You are on page 1of 28

Chapter 39

Gas-Condensate Reservoirs
Phillip L. Moses, Core Laboratories ~nc.* Charles W. Donohoe. Core Laboratories I~C

Introduction
The importance of gas-condensate reservoirs has grown continuously since the late 1930 s. Development and operation of these reservoirs for maximum recovery require engineering and operating methods significantly different from crude-oil or dry-gas reservoirs. The single most striking factor about gas-condensate systems (fluids) is that they exist either wholly or preponderantly as vapor phase in the reservoir at the time of discovery (the critical temperature of the system is lower than the reservoir temperature). This key fact nearly always governs the development and operating programs for recovery of hydrocarbons from such reservoirs; the properties of the fluids determine the best program in each case. A thorough understanding of fluid properties together with a good understanding of the special economics involved is therefore required for optimum engineering of gascondensate reservoirs. Other important aspects include geologic conditions. rock properties, well deliverability, well costs and spacing, well-pattern geometry, and plant costs. Engineers have a wealth of literature on gas-condensate reservoirs available for reference. From this mass of material, Refs. 1 through 5 are especially recommended for fundamental background, and Refs. 6 through 8 are recommended for information on properties of pure compounds and their simple mixtures related to gas-condensate systems. For information regarding reservoir engineering processes and data, Refs. 5 and 9 through 16 are recommended. The best single bibliography on gas-condensate reservoirs is that of Katz and Rzasa ; however, later pertinent literature listings will be found in Refs. 6 through 14. The collection of references in Refs. 11 and 12 is particularly recommended for case histories of various gascondensate operations. Petroleum production papers published by SPE (AIME) s and API have been indexed separately through the years 1985 and 19.53, respectively. The practicing field engineers should have the following minimum library on gas-condensate systems available for their use: either Ref. 1, 2, or 3; Refs. 5, 9, 13, and 15; and selected volumes of Refs. 11 and 12.

Properties and Behavior of Gas-Condensate Fluids


Sloan* described the general occurrence of petroleum in the earth: . think of all the hydrocarbons, beginning with the lightest, methane, to the heaviest asphaltic substances as a series of compounds of the same family, consisting of carbon and hydrogen in a limitless number of proportions. A hydrocarbon reservoir then. is a porous section of the sedimentary crust of the earth containing a group of hydrocarbons, which is probably unique and whose overall properties such as reservoir phase, gas/oil ratio, gasoline content, viscosity. etc., is the direct result of this composition, together with the temperature and pressure that happen to exist in this particular spot in the porous sediment. It is now easy to conceive of any possible combination of these hydrocarbons in a given reservoir, and it is also easy to visualize a reservoir fluid whose physical state may range from a completely dry gas in the reservoir, shading gradually through the wet gas, the condensate, the critical mixture, the highly compressible volatile liquid, the more stable light crude oil whose color is beginning to darken, the heavier crudes with decreasing solution gas, and ending with the semisolid asphalts and waxes with no measurable solution gas. The condensate reservoir that is the topic under discussion is therefore first a hydrocarbon reservoir. Due to the composition and proportion of the individual hydrocarbons in the mixtures, the content is gas phase at the temperature and pressure of the reservoir.

39-2

PETROLEUM ENGINEERING

HANDBOOK

TABLE 39.1~-HYDROCARBON

ANALYSES

AND PROPERTIES OF EXAMPLE CRUDE OILS AND GAS CONDENSATES Mole Fraction Crude Oil Crude Oil
8

Component Carbon dioxide Nitrogen Methane Ethane Propane Butanes Pentanes Hexanes Heptanes and heavier Molecular weight C, plus Specific gravity C, plus, 60 /6O F Viscosity C, plus, Saybolt universal seconds at lOOoF Tank-oil gravity, OAPI at 60/600F Producing gas/oil ratio, cu ft/bbl
-see Ref 12, D 327 See Ref. 2, Vbl I, Table 8 8, pp. 402-W Viscosity 01 residual 011 left in apparatus,

Condensate 843 0.00794 0.01375 0.76432 0.07923 0.04301 0.03060 0.01718 0.01405 0.02992 120 0.7397

Condensate 944 0.00130 0.00075 0.89498 0.04555 0.01909 0.00958 0.00475 0.00365 0.02015 144 0.7884

Condensate 1143 0.00695 0.01480 0.89045 0.04691 0.01393 0.00795 0.00424 0.00379 0.01098 143 0.7593

0.4404 0 0432 0.0405 0.0284 0.0174 0.0290 0.4011 287 0.9071 0.5345 0.0636 0.0466 0.0379 0.0274 0.0341 0.2559 247 0.8811 42 34.5 1,078

loo+
27.4 525

73 18,000+

53.2 43,000 f

61.1 69,000 a

approwmal~ng

Ihe hexanes-plus

material

Composition Ranges of Gas-Condensate

Systems

Approximate composition indices for gas-condensate systems are the gas/liquid ratio of produced fluids (sometimes called the GOR) or its reciprocal, the liquid/gas ratio, and the gravity of the tank liquid separated out under various surface conditions. These two indices vary widely; they do not necessarily prove whether a hydrocarbon system is in the vapor phase in the reservoir. Eilerts et al. (Vol. 1, Chaps. 1 and 8) show in a survey that the liquid/gas ratios of gas-condensate systems can vary from more than 500 (very rich) to less than 10 bbl/MMscf; tank condensate produced from the wells varied from less than 30 to more than 80API, and more than 85% was within the range of 45 to 65API. Eilerts et al. (Vol. 1) also quote a rule of thumb that a gascondensate system exists when the gas/liquid ratio exceeds 5,000 cu ftibbl (200 bbl/MMscf and less) and the liquid is lighter than 5OAPI. This appears to be on the conservative side because there is evidence that systems exist as single-phase vapor in the reservoir when the surface gas/liquid ratio is less than 4,000 cu ft/bbl (more than 250 bbl/MMscf) and the API gravity of the liquid in the stock tanks is lower than 40API. A more accurate representation of the composition of gas-condensate fluids is provided by fractional analyses of the well streams coming from the reservoirs. The contrast of the fluid composition with the total stream coming from crude-oil reservoirs is fairly large for the relative amounts of the lighter vs. heavier ends of the paraffinhydrocarbon series. For example. Eilerts et ul. (Vol. 1, Table 8.8) report a methane content from about 75 to 90 mol% for several gas-condensate systems, whereas Dodson and Standing report 44 and 53 mol%, respectively, for two crude-oil systems (see Table 39.1). The table, however, shows much lower heptanes-and-heavier content for the gas-condensate systems than for the crude oil. These are the two outstanding composition features of gas-condensate systems.

Pressure and Temperature Ranges of Gas-Condensate Reservoirs Gas-condensate reservoirs may occur at pressures below 2.000 psi and temperatures below l00F20 and probably can occur at any higher fluid pressures and temperatures within reach of the drill. Most known retrograde gas-condensate reservoirs are in the range of 3,000 to 8,000 psi and 200 to 400F. These pressure and temperature ranges, together with wide composition ranges, provide a great variety of conditions for the physical behavior of gas-condensate deposits. This emphasizes the need for very meticulous engineering studies of each gascondensate reservoir to arrive at the best mode of development and operation.

Phase and Equilibrium

Behavior

An understanding of the behavior of pure paraffin hydrocarbons and simple two-component or threecomponent systems (involving such compounds as methane, pentane, and decane) is of considerable benefit to the engineer working with gas-condensate reservoir problems. Excellent coverage is given this subject by Sage and Lacey and a more condensed discussion by Burcik. Occasional review of such material will assist the engineer concerned with more complex hydrocarbon mixtures. Chap. 23 describes the phase and equilibrium behavior of complex (multicomponent) hydrocarbon mixtures (see Fig. 23.14 and the accompanying discussion). Note that the critical state (critical point) is that state or condition at which the composition and all other intensive properties of the gas phase and the liquid phase become identical-i.e., the phases are indistinguishable. In gascondensate reservoirs, the portion of the phase diagram to the left of and above the critical point will not be involved.

GAS-CONDENSATE

RESERVOIRS

39-3

i
1T O50 I00 150 200 TEMPERATURE. F 250 300

Fig. 39.1-Phase

diagram of Eilerts Fluid 843.

Fig. 39.2-Phase

diagram of Eilerts Fluid

1143.

(discussed in The term retrograde condensation Chap. 23) is used more loosely than implied by its rigorous definition, In field practice, the term may imply any process where the amount of condensing liquid phase passes through a maximum, whether the process is isothermal or not. While Fig. 23.14 provides a simplified picture of the phase diagram, reservoir engineers will find that very few quantitative phase diagrams on naturally occurring gascondensate mixtures have been published. Figs. 39.1 through 39.3 come from extensive work and represent quantitative measurements on the flow streams from wells in the Chapel Hill, Carthage, and Seeligson fields in Texas. The critical points are not shown because they are at temperatures below those of interest to field operations. This emphasizes that the compositions of gas-condensate systems vary widely and strongly affect the form of the phase diagrams encountered in actual gas-condensate reservoirs. These three phase diagrams represent a reasonable spread in the properties of gas-condensate systems. from a gas/liquid ratio of about 18,000 to 69,000 cu ftibbl (56 to 14.5 bbl/MMscf). This does not mean, however, that all other gas-condensate systems would fall inside the limits of the properties suggested by these three phase diagrams. The three cases in Figs. 39.1 through 39.3 imply that the dewpoint boundary approaches zero pressure at a relatively high temperature. Other condensate systems are believed to approximate the qualitative picture shown in Fig. 23.14 more closely. Note that all three systems exhibit both cricondentherm and cricondenbar points (maximum temperature and pressure, respectively, beyond which there is no liquid present in the vapor); the critical temperatures all fall to the left of each diagram at lower temperatures and pressures than the maxima for the dewpoint boundaries. Liquid-content lines on phase diagrams can be represented by a number of different units. Figs. 39.1 through 39.3 use gallons per thousand cubic feet of separator gas.

The approximate behavior of condensate fluids while being produced from the reservoir into surface vessels can be represented advantageously on phase diagrams. In Fig. 39.2, for example, Line FT shows a flow path for fluids that starts at formation conditions (outside the dewpoint boundary, indicating that the formation fluids were all in vapor phase); proceeds to sandface pressure, Point S i , at the well; declines as the fluid rises from the bottom of the hole to the wellhead, Point WH; passes through the choke to separator conditions, Point S2 ; and reaches Point T, representing tank conditions. The phase diagram is thus helpful to the engineer in visualizing what happens to gas-condensate fluids as they flow from the formation to the wellbore and from there to surface equipment.

4,5OOf7777777

TEMPERATURE,

*F

Fig. 39.3-Phase

diagram of Eilerts Fluid 944

39-4

PETROLEUM ENGINEERING

HANDBOOK

Methods have been proposed by Organick and Eilerts et al. * for predicting the critical temperatures and pressures of hydrocarbon mixtures and for computing the phase diagrams (including dewpoint curves) of gascondensate fluids. The dependability of these methods for a wide range of gas-condensate compositions has not yet been established. For reservoir engineering work, direct laboratory measurements of phase diagrams or of pressure-depletion behavior are necessary because of the large recoveries at stake. Laboratory work may not be required for other problems. Gas/Liquid Ratios and Liquid Contents of Gas-Condensate Systems As discussed earlier, it is difficult to specify whether a hydrocarbon system is in the vapor phase in the reservoir from measurements of field gas/liquid ratio and tankoil gravity. Fluid production with tank-oil gravities as low as 30API and gas/liquid ratios as low as 3,000 cu ft/bbl may be from true gas-condensate systems; this possibility should always be checked by laboratory measurements of phase behavior for these and intermediate values. Liquid content and gas/liquid ratio can be direct reciprocals, depending on the type of problem considered. The terms must be carefully defined in each case because gas-condensate systems in the field frequently undergo different types of separating procedures that involve several stages before the final liquid phase (liquid means hydrocarbon liquid unless otherwise specified) reaches the tanks at atmospheric pressure. To study the properties of gas-condensate fluids at reservoir conditions, it is convenient to define gas/liquid ratios and liquid contents on the basis of the gas and liquid outputs of the first-stage separator through which the fluids pass. These two output streams then represent the total composition of the gas-condensate fluid in the reservoir if sampling, producing, and measuring conditions have been properly set and maintained. Other gas/liquid ratios may be reported, however, including the total gas output of all stages of separation divided by the tank-liquid volumes corresponding to the gas output: note that the total gas output would include a measurement of tank vapors as well as separator gas to represent the full composition of the wellstream. The gas/liquid ratio at stock-tank conditions may be roughly approximated when field facilities are not available for measurements. The gas and liquid flow rates from the high-stage separator are observed and a liquid sample collected from the separator in a stainless-steel cylinder of known volume. If all the cylinder contents are bled into a calibrated graduate at atmospheric pressure and the volume of the resultant liquid phase is compared with the original liquid volume, an approximate value of the liquidphase shrinkage may be determined. From this, the highstage gas/liquid ratio may be converted to stock-tank conditions. This procedure ignores the volume of gas liberated between high-stage separator and stock-tank conditions. This volume can be approximated by using a calibrated glass separator with gas meter attached in place of the graduate. Ignoring this gas volume adds further errors to those resulting from not simulating the existing field stage separation conditions. The higher the first-stage separation pressure, the greater the error in total gas volume of the gas/liquid ratio. This is only an approx-

imate method that may be used when there are no intermediate separator stages and stock tanks for individual well measurements and when the atmospheric temperature and pressure do not vary appreciably from stock-tank conditions. Gas/liquid ratios usually are reported in cubic feet per barrel of liquid (or thousands of cubic feet per barrel) and liquid contents or liquid/gas ratios in barrels of liquid per million standard cubic feet of gas. The separator streams used in the ratio must be specified. Properties of Separated Phases The properties of both liquid and gas phases separated from gas-condensate streams can vary considerably. One of the dominant properties of the gas is high methane content. Eilerts et al. 2 (Vol. I, Chap. 8) list the compositions of the gas and liquid phases of eight gas-condensate systems. Methane contents of the gas phases (simulated from field separators) varied from about 0.83 to 0.92 mole fraction; the hexanes and heavier (hexanes plus) varied from 0.004 to about 0.008 mole fraction. The liquid phases varied from about 0.1 to nearly 0.3 mole fraction methane; hexanes plus varied from about 0.4 to 0.7 mole fraction. In the absence of measured data, properties of the separated phases of gas-condensate systems (including volumetric and density behavior) can be approximated by methods described in Chaps. 20 through 23, especially Chaps. 20 and 22 (see also Refs. 9 and 14). Viscosities of Gas-Condensate Systems

The viscosity of a gas-condensate system is of interest in various reservoir calculations, particularly with respect to cycling operations and the representation of such reservoirs in computer models. Whenever possible, viscosity of the vapor phase at reservoir conditions should be measured directly. Carr et al. 23 presented a method to estimate the viscosities of gas systems from a knowledge of compositions or specific gravities (see also Chap. 20 and Ref. 14). Viscosities of separate gas and liquid phases at the surface conditions usually encountered can be obtained by direct measurement or by the use of the correlations for gas previously mentioned and the correlation of Chew and Connally24 for liquid (see also Chap. 22). Viscosity information on separated materials is needed mainly for separator or plant residue gases to be injected during cycling and for some types of reservoir calculations.

