You are on page 1of 48

By: Shaho Abdul qadir

INTRODUCTION
Natural gas is a mixture of hydrocarbon gases and impurities. The hydrocarbon gases normally found in natural gas are methane, ethane, propane, butanes, pentanes, and small amounts of hexanes, heptanes, octanes, and the heavier gases. The impurities found in natural gas include carbon dioxide, hydrogen sulfide, nitrogen, water vapor, and heavier hydrocarbons. Usually, the propane and heavier hydrocarbon fractions are removed for additional processing because of their high market value as gasoline-blending stock and chemicalplant raw feedstock. What usually reaches the transmission line for sale as natural gas is mostly a mixture of methane and ethane with some small percentage of propane. This chapter reviews those physical properties of natural gases that are important in solving gas well performance, gas production, and gas transmission problems. The properties of a natural gas may be determined either directly from laboratory tests or predictions from known chemical composition of the gas. In the latter case, the calculations are based on the physical properties of individual components of the gas and on physical laws, often referred to as mixing rules, relating the properties of the components to those of the mixture. COMPOSITION OF NATURAL CAS There is no one composition or mixture that can be referred to as the natural gas. Each gas stream produced has its own composition. Two wells from the same

18

reservoir may have different compositions. Also, each gas stream produced from a natural gas reservoir can change composition as the reservoir is depleted. Samples of the well stream should be analyzed periodically, since it may be necessary to change the production equipment to satisfy the new gas composition. Table 2.1 shows some typical natural gas streams. Well stream 1 is typical of an associated gas, that is, gas produced with crude oil. Well streams 2 and 3 are typical nonassociated low-pressure and high-pressure gases, respectively. Natural gas is normally regarded as a mixture of straight-chain or paraffin hydrocarbon gases. In straight-chain hydrocarbons, the carbon atoms are attached to form chains. However, cyclic and aromatic hydrocarbon gases are occasionally found in natural gas mixtures. In cyclic hydrocarbons, the carbon atoms are arranged to form rings. Figure 2.1 shows the structures of some straight-chain and cyclic hydrocarbons. The hydrocarbons listed in Table 2.1 are straight-chain or paraffinic compounds. PHASE BEHAVIOR Conventional gas reservoirs have been characterized in many different ways but most commonly on the basis of the surface-producing gas-oil ratio. Using this method, any well (or field) that produces at a gas-oil ratio (GOR) in excess of 100,000 cu ft per barrel of oil [standard cubic feet per stock tank barrel (scf/STB)] is considered a gas well; one producing with a GOR of 5000 to 100,000 scf/STB, a gas-condensate well; and one producing with a GOR of zero to several thousand scf/STB, an oil well. In practice, similar surface gas-oil ratios have been obtained for reservoirs containing a variety of hydrocarbon fluid compositions, existing over a wide range of reservoir pressures and temperatures, and producing with natural or artificial mechanisms. This has resulted in both technical and legal misunderstanding of the nature of conventional gas reservoirs. Therefore, the simplified classification described above is considered inadequate. Conventional gas reservoirs should be defined on the basis of their initial reservoir pressure and temperature on the usual pressure-temperature (P-T)

19

phase diagram (Fig. 2.2). P-T phase diagrams show the effects of pressure and temperature on the physical state of a hydrocarbon system. However, the phase diagram in Fig. 2.2 is for a specific composition. Although a different fluid would have a different phase diagram, the general configuration is similar. In Fig. 2.2, the area enclosed by the bubble point (BP) line A-S-C and the dew point (DP) line C-D-T-B to the lower left is the region of pressuretempera-

20

Component

Methane Ethane Propane /-butane n-butane i-pentane n-pentane Hexane Heptanes and heavier Carbon dioxide Hydrogen sulfide Nitrogen Total

TABLE 2.1 Typical Natural Gas Analyses Well No. 1 Well No. Well 2 No. 3 Mole Mole Mole Percent Percent Percen t 27.52 71.01 91.2 5 16.34 13.09 3.61 29.18 7.91 1.37 5.37 1.68 0.31 17.18 2.09 0.44 2.18 1.17 0.16 1.72 1.22 0.17 0.47 1.02 0.27

0.04 0.00 0.00 0.00 100.00

0.81 0.00 0.00 0.00 100.00

2.42 0.00 0.00

0.00 100. 00 Source: Courtesy Petroleum Extension Service. Note: Production from many wells will contain small quantities of carbon dioxide, hydrogen sulfide, and nitrogen.

