that compressed densities can be calculated for those liquids for which no experimental density data exist.
6.Readily adaptable to efficient computer and on-line microprocessor programming for use in process design and simulation calculations, and for liquid metering
in custody transfer operations.
7.Relatively simple equations.
8.Continous, single-valued functions over the whole range of interest.
9.Compressed density function that smoothly approaches the saturated density and whose lower limit exactly reduced to the saturated density.
10.Give accurate calculated densities near the critical point.
COSTALD Correlation
This section outline the extension of the generalized COSTALD method for saturated liquid densities of pure compounds and their bubble point mixtures, shown
above, to the calculation of compressed liquid densities. The resultant COSTALD correlation is applicable for pressure from saturation up to 68,950 kPa (10,000
psia) and temperature up to just below the critical temperature. With the partial exception of items 7 and 10, all of the desirable features listed above have been
incorporated into the COSTALD compressed liquid density correlation. The double sums and product of sums, which appear in the mixing rules, are quite tedious
to calculate by hand for mixtures of more than two or three components. However, these terms pose no difficulty for programed computer calculations.
(
V =V s 1−C ∙ ln
( B+ P
B+ P s ))
Where the B and C terms have been correlated by
[ ]
1 2 1
3 3 3
B=P s −1−9.070217(1−T r ) +62.45326(1−T r ) −135.1102 ( 1−T r ) + e1 (1−T r )
e 1=exp ( 4.79594+ 0.250047 ωSRK +1.14188 ω 2
SRK )
C=0.0861488+0.0344483 ω SRK
Note that in the above equation (V) as P approaches Ps the natural logarithm of (B+P)/(B+P) approaches zero and, consequently, the molar volume approaches
the saturated molar volume Vs. This is an essential feature of the correlation. The corresponding states principle was used to correlate B, which has units of
pressure, as a function of reduced temperature. Other criteria, the step-by-step procedures, and the experimental data sets used both for determining the final
form of the parameters B, e1, and C and for evaluation of the numerical constants are discussed in greater detail in reference.
In order to use equation Vs above, a knowledge of both the saturated molar volume Vs and the saturation pressure Ps at the system temperature is required, Vs
is determined by the procedures given above. The saturation pressure Ps, or Psm for bubble point mixtures*, may be calculated from the following generalized
Riedel vapor pressure equation developed by Hankinson, Estes, and Coker:
(0 ) (1 )
log ( P Rs )=P R + ω SRK P R
(0 )
P R =5.8031817 ∙ log ( T R ) +0.07608141 F
P(1)
R =4.86601 G
36.0
F=35.0− −96.376∙ log ( T R )+ T 6R
TR
G=log ( T R ) +0.03721754 F
Ps =( P Rs Pc )
Alternatively, any other suitable vapor pressure equation such as
C2
ln Ps =C1 + +C3 T +C 4 ln T
T
* Note *: The subscript m is added to denote a mixture property when the equation in this section are applied to mixture, e.g. Tcm, TRm, Pcm, PRm, Vsm, Vm,
etc.
In addition to the mixing rules, a knowledge of the mixture pseudocritical pressure Pcm is needed when Equations above for parameter B are being applied to
compressed liquid mixture. Thompson, et al recommended the following equations:
Z cm R T cm
Pcm = ¿
Vm
Z cm=0.291−0.080 ω SRK ,m
The accuracy of determining the density of mixtures, such as LNG, depends upon accurate compositional analyses, since the mole fractions of all components, xi,
are needed both in the COSTALD mixing rules and to calculate the mixture molecular weight Mm:
Mm
ρm =
Vm
M m= ∑ x i M i
i
The composition of liquid mixtures can be determined from chromatographic analysis, which yields accurate value of component weight fractions. Weigh
fractions are easily converted to mole fraction for insertion into the above equation.
COSTALD Accuracy:
Thomson, Brobst, and Hankinson tested the COSTALD compressed liquid density correlation against a total of 7690 data points of pure liquid compounds and
6926 data points of liquid mixtures. The results are given in below Table along with the corresponding results obtained from Yen-Woods correlation. The
COSTALD correlation gives significantly less error than the Yen-Woods correlation in all cases. These data covered temperatures from 50 to 600 K and pressure
from saturation to 68,950 kPa (10,000 psia). Of particular interest are the 319 data points for LNG/LPG mixtures for which COSTALD densities averaged 0.369
percent absolute error, while the Yen-Woods errors average nearly four times that or 1.46 percent.
Table: Average Absolute percent error and bias between calculated and experimental density of compressed liquids
Pure compounds
Non polar 6338 0.446 -0.269 1.49 -5.72
Polar and Quantum 1352 2.57 -9.75 3.48 -0.770
Mixtures
All mixtures 6926 1.61 9.50 2.51 -12.5
LNG/LPG mixtures 319 0.369 1.26 1.46 -6.82
HYSYS SOFTWARE
Peng-Robinson
See Also
Equation of State (EOS)
Peng-Robinson Calculation Methods
Peng-Robinson Parameters
The Peng-Robinson (PR) model is ideal for VLE calculations for hydrocarbon systems. Several enhancements to the original PR model were made to extend its
range of applicability and to improve its predictions for some non-ideal systems. However, in situations where highly non-ideal systems are encountered, the use
of Activity Models or advanced equations of state, such as CPA or PC-SAFT, is recommended.
The PR property package rigorously solves any single-, two-, or three-phase system with a high degree of efficiency and reliability and is applicable over a wide
range of conditions:
The PR property package also contains enhanced binary interaction parameters for all library hydrocarbon-hydrocarbon pairs (a combination of fitted and
generated interaction parameters), as well as for most hydrocarbon-non-hydrocarbon binaries, such as N2, CO2, H2S, H2, and H2O. For non-library or
hydrocarbon hypocomponents, HC-HC interaction parameters are generated automatically by HYSYS for improved VLE property predictions.
For Oil, Gas, or Petrochemical applications, the PR EOS is the generally recommended property package. The PR property package can be used for the following
simulations:
The PR equation of state applies a functionality to some specific component-component interaction parameters. Key components receiving special treatment
include He, H2, N2, CO2, H2S, H2O, CH3OH, EG, DEG, and TEG.
Where
V is molar volume
P is system pressure
T is system temperature
aP
A= 2
( RT )
bP
B=
RT
( R T ci
)
N
b=∑ x i 0.077796
i=1 Pci
[( ]
0.5
)( )
N N 2 2
(R T ci) (RT cj )
a=∑ ∑ x i x j 0.457235 α i 0.457235 αj ( 1−k ij )
i=1 j=1 P ci Pcj
α i =1+m i ( 1−T ri )
0.5 0.5
For the Peng-Robinson Equation of State, the enthalpy and entropy departure calculations use the following relations:
[ ] [
V + ( 2 +1 ) b
]
ID 0.5
H−H 1 da
=Z−1− 1.5 a−T ln
RT 2 bRT dt V − ( 2 −1 ) b
0.5
S−S0 ID
R
P
=ln ( Z−B )−ln 0 − 1.5
A
P 2 bRT adt
Tda
[ ]
ln
[
V + ( 20.5 +1 ) b
V −( 2 +1 ) b
0.5
]
Where
HID = Ideal Gas Enthalpy basis used by HYSYS changes with temperature according to the coefficients on the TDep tab for each individual component
a=∑ ∑ x i x j ( ai a j ) 0.5 (1−k ij )
i j
R T ci
b i=0.077796
P ci
2
mi=0.37464 +1.54226 ωi−0.26992ω i
H = Enthalpy
S = Entropy
Select the Peng-Robinson property package and click the Set Up tab. The following options are available when the Peng-Robinson package is selected:
Setting Options
Enthalpy Property Package EOS or Lee-Kessler
Density:
Costald: When this option is selected, you can specify the Chueh-Prausnitz or the Tait Equation.
Rackett: Recommended for petroleum and hydrocarbon liquid mixtures at low and moderate pressure.
Use EOS Density: When Use EOS Density is selected, you can specify volume translation parameters. For further information, refer to
Volume Translation (Peng-Robinson Property Package).
Modify Tc, Pc, for H2, He Modify or un-modified
Indexed Viscosity
HYSYS Viscosity:
If you select HYSYS Viscosity from the Indexed Viscosity drop-down list, a new Use Modified TRAPP Viscosity Model check box is
available. By default, this check box is cleared. Selecting this check box lets you use AspenTech's proprietary modified TRAPP viscosity
model to eliminate discontinuities observed in viscosity values. The modified model is based on Ely & Hanley (1981).
-or-
Indexed Viscosity
Peng-Robinson Options
HYSYS
-or-
Standard
If you select the Peng-Robinson Options field, a Smooth Water Alpha Function check box is available. This check box is cleared by default.
When the Smooth Water Alpha Function check box is cleared: Alpha function (1) is used when the T_stream/Tc_H2O < 0.85, which
attempts to match the steam table at low temperatures. Alpha function (2) is used when T_stream/Tc_H2O >=0.85, which attempts to
match the steam table at high temperatures. As a result, discontinuities occur for the water dew point curve at a T_stream/Tc_H2O of 0.85.