Gas-Condensate Well Tests and Sampling


Proper testing of gas-condensate wells is essential to ascertain the state of the hydrocarbon system at reservoir conditions and to plan the best production and recovery program for the reservoir. Without proper well tests and samples, it would be impossible to determine the phase conditions of the reservoir contents at reservoir temperature and pressure accurately and to estimate the amount of hydrocarbon materials in place accurately. Tests are made on gas-condensate wells for a number of specific purposes: to obtain representative samples for laboratory analysis to identify the composition and properties of the reservoir fluids; to make field determinations on gas and liquid properties; and to determine formation

GAS-CONDENSATE

RESERVOIRS

39-5

and well characteristics, including productivity. producibility, and injectivity. The first consideration for selccting wells for gas-condensate fluid samples is that they be far enough from the black-oil ring (if present) to minimize any chance that the liquid oil phase will enter the well during the test period. A second and highly important consideration is the selection of wells with as high productivities as possible so that minimum pressure drawdown will be suffered when the reservoir fluid samples are acquired. Well Conditioning Proper well conditioning is essential to obtain representative samples from the reservoir. The best production rates before and during the sampling procedure have to be considered individually for each reservoir and for each well. Usually the best procedure is to use the lowest rate that results in smooth well operation and the most dependable measurements of surface products. Minimum drawdown of bottomhole pressure during the conditioning period is desirable and the produced gas/liquid ratio should remain constant (within about 2%) for several days; the less-permeable reservoirs require longer periods. The farther the well deviates from constant produced gas/liquid ratio. the greater the likelihood that the samples will not be representative. Recombined separator samples from gas-condensate wells are considered more representative of the original reservoir fluid than subsurface samples. Accurate measurements of hydrocarbon gas and liquid production rates during the well-conditioning and wellsampling tests are necessary because the laboratory tests will later be based on fluid compositions recombined in the same ratios as the hydrocarbon streams measured in the field. The original reservoir fluid cannot be simulated in the laboratory unless accurate field measurements of all the separator streams are taken. (Gas/liquid ratios may be reported and used in several different forms, as discussed previously.) If the produced gas/condensate (gas/liquid) ratio from field measurements is in error by as little as 5 %, the dewpoint pressure determined in the laboratory may be in error by as much as 100 psi. Water production rates should be measured separately and produced water excluded as much as possible from hydrocarbon samples sent to the laboratory. Separator pressure and temperatures should remain as constant as possible during the well-conditioning period; this will help maintain constancy of the stream rates and thus of the observed hydrocarbon gas/liquid ratio. If the well is being prepared during a period when atmospheric temperatures vary considerably from night to day. reasonably consistent average temperatures and pressures on the several vessels during the conditioning period should be adequate. Field Sampling and Test Procedures After the conditioning period has proceeded long enough to show that producing conditions are steady. exacting measurement methods must be used to obtain representative samples. The mechanics of well sampling is partially covered in Chaps. 12 through 14, 16, and 17. The help of experienced laboratory personnel is advisable in

acquiring gas and condensate-liquid samples. Certain minimum items of information in addition to all stream rates are essential, including regular readings of the pressures and temperatures of all vessels sampled, and of tubing heads and casing heads where available, and a recorded history of the well conditions before and during sampling. along with the actual mechanics of the sampling steps. Other information acquired during the sampling period that would help to explain reservoir and well conditions should also be recorded because it is useful in interpreting the results of the tests. Care must be taken that the compositions of gas and liquid samples obtained are representative and are properly preserved for laboratory analyses. API RP 44? outlines appropriate sampling methods. For cases when the liquid-phase sample is obtained at a low temperature (from low-temperature separation equipment), triethylene-glycol/water mixtures are convenient for collecting the samples. Ten percent or more of the cylinder volume for liquid-phase samples should be gas to prevent excessive pressure that could result from temperature rise during subsequent shipment. This 10% gas cap can be effected by closing the cylinder sampleinlet valve when 90% of the glycoliwater mixture has been displaced and then carefully withdrawing nearly all the remaining mixture from the bottom of the cylinder without losing the oil phase. The volumes of fluids requested for laboratory testing should be acquired during the sampling period. plus a reasonable amount (25 % or more) of extra sample materials in separate containers for emergency use should some of the main samples be lost by leakage or other adversity between the field site and the laboratory. At the end of actual sampling mechanics in the field, the well should remain on stream for a reasonable period of time, and its producing rate, gas/liquid ratio, and various pressures and temperatures should be observed to confirm that they are consistent with the information developed before and during the sampling period. Any radical changes should be analyzed carefully to decide whether resampling may be necessary to ensure accuracy of the samples and well statistics obtained during the sampling period. Equipment is available for making some determinations of gas-condensate properties in the field. Among these properties are the gas/liquid ratios of several vessels simulating various separation conditions (numbers of stages, pressures and temperatures of the stages, and other conditions) and the gasoline content of the overhead gas at each stage. If hydrogen sulfide and carbon dioxide are present in the production streams, special sampling procedures should be used and the samples should be taken in stainless-steel cylinders. These corrosive gases could react with the sample cylinders during shipment. Field determinations of the hydrocarbon compositions of streams from gas-condensate wells can be made with appropriate fractionation equipment in mobile laboratories. Eilerts rt al. described such equipment and the test procedures for determining the effect of individual hydrocarbons on liquid/gas ratios at different separation pressures and temperatures. These tests can assist in determining optimum field separation conditions for given production objectives. They require special equipment and experienced personnel.

39-6

PETROLEUM

ENGINEERING

HANDBOOK

Measurements of gas-condensate well productivity, producibility, and injectivity are of considerable importance for planning overall field operations and size of plants for either gasoline recovery or condensate-liquids recovery and cycling, as bases for contracts for deliverability from a reservoir for pipeline purposes, and for various other needs. This topic is discussed more fully later; test procedures are described in Chap. 33 and in several published standards and regulations. 26-29

where K, = the equilibrium ratio for methane, y1 = methane in the vapor phase, mol%, and Xl = methane in the liquid phase, mol%. The experimental equilibrium ratio for methane is 7.71 for the temperature and pressure existing in the field separator at the time of sampling. The equilibrium ratios for each of the hydrocarbons methane through hexane are calculated in a similar manner. These data can then be compared with equilibrium ratios, such as those published in Ref. 16. If the equilibrium ratios compare favorably, then the samples are in equilibrium and the study should continue. If they do not compare well, then new samples should be obtained before proceeding. Recombination of Separator Samples

Sample Collection and Evaluation


In taking samples for recombination to evaluate a gascondensate reservoir, the samples of gas and samples of liquid usually are taken from the first stage of separation. A representative portion of all the hydrocarbons produced from the well will be contained in these two samples. The first step in the laboratory study is to evaluate the samples taken. The first test is to measure the bubblepoint of the separator liquid. The bubblepoint should correspond to the separator pressure at separator temperature at the time the samples were taken. The hydrocarbon composition of the separator samples should then be determined by chromatography or lowtemperature fractional distillation or a combination of both. An example of the composition of typical separator products are shown in Table 39.2. These compositions may be evaluated by calculation of the equilibrium ratio for each component (see Chap. 23). The equilibrium ratio for a component is the mole percent of that component in the vapor phase divided by the mole percent of the same component in the liquid phase. As an example, the equilibrium ratio for methane in Table 39.2 is calculated by the equation K, =yl/x, =83.01/10.76=7.71,

The samples are now ready to be recombined in the same ratio that they were produced. Because we have samples of first-stage separator gas and first-stage separator liquid, we must have the produced gas/liquid ratio in the same form. If the producing gas/liquid ratio was measured in the field in this form, then we can proceed directly with the recombination. If the ratio was measured in the field in the form of primary-separator gas per barrel of second-stage separator liquid or per barrel of stocktank liquid, then a laboratory shrinkage test must be run to simulate field separation conditions. The shrinkage obtained can then be used to convert the field-measured ratio to the form necessary for the recombination. Once the separator products have been recombined, the composition can be measured and compared with the calculated composition. This will check the accuracy of the physical recombination.

TABLE 39.2-HYDROCARBON

ANALYSES OF SEPARATOR

PRODUCTS AND CALCULATED Well Stream mol % 0.00 0.01 0.11 68.93 8.63 5.34 1.15 2.33 0.93 0.85 1.73 9.99 100.00 gal/l,000

WELL STREAM

SeDarator Liauid
Component Hydrogen sulfide C&bon dioxide Nitrogen Methane Ethane Propane iso-Butane n-Butane iso-Pentane n-Pentane Hexanes Heptanes plus (mol %) 0.00 0.00 0.01 10.76 6.17 8.81 2.85 7.02 3.47 3.31 8.03 49.57 100.00
plus

Separator Gas mol % 0.00 0.01 0.13 83.01 9.23 4.50 0.74 1.20 0.31 0.25 0.21 0.41 100.00 gal/l ,000 cf gas

cf gas

2.454 1.231 0.241 0.376 0.113 0.090 0.085 0.185 4.775

2.295 1.461 0.374 0.730 0.338 0.306 0.702 6.006 12.212

Total
Properties of heptanes

API gravity at 6O F 39.0 Density, g/cm3 at 60aR).8293 Molecular weight 160

103

0.827 158 0.699 1,230 3,944 1.191 805.19 171.4

Calculated separator gas gravity (air = 1.000) Calculated gross heating value for separator gas per cubic foot of dry gas at 14.65 psia and 60F, Btu Primary-separator-gas/separator-liquid ratio at 60F, scf/bbl* Primary-separator-liquid/stock-tank-liquid ratio at 60F, bbl Primary-separator-gas/well-stream ratio, MscWMMscf Stock-tank-liquid/well-stream ratio, bbl/MMscf
*Primary separator gas and primary separator liquid collected at 440 psig and 87 F.

GAS-CONDENSATE

RESERVOIRS

39-7

Dewpoint

and Pressure/Volume

Relations

The laboratory personnel will next measure the pressure/volume relations of the reservoir fluid at reservoir temperature with a visual cell. This is a constant-composition expansion and furnishes the dewpoint of the reservoir fluid at reservoir temperature and the total volume of the reservoir fluid as a function of pressure. The volume of liquid at pressures below the dewpoint as a percent of the total volume may also be measured. Phase diagrams can be developed dy measuring the liquid volumes at several other temperatures. Table 39.3 is an example of the dewpoint determination and pressure/volume relations of a gas-condensate reservoir fluid. Simulated Pressure Depletion Pressure depletion of gas-condensate reservoirs may be simulated in the laboratory by use of high-pressure visual cells. In these depletion studies made in the laboratory, the assumption is that the retrograde liquid that condenses in the reservoir rock will not achieve a sufficiently high saturation to become mobile. This assumption appears to be valid except for very rich gas-condensate reservoirs. For very rich gas-condensate reservoirs where the retrograde liquid may achieve a high enough saturation to migrate to producing wells, the gas/liquid relative permeability data should be measured for the reservoir rock system. These data can then be used to ad,just the predicted recovery from the reservoir. Table 39.4 is an example of a depletion study on a gascondensate reservoir fluid. Note from Table 39.4 that the dewpoint pressure of this reservoir fluid is 6,010 psig. The composition listed in the 6,010-psig-pressure column in Table 39.4 is the composition of the reservoir fluid at the dewpoint and exists in the reservoir in the gaseous state

TABLE 39.3-PRESSURE/VOLUME RELATIONS OF RESERVOIR FLUID AT 256OF (Constant-Composition Expansion)

Pressure (PSW
7,500 7,000 * 6,500 6,300 6,200 6,100 6,010+ 5.950 5,900 5,800 5,600 5,300 5,000 4,500 4,000 3,500 3,000 2,500 2,100 1,860 1,683 1,460 1,290 1,160 1,050
;Gas ev~ans~on

Relative Volume 0.9341 0.9523 0.9727 0.9834 0.9891 0.9942

Deviation Factor, z 1.328 1.264 * 1.19s 1.175 1.163 1 150 1.140f

1.oooo
1.0034 1.0076 1.0138 1.0267 1.0481 1.0749 1.1268 1.2024 1.3096 1.4689 1.7169 2.0191 2.2747 2.5150 2.9087 3.3173 3.7153 4.1342
factor = 1 545 Mscllbbl Mscfibb,

Reservoir preSSre
oewpolnl pressure
expansion Gas factor = 1 47,

TABLE 39.4--DEPLETION

STUDY AT 256F Reservoir Pressure, psig

6,010 Component Carbon dioxide Nitrogen Methane Ethane Propane iso-Butane n-Butane iso-Pentane n-Pentane Hexanes Heptanes plus 0.01 0.11 68.93 8.63 5.34 1.I5 2.33 0.93 0.85 1.73 9.99 100.00 Molecular weight of heptanes plus Density of heptanes plus Deviation factor, z Equilibrium gas Two-phase Well stream produced, cumulative % of initial 1.140 1.140 0.000 158 0.827

700 2,100 4,000 3,000 1,200 5,000 Hydrocarbon Analysis of Produced Well Stream, mol %

700*

0.01 0.12 70.69 8.67 5.26 1.10 2.21 0.86 0.76 1.48 8.84 100.00 146 0.817

0.01 0.12 73.60 8.72 5.20 1.05 2.09 0.78 0.70 1.25 6.48 100.00 134 0.805

0.01 0.13 76.60 8.82 5.16 1.01 1.99 0.73 0.65 1.08 3.82 100.00 123 0.794

0.01 0.13 77.77 8.96 5.16 1.Ol 1.98 0.72 0.63 1.Ol 2.62 100.00 115 0.784

0.01 0.12 77.04 9.37 5.44 1.10 2.15 0.77 0.68 1.07 2.25 100.00 110 0.779

0.01 0.11 75.13 9.82 5.90 1.26 2.45 0.87 0.78 1.25 2.42 100.00 109 0.778

Trace 0.01 11.95 4.10 4.80 1.57 3.75 2.15 2.15 6.50 63.02 100.00 174 0.837

1.015 1.016 6.624

0.897 0.921 17.478

0.853 0.851 32.927

0.865 0.799 49.901

0.902 0.722 68.146

0.938 0.612 77.902

39-8

PETROLEUM ENGINEERING

HANDBOOK

1.6

50

15

45 45

1.4

40

I 3

35

12

30

i I

25

10

20

09

15

08

10

07

0.6 0 1000 2000 3000 4000 osi 5000 6000 7000 8000

0 0 1000 2000 3000 4000 psi 5000 6000 7000 8000

Pressure.

Pressure.

Fig. 39.4-Deviation factor, z, of well stream during depletion at 256OF.

Fig. 39.5--Retrograde

condensation

during depletion.

The depletion study is performed by expanding the reservoir fluid in the cell by withdrawing mercury from the cell until the first depletion pressure is reached; this is 5,000 psig in the example. The fluid in the cell is brought to equilibrium and the volume of retrograde liquid is measured. The mercury is then reinjected into the cell and, at the same time, gas is removed from the top of the cell so that a constant pressure is maintained. Mercury is injected into the cell until the hydrocarbon or reservoir volume of the cell is the same as the volume when the test was begun at the dewpoint pressure. The gas volume removed from the cell is measured at the depletion pressure and reservoir temperature. The gas removed is charged to analytical equipment where its composition is determined and its volume is measured at atmospheric pressure and temperature. The composition determined is that listed in Table 39.4 under the heading 5,000 psig. The volume of gas produced in this manner is then divided by the standard volume of gas in the cell at the dewpoint pressure. The produced volume is presented in Table 39.4 as cumulative well stream produced. As mentioned earlier, as the gas is removed from the top of the cell, its volume is measured at the depletion pressure and reservoir temperature. From this volume, the ideal volume of this displaced volume may be calculated with the ideal-gas law. When the ideal volume is divided by the actual volume of the gas produced at standard conditions, we get the deviation factor, z, for the produced gas. This is listed in Table 39.4 under

Deviation Factor z, equilibrium gas and plotted in Fig. 39.4. The actual volume of gas remaining in the cell at this point is the gas originally in the cell at the dewpoint pressure minus the gas produced at the first depletion level. If we divide the actual volume remaining in the cell into the calculated ideal volume remaining in the cell at this first depletion pressure, we obtain the two-phase deviation factor shown in Table 39.4. We call this value a two-phase deviation factor because the material remaining in the cell after the first depletion level is actually gas and retrograde liquid and the actual gas volume we calculated above is the gas volume plus the vapor equivalent of the retrograde liquid. The two-phase z factor is significant in that it is the z factor of all the hydrocarbon material remaining in the reservoir. It is the two-phase z factor that should be used when a plz-vs.-cumulativeproduction plot is made in evaluating gas-condensate production. This series of expansions and constant-pressure displacements is repeated at each depletion pressure until an arbitrary abandonment pressure is reached. The abandonment pressure is considered arbitrary because no engineering or economic calculations have been made to determine this pressure for the purpose of the reservoirfluid study. In addition to the composition of the produced well stream at the final depletion pressure, the composition of the retrograde liquid was also measured. These data are included as a control composition in the event the study is used for compositional material-balance purposes.