ture combinations in which both gas and liquid phases will exist. The curves within the two-phase region show the gas-liquid percentages for any

21

temperature and pressure. The line A-S-C-T-B separates the two-phase region from the single- phase regions where all the fluids exist in a single phase. The bubble point line A-S-C separates the two phase region from the single-phase liquid region, while the dew point line C-D-T-B separates it from the singlephase gas region. Point C where the bubble point and dew point lines meet is called the critical point, and the corresponding temperature is called critical temperature Tc . Consider a reservoir initially at 3000 psia and 125F represented on the figure by point 1,. (The subscripts i and a depict the initial and abandonment conditions.) The pressure and temperature conditions are such that the initial state of the hydrocarbons is a liquid, that is, oil. Thus, point 1, delineates an oil reservoir. This type is called a bubble point reservoir, for as pressure declines in the reservoir (due to production) isothermally, the bubble point will be reached, in this case at 2550 psia, point S on the dashed line. Bubble point refers to the highest pressure at which the first bubble gas comes out of solution from the oil and exists as free gas in the reservoir. It is also known as the saturation pressure. Below this pressure, a free-gas phase will appear. Eventually the free gas evolved begins to flow to the well bore in ever-increasing quantities. Conversely, the oil flows in ever-decreasing quantities. The actual amount of free gas liberated will depend on the composition of the oil. Oils originally containing large amounts of the lighter hydrocarbons [high (API) gravity] will release large amounts of gas, while those oils originally containing only small amounts of the lighter hydrocarbons (low API gravity) will liberate much smaller amounts of gas.

22

Fig. 2 J Hydrocarbon gas molecule structures. (Courtesy Petroleum Extension Service.)

Other names for this type of oil reservoir are undersaturated, depletion, dissolved gas, solution gas drive, expansion, and internal gas drive.
If the same hydrocarbon mixture occurred at 2000 psia and 210F, point 2 in the figure, it would be an oil reservoir with an initial gas cap. Both phases exist in equilibrium. The slightest reduction in pressure causes liberation of gas from oil, making this a saturated oil reservoir. The initial pressure of the reservoir and the saturation pressure of the oil zone will be identical. Hence, if the initial reservoir pressure equals the reservoir saturation pressure, the reservoir has an initial gas cap. The oil zone will be produced as a bubble point reservoir, modified by the presence of the gas cap. The gas cap will be at the

23

dew point and may be either retrograde (discussed later) or nonretrograde (Fig. 2.3). Next, consider a reservoir at a temperature of 230F and with an initial pressure of 3300 psia, shown as point 3,- in Fig. 2.2. Since the initial conditions of pressure and temperature are to the right of the critical point and outside the phase envelope, this reservoir exists initially in the gaseous state. As production

Fig. 2J2 Pressure-temperature phase diagram of a reservoir fluid.

Fig. 23 Phase diagrams of a cap gas and oil zone fluid showing (a) retrograde cap gas and (b) nonretrograde cap gas. (After Craft and

Hawkins.)

24

begins from the reservoir and pressure declines, no change in the state of the reservoir fluids occurs until the dew point pressure is reached at 2700 psia, point D. Below this pressure a liquid condenses out of the reservoir fluid as a fog or dew. This is not considered to be a normal situation since, for most hydrocarbon fluids, a pressure reduction tends to increase the amount of gas. Therefore, this behavior is usually referred to as retrograde condensation, signifying that vaporization generally occurs during isothermal expansion rather than condensation. The condensation leaves the gas phase with a lower liquid content. As the condensed liquid adheres to the walls of the pore spaces of the rock, it is immobile. Thus, the gas produced at the surface will have a lower liquid content, and the producing GOR will rise. As the liquefiable portions of the reservoir fluids are usually the most valuable components, the loss of part of these fluids could substantially reduce the ultimate income from the property, which must be considered in an economic evaluation. Examination of Fig. 2.2 will show that for a reservoir fluid to exhibit the phenomenon of retrograde condensation, the initial conditions of pressure and temperature must exist outside the phase envelope to the right of the critical point C and to the left of point Tot within the phase envelope in the region marked X. The point T is called the cricondentherm and is the maximum temperature at which two phases can exist in equilibrium (300F for the example). The process of retrograde condensation continues until a point of maximum liquid volume is reached, 10% at 2250 psia (point E). In some cases, a sufficient volume of liquid will be condensed in the reservoir to provide mobility of the liquid phase. In such cases the surface fluid composition depends on the relative mobilities of the vapor and liquid in the reservoir. As production continues from point E to the abandonment pressure 3a, vaporization of the retrograde liquid occurs. This revaporization aids liquid recovery and may be evidenced by decreasing GOR on the surface. This example assumes that the reservoir fluid composition remains constant. Unfortunately, as retrograde condensation occurs, the reservoir fluid composition changes and the PT envelope shifts, increasing retrograde liquid condensation. Generally, for a particular initial hydrocarbon fluid, retrograde