An example of the results when the check box is cleared appear below.
When the Smooth Water Alpha Function check box is selected, HYSYS avoids discontinuities at T_stream/Tc_H2O of 0.85 by smoothing the
water alpha function. An example of the results when the check box is selected appear below.
For HYSYS Peng-Robinson property packages, when Use EOS Density is selected from the Density drop-down list, you can specify volume translation parameters.
When Use EOS Density is selected, Volume Translation information appears in the Parameters group. Volume Translation is a widely used empirical method to
improve the accuracy of the liquid density calculated by the equation of state.
1.On the Set Up tab of the Fluid Package form, from the Property Package Selection list, select the Peng-Robinson property package.
2.In the Options group, from the Density drop-down list, select Use EOS Density.
The Parameters group appears, with a matrix containing the volume correction constants for each component currently selected. The default value for each
component is 0.
3.Perform one of the following tasks:
Manually specify values in this matrix.
-or-
Select the desired volume translation method, and then click Estimate Vol. Trans. to have HYSYS estimate all the missing values. HYSYS offers three
methods of estimating the volume translation parameter:
*COSTALD: This is the default selection. The COSTALD model includes both temperature and pressure dependence.
*RACKETT: The RACKETT model incorporates temperature dependence.
*Peneloux: Uses the Peneloux volume translation method (Pederson, Karen S., and Peter L. Christensen. Phase Behavior of Petroleum Reservoir Fluids.)
For both the COSTALD and RACKETT model, the following equations are used for volume translation:
V =V eos +c
c=∑ (x i ∙ c i)
Where
Veos is calculated by the Equation of State model (PR).
ci is the volume translation parameter of component i. ci can be estimated using the COSTALD and RACKETT methods.
COSTALD Volume Translation Method
Where VmR, 0 and VmR, δ are functions of Tr for 0.25 < Tr < 1.0
[( ∑ )( )]
2 1
1
+3 ∑ x i ( V i ) ∑ x i (V i )
CTD ¿,CTD ¿, CTD 3 ¿,CTD 3
Vm = xi V i
4 i i i
ω=∑ x i ωi
i
Where
1
V CTD
ij T cij=( V i ¿ ,CTD
T ci V j ¿,CTD
T cj ) 2
i
Vcm is the critical molar volume
V cm =∑ x i V ci
i
Tr is the reduced temperature
T r=T /T c
The binary interaction parameter kij is estimated automatically using the following equation:
1/ 2
8( V ci V cj )
k ij=1− 3
( V 1/ci 3 +V 1/cj 3 )
Peneloux Volume Translation Method
For non-hydrocarbons and hydrocarbons lighter than C7, Peneloux et al. recommend that the volume translation parameter of component ci can be estimated
by the following (Jhaveri and Youngren. SPE Res. Eng. 1033-1040, Aug. 1988.):
For hydrocarbons C7 or heavier, Pedersen et al. recommend the following equation to estimate ci:
c i=V stdi−V eosi
Where:
Vstdi is the value of the volume of the component at 15 °C and atmospheric pressure.
Veosi is the volume of component calculated by the Equation of State model (Peng-Robinson) at 15 °C and atmospheric pressure.
The calculated results of the liquid mass density of components C5~C20 using the Peng-Robinson property package with the Peneloux volume correction are
shown in the following figure.
SRK
See Also
Equation of State (EOS)
Peng-Robinson (PR)
TST Mixing Rules
SRK Parameters
In many cases, the Soave-Redlich-Kwong (SRK) model provides comparable results to Peng-Robinson, but its range of application is significantly more limited:
The SRK property package is generally used for the following simulations:
TEG Dehydration
Cryogenic Gas Processing
Air Separation
Atm Crude Towers
Vacuum Towers
High H2 Systems
Reservoir Systems
Hydrate Inhibition
Crudes systems
Nitrogen & Helium Rejection
Methanol Injection
Mercury Removal
GOSP
Fractionation
LNG Processes
Light Ends Separation
The proprietary enhancements to the SRK property package allow the SRK equation of state (EOS) to correctly represent vacuum conditions and heavy
components (a problem with traditional EOS methods), as well as handle the light ends and high-pressure systems.
The SRK property package contains enhanced binary interaction parameters for all library hydrocarbon-hydrocarbon pairs (a combination of fitted and generated
interaction parameters), as well as for most hydrocarbon-nonhydrocarbon binaries. For non-library or hydrocarbon hypocomponent, HC-HC interaction
parameters are generated automatically by HYSYS for improved VLE property predictions.
SRK Parameters
Perform the steps for selecting the SRK property package and then click on the Parameters tab. The following options are available:
Setting Options
Enthalpy Property Package EOS or Lee-Kessler
Density
Costald: When this option is selected, you can specify the Chueh-Prausnitz or the Tait Equation.
Rackett: Recommended for petroleum and hydrocarbon liquid mixtures at low and moderate pressure.
Use EOS Density: When Use EOS Density is selected, you can specify volume translation parameters. For further information, refer to
Volume Translation (SRK Property Package).
Modify Tc, Pc, for H2, He Modify or un-modified
Indexed Viscosity HYSYS or Indexed
EOS Solution Methods Cubic EOS Analytical or Numerical Method
Phase Identification
Default - If Default you can modify the tuning factor
Venkatarathnam-Oellrich1 - Uses a thermodynamic method for determining the phase of a fluid from the partial derivatives of
pressure, volume and temperature without reference to saturated properties. Particularly useful for liquid-liquid or vapor-liquid-
liquid equilibria calculations.
Surface Tension Method HYSYS or API 10A3.2 Method
Thermal Conductivity API 12A3.2-1 or API 12A1.2-1 Method
When Use EOS Density is selected, Volume Translation information appears in the Parameters group. Volume Translation is a widely used empirical method to
improve the accuracy of the liquid density calculated by the equation of state.
1.On the Set Up tab of the Fluid Package form, from the Property Package Selection list, select the SRK property package.
2.In the Options group, from the Density drop-down list, select Use EOS Density.
The Parameters group appears, with a matrix containing the volume correction constants for each component currently selected. The default value for each
component is 0.
3.Perform one of the following tasks:
Manually specify values in this matrix.
-or-
Select the desired volume translation method, and then click Estimate Vol. Trans. to have HYSYS estimate all the missing values. HYSYS offers three
methods of estimating the volume translation parameter:
*COSTALD: This is the default selection. The COSTALD model includes both temperature and pressure dependence.
*RACKETT: The RACKETT model incorporates temperature dependence.
*Peneloux: Uses the Peneloux volume translation method (Pederson, Karen S., and Peter L. Christensen. Phase Behavior of Petroleum Reservoir Fluids.)
c=∑ (x i ∙ c i)
Where
Veos is calculated by the Equation of State model (SRK).
ci is the volume translation parameter of component i. ci can be estimated using the COSTALD and RACKETT methods.
For non-hydrocarbons and hydrocarbons lighter than C7, Peneloux et al. recommend that the volume translation parameter of component ci can be estimated
by the following (Peneloux, A., and Rauzy, E. A consistent correction for Redlich-Kwong-Soave volumes. Fluid Phase Equilib. 1982, 8, 7-23.):
For hydrocarbons C7 or heavier, Pedersen et al. recommend the following equation to estimate ci:
c i=V stdi−V eosi
Where:
Vstdi is the value of the volume of the component at 15 °C and atmospheric pressure.
Veosi is the volume of component calculated by the Equation of State model (SRK) at 15 °C and atmospheric pressure.
The calculated results of the liquid mass density of components C5~C20 using the SRK property package with the Peneloux volume correction are shown in the
following figure.
Standard Peng-Robinson EOS
The Standard Peng-Robinson equation-of-state is the original formulation of the Peng-Robinson equation of state with the standard alpha function (see Peng-
Robinson Alpha Functions). It is the basis for the Peng-Robinson property method and it is recommended for hydrocarbon processing applications such as gas
processing, refinery, and petrochemical processes. Its results are comparable to those of the standard Redlich-Kwong-Soave equation of state.
RT a
P= −
V m −b V m ( V m +b ) + b(V m −b)
Where
b=∑ x i b i
i
(3)
(1) k ij(2)
k ij=k + k T +
ij ij
T
(kij=kji)
a i=fcn (T ,Tci , Pci , ωci)
b i=fcn(Tci , Pci)
In terms of Z:
Z −( 1−B ) Z + ( A−2 B−3 B ) Z −( AB−B −B )=0
3 2 2 2 3
If the Peneloux volume correction is used, then the molar volume is calculated from:
V =V m−c
Where
Vm Molar volume calculated by the equation of state without correction
c=∑ xi c i
i
(the Peneloux volume correction term)
R Ti
c i=0.50033 (0.25969−z RAi )
pci
(the Peneloux volume correction term for pure components, calculated from the critical temperature and pressure and Rackett parameter)
For best results, the binary parameters kij and lij must be determined from regression of phase equilibrium data such as VLE data. The Aspen Physical Property
System also has built-in kij and lij for a large number of component pairs in the EOS-LIT databank from Knapp et al. These parameters are used automatically
with the Peng-Robinson property method. Values in the databank can be different than those used with other models such as Soave-Redlich-Kwong or Redlich-
Kwong-Soave, and this can produce different results.