GAS-CONDENSATE

RESERVOIRS

39-9

The volume of retrograde liquid measured during the course of the depletion study is shown in Fig. 39.5 and Table 39.5. The data are shown as a percent of hydrocarbon pore space. These are the data that should be used in conjunction with relative permeability data and water saturation data to determine the extent of retrograde liquid mobility. As mentioned earlier, this is a significant factor only with extremely rich gas-condensate reservoirs. Also obtained from the reservoir fluid study is Table 39.6. This table was calculated with the results of the laboratory depletion study described previously applied to a unit-volume reservoir. The unit volume chosen was 1,000 Mscf in place at the dewpoint pressure (note the 1,000 Mscf in Table 39.6 in the first column of numbers). Equilibrium ratios were then used to calculate the amount of stock-tank liquid, primary-separator gas, second-stage gas, and stock-tank gas contained in the unit-volume reservoir. The equilibrium ratios used were for the separator conditions listed at the bottom of Table 39.6. The separator conditions used for these calculations should be the conditions in use in the field or those conditions anticipated for the field. The relative amounts of gas and liquid produced will be a function of the surface separation conditions, among other things. These calculations may be made at a variety of conditions to determine optimum separator pressures and temperatures. For the purpose of this table, production was begun at the dewpoint pressure. The amount of total well effluent (well stream) produced from this unit-volume reservoir as a function of pressure is listed in the table. The amount of stock-tank liquid produced as a function of pressure is also listed. The primaryseparator gas, second-stage gas, and stock-tank gas are presented in a similar manner. Various other factors associated with the production of the gas and condensate from this reservoir are also presented in the table.

TABLE 39.5--RETROGRADE CONDENSATION DURING GAS DEPLETION AT 256 F

Pressure W9)
6,010 5,950

Retrograde Liquid Volume (% hydrocarbon pore space)


0.0

Trace
0.1 0.2 0.5 2.0 7.8 21.3 25.0 24.4 22.5 21.0 17.6

5,900
5,800 5,600 5,300 5,000 * 4,000 3,000 2,100 1,200 700 0
Dewpmt pressure First depletion level.

Table 39.6 shows the initial stock-tank liquid in place to be 181.74 bbl for this unit-volume reservoir. After production to 700 psig, 51.91 bbl had been produced. The difference between these two numbers (18 1.74 - 5 1.9 1), 129.83 bbl, is the amount of retrograde loss or liquid still unproduced at 700 psig expressed in terms of stock-tank barrels. The value of 181.74 bbl may be considered the recovery by pressure maintenance, assuming 100% conformance and 100% displacement efficiency. Table 39.7 furnishes the gravity of the stock-tank liquid that may be expected to be produced as a function of reservoir pressure. Also reported are the instantaneous gas/liquid ratios as a function of reservoir pressure.

TABLE 39.6-CALCULATED

CUMULATIVE

RECOVERY

DURINGDEPLETIONPER MMscf OF ORIGINALFLUID


Reservoir Pressure (wig) 4,000 3,000 2,100 174.78 21.83 145.16 5.17 5.38 344 163 73 35 27 17 8 3 404 250 178 890 329.27 31 .a9 283.78 8.03 8.73 674 331 155 73 42 27 13 5 767 468 325 1,322 499.01 39.76 440.02 10.51 11.85

Initial in Place Well stream, Mscf Normal temperature separation Stock-tank liquid, bbl Primary separator gas, Mscf Second-stage gas, Mscf Stock-tank gas, Mscf Total plant products in primary separator gas, gal Ethane Propane Butanes (total) Pentanes plus Total plant products in second-stage gas, gal Ethane Propane Butanes (total) Pentanes plus Total plant products in well stream, gal Ethane Propane Butanes (total) Pentanes plus
Primary separator at 450 psig and ,!YF, second-stage

6.010 0

5.000 66.24 10.08 53.18 2.26 2.29

1,200 681.46 47.36 608.25 13.21 15.51 1,474 749 374 177 70 47 23 10 1,626 979 674 2,037

700 779.02 51.91 696.75 14.99 18.05 1,709 873 441 206 80 54 27 11 1,880 1,137 789 2,249

1.OOo 181.74 777.15 38.52 38.45 1,841 835 368 179 204 121 53 23 2,295 1,461 1,104 7,352

0 0 0 0 0 0 0 0 0 0 0 0

126 58 26 12 12 3

1,050
526 256 122 55 36 17

153 95 70 408

1,171 707 486 1,680

separator a, 100 ps,gand75OF,

stock tank a, 75DF

39-10

PETROLEUM ENGINEERING

HANDBOOK

These data may be calculated without the benefit of rock propertles or interstitial water values. The assumption is that the retrograde liquid does not achieve significant mobility. Because only one phase is flowing, water and hydrocarbon liquid saturations do not enter into the calculations. The assumption that the retrograde liquid does not flow in the reservoir except in the drawdown area immediately around the wellbore appears to be good. Only with very rich reservoirs does movement of retrograde liquid add significantly to well production. It was mentioned earlier that the most popular form of material balance on a gas-condensate reservoir is the p/zvs.-cumulative-production curve. It was stated that the z factor to be used must be the two-phase : factor. The cumulative production must be the total production from the well. This includes. in most instances, the first-stage separator gas, second-stage separator gas, tank vapors. and the vapor equivalent of the stock-tank liquid. The most accurate production figures from a gas-condensate field are usually the sales-gas volumes. This usually includes the first- and second-stage separator gas. To make the p/zvs.-cumulative plot, the tank vapors and the vapor equivalent of the stock liquid must be accounted for. Without the benefit of laboratory data, the tank vapors must be estimated and the vapor equivalent of the stock-tank liquid calculated with an average or estimated number. Table 39.7 furnishes the data to make these calculations. If sales gas is the primary- and second-stage gas, and the average reservoir pressure is 5.000 psig, then the total well-stream volume can be calculated by dividing the sales volume by 0.83704. This factor accounts for the tank vapors and the vapor equivalent of the tank liquid. If the sales gas is only the first-stage gas, then the appropriate factor would be 0.80285.

curacy) on the basis of the composition of the gas-condensate system. Whenever possible, the predictions should be made with actual laboratory data because the better accuracy obtained at the reservoir conditions is justified by the large gas and liquid reserves involved in reservoir calculations. Predictions With Laboratory-Derived and Hydrocarbon Analysis Data

With the assumption that the liquid condensate in the reservoir during a pressure-depletion operation stays in place (does not build up sufficiently to provide liquid-phase permeability for flow), reservoir behavior can be predicted from the laboratory constant-composition depletion study discussed previously. Pertinent information is shown in Tables 39.3 through 39.6 and Figs. 39.4 and 39.5. Liquid-phase change in the reservoir is shown in Fig. 39.5 derived from Table 39.5. Note that the amount of liquid remaining in the reservoir passes through a maximum but does not return to zero, indicating that pressuredepletion operations leave some liquid hydrocarbons behind at abandonment pressure. Economic analyses of pressure-depletion operations are necessary for estimating the magnitude of this loss and its effect on development and operating policy for the reservoir. The ultimate recoveries by pressure depletion of wet gas. condensate, and plant products can be calculated for the reservoir described in Table 39.8 by use of the data given in Table 39.6. Gas in place ut original pressure: (500x 106)(1.545)(178. l)= 137,582 MMscf. Gas in place at dewpoint pressure: (500x106)(1.471)(178.1)=130,992 MMscf.

Operation by Pressure Depletion


Pressure-depletion gas-condensate reservoir behavior can be predicted from the laboratory data described previously, or if necessary, by various correlation and computation procedures that provide similar information (with less ac-

Wet gas produced to dewpoint pressure: 137,582130,992=6,590 MMscf.

TABLE 39.7-CALCULATED

INSTANTANEOUS

RECOVERY DURING DEPLETION

Reservoir Pressure (asiai


6,010 Normal temperature separation Stock-tank liquid gravity at 6OOF. OAPl Separator-qaslwell-stream ratio, Mscf/MMscf primary-separator gas only primary and second-stage separator gases Separator-gas/stock-tank-liquid ratio, scf/STB primary-separator gas only pnmary and second-stage separator gases Recovery from smooth well stream compositions, gal/min Ethane plus Propane plus Butanes plus Pentanes plus
Primary separator at 450 ps~g and 75T second-stage separator

5,000 51.7 802.85 837.04 5,277 5.502 10.953 8.648 7.209 6.158

4.000 55.4 847.45 874.26 7,828 8.076 9.175 6.856 5.434 4.437

3,000 60.4 897.28 915.77 13,774 14,058 7.509 5.164 3.752 2.800

2,100 64 6 920.44 935.04 19,863 20.178 6.851 4.469 3.057 2.108

1,200 67.5 922.04 936.84 22,121 22.476 6.970 4.479 2.990 1.959

700 68.6 907.14 925.38 19,475 19.867 7.574 4.963 3.349 2.171

49.3 777.15 815.67 4,276 4,488 12.212 9.917 8.456 7.352

at 100 pslg and 75OF. stock tank at 75OF

GAS-CONDENSATE

RESERVOIRS

39-11

Wet gas produced Sfom dewpoint pressure to abandonment: (130,992)(0.77902)= 102,045 MMscf.

TABLE 39.8-FORMATION AND FLUID DATA FOR A GAS-CONDENSATE RESERVOIR Original reservoir pressure, psig Dewpoint pressure, psig Assumed abandonment pressure, psig Average reservoir temperature, OF Hydrocarbon pore space (by volumelrics), cu ft Gas expansion factor (8,) of produced fluid at original pressure, Mscflbbl Gas expansion factor (B,) of produced fluid at dewpoint, Mscf/bbl 7,000 6,010 700 256 500x 10 1.545 1.471

Total wet gas produced: 6,590+ 102,045 = 108,635 MMscf. Condensate produced to dewpoint pressure: (6,590)(181.74)=1,197,667 Condensate producedfiom donment: (130,992)(51.91)=6,799,795 Total condensate produced: 1,197,667+6,799,795=7,997,462 Percent recoveries by pressure depletion from dewpoint pressure to abandonment: 102,045 Wet gas= ~ x 100=77.9%; 130,992 Condensate = 6,799,795 181.74x 130,992 The total plant products can be calculated in a similar manner, depending on the flow streams to be processed and the recovery efficiencies anticipated. x 100=28.6%. bbl. dewpoint pressure to aban-

Predictions With Vapor/Liquid Calculation and Correlations

Equilibrium

In the absence of direct laboratory data on a specific gascondensate system, pressure-depletion behavior can be estimated with vapor/liquid equilibrium ratios (i.e., equilibrium constants, equilibrium factors or K values) to compute the phase behavior when the composition of the total gas-condensate system is known. Correlations for estimating phase volumes must also be available. When multicomponent hydrocarbon gases and liquids exist together under pressure, part of the lighter hydrocarbons (light ends) are dissolved in the liquid phase, and part of the heavier hydrocarbons (heavy ends) are vaporized in the gas phase. A convenient concept to describe the behavior of specific components quantitatively is the equilibrium ratio. The ratios vary considerably with the pressure, temperature, and composition of the system involved The equilibrium ratio is defined as the mole fraction of a given constituent in the vapor phase divided by the mole fraction of the same constituent in the liquid phase, the two phases existing in equilibrium with each other. The equilibrium ratio is designated as K. The basis for this definition is discussed in Chap. 23 and by Standing. 9 Fig. 23.21 illustrates the behavior of equilibrium ratios for a particular system and shows the rather wide variation possible for a given constituent at different pressures. The

figure shows a tendency of the equilibrium ratios to converge isothermally to a value of K= 1 at a specific pressure. The pressure is roperly called the apparent convergence pressure. g The selection of equilibriumratio values for calculations usually is based on the system s apparent convergence pressure, which can change in a pressure-depletion process because of changing system composition with pressure decline. Large inaccuracies can occur in pressure-depletion calculations with equilibrium ratios when the heavier hydrocarbons (e.g., heptanes and heavier) are not adequately described. To obtain satisfactory results in calculating pressure-depletion behavior of a gas-condensate system, an extended analysis of the CT+ fraction should be made. A determination of the the molar distribution of CT+ through at least C!z=,is recommended. As can be observed in Table 39.4, the CT+ component of the subject gas-condensate fluid exhibited a change in molecular weight from 158 at a pressure of 6,010 psig to 109 at a pressure of 700 psig. The change in density of the C 7 + component was from 0.827 to 0.778 over the same pressure range. Table 39.4 also shows that at 700 psig, the molecular weight of the CT+ in the liquid phase is 174, compared to 109 in the gas phase, and the density is 0.837 in the liquid phase, compared to 0.778 in the gas phase. This change in composition of the C7+ fraction with pressure reduction leads to large errors in the vapor/ liquid split of the CT+ fraction when equilibrium ratios are used and in the resultant molecular weight and density of the calculated gas and liquid volumes. Should such an extended analysis of the CT+ component not be available, then a statistical split should be made that maintains the integrity of the average molecular weight and density of the CT+ component. Once the CT+ component has been divided into multiple pseudocomponents, the physical properties required to make reservoir flash calculations must be developed. Wbitson30 presents a method for determining the molar distribution of single-carbon-number (SCN) groups that are defined by their boiling points as a function of each group s molecular weight. To make the distribution, a three-parameter gamma probability function is used. Whitson also presents equations for calculating the required physical properties with the Watson3 characterization factor. This method can be easily programmed for a personal computer and permits rapid development of molar distribution and physical properties. A statistical expansion of the C7+ component of the gas-condensate fluid presented in Table 39.2 has been made with the teehnique Whitson described. The results of this expansion

39-12

PETROLEUM ENGINEERING

HANDBOOK

are presented in Table 39.9. The ability to calculate accurately the pressure-depletion performance of a gascondensate reservoir depends on proper characterization of the vapor/liquid equilibrium ratios (K values) of the hydrocarbon system. Equilibrium ratios for nonhydrocarbon components and hydrocarbons C, throu h C 10 can be found in the Engineering Data Book. 15 Hoffman et al. 32 and Cook et al. 33 have presented methods for developing K values for the pseudocomponents. Hoffman et al. s procedure can be programmed easily for a personal computer for rapid development of equilibrium ratios. An alternative method is to plot the methane and normal pentane K values as a function of their boiling points on a semilog graph for each depletion pressure to be calculated. An equation can be determined for a straight line connecting these two points. The K value for each of the other components and pseudocomponents can then be calculated for each pressure point with their individual boiling points. This method of obtaining K values was used in the earlier example calculation. There are some limitations on the accuracy of the data derived by these methods unless some measured data on similar hydrocarbon systems are available. However, the data should be usable for the quick, rough approximations often needed in the preliminary reservoir evaluation stage. The C t through Cc composition of the gas-condensate fluid presented in Table 39.2 was used to develop a K-value relationship for the extended C7+ compositions. The resultant relationship is presented in Fig. 39.6. Chap. 23 describes the general techniques of the use of vapor/liquid equilibrium ratios to compute the phase compositions and magnitudes of hydrocarbon gas/liquid mixtures. Standing also has an excellent presentation of this usage, including a discussion of the serious errors that can result in calculating the phase behavior of gascondensate systems. When these methods are used to estimate the pressure-depletion behavior of a gas-condensate reservoir, the following procedure is used. 1. Assume that the original (known) composition flashes from original pressure (and volume) to a lower pressure, at which the compositions and amounts (in moles) of the liquid and gas phases are computed with the best K values available. 2. Estimate the volume of each phase with the methods discussed below. 3. Assume that enough vapor-phase volume is removed (produced) at constant pressure to cause the remaining gas plus all the liquid to conform to the reservoir s original constant volume.

BOILING POINT CONDENSATE NO7 FLUID

COUPONENT CO2 N w

BOILING POINT OR 275 HO MO 217 ,"d, 462 482

E: I% NC4

CT.