25

loss increases at lower reservoir temperature, higher abandonment pressure, and for greater shifting of the phase envelope to the right. As a final illustration, consider a reservoir initially at 350F and 3600 psia, represented by point 4, in Fig. 2.2. Since the initial reservoir conditions exist to the right of the critical point C and outside the phase envelope, the reservoir fluid will be 100% gas. Furthermore, since the reservoir temperature exceeds the cricondentherm T, at no point in the isothermal depletion cycle (along path 4,--4a) is the phase envelope crossed. Therefore, the fluid in the reservoir never changes composition; it is always in the gaseous state. However, after the reservoir fluid leaves the reservoir and enters the wellbore, the temperature, as well as the pressure, will decline until surface temperature and pressure conditions are reached. The fluid produced through the well- bore and into surface separators at point 4S, though of the same composition as that in the reservoir, has entered the two-phase region due to the temperature and pressure decline along line 4,4S. This accounts for the production of a considerable amount of liquid at the surface from a gas in the reservoir; therefore, this reservoir is referred to as a wet gas reservoir. If the phase diagram is such that the separator conditions, point 4 S , lies outside the two-phase envelope in the single-phase (gas) region, then only gas will exist on the surface. No liquid will be formed in the reservoir or at the surface and the gas is called dry natural gas. The word dry indicates that the liquid does not contain enough of the heavier hydrocarbons to form a liquid at surface conditions. Nevertheless, it may contain liquid fractions, which can be removed by low- temperature separation or by natural gasoline plants.

As a starting point in the study of the properties of real gases, we consider a hypothetical fluid known as an ideal gas. An ideal gas is a fluid in which the

26

volume occupied by the molecules is insignificant with respect to the volume occupied by the total fluid, there are no attractive or repulsive forces between the molecules or between the molecules and the walls of the container, and all collisions of molecules are perfectly elastic, that is, there is no loss in internal energy upon collision. At low pressures, most gases behave like the ideal gas. In addition, under normal distribution pressures, natural gas follows the ideal gas laws quite closely. Under these conditions it is not normally necessary, therefore, that an accurate determination of any deviation from these laws be made. However, when gas pressures increase, a wide variation between the actual and ideal volumes of the gas may occur. To understand fully what happens when natural gas is subjected to changes in pressure and temperature, the fundamental gas laws must be reviewed. The nomenclature is as follows:

Boyle's Law
Robert Boyle (1627-1691), during the course of experiments with air, observed the following relation between pressure and volume: if the temperature of a given quantity of gas is held constant, the volume of gas varies inversely with the absolute pressure. This relation, written as an equation, is

In the application of Boyle's law, volume at a second set of pressure conditions is generally desired. A rearrangement of Eq. 2.1 gives the formula more readily used:

27

Example 2.1. A quantity of gas at a pressure of 50 psig has a volume of 1000 cu ft. If the gas is compressed to 100 psig, what volume would it occupy? Assume the barometric pressure is 14.73 psia and the temperature of the gas remains constant. Solution

Substituting in Eq. 2.2 would give

Charles' Law
About 100 years after the discovery of Boyle's law, Jacques A. Charles (1746 1823) and Joseph L. Gay-Lussac (1778-1850) independently discovered the law that is usually called Charles' law. This law is in two parts: 1. If the pressure on a particular quantity of gas is held constant, then, with any change of state, the volume will vary directly as the absolute temperature. Expressed as an equation,

Again, since the volume at a second set of temperature conditions is desired usually more than any other information, a handy arrangement of Eq. 2.3 is given as Eq. 2.4:

2. If the volume of a particular quantity of gas is held constant, then, with any change of state, the absolute pressure will vary directly as the absolute temperature:

28

In this instance, the pressure at a second temperature condition would be of more interest. Equation 2.5 could thus be written as

Example 2.2. (a) A given mass of gas has a volume of 500 cu ft when the temperature is 50F and the pressure is 10 psig. If the pressure remains the same, but the temperature is changed to 100F, what will be the volume of the gas? Atmospheric pressure may be taken as 14.73 psia. (b) What would be the new pressure of the gas in the above example if the volume remains the same and the temperature is increased from 50 to 100F as indicated?