Peng-Robinson EOS
The Peng-Robinson equation-of-state is the basis for the PR-BM property method. The model has been implemented with choices of different alpha functions
(see Peng-Robinson Alpha Functions) and has been extended to include advanced asymmetric mixing rules.
By default, the Peng-Robinson property method uses the literature version of the alpha function and mixing rules (see Standard Peng-Robinson). The PR-BM
property method uses the Boston-Mathias alpha function and standard mixing rules. These default property methods are recommended for hydrocarbon
processing applications such as gas processing, refinery, and petrochemical processes. Their results are comparable to those of the property methods that use
the standard Redlich-Kwong-Soave equation-of-state.
When advanced alpha function and asymmetric mixing rules are used with appropriately obtained parameters, the Peng-Robinson model can be used to
accurately model polar, non-ideal chemical systems. Similar capability is also available for the Soave-Redlich-Kwong model.
The equation for the Peng-Robinson model as used in the PR-BM property method is:
RT a
P= −
V m −b V m ( V m +b ) + b(V m −b)
Where
b=∑ x i b i
i
a=a 0+ a1
a 0=∑ ∑ x i x j ( ai a j ) 0.5 (1−k ij )
i j
(the standard quadratic mixing term, where kij has been made temperature-dependent)
(3)
(1)k ij
(2)
k ij=k + k T +
ij ij
T
(kij=kji)
( )
n n 3
1/ 3
a 1=∑ x i ∑ x j ( ( ai a j ) lij)
1 /2
i=1 j=1
(an additional, asymmetric term used to model highly non-linear systems)
(1) (2 ) l (3ij )
l ij =l ij +l ij T +
T
(In general, lij ≠ lji)
a i=fcn (T ,Tci , Pci , ωci)
b i=fcn(Tci , Pci)
If the Peneloux volume correction is used, then the molar volume is calculated from:
V =V m−c
Where
Vm Molar volume calculated by the equation of state without correction
c=∑ xi c i
i
(the Peneloux volume correction term)
R Ti
c i=0.50033 (0.25969−z RAi )
pci
(the Peneloux volume correction term for pure components, calculated from the critical temperature and pressure and Rackett parameter)
For best results, the binary parameters kij and lij must be determined from regression of phase equilibrium data such as VLE data. The Aspen Physical Property
System also has built-in kij and lij for a large number of component pairs in the EOS-LIT databank from Knapp et al. These parameters are used automatically
with the Peng-Robinson property method. Values in the databank can be different than those used with other models such as Soave-Redlich-Kwong or Redlich-
Kwong-Soave, and this can produce different results.
These expressions are derived by applying the critical constraints to the equation-of-state under these conditions:
α i=( T ci ) =1.0
The parameter α is a temperature function. It was originally introduced by Soave in the Redlich-Kwong equation-of-state. This parameter improves the
correlation of the pure component vapor pressure.
This equation for αi is still represented. The parameter mi can be correlated with the acentric factor:
2
m i=0.37464 +1.54226 ωi−0.26992ω i
The above Equations are the standard Peng-Robinson formulation. The Peng-Robinson alpha function is adequate for hydrocarbons and other nonpolar
compounds, but is not sufficiently accurate for polar compounds.
The HYSYS alpha function uses the same equation as the standard Peng-Robinson alpha function when ω < 0.49, and otherwise (ω >0.49) it uses equation 4 with
the following definition of m to correct the behavior for large values of acentric factor:
mi=0.379642+1.48503 ωi −0.164423 ω2i +0.016666 ω 3i
The Aspen Physical Property System calculates the critical compressibility factor (ZC) by:
Pc V c
Z c=
RTc
Acentric Factor ω
Definition method
Lee-Kesler method
Definition Method
When you use the definition method, the acentric factor is calculated from its definition:
( )
¿
Pi
ω i=−log10 −1.0
Pci
Lee-Kesler Method
The Lee-Kesler method depends on TB, TC, and PC. This method is recommended for hydrocarbons. Lee and Kesler reported that this method yields values of
acentric factors close to those selected by Passut and Danner (Ind. Eng. Chem. Process Des. Dev.12, 365, 1973).
Model (Type)
API liquid volume (Liquid volume)
Brelvi-O’Connell (Partial molar liquid volume of gases)
Chueh-Prausnitz (Liquid volume)
Clarke Aqueous electrolyte volume (Liquid volume)
COSTALD liquid volume (Liquid volume)
Debye-Huckel volume (Electrolyte liquid volume)
Liquid constant molar volume model (Liquid volume)
General pure component liquid molar volume (Liquid volume/Liquid density)
Rackett/Campbell-Thodos mixture liquid volume (Liquid volume)
Modified Rackett (Liquid volume)
General pure comment solid molar volume (Solid volume)
At high density, the Ritter equation is used (adapted from Ritter, Lenoir, and Schweppe, Petrol. Refiner 37 [11] 225 (1958)):
[ ]
−5 −1 /2
l 1 2 (1.2655 SG−0.5098+8.011× 10 T b )(T −519.67)
V =
p SG −
62.3636 Tb
Where SG is the specific gravity, Tb is the mean average boiling point in Rankine, T is the temperature of the system in Rankine, and the mass volume is
produced in units of cubic feet per pound-mass.
Brelvi-O’Connell:
The Brelvi-O'Connell model calculates partial molar volume of a supercritical component i at infinite dilution in pure solvent A. Partial molar volume at infinite
dilution is required to compute the effect of pressure on Henry's constant. (See Henry's Constant.)
The liquid molar volume of solvent is obtained from the Rackett model:
[ 1+(1−T r )2 /7]
R T c Z RA
¿ ,l
V =
r
Pc
[
Pvp is saturation pressure = Pcm exp −5.3727 ( 1+ ωm )
( 1
Tm )]
−1 for mixtures, or calculated using the General vapor pressure model for pure components.
Pcm =Z cm T cm /V cm
Tci is critical temperature of component i
Pci is critical pressure of component i
The Clarke model uses this equation to calculate the liquid molar volume for electrolyte solutions:
l l l
V m =V s +V e
Where
Vlm is liquid molar volume for electrolyte solution
Vls is liquid molar volume for solvent mixtures
Vle is liquid molar volume for electrolyte
For electrolytes:
V e =∑ x ca V ca
l
ca
V ca =V ∞ca + A ca
√ x solute
1+ √ x solute
x solute =∑ x ca
ca
x w =x nws + x solute =1
Where:
Xca is apparent mole fraction of electrolyte ca
Vca is liquid molar volume for electrolyte ca
The mole fractions xca are reconstituted arbitrarily from the true ionic concentrations, even if you use the apparent component approach. This technique is
explained in Electrolyte Calculation in Physical Property Methods.
The result is that electrolytes are generated from all possible combinations of ions in solution. The following equation is consistently applied to determine the
amounts of each possible apparent electrolyte nca:
nc na z factor
n ca=
∑ nc ' z c '
c'
Where
Nca is number of moles of apparent electrolyte ca
Zc is charge of c
Zfactor = Zc if ca and a have the same number of charges; otherwise 1
Nc is number of moles of cation c
Na is number of moles of anion
For example: given an aqueous solution of Ca2+, Na+, SO42-, Cl- four electrolytes are found: CaCl2, Na2SO4, CaSO4, and NaCl. The Clarke parameters of all four
electrolytes are used. You can rely on the default, which calculates the Clarke parameters from ionic parameters. Otherwise, you must enter parameters for any
electrolytes that may not exist in the components list. If you do not want to use the default, the first step in using the Clarke model is to add any needed
components for electrolytes not in the components list.
The true molar volume is obtained from the apparent molar volume:
a
l,t l,a n
V m =V m t
n
Where:
Vl,tm is liquid volume per number of true species
Vl,am is liquid volume per number of apparent species, Vlm of equation above Vlm=Vls+Vle
Na is number of apparent species
Nt is number of true species
Temperature Dependence
The temperature dependence of the molar volume of the solution is approximately equal to the temperature dependence of the molar volume of the solvent
mixture:
l
l l V s (T )
V m ( T )=V m (298.15 K ) l
V m (298.15 K )
Where
Where Vml(298.15K) is calculated from equation above Vlm=Vls+Vle
Parameter Applicable Components Symbol Default Units
name/Element
∞
VLCLK/1 Cation-Anion V ca † MOLE-VOLUME
VLCLK/2 Cation-Anion Aca 0.020 MOLE-VOLUME
†If VLCLK/1 is missing, it is calculated based on VLBROC and CHARGE as follows:
V ca =VLCLK / 1=[ VLBROC /1(c)+298.15 VLBROC /2(c ) ] z a +[VLBROC /1(a1)+298.15 VLBROC / 2(a)] z c
∞
If VLBROC/1 is missing, the default value of -0.0012 is used. See the Brelvi-O'Connell model for VLBROC and also Rackett/Campbell-Thodos Mixture Liquid
Volume for additional parameters used in the Rackett equation.