869

000,

200

400

Kc BOILING

803 POINT,'RANKlNE

1000

12M)

1403

Fig. 39.6-K-value

correlation for Condensate 7 depletion.

4. Subtract the number of moles of each component in the vapor represented by this gas removal from the original system composition. 5. With the new total composition from Step 4, consider the system flashed to the next lower pressure step and repeat the procedure. Removal of vapor phase alone is required by the assumption that fluid flowing into the wells will not be accompanied by any liquid phase at any step of the process. As indicated previously, the calculations require knowledge of the volume occupied by each phase at each pressure step. Methods to estimate these volumes are described in Chaps. 20 and 22 and also by Standing. 9 To estimate phase volumes, smoothed values should be used from curves drawn through the points computed from properties of the phase at each known composition.

TABLE 39.9-STATISTICAL

EXPANSION OF C,,

COMPONENT, CONDENSATE 7

C 7+ Mole fraction 0.0999 Molecular weight 158.0 Density, g/cm 0.827 Component C7 2 40 C ,I+ Mole Fraction 0.01685 0.01535 0.01235 0.00941 0.04594 Mole Weight 100.9 113.6 126.9 139.5 205.1 Density (g/cm3) 0.7486 0.7648 0.7813 0.7960 0.8641 Boiling Point (W 658 702 791 748 1,020

GAS-CONDENSATE

RESERVOIRS

39-13

These calculations are intended to approximate the experimental procedure used in the PVT cell during a laboratory pressure-depletion study. The number of pressure steps used in making such calculations is arbitrary but probably should conform to about SOO-psi intervals, with points usually closer together at the start and at the end of the calculations. The calculated depletion performance of Condensate 7 is presented in Table 39.10. The dewpoint pressure of 5,277 psig was calculated with an empirical relationship Nemeth and Kennedy j4 presented. The best method to determine the dewpoint pressure is by direct measurement, as in the laboratory PVT analysis. If these data are not available, then one must resort to estimation by empirical methods. such as that used in this example, or by gas/liquid production performance. In the latter choice, one must deplete the reservoir to a pressure below the dewpoint. In Table 39.10 a comparison of wet gas and condensate recoveries is made between the laboratory-measured and calculated depletion performance. As can be seen from the comparison, large errors are possible in the calculated data resulting from estimation of the dewpoint pressure and the physical properties of the reservoir fluid. Hydrocarbon/Liquid Condensation; Gas-Condensate Behavior Effect on

For some gas-condensate systems, large amounts of liquid can be condensed during pressure depletion, resulting in high liquid saturations in the formation pores. When this probability is indicated by either laboratory tests or calculations, the possibility of hydrocarbon/liquid flow through and out of the reservoir must be examined. Relative permeability information (usually curves showing k,/k, vs. liquid saturation in the formation) should be combined with viscosity data (pO/pR) to estimate the volumetric proportion of liquid in the flowing stream (thus removed from the reservoir), thereby affecting the remaining reservoir phase compositions at each of the depletion steps. The best k,gpu,/k,p., data to use are those determined in the laboratory with actual samples of the reservoir rock and hydrocarbon system in question. In the absence of such information, k,/k, can be estimated by the methods explained in Chap. 28; viscosity approximations may be made by the methods described by Carr et al. 23 After the amount of gas and liquid removed at each step has been estimated, the calculation procedures can be adjusted to obtain the desired behavior predictions. Pressure Drawdown at Wells; Effect on Well Productivity and Recovery The previous discussion has taken liquid condensation in the formation into account as though it occurred uniformly throughout the reservoir (uniform pressure at any instant of time). In low-permeability formations, however, there can be appreciable pressure drawdown at the producing wells because the pressures near the wellbores are much lower than in the main part of the reservoir. This tends to increase the early condensation of liquids around the wells considerably, thus decreasing the gas permeability and affecting the phase behavior of the system near the wells. This is important from at least two standpoints: (1) composition history of fluids produced from the reservoir may diverge from that predicted by assuming uniform pressure in the reservoir at any instant of time and

(2) adverse effects on the ability of the wells to produce may occur, potentially affecting the optimum well spacing and the rate of gas-condensate recovery from the zone as pressures decline. The effects of well-pressure drawdown on the composition history (and ultimate liquid recoveries) of gascondensate reservoir production have had little discussion in the literature. The general expectation would be that in lower-pressure areas around the wells, liquid hydrocarbons are precipitated earlier and in greater amounts than in the main volume of the reservoir. The main factors involved in this phenomenon are the richness of the gas condensate, the retrograde characteristics of the reservoir fluid, and the permeability of the reservoir rock. Normally, the area around the wellbore that is affected will be small and the condition will stabilize. Normal operating practices to restrict the pressure drawdown to reasonable values will alleviate the problem. In those reservoirs that exhibit extremely low permeability and contain fluids exhibiting condensable liquids of more than 200 bbl/MMscf, the problem can be severe. When separator samples are taken for the laboratory, the analysis procedure discussed previously should be followed to minimize the drawdown effect on the gas and liquid compositions. The effects on well productivity of precipitated liquid in the vicinity of the wellbore theoretically can be appreciable. Normally, estimates of future well productivity ignore the drawdown effects of production on liquidphase distribution in the reservoir. The greater liquid accumulations and lower gas permeabilities near the wells thus are ignored in theoretical predictions of well productivity (or extrapolations from early tests); these predictions then tend to show minimum decline rates. The operating engineer should be alert to this possibility whenever calculated well or reservoir rates approach undesirably close to the minimum necessary for the operating objectives of the project. Well productivity is discussed later. Relative Merits of Measured vs. Calculated Pressure-Depletion Behavior This chapter has emphasized that for purposes of reservoir analysis and prediction, measured properties and observed behavior of gas-condensate systems are much superior to the use of correlations or approximations. This applies in particular to the use of equilibrium ratios for simulating or predicting the pressure-depletion behavior of a reservoir. The problem is discussed and illustrated by Standing 9 in his Vapor Liquid Equilibria and GasCondensate Systems chapters. In particular, Standing s Fig. 36 shows that serious errors (in excess of 40%) can be incurred in the computation of the liquid volume of a gas-condensate system from errors of less than 10% in the equilibrium ratios for heptanes-plus and methane. The literature contains reports on the use of equilibrium ratios for calculating the reservoir behavior of gascondensate systems. Allen and Roe3 computed the pressure-depletion behavior of a gas-condensate reservoir and observed certain discrepancies with the actual behavior. These authors did not report laboratory-measured equilibrium ratios for the specific fluids involved, however; consequently, there were no means to compare computed fluid behavior with actual fluid behavior. All the observed discrepancies were assigned arbitrarily by Allen

39-14

PETROLEUM ENGINEERING

HANDBOOK

TABLE 39.10-CALCULATED

COMPOSITION OF PRODUCED STREAM, mol% Reservoir pressure (psig)

5,277 Carbon dioxide Nitrogen Methane Ethane Propane Iso-butane n-butane Iso-pentane n-pentane Hexanes Fraction C, Fraction C, Fraction C, Fraction C ,0 Fraction C , , + Heptanes-plus mol% molecular weight density Deviation factor, z equilibrium gas two-phase Gas FVF, Mscf/scf Retrograde liquid volume, % hydrocarbon pore space 0.01 Cl.11 68.93 8.63 5.34 1.15 2.33 0.93 0.85 1.73 1.685 1.535 1.235 0.941 4.594 100.000 9.990 156 0.825

5,000 0.01 0.11 70.74 8.67 5.28 1.12 2.26 0.89 0.81 1.62 1.55 1.38 1.09 0.81 3.66 100.00 8.49 155 0.822

4,000 0.01 0.13 74.77 a.77 5.13 1.06 2.10 0.79 0.71 1.35 1.21 1 .Ol 0.73 0.49 1.74 100.00 5.18 146 0.812

3.000 0.01 0.13 77.09 8.88 5.05 1.Ol 1.99 0.73 0.64 1.15 0.97 0.75 0.49 0.30 0.81 100.00 3.32 137 0.802

2,100 0.01 0.13 78.05 9.04 5.10 1.01 1.96 0.69 0.61 1.03 0.82 0.59 0.35 0.19 0.42 100.00 2.37 129 0.793

1,200 0.01 0.12 77.55 9.37 5.41 1.08 2.09 0.73 0.64 1.04 0.78 0.52 0.28 0.14 0.24 100.00 1.96 124 0.784

700 0.01 0.12 75.53 9.76 5.95 1.22 2.41 0.86 0.75 1.23 0.90 0.59 0.31 0.15 0.21 100.00 2.16 121 0.780 Trace 0.01 12.29 4.22 5.02 1.62 3.80 2.14 2.16 5.97 7.33 7.92 7.34 6.14 34.04 100.00 62.77 166 0.832

1.021 1.021 0.2561 0.000

0.987 1.009 0.2511 15.3

0.901 0.922 0.2201 26.96

0.861 0.845 0.1730 27.89

0.863 0.782 0.1211 26.43

0.899 0.695 0.0668 23.85

0.930 0.595 0.0380 21.95

Cumulative recovery per MMScf of original flurd Initial in place Well stream, Mscf Normal temperature separation * Stock-tank liquid, bbl Primary separator gas, Mscf Second-stage gas, Mscf Stock-tank gas, Mscf Total separator gas, Mscf 1.ooo Reservoir pressure (psig) 5.277 0.00 5,000 40.73 4,000 160.03 3,000 311.34 2,100 478.33 1,200 662.91 700 768.03

183.13 776.98 37.01 38.31 852.30

0.00 0.00 0.00 0.00 0.00

6.91 32.46 1.42 1.50 35.38

21.98 138.96 4.76 5.26 148.98

34.00 280.26 7.74 8.92 296.92

42.98 437.60 10.21 12.19 460.00

50.71 610.03 12.58 15.60 638.21

55.05 707.57 14.08 17.93 739.58

Comparison of Recovery Calculations Laboratory Depletion Gas in place at original pressure, MMscf Gas in place at dewpoint pressure, MMscf Wet gas produced to dewpornt pressure, MMscf Wet gas produced from dewpoint to abandonment, Total wet gas produced, MMscf Condensate produced to dewpoint pressure, bbl Condensate produced from dewpoint to abandonment, Total condensate produced, bbl bbl MMscf 137,582 130,992 6,590 102,045 108,635 1,197,667 5,297,156 6,494,823 Calculated Depletion 137,582 128,050 9,532 98,346 107,878 1,745,595 5,413,947 7,159,542

GAS-CONDENSATE

RESERVOIRS

39-15

and Roe to factors other than the possible inaccuracies of equilibrium ratios from correlations compared with actual measured ratios for the particular system composition and reservoir conditions involved. Some of these discrepancies were probably attributable to the equilibrium ratios used. Berrymanj6 compared calculated gas-condensate fluid performance with that actually obtained in the laboratory; however, he made observations on actual vapor/liquid equrlibrium in the laboratory cell and adjusted the literature equilibrium ratios to conform to this actual behavior. With the adjusted vapor/liquid equilibrium ratios, the calculated performance was found to match actual reservoir performance during early life satisfactorily. Rodgers ef ul. j7 provided detailed laboratory data, vapor/liquid equilibrium calculations, and actual reservoir performance for a small gas-condensate reservoir in Utah. The pressure range involved was moderate compared with most cases. Even at these moderate pressures, however, the literature-derived equilibrium ratios for heptanes-plus did not agree favorably with measured values for the system. The authors commented that the appearance of the data. clearly shows the need for improved techniques in establishing proper equilibrium data. On the basis of this experience and for the reasons Standing stated, it would appear desirable to use measured values of phase and volumetric behavior for a gas-condensate system in predicting the pressure-depletion behavior of a gas-condensate reservoir. As more data are obtained and better correlating methods developed, it is possible that equilibrium ratios may achieve suitable accuracy for reservoir-type calculations in the future. Numerous equation-of-state (EOS) calculation techniques have been developed that produce phase equilibrium data that can be used to perform depletion calculations for gascondensate reservoirs. Many are discussed in Refs. 38 through 40. The use of EOS methods, while more flexible and in many cases more accurate, requires sophisticated computer programs that may or may not be available or warranted. Continued improvement in techniques using EOS s may enhance the accuracy of calculated pressuredepletion performance.

Operation by Pressure Maintenance or Cycling


Pressure maintenance of a gas-condensate reservoir can exist by virtue of (1) an active water drive after moderate reduction of pressure from early production, (2) pressure maintenance through water injection operations, (3) injection of gas, or (4) combinations of all of these. From time to time, certain reservoirs may be encountered that have fluids near their critical points and that thereby may be candidates for special recovery methods, such as the injection of specially tailored gas compositions to provide miscibility and phase-change processes that could improve recovery efficiency. These usually are not regarded as gascondensate cases. Water Drive and Water Injection Pressure Maintenance Very few cases of gas-condensate reservoirs operated under natural water drive have been reported in the litera-

ture. To be attractive economically. a water drive would have to be sufficiently strong to maintain pressure high enough to minimize condensed hydrocarbon losses in the formation. Under these conditions, expenditures for cycling or other pressure-maintenance operations might not be justified: a careful engineering and economic analysis should be made if this possibility seems imminent. The analysis should include a geologic review of conditions surrounding the reservoir to estimate whether any indicated early water drive is apt to last for the life of the operation. There are also other considerations to be studied carefully. including the expenses of dewatering or working over invaded producing wells, the displacement efficiency of water moving gas. and the potential bypassing and loss of condensate fluids when wells become watered-out prematurely through permeable stringers [invasion efficiency (see Pages 39- 17 and 39- 18) of the natural flood]. Should this last possibility exist, use of a natural water drive would be of doubtful value if the amount of hydrocarbons in place is large. In any case, predictions of recovery by natural water drive should take into account the factors for water injection discussed below. The injection of water into a gas-condensate reservoir to maintain pressure is sometimes considered. A number of factors must be weighed carefully before a decision is reached. The mobility ratio (mobility of driving fluid over mobility of the driven fluid, water/gas) in this case is favorably low because of the very high mobility of the gas, thus tending to provide high areal sweep and pattern (@S-weighted) efficiencies. There is strong evidence, however. that displacement efficiency by the water is not high. While Buckley et al. 4 indicated that the displacement efficiency of water driving out gas can be as high as 80 to 85%, experiments and field observations by Geffen et al. indicate that it may be as low as 50%. This is offset to some extent by the improved area1 sweep efficiency enjoyed at a low mobility ratio. All things considered, the recovery of gas condensate in the vapor phase by water injection is likely to be appreciably lower than by cycling, and any consideration of water injection for gas-condensate recovery should be accompanied by detailed experimental work on cores from the specific reservoir involved. This will help to determine whether the water can, in fact, accomplish a high enough displacement efficiency to justify its use. Should water injection be decided on, gas and liquid recovery predictions for the reservoir can be made by combining the pattern (h&Gweighted). invasion, and displacement efficiencies with a knowledge of the condensable-liquids content of the gas-condensate system at the pressure chosen for pressure maintenance. As an example, an area1 sweep efficiency of 90% (based on an extremely low mobility ratio for water displacing gas) might be applied to the case cited on Page 39-24. Taking into account the thickness variations of the reservoir, this might provide a pattern (h&S-weighted) efficiency of about 95 % With an assumed invasion efficiency of 65 % within the invaded volume, water injection for this case would have swept out about 55% (product of the above three efhcienties) of the vapor phase in place at the start of injection. This compares with the actual recovery of more than 86% of the wet vapor by cycling operations. as discussed on Page 39-22.