(c)

29

This equation is known as Boyle's-Charles' law and as the simple gas law. It is one of the most widely used relations in gas measurement work, since it approximately represents the behavior of many gases under conditions close to ordinary atmospheric temperatures and pressures. One can substitute known values in the combined formula and solve for any one unknown value. In cases where one of the parameters, such as temperature, is not to be considered, it may be treated as having the same value on both sides of the formula. However, Eq. 2.7 is deficient in one important respect: It does not show the relations connecting volumes and masses of gases. Example 2.3. (a) How many cubic feet of an ideal gas, measured at standard conditions of 60F and 14,73 psia, are required to fill a 100-cu ft tank to a pressure of 40 psia when the temperature of the gas in the tank is 90F? Atmospheric pressure is 14.4 psia. (b) What would be the reading on the pressure gauge if the tank in the above example is cooled to 60F after being filled with the ideal gas?

A v o g a d r o ' s L a w

30

Avogadro's law

From Avogadro's law, it may be seen that the weight of a given volume of gas is a function of the weights of the molecules, and that there is some volume at which the gas would weigh, in pounds, the numerical value of its molecular weight. The volume at which the weight of the gas in pounds is equal to the numerical value of its molecular weight is known as the mole-volume. A pound- mole of an ideal gas occupies 378.6 cu ft at 60F and 14.73 psia. These conditions of temperature and pressure are commonly referred to as standard conditions.

The Ideal Gas Law


The equation of state for an ideal gas may be derived from a combination of Boyle's, Charles'/Gay Lussac's, and Avogadro's laws as

where p = absolute pressure, psia V = volume, cu ft T = absolute temperature, R n = number of pound-moles,where 1 lb-mol is the molecular weight of the gas (lb) R = the universal gas constant that, for the above units, has the value 10.732 psia cu ft/lb-mol R Equation 2.8 is only applicable at pressures close to atmospheric, for which it was experimentally derived, and at which gases behave as ideal. Since the number of pound-moles of a gas is equal to the mass of the gas divided by the molecular weight of the gas, the ideal gas law can be expressed as

31

where m = mass of gas, lb M = molecular weight of gas lbm/lb-mol Equation 2.9 may be rearranged to give the mass and density, p, of the gas:

and

32

Example 2.4. Using the fact that 1 pound-mole of an ideal gas occupies 378.6 scf, calculate the value of the universal gas constant, R. Solution p = 14.73 psia V = 378.6 scf n = 1 T = 520 R Using Eq. 2.8,

The numerical value of R depends on the units used to express temperature, pressure, and volume. Table 2.2 gives numerical values of R for various systems of units. Example 2.5. Repeat Example 2.3 using the ideal gas law, Eq. 2.8.

33

PROPERTIES OF GASEOUS MIXTURES Natural gas engineers invariably deal with gas mixtures and rarely with single- component gases. Since natural gas is a mixture of hydrocarbon compounds and because this mixture is varied in types as well as the relative amounts of the compound, the overall physical properties will vary. The overall physical properties of a natural gas determine the behavior of the gas under various processing conditions. If the composition of the gas mixture is known, the overall physical properties can be established from the physical properties of each pure component in the mixture using Kay's mixing rules. Physical properties that are most useful in natural gas processing are molecular weight, boiling point, freezing point, density, critical temperature, critical pressure, heat of vaporization, and specific heat. Table A.l is a tabulation of physical constant of a number of hydrocarbon compounds, other chemicals, and some common gases, taken from GPSA Engineering Data Book. Table 2.3 gives additional physical constants for the paraffin hydrocarbons methane through n-decane, including isobutane and isopentane.