Where:
VmR,0 and VmR,δ are functions or Tr for 0.25<Tr≤0.95.
For 0.95<Tr≤10, there is a linear interpolation between the liquid density at Tr = 0.95 and the vapor density at Tr = 1.05. This model can be used to calculate
saturated and compressed liquid molar volume. The compressed liquid molar volume is calculated using the Tait equation:
V m =V sat
(
m 1−C ln
( B+B+PP ))
sat
Where B and C are functions of T, ω, Tc, Pc and Psat is the saturated pressure at T.
Mixing rules:
V CTD
m =
1
4 [( (∑ xi V i +3∑ xi (V i )
i
¿ ,CTD ¿,CTD
i
2/ 3
)
) (∑ xi ( V i )
i
¿,CTD 1/ 3
)]
V m T c=∑ ∑ x i x j V ij T cij
CTD CTD
i j
ω=∑ x i ωi
i
Where
1/ 2
V CTD
ij T cij=( V ¿i ,CTD T ci V ¿j ,CTD T cj )
To improve results, the Aspen Physical Property System uses a special correlation for water when this model is used (1987). Changing the VSTCTD and OMGCTD
parameters for water will not affect the results of the special correlation.
Debye-Hückel Volume:
The Debye-Hückel model calculates liquid molar volume for aqueous electrolyte solutions.
k
Where
V*w is is the molar volume for water and is calculated from the ASME steam table.
Vk is calculated from the Debye-Hückel limiting law for ionic species:
∞
V m =V k +¿ z k ∨10
−3
( 3Avb ) ln (1+ b I 1 /2
)
Where
Vk∞ is partial molar ionic volume at infinite dilution
Zk is charge number of ion k
Av is Debye-Hückel constant for volume
b = 1.2
1
I= ∑
2 k
2
m k z k the ionic strength, with mk is molarity of ion k
Av is computed as follows:
A v =−2 x 106 A φ R 3
( ∂ l ln ε w 1 ∂ ρ w
∂P
+
Pw ∂ P )
Where
Aφ is Debye-Hückel constant for the osmotic coefficients (Pitzer, 1979)
( )
3/ 2
1 1/ 2 Q 2e
Aφ =
3
( (
2 π 10
−3
ρ w) N A)
εw k B T
Vk∞ = V1 + TV2
For liquid molar volume of mixtures, the Rackett mixture equation is always used. This is not necessarily consistent with the pure component molar volume or
density.
Many of these equations calculate density first, and return calculate liquid molar volume based on that density:
1
V ¿i ,l= ¿ ,l
ρi
DIPPR
DIPPR equation 105 is the default DIPPR equation for most substances:
¿ ,l C 1i
ρi =
1+(1−
T C
)
4i
for C 6 i ≤ T ≤ C7 i
C 3i
C 2i
In either case, linear extrapolation of ρi*,l versus T occurs outside of temperature bounds.
PPDS
PPDS Campbell-Thodos
The PPDS Campbell-Thodos model uses a form similar to the mixture Campbell-Thodos model:
2/7
1+(1−T r )
V i =C 1 [ C 2+ C3 (1−T r ) ]
¿ ,l
Where
Tr=T/Tci
Note: This equation uses the same parameter RACKET as the mixture Campbell-Thodos model, but it uses a different, incompatible definition of RACKET/3. Our
database does not include values for this parameter, but be aware that if you use this parameter in this equation, you cannot use the Campbell-Thodos mixture
equation.
Tmin and Tmax define the temperature range where the equation is applicable.
IK-CAPE
Rackett
Parameter RKTZRA is also used in other models that use the Rackett equation, and is the default for similar parameters in related equations.
The Rackett equation calculates liquid molar volume for all activity coefficient based and petroleum tuned equation of state based property methods. In the last
category of property methods, the equation is used in conjunction with the API model. The API model is used for pseudocomponents, while the Rackett model is
used for real components. (See API Liquid Volume.) Campbell-Thodos is a variation on the Rackett model which allows the compressibility term Zi*,RA to vary
with temperature.
Rackett
The equation for the Rackett model is:
2/7
RA 1+(1−T r )
l
R T c ( Zm )
V m=
Pc
Where:
T c =∑ ∑ x i x j V ci V cj ( T ci T cj )1/ 2 (1−k ij )/V 2cm
i j
Tc T
=∑ xi ci
Pc i P ci
Z m =∑ xi Z i
RA ¿ , RA
i
V cm =∑ x i V ci
i
T r=T /T c
1
( )
R T c [ 1+(1−T ) ] 2/7
= Z RA r
ρ Pc
Where:
Ρ is the density of the liquid mixture
R is universal gas constant
Tc is the critical temperature of the mixture
ZRA is an empirically derived constant
Tr is the reduced temperature, T/Tc
Campbell-Thodos
The Campbell-Thodos model uses the same equation as the Rackett model, above, except that ZmRA is allowed to vary with temperature:
m =∑ xi [ Zi
Z RA +d i (1−T r ) ]
¿ , RA
i
Campbell-Thodos uses a separate set of parameters, RACKET.
Note: The parameter RACKET is also used by the PPDS Campbell-Thodos equation in the General Pure Component Liquid Volume model, but with an
incompatible definition. Because of this, you cannot use these two models together.
Tmin and Tmax define the temperature range where the equation is applicable.
The Campbell-Thodos model is used when RACKET/3 is set to a value less than 0.11. Do not change this parameter unless you intend to use this model.
i j
k ij= A ij + Bij T +C ij T 2
Tc T ci
= ∑ xi
Pc i P ci
Z m =∑ xi Z ¿i , RA
RA
i
¿ , RA 2
Zi =ai +bi T + ci T
V cm =∑ x i V ci
i
T r=T /T c
Note: Reduced temperature Tr is always calculated using absolute temperature units.
Note: Above Tr=0.99 an extrapolation method is used to smooth the transition to constant molar volume equal to the critical volume.
Aspen Polynomial
The equation for the Aspen solids volume polynomial is:
¿ ,l 2 3 4
V i =C 1 i+C 2 i T +C 3i T +C 4 i T +C 5 i T
For C 6 i ≤ T ≤ C7 i
Linear extrapolation of Vi*,l versus T occurs outside of temperature bounds.
IK-CAPE Equation
The IK-CAPE equation is:
¿ ,l 2 3 4 5 6 7 8 9
V i =C 1 i+C 2 i T +C 3i T +C 4 i T +C 5 i T +C 6 i T +C7 i T +C 8 i T +C 9 i T +C10 i T
For C 11i ≤T ≤C 12i
Linear extrapolation of Vi*,l versus T occurs outside of temperature bounds.
DIPPR
The DIPPR equation is:
¿ ,l 2 3 4
ρi =C 1 i+ C2 i T +C 3i T +C 4 i T +C 5 i T for C 6 i ≤ T ≤ C7 i
Linear extrapolation of Vi*,l versus T occurs outside of temperature bounds.
Vi*,s = 1 / ρi*,s
(Other DIPPR equations may sometimes be used. See Pure Component Temperature-Dependent Properties for details.)
nTerms
ρ¿i ,l= ∑ Cmi τ m −1
m=1
Vi*,s = 1 / ρi*,s
The physical and transport properties that HYSYS calculates for a given phase are viscosity, density, thermal conductivity, and surface tension. The models used
for the transport property calculations are all pre-selected to yield the best fit for the system under consideration. For example, the corresponding states model
proposed by Ely and Hanley is used for viscosity predictions of light hydrocarbons (NBP<155), the Twu methodology for heavier hydrocarbons, and a
modification of the Letsou-Stiel method for predicting the liquid viscosities of non-ideal chemical systems.
A complete description of the models used for the prediction of the transport properties can be found in the references listed in each sub-section. All these
models are modified by AspenTech to improve the accuracy of the correlations.
=
In the case of multiphase streams, the transport properties for the mixed phase are meaningless and are reported as <empty>, although the single phase
properties are known. There is an exception with the pipe and heat exchanger operations. For three-phase fluids, HYSYS uses empirical mixing rules to determine
the apparent properties for the combined liquid phases.
Liquid Density
Saturated liquid volumes are obtained using a corresponding states equation developed by R. W. Hankinson and G. H. Thompson7 which explicitly relates the
liquid volume of a pure component to its reduced temperature and a second parameter termed the characteristic volume. This method is adopted as an API
standard. The pure compound parameters needed in the corresponding states liquid density (COSTALD) calculations are taken from the original tables published
by Hankinson and Thompson, and the API Data Book for components contained in HYSYS' library.