39-I 6

PETROLEUM ENGINEERING

HANDBOOK

These estimates of possible gas recoveries by either a natural water drive or water injection can be affected materially by the permeability distribution in the reservoir. The presence of large differences in permeability will result in premature water breakthrough. Flowing gas wells tend to load up when producing water and, depending on the vertical flow velocity and bottomhole flowing pressure, may cease to flow. This inability to flow results from sufficient water dropping out in the tubing to form a hydrostatic water column that exerts a pressure equal to the bottomhole pressure. It is difficult to obtain economical flow rates by artificial lift. This loss of productivity may result in premature abandonment of the project. The problems would be particularly serious for deeper reservoirs where the cost of removing water would be a significant factor. Yuster4 discusses possible remedial methods for drowned gas wells. Bennett and AuvenshineM discuss dewatering gas wells. Dunning and Eakin4 describe an inexpensive method to remove water from drowned gas wells with foaming agents. Generally, the use of water injection for maintaining pressure in a gas-condensate reservoir is unlikely to be attractive where a wide range of permeabilities exists in a layered reservoir and selective breakthrough of water into producing wells might be expected before an appreciable fraction of the gas condensate in place could be recovered. Reservoir Cycling, Gas Injection Dry-Gas Injection. Comparative economics determines whether a gas-condensate reservoir should be produced by pressure depletion or by pressure maintenance. The objective of using dry-gas injection in gascondensate reservoirs is to maintain the reservoir pressure high enough (usually above or near the dewpoint) to minimize the amount of retrograde liquid condensation. Dry field gases are miscible with nearly all reservoir gascondensate systems: methane normally is the primary constituent of dry field gas. Dry-gas cycling of gas-condensate reservoirs is a special case of miscible-phase displacement of hydrocarbon fluids for improving recovery. Experimentation has shown that the displacement of one fluid by another that is miscible with it is highly efficient on a microscopic scale; usually the efficiency is considered 100% or very nearly so. This is one of the factors that explain the effectiveness and attractiveness of cycling. Another advantage of cycling is that it provides a means to obtain liquid recoveries from reservoirs at economical rates while at the same time avoiding waste of the produced gas when a market for that gas is not available; the operation provides at its termination a reservoir of dry gas with a potentially greater economic value. Inert-Gas Injection. The demand for dry gas as a marketable commodity varies, and the economic aspects of retaining dry cycled gas in reservoirs for future use have a changing significance. Most conservation laws in the U.S. still provide for minimizing waste of condensable liquids that would result if gas-condensate reservoirs were depleted through the retrograde range in a manner that left large liquid volumes unrecoverable. The use of inert gas to replace voidage during cycling of gas-condensate reservoirs can be an economical altemative to dry natural gas. One of the first successful inert-

gas injection projects was in 1949 at Elk Basin, WY,46 where stack gas from steam boilers was used for injection. In 1959, the first successful use of internal combustion engine exhaust was seen in a Louisiana oil field.47 The first use of pure cryogenic produced nitrogen to prevent the retrograde loss of liquids from a gas-condensate fluid was in the Wilcox 5 sand in the Fordoche field located in Pointe Coupee Parish, LA.48 In the Fordoche field, the nitrogen was used as makeup gas. The nitrogen amounted to about 30% of the natural-gas/nitrogen mixture injected. Moses and Wilson s49 studies confirmed that the mixing of nitrogen with a gas-condensate fluid elevated the dewpoint pressure. Moses and Wilson also presented data to show that the mixing of a lean gas with a rich-gas condensate would also result in a fluid with a higher dewpoint pressure. The increase in dewpoint pressure was greater with nitrogen than with the lean gas. In the same study, results are presented from slim-tube displacement tests of the same gas-condensate fluid both by pure nitrogen and by a lean gas. In both displacements, more than 98% recovery of reservoir liquid was achieved. These test results were also observed by Peterson, 5o who used gascap gas material from the Painter field located in southwest Wyoming. The authors concluded that the observed results were obtained because of multiple-contact miscibility. Cryogenic-produced nitrogen possesses many desirable physical properties. 5 Those that make nitrogen most useful for a cycling fluid are that it is totally inert (noncorrosive) and that it has a higher compressibility factor than lean gas (requires less volume). The latter advantage is partially offset by increased compression requirements when compared with lean gas. Until the mid 1970 s, most inert-gas injection consisted of injection of combustion or boiler gas into oil zones. The need for an alternative source of gas for gas-condensate-cycling projects emerged because of the high cost of hydrocarbon gas needed to replace reservoir voidage. The combustion and boiler gas that had been used to displace oil miscibly contains byproducts (CO. 02, HzO, and NO, +) that are highly corrosive5* and decrease cost effectiveness. Economic parameters used to evaluate any process are by their nature representative only under the general economic conditions during which they are prepared. Therefore, there will be no attempt here to present representative economic data. However, one should be cognizant of and take into account those variables peculiar to a particular process when applying current economic parameters to compare different processes. Many factors affect the economics of a gas-cycling project. The major factors are product prices, makeup gas costs, liquid content of reservoir gas, and degree of reservoir heterogeneity. When inert-gas injection is considered, some important additional factors should also be considered. Donohoe and Buchanan and Wilson have discussed these factors. The use of inert gas as a cycling fluid offers both advantages and disadvantages. The major advantages are that it permits early sale of residue gas and liquids, resulting in greater discounted net income and that a higher recovery of total hydrocarbons is achieved because the reservoir contains large volumes of nitrogen rather than hydrocarbon gas at abandonment.

GAS-CONDENSATE

RESERVOIRS

39-17

Offsetting these advantages are some disadvantages: production problems and increased operating costs caused by corrosion if combustion or flue gas is used as cycling fluid; possible additional capital investments to remove the inert gas from the sales gas, to pretreat before compression, and/or to fund reinjection facilities; and early breakthrough of inert gas caused by high degrees of heterogeneity in the reservoir, resulting in excessive operating costs to obtain marketable sales gas. All these factors should be evaluated properly when the depletion method is selected. Calculation of Cycling Performance. Methods of calculating reservoir performance under gas-cycling operations generally fall into one of two categories: feasibility and/or sensitivity analysis or detailed design and evaluation. The calculation method selected usually is determined after consideration of the quality and quantity of data available and the ultimate use of the engineering study. When the potential of a gas-condensate reservoir for cycling is first considered, it is generally desirable to make calculations that require the use of some reasonably simplifying assumptions. In this manner, relatively rapid and inexpensive results can be obtained that define the approximate cycling rate, cycling life, ultimate recovery, and profitability. If, at the conclusion of these studies, it appears that gas cycling is feasible, more detailed and exacting studies can be made with mathematical simulators to evaluate the earlier results and to design the most advantageous distribution of injection and producing wells. Efficiency and Effectiveness of Cycling. The principal factors determining reservoir cycling efficiency have been used with interchangeable labels and definitions in the literature. It is therefore necessary to define the various efficiencies clearly. The engineer should define and explain terms carefully when reporting estimates on gascondensate reservoir behavior. Reservoir Cycling Efficiency. ER is defined as the reservoir wet hydrocarbons recovered during cycling divided by the reservoir wet hydrocarbons in place in the productive volume of the reservoir at the start of cycling. Both figures must be computed at the same pressure and temperature; e.g., at reservoir conditions or at standard conditions. The reservoir cycling efficiency can be visualized as the product of three other efficiencies: pattern (h@S-weighted), invasion, and displacement. A fourth efficiency factor, area1 sweep, can be evaluated for various injection patterns using analog or mathematical models. All efficiency terms used (except displacement efficiency) must be identified as to time-i.e., time of dry-gas breakthrough into first producing well, time of breakthrough into last well, end of cycling, or other suitable designation. Area1 Sweep Efficiency. EA is the area enclosed by the leading edge of the dry-gas front (outer limit of injected gas) divided by the total area of reservoir that was productive at the start of cycling. (Black oil, if present, is usually excluded from these areas.) Area of sweep can be estimated closely from analog or mathematical model studies (discussed later) or by observing the locations of wells developing dry-gas content during actual operations. The area1 sweep efficiency depends primarily on the injection and production well patterns and rates and the lateral

homogeneity of the formations from a permeability and porosity standpoint. Lesser factors affecting areal sweep efficiency include variations in water content of the pores; time of operation of the compression plant in relation to the input capacities of the wells and their locations in the reservoir; the activity, if any, of a natural water drive; and the presence and handling of black-oil wells if an oil ring exists in the reservoir. Mathematical model techniques (Chap. 48) provide a useful means for predicting the areal sweep efficiencies of gas-condensate reservoirs and, simultaneously, the rate of frontal advance of the injected dry gas. For such studies, a reasonable amount of subsurface data is needed on sand characteristics, reservoir fluid properties, properties of injected fluid, and geologic description. Pattern (hcpS- Weighted) Efficiency. E,, is the hydrocarbon pore space enclosed by the projection (through full reservoir thickness) of the leading edge of the dry-gas front divided by the total productive hydrocarbon pore space of the reservoir at start of cycling. (Black oil, if present, is usually excluded from these volumes.) The hydrocarbon volume contained within the dry-gasfront projection can be determined by outlining the farthest-advanced position of the front (from model studies or field observations) on a hydrocarbon isovol map (isovol maps are developed from data on sand thickness, porosity, and interstitial water content), determining the hydrocarbon volume enclosed by this line, and comparing the volume with total reservoir productive hydrocarbon pore space. Note that the definition specifies projection of the leading edge and avoids stating whether either the entire gross or entire microscopic PV s are invaded or displaced by the injected gas. For the special cases in which productive thickness, porosity, interstitial water content, and effective permeability are each uniform, the pattern (h&S-weighted) and areal sweep efficiencies are the same. The pattern (&S-weighted) efficiency in general depends on the same factors discussed for areal sweep efficiency. Expected pattern (&S-weighted) efficiencies of nearly 95 % have been predicted under favorable conditions. Invasion Efficiency. El is the hydrocarbon pore space invaded (contacted or affected) by the injected gas divided by the hydrocarbon pore space enclosed by the projection (through full reservoir thickness) of the leading edge of the dry-gas front. (Sometimes volumetric sweep efficiency, E,, =E, X El, is used.) The definition says nothing about the effectiveness of the invading fluid in forcing original fluid out of the pores contacted. The term vertical sweep efficiency has sometimes been used in the sense of invasion efficiency. This is misleading in that it uses a one-dimensional term (vertical) when dealing with a three-dimensional problem. Invasion efficiencies can be as high as 90% under favorable conditions. However, invasion is affected significantly by large variations in reservoir flow properties, These might be strictly lateral variations in horizontal permeability (and to a lesser extent in porosity and interstitial water content) of a singlebed reservoir that does not have any variations vertically at any location; strictly layering effects by which the reservoir may comprise several strata, each relatively uniform in properties but differing appreciably in permeability from all the others; or combinations of these extreme cases. Performance of cycling operations can vary ap-

39-18

PETROLEUM ENGINEERING

HANDBOOK

TABLE 39.11 -EFFICIENCY


Areai Sweep Efhoency Area enclosed by leadtng edge 01 ~n,ected-gas (dryugas, lronl dlwded by total area of re*erYoll r,rodctlve at 51111 01 Pattern IhoS-weIghted) Eil~ciency

TERMS USED IN RESERVOIR CYCLING OPERATIONS


lnvas~on Elflcency Hydrocarbon pore space invaded by (contacted Or affected by) dry gas dlwded by hydrocarbon pore S!XXX enclosed by Itw pro,ecmn (Ihrouqh full leservolr Displacement Efflclency Reservar Cycling Efflcencv

Hydrocarbon pore space enclosed by Ihe pro,ectlo jrhrough full resewxr Ihlckness) of leadmg edge of drygas front diwded bv total

volume Wet hydrocarbon swepl out of lndlvldual pores or Small groups 01 pores dwded bv

Reservmr we, hydrocarbons recovered drlg Cycling dwded by resewo~r we, hvdrocarbons I place at starI ai cycl,ng (calculated at same temperature and pressure)

sweep efllciency IReI 5 pages 657 77, and 777 Ret 51 Pages 246 and 247 and Rel 13 Pages 308-09)

sweep elficlency iReI 5. Pages 755 763 and 770 and Ret 13 pages 40s09) .

Elf,c~ency caused by permeab!My stral!flcatlon IRet 13. pages 408-09)

Displacement etficlency (Rel 56. Pages 130 and 136 and Ret 13 Pages 408-09)

sweep pages

ehxncy IRet 5 612 771. and 7881

Flood efiumcy (Rel Pages 358 and 374)

59

Conformance Flood coverage IRel 59 pages 358 and 374, Sweeping (&I 57)

laclot elhclency

lRel

56

Pages

130 and 136) factor (Ref D6placement Page 110) l&f 61

Conformance 571

Pattern elilclency (M 60 pages 63 64. 98 and 99 and Rel 54 Page 77)

Flushing elf,cencv (Ref 4 1, Pages 246 and 247)

preciably according to what combination of the two extremes may exist for a given reservoir. Mathematical models can handle reservoir heterogeneities, both horizontally and vertically, if the data are available. Maximum use of core analysis data, pressure buildup and drawdown analysis, and detailed analysis of downhole logs is required to ensure an accurate evaluation of a reservoir s potential as a cycling project. Displacement Efficiency. ED is the volume of wet hydrocarbons swept out of individual pores or small groups of pores divided by the volume of hydrocarbons in the same pores at the start of cycling; note that both volumes must be calculated at the same conditions of pressure and temperature. This term is used here because it has received wide acceptance in the literature (on immiscible as well as miscible processes) for the microscopic displacement of fluids. Displacement efficiency is controlled mainly by the miscibility of the driving and driven fluids and their mobilities. For a cycling operation in which the pressure is being maintained at or above the dewpoint, the displacement efficiency resulting from action of the dry gas against the wet-gas phase in the individual pores will be virtually 100% because of nearcomplete miscibility and the near-identical mobility ratios of the two fluids. If the pressure is well below the dewpoint, the displacement efficiency will be less than 100% because of the immobility of the condensed liquid and incompleteness of revaporization of the dry gas. Evaluation

of a case of this type requires trial calculations of vapor/liquid equilibrium to estimate the extent to which dry gas coming into contact with the condensed liquid would revaporize some of the components and carry them toward the producing wells. Thus the reservoir cycling efficiency is the product of the pattern (&S-weighted), invasion, and displacement efficiencies, as summarized in Table 39.11, along with the previous discussion, and usage of terms appearing in some of the literature. Permeability Distribution. Permeability variation, both laterally and vertically, can have a strong influence on recoveries by cycling. Vertical stratification of horizontal permeability is probably the primary factor controlling invasion efficiency. In reservoirs containing layers or regions of contrasting permeabilities, the leading edge of the dry-gas front (used in calculating invasion efficiency) is at a different position for each layer. Field observations usually establish the front on the basis of breakthrough in the most-permeable layer, whereas mathematical model studies may have been based on an average permeability of layers or a discrete number of layers. thus predicting later breakthrough. This possibility should be understood when model predictions of breakthrough time are compared with field observations. Detailed reservoir analysis is required in developing a mathematical model to ensure that the model used adequately reflects the properties of the reservoir.