34

35

Composition The composition of a natural gas mixture may be expressed as either the mole fraction, volume fraction, or weight fraction of its components. These may also be expressed as mole percent, volume percent, or weight percent by multiplying the fractional values by 100. Volume fraction is based on gas component volumes measured at standard conditions, so that volume fraction is equivalent to mole fraction. The mole fraction, yi is defined as

where

Volume fraction is defined as

where

Weight fraction to,, is defined as

where

36

It is easy to convert from mole fraction (or volume fraction) to weight fraction and vice versa. Apparent Molecular Weight Although, in a strict sense, a gas mixture does not have a unique molecular weight, it behaves as though it does. Thus, the concept of apparent or average molecular weight is quite useful in characterizing a gas mixture. The apparent molecular weight of a gas mixture is a pseudo property of the mixture and is defined as

where Ma = apparent molecular weight of mixture yi = mole fraction of component i Mi = molecular weight of component i The gas laws can be applied to gas mixtures by simply using apparent molecular weight instead of the single-component molecular weight in the formulas.

37

38

Therefore, the apparent molecular weight of the mixture is 17.08 lbm/lb-mol. Similarly, the apparent molecular weight of the gas mixture in Example 2.6 is 16.82 lbm/lb-mol.

BEHAVIOR OF REAL CASES


The ideal gas describes the behavior of most gases at pressure and temperature conditions close to atmospheric. Most natural gas engineers and operating personnel, at one time or another, become involved with the erratic behavior of natural gas under pressure. At moderate pressures, the gas tends to compress more than the ideal gas law indicates, particularly for temperatures close to the critical temperature. At high pressures the gas tends to compress less than the ideal gas law predicts. In most engineering problems the pressures of interest fall within the moderate range and the real gases are described as supercompressible. To correct for the deviation between the measured or observed volume and that calculated using the ideal gas law, an empirical factor z, called the gas deviation factor or the z-factor, is used. In the literature, this factor is sometimes referred to as the compressibility factor, which can result in confusion with another gas property. In order to avoid ambiguity, this factor

39

will be referred to as the gas deviation factor or z-factor throughout this text. The gas deviation factor is defined as

Real Gas Equation of State All gases deviate from ideal gas laws under most conditions. Numerous attempts have been made to account for these deviations of a real gas from theidealgas

40

equation of state. One of the more celebrated of these is the equation of van der Waals. More recent and more successful equations of state have been derived, for instance, the Beattie-Bridgeman and BenedictWebb-Rubin equations. But the real gas equation most commonly used in practice by the industry is

The units are the same as listed for Eq. 2.8; z, which is dimensionless, is

Also, using the ideal gas law (Eq. 2.8) and the definition of z (Eq. 2.16),

the gas deviation factor. The z-factor can be interpreted as a term by which the pressure must be corrected to account for the departure from the ideal gas equation by expressing Eq. 2.17 as

Equation 2.17 may be written, for a certain quantity of gas,

as

41

The z-factor is a function of both absolute pressure and absolute temperature; but for this book's purposes, the main interest lies in determining z as a function of pressure at constant reservoir or transmission temperature. The z(p) relationship obtained is then appropriate for the description of isothermal reservoir depletion or isothermal transmission problems. Equation 2.17 may be written in terms of specific volume v or density p and gas gravity y g
or

where v = specific volume, cu ft/lbm m = mass of gas, Ibm M= molecular weight of gas, lbm/lb-mol p = density of the gas, Ibm/cu ft Comparing the density of a gas, at any pressure and temperature, to the density of air at the same conditions gives

And, in particular, at standard conditions:

The Theorem of Corresponding States Beginning with his thesis in 1873, J. D. van der Waals proposed his theorem of corresponding states. Before stating this theorem and discussing some of its uses, the following terms will be defined.

42

Critical pressure is that pressure which a gas exerts when in equilibrium with the liquid phase and at the critical temperature. It may also be defined as the saturation pressure corresponding to the critical temperature. Critical temperature is that temperature (of a gas) above which a gas cannot be liquefied by the application of pressure alone, regardless of the amount of pressure. Critical volume is the volume of 1 pound-mass of gas at the critical temperature and pressure; that is, the specific volume of the gas at critical temperature and critical pressure. Reduced temperature, reduced pressure, and reduced volume are the ratios of the actual temperature, pressure, and specific volume to the critical temperature, critical pressure, and critical volume, respectively:

