Equations of State and PVT Analysis Applications for Improved Reservoir Modeling Equations of State and PVT Analysis Applications for Improved Reservoir Modeling Second Edition Tarek Ahmed, PhD, PE AMSTERDAM • BOSTON • HEIDELBERG • LONDON NEW YORK • OXFORD • PARIS • SAN DIEGO SAN FRANCISCO • SINGAPORE • SYDNEY • TOKYO Gulf Professional Publishing is an imprint of Elsevier Gulf Professional Publishing is an imprint of Elsevier 50 Hampshire Street, 5th Floor, Cambridge, MA 02139, USA The Boulevard, Langford Lane, Kidlington, Oxford, OX5 1GB, UK Copyright # 2016, 2007 Elsevier Inc. All rights reserved. No part of this publication may be reproduced or transmitted in any form or by any means, electronic or mechanical, including photocopying, recording, or any information storage and retrieval system, without permission in writing from the publisher. Details on how to seek permission, further information about the Publisher’s permissions policies and our arrangements with organizations such as the Copyright Clearance Center and the Copyright Licensing Agency, can be found at our website: www.elsevier.com/ permissions. This book and the individual contributions contained in it are protected under copyright by the Publisher (other than as may be noted herein). Notices Knowledge and best practice in this field are constantly changing. As new research and experience broaden our understanding, changes in research methods, professional practices, or medical treatment may become necessary. Practitioners and researchers must always rely on their own experience and knowledge in evaluating and using any information, methods, compounds, or experiments described herein. In using such information or methods they should be mindful of their own safety and the safety of others, including parties for whom they have a professional responsibility. To the fullest extent of the law, neither the Publisher nor the authors, contributors, or editors, assume any liability for any injury and/or damage to persons or property as a matter of products liability, negligence or otherwise, or from any use or operation of any methods, products, instructions, or ideas contained in the material herein. British Library Cataloguing in Publication Data A catalogue record for this book is available from the British Library Library of Congress Cataloging-in-Publication Data A catalog record for this book is available from the Library of Congress ISBN: 978-0-12-801570-4 For information on all Gulf Professional publications visit our website at http://www.elsevier.com/ This book is dedicated to my wife Wendy About the Author Tarek Ahmed, PhD, PE, served as a professor of petroleum engineering at Montana Tech of the University of Montana from 1980 to 2002. He has worked for Anadarko Petroleum, Baker Hughes, and Talisman Energy. In 2012, Dr. Ahmed founded the consulting company Tarek Ahmed & Associates, Ltd. Dr. Ahmed has authored or coauthored several textbooks, including Reservoir Engineering Handbook and Hydrocarbon Phase Behvior, among others. xiii Preface The primary focus of this book is to present the basic fundamentals of equations of state and PVT laboratory analysis, and their practical applications in solving reservoir engineering problems. The book is arranged so it can be used as a textbook for senior and graduate students, or as a reference book for practicing petroleum engineers. Chapter 1 reviews the principles of hydrocarbon phase behavior and illustrates the use of phase diagrams in characterizing reservoirs and hydrocarbon systems. Chapter 2 presents numerous mathematical expressions and graphical relationships that are suitable for characterizing the undefined hydrocarbon-plus fractions. Chapter 3 provides a comprehensive and updated review of natural gas properties and the associated, well-established correlations that can be used to describe the volumetric behavior of gas reservoirs. Chapter 4 discusses the PVT properties of crude oil systems and illustrates the use of laboratory data to generate the properties of crude oil that can be used to perform reservoir engineering studies. Chapter 5 reviews developments and advances in the field of empirical cubic equations of state, and demonstrates their practical applications in solving phase equilibria problems. xv Acknowledgments It is my hope that the information presented in this textbook will improve the understanding of the subject of equations of state and phase behavior. Much of the material on which this book is based was drawn from the publications of the Society of Petroleum Engineers, the American Petroleum Institute, and the Gas Processors Suppliers Associations. Tribute is paid to the educators and authors who have made numerous and significant contributions to the field of phase equilibria. I would like to especially acknowledge the significant contributions that have been made to the field of phase behavior and equations of state to Donald Katz, M. B. Standing, Curtis Whitson, and Bill McCain. Special thanks to the editorial and production staff of Elsevier, in particular, Anusha Sambamoorthy, Kattie Washington, and Katie Hammon. I hope my children Carsen and Asia will follow my footsteps and study petroleum engineering. xvii Chapter 1 Fundamentals of Hydrocarbon Phase Behavior A PHASE IS DEFINED AS ANY homogeneous part of a system that is physically distinct and separated from other parts of the system by definite boundaries. For example, ice, liquid water, and water vapor constitute three separate phases of the pure substance H2O, because each is homogeneous and physically distinct from the others; moreover, each is clearly defined by the boundaries existing between them. Whether a substance exists in a solid, liquid, or gas phase is determined by the temperature and pressure acting on the substance. It is known that ice (solid phase) can be changed to water (liquid phase) by increasing its temperature and, by further increasing the temperature, water changes to steam (vapor phase). This change in phases is termed phase behavior. Hydrocarbon systems found in petroleum reservoirs are known to display multiphase behavior over wide ranges of pressures and temperatures. The most important phases that occur in petroleum reservoirs are a liquid phase, such as crude oils or condensates, and a gas phase, such as natural gases. The conditions under which these phases exist are a matter of considerable practical importance. The experimental or the mathematical determinations of these conditions are conveniently expressed in different types of diagrams, commonly called phase diagrams. The objective of this chapter is to review the basic principles of hydrocarbon phase behavior and illustrate the use of phase diagrams in describing and characterizing the volumetric behavior of single-component, twocomponent, and multicomponent systems. SINGLE-COMPONENT SYSTEMS The simplest type of hydrocarbon system to consider is that containing one component. The word component refers to the number of molecular or atomic species present in the substance. A single-component system is Equations of State and PVT Analysis. http://dx.doi.org/10.1016/B978-0-12-801570-4.00001-5 Copyright # 2016 Elsevier Inc. All rights reserved. 1 2 CHAPTER 1 Fundamentals of Hydrocarbon Phase Behavior composed entirely of one kind of atom or molecule. We often use the word pure to describe a single-component system. The qualitative understanding of the relationship between temperature T, pressure p, and volume V of pure components can provide an excellent basis for understanding the phase behavior of complex petroleum mixtures. This relationship is conveniently introduced in terms of experimental measurements conducted on a pure component as the component is subjected to changes in pressure and volume at a constant temperature. The effects of making these changes on the behavior of pure components are discussed next. Suppose a fixed quantity of a pure component is placed in a cylinder fitted with a frictionless piston at a fixed temperature T1. Furthermore, consider the initial pressure exerted on the system to be low enough that the entire system is in the vapor state. This initial condition is represented by point E on the pressure-volume phase diagram (p-V diagram) as shown in Fig. 1.1. Consider the following sequential experimental steps taking place on the pure component: 1. The pressure is increased isothermally by forcing the piston into the cylinder. Consequently, the gas volume decreases until it reaches point F on the diagram, where the liquid begins to condense. The corresponding pressure is known as the dew point pressure, pd, and defined as the pressure at which the first droplet of liquid is formed. 2. The piston is moved further into the cylinder as more liquid condenses. This condensation process is characterized by a constant pressure and represented by the horizontal line FG. At point G, traces of gas remain Van der Waals’ critical point condition H c C pc c c c c T1 < T2 < T3 < TC < T4 Critical point Liquid + Vapor G F B Vc c V n FIGURE 1.1 Typical pressure-volume diagram for pure component. A T4 TC T3 T E 2 T1 Single-Component Systems 3 and the corresponding pressure is called the bubble point pressure, pb, and defined as the pressure at which the first sign of gas formation is detected. A characteristic of a single-component system is that, at a given temperature, the dew point pressure and the bubble point pressure are equal. 3. As the piston is forced slightly into the cylinder, a sharp increase in the pressure (point H) is noted without an appreciable decrease in the liquid volume. That behavior evidently reflects the low compressibility of the liquid phase. By repeating these steps at progressively increasing temperatures, a family of curves of equal temperatures (isotherms) is constructed as shown in Fig. 1.1. The dashed curve connecting the dew points, called the dew point curve (line FC), represents the states of the “saturated gas.” The dashed curve connecting the bubble points, called the bubble point curve (line GC), similarly represents the “saturated liquid.” These two curves meet a point C, which is known as the critical point. The corresponding pressure and volume are called the critical pressure, pc, and critical volume, Vc, respectively. Note that, as the temperature increases, the length of the straight-line portion of the isotherm decreases until it eventually vanishes and the isotherm merely has a horizontal tangent and inflection point at the critical point. This isotherm temperature is called the critical temperature, Tc, of that single component. This observation can be expressed mathematically by the following relationship: @p @V ¼ 0, at the critical point (1.1) 2 @ p ¼ 0, at the critical point @V 2 Tc (1.2) Tc Referring to Fig. 1.1, the area enclosed by the area AFCGB is called the twophase region or the phase envelope. Within this defined region, vapor and liquid can coexist in equilibrium. Outside the phase envelope, only one phase can exist. The critical point (point C) describes the critical state of the pure component and represents the limiting state for the existence of two phases, that is, liquid and gas. In other words, for a single-component system, the critical point is defined as the highest value of pressure and temperature at which two phases can coexist. A more generalized definition of the critical point, which is applicable to a single- or multicomponent system, is this: The critical point is the point at which all intensive properties of the gas and liquid phases are equal. 4 CHAPTER 1 Fundamentals of Hydrocarbon Phase Behavior An intensive property is one that has the same value for any part of a homogeneous system as it does for the whole system; that is, a property independent of the quantity of the system. Pressure, temperature, density, composition, and viscosity are examples of intensive properties. Many characteristic properties of pure substances have been measured and compiled over the years. These properties provide vital information for calculating the thermodynamic properties of pure components as well as their mixtures. The most important of these properties include n n n n n n n n Critical pressure, pc. Critical temperature, Tc. Critical volume, Vc. Critical compressibility factor, Zc. Boiling point temperature, Tb. Acentric factor, ω. Molecular weight, M. Specific gravity, γ. Those physical properties needed for hydrocarbon phase behavior calculations are presented in Table 1.1 for a number of hydrocarbon and nonhydrocarbon components. Another means of presenting the results of this experiment is shown in Fig. 1.2, in which the pressure and temperature of the system are the independent parameters. Fig. 1.2 shows a typical pressure-temperature diagram (p-T diagram) of a single-component system with solids lines that clearly represent three different phase boundaries: vapor-liquid, vapor-solid, and liquid-solid phase separation boundaries. As shown in the illustration, line AC terminates at the critical point (point C) and can be thought of as the dividing line between the areas where liquid and vapor exist. The curve is commonly called the vapor pressure curve or the boiling point curve. The corresponding pressure at any point on the curve is called the vapor pressure, pv, with a corresponding temperature termed the boiling point temperature. The vapor pressure curve represents the conditions of pressure and temperature at which two phases, vapor and liquid, can coexist in equilibrium. Systems represented by a point located below the vapor pressure curve are composed only of the vapor phase. Similarly, points above the curve represent systems that exist in the liquid phase. These remarks can be conveniently summarized by the following expressions: Table 1.1 Physical Properties for Pure Components See Note No. ! Number Compound Formula 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 Methane Ethane Propane Isobutane n-Butane Isopentane n-Pentane Neopentane n-Hexane 2-Methylpentane 3-Methylpentane Neohexane 2.3-Dimethylbutane n-Heptane 2-Methylhexane 3-Methylhexane 3-Ethylpentane 2.2-Dimethylpetane 2.4-Dimethylpentane 3.3-Dimethylpentane Triptane n-Octane Dilsobutyl Isooctane n-Nonane n-Decane Cyclopentane Methylcyclopentane Cyclohexane Methyl cyclohexane Ethene (ethylene) Propene (propylene) CH4 C2H6 C3H8 C4H10 C4H10 C5H12 C5H12 C5H12 C6H14 C6H14 C6H14 C6H14 C6H14 C7H16 C7H16 C7H16 C7H16 C7H16 C7H16 C7H16 C7H16 C8H18 C8H18 C8H18 C9H20 C8H22 C5H10 C6H12 C6H12 C7H14 C6H4 C6H6 A B C D Critical Constants Molar Mass (Molecular Weight) Boiling Point (°F) 14.696 psia Vapor Pressure (psia), 100°F Freezing Point (°F) 14.696 psia Refractive Index, πD Pressure 60°F (psia) 16.043 30.070 44.097 58.123 58.123 72.150 72.150 72.150 86.177 86.177 86.177 86.177 86.177 100.204 100.204 100.204 100.204 100.204 100.204 100.204 100.204 114.231 114.231 114.231 128.258 14.285 70.134 84.161 84.161 98.188 28.054 42.081 258.73 127.49 43.75 10.78 31.08 82.12 96.92 49.10 155.72 140.47 145.89 121.52 136.36 209.16 194.09 197.33 200.25 174.54 176.89 186.91 177.58 258.21 228.39 210.63 303.47 345.48 120.65 161.25 177.29 213.68 154.73 53.84 (5000)* (800)* 188.64 72.581 51.706 20.445 15.574 36.69 4.9597 6.769 6.103 9.859 7.406 1.620 2.272 2.131 2.013 3.494 3.293 2.774 3.375 0.53694 1.102 1.709 0.17953 0.06088 9.915 4.503 3.266 1.609 (1400)* 227.7 296.44* 297.04* 305.73* 255.28 217.05 255.82 201.51 2.17 139.58 244.62 – 147.72 199.38 131.05 180.89 – 181.48 190.86 182.63 210.01 12.81 70.18 132.11 161.27 64.28 21.36 136.91 224.40 43.77 195.87 –272.47* 301.45* 1.00042* 1.20971* 1.29480* 1.3245* 1.33588* 1.35631 1.35992 1.342* 1.37708 1.37387 1.37888 1.37126 1.37730 1.38989 1.38714 1.39091 1.39566 1.38446 1.38379 1.38564 1.39168 1.39956 1.39461 1.38624 1.40748 1.41385 1.40896 1.41210 1.42862 1.42538 (1.228)* 1.3130* 666.4 706.5 616.0 527.9 550.6 490.4 488.6 464.0 436.9 436.6 453.1 446.8 453.5 396.8 396.5 408.1 419.3 402.2 396.9 427.2 428.4 360.7 360.6 372.4 331.8 305.2 653.8 548.9 590.8 503.5 731.0 668.6 Temperature Volume (°F) (ft3/lb-m) 116.67 89.92 206.06 274.46 305.62 369.10 385.8 321.13 453.6 435.83 448.4 420.13 440.29 512.7 495.00 503.80 513.39 477.23 475.95 505.87 496.44 564.22 530.44 519.46 610.68 652.0 461.2 499.35 536.6 570.27 48.54 197.17 0.0988 0.0783 0.0727 0.0714 0.0703 0.0679 0.0675 0.0673 0.0688 0.0682 0.0682 0.0667 0.0665 0.0691 0.0673 0.0646 0.0665 0.0665 0.0668 0.0662 0.0636 0.0690 0.0676 0.0656 0.0684 0.0679 0.0594 0.0607 0.0586 0.0600 0.0746 0.0689 Number 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 (Continued) Table 1.1 Physical Properties for Pure Components Continued See Note No. ! A B Boiling Point (°F) 14.696 psia Vapor Pressure (psia), 100°F Freezing Point (°F) 14.696 psia 20.79 38.69 33.58 19.59 85.93 51.53 24.06 93.31 120.49* 176.18 231.13 277.16 291.97 282.41 281.07 293.25 306.34 148.44 172.90 312.68 109.257* 76.497 14.11 27.99 317.8 422.955* 297.332* 320.451 29.13 212.000* 452.09 121.27 62.10 45.95 49.87 63.02 19.12 36.53 59.46 16.48 – 3.225 1.033 0.3716 0.2643 0.3265 0.3424 0.2582 0.1884 4.629 2.312 – – 394.59 85.46 211.9 – – – – 157.3 0.9501 – 906.71 301.63* 218.06 157.96 220.65 265.39 213.16 164.02 230.73 114.5 41.95 139.00 138.966 13.59 54.18 55.83 23.10 140.814 143.79 173.4 337.00* 69.83* 121.88* 103.86* 107.88* – 435.26* 361.820* 346.00* 149.73* 32.00 – 173.52* Number Compound Formula Molar Mass (Molecular Weight) 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 1-Butene (butylene) cis-2-Butene trans-2-Butene Isobutene 1-Pentene 1.2-Butadiene 1.3-Butadiene Isoprene Acetylene Benzene Toluene Ethylbenzene o-Xylene m-Xylene p-Xylene Styrene Isopropylbenzene Methyl alcohol Ethyl alcohol Carbon monoxide Carbon dioxide Hydrogen sulfide Sulfur dioxide Ammonia Air Hydrogen sulfide Oxygen Nitrogen Chlorine Water Helium Hydrogen chloride C4H8 C4H8 C4H8 C4H8 C5H10 C4H6 C4H6 C5H8 C2H2 C6H6 C7H8 C8H10 C8H10 C8H10 C8H10 C8H8 C9H12 CH4O C2H6O CO CO2 H2S SO2 NH3 N2+O2 H2 O2 N2 Cl2 H2O He HCl 56.108 56.108 56.108 56.108 70.134 54.092 54.092 68.119 26.038 78.114 92.141 106.167 106.167 106.167 106.167 104.152 120.194 32.042 46.069 28.010 44.010 34.08 64.06 17.0305 2.0159 2.0159 31.9988 28.0134 70.906 18.0153 4.0026 36.461 C D Critical Constants Refractive Index, πD Pressure 60°F (psia) 1.3494* 1.3665* 1.3563* 1.3512* 1.37426 – 1.3875* 1.42498 – 1.50396 1.49942 1.49826 1.50767 1.49951 1.49810 1.54937 1.49372 1.33034 1.36346 1.00036* 1.00048* 1.00060* 1.00062* 1.00036* 1.00028* 1.00013* 1.00027* 1.00028* 1.3878* 1.33335 1.00003* 1.00042* 583.5 612.1 587.4 580.2 511.8 (653)* 627.5 (558)* 890.4 710.4 595.5 523.0 541.6 512.9 509.2 587.8 465.4 1174 890.1 507.5 1071 1300 1143 1646 546.9 188.1 731.4 493.1 1157 3198.8 32.99 1205 Temperature Volume (°F) (ft3/lb-m) 295.48 324.37 311.86 292.55 376.93 (340)* 305 (412)* 95.34 552.22 605.57 651.29 674.92 651.02 649.54 (703)* 676.3 463.08 465.39 220.43 87.91 212.45 315.8 270.2 221.31 399.9 181.43 232.51 290.75 705.16 450.31 124.77 0.0685 0.0668 0.0679 0.0682 0.0676 (0.065)* 0.0654 (0.065)* 0.0695 0.0531 0.0550 0.0565 0.0557 0.0567 0.0570 0.0534 0.0572 0.0590 0.0581 0.0532 0.0344 0.0461 0.0305 0.0681 0.0517 0.5165 0.0367 0.0510 0.0280 0.0497s 0.2300 0.0356 Number 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 Table 1.1 Physical Properties for Pure Components Continued E F G H Density of Liquid 14.696 psia, 60°F Number Relative Density (Specific Gravity) 60°F/60°F 1b-m/gal. gal./lb-m 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 (0.3)* 0.35619* 0.50699* 0.56287* 0.58401* 0.62470 0.63112 0.59666* 0.66383 0.65785 0.66901 0.65385 0.66631 0.68820 0.68310 0.69165 0.70276 0.67829 0.67733 0.69772 0.69457 0.70696 (2.5)* 2.9696* 4.2268* 4.6927* 4.4690* 5.2082 5.2617 4.9744* 5.5344 5.4846 5.5776 5.4512 5.551 5.7376 5.6951 5.7664 5.8590 5.6550 5.6470 5.8170 5.7907 5.8940 (6.4172)* 10.126* 10.433* 12.386* 11.937* 13.853 13.712 14.504* 15.571 15.713 15.451 15.809 15.513 17.464 17.595 17.377 17.103 17.720 17.745 17.226 17.304 19.381 – – 0.00162* 0.00119* 0.00106* 0.00090 0.00086 0.00106* 0.00075 0.00076 0.00076 0.00076 0.00076 0.00068 0.00070 0.00070 0.00069 0.00070 0.00073 0.00067 0.00068 0.00064 23 24 25 26 0.69793 0.69624 0.72187 0.73421 5.8187 5.8046 6.0183 6.1212 19.632 19.679 21.311 23.245 0.00067 0.00065 0.00061 0.00057 I J Ideal Gas 14.696 psia, 60°F Specific Heat 60°F 14.696 psia Btu/ (1 bm, °F) Compressibility Factor of Real Gas, Z 14.696 psia, 60°F Relative Density (Specific Gravity) Air 5 1 ft3 gas/ lb-m 0.0104 0.0979 0.1522 0.1852 0.1995 0.2280 0.2514 0.1963 0.2994 0.2780 0.2732 0.2326 0.2469 0.3494 0.3298 0.3232 0.3105 0.2871 0.3026 0.2674 0.2503 0.3977 0.9980 0.9919 0.9825 0.9711 0.9667 – – 0.9582 – – – – – – – – – – – – – – 0.5539 1.0382 1.5226 2.0068 2.0068 2.4912 2.4912 2.4912 2.9755 2.9755 2.9755 2.9755 2.9755 3.4598 3.4598 3.4598 3.4598 3.4598 3.4598 3.4598 3.4598 3.9441 0.3564 0.3035 0.4445 0.4898 – – – – 3.9441 3.9441 4.4284 4.9127 Temperature Coefficient Acentric of Density, Factor, 1/°F ω ft3 gas/ gal. liquid Cp. I Deal Gas Cp. Liquid Number 23.654 (59.135)* 12.620 37.476* 8.6059 36.375* 6.5291 30.639* 5.25596 27.393 5.2596 27.674 5.2596 26.163* 4.4035 24.371 4.4035 24.152 4.4035 24.561 4.4035 24.462 3.7872 21.729 3.7872 21.568 3.7872 21.838 3.7872 22.189 3.7872 21.416 3.7872 21.386 3.7872 22.030 3.7872 21.930 3.220 19.580 3.3220 19.330 3.3220 19.283 0.52669 0.40782 0.38852 0.38669 0.38448 0.38825 0.39038 0.38628 0.38526 0.37902 0.37762 0.38447 0.38041 0.37882 0.38646 0.38594 0.39414 0.38306 0.37724 0.38331 0.37571 0.38222 – 0.97225 0.61996 0.57066 0.53331 0.54363 0.55021 0.53327 0.52732 0.51876 0.51308 0.52802 0.52199 0.51019 0.51410 0.51678 0.52440 0.50138 0.49920 0.52406 0.51130 0.48951 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 2.9588 2.9588 2.9588 2.6671 0.38246 0.38222 0.38246 0.38179 0.52244 0.48951 0.52244 0.52103 23 24 25 26 17.807 17.807 17.807 16.326 (Continued) Table 1.1 Physical Properties for Pure Components Continued E F G H Density of Liquid 14.696 psia, 60°F Number Relative Density (Specific Gravity) 60°F/60°F 1b-m/gal. gal./lb-m 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 0.75050 0.75349 0.78347 0.77400 – 0.52095* 0.60107* 0.62717* 0.60996* 0.60040* 0.64571 0.65799* 0.62723* 0.68615 (0.41796) 0.88448 0.87190 0.87168 0.88467 6.2570 6.2819 6.5319 6.4529 – 4.3432* 5.0112* 5.2288* 5.0853* 5.0056* 5.3834 5.4857* 5.2293* 5.7205 (3.4842) 7.3740 7.2691 7.2673 7.3756 11.209 13.397 12.885 15.216 – 9.6889* 11.197* 10.731* 11.033* 11.209* 13.028 9.8605 10.344* 11.908 (7.473) 10.593 12.676 14.609 14.394 Temperature Acentric Coefficient Factor, of Density, ω 1/°F 0.00073 0.00069 0.00065 0.00062 – 0.00173* 0.00112* 0.00105* 0.00106* 0.00117* 0.00089 0.00101* 0.00110* 0.00082* – 0.00067 0.00059 0.00056 0.00052 0.1950 0.2302 0.2096 0.2358 0.0865 0.1356 0.1941 0.2029 0.2128 0.1999 0.2333 0.2840 0.2007 0.1568 0.1949 0.2093 0.2633 0.3027 0.3942 I J Ideal Gas 14.696 psia, 60°F Specific Heat 60°F 14.696 psia Btu/ (1 bm, °F) Compressibility Factor of Real Gas, Z 14.696 psia, 60°F Relative Density (Specific Gravity) Air 5 1 ft gas/ lb-m ft3 gas/ gal. liquid – – – – 0.9936 0.9844 0.9699 0.9665 0.9667 0.9700 – (0.969) (0.965) – 0.9930 – – – – 2.4215 2.9059 2.9059 3.3902 0.9686 1.4529 1.9373 1.9373 1.9373 1.9373 2.4215 1.8677 1.8677 2.3520 0.8990 2.6971 3.1814 3.6657 3.6657 5.4110 4.5090 4.5090 3.8649 13.527 9.0179 6.7636 6.7636 6.7636 6.7636 5.4110 7.0156 7.0156 5.5710 14.574 4.8581 4.1184 3.5744 3.5744 33.856 28.325 29.452 24.940 – 39.167* 33.894* 35.366* 34.395* 33.856* 29.129 38.485* 36.687* 31.869 – 35.824 29.937 25.976 26.363 3 Cp. I Deal Gas Cp. Liquid Number 0.27199 0.30100 0.28817 0.31700 0.35697 0.35714 0.35446 0.33754 0.35574 0.37690 0.36351 0.34347 0.34120 0.35072 0.39754 0.24296 0.26370 0.27792 0.28964 0.42182 0.44126 0.43584 0.44012 – 0.57116 0.54533 0.54215 0.54215 0.54839 0.51782 0.54029 0.53447 0.51933 – 0.40989 0.40095 0.41139 0.41620 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 0.86875 0.86578 0.91108 0.86634 0.79626 0.79399 0.78939* 0.81802* 0.80144* 1.3974* 0.61832* 0.87476* 0.071070* 1.1421* 0.80940* 1.4244* 1.00000 0.12510* 0.85129* 7.2429 7.2181 7.5958 7.228 6.6385 6.6196 6.5812* 6.8199* 6.6817* 11.650* 5.1550* 7.2930* 0.59252* 9.5221* 6.7481* 11.875* 8.33712 1.0430* 7.0973* 14.658 14.708 13.712 16.641 4.8267 6.9595 4.2561* 6.4532* 5.1005* 5.4987 3.3037* 3.9713* 3.4022* 3.3605* 4.1513* 5.9710* 2.1609 3.8376* 5.1373* 0.00053 0.00056 0.00053 0.00055 0.00066 0.00058 – 0.00583* 0.00157* – – – – – – – 0.00009 – 0.00300* 0.3257 0.3216 (0.2412) 0.3260 0.5649 0.6438 0.0484 0.2667 0.0948 0.2548 0.2557 – 0.2202 0.0216 0.0372 0.0878 0.3443 0. 