GAS-CONDENSATE

RESERVOIRS

39-19

TABLE 39.12-CALCULATIONS ILLUSTRATING THE DILUTION CAUSED BY WEIGHTED-AVERAGE PERMEABILITY PROFILE-BASED ON 16 WELLS (COTTON VALLEY BODCAW GAS-CONDENSATE RESERVOIR)

1866 ,860 1855 1825 I8 10

14 64 37 20 74 20 36 50 18100

77 4 78 9 80 4 84 3 86 2 88 2 90 4 1000 105 7

15 39 19 20 22 96 57 35 33 37 38 45 45 52 106 67 150 2173 239 0

3 14 6 16 1195 14 78 25 83 28 53 33 82 36 43 48 50 40 76 55 88

1 23 1 36 1 19 76 3 77 4 78 4 80 9 81 9 83 0 84 0 88 3 90 8 91 9 92 9 93 7 94 5 96 2 96 8 96 3 97 2 97 6 97 8 99 4 1000

89 30 95 20 10280 10740 11090 1,430 118 10 122 20 126 70 131 60 139 60 148 30 159 20 275 30 503 50

55 40 57 41 66 84 70 49 77 54 79 24 84 15 87 29 90 31 91 76 94 40 98 17 99 30 99 74 10000

22 46 20 76 15 85 12 71 9 69 8 24 5 60 1 88 0 70 0 26 0 00

0 0 0 1 0

89 38 27 52 62

There can be several sources of comparative permeability information for reservoir layers, including direct measurements of permeabilities on cores removed from wells, formation tests during drilling and completion, comparative transmissibilities from carefully run injection profiles, and flow, drawdown, and buildup tests on wells completed in different layers. If different kinds of information are to be used together, they should all be adjusted to the same units for calculating the effects of permeability variation on gas-condensate reservoir performance. Much discussion has been published regarding the effects of permeability variation on the recoveries of hydrocarbons from reservoirs. Discussions with particular reference to as-condensate reservoirs have been provided by Muskat, B +I Standing et al., 65 Miller and Lents, 66 and others. 67-70 Generally, the proposals to account for the effect of permeability variations on gas-condensate reservoir performance use two different methods of wellto-well averaging of horizontal permeabilities. The first method averages all high permeabilities from all wells together (irrespective of vertical positions of the highpermeability samples in the section) and all low permeabilities from all wells in another group, with intermediate permeabilities classified into one or more subgroups. Each of the average permeabilities is regarded as a single stratum continuous throughout the reservoir. This type of averaging would appear to give maximum probability of computed early breakthroughs of dry gas to producing wells. In the second method, permeabilities are averaged from well to well according to vertical position in the sec-

tion. For example, permeabilities in the top 10% of each well s productive section might all be averaged together, the next 10% together, and so on to the bottom. This procedure maintains layers in their relative vertical positions in the reservoir, and thus, by averaging laterally, the effects of any individual high-permeability samples tend to be damped out unless high-permeability streaks are actually persistent in one or more layers of the section. Either of these methods can be used in solutions presented by Muskat, 5XA who used the stratification ratio to develop mathematical means of evaluating the effects of vertical variation of permeability on cycling. Stratification ratio is the ratio of the permeability of the mostpermeable recognizable layer in the section to that of the least-permeable layer in the same section (these permeabilities are the layer average in each case, determined by whatever means, rather than individual high or low permeabilities from single plugs or cores from the layer). The Muskat development also assumes simple parallel superposition of layers of different horizontal permeabilities with no crossflow between. The resultant correlations are presented graphically in the references. Miller and Lents66 used the second type of lateral permeability averaging in their analysis of the Cotton Valley Bodcaw reservoir. Their work should be reviewed for an understanding of the detailed procedure used. The table of permeabilities they developed (rearranged in descending order of magnitude) for illustrating the calculation of dilution behavior of the subject reservoir with time is shown here as Table 39.12. The calculation assumes no

39-20

PETROLEUM ENGINEERING

HANDBOOK

sweep is sufficiently great in length. Few reservoirs conform to a parallel deposition of lens, each of different uniform permeability, unless one wishes to subscribe to the worst possible consequences for cycling, which can condemn the application of such a program in a rich gas-condensate field. Such unpublished information as has come to our attention tends to substantiate the belief that most reservoirs are not composed of continuous layers of contrasting pcrmeabilities (with no crossflow) that would tend to produce quick breakthrough during injection operations. Hurst s viewpoint should therefore be considered seriously by the engineer predicting the behavior of cycling projects, because overemphasis on the permeability variation within a reservoir could produce too pessimistic a view of possible recoveries and thereby condemn cycling in gascondensate reservoirs that might, in fact, yield profitable cycling performance. The second method for lateral averaging of permeabilities is recommended, whether the Miller and Lents66 analysis or other techniques are applied to the handling of permeability variation in gas-condensate reservoirs. Proper consideration for pattern (&S-weighted) efficiency must be given in each case. Prediction of Cycling Operations with Model StudiesAnalog Techniques. The steady-state flow of fluids through porous media, when governed by Darcy s law, is analogous to the flow of current through an electrical conductor governed by Ohm s law. Thus steady-state electrical-model studies have been used quite successfully in the prediction of gas-condensate cycling operations. The fundamental analogy between an electrical model of a gas-condensate reservoir and the flow system of the reservoir depends on the equivalence of electrical charge to reservoir fluid, current flow to fluid flow, specific conductivity to fluid mobility, and potential (voltage) distribution in the model to a function ap, (not to pressure distribution in the reservoir, as in an oil/water system) defined by Muskat as

Fig. 39.7~-Boundary of invaded area predicted by early potentiometric model studies.

crossflow, and the reservoir is treated as though it were composed of alternating layers of variable porosity and permeability. It is also assumed that parallel flow occurs simultaneously in the various layers with the same potential distribution throughout the layers. The injection wells are treated as a line source, and the producing wells as a line sink. Hence, the calculations in the table predict the percentage of original reservoir hydrocarbon volume at constant pressure produced at the instant each layer has been displaced and the percentage of dry gas (and wet gas) in the producing stream as more and more layers are displaced (breakthrough). The recovery to any stage of dilution in the produced gas can then be predicted; the recovery Miller and Lents calculated (supported by later production history, as shown by Brinkley ss5 Fig. 7) is in good agreement with predictions from Muskat s correlations. Very little has been published comparing the actual behavior and final recoveries of gas-condensate reservoirs with those predicted with the different methods of accounting for permeability variation. Stelzer63 reports on the performance of the Paluxy gas-condensate reservoir of the Chapel Hill field, TX, the cycling behavior of which had been predicted earlier by Marshall and Oliver. 58 This analysis is discussed further later. In a discussion of Stelzer s paper, Hurst takes the position that permeability variation or stratification in a reservoir can be of minor significance in controlling the ultimate recovery by cycling: The lithological nature of a reservoir is such that with the interspersion of shale throughout, it can virtually reproduce the configuration of a uniform sand if the

where pg = gas density, px = gas viscosity, and p = pressure. This analogy holds, provided the sources, sinks, and boundary conditions are made equivalent in shape and distribution. Steady-state models can be divided into two general classes: electronic and electrolytic. The former depends on the movement of electrons through resistive solids, such as metal sheets, carbon paper, and graphiteimpregnated cloth or rubber sheeting. Electrons are introduced at one boundary and move into the model to displace free electrons throughout the entire body of the model. The electrons moving out of the model at the other boundary produce a current that causes a potential drop in the solid resistive medium in accordance with Ohm s

GAS-CONDENSATE

RESERVOIRS

law. As a result, the movement of the equivalent fluid interface can be traced. In the case of a graphiteimpregnated cloth model, the reservoir is represented by layers of cloth, the number of layers of which are some function of the permeability/net-thickness product (kh) of the producing strata. The shape of each layer of cloth conforms to the shape of the kh range it represents. Copper electrodes are fixed in the cloth model at positions corresponding to the wells in the reservoir and direct currents are passed through these electrodes in proportion to the well flow rates. The electrodes are not usually scaled to the actual well diameters. Electrolytic models depend on the mobility of the ions in the medium. Because the velocity of an ion in an electrolyte system is proportional to the potential gradient, just as the velocity of a liquid particle in a porous medium is proportional to the pressure gradient, an electrolytic model can be set up that provides a good analogy to singlephase flow in a porous system. The ions are moved into the model across one or more boundaries and displace ions throughout the entire medium, causing ions to leave through other boundaries. The flowing current and potential drop are established in exactly the same way as in the electronic models. Electrolytic models can be divided into three major types: gel, blotter, and liquid. Although the first two types can be used to determine the area1 sweep patterns in twodimensional uniform media, the potentiometric model that uses a liquid electrolyte is the most flexible and accurate. In this type, the fluid conductivity of the porous medium is usually represented by an open container that has its bottom shaped to produce electrolyte depths proportional to the kh of the producing strata and its sides shaped to conform to the productive limits of the strata. This construction implies that there is no vertical variation in permeability and no bedding at any location in the reservoir, as represented by the model. Copper electrodes (not scaled to well diameter) are fixed in the model at positions corresponding to the locations of the wells in the reservoir, and alternating currents of proper phase are passed through these electrodes. The magnitudes of these currents are made proportional to the production and injection rates to be used in the reservoir. The direction of current flow at every point in the model is considered analogous to the direction taken by the flowing fluid in the reservoir. The general assumptions applicable to steady-state analog techniques are that (1) a vertical and discrete interface exists between the displacing and the displaced phases; (2) because the history of advance of only one front can be traced at any one time, if two interfaces or fronts are present (such as gas/gas and gas/water), one is considered a stationary boundary; (3) average reservoir pressure is constant regardless of the injection or production schedule (this avoids compressibility effects in the model study); and (4) gravitational effects are neglected. In addition, if the mobility ratio of the system is not (near) unity or infinity, the necessary procedures become tedious and costly. An example case history by Marshall and Oliver5* reported results of a potentiometric model study of the Paluxy sand reservoir of the Chapel Hill field. Smith County, TX. This gas-condensate reservoir is bounded on the north by a gas/water contact, on the west by a fault, and on the south and east by a pinchout. It was assumed

A B C D-W E F G HI -C. ---

I, WALTON #I (INJ) I. WALTON #Z (INJ) W. WALTON 8 #I WALTON #I S. WALTON # I H CAMPBELL #I B MOSLEY #I M WARREN #l-A G FINCH #I PHASE PHASF I II
III

----PHASE

Fig. 39.8-Boundaries of invaded areas predicted by later potenliometric model studies.

that the gas/water contact was a fixed impermeable boundary; that the permeability, porosity, and interstitial water content were each uniform throughout the producing zone; that the reservoir volume rate of dry-gas injection was equal to the corresponding rate of gas-condensate production; and that gravity effects were negligible. Fig. 39.7 shows the final dry-gas/wet-gas interface position at time of breakthrough into Well 1 (determined after several trials of well arrangement and production- and injectionrate schedules) that yielded an optimum pattern (h&Tweighted) efficiency prediction of 83 %. Injection was into Wells 1 and A with production from Wells 2 through 4 and B as indicated in Fig. 39.7. This program provided a sustained capacity of 35 MMscf/D for the life of the operation. Stelzer63 reported a comparison of model study predictions with actual performance for this reservoir. Actual gas injection was begun in accordance with the north/ south sweep indicated by the model study. During the initial period (first 15 months after cycling began) the production- and injection-rate program predicted by the initial model study was followed quite closely. New structural data revealed in the drilling of additional wells, however, required some changes in the isopach map of the Paluxy sand. The results from a second model study, which incorporated these changes plus injection into only Wells A and B, are shown in Fig. 39.8. Three interface boundaries (dry-gas fronts) are shown for three

39-22

PETROLEUM ENGINEERING

HANDBOOK

ND OF PHASE IO SAME PHASES 0 0 1 IO I 20 30

l (ADUSTED RATES

TO AS,
I

INJECTION

II AND III1

40

50

60

70

ACTUAL RESERVOIR OPERATING TIMEMONTHS AFTER START OF CYCLING

Fig. 39.9-Comparison of predicted with actual times of first drygas breakthrough, Paluxy gas-condensate reservoir, Chapel Hill field. TX.

indicates an invasion efficiency a little greater than 90% and implies that more complete invasion of lowpermeability regions behind the dry-gas front was accomplished during the later stages of cycling. The agreement of predicted breakthrough times within 10% of actual breakthrough times illustrates the great utility of potentiometric models in planning cycling operations. Small further improvement in the pattern (k&S-weighted) and invasion efficiencies was to be expected before abandonment of the reservoir in this case. Stelzer s63 figures (at the start of cycling) of 78.4 Bscf of gas in place and 74 bbl of condensable liquids in the vapor phase of the reservoir per 1 MMscf of gas indicate that 5,800,OOO bbl of condensable liquids is in the reservoir vapor phase at the start of cycling. Using the modelderived pattern (/#-weighted) efficiency of 88% (end of Schedule 3), 5,100,OOO bbl of liquids was subject to removal by dry-gas invasion. Stelzer s Fig. 5 shows that about 4,640,OOO bbl of liquid products were recovered between the start of cycling and the breakthrough of gas at Well 1. This provides an invasion efficiency of 91% at that time, based on 100% displacement efficiency. Thus the product of the pattern (k&Y-weighted) and invasion efficiencies represents a reservoir cycling efficiency of 80% at the time of breakthrough into Well 1. In addition, later operations increased the cumulative recovery during cycling to more than 5 million bbl of condensable liquids, thus bringing final reservoir cycling efficiency to more than 86 % This is considered very good. Prediction of Cycling Operations With Mathematical Reservoir Simulators. The use of mathematical reservoir simulators to calculate reservoir performance during gascycling operations yields results superior to those obtained by the more simplified calculation procedures. Use of these simulators removes the necessity of making the assumptions required in an analog model. Some assumptions are required, however, which should be understood to perform a reservoir simulation study properly. The theory of reservoir simulation is presented in Chap. 48. Coats7 presents a good discussion of reservoir simulation studies of gas-condensate reservoirs. One must keep in mind that the results from a mathematical reservoir simulator depend on the quality of the data used to prepare the reservoir model. If good data are not available, one should consider whether the expense and time required to perform a mathematical reservoir simulation are justified. Data Requirements for Gas-Condensate Cycling Study. To evaluate properly the potential of cycling a gas-condensate reservoir, the following data are required. 1. Geologic data-maps and cross sections showing net effective sand thickness, structural contours on the top and base of the productive formation, location of gas/ water interface originally and at the date the model study begins, and location of dry-gas/wet-gas interface at the start of study-and general information on lithology and lenticularity of the productive strata, such as extent of fissures, fractures, caverns, and other special conditions. If a black-oil ring is present, its size and extent should be shown. 2. Physical properties of the reservoir rock-isoporosity map (or average porosity), effective or specific isopermeability map (or average values), and interstitial water content.

production- and injection-rate schedules. The first schedule was maintained for the first 15 months of cycling; the second was continued until breakthrough of dry gas into Well E; the final schedule was maintained until first breakthrough at Well 1. There was close agreement between the model rates used and actual reservoir rates. The second model study indicated a pattern (h&G weighted) efficiency of 88 % , a 5 % increase over that obtained by the initial study. Stelzer estimated the amount of reservoir gas in place at start of cycling to be 78.4 Bscf. The new model study thus implies an additional 4 Bscf of predicted recoverable gas as a result of better reservoir definition and better operating schedules. The data in Fig. 39.9 compare model (predicted) breakthrough times with the actual times to dry-gas appearance in corresponding field wells. (Phases 1, 2, and 3 of actual behavior correspond to Schedules 1, 2, and 3 of the model study.) Field data on breakthrough were taken from breaks in content curves of isobutanes-plus; the dashed line shows the cumulative well-by-well breakthrough behavior of the dry-gas flood. Because predicted and actual injection and production rates were nearly equal and constant during the period shown (except for Phase 1, which was adjusted to the same average rates), times on the plot are directly proportional to cumulative reservoir volumes of gas. Therefore, the lower light line represents a hypothetical invasion efficiency of 100% that would prevail if actual breakthrough times coincided with those predicted by the model [and the area1 and pattern @@S-weighted) sweeps were identical with model predictions]. The upper light line represents an arbitrary invasion efficiency of 80% [assuming that predicted and actual pattern (h&-weighted) efficiencies are identical]. The straight heavy line from the origin through the last well to experience breakthrough

GAS-CONDENSATE

RESERVOIRS

39-23

3. Fluid characteristics (produced, and injected where applicable)-fluid composition. retrograde dewpoint pressure of reservoir fluids, gas FVF or specific volume vs. pressure, deviation factor, condensate content of reservoir fluid. viscosity, and densities of liquid and gas phases, all from original reservoir pressure through the range of interest (usually to abandonment conditions). 4. Amount of original fluids in place (derivable from data in Points I through 3). 5. Reservoir pressure history (volumetrically weighted) from discovery to present. If this is not available, isobaric contour maps at the various pressure survey dates should be supplied. 6. Condensate. gas, and water production data, from the date of discovery. 7. Proposed future production rates. 8. Gas- and/or water-injection data, past and future projections. 9. Productivity, injectivity, and backpressure test data on wells. Ultimate Recovery of Gas and Condensate Liquids by Cycling. The same reservoir for which pressure-depletion calculations were made previously can be used to illustrate the effectiveness of a cycling operation. Table 39.8 lists the basic data for predicting the ultimate recoveries of wet gas, condensate, and plant products during cycling at original reservoir pressure (to avoid serious drawdown effects) followed by pressure depletion to abandonment pressure. Productive thickness, porosity. and interstitial water content are each assumed uniform. Consequently, the 79.0% areal sweep efficiency obtained by a potentiometric model study is also the pattern (@S-weighted) efficiency. The invasion efficiency is assumed to be 90% because permeability variations are moderate. Because a dry-gas/wet-gas cycling operation is a miscible flood, the displacement efticiency is essentially 100%. Therefore, the reservoir cycling efficiency would be 7 1.1%. To simplify the example. it is assumed that after cycling, the unswept pore space both inside and outside the dry-gas front will pressure deplete in the same manner as predicted previously for the noncycling case: it will also be assumed economical to recover the butanes-plus from the gas produced.