For any substance, the absolute magnitude of pressure or temperature is not really what counts in locating a state point, that is, in the liquid region or two-phase region or superheat region, etc. The value of the pressure and temperature relative to a corresponding critical value really counts. Thus, the reduced coordinates are most important. The physical characteristics of a substance are controlled by the relative nearness of any state point to the critical point. If the pressure relative to the critical pressure and the temperature relative to the critical temperature are the same for two different substances, then the substances are in corresponding states and any other property, like the density relative to the critical density, will be the same for both substances. This is the theorem or principle of corresponding states. Stated in other words, the deviation of a real gas from the ideal gas law is the same for different gases at the same corresponding conditions of reduced temperature and reduced pressure. The theorem of corresponding states

43

is accurate within several percent for widely dissimilar types of substances and much more accurate than this for restricted substances having physiochemically similar characteristics. The theorem of corresponding states has many practical applications. The most common use is in evaluating the deviation of real gases from the equation of state for an ideal gas. This permits evaluation of other thermodynamic property deviations from ideal gas relationships. Equation 2.21 can be expressed in the form

or By the theorem of corresponding states, reduced pressure and temperature define the reduced density, or

Simultaneous solution of Eqs. 2.27 and 2.28 yields the general principle If the form of the algebraic function/ in Eq. 2.28 is known, then z c can be defined. Using Eq. 2.29, the z-factor for any gas mixture is defined solely by reduced temperature and reduced pressure:

Reduced vapor pressure, reduced enthalpy, reduced entropy, etc. are examples of this theorem's uses for the purposes of generalizing results. The generalization of vapor pressure data permits description of such data for many gases on one chart. Also, the pressure-density-temperature relationship may be generalized for many different liquids by use of reduced coordinates.

44

Determination of z-factor Experimental Determination Some quantity of gas (n moles) is charged into a cylinder, the volume of which can be altered by the movement of a piston. The container is maintained at the desired temperature T throughout the experiment. If V0 is the gas volume at an atmospheric pressure of 14.7 psia then, applying the real gas law Eq. 2.17, 14.7 V 0 = nRT since z= 1 at atmospheric pressure. At any higher pressure p for which the corresponding volume of the gas is V, pV = ZnRT . Dividing these equations gives z = pV/ 14.7 V o . By varying p and measuring V, the isothermal Z(p) function can be readily obtained. This is the most satisfactory method of determining the function; but in the majority of cases, the time and expense involved are not warranted since reliable methods of direct calculation are available. The z-factor Correlation of Standing and Katz In 1941, Standing and Katz presented a z-factor chart (Fig. 2.4) based on binary mixtures and saturated hydrocarbon vapor data. Figure 2.4 is a correlation of the z-factor as a function of reduced temperature and pressure. This chart is generally reliable for sweet and natural gases and correctable for those containing hydrogen sulfide and carbon dioxide. It has become one of the most widely accepted correlations in the petroleum industry. This correlation requires a knowledge of the composition of the gas or, at least, the gas gravity. In order to use the StandingKatz correlation, it is first necessary from a knowledge of the gas compositionto determine the pseudo critical pressure and temperature, or the apparent molecular weight of the mixture. These pseudo- properties are given by Kay's mixing rules as

45

where

Fig. 2.4 Gas deviation factor for natural gases. (After Standing and Katz.)

46

In cases where the composition of a natural gas is not available, the pseudo- critical pressure and pseudocritical temperature may be approximated from Fig. 2.5 and a knowledge of the gas gravity (Brown et al., Carr et ah). Useful correlations derived from Fig. 2.5 and more recent data from other sources by Thomas, Hankinson, and Phillips are:

The allowable concentrations of sour gases and other nonhydrocarbons for the above equations are 3% H2S, 5% N2, or a total impurity content of 7%.

Gas gravity. Air = 1

Fig. 2.5 Pseudocritical properties of miscellaneous natural gases. (After Brown et al.)