0.1259 – – – – – – 0.9959 0.9943 0.9846 0.9802 0.9877 1.0000 1.0006 0.9992 0.9997 (0.9875) – 1.0006 0.9923 3.6657 3.6657 3.5961 4.1500 1.1063 1.5906 0.9671 1.5196 1.1767 2.2118 0.5880 1.0000 0.06960 1.1048 0.9672 2.4482 0.62202 0.1382 1.2589 3.5744 3.5744 3.6435 3.1573 11.843 8.2372 13.548 8.6229 11.135 5.9238 22.283 13.103 188.25 11.859 13.546 5.3519 21.065 94.814 10.408 25.889 25.800 27.675 22.804 78.622 54.527 89.16* 58.807* 74.401* 69.012* 114.87* 95.557* 111.54* 112.93* 91.413* 63.554* 175.62 98.891* 73.806* 0.27427 0.27471 0.27110 0.29170 0.32316 0.33222 0.24847 0.19911 0.23827 0.14804 0.49677 0.23988 3.4038 0.21892 0.24828 0.11377 0.44457 1.2404 0.19086 0.40545 46 0.40255 47 0.41220 48 0.42053 49 0.59187 50 0.56610 51 – 52 – 53 0.50418 54 0.32460 55 1.1209 56 – 57 – 58 – 59 – 60 – 61 0.99974 62 – 63 – 64 Note: Numbers in this table do not have accuracies greater than 1 part in 1000; in some cases extra digits have been added to calculated values to achieve consistency or to permit recalculation of experimental values. Source: Courtesy of the Gas Processors Suppliers Association. Published in the GPSA Engineering Data Book, 10th edition, 1987. 10 CHAPTER 1 Fundamentals of Hydrocarbon Phase Behavior D pc r po Va id + Liquid qu Solid Li Solid + L iq uid C J Pressure I or + lid A Vapor p Va So B T1 T2 Tc Temperature n FIGURE 1.2 Typical pressure-temperature diagram for a single-component system. If p < pv ! the system is entirely in the vapor phase; If p > pv ! the system is entirely in the liquid phase; If p ¼ pv ! the vapor and liquid coexist in equilibrium; where p is the pressure exerted on the pure substance. It should be pointed out that these expressions are valid only if the system temperature is below the critical temperature Tc of the substance. The lower end of the vapor pressure line is limited by the triple point A. This point represents the pressure and temperature at which solid, liquid, and vapor coexist under equilibrium conditions. The line AB is called the sublimation pressure curve of the solid phase, and it divides the area where solid exists from the area where vapor exists. Points above AB represent solid systems, and those below AB represent vapor systems. The line AD is called the melting curve or fusion curve and represents the change of melting-point temperature with pressure. The fusion (melting) curve divides the solid phase area from the liquid phase area, with a corresponding temperature at any point on the curve termed the fusion or melting-point temperature. Note that the solid-liquid curve (fusion curve) has a steep slope, which indicates that the triple point for most fluids is close to their normal meltingpoint temperatures. For pure hydrocarbons, the melting point generally increases with pressure so the slope of the line AD is positive. Water is the exception in that its melting point decreases with pressure, so in this case, the slope of the line AD is negative. Single-Component Systems 11 Each pure hydrocarbon has a p-T diagram similar to that shown in Fig. 1.2. In addition, each pure component is characterized by its own vapor pressures, sublimation pressures, and critical values, which are different for each substance, but the general characteristics are similar. If such a diagram is available for a given substance, it is obvious that it could be used to predict the behavior of the substance as the temperature and pressure are changed. For example, in Fig. 1.2, a pure component system is initially at a pressure and temperature represented by the point I, which indicates that the system exists in the solid phase state. As the system is heated at a constant pressure until point J is reached, no phase changes occur under this isobaric temperature increase and the phase remains in the solid state until the temperature reaches T1. At this temperature, which is identified as the melting point at this constant pressure, liquid begins to form and the temperature remains constant until all the solid has disappeared. As the temperature is further increased, the system remains in the liquid state until the temperature T2 is reached. At T2 (which is the boiling point at this pressure), vapor forms and again the temperature remains constant until all the liquid has vaporized. The temperature of this vapor system now can be increased until the point J is reached. It should be emphasized that, in the process just described, only the phase changes were considered. For example, in going from just above T1 to just below T2, it was stated that only liquid was present and no phase change occurred. Obviously, the intensive properties of the liquid are changed as the temperature is increased. For example, the increase in temperature causes an increase in volume with a resulting decrease in the density. Similarly, other physical properties of the liquid are altered, but the properties of the system are those of a liquid and no other phases appear during this part of the isobaric temperature increase. A method that is particularly convenient for plotting the vapor pressure as a function of temperature for pure substances is shown in Fig. 1.3. The chart is known as a Cox chart. Note that the vapor pressure scale is logarithmic, while the temperature scale is entirely arbitrary. Review of Pure Component Volumetric Properties n Density “ρ,” expressed in lbm/ft3, is defined as the ratio of mass “m” to the volume, ie, ρ¼ n m V Number of moles “n” is defined as the ratio of mass “m” to the molecular weight “M,” or n¼ m M 12 CHAPTER 1 Fundamentals of Hydrocarbon Phase Behavior n FIGURE 1.3 Vapor pressure chart for hydrocarbon components. (Courtesy of the Gas Processors Suppliers Association. Published in the GPSA Engineering Data Book, 10th edition, 1987.) n Specific volume “v,” expressed in ft3/lbm, is defined as the ratio of volume to mass; ie, reciprocal of density: v¼ n V 1 ¼ m ρ Molar volume “Vm” is defined as the ratio of volume “V” number of moles “n,” ie, V M Vm ¼ ¼ n ρ n Molar density “ρm” is defined by the relationship ρm ¼ 1 ρ ¼ Vm M Single-Component Systems 13 EXAMPLE 1.1 A pure propane is held in a laboratory cell at 80°F and 200 psia. Determine the “existence state” (ie, as a gas or liquid) of the substance. Solution From a Cox chart, the vapor pressure of propane is read as pv ¼ 150 psi, and because the laboratory cell pressure is 200 psi (ie, p > pv), this means that the laboratory cell contains a liquefied propane. The vapor pressure chart as presented in Fig. 1.3 allows a quick estimation of the vapor pressure pv of a pure substance at a specific temperature. For computer applications, however, an equation is more convenient. Lee and Kesler (1975) proposed the following generalized vapor pressure equation: pv ¼ pc exp ðA + ωBÞ (1.3) With the coefficients A and B as defined below: A ¼ 5:92714 6:09648 1:2886 ln ðTr Þ + 0:16934ðTr Þ6 Tr (1.4) B ¼ 15:2518 15:6875 13:4721 ln ðTr Þ + 0:4357ðTr Þ6 Tr (1.5) The term Tr, called the reduced temperature, is defined as the ratio of the absolute system temperature to the critical temperature of the component, or Tr ¼ T Tc where Tr ¼ reduced temperature T ¼ substance temperature, °R Tc ¼ critical temperature of the substance, °R pc ¼ critical pressure of the substance, psia ω ¼ acentric factor of the substance The acentric factor “ω” is a concept that was introduced by Pitzer (1955), and has proven to be very useful in the characterization of substances. It has become a standard for the proper characterization of any single pure component, along with other common properties, such as molecular weight, critical temperature, critical pressure, and critical volume. The acentric factor ω is a unique correlating parameter that is used as a measure of the centricity or the deviation of the component molecular shape from that of a spherical. The shape of the Argon molecule is considered completely spherical and is assigned an acentric factor of zero. This important property is defined by the following expression: pv ω ¼ log 1 (1.6) pc T¼0:7Tc 14 CHAPTER 1 Fundamentals of Hydrocarbon Phase Behavior where pc ¼ critical pressure of the substance, psia pv ¼ vapor pressure of the substance at a temperature equal to 70% of the substance critical temperature (ie, T ¼ 0.7Tc), psia The acentric factor frequently is used as a third parameter in corresponding states and equation-of-state correlations. Values of the acentric factor for pure substances are tabulated in Table 1.1. EXAMPLE 1.2 Calculate the vapor pressure of propane at 80°F by using the Lee and Kesler correlation. Solution Obtain the critical properties and the acentric factor of propane from Table 1.1: Tc ¼ 666:01°R pc ¼ 616:3psia ω ¼ 0:1522 Calculate the reduced temperature: Tr ¼ T 540 ¼ ¼ 0:81108 Tc 666:01 Solve for the parameters A and B by applying Eqs. (1.4), (1.5), respectively, to give: A ¼ 5:92714 6:09648 1:2886 ln ðTr Þ + 0:16934ðTr Þ6 Tr A ¼ 5:92714 6:09648 1:2886 ln ð0:81108Þ + 0:16934ð0:81108Þ6 ¼ 1:273590 0:81108 B ¼ 15:2518 15:6875 13:4721 ln ðTr Þ + 0:4357ðTr Þ6 Tr B ¼ 15:2518 15:6875 13:4721 ln ð0:81108Þ + 0:4357ð0:81108Þ6 ¼ 1:147045 0:81108 Solve for pv by applying Eq. (1.3): pv ¼ pc exp ðA + ωBÞ pv ¼ ð616:3Þ exp ½1:27359 + 0:1572ð1:147045Þ ¼ 145psia The densities of the saturated phases of a pure component (ie, densities of the coexisting liquid and vapor) may be plotted as a function of temperature, as shown in Fig. 1.4. Note that, for increasing temperature, the density of the saturated liquid is decreasing, while the density of the saturated vapor Single-Component Systems 15 Density A x x Densit y curv e of s Avera d liquid ge de (rv + r L) / 2 = Critical density rc aturate nsity ravg = a + bT C rc = a + bTc B sity x Den e curv por d va rate tu of sa Temperature TC n FIGURE 1.4 Typical density-temperature diagram. increases. At the critical point C, the densities of vapor and liquid converge and become identical with a value equal to the critical density of the pure substance “ρc.” It should be pointed out that at the critical point C, all other properties of the phases become identical, such as viscosity and density. Fig. 1.4 illustrates a useful observation that states that the arithmetic average of the densities of the liquid and vapor phases is a linear function of the temperature and can be represented by a straight-line relationship that passes through the critical point. Mathematically, this straight-line relationship is expressed as ρv + ρL ¼ ρavg ¼ a + bT 2 (1.7) where ρv ¼ density of the saturated vapor, lb/ft3 ρL ¼ density of the saturated liquid, lb/ft3 ρavg ¼ arithmetic average density, lb/ft3 T ¼ temperature, °R a, b ¼ intercept and slope of the straight line Because, at the critical point, ρv and ρL are identical, Eq. (1.7) can be expressed in terms of the critical density as follows: ρc ¼ a + bTc (1.8) where ρc ¼ critical density of the substance, lb/ft3. Combining Eq. (1.7) with (1.8) and solving for the critical density gives a + bT ρ ρc ¼ a + bTc avg 16 CHAPTER 1 Fundamentals of Hydrocarbon Phase Behavior This density-temperature diagram is useful in calculating the critical volume from density data. The experimental determination of the critical volume sometimes is difficult, as it requires the precise measurement of a volume at a high temperature and pressure. However, it is apparent that the straight line obtained by plotting the average density versus temperature intersects the critical temperature at the critical density. The molal critical volume is obtained by dividing the molecular weight by the critical density: Vc ¼ where M ρc Vc ¼ critical volume of pure component, ft3/lbm-mol M ¼ molecular weight, lbm/lbm-mol ρc ¼ critical density, lbm/ft3 Fig. 1.5 shows the saturated densities for a number of fluids of interest to the petroleum engineer. Note that, for each pure substance, the upper curve is (Specific gravity) Density grams in c.c. Hydrocarbon fluid densities Temperature (°F) Identification A 8 mol% CH4 – 92 mol% C3H8 (370 – 739 lbs.) B 50.25 wt% C2H6 – 49.75 wt% n-C7H16 C 19.2 wt% CH4 – 80.8 wt% C6H14 (2412 – 2506 lbs.) D 7.15 wt% CH4 – 92.85 wt% n-C5H12 (8540 – 1043 lbs.) E National Standard Petroleum Tables F 7.15 wt% CH4 – 92.85 wt% n % C5H12 (at 3000 lbs.) G 9.78 wt% C2H6 – 90.22 wt% n-C7H16 H 75.45 mol% C3H5 – 24.55% n-C4H10 I 65.77 mol% C2H6 – 34.23% n-C4H10 n FIGURE 1.5 Fluid densities. (Courtesy of the Gas Processors Suppliers Association. GPSA Engineering Book, 10th edition, 1987.) Single-Component Systems 17 termed the saturated liquid density curve, while the lower curve is labeled the saturated vapor density curve with both curves meet and terminate at the critical point “C.” EXAMPLE 1.3 Calculate the saturated liquid and gas densities of n-butane at 200°F. Solution From Fig. 1.5, read both values of liquid and vapor densities at 200°F to give Liquid densityρL ¼ 0:475g=cm3 Vapor density ρv ¼ 0:035g=cm3 The density-temperature diagram also can be used to determine the state of a single-component system, ie, the pure component exits in a liquefied state, vaporous state, or two-phase state. Suppose that the overall (total) density of the system, ρt, is known at a given temperature. If this overall density is less than or equal to the pure component vapor density ρv, it is obvious that the system is composed entirely of vapor. Similarly, if the overall density ρt is greater than or equal to the pure component liquid density ρL, the system is composed entirely of liquid. If, however, the overall density is between ρL and ρv, it indicates that both liquid and vapor are present. If both phases are present, the volume and weight of each phase present can be calculated by applying volume and mass balances as shown below: mL + mv ¼ mt VL + Vv ¼ Vt where mL, mv, and mt ¼ the mass of the liquid, vapor, and total system, respectively VL, Vv, and Vt ¼ the volume of the liquid, vapor, and total system, respectively Combining the above two equations and introducing the density into the resulting equation gives mt mv mv + ¼ Vt ρL ρv (1.9) 18 CHAPTER 1 Fundamentals of Hydrocarbon Phase Behavior EXAMPLE 1.4 Ten pounds of a hydrocarbon are placed in a 1 ft3 vessel at 60°F. The densities of the coexisting liquid and vapor are known to be 25 and 0.05 lb/ft3, respectively, at this temperature. Calculate the weights and volumes of the liquid and vapor phases. Solution Step 1. Calculate the density of the overall system: ρt ¼ mt 10 ¼ ¼ 10lb=ft3 Vt 1:0 Step 2. As the overall density of the system is between the density of the liquid and the density of the gas, the system must be made up of both liquid and vapor. Step 3. Calculate the weight of the vapor from Eq. (1.9): 10 mv mv + ¼1 25 0:05 Solving the above equation for mv and applying the mass balance to solve for mL gives mv ¼ 0:030lb mL ¼ 10 mv ¼ 10 0:03 ¼ 9:97lb Step 4. Calculate the volume of the vapor and liquid phases: Vv ¼ mv 0:03 ¼ ¼ 0:6ft3 ρv 0:05 VL ¼ Vt Vv ¼ 1 0:6 ¼ 0:4ft3 EXAMPLE 1.5 A utility company stored 58 million pounds of propane, ie, mt ¼ 58,000,000 lb, in a washed-out underground salt cavern of volume 480,000 bbl (Vt ¼ 480,000 bbl) at a temperature of 110°F. Estimate the weight and volume of the existing propane phases in the underground cavern. Solution Step 1. Calculate the volume of the cavern in ft3: Vt ¼ ð480, 000Þð5:615Þ ¼ 2, 695,200ft3 Step 2. Calculate the total density of the system: ρt ¼ mt 58,000, 000 ¼ ¼ 21:52lb=ft3 Vt 2, 695,200 Single-Component Systems 19 Step 3. Determine the saturated densities of propane from the density chart of Fig. 1.5 at 110°F: ρL ¼ 0:468g=cm3 ¼ ð0:468Þð62:4Þ ¼ 29:20lb=ft3 ρv ¼ 0:030g=cm3 ¼ ð0:03Þð62:4Þ ¼ 1:87lb=ft3 Step 4. Test for the existing phases. Because ρv < ρt < ρL 1:87 < 21:52 < 29:20 both liquid and vapor are present. Step 5. Solve for the weight of the vapor phase by applying Eq. (1.9) 58, 000,000 mv mv + ¼ 2, 695,200 29:20 1:87 mv ¼ 1,416, 345lb Vv ¼ mv 1,416, 345 ¼ ¼ 757,404ft3 ρ 1:87 Step 6. Solve for the volume and weight of propane: VL ¼ Vt Vv ¼ 2,695, 200 757, 404 ¼ 1,937, 796ft3 The above suggests that the propane in the liquid state occupies 72% of total cavern volume. In terms of the propane mass: mL ¼ mt mv ¼ 58,000, 000 1,416, 345 ¼ 56,583, 655lb ð98%of totalweightÞ: Pure Component Saturated Liquid Density Rackett (1970) proposed a simple generalized equation for predicting the saturated liquid density, ρL, of pure compounds. Rackett expressed the relation in the following form: ρL ¼ Mpc RTc Zca with the exponent “a” given as a ¼ 1 + ð1 Tr Þ2=7 where M ¼ molecular weight of the pure substance pc ¼ critical pressure of the substance, psia Tc ¼ critical temperature of the substance, °R Zc ¼ critical gas compressibility factor (1.10) 20 CHAPTER 1 Fundamentals of Hydrocarbon Phase Behavior R ¼ gas constant, 10.73 ft3 psia/lb-m, °R Tr ¼ T=Tc , reduced temperature T ¼ temperature, °R Spencer and Danner (1973) modified Rackett’s correlation by replacing the critical compressibility factor Zc in Eq. (1.9) with a unique parameter called Rackett’s compressibility factor “ZRA,” which is constant for each compound. Spencer and Danner introduced the parameter ZRA into Eq. (1.10) to give the following modification of the Rackett equation: ρL ¼ Mpc RTc ðZRA Þa (1.11) with the exponent “a” as defined previously by a ¼ 1 + ð1 Tr Þ2=7 The values of ZRA are given in Table 1.2 for selected components. The Rackett parameter ZRA can also be estimated from a correlation proposed by Yamada and Gunn (1973) as ZRA ¼ 0:29056 0:08775ω (1.12) where ω is the acentric factor of the compound. The Rackett’s compressibility factor can be approximated as a function of the number of carbon atoms “n” for components heavier than heptane; eg, n ¼ 8,9,…, by the following simple relationship: ZRA ¼ 0:37748n0:16941 Table 1.2 Values of ZRA for Selected Pure Components Carbon dioxide Nitrogen Hydrogen sulfide Methane Ethane Propane i-Butane n-Butane i-Pentane 0.2722 0.2900 0.2855 0.2892 0.2808 0.2766 0.2754 0.2730 0.2717 n-Pentane n-Hexane n-Heptanes i-Octane n-Octane n-Nonane n-Decane n-Undecane 0.2684 0.2635 0.2604 0.2684 0.2571 0.2543 0.2507 0.2499 Single-Component Systems 21 EXAMPLE 1.6 Calculate the saturated liquid density of propane at 160°F by using 1. The Rackett correlation. 2. The modified Rackett equation. Solution 1. The Rackett Approach: Find the critical properties of propane from Table 1.1, to give Tc ¼ 666:06°R pc ¼ 616:0psia M ¼ 44:097 vc ¼ 0:0727ft3 =lb Calculate Zc by applying the real gas equation of state (EOS): Z¼ pV pV pvM ¼ ¼ nRT ðm=MÞRT RT where V ¼ substance volume, ft3 v ¼ substance specific volume; V/m, ft3/lb Using the critical properties of propane to determine Zc, to give Zc ¼ pc vc M ð616:0Þð0:0727Þð44:097Þ ¼ ¼ 0:2763 RTc ð10:73Þð666:06Þ Calculate the reduced temperature Tr and determine the saturated liquid density by applying the Rackett equation, Eq. (1.10): Tr ¼ T 160 + 460 ¼ ¼ 0:93085 Tc 666:06 a ¼ 1 + ð1 Tr Þ2=7 ¼ 1 + ð1 0:93085Þ2=7 ¼ 1:4661 ρL ¼ ρL ¼ Mpc RTc Zca ð44:097Þð616:0Þ ð10:73Þð666:06Þð0:2763Þ1:4661 ¼ 25:05lb=ft3 2. The Modified Rackett Approach: Using the Rackett compressibility factor of ZRA ¼ 0.2766 from Table 1.2 to calculate the liquid density, gives ρL ¼ ð44:097Þð616:0Þ ð10:73Þð666:06Þð0:2766Þ1:4661 ¼ 25:01lb=ft3 22 CHAPTER 1 Fundamentals of Hydrocarbon Phase Behavior TWO-COMPONENT SYSTEMS A unique feature of the single-component system is that, at a fixed temperature, two phases (vapor and liquid) can exist in equilibrium at only the vapor pressure. For a binary system, two phases can exist in equilibrium at various pressures at the same temperature. The following discussion concerning the description of the phase behavior of a two-component system involves many concepts that apply to the more complex multicomponent mixtures of oils and gases. An important characteristic of binary systems is the variation of their thermodynamic and physical properties with the composition. Therefore, it is necessary to specify the composition of the mixture in terms of mole or weight fractions. It is customary to designate one of the components as the more volatile component and the other the less volatile component, depending on their relative vapor pressure at a given temperature. Suppose that the examples previously described for a pure component are repeated, but this time we introduce into the cylinder a binary mixture of a known overall composition “zi.” Consider that the initial pressure p1 exerted on the system, at a fixed temperature of T1, is low enough that the entire system exists in the vapor state. This initial condition of pressure and temperature acting on the mixture is represented by point 1 on the p/V diagram of Fig. 1.6. As the pressure is increased isothermally, it reaches pb Constant temperature Liquid Pressure 4 Bubble point Bubble point pressure 3 pd Liqu id + Dew point Vap or Dew point pressure 2 Va por 1 Volume n FIGURE 1.6 Pressure-volume isotherm for a two-component system. Two-Component Systems 23 point 2, at which point an infinitesimal amount of liquid is condensed. The pressure at this point is called the dew point pressure, pd, of the mixture. It should be noted that, at the dew point pressure, the composition of the vapor phase is equal to the overall composition of the binary mixture. As the total volume is decreased by forcing the piston inside the cylinder, a noticeable increase in the pressure is observed as more and more liquid is condensed. This condensation process is continued until the pressure reaches point 3, at which point traces of gas remain. At point 3, the corresponding pressure is called the bubble point pressure, pb. Because, at the bubble point, the gas phase is only of infinitesimal volume, the composition of the liquid phase therefore is identical with that of the whole system. As the piston is forced further into the cylinder, the pressure rises steeply to point 4 with a corresponding decreasing volume. Repeating the above process at progressively increasing temperatures, a complete set of isotherms is obtained on the p/V diagram of Fig. 1.7 for a binary system consisting of n-pentane and n-heptane. The bubble point curve, as represented by line AC, represents the locus of the points of pressure and volume at which the first bubble of gas is formed. The dew point curve (line BC) describes the locus of the points of pressure and volume at which the first droplet of liquid is formed. The two curves meet at the critical point (point C). The critical pressure, temperature, and volume are given by pc, Tc, and Vc, respectively. Any point within the phase envelope (line ACB) represents a system consisting of two phases. Outside the phase envelope, only one phase can exist. Liquid C 45 0° 425° Bubble point Pressure 400 300 Critical point: 45 4° F Vapor 400° Liquid + Vapor Dew point curve 350° 200 300° B A 0.1 0.2 0.3 Volume (ft3/Ib) n FIGURE 1.7 Pressure-volume diagram for a binary system. 