Total separator gas produced (see Table 39.6): During cycling, 777.15+38.52+38.45 l,ooO =0.85412x93,135=79,548 During depletion, 696.75+ 14.99+18.05 1,000 =0,72979x29,491 Total : 79,548+21,522= 101,070 MMscf. x29,491 MMscf.

x93,135

=21,522

MMscf.

Total condensate produced: During cycling, 181.74x93,135=16,926,355 During depletion, 51.91 x29,491 = 1,530,878 bbl Total: 16,926,355 + 1,530,878 = 18,457,233 bbl These figures represent a significant improvement over the recoveries previously estimated for pressure-depletion alone. Noninjection-Gas Requirements in Cycling Operations. The noninjection-gas requirements for cycling can affect the amount of gas available for injection. The amount of gas to be cycled is determined by the optimum pressure level to be maintained and the efficiency of reservoir fluid recovery to be achieved; the amount of gas readily available, including sources and costs; and the design and operating programs for surface facilities. The amount of gas that is economical to cycle through a gas-condensate reservoir varies with many factors, including richness of the vapor at reservoir cycling pressure, size and cost of the plant, and the price of the field products and of dry gas. Miller and Lents% expected to cycle the equivalent of about 115 % of the gas in place to recover some 85 % of the wet-gas reserve of the Cotton Valley Bodcaw reservoir. While Brinkley 55 indicated cycling-gas volumes of as much as 130% of original wet gas in place for various reservoirs, no general correlation has been prebented on the amount of gas that is economically sound to cycle; this should be the subject of a detailed engineering analysis in each case. The makeup gas needed for constant-pressure cycling is mainly the volume required to replace shrinkage by liquid recovery and the amount consumed bbl.

Reservoir Mvt gas produced during cycling period (original reservoir comnposition): 130.992x0.711=93.135 MMscf.

Reser\vir wet gas produced by pressure depletim ufter cycling (changing cornposition, as shown in pressuredepletion example): 102,045x(1,000-0.711)=29,491 MMscf

Resertjoir tvet gas produced at ahundomnentpressure, 700 p.sig. 93.135+29,491= 122.626 MMscf.

39-24

PETROLEUM

ENGINEERING

HANDBOOK

for various fuel needs. For some composition, temperature, and pressure ranges, the removal of high-molecularweight constituents from the produced wet gas may result in a higher compressibility factor for the injected dry gas; hence, the greater volume per mole injected may require little or no makeup gas for constant-pressure cycling. The amount of gas not available for injection because of consumption for operating needs should be taken into account in determining makeup gas requirements if pressure is to be maintained. The amount of fuel for compression and treatment plants depends mainly on the total amount of gas to be returned to the reservoir and the discharge pressure for the plant. Discharge pressure, in turn, depends on the total rate of injection demanded and the number of injection wells and their intake capacities throughout the life of the operation. Other factors affecting the amount of gas required for overall operations are type of plant, type of liquid-recovery system used, and auxiliary field requirements (such as for drilling. completion, and well testing; camp fuel and power for maintenance shops, general service facilities, employee housing; and other factors that vary from one case to another). Moores4 reports that gas fuel consumption for the compression plant alone varies from 7 to 12 cu ft/bhp-hr; this is probably for gases with heat values of about 1,000 Btu/cu ft. Horsepower requirements per million standard cubic feet of gas compressed per day are correlated in Ref. 16 (Compressor section). An example based on Refs. 16 and 52 shows that, with 8 cu ft/bhp-hr, a compression ratio of 15.0 (compressing from, say, 461 to 7,000 psia) requiring three stages of compression with a ratio per stage of 2.47, and a specificheat ratio of 1.25, the cubic feet of compressor fuel used per million cubic feet of gas compressed can be calculated as follows. For a gas of 0.65 specific gravity and a stage compression ratio of 2.5. the chart in Ref. 16 reads 22 bhp. The allowance factor for interstage pressure drop (three compression stages) is 1.1. Fuel used per million cubic feet of gas compressed = bhp x cu ft of fuel/bhp-hr x ratio/stage x number of stages x allowance factor. Or compressor fuel consumption is m,.=22x8x24x2.47x3x1.1=34.4 MscfiMMscf.

resulting from imperfect seals or corrosion in well tubings, casings, and cement jobs. Remedial workover operations should be planned immediately when there is evidence of appreciable loss of gas between the compression plant and the reservoir sandface or between the outflow-well sandface and the plant intake. Combination Recovery Procedures

Partial water drive-conditions of natural water influx at rates too low to maintain pressure completely at the desired production rates-can exist for gas-condensate reservoirs. In such cases, operation may be by partial water drive and depletion, supplemental water injection, or partial water drive and cycling. Prediction of reservoir behavior and recovery under these conditions requires knowledge or assumptions about the aquifer and the water drive it supplies. This information can be deduced from a study of geologic conditions and early producing history of the reservoir; sometimes the deductions are accurate, sometimes not. Projections of water drive magnitude into the future at selected reservoir pressure levels can be made by methods developed in Refs. 72 and 73. If sufficient early producing history of a reservoir is available, it can usually be matched (simulated) by a mathematical reservoir simulation study. The future behavior of the reservoir can then be predicted under the following producing methods: (1) producing history and ultimate recovery of gas and liquids under partial water drive and pressure depletion at the selected production rate; (2) amount of supplemental water injection required to maintain reservoir pressure fully at the selected pressure level and production rate; and (3) size of cycling plant required to maintain pressure at the selected pressure level and production rate.

General Operating Problems: Well Characteristics and Requirements


As with any complex operation, gas-condensate recovery projects have many operating problems. Those pertaining to the plant, lines, and other surface facilities are best left to experienced plant and maintenance personnel, except as they affect reservoir operation (e.g., compressor-oil or corrosion-products carry-over into wells). Operating difficulties occurring at and below the wellhead are often concerns of the reservoir engineer and have an important bearing on the effectiveness of reservoir operation, whether by pressure depletion or by pressure maintenance. Among these are the maintenance of injection and production wells in good mechanical condition, the protection of wells against excessive corrosion, the general maintenance of well injectivity and well productivity (which are often interrelated), and the formation of hydrates that can interfere with the general injection and/or production operation. Well Productivity and Testing It is essential to maintain the producing capacities of gascondensate wells above minimum levels for good economic performance. Much has been written about the productivities of gas and gas-condensate wells, their general producing characteristics, and the optimum methods for testing and reporting their productivities. Loss of productivity of gas-condensate wells can occur from reservoir

This compares favorably with the factor presented in Moore s54 Fig. 8. For an example reservoir originally containing 130,992 MMscf of wet gas, which might be cycled the equivalent of 1 l/4times, the approximate compressor fuel consumption would be 130.992x 1.25x34.4=5.633 MMscf.

This is approximately 3 % of the gas handled through the plant. Treatment plant fuel and other plant needs added to compressor fuel bring the range of consumption inside the plant fence to 3 to 7 % of the gas handled by a cycling plant. In addition to these needs and others mentioned earlier, possible gas losses can occur in a cycling operation: gas used in blowing down wells, should this be necessary for cleaning or treating purposes; small gas leaks at compressor plants and in field lines; and gas leaks

GAS-CONDENSATE

RESERVOIRS

39-25

pressure decline (including possible effects from condensation of liquids in the reservoir and consequent reduction of effective gas permeability), from the invasion of water into producing wells, from solid precipitates in the pore space, from formation damage during well killing or workover operations, and from mechanical failure of downhole equipment. The engineer must have indices at his disposal that show the productivity histories of wells and whether productivity decline is excessive for prevailing producing conditions. Productivity Testing. In making productivity tests on wells, orderly well-conditioning and overall test procedures should be used. as suggested in Chap. 33 or in standards recommended by Texas, 26 New Mexico, Kansas, 28 and the Interstate Oil Compact Commission. 2y It is common to use wellhead pressures in determining well productivity (or injectivity) characteristics with arbitrary correction procedures for estimating BHP s from the observed surface pressures. No fully satisfactory methods have been devised for making accurate estimates of gas-condensate well BHP s, either static or flowing. Calculated static pressures can have serious uncertainties because of unknown amounts of liquid hydrocarbons or water in the wellbore and tubing and unknown temperature distribution. Calculated flowing pressures can have uncertainties because of inaccuracies in the detailed temperature distribution and the particular friction factor assumed for each specific case. Lesem er ~1. ~ provide helpful charts for approximating the temperature distribution in flowing gas wells. Errors and uncertainties of the above nature become worse as well depths increase. Consequently, for best results, downhole pressure measurements with accurate gauges should be used. Where this is not feasible, BHP s may be estimated from surface pressure readings for gas-condensate wells with better accuracy than is usually true for oil wells. Chaps. 33 and 34 discuss methods for making such estimates. For these methods, measured fluid properties (e.g., density) should be used whenever available in preference to calculated or correlation values. For gas and gas-condensate wells, a plot of static and producing BHP s vs. producing rates (in millions of standard cubic feet per day) is not a straight line. Smooth curves with closer approximations to straight lines can be obtained by plotting squares of the static and producing well BHP s (absolute) vs. producing rate. A rough analogy to oilwell behavior is then obtained by plotting the differences in squares of the static and producing pressures vs. the corresponding producing rates (usually on log-log paper). If several pressures are obtained on a well at different rates, these procedures do not always yield straight-line relationships (see Chap. 33 and Ref. 75); however, they provide reasonable indices for limited extrapolation to future well behavior and for comparison of current with past well behavior. Estimation of future well productivity can be made by modifying initial well productivity to account for the changes in reservoir pressure and gas permeability as pressure declines and liquid is deposited in the pores. For no loss of gas permeability, a new productivity line can be drawn on the plot of pressure squared vs. rate, parallel to the original productivity line and through the square of the new static pressure selected: this yields an estimate of flowing rate for any

flowing pressure selected. If the original curve for rate vs. difference in squares of static and flowing pressures is used, rates can be estimated for any future flowing pressure by using the proper (future) static pressure; lowpermeability wells would require special adjustment of earlier isochronal test data obtained (see Chap. 33 and Ref. 7.5). These methods yield approximations of future productivity as affected by pressure decline in the absence of fluid-phase and viscosity changes in the reservoir. If gas permeability, k,, is likely to be seriously affected by condensation of liqutds in the pores (and gas viscosity by pressure decline), then the change in gas mobility k,/p,, must be approximated and radial-flow calculations made (see Chap. 35) to estimate the new productivity curve corresponding to the static pressure selected for prediction. Normally, the two aforementioned types of productivity estimates ignore the drawdown effects of production on liquid-phase distribution in the reservoir and any consequent additional reduction of gas permeability near the producing wells; minimum calculated reduction of productivity should, therefore, result from these two estimating methods. Large deviations from such estimates, based on a well s early characteristics, would indicate that the well should be analyzed for productivity troubles. Excessive Productivity Loss. If the capacity of a producing well declines abnormally compared with that predicted from its original productivity (in the absence of excessive water production), and if appreciable liquid condensation around the wellbore within the formation is suspected, efforts to improve well productivity should be made. These could include the short-term injection of dry gas into the well (several days to several weeks) to evaporate part of the liquid, followed by immediate production to remove some of the vaporized liquid block. Loss of well productivity caused by excessive water production has been discussed briefly. In some cases, well workover operations would be justified to reduce or to shut off water entry. Other factors that can influence well productivity are deposits on the sandface or in the pores near the wellbore, perhaps caused by salts precipitated from reservoir water: any mechanical damage resulting from killing the well for pulling equipment or workover: mechanical failure of downhole equipment; and possible hydrates (see Chap. 33). In case of well productivity injury for mechanical reasons, conventional methods of well repair should be undertaken on the basis of the particular difficulty involved. Various means are available for stimulating lowproductivity wells; see Chaps. 54 and 55 and discussions by Clinkenbeard et al. 76 Well Injectivity Maintenance of well injectivity is essential for the economic operation of cycling programs. Injectivity decline can be caused by sandface plugging or by buildup of reservoir pressure. Lnjectivity Testing. The characterization of gas-injection wells is similar to that for gas-producing wells. In either case, analysis is made on the basis of plots of rates vs. the squares of BHP s or rates vs. differences of squares

39-26

PETROLEUM

ENGINEERING

HANDBOOK

of pressures. Consequently, after suitable well conditioning. as previously described, injectivity testing should consist of a series of injection rates at different pressures to establish the early injectivity performance of the well when well conditions are known to be good and the sandface is clean. If facilities are not available for obtaining a range of injection rates and pressures, it is sometimes acceptable to obtain production rates and pressures for the injection well through a reasonable range and use the pressure-squared relationship for extrapolating across the zero-rate axis into higher injectionpressure ranges to approximate well characteristics. Plots of production rate vs. difference in squares of pressure can also be adapted to estimate later well-injectivity behavior. As in the case of producing wells, if injectivity declines with time, analysis of well conditions is required to decide whether corrective procedures should be used. If a gascondensate reservoir is being operated essentially at constant pressure, then the obvious index of injectivity decline is whether the rate for each injection well remains constant at the injection-well pressure. Injection-rate decline at constant well pressure or injection-pressure rise at constant irrjection rate shows that injectivity is declining. Injection-Well Plugging. Plugging of the sandface can occur in injection wells. This may result from liquid carryover from the compressors (probably lubricating oil components) or from corrosion products from surface lines or well equipment. Carry-over of lubricating oils from compressors can be serious. Usually. the remedy is to install high-efficiency aftercoolers, scrubbers, and/or mist extractors on the discharge side of the compressors. A particularly el fective combination for this is the use of drips or collectors, followed by plate or screen impaction-type mist elitninators. followed by combination fibrous and wire-mesh filter elements. When liquid-blocking of the sand around an injection wellbore cannot be relieved by backflowing (as mentioned later), consideration can be given to slugging the well with suitable volatile solvents. The solvent used should preferably be miscible with both the normal injection gas and the liquid that is suspected to be blocking the pores. While propane is a good solvent for many hydrocarbon liquids, some lubricating oils have constituents not soluble or miscible with propane. In these cases, other solvents (possibly nonhydrocarbons) should be used. Sometimes solvent injection is followed immediately by resumption of dry-gas injection. If successful, this dissolves part or all of the liquid block and spreads out the materials in the reservoir sufficiently to relieve the problem. In other cases, the solvent is injected into the formation for short periods and then produced back out to provide a type of washing intended to remove the liquid accumulation from the formation. Corrosion products from steel lines between compressor discharge and the sandface can also provide serious well plugging. All well piping and casing and all surface lines should be cleaned thoroughly before they are installed to avert as much as possible the transportation of fine corrosion products to the sandface when injection starts. For continued protection during the life of injection equipment, liquid carry-over and mist-elimination measures should be combined with adequate control of corrosive

agents in the field gas. Sometimes the use of internally coated or lined pipe is justified. These and other corrosioncontrol procedures are best carried out with the help of a competent corrosion engineer. Corrosion products that plug the sandface are sometimes removed by backflowing the injection well to blow the material off the sand and out of the well. Where this is feasible, such complete removal of the plugging agents from the borehole is believed to be the best for the well. Other remedies may include treating the well with inhibited hydrochloric acid to dissolve the corrosion products. Sometimes the acid is pushed back into the formation and injection is started immediately without backwashing or backflowing of the well. If repeated periodically, this procedure is questionable because it is possible to develop plugging farther away from the well face that could ultimately hinder injection and be difficult to correct. Number of Wells Required The number of wells used in exploiting gas-condensate reservoirs has varied from the equivalent of less than I60 acres/well to more than 640 acres/well. Bennett discussed the general problem and pointed out that the first wells are drilled to determine the upper and lower limits of condensate production; to determine the extent of the pool, the net pay, thickness, porosity. etc.; and to provide suitable production or injection wells to fit a final pattern, which will not necessarily have a regular geometrical design. The number of wells to be drilled for gas-condensate operations must be analyzed for each specific case. Important factors to be considered are (I) contract commitments to deliver gas and products, (2) capacity of plant to be served, (3) productivities and injectivities of the wells, (4) maximum practical pattern (&S-weighted) efficiencies, controlled by number and location of wells (reservoir geometry is an important consideration), (5) amount of recoverable hydrocarbons and their value, and (6) project costs, including well-development costs. Items 3 through 5 must be balanced against Items 1. 2, and 6 to ensure that the economic objectives and contract commitments of the project are met. If wells are low in capacity, extra wells may be needed to meet production requirements during periods of well repair or workover.