47

The next step is to calculate the pseudoreduced pressure and temperature:

where p and T are the absolute pressure and absolute temperature at which z- factor is required. With these two reduced parameters the StandingKatz correlation, which consists of a set of isotherms giving z as a function of the pseudoreduced pressure, can be used to determine the z-factor.
48 Properties of Natural Gases and Condensate Systems

For comparison, from Fig. 2.5 for yg = 0.686, p pc = 667 psia, T pc = 376 R; and using Eqs. 2.33 and 2.34, ppc - 667 psia, T pc = 381 R. The Standing- Katz z-factor chart is generally reliable for sweet natural gases with small amounts of nonhydrocarbon components, say, less than 5% by volume. For sour natural gases, the Standing-Katz z-factor chart may be used with appropriate adjustment of the pseudocritical temperature

and pseudocritical pressure. The pseudocritical temperature adjustment factor e3 is given by

where

Wichert and Aziz have developed a chart (Fig. 2.6) that gives values of e 3 to be used in adjusting the pseudocritical temperatures and pressure as follows:

where

Using the adjusted pseudocritical temperature and pressure, the reduced temperature and reduced pressure are calculated for use in Fig. 2.4 to predict sour gas z-factors.

49

Fig. 2.6 Pseudocritical temperature adjustment factor, e3, R. (After Wichert and Aziz.)

50

Using Eq. 2.39,

The reduced properties at 180F and 2500 psia become

Using these values in Fig. 2.4 gives a gas deviation factor z = 0.850. In using the mole fraction analysis outlined above, the C7 + fraction is generally considered as a single component. In Example 2.10, the physical properties of octane (CgH18) are used for the C7 + fraction. This is a common practice. If the molecular weight and specific gravity of the C7 + fraction are known, the pseudocritical temperature and pseudocritical pressure of this fraction may be obtained by Fig. 2.7. Figures A.l to A.9 in the Appendix give plots of z-factor versus pressure for methane and gases of various gravities. These may be used instead of the Standing and Katz correlation (Fig. 2.4). Direct Calculation of z-factors The StandingKatz z-factor correlation is very reliable and has been used with confidence by the industry for more than 35 years. With the advent of computers, however, the need arose to find some convenient technique for calculating z- factors for use in natural gas engineering programs, rather than feeding in the entire chart from which z-factors could be retrieved by 1. The Hall- Yarborough Method (1973) The Hall-Yarborough equations, developed using the Starling-Carnahan equation of state, are where

table lookup. Takacs (1976) has compared eight different methods for calculating z-factors, which have been developed over the years. These fall into two main categories: those that attempt to curve-fit the Standing-Katz isotherms analytically and those that compute z-factors using an equation of state. Some of these are given below. P pr = the pseudoreduced pressure t = the reciprocal, pseudoreduced temperature (T pc /T)

51

The nonlinear equation (Eq. 2.41) can be solved for / using Newton Raphson iterative technique. Substitution of the correct value of y in Eq. 2.40 will give the z-factor. 2. Dranchuk, Purvis and Robinson Method (1974) This method fits the StandingKatz z-factor correlation by means of an eight-coefficient BenedictWebb Rubin type equation of state. The z-factor equation is

Fig. 2.7 Correlation charts for estimation of the pseudocritical temperature and pressure of heptanes plus fractions from molecular weight and specific gravity. (Courtesy Mathews, Roland, and Katz.)

Molecular weight

3. Gopal Method (1977) This method fits straightline equations to different portions of the z-factor chart. It uses a general equation of the form

The values of constants A, B, C, and D for various combinations of p r and T r are shown in Table 2.4. Note that above pr of 5.4, an equation of a different form is used.

Supercompressibility Sometimes, another term is used by engineers to compensate for the deviation from the ideal gas law. This term is the supercompressibility factor F pv and is

53

used primarily for high-pressure calculations. A mathematical relationship exists between supercompressibility and gas deviation factor, as follows:

The equation of state for real gases can now be written in terms of the supercompressibility factor as

Tables of supercompressibilities for natural gases are available from a number of sources (see Chapter 6) and can be readily used with the above equation to predict PVT relationships.

Other Equations of State


An equation of state gives the relationship between the pressure p, molar volume v, and absolute temperature T of a fluid. The functional form may be written as Alternatively, Eq. 2.47 may be written explicitly in terms of p: or explicitly in terms of v or T. The ideal gas, and one form of real gas, equations of state have been discussed earlier, Eqs. 2.8 and 2.17, respectively. The gas deviation factor in the real gas equation of state presented earlier, Eq. 2.17, is not a constant, and needs to be obtained by graphical or numerical techniques. This is a major limitation, which makes direct mathematical manipulations difficult. Many other equations of state have been developed with a constant correction factor thus permitting direct mathematical manipulations like differentiation and integration. Van der Waals' Equation of State This is an attempt to modify the ideal gas law so that it will be applicable to nonideal gases. For one mole of a pure gas, Van der Waals' equation of state is written:

54

where a and b are constants. Table 2.5 gives values of a and b for common substances. The quantity a/v2 accounts for the attractive forces between the molecules. Actual external pressure would need to be larger to produce the same volume than if no attraction existed. The constant b represents the volume of the molecules themselves. Actual volume available to the gas is less than overall volume of gas by the amount taken up by the molecules. If n moles of gas are involved, Eq. 2.49 becomes

When V is large (low pressure and high temperature), van der Waals' equation reduces to the ideal gas law. If the constants a and b are not known, it is possible to estimate their values from values of critical pressure P c and critical volume Vc using

Van der Waals' equation has limited application in engineering. It is accurate only at low pressures.

55

TABLE 2.5
Van der Waals Constants

Substance Methane

a Liters-atm mole 2.253

b Liters/mole 0.04278 0.06380 0.08445 0.1142 0.1226 0.1417 0.1460 0.1735 0.2654 0.03913 0.04267 0.04287 0.02370 0.03049 0.02661 0.05714 0.08272

Ethane Propane fsobutane n-Butane Isopentane n-Pentane n-Hexane n-Heptane Nitrogen Carbon dioxide Hydrogen sulfide Helium Water Hydrogen Ethylene Propylene

5.489 8.664 12.87 14.47 18.05 19.01 24.39 31.51 1.390 3.592 4.431 0.03412 5.464 0.2444 4.471 8.379

56

Benedict-Webb-Rubin (B-W-R) Equation of State


A powerful equation of state describing the behavior of pure, light hydrocarbons over single and two-phase regions, both below and above critical pressure, was developed by Benedict, Webb, and Rubin as early as 1940. The B-W-R equation of state is a modification of the BeattieBridgeman equation of state and gives better PVT predictions than the Beattie-Bridgeman equation. et al., inand a Condensate series of articles in 1951, 57Benedict Properties of Natural Gases Systems showed the accuracy of the B-W-R equation when applied to light hydrocarbons and their simple mixtures and demonstrated the equation's utility for making vapor-liquid equilibrium calculations. Since natural gases are primarily mixtures of light hydrocarbons, the B-W-R has been used quite often in computing thermodynamic properties and phase equilibria of natural gases. The B-W-R equation expressing pressure as a function of temperature and molar density is

The parameters A 0 , B 0) C 0 , a, b, c, a, and y are constants for pure substances and functions of composition for mixtures. The purecomponent values of the eight constants are listed in Table 2.6. Benedict et al. proposed the following mixing rules to enable Eq. 2.53 be used for mixtures:

P m is the molar density in lb-mole/cu ft Redlich-Kwong (R-K) Equation of State The original Redlich-Kwong equation of state is

58

For temperature in R and pressure in psia,

The R-K equation involves only two empirical constants, as opposed to the eight required in the B-W-R equation. However, to simplify calculations with the R-K equation, especially for application to mixtures, other constants have been defined:
LJU CO

<
H

The original R-K equation was modified by Soave in 1972 by replacing the term 1 /T l/2 with a more general temperature dependent term a:

59

A gas deviation factor z can be calculated by writing Eq. 2.60 as where, for pure substances,

For mixtures, the constants A and B are calculated using

60

where T Bi is the boiling point in R of component i at 14.7 psia. Values of o>, may be obtained from Table A.l. Then,

61 Properties of Natural Gases and Condensate Systems

and

COMPRESSIBILITY OF NATURAL GASES The coefficient of isothermal compressibility of a gas is given by

For an ideal gas,

For a real gas,

and

Thus,

or

In terms of reduced variables,

or and

Using Eqs. 2.42 and 2.43 along with Eq. 2.71, Mattar, Brar, and Aziz have obtained an expression for c pr given by

where

Thus,

62

Eqs. 2.76 and 2.77 can be used directly in computer calculations. These equations were used to develop Figs. 2.8 and 2.9 for manual calculations of the coefficient of isothermal compressibility.

Fig. 2.8 Variation of c,Tr with reduced temperature and pressure (1.05 ^ Tr s 1.4; 0.2 *s pr 15.0). (After Mattar, Brar, and

Aziz.)

63

Fig. 2.9 Variation of crTr with reduced temperature and pressure (1.4 Tr 3.0; 0.2 pr 15.0). (After Mattar, Brar, and

Aziz.)

64

You might also like