0.4 24 CHAPTER 1 Fundamentals of Hydrocarbon Phase Behavior If the bubble point pressure and dew point pressure for the various isotherms on a p-V diagram are plotted as a function of temperature, a p-T diagram similar to that shown in Fig. 1.8 is obtained. Fig. 1.8 indicates that the pressure-temperature relationships no longer can be represented by a simple vapor pressure curve, as in the case of a single-component system, but take on the form illustrated in the figure by the phase envelope ACB. The dashed lines within the phase envelope are called quality lines; they describe the pressure and temperature conditions of equal volumes of liquid. Obviously, the bubble point curve and the dew point curve represent 100% and 0% liquid, respectively. Fig. 1.9 demonstrates the effect of changing the composition of the binary system on the shape and location of the phase envelope. Two of the lines shown in the figure represent the vapor pressure curves for two pure components, designed as 1 for the lighter component and 2 for the other component. Four phase boundary curves (phase envelopes); labeled A through D, represent various mixtures of the two components with increasing the concentration component 2. These curves pass continuously from the vapor pressure curve of the one pure component to that of the other as the composition is varied. The points labeled A through D represent the critical points of the mixtures. It should be noted by examining Fig. 1.9 that, when one of the constituents becomes predominant, the binary mixture tends to exhibit a relatively narrow phase envelope and displays critical properties close to the predominant component. The size of the phase envelope enlarges noticeably as the composition of the mixture becomes evenly distributed between the two components. Cricondenbar Liquid % 100 Pressure A ble Bub t cur poin ve C Vapor + Liquid 75% Cricondentherm Pc 50% B 25% ur ve 0% oint c Dew p Vapor Temperature n FIGURE 1.8 Typical pressure-temperature diagram for a two-component system. Tc Two-Component Systems 25 B A C D Critical point Pressure Critical point Temperature n FIGURE 1.9 Phase diagram of binary mixture. Pressure-Composition Diagram “p-x” for Binary Systems The pressure-composition diagram, commonly called the p-x diagram, is another means of describing the phase behavior of a binary system, as its overall composition changes at a constant temperature. It is constructed by plotting the dew point and bubble point pressures as a function of composition. The bubble point and dew point lines of a binary system are drawn through the points that represent these pressures as the composition of the system is changed at a constant temperature. Fig. 1.10 represents a typical pressure-composition diagram for a two-component system. Component 1 is described as the more volatile fraction and component 2 as the less volatile fraction. Point A in the figure represents the vapor pressure (dew point, bubble point) of the more volatile component, while point B represent that of the less volatile component. Assuming a composition of 75% by weight of component 1 (ie, the more volatile component) and 25% of component 2, this mixture is characterized by a dew point pressure represented as point C and a bubble point pressure of point D. Different combinations of the two components produce different values for the bubble point and dew point 26 CHAPTER 1 Fundamentals of Hydrocarbon Phase Behavior A Constant temperature D Bub ble poin t cu Pressure Liquid Liquid + Vapor C Vapor r ve X Dew point Y cur ve B wv 0 25 50 wt% of less volatile component wL 75 100 n FIGURE 1.10 Typical pressure-composition diagram for a two-component system. Composition expressed in terms of wt% of the less volatile component. pressures. The curve ADYB represents the bubble point pressure curve for the binary system as a function of composition, while the line ACXB describes the changes in the dew point pressure as the composition of the system changes at a constant temperature. The area below the dew point line represents vapor, the area above the bubble point line represents liquid, and the area between these two curves represents the two-phase region, where liquid and vapor coexist. It should be pointed out that the composition may be expressed equally well in terms of weight percent of the more volatile component, in which case the bubble point and dew point lines have the opposite placement; that is, dew point curve above the bubble point curve. As shown in Fig. 5.10, the p-x diagram can also be displayed in terms of mole or weight percent of the more volatile component. The points X and Y at the extremities of the horizontal line XY represent the composition of the coexisting of the vapor phase (point X) and the liquid phase (point Y) that exist in equilibrium at the same pressure. In other words, at the pressure represented by the horizontal line XY, the compositions of the vapor and liquid that coexist in the two-phase region are given by wv and wL, and they represent the weight percentages of the less volatile component in the vapor and liquid, respectively. In the p-x diagram shown in Fig. 1.11, the composition is expressed in terms of the mole fraction of the more volatile component. Assume that a binary system with an overall composition of z exists in the vapor phase state as represented by point A. If the pressure on the system is increased, no phase change occurs until the dew point, B, is reached at pressure P1. At this dew point pressure, an infinitesimal amount of liquid forms whose composition is given by x1. The composition of the vapor still is equal to the original Two-Component Systems 27 C p3 Liquid Pressure t oin le p b Bub ve cur p2 Liquid + Vapor ap id iqu p1 XB or +V oint p Dew L e curv XA Vapor 0% x1 x2 z y2 Mole fraction of more volatile component n FIGURE 1.11 Pressure-composition diagram illustrating an isothermal compression through the two-phase region. composition z. As the pressure is increased, more liquid forms and the compositions of the coexisting liquid and vapor are given by projecting the ends of the straight, horizontal line through the two-phase region of the composition axis. For example, at p2, both liquid and vapor are present and the compositions are given by x2 and y2. At pressure p3, the bubble point, C, is reached. The composition of the liquid is equal to the original composition z with an infinitesimal amount of vapor still present at the bubble point with a composition given by y3. As indicated already, the extremities of a horizontal line through the twophase region represent the compositions of coexisting phases. Burcik (1957) points out that the composition and the amount of each phase present in a two-phase system are of practical interest and use in reservoir engineering calculations. At the dew point, for example, only an infinitesimal amount of liquid is present with a finite mole fractions of the two components. Introducing the following nomenclatures: n ¼ total number of moles in the binary system nL ¼ number of moles of liquid nv ¼ number of moles of vapor z ¼ mole fraction of the more volatile component in the system x ¼ mole fraction of the more volatile component in the liquid phase y ¼ mole fraction of the more volatile component in the vapor phase y3 100% 28 CHAPTER 1 Fundamentals of Hydrocarbon Phase Behavior An equation for the relative amounts of liquid and vapor in a two-phase system may be derived on one of the two components (the more volatile or less volatile). Selecting and developing molar material balance on the more volatile component, let n ¼ nL + nv n z ¼ moles of the more volatile component in the system nL x ¼ moles of the more volatile component in the liquid nv y ¼ moles of the more volatile component in the vapor A material balance on the more volatile component gives nz ¼ nL x + nv y (1.13) and nL ¼ n nv Combining these two expressions gives nz ¼ ðn nv Þx + nv y Rearranging the above expression, gives nv z x ¼ n yx (1.14) Similarly, if nv is eliminated in Eq. (1.13) instead of nL, it gives nL z y ¼ n xy (1.15) The geometrical interpretation of Eqs. (1.14), (1.15) is shown in Fig. 1.12, which indicates that these equations can be written in terms of the two segments of the horizontal line AC. Because z x ¼ the length of segment AB, and y x ¼ the total length of horizontal line AC, Eq. (1.14) becomes nv z x AB ¼ ¼ n y x AC (1.16) Similarly, Eq. (1.15) becomes nL BC ¼ n AC (1.17) Eq. (1.16) suggests that the ratio of the number of moles of vapor to the total number of moles in the system is equivalent to the length of the line segment Two-Component Systems 29 e urv nt c oi le p b Bub Liquid Liquid + Vapor B Pressure A C Liquid + Vapor Vapor Dew point curve x 0% z y Mole fraction of more volatile component n FIGURE 1.12 Geometrical interpretation of equations for the amount of liquid and vapor in the two-phase region. AB that connects the overall composition to the liquid composition divided by the total length psia. This rule is known as the inverse lever rule. Similarly, the ratio of the number of moles of liquid to the total number of moles in the system is proportional to the distance from the overall composition to the vapor composition BC divided by the total length AC. It should be pointed out that the straight line that connects the liquid composition with the vapor composition, that is, line AC, is called the tie line. Note that results would have been the same if the mole fraction of the less volatile component had been plotted on the phase diagram instead of the mole fraction of the more volatile component. EXAMPLE 1.7 A system is composed of 3 mol of isobutene and 1 mol of n-heptanes. The system is separated at a fixed temperature and pressure and the liquid and vapor phases recovered. The mole fraction of isobutene in the recovered liquid and vapor are 0.370 and 0.965, respectively. Calculate the number of moles of liquid nl and vapor nv recovered. Solution Step 1. Given x ¼ 0.370, y ¼ 0.965, and n ¼ 4, calculate the overall mole fraction of isobutane in the system: 3 z ¼ ¼ 0:750 4 100% 30 CHAPTER 1 Fundamentals of Hydrocarbon Phase Behavior Step 2. Solve for the number of moles of the vapor phase by applying Eq. (1.14): zx 0:750 0:370 nv ¼ n ¼4 ¼ 2:56mol of vapor yx 0:965 0:375 Step 3. Determine the quantity of liquid: nL ¼ n nv ¼ 4 2:56 ¼ 1:44mol of liquid The quantity of nL also could be obtained by substitution in Eq. (1.15): zy 0:750 0:965 nL ¼ n ¼4 ¼ 1:44 xy 0:375 0:965 If the composition is expressed in weight fraction instead of mole fraction, similar expressions to those expressed by Eqs. (1.14), (1.15) can be derived in terms of weights of liquid and vapor. Let mt ¼ total mass (weight) of the system mL ¼ total mass (weight) of the liquid mv ¼ total mass (weight) of the vapor wo ¼ weight fraction of the more volatile component in the original system wL ¼ weight fraction of the more volatile component in the liquid wv ¼ weight fraction of the more volatile component in the vapor A material balance on the more volatile component leads to the following equations: mv wo wL ¼ mt wv wL mL wo wv ¼ mt wL wv THREE-COMPONENT SYSTEMS The phase behavior of mixtures containing three components (ternary systems) is conveniently represented in a triangular diagram, such as that shown in Fig. 1.13. Such diagrams are based on the property of equilateral triangles that the sum of the perpendicular distances from any point to each side of the diagram is a constant and equal to the length on any of the sides. Thus, the composition xi of the ternary system as represented by point A in the interior of the triangle of Fig. 1.13 is Component 1 x1 ¼ L1 LT Component 2 x2 ¼ L2 LT Component 3 x3 ¼ L3 LT Three-Component Systems 31 L3 L2 t3 en on mp co 0% 0% co mp on en t2 100% component 1 L1 100% component 3 0% component 1 100% component 2 n FIGURE 1.13 Properties of the ternary diagram. where LT ¼ L1 + L2 + L3 Typical features of a ternary phase diagram for a system that exists in the two-phase region at fixed pressure and temperature are shown in Fig. 1.14. Any mixture with an overall composition that lies inside the binodal curve (phase envelope) will split into liquid and vapor phases. The line that connects the composition of liquid and vapor phases that are in equilibrium is called the tie line. Any other mixture with an overall composition that lies on that tie line will split into the same liquid and vapor compositions. Only the amounts of liquid and gas change as the overall mixture composition changes from the liquid side (bubble point curve) on the binodal curve to the vapor side (dew point curve). If the mole fractions of component i in the liquid, vapor, and overall mixture are xi, yi, and zi, the fraction of the total number of moles in the liquid phase nl is given by nl ¼ y i zi yi xi This expression is another lever rule, similar to that described for binary diagrams. The liquid and vapor portions of the binodal curve (phase envelope) meet at the plait point (critical point), where the liquid and vapor phases are identical. 32 CHAPTER 1 Fundamentals of Hydrocarbon Phase Behavior co V po apo sit r io n m “y” Two-phase region Ti e lin e 1 Liquid composition “x” Cri t “pl ical p ait o po int int ” Phase envelop “binodal curve” Single phase region 3 2 n FIGURE 1.14 Ternary phase diagram at a constant temperature and pressure for a system that forms a liquid and a vapor. MULTICOMPONENT SYSTEMS The phase behavior of multicomponent hydrocarbon systems in the twophase region, that is, the liquid-vapor region, is very similar to that of binary systems. However, as the system becomes more complex with a greater number of different components, the pressure and temperature ranges in which two phases lie increase significantly. The conditions under which these phases exist are a matter of considerable practical importance. The experimental or the mathematical determinations of these conditions are conveniently expressed in different types of diagrams, commonly called phase diagrams. One such diagram is called the pressure-temperature diagram. PRESSURE-TEMPERATURE DIAGRAM Fig. 1.15 shows a typical pressure-temperature diagram (p-T diagram) of a multicomponent system with a specific overall composition. Although a different hydrocarbon system would have a different phase diagram, the general configuration is similar. These multicomponent p-T diagrams are essentially used to classify reservoirs, specify the naturally occurring hydrocarbon systems, and describe the phase behavior of the reservoir fluid. Pressure-Temperature Diagram 33 4000 pcb D Liquid phase 2 Gas phase d i qu 3500 0% li 10 3000 (Pc)mixture 3 “C” Critical point rv e e rv cu int id po liqu w De 0% int cu 2500 lin Bu y bb lit le 2000 ua po Q es Pressure (psia) 1 1500 B 1000 Two-phase region E 500 0 –100 0 100 200 300 400 500 Temperature (°F) n FIGURE 1.15 Typical p-T diagram for a multicomponent system. To fully understand the significance of the p-T diagrams, it is necessary to identify and define the following key points on the p-T diagram: n n n n n Cricondentherm (Tct): The cricondentherm is the maximum temperature above which liquid cannot be formed regardless of pressure (point E). The corresponding pressure is termed the cricondentherm pressure, pct. Cricondenbar (pcb): The cricondenbar is the maximum pressure above which no gas can be formed regardless of temperature (point D). The corresponding temperature is called the cricondenbar temperature, Tcb. Critical point: The critical point for a multicomponent mixture is referred to as the state of pressure and temperature at which all intensive properties of the gas and liquid phases are equal (point C). At the critical point, the corresponding pressure and temperature are called the critical pressure, pc, and critical temperature, Tc, of the mixture. Phase envelope (two-phase region): The region enclosed by the bubble point curve and the dew point curve (line BCA), where gas and liquid coexist in equilibrium, is identified as the phase envelope of the hydrocarbon system. Quality lines: The lines within the phase diagram are called quality lines. They describe the pressure and temperature conditions for equal liquid percentage by volumes (% of liquid). Note that the quality lines converge at the critical point (point C). A 600 700 Tct (Tc)mixture 800 34 CHAPTER 1 Fundamentals of Hydrocarbon Phase Behavior n n Bubble point curve: The bubble point curve (line BC) is defined as the line separating the liquid phase region from the two-phase region. Dew point curve: The dew point curve (line AC) is defined as the line separating the vapor phase region from the two-phase region. CLASSIFICATION OF RESERVOIRS AND RESERVOIR FLUIDS Petroleum reservoirs are broadly classified as oil or gas reservoirs. These broad classifications are further subdivided depending on 1. The composition of the reservoir hydrocarbon mixture. 2. Initial reservoir pressure and temperature. 3. Pressure and temperature of the surface production. 4. Location of the reservoir temperature with respect to the critical temperature and the cricondentherm. In general, reservoirs are conveniently classified on the basis of the reservoir temperature “T” as compared with the critical temperature of the hydrocarbon mixture “(Tc)mix.” It should be noted that perhaps the most important components are the Methane and the Plus-Fraction, eg, C7+. In general, these two components impact the size of the phase envelope as well as the critical temperature of the mixture. Accordingly, reservoirs can be classified into basically three types: 1. Oil reservoirs: If the reservoir temperature, T, is less than the critical temperature, Tc, of the reservoir fluid, the reservoir is classified as an oil reservoir. 2. Gas reservoirs: If the reservoir temperature is greater than the critical temperature of the hydrocarbon fluid, the reservoir is considered a gas reservoir. 3. Near-critical reservoirs: If the reservoir temperature is very close to the critical temperature of the hydrocarbon fluid, the reservoir is classified as near-critical reservoir. Fig. 1.16 shows a flow diagram that summarizes the three reservoir classifications. The impact of fluid composition on the shape and size of the phase envelope as well as the reservoir type identification are illustrated in Fig. 1.17. Classification of Oil Reservoirs Depending on initial reservoir pressure, pi, oil reservoirs can be subclassified into the following categories: 1. Undersaturated oil reservoir: If the initial reservoir pressure, pi (as represented by point 1 in Fig. 1.15), is greater than the bubble point pressure, pb, of the reservoir fluid, the reservoir is an undersaturated oil reservoir. Classification of Reservoirs and Reservoir Fluids 35 Classification of reservoirs is defined by T & (Tc)hydrocarbon % of C1&C7+ in the system will impact (Tc)mixture (Tc)mixture = f(composition) Gas reservoirs T > (Tc)mixture Oil reservoirs T < (Tc)mixture Near critical reservoirs T ≈ (Tc)mixture Special case: Critical mixture separates the gas-cap from the oil rim Critical mixture T = (Tc)mixture n FIGURE 1.16 Reservoir classifications. 2. Saturated oil reservoir: When the initial reservoir pressure is equal to the bubble point pressure of the reservoir fluid, as shown in Fig. 1.15 by point 2, the reservoir is a saturated oil reservoir. 