Economics of Gas-Condensate Reservoir Operation


Arthur and Boatright and Dixon79 published discussions on the economics of cycling gas-condensate reservoirs. Arthur concluded that the most profitable method of operation depends on many factors. and the answer cannot be generalized. The following factors adapted from Arthur s list are considered important. 1. Reservoir formation and fluid characteristics, including occurrence or absence of black oil, size of reserves of products, properties and composition of reservoir hydrocarbons, productivities and injectivities of wells, permeability variation (controls the degree of bypassing of injected gas), and degree of natural water drive existing. 2. Reservoir development and operating costs. 3. Plant installation and operating costs. 4. Market demand for gas and liquid petroleum proaucts.

GAS-CONDENSATE

RESERVOIRS

39-27

5. Future relative value of the products. 6. Existence or absence of competitive producing conditions between operators in the same reservoir. 7. Severance, ad valorem, and income taxes. 8. Special hazards or risks (limited concession or lease life, political climate, and others). 9. Overall economic analysis. In choosing between pressure depletion and pressure maintenance as operating methods for a gas-condensate reservoir, detailed analyses must be made for predicting optimum economics. Cycling and gas processing procedures require sizable plant expenditures. Possible processing methods, whether reservoir fluids are cycled or not, include stabilization. compression, absorption, and fractionation. The last two recover appreciably more condensables from wet gas than do the first two. If the removal of ethane from a gas stream is desirable for economic or other reasons, fractionation should be used. When reservoir characteristics appear favorable for recovery of condensable hydrocarbons, it must be considered whether cycling would be economical. The primary comparison is between value of the estimated additional recovery of liquid products by cycling and the actual cycling costs, taking into account deferment of gas income and other factors. Economic analyses of cycling and noncycling are required and must be carried out in detail for maximum dependability with information factors and assumptions pertinent to each particular case. General information on valuation of oil and gas properties is given in Chap. 41. Economic comparisons are of no value unless reasonably accurate predictions of physical reservoir behavior can be made. Consequently. in the gas-condensate reservoir case. the information given previously would have to be expanded to include schedules of annual production and injection volumes derived from the physical characteristics of the reservoir and from the external factors that would affect production rates. Schedules of investment, anticipated prices of products. operating costs, and taxes would also be required to complete the detailed information needed to make comparative economic analyses.

0s = gas density. g/cm3 4 = porosity, X +,s = flow potential, psi

References

13. I?. IS. 16. D~,qirzerrirlg Daicr Book. ninth edItion. Gas Processors Suppliers Astn. and Natural Gas Assn. of America. Tulsa. OK (1981). 17. Katz. D .L. and Rzaaa. M .J. : B&liogrtrphJ .fiw Phwicd Ba/w\?or of Hw/mcarbons Under Prr.wurr md Rrlorrrl Phr~romv~~r. J W. Edwards Publisher Inc.. Ann Arbor (1946). 18. Genercrl Inde.~ to Pmolrum Puhlrcrrtiom of SPE-AIME. SPE, RIchardson, TX (1921-85) 1-5. 19. fnrfe.r c~Di~~isiori of Producrion Prqwrs. 192 7- 1953. API. New York City (1954). 20. Sloan. J.P.: Phase Behavior of Natural Gas and Condenaate Systems. Pet. Eq. (Feb. 1950) 22. No. 2. B-54-8-64. Prediction of Volumetric and 21. Dodson, C.R. and Standing. M.B.: Phase Behavior of Naturally Occurrq Hydrocarbon Systems. Drill. trnd Pmcl. Pruc~.. API (194 I) 326-40. 22. Organick. E.L.: Prediction of Critical Temperatures and Critical Pressures of Complex Hydrocarbon Mixtures. Clirw. &q. Prqq. (1953) 49, No. 6, 81-97. 23. Carr. N.L.. Kobayashi, R.. and Burrows. D.B.: Vlscnsity of Hydrocarbon Gases Under Pressure. J. Per. Tdz. (Oct. 1954) 47-55: Tiww.. AIME. 201. 24. Chew, J.N. and Connally, C.A. Jr.: A Viscosity Correlatmn for Gas-Saturated Crude Oils. Trcrn.s.. AIME (19.59) 216. 23-25. 25. API Recommended Practice for Sampling Petroleum Reservoir Fluids. API RP 44. first edttion. Dallas (Jan. 1966) 26. Bock-Prc~ssurz Tesr for IV&~& Co.\ l+ c,l/\ Texas Railroad Commission. Austin (1985). 27. Mutzual for Back Pwssuw Tcsr,fi~r Nuruml Grrs We//.\, New Mexico Oil Conservation Commission. Santa Fe (1966). 28. Munuul ofBud Pressurc~ Tcdna rf Gav W~,i/s. Kansas State Corp. Commission, Topeka (1959). 29 A Su,qp~fed Mumud for Standurtl Buck- Pw wuc Tc.viqq Mm ml.~ Interstate Oil Compact Commission. Oklahoma City (1986). 30. Whitson. C.H.: Characterizing Hydrocarbon Plus Fraction\. paper EUR I83 presented at the 1980 SPE European Offshore Pctroleum Conference and Exhibition. London. Oct. 2 I-24. 71. Watson, K.M.. Nelson. E.F.. and Murphy. G.B.: Characterization nf Petroleum Fractions. fncl. ,%K. Chrw. 11935) 27. 1460-64.

Nomenclature
8, = gas expansion factor (gas FVF) E, = area1 sweep efficiency ED = displacement efficiency E, = invasion efficiency E,, = pattern (h4.5weighted) efficiency ER = reservoir cycling efficiency El/ = volumetric sweep efficiency h= net pay thickness, ft k= permeability. md k,, = relative permeability to gas, fraction k,, = relative permeability to oil, fraction K= equilibrium ratio P= pressure, psi S= hydrocarbon fluid saturation of the pore space, % layer number deviation factor (compressibility factor) gas viscosity, cp oil viscosity, cp

39-28

PETROLEUM

ENGINEERING

HANDBOOK

32. Hoffman, A.E.. Grump, J.S.. and Hocott. CR.: Equilibrnnn Constants for a Gas-Condensate System, Trans., AIME (1953) 198. l-10. 33. Cook, A.B., Walker, C.J., and Spencer, G.B.: Realistic K Values Hydrocarbons for Calculating Oil Vaporization During Gas OfC,, e;r. Tech:(July 1969) 9Oi-15; Cycling at High Pressures, .I t Tram.. AIME. 246. A Correlation of Dcwpoint 34. Nemeth. L.K. and Kennedy. H.T.: Pressure With Fluid Composition and Temperature. Sot. Per. Enx. I. (June 1967) 99-104. Performance Characteristics of a 35 Allen. F.H. and Roe, R.P.: Volumetric Condensate Reservoir, Trans., AIME (1950) 189, 83-90. Predicted Performance of a Gas-Condensate 36. Berryman, J.E.: System. Washington Field, Louisiana, J. Per. Tech. (April 1957) 102-07: 7-runs.. AIME, 210. Comparison Be37. Rodgers, J.K., Harrison. N.H.. and Regier, S.: ~ween the Predicted and Actual Production History of a Condensate Reservoir. J. Per. Tech. (June 1958) 127-31: Trans., AIME. 213. On the Thermodynamics of 38. Redlich, 0. and Kwong, J.N.S.: Solutions V. an Equation of State Fuaacities of Gaseous Solutions. Chem. Review (1949) 44. 233. A New Two-Constant Equation 39. Peng. D.Y. and Robinson, D.B.: of State. Ind. Eng. Chum Fundamentals (1976) 15. 15-59. Cubic Equations of State-Which? Ind. GI,~. Chrm. 40. Martin, J.J.: Fundamentals (May 1979) 18, 81, 41. Petroleum Conservafion, S.E. Buckley ef al. (eds.), AIME, New York City (1951). 42. Geffen. T.M. @al.: Efticiency of Gas Displacement From Porous Media by Liquid Flooding. Trans., AIME (1952) 195, 29-38. 43. Yuster, S.T.: The Rehabilitation of Drowned Gas Wells, Drill. and Prod. Prac.. API ( 1946) 209- 16. 44. Bennett. E.N. and Auvcnshine. W.L.: Dewatering of Gas Wells, Drill. and Prod. Prac.. API (1956) 224-30. Foaming Agents arc Low-Cost 45. Dunning, H.N. and Eakin, J.L.: Treatment for Tired Gassers. Oil and GasJ. (Feb. 2. 1959) 57. No. 6, 108-10. Eight Years of 46. Bates, G.O., Kilmer, J.W., and Shirley, H.T.: Experience with Inert Gas Equipment. paper 57-PET-34 presented at the 1957 ASME Petroleum Mechanical Engineermg Conference. Sept. Fourteen Years of Progress tn Catalytic Treating 47. Barstow, W.F.: of Exhaust Gas. paper SPE 457 presented at the 1973 SPE Annual Meeting. Las Vegas, Sept. 30-Oct. 3. Unique Enhanced Oil and Gas 48. Eckies. W.W. and Holden, W.W.: Recovery Project for Very High Pressure Wilcox Sands Uses Cryogenic Nitrogen and Methane Mixture, paper SPE 9415 presented at the 1980 SPE Annual Technical Conference and Exhibition, Dallas. Sept. 21-24. Phase Equilibrium Considerations 49. Moses, P.L. and Wilson, K.: in Utilizing Nitrogen for Improved Recovery From Retrograde Condensate Reservoirs, paper SPE 7493 presented at the 1978 SPE Annual Technical Conference and Exhibition, Houston. Oct. l-4. Optimal Recovery Experiments with Nz and 50. Peterson, A.V.: co, \. Pet. Enx. Inrl. (Nov. 1978) 40-50. Physical Prop&es of Nitrogen for Use in Petroleum Reservoirs, 51. Eu[/. 1 Air Products and Chemical Inc.. Allentown. PA (1977). Enhanced-Recovery Inert Gas Processes Compared, 52. Wilson, K.: 011 and Gas J. (July 31, 1978) 162-72. Economic Evaluation of 53. Donohoe. C.W. and Buchanan, R.D.: Cycling Gas-Condensate Reservotrs With Nitrogen. paper SPE 7494 presented at the 1978 SPE Annual Technical Conference and Exhibition, Houston, Oct. 1-4. 54. Proc.. Ninth Oil Recovery Conference, Symposium on Natural Gas m Texas. College Station, TX (1956).

s.5. BrinkIcy, T.W.. Calculation of Rate and Ultimate Recovery from Gas Condensate Reservoirs. paper 1028-G presented at the 1958 SPE Petroleum Conference on Production and Reservoir Engineering. Tulsa, OK, March 20-2 I. 56. Patton, C.E. Jr.: Evaluation of Pressure Matntenance by Internal Gas Injection in Volumetrically Controlled Rcscrvoirs. Trrr,r.s.. AIME (1947) 170, 112-55. 57. API Standing Subcommittee on Secondary Recovery Methods, Circ. D-294. API (March 1949) Appendix B 58. Marshall, D.L. and Oliver, L.R.: Some Uses and Limitations 01 Model Studies in Cycling, Truns., AIME (1948) 174. 67-87. 59. Calhoun, J.C. Jr.: Fitndarnenrals r$Rrsrrwir EnKineerrnX, U. of Oklahoma Press. Norman (1953) 358, 374. 60. Hock, R.L.: Determination of Cycling Efficiencies in Cotton Valley Field Gas Reservoir, Oil alrd Gus J. (Nov. 4, 1948) 47. No. 27, 63-99. 61. Calhoun. J.C. Jr.: A Resume of the Factors Governing Interpretation of Waterflood Performance, paper presented at the 1956 SPE-AIME North Texas Section Secondary Recovery Symposium, Wichita Falls, Nov. 19-20. 62. Pirson, S.J.: Oil Reservoir Emginrering, McGraw-Hill Book Co. Inc.. New York City (1958) 406. 63. Stelzer, R.B.: Model Study vs. Field Performance Cycling the Paluxy Condensate Reservoir, Drill. und Prod. Prur., API (1956) 336-42. 64. Muskat, M.: Effect of Permeability Stratification in Cycling Operations, Trans., AIME (1949) 179. 3 13-28. 65. Standing. M.B., Linblad. E.N.. and Parsons. R.L.: Calculated Recoveries by Cycling from a Retrograde Reservoir of Variable Permeability, Trans., AIME (1948) 174, 165-90. 66. Miller, M.G. and Lents, M.R.: Performance of Bodcaw Reservoir. Cotton Valley Field Cycling Project, New Methods of Predicting Gas-Condensate Reservoir Performance Under Cycling Operations Compared to Field Data. Drill. and Prod. Prac.. API (I 946) 128-49. 67. Law, J.: A Statistical Approach to the Interatttial Heterogeneity of Sand Reservoirs, Trans.. AIME (1945) 155, 202-22. 68. Hurst, W. and van Everdingen. A.F.: Performance of Distillate Reservoirs in Gas Cycling, Trans., AIME (1946) 16.5, 36-51. 69. Cardwell. W.T. Jr. and Parsons, R.L.: Average Permcabthttes of Heterogeneous Oil Sands, Trcr,~s., AIME (1945) 160, 34-42 70. Sheldon, W.C.: *Calculating Recovery by Cycltng a Retrograde Condensate Reservoir, .I. Pel. Tech. (Jan 19.59) 29-34. Simulation of Gas Condensate Reservoir Perform71. Coats, K.H.: ance. paper SPE 10512 presented at the 1982 SPE Reservoir Simulation Symposium. New Orleans. Jan. 3 I-Feb. 3. 72. Hurst, W.: Water Influx into a Reservoir and Its Application to the Equation of Volumetric Balance, Trcrris.. AIME (1943) 151. 57-72. 73. van Everdingen, A.F. and Hurst, W.: Application of Laplace Transformation to Flow Patterns in Reservoirs. Tram\. . AIME (1949) 186, 305-24. 74. Lesem, L.B. et ai. : A Method of Calculating the Distribution of Temperature in Flowing Gas Wells, J. Per. Tech. (June 1957) 169-76; Trans., AIME, 210. 75. Tek. M.R. 1Grove, M.L., and Pocttmann. F.H.: Method for Predicting the Back-Pressure Behavior of Low Permeability Natural Gas Wells, J. Pet. Tech. (Nov. 1957) 302-09: Truns.. AIME, 210. 76. Clinkenbeard. P., Bozeman, J.F., and Davidson. R.D.: Gas Well Stimulation Increases Production and Profits, J. Per. Tech. (Nov. 1958) 21-24. 77. Bennett. E.O.: Factors Influencing Spacing in Condensate Fields. Pet. Eq. (1944) 15, No. IO. 158-62. 78. Arthur, M.G.: Economics of Cycling, Drill. and Prod. Pram., API (1948) 144-59. 79. Boatright. B.B. and Dixon, P.C.: Practical Economics of Cyclmg. Drill. and Prod. Pram., API (1941) 221-27.

You might also like