The parameters for hypothetical components are based on the API gravity and the generalized Lu equation. Although the COSTALD method was developed for
saturated liquid densities, it can be applied to sub-cooled liquid densities, i.e., at pressures greater than the vapor pressure, using the Chueh and Prausnitz
correction factor for compressed fluids. It is used to predict the density for all systems whose pseudo-reduced temperature is below 1.0. Above this
temperature, the equation of state compressibility factor is used to calculate the liquid density.
Hypocomponents generated in the Oil Characterization Environment have their densities either calculated from internal correlations or generated from input
curves. Given a bulk density, the densities of the hypocomponent are adjusted such that:
1.0
ρbulk =
xi
∑ ρo
i
The characteristic volume for each hypocomponent is calculated using the adjusted densities and the physical properties. The calculated characteristic volumes
are then adjusted such that the bulk density calculated from the COSTALD equation matches the density calculated using the above equation. This ensures that a
given volume of fluid contains the same mass whether it is calculated with the sum of the component densities or the COSTALD equation.
The Rackett model can be used to calculate liquid density. The Rackett method is based on API (American Petroleum Institute) Procedure 6A3.6 and is
recommended for petroleum and hydrocarbon liquid mixtures at low and moderate pressure.
ρ
=
( )
1 RTc
Pc
Z RA
¿¿
Where:
Ρ is the density of the liquid mixture
R is universal gas constant
Tc is the critical temperature
Pc is the critical pressure
ZRA is an empirically derived constant
Tr is the reduced temperature
vapor Density
The density for all vapor systems at a given temperature and pressure is calculated using the compressibility factor given by the equation of state or by the
appropriate vapor phase model for Activity Models.
Viscosity
HYSYS automatically selects the model best suited for predicting the phase viscosities of the system under study. The model selected is from one of the three
available in HYSYS: a modification of the NBS method (Ely and Hanley), Twu's model, or a modification of the Letsou-Stiel correlation. HYSYS selects the
appropriate model using the following criteria:
All of the models are based on corresponding states principles and are modified for more reliable application. Internal validation showed that these models
yielded the most reliable results for the chemical systems shown. Viscosity predictions for light hydrocarbon liquid phases and vapor phases were found to be
handled more reliably by an in-house modification of the original Ely and Hanley model, heavier hydrocarbon liquids were more effectively handled by Twu's
model, and chemical systems were more accurately handled by an in-house modification of the original Letsou-Stiel model.
A complete description of the original corresponding states (NBS) model used for viscosity predictions is presented by Ely and Hanley in their NBS publication.
The original model is modified to eliminate the iterative procedure for calculating the system shape factors. The generalized Leech-Leland shape factor models
are replaced by component specific models. HYSYS constructs a PVT map for each component using the COSTALD for the liquid region. The shape factors are
adjusted such that the PVT map can be reproduced using the reference fluid.
The shape factors for all the library components are already regressed and included in the Pure Component Library. Hypocomponent shape factors are regressed
using estimated viscosities. These viscosity estimations are functions of the hypocomponent Base Properties and Critical Properties.
Hypocomponents generated in the Oil Characterization Environment have the additional ability of having their shape factors regressed to match kinematic or
dynamic viscosity assays.
The general model employs CH4 as a reference fluid and is applicable to the entire range of non-polar fluid mixtures in the hydrocarbon industry. Accuracy for
highly aromatic or naphthenic crudes is increased by supplying viscosity curves when available, since the pure component property generators were developed
for average crude oils. The model also handles H2O and acid gases as well as quantum gases.
Although the modified NBS model handles these systems very well, the Twu method was found to do a better job of predicting the viscosities of heavier
hydrocarbon liquids. The Twu model16 is also based on corresponding states principles, but has implemented a viscosity correlation for n-alkanes as its
reference fluid instead of CH4. A complete description of this model is given in the paper entitled “Internally Consistent Correlation for Predicting Liquid
Viscosities of Petroleum Fractions”21.
For chemical systems the modified NBS model of Ely and Hanley is used for predicting vapor phase viscosities, whereas a modified form of the Letsou-Stiel model
is used for predicting the liquid viscosities. This method is also based on corresponding states principles and was found to perform satisfactorily for the
components tested.
The shape factors contained in the HYSYS Pure Component Library are fit to match experimental viscosity data over a broad operating range. Although this yields
good viscosity predictions as an average over the entire range, improved accuracy over a narrow operating range can be achieved by using the Tabular features.
Liquid Phase Mixing Rules for Viscosity
The estimates of the apparent liquid phase viscosity of immiscible Hydrocarbon Liquid - Aqueous mixtures are calculated using the following "mixing rules":
If the volume fraction of the hydrocarbon phase is greater than or equal to 0.5, the following equation is used:
3.16 (1−v oil )
μeff =μoil e
where:
μeff is apparent viscosity
μoil is viscosity of hydrocarbon phase
Voil is volume fraction Hydrocarbon phase
If the volume fraction of the hydrocarbon phase is less than 0.33, the following equation is used:
where:
[
μeff = 1+2.5 v oil
( μoil+0.4 μ H 2 O
μ oil + μ H 2O )]
μH 2 O
If the volume of the hydrocarbon phase is between 0.33 and 0.5, the effective viscosity for combined liquid phase is calculated using a weighted average
between above two Equations.
Molecular weight:
MW eff =∑ x i MW i
Mixture density:
1
ρ eff =
∑ () xi
ρi
Mixture specific heat:
C P =∑ x i C p
eff i
Thermal Conductivity
As in viscosity predictions, a number of different models and component specific correlations are implemented for prediction of liquid and vapor phase thermal
conductivities. The text by Reid, Prausnitz and Poling was used as a general guideline in determining which model was best suited for each class of components.
For hydrocarbon systems the corresponding states method proposed by Ely and Hanley is generally used. The method requires molecular weight, acentric factor
and ideal heat capacity for each component. These parameters are tabulated for all library components and may either be input or calculated for hypothetical
components. It is recommended that all of these parameters be supplied for non-hydrocarbon hypotheticals to ensure reliable thermal conductivity coefficients
and enthalpy departures.
The modifications to the method are identical to those for the viscosity calculations. Shape factors calculated in the viscosity routines are used directly in the
thermal conductivity equations. The accuracy of the method depends on the consistency of the original PVT map.
Vapor phase:
For vapor phase thermal conductivity predictions, the Misic and Thodos, and Chung et al. methods are used (except for H2O, C1, H2, CO2, NH3 which use a
polynomial for pure components). The effect of higher pressure on thermal conductivities is taken into account by the Chung et al. method.
The vapor phase thermal conductivity is calculated using the following mixing rules:
∑ ( λi MW i )
1
( )
3
λ mix= 1
( )
∑ MW i 3
where:
λmix is the vapor thermal conductivity of the mixture
λi is the vapor thermal conductivity of component i
MWi is molecular weight of component i
Liquid phase:
For liquid phase thermal conductivity predictions, the following methods are used:
The liquid phase thermal conductivity is calculated using the following mixing rule:
(∑ xi λ i )
1 /3 3
λ mix=
i
Where:
λmix is the liquid thermal conductivity of the mixture
λi is the liquid thermal conductivity of component i
xi is mole fraction of component i
As with viscosity, the thermal conductivity for two liquid phases is approximated by using empirical mixing rules for generating a single pseudo liquid phase
property. The thermal conductivity for this pseudo liquid phase is calculated by the following equation:
λ Lmix =∑ ∑ φLi φLj γ Lij
i j
Where:
λLmix is the liquid thermal conductivity of the combined two liquid phases
2
λ Lij=
( 1/ λ Li ) + ( 1/ λ Lj )
λLi is the liquid thermal conductivity of liquid phase i at temperature T
λLj is the liquid thermal conductivity of liquid phase j at temperature T
x Li V Li
φ Li =
∑ x Lk V Lk
k=1
xLi is molar phase fraction of liquid phase i
VLi is molar volume of liquid phase i
xLk is molar phase fraction of liquid phase k
VLk is molar volume of liquid phase k
Surface Tension
Surface tensions for hydrocarbon systems are calculated using a modified form of the Brock and Bird equation17. The equation expresses the surface tension (σ)
as a function of the reduced and critical properties of the component. The basic form of the equation was used to regress parameters for each family of
components.
2/ 3 1/ 3 a
σ =Pc T c Q(1−T r ) ×b
Where:
σ is surface tension (dynes/cm2)
Q=0.1207 1+
( T BR × ln Pc
1.0−T BR )−0.281
For aqueous systems, HYSYS employs a polynomial to predict the surface tension.
Heat Capacity
Heat Capacity is calculated using a rigorous Cv value whenever HYSYS can. The method used is given by the following equations:
( )
2
dV
−T ×
dT
C p−C v =
dV
dP
However, whenever this equation fails to provide an answer, HYSYS falls back to the semi-ideal Cp/Cv method by computing Cp/Cv as Cp/(Cp-R), which is only
approximate and valid for ideal gases. Examples of when HYSYS uses the ideal method are:
HYSYS has the ability to interpret and produce a wide assortment of flow rate data. It can accept several types of flow rate information for stream specifications
as well as report back many different flow rates for streams, their phases and their components. One drawback of the large variety available is that it often leads
to some confusion as to what exactly is being specified or reported, especially when volumetric flow rates are involved.