3. Gas-cap reservoir: When two phases (gas and oil) initially exit in equilibrium in a reservoir; the reservoir is defined as a gas-cap reservoir. Because the composition of the gas and oil zones are entirely different from each other, each phase will be represented separately by individual phase diagrams with the oil zone exits at its bubble point with the gas cap its dew point. Based on the gas-cap composition, the gas cap may be either retrograde gas or dry/wet gas, as illustrated in Fig. 1.18. It should be pointed out that the main condition for a reservoir to be initially considered in equilibrium in that the reservoir pressure “pi,” bubble point pressure “pb,” and dew point pressure “pd” are all equal at gas-oil contact (GOC) (ie, pi ¼ pb ¼ pd at GOC); this is an essential requirement to signify that the reservoir system is in equilibrium as shown in Fig. 1.20 Classification of Crude Oil Systems Crude oil is a mixture of small and large hydrocarbon. Depending on the mixture of hydrocarbon molecules, crude oil varies in color, composition, and consistency. Different oil-producing areas yield significantly different Composition defines the size of phase envelop Oil reservoir T < Tc Comp. Mix 2 Pc, psia Tc, °R 4000 CO2 0.000 1071 547.9 3500 N2 0.000 493 227.6 C1 0.350 667.8 343.37 C2 0.100 707.8 550.09 C3 0.060 616.3 666.01 i-C4 0.020 529.1 734.98 n-C4 0.020 550.7 765.65 i-C5 0.001 490.4 829.1 n-C5 0.002 488.6 845.7 C6 0.030 436.9 913.7 C7+ 0.417 320.3 1139.4 T Low % of methane High % of C7+ C Pressure (psia) 3000 2500 Two-phase region 2000 1500 1000 500 0 –100 TC 0 100 200 300 400 Temperature (°F) 500 600 700 800 Gas reservoir T > Tc Mix 1 Pc, psia Tc, °R CO2 0.009 1071 547.9 N2 0.003 493 227.6 C1 0.535 667.8 343.37 C2 0.115 707.8 550.09 C3 0.088 616.3 666.01 i-C4 0.023 529.1 734.98 n-C4 0.023 550.7 765.65 i-C5 0.015 490.4 829.1 n-C5 0.015 488.6 845.7 C6 0.015 436.9 913.7 C7+ 0.159 320.3 1139.4 High % of methane Low % of C7+ Liquid Pressure Comp. C 75 50 25 0 TC Temperature n FIGURE 1.17 Impact of composition on the size of phase envelope. Retrograde gas cap Dry or wet gas cap Pressure Pressure C Gas-cap phase envelope Oil zone phase envelope pb = pd = pi C Gas-cap phase envelope nt c urv e C pb = pd = pi Bubble point curve Oil zone phase envelope ble poi Dew point curve Bubble point curve Bub Dew point curve T Temperature n FIGURE 1.18 Classifications of the gas-cap reservoir. Temperature Dew point curve T C Classification of Reservoirs and Reservoir Fluids 37 varieties of crude oil that ranges from light to heavy crude oil systems. These variations in the type of crude oil are related to the ability of oil to flow (viscosity), color, and density. The term “sweet” is used to describe crude oil that is low in sulfur compounds such as hydrogen sulfide, and the term “sour” is used to describe crude oil containing high sulfur compounds. Crude oils cover a wide range in physical properties and chemical compositions, and it is often important to be able to group them into broad categories of related oils. In general, crude oils are commonly classified into the following types n n n n Ordinary black oil Low-shrinkage crude oil High-shrinkage (volatile) crude oil Near-critical crude oil. The above classification essentially is based on the physical and PVT properties exhibited by the crude oil, including n n n n n Physical properties, such as API gravity of the stock-tank liquid Composition Initial producing gas-oil ratio (GOR) Appearance, such as the color of the stock-tank liquid Pressure-temperature phase diagram. Three of the above properties generally are available that include initial GOR, API gravity, and color of the separated liquid. The initial producing GOR perhaps is the most important indicator of fluid type. Color has not been a reliable means of differentiating clearly between gas-condensates and volatile oils, but in general, dark colors indicate the presence of heavy hydrocarbons. No sharp dividing lines separate these categories of hydrocarbon systems, only laboratory studies could provide the proper classification. In general, reservoir temperature and composition of the hydrocarbon system greatly influence the behavior of the system. The fluid physical characteristics of the four crude oil systems are summarized below. 1. Ordinary black oil: A typical p-T phase diagram for ordinary black oil is shown in Fig. 1.19. Note that quality lines that are approximately equally spaced characterize this black oil phase diagram. Following the pressure reduction path, as indicated by the vertical line EF in Fig. 1.19, the liquid-shrinkage curve, shown in Fig. 1.20, is prepared by plotting the 38 CHAPTER 1 Fundamentals of Hydrocarbon Phase Behavior Ordinary black oil Quality lines are evenly spaced Liquid phase 4000 1 Reservoir pressure decline path E 3500 Gas phase 00% 1 80% rv e C % cu 50 po int 2500 % 25 F Bu bb le 2000 rve cu int d po iqui w L De 0% Pressure (psia) 3000 1500 A 1000 Two-phase region E Separator 500 0% 0 –100 B 0 100 200 300 400 Temperature (°F) 500 600 700 Tct 800 (Tc)mixture n FIGURE 1.19 A typical p-T diagram for an ordinary black oil. 100% Liquid volume E F Residual oil 0% Pressure n FIGURE 1.20 Liquid-shrinkage curve for black oil. Bubble point pressure pb Classification of Reservoirs and Reservoir Fluids 39 liquid volume percent as a function of pressure. The liquid-shrinkage curve approximates a straight line except at very low pressures. When produced, ordinary black oils usually yield GORs between 200 and 700 scf/STB and oil gravities of 15–40 API. The stock-tank oil usually is brown to dark green in color. 2. Low-shrinkage oil: A typical p-T phase diagram for low-shrinkage oil is shown in Fig. 1.21. The diagram is characterized by quality lines that are closely spaced near the dew point curve. The liquid-shrinkage curve, given in Fig. 1.22, shows the shrinkage characteristics of this category of crude oils. The other associated properties of this type of crude oil are n Oil formation volume factor less than 1.2 bbl/STB. n GOR less than 200 scf/STB. n Oil gravity less than 35° API. n Black or deeply colored. n Substantial liquid recovery at separator conditions as indicated by point G on the 85% quality line of Fig. 1.22. 3. Volatile crude oil: The phase diagram for a volatile (high-shrinkage) crude oil is given in Fig. 1.23. Note that the quality lines are close together near the bubble point, and at lower pressures they are more widely spaced. This type of crude oil is commonly characterized by a high liquid shrinkage Quality lines are distant from the bubble point curve x Reservoir pressure XE decline path 0% 10 Liquid Gas Two-phase region C Pressure % 95 % 85 % A 70 xF % 50 Separator 0% B Temperature n FIGURE 1.21 A typical phase diagram for a low-shrinkage oil. 40 CHAPTER 1 Fundamentals of Hydrocarbon Phase Behavior 100% E Liquid volume F Residual oil 0% Pressure Bubble point pressure pb n FIGURE 1.22 Oil-shrinkage curve for low-shrinkage oil. n FIGURE 1.23 A typical p-T diagram for a volatile crude oil. immediately below the bubble point, shown in Fig. 1.24. The other characteristic properties of this oil include n Oil formation volume factor greater than 1.5 bbl/STB. n GORs between 2000 and 3000 scf/STB. n Oil gravities between 45° and 55° API. n Lower liquid recovery of separator conditions, as indicated by point G in Fig. 1.23. n Greenish to orange in color. Classification of Reservoirs and Reservoir Fluids 41 Liquid volume 100% E F Residual oil 0% Pressure Bubble point pressure pb n FIGURE 1.24 A typical liquid-shrinkage curve for a volatile crude oil. Solution gas released from a volatile oil contains significant quantities of stock-tank liquid (condensate) when the solution gas is produced at the surface. Solution gas from black oils usually is considered “dry,” yielding insignificant stock-tank liquid when produced to surface conditions. For engineering calculations, the liquid content of released solution gas perhaps is the most important distinction between volatile oils and black oils. Another characteristic of volatile oil reservoirs is that the API gravity of the stock-tank liquid increases in the later life of the reservoirs. 4. Near-critical crude oil: If the reservoir temperature, T, is near the critical temperature, Tc, of the hydrocarbon system, as shown in Fig. 1.25, the hydrocarbon mixture is identified as a near-critical crude oil. Because all the quality lines converge at the critical point, an isothermal pressure drop (as shown by the vertical line EF in Fig. 1.25) may shrink the crude oil from 100% of the hydrocarbon pore volume at the bubble point to 55% or less at a pressure 10–50 psi below the bubble point. The shrinkage characteristic behavior of the near-critical crude oil is shown in Fig. 1.26. This high shrinkage creates high gas saturation in the pore space and because of the gas-oil relative permeability characteristics of most reservoir rocks; free gas achieves high mobility almost immediately below the bubble point pressure. The near-critical crude oil is characterized by a high GOR, in excess of 3000 scf/STB, with an oil formation volume factor of 2.0 bbl/STB or 42 CHAPTER 1 Fundamentals of Hydrocarbon Phase Behavior 4000 Near critical oil 0% liq 10 E cu Two-phase region int 2500 % 80 po % 50 Bu bb le 2000 C rve cu int d po iqui l w De 0% rve 3000 Pressure (psia) Reservoir pressure decline path d ui 3500 1500 % 25 A F 1000 500 Separator 0% B 0 –100 – 0 Temperature (ºF) n FIGURE 1.25 A schematic phase diagram for the near-critical crude oil. Liquid volume 100% E ≥ 50% F Residual oil 0% Pressure n FIGURE 1.26 A typical liquid-shrinkage curve for the near-critical crude oil. Bubble point pressure pb Classification of Reservoirs and Reservoir Fluids 43 How near are the quality lines to the bubblepoint curve will determine the type of crude oil 100% E ge oil -shrinka (A) Low Liquid volume ary din ) Or k oil balc (B il age o hrink igh-s (C) H ear (D) N al oil critic 0% Pressure Bubble point pressure pb n FIGURE 1.27 Liquid shrinkage for crude oil systems. higher. The compositions of near-critical oils usually are characterized by 12.5–20 mol% heptanes-plus, 35% or more of ethane through hexanes, and the remainder methane. It should be pointed out that nearcritical oil systems essentially are considered the borderline to very rich gas-condensates on the phase diagram. Fig. 1.27 compares the characteristic shape of the liquid-shrinkage curve for each crude oil type. Gas Reservoirs In general, if the reservoir temperature is above the critical temperature of the hydrocarbon system, the reservoir is classified as a natural gas reservoir. Natural gases can be categorized on the basis of their phase diagram and the prevailing reservoir condition into four categories: 1. Retrograde gas reservoirs. 2. Near-critical gas-condensate reservoirs. 3. Wet gas reservoirs. 4. Dry gas reservoirs. In some cases, when condensate (stock-tank liquid) is recovered from a surface process facility, the reservoir is mistakenly classified as a retrograde gas reservoir. Strictly speaking, the definition of a retrograde gas reservoir depends only on reservoir temperature. 44 CHAPTER 1 Fundamentals of Hydrocarbon Phase Behavior Retrograde Gas Reservoirs If the reservoir temperature, T, lies between the critical temperature, Tc, and cricondentherm, Tct, of the reservoir fluid, the reservoir is classified as a retrograde gas-condensate reservoir. This category of gas reservoir has a unique type of hydrocarbon accumulation, in that the special thermodynamic behavior of the reservoir fluid is the controlling factor in the development and the depletion process of the reservoir. When the pressure is decreased on these mixtures, instead of expanding (if a gas) or vaporizing (if a liquid) as might be expected, they vaporize instead of condensing. Consider that the initial condition of a retrograde gas reservoir is represented by point 1 on the pressure-temperature phase diagram of Fig. 1.28. Because the reservoir pressure is above the upper dew point pressure, the hydrocarbon system exists as a single phase (ie, vapor phase) in the reservoir. As the reservoir pressure declines isothermally during production from the initial pressure (point 1) to the upper dew point pressure (point 2), the attraction between the molecules of the light and heavy components move farther apart. As this occurs, attraction between the heavy component molecules becomes more effective; therefore, liquid begins to condense. This retrograde condensation process continues with decreasing pressure until the liquid dropout (LDO) reaches its maximum at point 3. Further reduction in pressure permits the heavy molecules to commence the normal vaporization process. This is the process whereby fewer gas molecules strike the liquid surface and more molecules leave than enter the liquid phase. The vaporization process continues until the reservoir pressure n FIGURE 1.28 A typical phase diagram of a retrograde system. Classification of Reservoirs and Reservoir Fluids 45 n FIGURE 1.29 A typical liquid dropout curve. reaches the lower dew point pressure. This means that all the liquid that formed must vaporize because the system essentially is all vapor at the lower dew point. Fig. 1.29 shows a typical liquid-shrinkage volume curve for a relatively rich condensate system. The curve is commonly called the liquid dropout curve. The maximum LDO is 26.5%, which occurs when the reservoir pressure drops from a dew point pressure of 5900–2800 psi. In most gas-condensate reservoirs, the condensed liquid volume seldom exceeds more than 15–19% of the pore volume. This liquid saturation is not large enough to allow any liquid flow. It should be recognized, however, that around the well bore, where the pressure drop is high, enough LDO might accumulate to give two-phase flow of gas and retrograde liquid. The associated physical characteristics of this category are n n n GORs between 8000 and 70,000 scf/STB. Generally, the GOR for a condensate system increases with time due to the LDO and the loss of heavy components in the liquid. Condensate gravity above 50° API. Stock-tank liquid is usually water-white or slightly colored. It should be pointed out that the gas that comes out of the solution from a volatile oil and remains in the reservoir typically is classified a retrograde gas and exhibits the retrograde condensate with pressure declines. 46 CHAPTER 1 Fundamentals of Hydrocarbon Phase Behavior There is a fairly sharp dividing line between oils and condensates from a compositional standpoint. Reservoir fluids that contain heptanes and are in concentration of more than 12.5 mol% almost always are in the liquid phase in the reservoir. Oils have been observed with heptanes and heavier concentrations as low as 10% and condensates as high as 15.5%. These cases are rare, however, and usually have very high tank liquid gravities. Near-Critical Gas-Condensate Reservoirs If the reservoir temperature is near the critical temperature, as shown in Fig. 1.30, the hydrocarbon mixture is classified as a near-critical gascondensate. The volumetric behavior of this category of natural gas is described through the isothermal pressure declines, as shown by the vertical line 1–3 in Fig. 1.30 and the corresponding LDO curve of Fig. 1.31. Because all the quality lines converge at the critical point as shown in Fig. 1.31, a rapid liquid buildup immediately occurs below the dew point as the pressure is reduced to point 2. This behavior can be justified by the fact that several quality lines are crossed very rapidly by the isothermal reduction in pressure. At the point where the liquid ceases to build up and begins to shrink again, the reservoir goes from the retrograde region to a normal vaporization region. Liquid phase Gas phase Reservoir pressure e urv nt c decline path poi D Pressure 1 Critical point “C” e v ur tc 0% ew 20% 50% 0% 2 n oi le p ve Quality lines b ub ur tc B n oi w 3 De Temperature n FIGURE 1.30 A typical phase diagram for a near-critical gas-condensate reservoir. p Classification of Reservoirs and Reservoir Fluids 47 Liquid volume 100% ≈50% 3 2 ropout d Liquid curve 1 0% Pressure n FIGURE 1.31 Liquid-shrinkage (dropout) curve for a near-critical gas-condensate system. Wet Gas Reservoirs A typical phase diagram of a wet gas is shown in Fig. 1.32, where the reservoir temperature is above the cricondentherm of the hydrocarbon mixture. Because the reservoir temperature exceeds the cricondentherm of the hydrocarbon system, the reservoir fluid always remains in the vapor phase region as the reservoir is depleted isothermally, along the vertical line AB. Liquid phase Gas phase Reservoir pressure decline path at constant temperature A Pressure Critical point “C” Two-phase region Liquid 75 50 25 5 0 Separator Temperature n FIGURE 1.32 Phase diagram for a wet gas. Gas B 48 CHAPTER 1 Fundamentals of Hydrocarbon Phase Behavior However, as the produced gas flows to the surface, the pressure and temperature of the gas decline. If the gas enters the two-phase region, a liquid phase condenses out of the gas and is produced from the surface separators. This is caused by a sufficient decrease in the kinetic energy of heavy molecules with temperature drop and their subsequent change to liquid through the attractive forces between molecules. Wet gas reservoirs are characterized by the following properties n n n n GORs between 60,000 and 100,000 scf/STB. Stock-tank oil gravity above 60° API. Liquid is water-white in color. Separator conditions (ie, separator pressure and temperature) lie within the two-phase region. It should be noted that when applying the material balance equation for a wet gas. Reservoir, as expressed in terms of cumulative gas production “GP” as a function of p/Z, the cumulative surface condensate volume “Np STB” or condensate production rate “Qo STB/day” must be converted into an equivalent gas volume and added to the cumulative separator gas production “GP” or to the separator gas production rate qgas. The equivalent gas volume “Veq” as expressed in scf/STB, is an important property that can be used to convert a liquid volume into an equivalent gas volume. The working basic equations to develop an expression for Veq are based on the definition of the number of moles “n” and the gas EOS Number of moles n: The number of moles (liquid or gas) is defined as the ratio of mass “m” to the molecular weight “M” with the fluid mass “m” given by fluid volume times its density. Applying the definition to the liquid phase at stock-tank condition, gives no ¼ mo Vo ρo ð5:615Þð62:4γ o Þ ¼ ¼ Mo Mo Mo Or: no ¼ mo Vo ρo ð5:615Þð62:4γ o Þ 350:376γ o ¼ ¼ ¼ Mo Mo Mo Mo where Vo ¼ liquid volume, STB Mo ¼ molecular weight of liquid γ o ¼ liquid specific gravity, 60°/60° Classification of Reservoirs and Reservoir Fluids 49 Gas EOS: The gas EOS is expressed by PV ¼ ZnRT Solving the EOS for volume “V” at stock-tank conditions, ie, at 14.7 psi and 60°F, with gas constant “R” of the value 10.73, gives Vsc ¼ Zsc RTn ð1Þð10:73Þð60 + 460Þn ¼ psc 14:7 The above indicates that 1 mol occupies an equivalent volume of 379.4 scf at standard conditions. This suggests that no moles of liquid can be converted into an equivalent gas volume “Veq” by the expression: 350:376γ o γ 133,000 o Veq ¼ 379:4no ¼ 379:4 Mo Mo where Veq ¼ equivalent gas volume in scf/STB The liquid specific gravity γ o is related to the API gravity by the expression: γo ¼ 141:5 API + 131:5 For condensate and light hydrocarbon systems; the molecular weight “Mo” can be approximated in terms of the API gravity by the following relationship: Mo ¼ 6084 API 5:9 Combining the above two expressions and solving for the ratio γ o/Mo, gives γo ¼ 0:001892 + 7:35 105 API 4:52 108 ðAPIÞ2 Mo The ratio γ o/Mo in the equivalent gas volume equation can be replaced by the above expression to give Veq ¼ 252 + 9:776API 0:006ðAPIÞ2 The following example illustrates the practical use of the equivalent gas volume in converting liquid volume into equivalent gas volume. 50 CHAPTER 1 Fundamentals of Hydrocarbon Phase Behavior EXAMPLE 1.8 The following data is available on a wet-shale gas reservoir: pi ¼ 3200psia T ¼ 200°F Φ ¼ 8% Swi ¼ 30% Condensate daily flow rate Qo ¼ 400 STB/day API gravity of the condensate ¼ 50° Daily HP separator gas rate qgsep ¼ 4.20 MMscf/day Daily IP separator gas rate qgsep ¼ 1.20 MMscf/day Average separator gas specific gravity ¼ 0.65 Daily stock-tank gas rate qgst ¼ 0.15 MMscf/day Based on the configuration of the surface facilities as shown below, calculate the daily well stream gas flow rate in MMscf/day 4.2 MMscf/day 1.2 MMscf/day 0.15 MMscf/day Qo= 400 STB/day Gas Gas Solution Step 1. Calculate the equivalent gas volume “Veq”: Veq ¼ 252 + 9:776API 0:006ðAPIÞ2 Veq ¼ 252 + ð9:776Þð50Þ 0:006ð50Þ2 ¼ 725:8scf=STB Step 2. Calculate total well rate “Qg”: Qg ¼ 4:2 + 1:2 + 0:15 + ð400Þð725:8=1:0e6Þ ¼ 5:840MMscf=day The equivalent gas volume equation can be further simplified to combine with the above. Dry Gas Reservoirs The hydrocarbon mixture exists as a gas both in the reservoir and the surface facilities. The only liquid associated with the gas from a dry gas reservoir is water. Fig. 1.33 is a phase diagram of a dry gas reservoir. Usually, a system Classification of Reservoirs and Reservoir Fluids 51 Gas phase Liquid phase Reservoir pressure decline path at constant temperature A Pressure Liquid Critical point “C” Separator 75 50 25 0 Gas Temperature n FIGURE 1.33 Phase diagram for a dry gas. that has a GOR greater than 100,000 scf/STB is considered to be a dry gas. The kinetic energy of the mixture is so high and attraction between molecules so small that none of them coalesce to a liquid at stock-tank conditions of temperature and pressure. It should be pointed out that the listed classifications of hydrocarbon fluids might be also characterized by the initial composition of the system. It should be pointed out that the heavy components (eg, C7+) in the hydrocarbon mixtures have the strongest effect on fluid characteristics. Numerous laboratory compositional analyses on retrograde and crude oil samples suggests that the phase boundary separating the retrograde gas phase and other fluid systems is based on the overall mole % of C7+ in the hydrocarbon system. In general, when Mole % C7+ < 12.5: the hydrocarbon system is classified as gas phase Mole % C7+ 12.5–20: the hydrocarbon system is classified as near critical Mole % C7+ > 20: the hydrocarbon system is classified as oil phase The ternary diagram shown in Fig. 1.34 with equilateral triangles can be conveniently used to roughly define the compositional phase boundaries that separate different types of hydrocarbon systems. Fluid samples obtained from a new field discovery may be instrumental in defining the existence of a two-phase, that is, gas-cap system with an overlying gas cap or underlying oil rim. As the compositions of the gas and oil zones are completely different from each other, both systems may be represented separately by individual phase diagrams, which bear little relation to B 52 CHAPTER 1 Fundamentals of Hydrocarbon Phase Behavior C1+N2 A A: Dry gas B: Wet gas C: Retrograde gas D: Near critical E: Volatile oil F: Ordinary oil G: Low shrinkage oil D C B E F G C7+ C2-C6+CO2 n FIGURE 1.34 Compositions of various reservoir fluid systems. each other or to the composite. The oil zone will be at its bubble point and produced as a saturated oil reservoir but modified by the presence of the gas cap. Depending on the composition and phase diagram of the gas, the gascap gas may be a retrograde gas cap, as shown in Fig. 1.35, or dry or wet, as shown in Fig. 1.36. Therefore, a discovery well drilled through a saturated reservoir fluid usually requires further field delineation to substantiate the presence of a second equilibrium phase above (ie, gas cap) or below (ie, oil rim) the tested well. This may entail running a repeat formation tester (RFT) tool to determine fluid-pressure gradient as a function of depth; or a new well drilled updip or downdip to that of the discovery well may be required. When several samples are collected at various depths, they exhibit PVT properties as a function of the depth expressed graphically to locate the GOC. The variations of PVT properties can be expressed graphically in terms of the compositional changes of C1 and C7+ with depth and in terms of well-bore and stock-tank densities with depth, as shown in Table 1.3 and Figs. 1.37–1.39. Customized plot 4000 Pb 3500 Phase envelope of the oil rim At GOC Pb = Pd 3000 Pd Pressure (psia) 2500 2000 1500 1000 500 0 −100 0 100 200 300 400 500 600 700 Temperature (°F) n FIGURE 1.35 Retrograde gas-cap reservoir. Phase envelope of the gas-cap gas Pressure Critical point C Gas-oil contact Phase envelope of the oil rim P C Critical point T Temperature n FIGURE 1.36 Dry gas-cap reservoir. 800 Table 1.3 Locating the GOC Depth (ft) Pressure (psia) Temperature (°F) Phase Saturation Pressure (psia) Molecular Weight Density (lb/ft3) GOR (scf/STB) STO Density (lb/ft3) Viscosity (cP) 5200.0 5248.3 5296.6 5344.8 5393.1 5441.4 5489.7 5537.9 5586.2 5634.5 5682.8 5731.0 5779.3 5827.6 5875.9 5924.1 5972.4 6020.7 6034.4 6034.4 6069.0 6117.2 6165.5 6213.8 6262.1 6286.0 6310.3 6358.6 6406.9 6455.2 6503.4 6551.7 6600.0 3586.8 3591.6 3596.5 3601.4 3606.3 3611.1 3616.1 3621.0 3625.9 3630.9 3635.9 3640.9 3645.9 3650.9 3656.0 3661.1 3666.2 3671.4 3672.9 3672.9 3681.2 3693.2 3705.4 3717.9 3730.6 3737.0 3743.5 3756.6 3769.9 3783.4 3796.9 3810.7 3824.5 163.71 164.43 165.16 165.88 166.61 167.33 168.05 168.78 169.50 170.23 170.95 171.68 172.40 173.12 173.85 174.57 175.30 176.02 176.23 176.23 176.74 177.47 178.19 178.92 179.64 180.00 180.37 181.09 181.81 182.54 183.26 183.99 184.71 Gas Gas Gas Gas Gas Gas Gas Gas Gas Gas Gas Gas Gas Gas Gas Gas Gas Gas Gas Oil Oil Oil Oil Oil Oil Oil Oil Oil Oil Oil Oil Oil Oil 3070.0 3089.9 3110.9 3133.1 3156.6 3181.5 3208.0 3236.1 3266.1 3298.1 3332.4 3369.0 3408.3 3450.6 3496.0 3545.1 3598.1 3655.7 3672.8 3673.0 3640.0 3597.5 3557.9 3520.2 3483.8 3466.2 3448.6 3414.2 3380.5 3347.4 3314.8 3282.7 3251.1 23.0 23.1 23.1 23.2 23.3 23.3 23.4 23.5 23.6 23.7 23.8 23.8 23.9 24.1 24.2 24.3 24.4 24.6 24.6 60.1 62.7 66.3 70.0 73.7 77.5 79.4 81.3 85.1 89.0 92.8 96.7 100.5 104.4 14.4365 14.4714 14.5079 14.5462 14.5865 14.6288 14.6735 14.7207 14.7709 14.8242 14.8812 14.9423 15.0082 15.0796 15.1575 15.2430 15.3379 15.4443 15.4767 34.4996 35.1887 36.0601 36.8474 37.5643 38.2196 38.5237 38.8198 39.3702 39.8754 40.3393 40.7656 41.1576 41.5185 43485.8 42419.5 41365.4 40322.1 39288.6 38263.5 37245.5 36233.1 35224.5 34218.1 33211.7 32203.1 31189.6 30168.0 29134.8 28085.5 27014.6 25914.9 25599.2 1793.4 1656.5 1495.0 1359.2 1243.2 1142.9 1098.1 1055.5 978.8 911.1 851.1 797.6 749.7 706.7 48.2664 48.2697 48.2733 48.2771 48.2814 48.2859 48.2910 48.2964 48.3025 48.3091 48.3164 48.3245 48.3334 48.3434 48.3545 48.3670 48.3810 48.3968 48.4016 49.2422 49.3460 49.4877 49.6253 49.7583 49.8863 49.9478 50.0089 50.1257 50.2367 50.3419 50.4413 50.5352 50.6236 0.0278 0.0279 0.0280 0.0281 0.0281 0.0282 0.0283 0.0284 0.0285 0.0286 0.0287 0.0288 0.0290 0.0291 0.0293 0.0295 0.0296 0.0299 0.0299 0.1404 0.1558 0.1800 0.2076 0.2391 0.2750 0.2946 0.3156 0.3612 0.4119 0.4678 0.5290 0.5954 0.6670 Determination of GOC from pressure gradients 7500. De w po in 7600. oir Reser v tp re s su re g pressu 7700. ; psi / ft 7800. ient Depth/ft re grad g 7900. Sat PRes P pd = pb = pi 8000. o psi / ft o 8100. 3300 3400 3500 3600 t in po e le ur bb ss u B pre 3700 3800 Pressure (psia) n FIGURE 1.37 Determination of gas-oil contact from pressure gradients. Saturation pressure & %C7+ Saturation pressure & %C1 pd pd C1 C7+ GOC pb Less gas in solution, ie, less GOR pb Depth n FIGURE 1.38 Compositional changes of C1 and C7+ with depth. Less gas in solution, ie, less GOR Depth 3900 56 CHAPTER 1 Fundamentals of Hydrocarbon Phase Behavior Stock-tank density Depth GOC Bottom-hole density Density n FIGURE 1.39 Variation of well-bore and stock-tank density with depth. Defining the fluid contacts, that is, GOC and water-oil contact (WOC), are extremely important when determining the hydrocarbon initially in place and planning field development. The uncertainty in the location of the fluid contacts can have a significant impact on the reserves estimate. Contacts can be determined by 1. Electrical logs, such as resistively tools. 2. Pressure measurements, such as a repeat formation tester (RFT) or a modular formation dynamic tester (MDT). 3. Possibly by interpreting seismic data. Normally, unless a well penetrates a fluid contact directly, there remains doubt as to its locations. The RFT is a proprietary name used by Schlumberger for an open-hole logging tool used to establish vertical pressure distribution in the reservoir (ie, it provides a pressure-depth profile in the reservoir) and to obtain fluid samples. At the appraisal stage of a new field, the RFT survey provides the best-quality pressure data and routinely is run to establish fluid contacts. The surveys are usually straightforward to interpret compared to the drill stem tests (DSTs), because no complex buildup analysis is required to determine the reservoir pressure nor are any extensive depth corrections to be applied, as the gauge depth is practically coincidental with that of the RFT probe. Classification of Reservoirs and Reservoir Fluids 57 The RFT tool is fitted with a pressure transducer and positioned across the target zone. The device is placed against the side of the borehole by a packer. A probe that consists of two pretest chambers, each fitted with a piston, is pushed against the formation and through the mud cake. A pressure drawdown is created at the probe by withdrawing the pistons in the pretest chambers. These two chambers operate in series: The first piston is withdrawn slowly (taking about 4 s to withdraw 10 cm3 of fluid); the second piston is withdrawn at a faster rate, 5 s for 10 cm3. Before the tool is set against the formation, the pressure transducer records the mud pressure at the target depth. As the first pretest chamber is filled (slowly) with fluid, the first main pressure drop Δp1 is observed, followed by Δp2 as the second pretest chamber is filled. The tighter the formation, the larger Δp1 and Δp2 are. The recording, therefore, gives a qualitative indication of the permeability of the reservoir. Because the flow rate and pressure drop are known, the actual formation permeability can be calculated, and this is done by the CSU (the computer in the logging unit). Caution should be attached to the permeability values, because a very small part of the reservoir is being tested and the analysis assumes a particular flow regime, which may not be representative in some cases. The reliability of the values should be regarded as indicating the order of magnitude of the permeability. Once both drawdowns have occurred, fluid withdrawal stops and the formation pressure is allowed to recover. This recovery is recorded as the pressure buildup, which should stabilize at the true formation fluid pressure at the depth. Note that the drawdown pressure drops, Δp1 and Δp2, are relative to the final pressure buildup and not to the initial pressure, which was the mud pressure. The rate of pressure buildup can be used to estimate permeability; the test therefore gives three permeability estimates at the same sample point. A tight formation leads to very large pressure drawdowns and a slow buildup. If the pressure has not built up within 4–5 min, the pressure test usually is abandoned for fear of the tool becoming stuck. The tool is retracted and the process repeated at a new depth. Once the tool is unset, the pressure reading should return to the same mud pressure as prior to setting. This is used as a quality control check on gauge drift. There is no limit as to the number of depths at which pressure samples may be taken. If a fluid sample is required, this can be done by diverting the fluid flow during sampling to sample chambers in the tool. 58 CHAPTER 1 Fundamentals of Hydrocarbon Phase Behavior It should be pointed out that, when plotting the pressure against depth, the unit of pressures must be kept consistent; that is, either in absolute pressure (psia) or gauge pressure (psig). The depth must be the true vertical depth, preferably below the subsurface datum depth. The basic principle of pressure versus depth plotting is illustrated in Fig. 1.40. This illustration shows two wells: n n Well 1 penetrates the gas cap with a recorded gas pressure of pg and a measured gas density of ρg. Well 2 penetrates the oil zone with a recorded oil pressure of po and oil density of ρo. The gas, oil, and water gradients can be calculated from: dpg ρg ¼ ¼ γg dh 144 dpo ρo ¼ ¼ γo dh 144 dpw ρw ¼ ¼ γw dh 144 Well-1 Pressure (dpg/dh)gas Well-2 Known gas column GOC Depth GOC Known oil column (dpo/dh)oil WOC WOC Oil-down-to “ODT” Water gradient 0.45 psi/ft n FIGURE 1.40 Pressure gradient vs. depth. Fault Gas-Oil Contact 59 where dp/dh ¼ fluid gradient, psi/ft γ g ¼ gas gradient, psi/ft γ o ¼ oil gradient, psi/ft γ w ¼ water gradient, psi/ft ρg ¼ gas density, lb/ft3 ρo ¼ oil density, lb/ft3 ρw ¼ water density, lb/ft3 GAS-OIL CONTACT As shown in Fig. 1.40, the intersection of the gas and oil gradient lines locates the position of the GOC. Note that the oil can be seen only down to a level termed the oil-down-to (ODT) level. The level is assigned to the oil-water contact unless a well is drilled further downdip to locate the WOC. Fig. 1.41 shows another case where a well penetrated an oil column. Oil can be seen down to the ODT level. However, a gas cap may exist. If the pressure of the oil is recorded as po and the bubble point pressure is measured as pb at this specific depth, the GOC can be determined from A discovery well penetrates an oil column what is the possibility of a gas cap updip ? Gas zone ! GOC ? o b D o o po Known oil column ; psi / ft o n FIGURE 1.41 A well in the oil rim. WOC ? Oil-down-to “ODT” 60 CHAPTER 1 Fundamentals of Hydrocarbon Phase Behavior ΔD ¼ po pb po pb ¼ ðdp=dhÞoil γo (1.18) If the calculated value of ΔD locates the GOC within the reservoir, then there is a possibility that a gas cap exists, but this is not certain. Eq. (1.18) assumes that the PVT properties, including the bubble point pressure, do not vary with depth. The only way to be certain that a gas cap exists is to drill a crestal well. A similar scenario that can be encountered is illustrated in Fig. 1.42. Here, a discovery well penetrates a gas column with gas as seen down to the gasdown-to (GDT) level, which allows the possibility of a downdip oil rim. If the field operator feels that an oil rim exists, based on similarity with other fields in the area, the distance to GOC can be estimated from ΔD ¼ pb pg pb pg ¼ ðdp=dhÞgas γg A discovery well penetrates a gas column what is the possibility of an oil rim? pg b Gas-down-to “GDT” g D g GOC ? g ; psi / ft g n FIGURE 1.42 A discovery well in gas column. Oil zone ! Gas-Oil Contact 61 Well-1 Oil gradient ª 0.33 psi/ft Known-gas-column Well-2 Possible GOC Gas-down-to Possible oil zone ΔD Possible OWC Possible GWC b a n FIGURE 1.43 Two discovery wells. Fig. 1.43 shows a case where two wells are drilled: The first well is in the gas column and the second penetrates the water zone. Of course, only two possibilities could exist: 1. The first possibility is shown in Fig. 1.43, which suggests a known gas column with an underlying water zone. Also, the gas-water contact (GWC) has not been established. A “possible” GWC can be calculated from the following relationship: D1 ¼ pg pw ΔDðdp=dhÞwater pg ½pw ΔDγ water ¼ γ water γ gas ðdp=dhÞwater ðdp=dhÞgas γ ¼ fluid gradient, psi/ft, that is, ρ/144 ρ ¼ fluid density, lb/ft3 D1 ¼ vertical distance between gas well and GWC, ft ΔD ¼ vertical distance between the two wells dp/dh ¼ γ ¼ fluid gradient, psi/ft 62 CHAPTER 1 Fundamentals of Hydrocarbon Phase Behavior 2. The second possibility is shown in Fig. 1.43, where a possible oil zone exists between the two wells. Assuming a range of oil gradients, such as 0.28–0.38 psi/ft, the maximum possible oil thickness can be estimated graphically as shown in Fig. 1.45B, by drawing a line originating at the GDT level with a slope in the range of 0.28–0.38 psi/ft to intersect with the water gradient line. The illustration in Fig. 1.44 shows again two discovery wells; one well penetrated the gas cap and the other well penetrated the water zone, with a vertical distance between the two wells of ΔD. If the bubble point pressure is known or assumed, the maximum oil column thickness as well as the depth to different contacts can be estimated, from the following expressions: Possible depth to GOC “D1” ΔD1 ¼ pb pg @pg =@D g Possible depth to WOC “D2” ΔD3 ¼ ðpw pb Þ ½ð@po =@DÞðΔD ΔD1 Þ ½ð@pw =@DÞ ð@po =@DÞ o ; psi / ft g ; psi / ft w ; psi / ft o Two discovery wells Maximum Possible oil thickness g w Gas DD1 g g g DD DD3 w b o w n FIGURE 1.44 Two discovery wells — maximum possible oil thickness. b o Gas-Oil Contact 63 Possible maximum oil zone thickness “D2” ΔDðdpw =dhÞ ðpw pb Þ pb pg ðdpw =dhÞ= dpg =dh ΔD2 ¼ ðdpw =dhÞ ðdpo =dhÞ ΔDγ w ðpw pb Þ pb pg γ w =γ g ¼ γw γo Or simply ΔD2 ¼ ΔD ΔD1 ΔD3 where ΔD2 ¼ maximum possible oil column thickness, ft ΔD ¼ vertical distance between the two wells, ft dp/dh ¼ fluid-pressure gradient, psi/ft EXAMPLE 1.9 Fig. 1.45 shows two wells, one penetrated the gas zone at 4950 ft true vertical depth (TVD) and the other in the water zone at 5150 ft. Pressures and PVT data follow: pg ¼ 1745psi ρg ¼ 14:4lb=ft3 pw ¼ 1808psi ρw ¼ 57:6lb=ft3 A nearby field has a bubble point pressure of 1750 psi and an oil density of 50.4 lb/ft3 (ρo ¼ 50.4 lb/ft3). Calculate 1. Possible depth to GOC. 2. Maximum possible oil column thickness. 3. Possible depth to WOC. Solution Step 1. Calculate the pressure gradient of each phase: ρg 14:4 dp ¼ ¼ ¼ 0:1psi=ft Gas phase dh gas 144 144 dp ρ 50:4 Oil phase ¼ o ¼ ¼ 0:35psi=ft dh oil 144 144 dp ρ 57:6 ¼ w ¼ ¼ 0:4psi=ft Water phase dh water 144 144 Step 2. Calculate the distance to the possible GOC: ΔD1 ¼ pb pg 1750 1745 ¼ ¼ 50ft 0:1 @pg =@D g GOC ¼ 4950 + 50 ¼ 5000ft 64 CHAPTER 1 Fundamentals of Hydrocarbon Phase Behavior o ; psi / ft g ; psi / ft w ; psi / ft o g w Gas zone DD1 DD2 Possible oil rim DD3 n FIGURE 1.45 Illustration for Example 1.9. Step 3. Possible depth to WOC: ΔD3 ¼ ðpw pb Þ ð@po =@DÞðΔD ΔD1 ð1808 1750Þ 0:35ð200 50Þ ¼ ½ð@pw =@DÞ ð@po =@DÞ 0:4 0:35 ¼ 110ft WOC ¼ 5150 110 ¼ 5040ft Step 4. Calculate possible maximum oil rim thickness: ΔDγ w ðpw pb Þ pb pg γ w =γ g ΔD2 ¼ γw γo ΔD2 ¼ maximum oil thickness ¼ 200ð0:4Þ ð1808 1750Þ ð1750 1745Þð0:4=0:1Þ ¼ 40ft ð0:4Þ ð0:35Þ or simply as ΔD2 ¼ ΔD ΔD2 ΔD3 ¼ 200 50 110 ¼ 40ft Phase Rule 65 UNDERSATURATED GOC For gas-cap reservoirs with significant composition and temperature variations with depth, the compositional variations with depth could result in changes in the type of hydrocarbon systems ranging from retrograde gas at the top of the hydrocarbon column to hydrocarbon liquid with variable volatility. In some cases, the transition from the dew point gas to the bubble point oil occurs without an existing gas cap. A possible explanation for this occurrence, as illustrated schematically in Fig. 1.46, is attributed to the possibility that the temperature gradient coincided with the hydrocarbon critical temperature in the transition between gas and liquid system to form a critical mixture. This critical mixture separated the gas and liquid without forming a distinct GOC; labeled as undersaturated GOC PHASE RULE It is appropriate at this stage to introduce and define the concept of the phase rule. Gibbs (1948 [1876]) derived a simple relationship between the number of phases, P, in equilibrium, the number of components, C, and the number Pressure & temperature Gas cap Critical region Depth Oil rim rvoir Rese n bo s ar int c o o dr al p Hy itic cr re eratu temp n FIGURE 1.46 Undersaturated gas-oil contact. 66 CHAPTER 1 Fundamentals of Hydrocarbon Phase Behavior of independent variables, F, that must be specified to describe the state of the system completely. Gibbs proposed the following fundamental statement of the phase rule: F¼CP+2 (1.19) where F ¼ number of variables required to determine the state of the system at equilibrium or number of degrees of freedom (such as pressure, temperature, density) C ¼ number of independent components P ¼ number of phases A phase was defined as a homogeneous system of uniform physical and chemical compositions. The degrees of freedom, F, for a system include the intensive properties such as temperature, pressure, density, and composition (concentration) of phases. These independent variables must be specified to define the system completely. In a single component (C ¼ 1), two-phase system (P ¼ 2), there is only one degree of freedom (F ¼ 1 2 + 2 ¼ 1); therefore, only pressure or temperature needs to be specified to determine the thermodynamic state of the system. The phase rule as described by Eq. (1.19) is useful in several ways. It indicates the maximum number of equilibrium phases that can coexist and the number of components present. It should be pointed out that the phase rule does not determine the nature, exact composition, or total quantity of the phases. Furthermore, it applies only to a system in stable equilibrium and does not determine the rate at which this equilibrium is attained. The importance and the practical application of the phase rule are illustrated through the following examples. EXAMPLE 1.9 For a single-component system, determine the number of degrees of freedom required for the system to exist in the single-phase region. Solution Applying Eq. (1.19) gives F ¼ 1 1 + 2 ¼ 2. Two degrees of freedom must be specified for the system to exist in the single-phase region. These must be the pressure p and the temperature T. Problems 67 EXAMPLE 1.10 What degrees of freedom are allowed for a two-component system in two phases? Solution Because C ¼ 2 and P ¼ 2, applying Eq. (1.19) yields F ¼ 2 2 + 2 ¼ 2. The two degrees of freedom could be the system pressure p and the system temperature T, and the concentration (mole fraction), or some other combination of T, p, and composition. EXAMPLE 1.11 For a three-component system, determine the number of degrees of freedom that must be specified for the system to exist in the one-phase region. Solution Using the phase rule expression gives F ¼ 3 1 + 2 ¼ 4, four independent variables must be specified to fix the system. The variables could be the pressure, temperature, and mole fractions of two of the three components. From the foregoing discussion it can be observed that hydrocarbon mixtures may exist in either the gas or liquid state, depending on the reservoir and operating conditions to which they are subjected. The qualitative concepts presented may be of aid in developing quantitative analyses. Empirical equations of state are commonly used as a quantitative tool in describing and classifying the hydrocarbon system. These equations of state require n n Detailed compositional analysis of the hydrocarbon system. Complete description of the physical and critical properties of the mixture of the individual components. PROBLEMS 1. The following is a list of the compositional analysis of different hydrocarbon systems, with the compositions expressed in terms of mol%. Classify each of these systems. 68 CHAPTER 1 Fundamentals of Hydrocarbon Phase Behavior Component System 1 System 2 System 3 System 4 C1 C2 C3 C4 C5 C6 C7+ 68.00 9.68 5.34 3.48 1.78 1.73 9.99 25.07 11.67 9.36 6.00 3.98 3.26 40.66 60.00 8.15 4.85 3.12 1.41 2.47 20.00 12.15 3.10 2.51 2.61 2.78 4.85 72.00 2. A pure component has the following vapor pressure: T, °F pv , psi 104 46:09 140 135:04 176 345:19 212 773:75 a. Plot these data to obtain a nearly straight line. b. Determine the boiling point at 200 psi. c. Determine the vapor pressure at 250°F. 3. The critical temperature of a pure component is 260°F. The densities of the liquid and vapor phase at different temperatures are T, °F ρL , lb=ft3 ρv , lb=ft3 86 40:280 0:886 122 38:160 1:691 158 35:790 2:402 212 30:890 5:054 Determine the critical density of the substance. 4. Using the Lee and Kesler vapor correlation, calculate the vapor pressure of i-butane at 100°F. Compare the calculated vapor pressure with that obtained from the Cox charts. 5. Calculate the saturated liquid density of n-butane at 200°F using a. The Rackett correlation. b. The modified Rackett correlation. Compare the calculated two values with the experimental value determined from Fig. 1.5. 6. What is the maximum number of phases that can be in equilibrium at constant temperature and pressure in one-, two-, and threecomponent systems? 7. For a seven-component system, determine the number of degrees of freedom that must be specified for the system to exist in the following regions: a. One-phase region. b. Two-phase region. References 69 8. Fig. 1.9 shows the phase diagrams of eight mixtures of methane and ethane along with the vapor pressure curves of the two components. Determine a. Vapor pressure of methane at 160°F. b. Vapor pressure of ethane at 60°F. c. Critical pressure and temperature of mixture 7. d. Cricondenbar and cricondentherm of mixture 7. e. Upper and lower dew point pressures of mixture 6 at 20°F. f. The bubble point and dew point pressures of mixture 8 at 60°F. 9. Using Fig. 1.9, prepare and identify the different phase regions of the pressure-composition diagram for the following temperatures: a. 120°F. b. 20°F. 10. Using Fig. 1.9, prepare the temperature-composition diagram (commonly called the T/X diagram) at the following pressures: a. 300 psia. b. 700 psia. c. 800 psia. 11. Derive equations for the amounts of liquid and vapor present in the two-phase region when the composition is expressed in terms of a. Weight fraction of the more volatile component. b. Weight fraction of the less volatile component. 