In the following sections, the available flow rates are listed, each corresponding density basis is explained, and the actual formulation of the flow rate
calculations is presented. For volumetric flow rate data that is not directly accepted as a stream specification, a final section is provided that outlines techniques
to convert your input to mass flow rates.
Many types of flow rates appear in HYSYS output. However, only a subset of these are available for stream specifications.
The flow rate types available through the numerous reporting methods - property views, workbook, PFD, specsheets, and so forth are:
Molar Flow
Mass Flow
Std Ideal Liq Vol Flow
Liq Vol Flow @Std Cond
Actual Volume Flow
Std Gas Flow
Actual Gas Flow
The following flow rate types are available for stream specifications:
Molar Flows
Mass Flows
LiqVol Flows
The volumetric flow rate reference state is defined as 60 °F and 1 atm when using Field units or 15 °C and 1 atm when using SI units.
All calculations for volumetric stream flows are based on density. HYSYS uses the following density basis:
The Standard and Actual liquid densities are calculated rigorously at the appropriate T and P using the internal methods of the chosen property package. Flow
rates based upon these densities automatically take into account any mixing effects exhibited by non-ideal systems. Thus, these volumetric flow rates may be
considered as "real world".
Contrary to the rigorous densities, the Standard Ideal Liquid Mass density of a stream does not take into account any mixing effects due to its simplistic
assumptions. Thus, flow rates that are based upon it do not account for mixing effects and are more empirical in nature. The calculation is as follows:
1
Ideal Density Stream=
x
∑ ρ Ideal
i
i i
Where:
Xi is molar fraction of component i
ρi Ideal is pure component Ideal liquid density
HYSYS contains Ideal Liquid densities for all components in the Pure Component Library. These values are determined in one of three ways, based on the
characteristics of the component, as described below:
Case 1 - For any component that is a liquid at 60 °F and 1 atm, the data base contains the density of the component at 60 °F and 1 atm.
Case 2 - For any component that can be liquified at 60 °F and pressures greater than 1 atm, the data base contains the density of the component at 60 °F and
Saturation Pressure.
Case 3 - For any component that is non-condensable at 60 °F under any pressure, i.e., 60 °F is greater than the critical temperature of the component, the data
base contains GPA tabular values of the equivalent liquid density. These densities were experimentally determined by measuring the displacement of
hydrocarbon liquids by dissolved non-condensable components.
For all hypothetical components, the Standard Liquid density (Liquid Mass Density @Std Conditions) in the Base Properties is used in the Ideal Liquid density ( Std
Ideal Liq Mass Density) calculation. If a density is not supplied, the HYSYS estimated liquid mass density (at standard conditions) is used. Special treatment is
given by the Oil Characterization feature to its hypocomponent such that the ideal density calculated for its streams match the assay, bulk property, and flow
rate data supplied in the Oil Characterization Environment.
The various procedures used to calculate each of the available flow rates are detailed below, based on a known molar flow.
Molar Flow Rate
Molar Mass
Even if a stream is all vapour, it still have a Liq Volume flow, based upon the stream's Standard Ideal Liquid Mass density, whose calculation is detailed in the
previous section.
This volumetric flow rate is calculated using the ideal density of the stream and thus is somewhat empirical in nature.
Total Molar Flow × MW Stream
LiqVolFlow=
StdLiqDensity Stream
This volumetric flow rate is calculated using a rigorous liquid density calculation at the actual stream T and P conditions, and reflects non-ideal mixing effects.
Molar Flow × MW
ActualVolume Flow=
Density
Standard gas flow is based on the molar volume of an ideal gas at standard conditions. It is a direct conversion from the stream's molar flow rate, based on the
following:
This volumetric flow rate is calculated using a rigorous vapor density calculation at the actual stream T and P conditions, and reflects non-ideal mixing and
compressibility effects.
Molar Flow × MW
ActualGas Flow=
Density
Volumetric Flow Rates as Specifications
If you require that the flow rate of your stream be specified based on actual density or standard density as opposed to Standard Ideal Mass Liquid density, you
must use one of the following procedures:
Specified variables can only be re-specified by you or through Recycle, Adjust, or Spread Sheet operations. They do not change through any heat or material
balance calculation.
Rigorous three phase calculations are performed for all equations of state and activity models with the exception of Wilson's equation, which only performs two
phase vapor-liquid calculations. As with the Wilson Equation, Steam property package only supports two phase equilibrium calculations.
HYSYS uses internal intelligence to determine when it can perform a flash calculation on a stream, and then what type of flash calculation needs to be performed
on the stream. This is based completely on the degrees of freedom concept. Once the composition of a stream and two property variables are known, (vapor
fraction, temperature, pressure, enthalpy or entropy) one of which must be either temperature or pressure, the thermodynamic state of the stream is defined.
When HYSYS recognizes that a stream is thermodynamically defined, it performs the correct flash automatically in the background. You never have to instruct
HYSYS to perform a flash calculation.
Property variables can either be specified by you or back-calculated from another unit operation. A specified variable is treated as an independent variable. All
other stream properties are treated as dependent variables and are calculated by HYSYS.
If a flash calculation is performed on a stream, HYSYS knows all the property values of that stream, i.e., thermodynamic, physical and transport properties.
In this manner, HYSYS also recognizes when a stream is overspecified. For example, if you specify three stream properties plus composition, HYSYS prints out a
warning message that an inconsistency exists for that stream. This also applies to streams where an inconsistency is created through HYSYS calculations.
For example, if a stream Temperature and Pressure are specified in a flowsheet, but HYSYS back-calculates a different temperature for that stream as a result of
an enthalpy balance across a unit operation, HYSYS generates an Inconsistency message.
Note: HYSYS automatically performs the appropriate flash calculation when it recognizes that sufficient stream information is known. This information is either
specified by the user or calculated by an operation.
Note: Depending on the known stream information, HYSYS performs one of the following flashes: T-P, T-VF, T-H, T-S, P-VF, P-H, or P-S.
The independent variables for this type of flash calculation are the temperature and pressure of the system, while the dependent variables are the vapor
fraction, enthalpy, and entropy.
With the equations of state and activity models, rigorous calculations are performed to determine the co-existence of immiscible liquid phases and the resulting
component distributions by minimization of the Gibbs free energy term. For vapor pressure models or the semi-empirical methods, the component distribution
is based on the Kerosene solubility data (Figure 9A1.4 of the API Data Book).
If the mixture is single-phase at the specified conditions, the property package calculates the isothermal compressibility (dv/dp) to determine if the fluid behaves
as a liquid or vapor. Fluids in the dense-phase region are assigned the properties of the phase that best represents their current state .
Note: The material solids appear in the liquid phase of two-phase mixtures, and in the heavy (aqueous/slurry) phase of three-phase systems. Therefore, when a
separator is solved using a T-P flash, the vapor phase is identical regardless of whether or not solids are present in the feed to the flash drum.
Note: Use caution in specifying solids with systems that are otherwise all vapor. Small amounts of non-solids may appear in the "liquid" phase.
vapor fraction and either temperature or pressure are the independent variables for this type of calculation. This class of calculation embodies all fixed quality
points including bubble points (vapor pressure) and dew points.
To perform bubble point calculation on a stream of known composition, simply specify the vapor Fraction of the stream as 0.0 and define the temperature or
pressure at which the calculation is desired. For a dew point calculation, simply specify the vapor Fraction of the stream as 1.0 and define the temperature or
pressure at which the dew point calculation is desired. Like the other types of flash calculations, no initial estimates are required.
The vapor fraction is always shown in terms of the total number of moles. For example, the vapor fraction (VF) represents the fraction of vapor in the stream,
while the fraction, (1.0 - VF), represents all other phases in the stream (i.e., a single liquid, 2 liquids, a liquid and a solid).
Dew Points
Given a vapor fraction specification of 1.0 and either temperature or pressure, the property package calculates the other dependent variable (P or T). If
temperature is the second independent variable, HYSYS calculates the dew point pressure. Likewise, if pressure is the independent variable, then the dew point
temperature is calculated. Retrograde dew points may be calculated by specifying a vapor fraction of -1.0. It is important to note that a dew point that is
retrograde with respect to temperature can be normal with respect to pressure and vice versa.
A vapor fraction specification of 0.0 defines a bubble point calculation. Given this specification and either temperature or pressure, the property package
calculates the unknown T or P variable. As with the dew point calculation, if the temperature is known, HYSYS calculates the bubble point pressure and
conversely, given the pressure, HYSYS calculates the bubble point temperature. For example, by fixing the temperature at 100 °F, the resulting bubble point
pressure is the true vapor pressure at 100 °F.
Bubble and dew points are special cases of quality point calculations. Temperatures or pressures can be calculated for any vapor quality between 0.0 and 1.0 by
specifying the desired vapor fraction and the corresponding independent variable. If HYSYS displays an error when calculating vapor fraction, then this means
that the specified vapor fraction doesn't exist under the given conditions, i.e., the specified pressure is above the cricondenbar, or the given temperature lies to
the right of the cricondentherm on a standard P-T envelope.