12. A 20 ft3 capacity tank is evacuated and thermostated at 60°F. Five pounds of liquid propane is injected. What is the pressure in the tank and the proportions of liquid and vapor present? Repeat calculations for 100 lb of propane injected. REFERENCES Burcik, E., 1957. Properties of Petroleum Reservoir Fluids. International Human Resources Development Corporation (IHRDC), Boston, MA. Gibbs, J.W., 1948. The Collected Works of J. Willard Gibbs (Connecticut Academy of Arts and Sciences, Trans.). vol. 1. Yale University Press, New Haven, CT (original text published 1876). Lee, B.I., Kesler, M.G., 1975. A generalized thermodynamics correlation based on threeparameter corresponding states. AIChE J. 21 (3), 510–527. Pitzer, K.S., 1955. The volumetric and thermodynamics properties of fluids. J. Am. Chem. Soc. 77 (13), 3427–3433. Rackett, H.G., 1970. Equation of state for saturated liquids. J. Chem. Eng. Data 15 (4), 514–517. Spencer, F.F., Danner, R.P., 1973. Prediction of bubblepoint density of mixtures. J. Chem. Eng. Data 18 (2), 230–234. Yamada, T., Gunn, R., 1973. Saturated liquid molar volumes: the Rackett equation. J. Chem. Eng. Data 18 (2), 234–236. Chapter 2 Characterizing Hydrocarbon-Plus Fractions Reservoir fluids contain a variety of substances of different chemical structure that include hydrocarbon and nonhydrocarbon components. Because of the enormous number components that constitute the hydrocarbon system, a complete chemical composition as well as identification of various components that constitute the hydrocarbon fluid are largely speculative. The constituents of a natural reservoir fluid form hydrocarbon spectrum from the lightest one; eg, nitrogen or methane, through intermediate to very large molecular weight fractions. The relative proportion of these different parts of the fluid can vary over a wide range which results in petroleum fluids showing very different features: eg, dry gas, retrograde, volatile oil, and others. CRUDE OIL ASSAY Each crude oil has unique molecular and chemical characteristics. These variances translate into crucial differences in crude oil quality. A crude oil assay is essentially the evaluation of a crude oil by physical and chemical testing. It provides data that gives information on the suitability of the crude for its end use and also helps predict the commercial value of the crude. In addition, Information obtained from the petroleum assay is used for detailed equations of state tuning and validation. Crude oil bulk properties are useful for initial screening and tentative identification of genetically related oils. These bulk properties include several laboratory measurements that can be used to quantify the quality of the crude oil. A brief discussion of these properties is given below: n The American Petroleum Institute (API) gravity: The API gravity is a standard industry property that is designed to quantify the quality of the crude oil. This oil property is defined in a manner, as given below, that suggests a higher value indicates a light oil and a lower value indicates a heavy oil. API ¼ 141:5 131:5 γ where γ is the liquid specific gravity. Equations of State and PVT Analysis. http://dx.doi.org/10.1016/B978-0-12-801570-4.00002-7 Copyright # 2016 Elsevier Inc. All rights reserved. 71 72 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions n n n n n n It should be noted that the API gravity of water is 10 because the water specific gravity is 1, which suggests that the API gravity for very heavy crudes could be less than 10 and varies between 10 and 30 for heavy crudes, to between 30 and 40 for medium crudes, and to above 40 for light crudes. The cloud point: The cloud point is defined as the lowest temperature at which wax crystals begin to form by a gradual cooling under standard conditions. At this temperature the oil becomes cloudy and the first particles of wax crystals are observed. Low cloud point products are desirable under low-temperature conditions. Cloud points are measured for oils that contain paraffins in the form of wax and therefore for light fractions (naphtha or gasoline) no cloud point data are reported. Wax crystals can cause severe flow assurance problems including plugging flow lines and reduction in well flow potential. The Reid vapor pressure (RVP): The RVP is the measure of vapor pressure of lightweight hydrocarbons present above the crude oil surface, the higher the value of RVP, the higher the percentage of light hydrocarbons in the crude oil samples. The pour point: The pour point of a petroleum fraction is defined as the lowest temperature at which the oil can be stored and still be capable of flowing through a pipeline without stirring. The sulfur content: The sulfur content in natural gas is usually found in the form of hydrogen sulfide (H2S). Some natural gas contains H2S as high as 30% by volume. The amount of sulfur in a crude is expressed as a percentage of sulfur by weight, and varies from 0.1% to 6%. Crude oils with less than 1 wt.% sulfur are called low-sulfur or sweet crude oil, and those with more than 1 wt.% sulfur are called high-sulfur or sour crude oil. The flash point: The flash point of a liquid hydrocarbon or any of its derivatives is defined as the lowest temperature at which sufficient vapor is produced above the liquid to form a mixture with air that can produce a spontaneous ignition if a spark is present. Flash point is an important parameter for safety considerations, especially during storage and transportation of volatile petroleum products (ie, LPG, light naphtha, gasoline). The surrounding temperature around a storage tank should always be less than the flash point of the fuel to avoid the possibility of ignition. The refractive index (RI): The RI represents the ratio of the velocity of light in a vacuum to that in the oil. The property is considered as the degree to which light bends (refraction) when passing through a medium. Values of RI can be measured very accurately and are used to correlate density and other properties of hydrocarbons with high reliability. Information obtained from RI measurements can be used to Crude Oil Assay 73 n n n n n n n determine the behavior of asphaltenes; specifically, their tendency to precipitate from crude oil. The freeze point: The freeze point is defined as the temperature at which the hydrocarbon liquid solidifies at atmospheric pressure. The smoke point: The smoke points is a property that is designed to describe the tendency of a fuel to burn with a smoky flame. A higher amount of aromatics in a fuel causes a smoky characteristic for the flame and energy loss due to thermal radiation. This property refers to the height of a smokeless flame of fuel as expressed in millimeters beyond which smoking takes place; ie, a high smoke point indicates a fuel with low amount of aromatics. The aniline point: This property represents the minimum temperature for complete miscibility of equal volumes of aniline and petroleum oil. The aniline point is important in characterization of petroleum fractions and analysis of molecular type. The Conradson carbon residue (CCR): This property measures the tendency and ability of the crude oil to form coke. It is determined by destructive distillation of a sample to the remaining coke residue. The CCR property is expressed as the weight of the remaining coke residue as a percentage of the original sample. The acid number: The acid number is a property that determines the organic acidity of a refinery stream. The gross heat of combustion (high heating value): This property is a measure of the amount of heat produced by the complete burning of a unit quantity of fuel. The net heat of combustion (lower heating value): This is obtained by subtracting the latent heat of vaporization of the water vapor formed by the combustion from the gross heat of combustion or higher heating value. There are different ways to classify the components of the reservoir fluids. Normally, the constituents of a hydrocarbon system are classified into the following two categories: n n Well-defined petroleum fractions Undefined petroleum fractions Well-Defined Components Pure fractions are considered well-defined components because their physical properties were measured and compiled over the years. These properties include specific gravity, normal boiling point, molecular weight, critical properties, and acentric factor. Well-defined fractions are grouped in the following three sets: 74 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions n n n Nonhydrocarbon fractions that include CO2, N2, H2S, etc. Methane C1 through normal pentane n-C5. Hydrocarbon components that include hexanes, heptanes, and other heavier components, (ie, C6, C7, C8, C9, etc.) where the number of isomers rises exponentially. For example, Hexanes “C6” refers to the hexane petroleum hydrocarbon distillation fraction that contains a high proportion of n-hexane in addition to 2-methylpentane, 3-methylpentane, and smaller amounts of other components. Similarly, the Heptanes refer to the heptane petroleum hydrocarbon fraction which contains isomers of the empirical formula C7H16 with the most prevalent components being n-heptane, dimethylcyclopentanes, 3-ethylpentane, methylcyclohexane, and 3-methylhexane. However, the identification and characterizing of each one of these components for us in EOS applications can be a difficult task. Therefore, these components are expressed as a group that consists of a number of components that have boiling points within a certain range. These groups are commonly denoted as single carbon number (SCN) and called pseudofractions because they represent a group of components; eg, C6 group, C7 group, and so forth. The tabulated laboratory data shown in Table 2.1 illustrates both the concept of SCN and pseudofractions. Katz and Firoozabadi (1978) presented a generalized set of physical properties for the hydrocarbon groups C6 through C45 that are expressed as a SCN, such as the C6-group, C7-group, C8-group, and so on. These generalized properties, as shown in Table 2.2, were generated by analyzing the physical properties of 26 condensates and crude oil systems. The tabulated properties of these groups include the average boiling point, specific gravity, and molecular weight as well as critical properties. Ahmed (1985) conveniently correlated Katz and Firoozabadi’s tabulated physical properties with SCN as represented by the number of carbon atoms “n” by using a regression model. The generalized equation has the following form: θ ¼ a1 + a2 n + a3 n2 + a4 n3 + a5 n where θ ¼ any physical property, such as Tc, pc, or Vc n ¼ effective number of carbon atoms of SCN group, eg, 6, 7, …, 45 a1–a5 ¼ coefficients of the equation as tabulated in Table 2.3 Crude Oil Assay 75 Table 2.1 The Concept of Single Carbon Number Wellstream Single carbon number group “SCN” C6 group C7 group C8 group C9 group C10 group * Component Mole % GPM* Hydrogen sulfide Nitrogen Carbon dioxide Methane Ethane Propane Iso-butane N-Butane 2,2-Dimethylpropane Iso-pentane N-Pentane 2,2-Dimethylbutane Cyclopentane 2,3-Dimethylbutane 2-Methylpentane 3-Methylpentane Other hexanes n-Hexane Methylcyclopentane Benzene Cyclohexane 2-Methylhexane 3-Methylhexane 2,2,4-Trimethylpentane Other heptanes n-Heptane Methylcyclohexane Toluene Other C-8s n-Octane Ethylbenzene m- and p-Xylene o-Xylene Other C-9s n-Nonane Other C-10s n-Decane Undecanes plus Total 0.002 0.084 1.163 65.410 12.107 5.329 1.280 2.039 0.012 0.899 0.902 0.031 0.006 0.060 0.370 0.221 0.000 0.560 0.098 0.066 0.144 0.235 0.211 0.000 0.213 0.471 0.329 0.286 0.815 0.388 0.088 0.354 0.113 0.547 0.299 0.742 0.247 3.880 100.000 0.000 0.000 0.000 0.000 3.220 1.459 0.416 0.639 0.004 0.327 0.325 0.013 0.002 0.025 0.153 0.090 0.000 0.229 0.035 0.018 0.049 0.109 0.096 0.000 0.092 0.216 0.132 0.095 0.380 0.198 0.034 0.136 0.043 0.285 0.167 0.425 0.151 3.131 12.692 Gallon-per-thousand cubic feet 76 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions Table 2.2 Generalized Physical Properties Group SCN Tb (°R) γ Kw M Tc (°R) Pc (psia) ω Vc (ft3/lb) C6 C7 C8 C9 C10 C11 C12 C13 C14 C15 C16 C17 C18 C19 C20 C21 C22 C23 C24 C25 C26 C27 C28 C29 C30 C31 C30 C33 C34 C35 C36 C37 C38 C39 C40 C41 C42 C43 C44 607 658 702 748 791 829 867 901 936 971 1002 1032 1055 1077 1101 1124 1146 1167 1187 1207 1226 1244 1262 1277 1294 1310 1326 1341 1355 1368 1382 1394 1407 1419 1432 1442 1453 1464 1477 0.69 0.727 0.749 0.768 0.782 0.793 0.804 0.815 0.826 0.836 0.843 0.851 0.856 0.861 0.866 0.871 0.876 0.881 0.885 0.888 0.892 0.896 0.899 0.902 0.905 0.909 0.912 0.915 0.917 0.92 0.922 0.925 0.927 0.929 0.931 0.933 0.934 0.936 0.938 12.27 11.96 11.87 11.82 11.83 11.85 11.86 11.85 11.84 11.84 11.87 11.87 11.89 11.91 11.92 11.94 11.95 11.95 11.96 11.99 12.00 12.00 12.02 12.03 12.04 12.04 12.05 12.05 12.07 12.07 12.08 12.08 12.09 12.10 12.11 12.11 12.13 12.13 12.14 84 96 107 121 134 147 161 175 190 206 222 237 251 263 275 291 300 312 324 337 349 360 372 382 394 404 415 426 437 445 456 464 475 484 495 502 512 521 531 914 976 1027 1077 1120 1158 1195 1228 1261 1294 1321 1349 1369 1388 1408 1428 1447 1466 1482 1498 1515 1531 1545 1559 1571 1584 1596 1608 1618 1630 1640 1650 1661 1671 1681 1690 1697 1706 1716 476 457 428 397 367 341 318 301 284 268 253 240 230 221 212 203 195 188 182 175 168 163 157 152 149 145 141 138 135 131 128 126 122 119 116 114 112 109 107 0.271 0.310 0.349 0.392 0.437 0.479 0.523 0.561 0.601 0.644 0.684 0.723 0.754 0.784 0.816 0.849 0.879 0.909 0.936 0.965 0.992 1.019 1.044 1.065 1.084 1.104 1.122 1.141 1.157 1.175 1.192 1.207 1.226 1.242 1.258 1.272 1.287 1.300 1.316 5.6 6.2 6.9 7.7 8.6 9.4 10.2 10.9 11.7 12.5 13.3 14 14.6 15.2 15.9 16.5 17.1 17.7 18.3 18.9 19.5 20.1 20.7 21.3 21.7 22.2 22.7 23.1 23.5 24 24.5 24.9 25.4 25.8 26.3 26.7 27.1 27.5 27.9 Source: Permission to publish from the Society of Petroleum Engineers of the AIME. # SPE-AIME. Group 1. TBP 77 Table 2.3 Coefficients of the Equation θ M Tc (°R) pc (psia) Tb (°R) ω γ Vc (ft3/ lbm-mol) a1 a2 a3 a4 a5 131.11375 24.96156000 0.34079022 2.49411840E-03 468.32575000 926.602244514 39.729362915 0.722461850 0.005519083 1366.431748654 311.2361908 14.6869301 0.3287671 0.0027346 1690.9001135 427.2959078 50.08577848 0.88693418 6.75667E-03 551.2778516 0.31428163 7.80028E-02 1.39205E-03 1.02147E-05 0.991028867 0.86714949 3.4143E-03 2.8396E-05 2.4943E-08 1.1627984 0.232837085 0.974111699 0.009226997 3.63611E-05 0.111351508 Undefined Components The undefined petroleum fractions are those heavy components lumped together and identified as the plus fraction; for example, the C7+ fraction. Nearly all naturally occurring hydrocarbon systems contain a quantity of heavy fractions that are not well defined and not mixtures of discretely identified components. A proper description of the physical properties of the plus fractions and other undefined petroleum fractions in hydrocarbon mixtures is essential in performing reliable phase-behavior calculations and compositional modeling studies. The proper description of these physical properties require the use of thermodynamic property prediction models, such as an equation of state, which in turn requires that one must be able to provide the acentric factor, critical temperature, and critical pressure for both the defined and undefined (heavy) fractions in the mixture. The problem of how to adequately characterize these undefined plus fractions in terms of their critical properties and acentric factors has been long recognized in the petroleum industry. Typically, undefined components are identified by a true boiling point (TBP) distillation analysis and characterized by an average normal boiling point, specific gravity, and molecular weight. In general, the data available that can be used to characterize the plus fractions are divided into the following three distinct groups: n n n TBP Stimulated distillation by gas chromatography (GC) The plus fraction characterization methods A brief discussion of the three groups is given next. GROUP 1. TBP Boiling point of a pure petroleum fraction is the temperature at which it starts boiling and transfers from a liquid state to a vapor state at a pressure of 1 atm. In other words, it is the temperature at which the vapor pressure is 78 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions equal to atmospheric pressure for the petroleum fraction. There are various standard distillation methods through which the boiling point curve can be obtained. In the TBP distillation test, the oil is distilled and the temperature of the condensing vapor as well as the volume of liquid formed are recorded. The TBP test data are used to construct a distillation curve of liquid volume percent distilled versus condensing temperature. In this distillation process, the Heptanes-plus fraction “C7+” is subjected to a standardized analytical distillation, first at atmospheric pressure, then in a vacuum at a pressure of 40 mmHg. Usually the temperature is taken when the first droplet distills over. Ten fractions (cuts) are then distilled off, the first one at 50°C and each successive one with a boiling range of 25°C. For each distillation cut of the C7+, the volume, specific gravity, and molecular weight are determined. Cuts obtained in this manner are identified by the boiling point ranges in which they were collected. The condensing temperature of the vapor at any point in the test will be close to the boiling of the material condensing at that point. A key result from this distillation test is the boiling point curve that is defined as the boiling point of the petroleum fraction versus the percentage volume distilled. The initial boiling point (IBP) is defined as the temperature at which the first drop of liquid leaves the condenser tube of the distillation apparatus with the final boiling point (FBP) defined as the highest temperature recorded in the test. The difference between FBP and IBP is called boiling point range. It should be pointed out that oil fractions tend to crack at a temperature of approximately 650°F and one atmosphere. As a result, when approaching this cracking temperature, the pressure is gradually reduced to as low as 40 mmHG to avoid cracking the sample and distorting TBP measurements. A typical TBP curve is shown in Fig. 2.1. It should be noted that each distilled cut includes a large number of components with close boiling points. Based on the data from the TBP distillation, the C7+ fraction is divided into pseudocomponents or hypothetical components suitable for equations of state modeling application. These fractions are commonly collected within the temperature range of two consecutive distillation volumes, as shown in Fig. 2.1. The distilled cut is referred to as pseudocomponent. Each pseudocomponent, eg, pseudocomponent 1, is represented by a volumetric portion of the TBP curve with an average normal boiling point “Tbi.” The average boiling point of each cut is taken at the mid-volume percent of the cut; that is, between Vi and Vi1. This is given by the integration ð Vi Tb ðV ÞdV ðTb Þi + ðTb Þi1 Tb1 ¼ Vi1 2 Vi Vi1 Group 1. TBP 79 Pseudocomponent Tbi from TBP curve Temperature Vi Tb1 = ∫V T dv (T ) + (T ) i−1 Vi − Vi−1 ≈ b i b i−1 2 (Tb)i (Tb)i–1 Pseudocomponent 1 Vi–1 Vi Volume distilled n FIGURE 2.1 Typical distillation curve. where Vi Vi1 ¼ volumetric cut of the TBP curve associated with pseudocomponent i Tbt(V) ¼ TBP temperature at liquid volume percent distilled “V” The pseudocomponents from the TBP test are identified primarily by the following three laboratory measured physical properties: n n n The boiling point “Tb.” The molecular weight “M,” is measured by the freezing point depression experiment. In this test, the molecular weight of the distillation cut is measured by comparing its freezing point with the freezing point of the pure solvent with a known molecular weight. The fraction specific gravity “γ i” of each distillation cut, is measured by using a pycnometer or electronic densitometer. The mass (mi) of each distillation cut is measured directly during the TBP analysis. The cut is quantified in terms of number of moles ni, which is given as the ratio of the measured mass (mi) to the molecular weight (Mi); ie, ni ¼ mi Mi The density (ρi) and specific gravity (γ i) is then calculated from the measured mass and volume of the distillation cut as ρi ¼ mi Vi 80 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions Or in terms of the specific gravity as γi ¼ ρi 62:4 The average molecular weight and density of the C7+ from the TBP experiment with N distillation cuts are given by XN MC7 + ¼ Xi¼1 N mi n i¼1 i and XN ρC7 + ¼ Xi¼1 N mi V i¼1 i The calculated molecular weight and density from the distillation test should be compared with those of the measured C7+ values, any discrepancies are attributed and reported as lost materials resulting from the distillation test. Fig. 2.2 shows a typical graphical presentation of the molecular weight, specific gravity, and the TBP as a function of the weight percent of liquid vaporized. The molecular weight, M, specific gravity, γ, and boiling point temperature, Tb, are considered the key properties that reflect the chemical makeup Tb, M Sp.gr 1 1600 0.9 1400 0.8 1200 Tb 800 M 600 0.6 0.5 0.4 0.3 400 0.2 200 0 0.1 0 20 40 60 80 Mass % n FIGURE 2.2 TBP, specific gravity, and molecular weight of a stock-tank oil sample. 0 100 Specific gravity 0.7 1000 Group 1. TBP 81 of petroleum fractions. Watson et al. (1935) introduced a widely used characterization factor, commonly known as the universal oil products or Watson characterization factor, based on the normal boiling point and specific gravity. This characterization parameter is given by the following expression: 1=3 Kw ¼ Tb γ (2.1) where Kw ¼ Watson characterization factor Tb ¼ normal boiling point temperature, °R γ ¼ specific gravity The average heptanes-plus characterization factor “KC7 + ” can be estimated from the TBP distillation data by applying the following relationship: XN h 1=3 i Tbi =γ i mi i¼1 K C7 + ¼ XN m i¼1 i where mi is the mass of each individual cut. The Watson characterization factor “Kw” provides a qualitative measure of the paraffinicity of a crude oil. In general, Kw varies roughly from 8.5 to 13.5 as follows: n n n For paraffinic hydrocarbon systems, Kw ranges from 12.5 to 13.5. For naphthenic hydrocarbon systems, Kw ranges from 11.0 to 12.5. For aromatic hydrocarbon systems, Kw ranges from 8.5 to 11.0. The significance of the Watson characterization property is based on the observation that over reasonable ranges of boiling point, Kw is almost constant for C7+ distillation cuts. Whitson (1980) suggests that the Watson factor can be correlated with the molecular weight M and specific gravity γ as given by the following expression: 0:15178 M Kw 4:5579 0:84573 γ (2.2) Whitson pointed out that Eq. (2.2) becomes less reliable when the molecular weight “M” of the distillation cut exceeds 250. However; the limitation on the molecular weight can be removed by expressing the above relationship in the following modified form: 0:149405 M Kw 3:33475 0:935227 + 3:099934 γ (2.2A) 82 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions If the molecular weight and specific gravity of the C7+ are available data, the Watson characterization factor is estimated from the above expression as 0:149405 M ðKw ÞC7 + 3:33475 0:935227 + 3:099934 γ C7 + Whitson and Brule (2000) observed that a hydrocarbon mixture with multiple hydrocarbon samples, the ðKw ÞC7 + as correlated with MC7 + and γ C7 + and expressed by Eq. (2.2) often exhibits a constant value for all PVT samples and distillation cuts. The authors suggest that a plot of molecular weight versus specific gravity of the plus fractions is useful for checking the consistency of C7+ molecular weight and specific gravity measurements. Austad (1983) and Whitson and Brule (2000) illustrated this observation by plotting M versus γ for each C7+ fractions from data on two North Sea fields. Whitson and Brule’s observations are shown schematically in Figs. 2.3 and 2.4. The schematic plot of MC7 + versus γ C7 + of the volatile oil shown in Fig. 2.3 has an average ðKw ÞC7 + ¼ 11:90 0:01 for a range of molecular weights from 220 to 255. Data for the gas condensate, as shown schematically in Fig. 2.4, indicate an average ðKw ÞC7 + ¼ 11:99 0:01 for a range of molecular weights from 135 to 150. Whitson and Brule concluded n FIGURE 2.3 Volatile oil specific gravity versus molecular weight of C7+. Group 1. TBP 83 n FIGURE 2.4 Condensate-gas specific gravity versus molecular weight of C7+. that the high degree of correlation for these two fields suggests accurate molecular weight measurements by the laboratory. In general, the spread in ðKw ÞC7 + values exceeds 0.01 when measurements are performed by a commercial laboratory. Whitson and Brule stated that, when the characterization factor for a field can be determined, Eq. (2.8) is useful for checking the consistency of C7+ molecular weight and specific gravity measurements. Significant deviation in ðKw ÞC7 + , such as 0.03 could indicate possible error in the measured data. Because molecular weight is more prone to error than determination of specific gravity, an anomalous ðKw ÞC7 + usually indicates an erroneous molecular weight measurement. The implication of assuming a constant Kw, eg, ðKw ÞC7 + , is that it can be used to estimate “Mi” or “γ i” of a distillation cut if they are not measured. Assuming a constant Kw for each fraction, Eq. (2.2A) is rearranged to solve for the molecular weight “Mi” and specific gravity “γ i”, to give Mi ¼ 315:5104753 106 ðKw 3:099934Þ6:6932358 γ 6:2596977 i or γi ¼ 3:624849522 Mi0:159752123 ðKw 3:099934Þ1:0692586 (2.3) 84 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions It should be noted that the molecular weight “M,” specific gravity “γ,” and the Watson characterization factor “Kw” of pseudocomponents provide the key data needed for further characterization of these fractions in terms of their critical properties and acentric factors. The following simplified two sets of correlations for estimating Tc, pc, and ω of pseudocomponents were generated by matching data from several literature published data. The first set was correlated as a function of the ratio “M/γ” while the second set was correlated as a function of “Tb/γ,” to give Set 1. Properties as a function of the ratio “M/γ”: 0:351579776 M M Tc ¼ 231:9733906 0:352767639 233:3891996 (2.3A) γ γ 0:885326626 M M 0:106467189 + 49:62573013 pc ¼ 31829 γ γ M M ω ¼ 0:279354619 ln + 0:00141973 1:243207091 γ γ vc ¼ 0:054703719 (2.3B) (2.3C) 0:984070268 M M + 87:7041106 0:001007476 (2.3D) γ γ 0:664115311 M M 379:4428973 1:40066452 Tb ¼ 33:72211 ðM=γ Þ γ γ (2.3E) Set 2. Properties as a function of the ratio “Tb/γ”: Tc ¼ 1:379242274 0:967105064 Tb Tb + 0:010009512 (2.3F) + 0:012011487 γ γ Tb Tb 0:00867297 + 0:099138742 pc ¼ 2898:631565exp 0:002020245 γ γ (2.3G) 1:250922701 Tb Tb ω ¼ 0:000115057 0:92098982 (2.3H) + 0:00068975 γ γ 3 2 Tb vc ¼ 8749:94 106 Tγb 9:75748 106 Tγb + 0:014653 γ 595:7922076 Tb γ 0:001617123Tb 132:1481762 M ¼ 52:11038728 exp γ 4:60231684 (2.3I) (2.3J) Group 1. TBP 85 If the Tb is not known, the molecular weight can be estimated from the following expression: M ¼ 0:048923 exp ð9:88378γ Þ 33:085468γ + 39:598437 γ (2.3K) where Tb ¼ boiling point in °R Tc ¼ critical temperature in °R vc ¼ critical volume, ft3/lbm-mol M ¼ molecular weight Notice, to express critical volume in ft3/lbm, divide vc by the molecular weight “M.” Vc ¼ vc M The main issue that should be addressed after obtaining the distillation curve is how to break down or cut the entire boiling range into a number of cut-point ranges, which are used to define pseudocomponent. The determination of the number of cuts is essentially arbitrary; however, the number of pseudocomponents has to be sufficient to reproduce the volatility properties in fractions with wide boiling point ranges. The following provide a brief summary of industry experience and guidelines: 1. Pedersen et al. (1982) suggest selecting a small number of pseudocomponents with approximately the same weight fraction. 2. As a rule of thumb, use between 10 and 15 fractions if the TBP curve is relatively “steep” and use 5–8 fractions if the curve is relatively “flat.” 3. To better reproduce the shape of the TBP curve and to reflect the relative distribution of the light, middle, and heavy ends in pseudocomponent, Miquel et al. (1992) pointed out that the pseudocomponent breakdown can be made from either equal volumetric fractions or regular temperature intervals in the TBP curve. To reproduce the shape of the TBP curve better and reflect the relative distribution of light, middle, and heavy components in the fraction, a breakdown based on equal TBP temperature intervals is recommended. On the other hand, when equal volumetric fractions are taken, information concerning the light initial component and heavy end component could be lost. 4. Most commercial process simulators (eg, the HYSYS model) generate pseudocomponents based on boiling point ranges 86 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions representing the oil fractions. Typical boiling point ranges for pseudocomponents in a process simulator are shown below: Boiling Point Range IBP to 800°F 800–1200°F 1200–1650°F Total no. of pseudocomponents Suggested Number of Pseudocomponents 30 10 8 48 It should be pointed out that a complete TBP curve is rarely available for an entire range of percent distilled. In many cases, the data is obtained up to the 50% or 70% distilled point for any petroleum fractions or crude oils. This incomplete distillation curve is attributed to the fact that the petroleum fractions or crude oils contain more heavy hydrocarbons toward the end of the distillation curve; therefore it is more difficult is to achieve accurate boiling points for that fraction or crude oil. However, it is important for accurate fluid characterizations and in the process of equations of state tuning that the distillation curve should be extended to 90–95% distilled point. As conceptually shown in Fig. 2.5, Riazi and Daubert (1987) developed a (1/b) Temperature Tbi = To 1+ a 1 ln 1–Vi b Tbi To ≈50–70% Volume distilled “Vi” n FIGURE 2.5 Mathematical representation of TBP curve. Group 2. Stimulated Distillation by GC 87 mathematical representation of the TBP curve that can be used to complete the TBP curves up to the 95% point. The mathematical representation is given by the following equation: ( Tbi ¼ To 1 + a 1 ln b 1 Vi ð1=bÞ ) where, Tbi ¼ TBP temperature and any distilled volume Vi Vi ¼ distilled volume fraction To ¼ IBP (ie, at Vi ¼ 0) a, b ¼ correlating coefficients The coefficients a and b are the two parameters that can be determined by regression or alternatively by linearizing the above equation as follows: ln Tbi To 1 a 1 1 + ¼ ln ln ln To b b b 1 Vi The above relationship indicates that a plot of ln ½Tbi To =To versus ln ½1=1 Vi would produce a straight line with a slope of (1/b) and intercept of [(1/b)ln(a/b)] and used to calculate the parameters a and b. The mathematical expression can then be used to predict and complete the TBP curve up to 95% distilled volume. GROUP 2. STIMULATED DISTILLATION BY GC Normal TBP distillation involves a long procedure and is costly. GC has been used widely as a fast and reproducible method for simulated distillation (SD) to analyze the carbon number distribution of various hydrocarbon components in the crude oil. GC is a common type of chromatography used for separating and analyzing compounds that can be vaporized without decomposition. Like all other chromatographic techniques, GC consists of a mobile and a stationary phase. The mobile phase is carrier gas comprised of an inert gas (ie, helium, argon, or nitrogen). The stationary phase consists of a packed capillary column with the packing acting as a stationary phase that is capable of absorbing and retaining hydrocarbon components for a specific time before releasing them sequentially based on their boiling points. This causes each compound to liberate at a different time, known as the retention time of the compound. GC is then similar to fractional distillation, as both processes separate the components of a mixture primarily based on boiling point. 88 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions The specific process of obtaining the SD curve from a GC experiment is based on injecting a small oil sample into a stream of carrier gas, which displaces that sample into a column packed with materials that absorbs the heavier components of the oil. The column is placed in a temperature programmable oven; as the temperature of the column is raised, the heavier components are released sequentially based on their boiling point and consequently removed by the carrier gas into a detector that senses changes in the thermal conductivity of the carrier gas. For a given program of heating, the retention time in the column is proportional to the normal boiling point of the released components. These are correlated with the retention times of a calibration curve obtained by conducting the separation under the same conditions of a known mixture of hydrocarbons, usually n-alkanes, covering the boiling range expected in the sample. The resulting calibration curve relates carbon number and boiling point to retention time and allows for the determination of boiling points for the components in the hydrocarbon sample. As shown in Fig. 2.6 and Table 2.4, all components detected by the GC between two neighboring n-alkanes are commonly grouped together and n FIGURE 2.6 GC data calibration sample of n-alkanes. Group 2. Stimulated Distillation by GC 89 Table 2.4 Boiling Point Data n-Alkane Retention Time (min) Boiling Point (°C) C5 C6 C7 C8 C9 C10 C11 C12 C14 C15 C16 C17 C18 C20 C24 C28 C32 C36 C40 0.09764 0.11994 0.16272 0.23050 0.31882 0.41783 0.51717 0.61712 0.79714 0.87985 0.95957 1.03281 1.10226 1.23156 1.45693 1.64895 1.81618 1.96389 2.10297 21.9 54.7 84.1 111.9 136.3 159.7 181.9 201.9 220.8 239.7 256.9 273.0 288.0 301.9 376.9 416.9 451.9 481.9 508.0 reported as a pseudocomponent with a SCN equal to the higher n-alkane; eg, between n-alkanes of C8 and C9 is designated as pseudocomponent with SCN of C9. As shown in Fig. 2.7, the distillation curve resulting from the plot of GC data in terms of boiling point as a function of the retention time or SCN of n-alkane is called a SD curve and represents boiling points of compounds in a petroleum mixture at atmospheric pressure. It should be pointed out that SD curves are very close to actual boiling points shown by TBP curves. The main advantage of this SD process is that it requires smaller samples and also is less expensive than TBP distillation. The GC is used for gas and oil samples generating distillation curves as well as the composition of hydrocarbon mixtures in terms of their weight fractions “wi”. Conversion from weight fraction “wi” to mole fraction “xi” requires molecular weights “Mi” of all components and can be achieved from wi =Mi xi ¼ X ðwi =Mi Þ i 90 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions Stimulated distillation curve 600 Boiling point (°C) 500 400 300 200 100 0 0 0.5 1 1.5 2 2.5 Retention time (min) n FIGURE 2.7 Stimulated distillation curve. Similarly, if the volume fraction vi and density ρi are available, they can be converted into mole fraction from the following relationship: vi ρ =Mi xi ¼ X i ðv ρ =Mi Þ i i i However, it should be pointed out that a major drawback of GC analysis is that no physical fractions are collected during this chromatographic analysis test and, therefore, their molecular weights “Mi” and specific gravities are not measured. As discussed later in this chapter, to describe these fractions in terms of their critical properties and acentric factors, the molecular weight and specific gravity must be known and used as input for most of the widely used property correlations. Ahmed (2014) pointed out that without the availability of molecular weight and specific gravity of pseudocomponents, the GC stimulated distillation data could still be used to characterize these fractions. The recommended proposed methodology is based on reproducing the GC stimulated boiling point data by a corresponding set of pseudocomponents with equivalent average boiling point temperatures to that of the GC stimulated boiling point data. These pseudocomponents are identified by a set of effective carbon numbers (ECN) that can be used to estimate their physical and critical properties. This approach relies on matching the boiling point temperature “Tb” by using Microsoft Solver and regressing on the ECN as given by the following relationship: Tb ¼ 427:2959078 + 50:08577848ðECNÞ 0:88693418ðECNÞ2 + 0:00675667ðECNÞ3 551:2778516 ðECNÞ Group 2. Stimulated Distillation by GC 91 where Tb is expressed in °R. The physical, critical properties, and acentric factor of a pseudocomponent can then be calculated from the following set of expressions: M ¼ 131:11375 + 24:96156ðECNÞ 0:34079022ðECNÞ2 + 0:002494118ðECNÞ3 + 468:32575 ðECNÞ γ ¼ 0:86714949 + 0:0034143408ðECNÞ 2:839627 1:1627984 105 ðECNÞ2 + 2:49433308 108 ðECNÞ3 ðECNÞ Tc ¼ 926:6022445 + 39:729363ðECNÞ 0:7224619ðECNÞ2 + 0:005519083ðECNÞ3 1366:4317487 ðECNÞ pc ¼ 311:236191 14:68693ðECNÞ + 0:3287671ðECNÞ2 1690:900114 0:0027346ðECNÞ3 + ðECNÞ Tb ¼ 427:295908 + 50:0857785ðECNÞ 0:8869342ðECNÞ2 + 0:00675667ðECNÞ3 551:277852 ðECNÞ ω ¼ 0:3142816 + 0:0780028ðECNÞ 0:00139205ðECNÞ2 + 1:02147 0:99102887 105 ðECNÞ3 + ðECNÞ vc ¼ 0:2328371 + 0:9741117ðECNÞ 0:009227ðECNÞ2 + 36:3611 0:1113515 106 ðECNÞ3 + ðECNÞ where, ECN ¼ effective carbon number Tc ¼ critical temperature of the fraction with a carbon number “n,” °R Tb ¼ boiling point temperature of the fraction with a carbon number “n,” °R pc ¼ critical pressure of the fraction with a carbon number “n,” psia vc ¼ critical volume of the fraction with a carbon number “n,” ft3/lbm-mol M ¼ molecular weight ω ¼ acentric factor γ ¼ specific gravity It should be pointed out that the number of carbon atoms “n” that is equivalent to that of n-alkane is calculated from n ¼ INT½ðECNÞ + 1 92 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions where, ECN ¼ effective carbon number n ¼ equivalent carbon number of n-alkane INT ¼ integer As a validation of Ahmed’s approach, assume the tabulated data listed in Table 2.24 are closely approximate the TBP. The objective is to estimate the ECN for each of the listed boiling point and compare with n-alkane carbon number and to calculate the physical and critical properties of each carbon number. Step 1. Using regression (Microsoft Solver), determine the ECN from the expression Tb ¼ 427:2959078 + 50:08577848ðECNÞ 0:88693418ðECNÞ2 + 0:00675667ðECNÞ3 551:2778516 ðECNÞ n ¼ INT½ðECNÞ + 1 To give: n Tb (°C) Tb (°R) ECN 5 6 7 8 9 10 11 12 14 15 16 17 18 20 24 28 32 36 40 21.9 54.7 84.1 111.9 136.3 159.7 181.9 201.9 220.8 239.7 256.9 273 288 310.9 376.9 416.9 451.9 481.9 508 530.82 589.86 642.78 692.82 736.74 778.86 818.82 854.82 888.84 922.86 953.82 982.8 1009.8 1051.02 1169.82 1241.82 1304.82 1358.82 1405.8 4.8 5.8 6.7 7.7 8.7 9.7 10.7 11.7 12.6 13.6 14.6 15.6 16.5 18.1 23.2 27.1 31.1 35 38.9 Group 3. The Plus Fraction Characterization Methods 93 Step 2. Calculate the physical and critical properties of each carbon number using the proposed working relationships and the values of ECN given above, to give n M Tc (°R) pc (psi) Tb (°R) W γ Vc (ft3/lbm-mol) 5.000 6.000 7.000 8.000 9.000 10.000 11.000 12.000 14.000 15.000 16.000 17.000 18.000 20.000 24.000 28.000 32.000 36.000 40.000 79.151 84.980 94.678 106.586 119.791 133.750 148.123 162.688 191.849 206.271 220.515 234.547 248.342 275.171 325.661 372.107 415.049 455.213 493.416 834.591 912.423 975.996 1030.221 1077.845 1120.526 1159.332 1194.988 1258.753 1287.520 1314.526 1339.947 1363.928 1408.036 1483.330 1544.968 1596.289 1640.091 1678.898 546.140 516.176 465.156 424.744 391.568 363.599 339.539 318.518 283.332 268.402 254.890 242.602 231.383 211.673 180.771 158.115 141.146 127.974 117.046 546.140 605.461 658.000 705.768 749.899 791.089 829.797 866.342 933.821 965.074 994.834 1023.198 1050.252 1100.727 1188.915 1262.975 1325.995 1380.843 1430.278 0.240 0.271 0.309 0.350 0.393 0.436 0.479 0.522 0.604 0.643 0.681 0.718 0.753 0.820 0.938 1.038 1.122 1.194 1.257 0.651 0.693 0.724 0.747 0.766 0.782 0.796 0.807 0.826 0.835 0.842 0.849 0.855 0.866 0.885 0.900 0.912 0.922 0.931 4.434 5.306 6.162 7.002 7.826 8.633 9.424 10.200 11.704 12.433 13.147 13.846 14.530 15.855 18.338 20.610 22.685 24.577 26.298 GROUP 3. THE PLUS FRACTION CHARACTERIZATION METHODS It is well known that C7+ characterization methodology has a significant impact on EOS predictions of reservoir hydrocarbon fluid behavior. The basis for most characterization methods is the TBP data that includes mass, mole, and volume fractions of each distillation cuts with a measured molecular weight, specific gravity, and boiling point. Each distillation cut is usually treated as pseudocomponent that can be characterized by a critical pressure, critical temperature, and acentric factor. If the distillation cuts are collected within a range of neighboring normal alkanes boiling points, they are referred to as SCN groups. Correlations for estimating their critical properties are commonly used based on the molecular weight and specific gravity of the SCN group. Based on the type of existing laboratory measured data on the plus fraction, two approaches commonly are used to generate properties of the plus fractions: 94 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions n n n Generalized correlations, PNA determination using analytical correlations, and PNA determination using graphical correlations. These two techniques of characterizing the undefined petroleum fractions are detailed next. Generalized Correlations To use any of the thermodynamic property prediction models, such as an equation of state, to predict the phase and volumetric behavior of complex hydrocarbon mixtures, one must be able to provide the acentric factor, ω, critical temperature, Tc, and critical pressure, pc, for both the defined and undefined (heavy) fractions in the mixture. The problem of how to adequately characterize these undefined plus fractions in terms of their critical properties and acentric factors has been long recognized in the petroleum industry. Whitson (1984) presented excellent documentation on the influence of various heptanes-plus (C7+) characterization schemes on predicting the volumetric behavior of hydrocarbon mixtures by equations of state. Numerous correlations are available for estimating the physical properties of petroleum fractions. Most of these correlations use the specific gravity, γ, and the boiling point, Tb, correlation parameters. However, coefficients of these correlations were determined from matching pure component physical and critical properties data, which might cause abnormality when applied to characterize the undefined fractions. Several of these correlations are presented next and referred to by the names of the authors, including: n n n n n n n n n n n n n n Riazi and Daubert Cavett Kesler-Lee Winn and Sim-Daubert Watansiri-Owens-Starling Edmister Critical compressibility factor correlations Rowe Standing Willman-Teja Hall-Yarborough Magoulas-Tassios Twu Silva and Rodriguez Group 3. The Plus Fraction Characterization Methods 95 Riazi and Daubert Generalized Correlations Riazi and Daubert (1980) developed a simple two-parameter equation for predicting the physical properties of pure compounds and undefined hydrocarbon mixtures. The proposed generalized empirical equation is based on the use of the normal boiling point and specific gravity as correlating parameters. The basic equation is θ ¼ aTbb γ c (2.4) where θ ¼ any physical property (Tc, pc, Vc or M) Tb ¼ normal boiling point, °R γ ¼ specific gravity M ¼ molecular weight Tc ¼ critical temperature, °R pc ¼ critical pressure, psia Vc ¼ critical volume, ft3/lbm a, b, c ¼ correlation constants are given in Table 2.5 for each property The expected average errors for estimating each property are included in the table. Note that the prediction accuracy is reasonable over the boiling point range of 100–850°F. Riazi and Daubert (1987) proposed improved correlations for predicting the physical properties of petroleum fractions by taking into consideration the following factors: accuracy, simplicity, generality, and availability of input parameters; ability to extrapolate; and finally, comparability with similar correlations developed in recent years. The authors proposed the following modification of Eq. (2.4), which maintains the simplicity of the previous correlation while significantly improving its accuracy: θ ¼ aθb1 θc2 exp ½dθ1 + eθ2 + f θ1 θ2 Table 2.5 Correlation Constants for Eq. (2.4) Deviation (%) θ M Tc (°R) pc (psia) Vc (ft3/lb) a b 5 4.56730 10 24.27870 3.12281 109 7.52140 103 c 2.19620 1.0164 0.58848 0.3596 2.31250 2.3201 0.28960 0.7666 Average Maximum 2.6 1.3 3.1 2.3 11.8 10.6 9.3 9.1 96 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions where θ ¼ any physical property a–f ¼ constants for each property Riazi and Daubert stated that θ1 and θ2 can be any two parameters capable of characterizing the molecular forces and molecular size of a compound. They identified (Tb,γ) and (M,γ) as appropriate pairs of input parameters in the equation. The authors finally proposed the following two forms of the generalized correlation. In the first form, the boiling point, Tb, and the specific gravity, γ, of the petroleum fraction are used as correlating parameters: θ ¼ aTbb γ c exp ½dTb + eγ + fTb γ (2.5) The constants a–f for each property θ are given in Table 2.6. In the second form, the molecular weight, M, and specific gravity, γ, of the component are used as correlating parameters. Their mathematical expression has the following form: θ ¼ aðMÞb γ c exp ½dM + eγ + f γM (2.6) In developing and obtaining the coefficients a–f, as