Note: HYSYS calculates the retrograde condition for the specified vapor quality if the vapor fraction is input as a negative number.
Enthalpy Flash
Given the enthalpy and either the temperature or pressure of a stream, the property package calculates the unknown dependent variables. Although the
enthalpy of a stream cannot be specified directly, it often occurs as the second property variable as a result of energy balances around unit operations such as
valves, heat exchangers and mixers.
Note: If a specified amount of energy is to be added to a stream, this may be accomplished by specifying the energy stream into either a Cooler/Heater or
Balance operation.
If HYSYS responds with an error message, and cannot find the specified property (temperature or pressure), this probably means that an internally set
temperature or pressure bound was encountered. Since these bounds are set at quite large values, there is generally some erroneous input that is directly or
indirectly causing the problem, such as an impossible heat exchange.
Entropy Flash
Given the entropy and either the temperature or pressure of a stream, the property package calculates the unknown dependent variables.
Electrolyte Flash
The electrolyte stream flash differs from the HYSYS material stream flash to handle the complexities of speciation for aqueous electrolyte systems.
The HYSYS OLI Interface package is an interface to the OLI Engine (OLI Systems) that enables simulations within HYSYS using the full functionality and capabilities
of the OLI Engine for flowsheet simulation.
When the OLI_Electrolyte property package is associated with material streams, the streams exclusively become electrolyte material streams in the flowsheet.
That is, the stream conducts a simultaneous phase and reaction equilibrium flash. For the model used and the reactions involved in the flash calculation, refer to
the HYSYS OLI Interface Reference Guide.
An electrolyte material stream in HYSYS can perform the following type of flashes:
TP Flash
PH Flash
TH Flash
PV Flash
TV Flash
Due to the involvement of reactions in the stream flash, the equilibrium stream flash may result in a different molar flow and composition from the specified
value. Therefore, mass and energy are conserved for an electrolyte material stream against the HYSYS stream for mass, molar and energy balances.
Limitations exist in the HYSYS OLI Interface package in the calculation of the stream flash results. The calculation for the electrolyte flash results must fall within
the following physical ranges to be valid.
Handling of Water
Water is handled differently depending on the correlation being used. The PR and PRSV equations are enhanced to handle H2O rigorously whereas the semi-
empirical and vapor pressure models treat H2O as a separate phase using steam table correlations.
In these correlations, H2O is assumed to form an ideal, partially-miscible mixture with the hydrocarbons and its K value is computed from the relationship:
Po
Kw=
(x s P)
Where
P° is vapor pressure of H20 from Steam Table
P is system pressure
Xs is solubility of H2O in hydrocarbon liquid at saturation conditions
The value for xs is estimated by using the solubility data for kerosene as shown in Figure 9A1.4 of the API Data Book2. This approach is generally adequate when
working with heavy hydrocarbon systems. However, it is not recommended for gas systems.
For three phase systems, only the PR and PRSV property package and Activity Models allow components other than H2O in the second liquid phase. Special
considerations are given when dealing with the solubilities of glycols and CH3OH. For acid gas systems, a temperature dependent interaction parameter was
used to match the solubility of the acid component in the water phase.
The PR equation considers the solubility of hydrocarbons in H2O, but this value may be somewhat low. The reason for this is that a significantly different
interaction parameter must be supplied for cubic equations of state to match the composition of hydrocarbons in the water phase as opposed to the H2O
composition in the hydrocarbon phase. For the PR equation of state, the latter case was assumed more critical. The second binary interaction parameter in the
PRSV equation allows for an improved solubility prediction in the alternate phase.
With the activity coefficient models, the limited mutual solubility of H2O and hydrocarbons in each phase can be taken into account by implementing the
insolubility option (please refer to Section A.3.2 - Activity Models). HYSYS generates, upon request, interaction parameters for each activity model (with the
exception of the Wilson equation) that are fitted to match the solubility of H2O in the liquid hydrocarbon phase and hydrocarbons in the aqueous phase based
on the solubility data referred to in that section.
The Peng-Robinson and SRK property packages will always force the water rich phase into the heavy liquid phase of a three phase stream. As such, the aqueous
phase is always forced out of the bottom of a three phase separator, even if a light liquid phase (hydrocarbon rich) does not exist. Solids are always carried in the
second liquid phase.
Supercritical Handling
HYSYS reports a vapor fraction of zero or one, for a stream under supercritical conditions. Theoretically, this value doesn’t have any physical meaning for a
supercritcial fluid, since there is no distinction of liquid or vapor phases in a supercritical region. However, it is important to determine if a supercritical fluid is
liquid-like or a vapor-like fluid. This is because some of the properties reported in HYSYS are calculated using certain sets of specific phase models. In other
words, phase identification has to be carried out in order to decide which model to use to calculate these properties.
In HYSYS, all flash results go through a phase order function to identify the phase type. Different packages have their own different order.
For example, the following criteria are used to identify phase types for the PR, SRK, SourPR, and Sour SRK cubic equations of state at supercritical region:
1.If the compressibility factor (Z) is greater than 0.3, and the isothermal compressibility factor (beta) is greater than 0.75/P (P in kPa), a vapor fraction of 1.0 is
assigned to the stream.
2.If Z is greater than 0.75 and the sum of composition of light compounds (NBP<230K) is greater than the sum of composition of heavy compounds, a vapor
fraction of 1.0 is assigned to the stream.
Otherwise, vapor fraction of 0 is assigned to the stream and liquid correlations are used.
Solids
HYSYS does not check for solid phase formation of pure components within the flash calculations, however, incipient solid formation conditions for CO2 and
hydrates can be predicted with the Utility Package.
Solid materials such as catalyst or coke can be handled as user-defined, solid type components. The HYSYS property package takes this type of component into
account in the calculation of the following stream variables: stream total flow rate and composition (molar, mass and volume), vapor fraction, entropy, enthalpy,
specific heat, density, molecular weight, compressibility factor, and the various critical properties. Transport properties are computed on a solids-free basis. Note
that solids are always carried in the second liquid phase, i.e., the water rich phase.
Solids do not participate in vapor-liquid equilibrium (VLE) calculations. Their vapor pressure is taken as zero. However, since solids do have an enthalpy
contribution, they have an effect on heat balance calculations. Thus, while the results of a Temperature flash are the same whether or not such components are
present, an Enthalpy flash is affected by the presence of solids.
Stream Information
When a flash calculation occurs for a stream, the information that is returned depends on the phases present within the stream. The following table shows the
stream properties that are calculated for each phase:
Notes:
B - Physical property queries are allowed on the feed phase of single phase streams.
C - Physical property queries are allowed on the feed phase only for streams containing vapor and/or liquid phases.
D - Physical property queries are allowed on the feed phase of liquid streams with more than one liquid phase.
See Also
Calculating Greenhouse Gas Emissions in Aspen HYSYS
Aspen HYSYS Greenhouse Gas (GHG) Emissions Calculations let process engineers estimate the Greenhouse Gas Emissions associated with a process. Two
sources of GHG emissions are considered:
In addition, engineers can use GHG calculations to check the capture efficiency of various chemical and physical solvents in term of “carbon loading.” Aspen
HYSYS includes a number of "CO2 Capture" sample models to support the simulation of acid gas scrubbing and CO2 capture systems based on amines and
physical solvents.
Greenhouse gas emissions are reported in terms of carbon equivalents, written as “CO2e”. This is a measure of the total global warming impact of volatile
emissions over a given time span, most commonly 100 years. The CO2e is the sum of the product of the mass flow rate of a given emission and the “GWP”
(Global Warming Potential). The GWP is determined by evaluating the total warming resulting from a given gas over a fixed period, relative to the amount of
warming caused by carbon dioxide. This depends on the radiative properties of the gas, as well as the reaction products of the gas as it decomposes in the upper
atmosphere. The GWP values are estimates which continue to be refined over time, however there are a few well accepted sources including the
Intergovernmental Panel on Climate Change (IPCC), the European Union, and the US EPA.
There are many known greenhouse gasses, including CO2, methane, nitrous oxide, sulfur hexafluoride, and a wide variety of chlorofluorocarbons. The tables
below list the factors for various gasses.
Note: The calculations are for conceptual design purposes only and are not intended to meet the full reporting criteria of any governing body.
HYSYS uses three stream property correlations corresponding to the CO2e based on the well-known 2nd and 4th reports of the IFPP.
Less commonly, the CO2 loading may be reported on a concentration basis, for example apparent moles of CO2 per volume or mass of solvent.
Aspen HYSYS includes "CO2 Capture" sample models to support simulation of acid gas scrubbing and CO2 capture systems based on amines and physical
solvents. Use the CO2 Loading property to report the ratio of the sum of apparent molar flow rates for the specified components to the sum of apparent molar
flow rates for the base components, for example, the moles of CO2 captured in a mixture of amine solvents.
These correlations may be added as stream property correlations under the Standard level:
Name Description
CO2 Loading (Set parameter: Mole or Mass)
Molar loading is the apparent molar ratio of a component or group of components to a solvent or group of solvents
(dimensionless molar ratio).
Mass loading is the apparent mass ratio of a component or group of components to a solvent or group of solvents
(dimensionless).
Set choice as a parameter in the Correlation Manager.
CO2 apparent Wt. Conc. Apparent molar concentration (apparent moles / volume of solvent).
CO2 apparent Mole Conc. Apparent weight concentration (apparent moles / mass of solvent).
API Publication 955, A New Correlation of NH3, CO2 and H2S Volatility Data From Aqueous Sour Water Systems, March 1978.
API Technical Data Book, Petroleum Refining, Fig. 9A1.4, p. 9-15, 5th Edition (1978).
Chao, K. D. and Seader, J. D., A.I.Ch.E. Journal, pp. 598-605, December 1961.
Ely, J.F. and Hanley, H.J.M., "A Computer Program for the Prediction of Viscosity and Thermal Conductivity in Hydrocarbon Mixtures", NBS Technical Note 1039.
Gambill, W.R., Chem. Eng., March 9, 1959.
Grayson, H. G. and Streed, G. W., "vapor-Liquid Equilibria for High Temperature, High Pressure Systems", 6th World Petroleum Congress, West Germany, June
1963.
Hankinson, R.W. and Thompson, G.H., A.I.Ch.E. Journal, 25, No. 4, p. 653 (1979).
Hayden, J.G. and O’Connell, J.P., Ind. Eng. Chem., Process Des. Dev. 14, 209 (1975).
Jacobsen, R.T and Stewart, R.B., 1973. "Thermodynamic Properties of Nitrogen Including Liquid and vapor Phases from 63 K to 2000K with Pressure to 10 000
Bar." J. Phys. Chem. Reference Data, 2: 757-790.
Kabadi, V.N., and Danner, R.P. A Modified Soave-Redlich-Kwong Equation of State for Water-Hydrocarbon Phase Equilibria, Ind. Eng. Chem. Process Des. Dev.
1985, Volume 24, No. 3, pp 537-541.
Keenan, J. H. and Keyes, F. G., Thermodynamic Properties of Steam, Wiley and Sons (1959).
Knapp, H., et al., "Vapor-Liquid Equilibria for Mixtures of Low Boiling Substances", Chemistry Data Series Vol. VI, DECHEMA, 1989.
Passut, C. A.; Danner, R. P., “Development of a Four-Parameter Corresponding States Method: vapor Pressure Prediction”, Thermodynamics - Data and
Correlations, AIChE Symposium Series; p. 30-36, No. 140, Vol. 70.
Peng, D. Y. and Robinson, D. B., "A Two Constant Equation of State", I.E.C. Fundamentals, 15, pp. 59-64 (1976).
Perry, R. H.; Green, D. W.; “Perry’s Chemical Engineers’ Handbook Sixth Edition”, McGraw-Hill Inc., (1984).
Prausnitz, J.M., Lichtenthaler, R.N., Azevedo, E.G., "Molecular Thermodynamics of Fluid Phase Equilibria", 2nd. Ed., McGraw-Hill, Inc. 1986.
Reid, C.R., Prausnitz, J.M., and Sherwood, T.K., "The Properties of Gases and Liquids", McGraw-Hill Book Company, 1977.
Reid, R.C., Prausnitz, J.M., and Poling, B.E., "The Properties of Gases & Liquids", McGraw-Hill, Inc., 1987.
Soave, G., Chem Engr. Sci., 27, No. 6, p. 1197 (1972).
Stryjek, R., Vera, J.H., J. Can. Chem. Eng., 64, p. 334, April 1986.
Twu, C.H., I.E.C. Proc Des & Dev, 24, p. 1287 (1985).
Woelflin, W., "Viscosity of Crude-Oil Emulsions", presented at the spring meeting, Pacific Coast District, Division of Production, Los Angeles, Calif., Mar. 10, 1942.
Zudkevitch, D., Joffee, J. "Correlation and Prediction of Vapor-Liquid Equilibria with the Redlich-Kwong Equation of State", AIChE Journal, Volume 16, No. 1,
January pp. 112-119.
SELF NOTES:
PV Pv
Z= =
nRT RT
Where:
P is the system pressure
T is the system pressure
N is number of moles
V is the system volume
v is the molar volume
R is universal gas constant
R = 8.31446261815324 m³Pa/(K.mol) or J/(K.mol) or L.kPa/(K.mol)
= 10.731577089016 psi.ft³/(lbmol.°R)
At standard conditions, Z = 1
For ideal gas (Z=1): Z=PsV/nRTs
COSTALD method (of Hankinson & Thomson - 1979) (Preferred method of calculation)
Saturation density could be calculated by COSTALD method
If 0.25 < Tr=T/Tc,m < 0.95
( ( )( ))
N N 2 N 1
1
¿
V =
4
∑ x i V ¿i +3 ∑ xi V ¿i 3 ∑ x i V ¿i 3
i =1 i=1 i=1
∑ ∑ ( xi x j √V ¿ i T ci V ¿ j T cj )
T c ,m= i=1 j=1
¿
V m
or
(∑ )
N 2
xi √ T c, i V
¿
i
i=1
T c ,m=
V ¿m
¿ (0) (δ)
V s =V ∙V R (1−ω SRK ,m ∙V R )
where:
V* is a characteristic volume analogous to the critical volume, usually regressed by fitting the equations to experimental density data, using known values of Tc
and ωSRK, but it can be calculated from a single density point or estimated from a generalized correlation of the form
RTs
V ¿= (k 1 +k 2 ω SRK +k 3 ω 2SRK )
Ps
1 2 4
(0) 3 3 3
V =1−1.52816 (1−T r ) +1.43907 (1−T r ) −0.81446 ( 1−T r ) + 0.190454(1−T r )
R
is the spherical molecule function,
2 3
−0.296123+0.386914 T r −0.0427258 T r −0.0480645T r
(δ)
V =
R
T r−1.00001
is the function, when multiplying by VR(0), which gives the deviation function,
N
ω SRK ,m =∑ x i ω SRK ,i
i=1
Is the Phillips acentric factor based on the Soave-Redlich-Kwong equation of state.
Compressed liquid density
COSTALD method
If 0.95 > Tr >= 1, there is a linear interpolation between the liquid density at Tr =0.95 and the vapor density at Tr = 1,05. This model can be used to calculate
saturated and compressed molar volume, which is calculated using Tait-like equation:
Molar volume Vm is calculated as below:
(
V =V s 1−C ∙ ln
( B+B+PP )) s
where
[ ]
1 2 1
B=P s −1−9.070217(1−T r ) 3 +62.45326(1−T r ) 3 −135.1102 ( 1−T r ) + e1 (1−T r ) 3
e 1=exp ( 4.79594+ 0.250047 ωSRK +1.14188 ω 2
SRK )
C=0.0861488+0.0344483 ω SRK
In order to use equation Vs above, a knowledge of both the saturated molar volume Vs and the saturation pressure Ps at the system temperature is required, Vs
is determined by the procedures given above. The saturation pressure Ps, or Psm for bubble point mixtures*, may be calculated from the following generalized
Riedel vapor pressure equation developed by Hankinson, Estes, and Coker:
log ( P Rs )=P(0R )+ ω SRK P(1R )
P(0R )=5.8031817 ∙ log ( T R ) +0.07608141 F
P(1)
R =4.86601 G
36.0 6
F=35.0− −96.376∙ log ( T R )+ T R
TR
G=log ( T R ) +0.03721754 F
Ps =( P Rs Pc )
Alternatively,
C2
ln Ps =C1 + +C3 T +C 4 ln T
T
* Note *: The subscript m is added to denote a mixture property when the equation in this section are applied to mixture, e.g. Tcm, TRm, Pcm, PRm, Vsm, Vm,
etc.
In addition to the mixing rules, a knowledge of the mixture pseudocritical pressure Pcm is needed when Equations above for parameter B are being applied to
compressed liquid mixture. Thompson, et al recommended the following equations:
Z cm R T cm
Pcm = ¿
Vm
Z cm=0.291−0.080 ω SRK ,m
The accuracy of determining the density of mixtures, such as LNG, depends upon accurate compositional analyses, since the mole fractions of all components, xi,
are needed both in the COSTALD mixing rules and to calculate the mixture molecular weight Mm:
Mm
ρm =
Vm
M m= ∑ x i M i
i
The composition of liquid mixtures can be determined from chromatographic analysis, which yields accurate value of component weight fractions. Weigh
fractions are easily converted to mole fraction for insertion into the above equation.
In terms of accuracy and usefulness, COSTALD correlation for compressed liquid density is better than other correlations (such as Hudleston (1937), Lyckman, et
al (1965, 1969), Yen-Woods (1966), Generalized Lu equation extended by Ewbank & Harden (1967), McCarty-Klosek-McKinley (1980), Tait (1988)…)