given in Table 2.7, the authors used data on the properties of 38 pure hydrocarbons in the carbon number range 1–20, including paraffins, olefins, naphthenes, and aromatics in the molecular weight range 70–300 and the boiling point range 80–650°F. Table 2.6 Correlation Constants for Eq. (2.5) θ a b c d e f M Tc (°R) pc (psia) Vc (ft3/lb) 581.96000 10.6443 6.16200 106 6.23300 104 0.97476 0.81067 0.48440 0.75060 6.51274 0.53691 4.08460 1.20280 5.43076 104 5.17470 104 4.72500 103 1.46790 103 9.53384 0.54444 4.80140 0.26404 1.11056 103 3.59950 104 3.19390 103 1.09500 103 Table 2.7 Correlation Constants for Eq. (2.6) θ Tc (°R) Pc (psia) Vc (ft3/lb) Tb (°R) a 544.40000 4.52030 104 1.20600 102 6.77857 b 0.299800 0.806300 0.203780 0.401673 c 1.05550 1.60150 1.30360 1.58262 d 4 1.34780 10 1.80780 103 2.65700 103 3.77409 103 e f 0.616410 0.308400 0.528700 2.984036 0.00000 0.00000 2.60120 103 4.25288 103 Group 3. The Plus Fraction Characterization Methods 97 Cavett’s Correlations Cavett (1962) proposed correlations for estimating the critical pressure and temperature of hydrocarbon fractions. The correlations received wide acceptance in the petroleum industry due to their reliability in extrapolating conditions beyond those of the data used in developing the correlations. The proposed correlations were expressed analytically as functions of the normal boiling point, TbF, in °F and API gravity. Cavett proposed the following expressions for estimating the critical temperature and pressure of petroleum fractions: Tc ¼ a0 + a1 ðTbF Þ + a2 ðTbF Þ2 + a3 ðAPIÞðTbF Þ + a4 ðTbF Þ3 + a5 ðAPIÞðTbF Þ2 + a6 ðAPIÞ2 ðTbF Þ2 (2.7) log ðpc Þ ¼ b0 + b1 ðTbF Þ + b2 ðTbF Þ2 + b3 ðAPIÞðTbF Þ + b4 ðTbF Þ3 + b5 ðAPIÞðTbF Þ2 + b6 ðAPIÞ2 ðTbF Þ + b7 ðAPIÞ2 ðTbF Þ2 (2.8) where Tc ¼ critical temperature, °R pc ¼ critical pressure, psia TbF ¼ normal boiling point, °F API ¼ API gravity of the fraction Note that the normal boiling point in the above relationships is expressed in °F. The coefficients of Eqs. (2.7) and (2.8) are tabulated in the table below. However, Cavett presented these correlations without reference to the type and source of data used for their development. i ai bi 0 1 2 3 4 5 6 7 768.0712100000 1.7133693000 0.0010834003 0.0089212579 0.3889058400 106 0.5309492000 105 0.3271160000 107 2.82904060 0.94120109 103 0.30474749 105 0.20876110 104 0.15184103 108 0.11047899 107 0.48271599 107 0.13949619 109 Kesler and Lee Correlations Kesler and Lee (1976) proposed a set of equations to estimate the critical temperature, critical pressure, acentric factor, and molecular weight of 98 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions petroleum fractions. These relationships, as documented below, use specific gravity “γ” and boiling point “Tb” as input parameters for their proposed expressions: 0:0566 2:2898 0:11857 103 Tb 0:24244 + + γ γ γ2 3:648 0:47227 1:6977 7 2 10 1010 Tb3 + 1:4685 + T 0:42019 + + b γ γ2 γ2 (2.9) lnðpc Þ ¼ 8:3634 Tc ¼ 341:7 + 811:1γ + ½0:4244 + 0:1174γ Tb + ½0:4669 3:26238γ 105 (2.10) Tb M ¼ 12272:6 + 9486:4γ + ½4:6523 3:3287γ Tb + 1 0:77084γ 0:02058γ 2 720:79 107 181:98 1012 + 1 0:80882γ 0:02226γ 2 1:8828 1:3437 Tb Tb Tb Tb3 (2.11) The molecular weight equation was developed by regression analysis to match laboratory data with molecular weights ranging from 60 to 650. In addition, Kesler and Lee developed two expressions for estimating the acentric factor that use the Watson characterization factor and the reduced boiling point temperature as correlating parameters. The two correlating parameters are defined by the following two parameters: n n the Watson characterization factor “Kw”; ie, Kw ¼ (Tb)1/3/γ the reduced boiling point “θ” as defined by θ ¼ Tb/Tc where the boiling point Tb and critical temperature Tc are expressed in °R. Kessler and Lee proposed the following two expressions for calculating the acentric factor that are based on the value of the reduced boiling point: For θ > 0:8 : ω ¼ 7:904 + 0:1352Kw 0:007456ðKw Þ2 + 8:359θ + ln For θ < 0:8 : ω ¼ where 1:408 0:01063Kw θ (2.12) pc 6:09648 + 1:28862 ln ½θ 0:169347θ6 5:92714 + θ 14:7 15:6875 15:2518 13:4721 ln ½θ + 0:43577θ6 θ (2.13) pc ¼ critical pressure, psia Tc ¼ critical temperature, °R Tb ¼ boiling point, °R Group 3. The Plus Fraction Characterization Methods 99 ω ¼ acentric factor M ¼ molecular weight γ ¼ specific gravity Kesler and Lee stated that Eqs. (2.9) and (2.10) give values for pc and Tc that are nearly identical with those from the API data book up to a boiling point of 1200°F. Winn and Sim-Daubert Correlations Winn (1957) developed convenient nomographs to estimate various physical properties including molecular weight and the pseudocritical pressure for petroleum fractions. The input data in both nomographs are boiling point (or Kw) and the specific gravity (or API gravity). Sim and Daubert (1980) concluded that the Winn’s monograph is the most accurate approach for characterizing petroleum fractions. Sim and Daubert developed analytical relationships that closely matched the monograph graphical data. The authors used specific gravity and boiling point as the correlating parameters for calculating the critical pressure, critical temperature, and molecular weight. Sim and Daubert relationships as given below can be used to estimate the pseudocritical properties of the undefined petroleum fraction: pc ¼ 3:48242 109 Tb2:3177 γ 2:4853 (2.14) Tc ¼ exp 3:9934718T 0:08615 γ 0:04614 b (2.15) M ¼ 1:4350476 105 Tb2:3776 γ 0:9371 (2.16) where pc ¼ critical pressure, psia Tc ¼ critical temperature, °R Tb ¼ boiling point, °R Watansiri-Owens-Starling Correlations Watansiri et al. (1985) developed a set of correlations to estimate the critical properties and acentric factor of coal compounds and other undefined hydrocarbon components and their derivatives. The proposed correlations express the critical and physical properties of the undefined fraction as a function of the fraction normal boiling point, specific gravity, and molecular weight. These relationships have the following forms: lnðTc Þ ¼ 0:0650504 0:0005217Tb + 0:03095 ln ½M + 1:11067 lnðTb Þ h i + M 0:078154γ 1=2 0:061061γ 1=3 0:016943γ (2.17) 100 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions lnðvc Þ ¼ 76:313887 129:8038γ + 63:1750γ 2 13:175γ 3 + 1:10108 ln ½M + 42:1958 ln ½γ (2.18) 0:8 Tc M Tb 0:08843889 8:712 lnðpc Þ ¼ 6:6418853 + 0:01617283 Vc M Tc (2.19) 8 9 > 5:12316667 104 Tb + 0:281826667ðTb =MÞ + 382:904=M > > > > > > > > > 2 5 4 > > > + 0:074691 10 ðTb =γ Þ 0:12027778 10 ðTb ÞðMÞ > > > > > > > > > 2 4 > > > > < + 0:001261ðγ ÞðMÞ + 0:1265 10 ðMÞ =5T b ω¼ 1=3 ð T Þ > > 9M b 2 > > > > + 0:2016 104 ðγ ÞðMÞ 66:29959 > > > > M > > > > > > > > 2=3 > > > > T > > b > > : 0:00255452 2 ; γ (2.20) where ω ¼ acentric factor pc ¼ critical pressure, psia Tc ¼ critical temperature, °R Tb ¼ normal boiling point, °R vc ¼ critical volume, ft3/lb-mol The proposed correlations produce an average absolute relative deviation of 1.2% for Tc, 3.8% for vc, 5.2% for pc, and 11.8% for ω. Edmister’s Correlations Edmister (1958) proposed a correlation for estimating the acentric factor, ω, of pure fluids and petroleum fractions. The equation, widely used in the petroleum industry, requires boiling point, critical temperature, and critical pressure. The proposed expression is given by the following relationship: ω¼ 3½ log ðpc =14:70Þ 1 7½ðTc =Tb 1Þ where ω ¼ acentric factor pc ¼ critical pressure, psia Tc ¼ critical temperature, °R Tb ¼ normal boiling point, °R (2.21) Group 3. The Plus Fraction Characterization Methods 101 If the acentric factor is available from another correlation, the Edmister equation can be rearranged to solve for any of the three other properties (providing the other two are known). Critical Compressibility Factor Correlations The critical compressibility factor is defined as the component compressibility factor calculated at its critical point by using the component critical properties; ie, Tc, pc, and Vc. This property can be conveniently computed by the real gas equation of state at the critical point, or Zc ¼ p c vc RTc (2.22) where R ¼ universal gas constant, 10.73 psia ft3/lb-mol, °R vc ¼ critical volume, ft3/lb-mol If the critical volume, Vc, is given in ft3/lbm, Eq. (2.22) is written as Zc ¼ pc Vc M RTc where M ¼ molecule weight Vc ¼ critical volume, ft3/lb The accuracy of Eq. (2.22) depends on the accuracy of the values of pc, Tc, and vc. The following table presents a summary of the critical compressibility estimation methods that have been published over the years: Method Haugen Reid, Prausnitz, and Sherwood Salerno et al. Nath Year Zc Equation No. 1959 1977 Zc ¼ 1=ð1:28ω + 3:41Þ Zc ¼ 0:291 0:080ω (2.23) (2.24) 1985 1985 Zc ¼ 0:291 0:080ω 0:016ω2 Zc ¼ 0:2918 0:0928ω (2.25) (2.26) Rowe’s Characterization Method Rowe (1978) proposed a set of correlations for estimating the normal boiling point, the critical temperature, and the critical pressure of the heptanes-plus fraction C7+. The prediction of the C7+ properties is based on the assumption that the plus fraction behaves as a normal paraffin hydrocarbon fraction that has a number of carbon atoms “n” with identical critical properties to that of 102 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions the plus fraction. Rowe proposed the following set of formulas for characterizing the C7+ fraction in terms of the critical temperature, critical pressure, and boiling point temperature that are equivalent to properties of a normal paraffin hydrocarbon fraction with a number of carbon atoms “n” that can be calculated from the molecular weight of the C7+ fraction by the following relationship: n¼ MC7 + 2:0 14 ðTc ÞC7 + ¼ 1:8½961 10a (2.27) (2.28) The coefficient “a” as defined by a ¼ 2:95597 0:090597n2=3 The author correlated the critical pressure and the boiling point temperature of the C7+ fraction in terms of the critical temperature and number of carbon atoms by the following expressions: ðpc ÞC7 + ¼ 10ð4:89165 + Y Þ ðTc ÞC7 + (2.29) with the parameter “Y” as given by Y ¼ 0:0137726826 n + 0:6801481651 ðTb ÞC7 + ¼ 0:0004347ðTc Þ2C7 + + 265 (2.30) where ðTc ÞC7 + ¼ critical temperature of C7+, °R ðTb ÞC7 + ¼ boiling point temperature of C7+, °R ðpc ÞC7 + ¼ critical pressure of the C7+ in psia MC7 + ¼ molecular weight of the heptanes-plus fraction Soreide (1989) proposed an alternative methodology of calculating the boiling point that was generated from the analysis of 843 TBP fractions taken from 68 reservoir C7+ samples. Soreide proposed the boiling point temperature correlation is expressed as a function of molecular weight and specific gravity of the plus fraction: Tb ¼ 1928:3 1:695 105 γ 3:266 exp 4:922 103 M 4:7685γ + 3:462 103 Mγ M0:03522 where Tb is expressed in °R. Group 3. The Plus Fraction Characterization Methods 103 Standing’s Correlations Matthews et al. (1942) presented graphical correlations for determining the critical temperature and pressure of the heptanes-plus fraction. Standing (1977) expressed these graphical correlations more conveniently in mathematical forms as follows: h i h i ðTc ÞC7 + ¼ 608 + 364log ðMÞC7 + 71:2 + 2450log ðMÞC7 + 3800 log ðγ ÞC7 + (2.31) h i ðpc ÞC7 + ¼ 1188 431log ðMÞC7 + 61:1 n h ih io + 2319 852 log ðMÞC7 + 53:7 ðγ ÞC7 + 0:8 (2.32) where ðMÞC7 + and ðγ ÞC7 + are the molecular weight and specific gravity of the C7+. Willman-Teja Correlations Willman and Teja (1987) proposed correlations for determining the critical pressure and critical temperature of the n-alkane homologous series. The authors used the normal boiling point and the number of carbon atoms of the n-alkane as a correlating parameter. The authors used a nonlinear regression model to generate a set of relationships that best match the critical properties data reported by Bergman et al. (1977) and Whitson (1980). Willman and Teja introduced the ECN parameter “n” into their proposed correlations and suggested that this correlating parameter can be calculated by matching the boiling point of the undefined petroleum fraction. Ahmed (2014) suggested that the ECN can be best estimated by regression (eg, Microsoft Solver) to match the boiling point temperature “Tb” as given by following expression: Tb ¼ 434:38878 + 50:125279n 0:9097293n2 + 7:0280657 103 n3 601:85651=n where Tb is expressed in °R. The critical temperature and pressure can then be calculated from the following expressions: h i Tc ¼ Tb 1 + ð1:25127 + 0:137252nÞ0:884540633 Pc ¼ 339:0416805 + 1184:157759n ½0:873159 + 0:54285n1:9265669 where n ¼ effective carbon number Tb ¼ average boiling point of the undefined fraction, °R (2.33) (2.34) 104 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions Tc ¼ critical temperature of the undefined fraction, °R Pc ¼ critical pressure of the undefined fraction, psia Hall and Yarborough Correlations Hall and Yarborough (1971) proposed a simple expression to determine the critical volume of a fraction from its molecular weight and specific gravity: vc ¼ 0:025M1:15 γ 0:7935 (2.35) where vc is the critical volume as expressed in ft3/lb-mol. Note that, to express the critical volume in ft3/lb, the relationship is given by vc ¼ MVc where M ¼ molecular weight Vc ¼ critical volume in ft3/lb vc ¼ critical volume in ft3/lb-mol The critical volume also can be calculated by applying the real gas equation of state at the critical point of the component as pV ¼ Z m M RT Applying real gas equation at the critical point, gives Vc ¼ Zc RTc pc M Magoulas and Tassios Correlations Magoulas and Tassios (1990) correlated the critical temperature, critical pressure, and acentric factor with the specific gravity, γ, and molecular weight, M, of the fraction as expressed by the following relationships: Tc ¼ 1247:4 + 0:792M + 1971γ lnð pc Þ ¼ 0:01901 0:0048442M + 0:13239γ + 27000 707:4 + M γ 227 1:1663 + 1:2702 lnðMÞ M γ ω ¼ 0:64235 + 0:00014667M + 0:021876γ where Tc ¼ critical temperature, °R pc ¼ critical pressure, psia 4:559 + 0:21699 ln ðMÞ M Group 3. The Plus Fraction Characterization Methods 105 Twu’s Correlations Twu (1984) developed a suite of critical properties, based on perturbationexpansion theory with normal paraffins as the reference states, which can be used for determining the critical and physical properties of undefined hydrocarbon fractions, such as C7+. The methodology is based on selecting (finding) a normal paraffin fraction with a boiling temperature TbP identical to that of the hydrocarbon-plus fraction “TbC + ,” such as C7+. The methodology requires the availability of the boiling point temperature of the plus fraction “TbC + ,” molecular weight of the plus fraction “MC + ,” as well as its specific gravity “γ C + .” If the boiling point temperature is not available, it can be estimated from the correlation proposed by Soreide (1989): a TbC + ¼ a1 + a2 ðMC + Þa3 γ C + 4 exp a5 MC + + a6 γ C + + a7 MC + γ C + where a1 ¼ 1928.3 a2 ¼ 1.695(105) a3 ¼ 0.03522 a4 ¼ 3.266 a5 ¼ 4.922(103) a6 ¼ 4.7685 a7 ¼ 3.462(103) The Twu approach is performed by applying the following two steps: Step 1. Calculate the properties of the normal paraffins, ie, TcP, pcP, γ P, and vcP, from the following set of working expressions: n Critical temperature of normal paraffins, TcP, in °R: " TcP ¼ TbC + 2 3 A1 + A2 TbC + + A3 TbC + A4 TbC + + + A5 ðA6 TbC + Þ13 where A1 ¼ 0.533272 A2 ¼ 0.191017(103) A3 ¼ 0.779681(107) A4 ¼ 0.284376(1010) A5 ¼ 0.959468(102) A6 ¼ 0.01 n Critical pressure of normal paraffins, pcP, in psia: 2 4 pcP ¼ A1 + A2 α0:5 i + A3 αi + A4 αi + A5 αi 2 #1 106 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions With the parameter αi as defined by the expression αi ¼ 1 TbC + TcP where A1 ¼ 3.83354 A2 ¼ 1.19629 A3 ¼ 34.8888 A4 ¼ 36.1952 A5 ¼ 104.193 n Specific gravity of normal paraffins, γ P: γ P ¼ A1 + A2 αi + A3 α3i + A4 α12 i where A1 ¼ 0.843593 A2 ¼ 0.128624 A3 ¼ 3.36159 A4 ¼ 13749.5 n Critical volume of normal paraffins, vcP, in ft3/lbm-mol: vcP ¼ 1 + A1 + A2 αi + A3 α3i + A4 α14 i 8 where A1 ¼ 0.419869 A2 ¼ 0.505839 A3 ¼ 1.56436 A4 ¼ 9481.7 Step 2. Calculate the properties of the plus fraction by applying the following relationships: n Critical temperature of the plus fraction, TC + , in °R: TC + ¼ TcP 1 + 2fT 2 1 2fT With the function “fT” as given by fT ¼ exp 5 γ p γ C + 1 " # ! A1 A3 exp 5 γ p γ C + 1 + A2 + 0:5 0:5 TbC TbC + + where A1 ¼ 0.362456 A2 ¼ 0.0398285 A3 ¼ 0.948125 Group 3. The Plus Fraction Characterization Methods 107 n Critical volume of the plus fraction, vC + , in ft3/lbm-mol: vC + ¼ vcP 1 + 2fv 2 1 2fv with ! " # h i o A h i o A3 n 1 2 2 2 2 exp 4 γ p γ C + 1 fv ¼ exp 4 γ p γ C + 1 + A2 + 0:5 0:5 TbC TbC + + n where A1 ¼ 0.466590 A2 ¼ 0.182421 A3 ¼ 3.01721 n Critical pressure of the plus fraction, pC + , in psia: pC + ¼ pcP TC + TcP vcP vC + 1 + 2fp 2 1 2fp with n i o fp ¼ exp 0:5 γ p γ C + 1 n h h i o exp 0:5 γ p γ C + 1 ! A2 A1 + 0:5 + A3 TbC + TbC + !! ! A5 + A4 + 0:5 + A6 TbC + TbC + where A1 ¼ 2.53262 A2 ¼ 46.19553 A3 ¼ 0.00127885 A4 ¼ 11.4277 A5 ¼ 252.14 A6 ¼ 0.00230535 Silva and Rodriguez Correlations Silva and Rodriguez (1992) suggested the use of the following two expressions to estimate the boiling point temperature and specific gravity from the molecular weight: M Tb ¼ 460 + 447:08723 ln 64:2576 Use the preceding calculated value of Tb to calculate the specific gravity of the fraction from the following expression: γ ¼ 0:132467 ln ðTb 460Þ + 0:0116483 where the boiling point temperature Tb is expressed in °R. 108 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions Sancet Correlations Sancet (2007) presented a set expressions to estimate the critical properties and boiling point from the molecular weight “M”; these correlations have the following forms: pc ¼ 653 exp ð0:007427MÞ + 82:82 Tc ¼ ½383:5 ln ðM 4:075Þ 778:5 h i Tb ¼ 0:001241ðTc Þ1:869 + 194 where the boiling point temperature Tb and Tc are expressed in °R. Comparison of Molecular Weight Methods Correlations Mudgal (2014) compare several methodologies for predicting the molecular weight from the boiling point temperature data. The author estimated the molecular weight by using Lee-Kesler, Riazi-Daubert, Winn nomographs, Goossens, and Twu methods. Mudgal concluded that Goossens and Twu methods yielded appropriate estimations of the molecular weight. In addition, Ahmed (2014) applied the concept of the ECN and equivalent carbon number “n” to match Mudgal’s boiling point temperature “Tb”; as given previously in the chapter by Tb ¼ 427:2959078 + 50:08577848ðECNÞ 0:88693418ðECNÞ2 + 0:00675667ðECNÞ3 551:2778516 ðECNÞ with ‘n’ as given by n ¼ INT½ðECNÞ + 1 The equivalent carbon number “n” is used to estimate the molecular weight from M ¼ 131:11375 + 24:96156n 0:34079022n2 + 0:002494118n3 + ð468:32575=nÞ The above expression shows a good agreement with Twu’s methodology; the tabulated comparison is shown in Table 2.8 and presented in a graphical form in Fig. 2.8. Group 3. The Plus Fraction Characterization Methods 109 Table 2.8 Comparison Molecular Weight Correlations LeeKesler RiazDaubert Winn Twu Goossens Ahmed Fractions Tb (°K) Tb (°R) M M M M M n M 1 2 3 4 5 6 7 8 9 10 Residue 363.15 403.15 453.15 473.15 498.15 548.15 583.15 593.15 613.15 644.35 764.83 653.67 725.67 815.67 851.67 896.67 986.67 1049.67 1067.67 1103.67 1159.83 1376.694 87.78 106.15 131.79 143.32 155.18 190.82 219.15 225.15 241.19 268.56 434.2 92.68 117.51 149.16 160.25 178.52 217.03 248.41 256.75 274.79 305.92 450.25 91.02 112.55 139.77 148.29 163.66 194.13 219.9 225.48 239.71 265.62 414.18 94.3 115.35 144.77 155.62 173.02 209.98 240.97 249.21 267.45 300.34 474.49 92 112 143.5 156 175 225.5 242 250 251 270 345 7 8.4 10.6 11.6 12.9 15.7 18 18.7 20.2 22.7 36.5 94.67761 111.7485 142.3398 156.8496 175.8379 216.263 248.3421 257.8502 277.7957 309.7129 460.0781 Molecular weight vs. boiling point Lee-Kesler Riaz-Daubert Winn Nomogram Twu Goossens Ahmed Silva-Rodriguez 1200 1300 1400 600 500 400 M 300 200 100 0 600 700 800 900 n FIGURE 2.8 Comparison of molecular weight estimation methods. 1000 1100 Tb (°R) 1500 110 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions EXAMPLE 2.1 Estimate the critical properties and the acentric factor of the heptanes-plus fraction, that is, C7+, with a measured molecular weight of 150 and specific gravity of 0.78 by using: n n the Riazi-Daubert equation ((2.6)) and Edmister’s equation for calculating the acentric factor Eqs. (2.3A)–(2.3E) Solution Applying the Riazi-Daubert equation: n θ ¼ aðMÞb γ c exp ½dM + eγ + f γM Tc ¼ 544:2ð150Þ0:2998 ð0:78Þ1:0555 exp 1:3478 104 ð150Þ 0:61641ð0:78Þ + 0 ¼ 1139:4°R pc ¼ 4:5203 104 ð150Þ0:8063 ð0:78Þ1:6015 exp 1:8078 103 ð150Þ 0:3084ð0:78Þ + 0 ¼ 320:3 psia Vc ¼ 1:206 102 ð150Þ0:20378 ð0:78Þ1:3036 exp 2:657 103 ð150Þ + 0:5287ð0:78Þ2:6012 103 ð150Þð0:78Þ ¼ 0:06035 ft3 =lb Tb ¼ 6:77857ð150Þ0:401673 ð0:78Þ1:58262 exp 3:77409 103 ð150Þ + 2:984036ð0:78Þ 4:25288 103 ð150Þð0:78Þ ¼ 825:26°R Use Edmister’s equation (Eq. (2.21)) to estimate the acentric factor: ω¼ 3½ log ðpc =14:70Þ 1 7½Tc =Tb 1 ω¼ 3½ log ð320:3=14:7Þ 1 ¼ 0:5067 7½1139:4=825:26 1 Applying Eqs. (2.3A)–(2.3E) 0:351579776 M M 233:3891996 ¼ 1170°R Tc ¼ 231:9733906 0:352767639 γ γ n 0:885326626 M M pc ¼ 31829 0:106467189 + 49:62573013 ¼ 331:6 psi γ γ M M + 0:00141973 1:243207091 ¼ 0:499 ω ¼ 0:279354619 ln γ γ 0:984070268 M vc ¼ 0:054703719 Mγ + 87:7041 106 γ 0:001007476 ¼ 9:69 ft3 =lbm mol ¼ 0:0646 ft3 =lbm Group 3. The Plus Fraction Characterization Methods 111 0:664115311 M M 379:4428973 Tb ¼ 33:72211 1:40066452 ¼ 837°R γ γ ðM=γ Þ n Silva and Rodriguez Correlations Boiling point temperature and specific gravity of the cut are not available: M ¼ 840°R Tb ¼ 460 + 447:08723 ln 64:2576 γ ¼ 0:132467 ln ðTb 460Þ + 0:0116483 ¼ 0:7982 The following table summarizes the results for this example. Method Tc (°R) pc (psia) Vc (ft3/lb) Tb (°R) ω Riazi-Daubert Silva and Rodriguez Sancet Eqs. (2.3A)–(2.3E) 1139 – 1133 1170 320 – 297 331 0.0604 – – 0.0646 825 840 828 837 0.507 – – 0.499 EXAMPLE 2.2 Estimate the critical properties, molecular weight, and acentric factor of a petroleum fraction with a boiling point of 198°F (658°R) and specific gravity of 0.7365 by using the following methods: (a) (b) (c) (d) (e) (f) (g) (h) Riazi-Daubert (Eq. (2.5)) Riazi-Daubert (Eq. (2.5)) Cavett Kesler-Lee Winn and Sim-Daubert Watansiri-Owens-Starling Willman and Teja Eqs. (2.3F)–(2.3J) Solution (a) Using Riazi-Daubert (Eq. (2.4)) θ ¼ aTbb γ c M ¼ 4:5673 105 ð658Þ2:1962 ð0:7365Þ1:0164 ¼ 96:4 Tc ¼ 24:2787ð658Þ0:58848 ð0:7365Þ0:3596 ¼ 990:67°R pc ¼ 3:12281 109 ð658Þ2:3125 ð0:7365Þ2:3201 ¼ 466:9 psia Vc ¼ 7:5214 103 ð658Þ0:2896 ð0:7365Þ0:7666 ¼ 0:06227 ft3 =lb 112 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions Solve for Zc by applying Eq. (2.22): Zc ¼ pc Vc M ð466:9Þð0:06227Þð96:4Þ ¼ ¼ 0:26365 RTc ð10:73Þð990:67Þ Solve for ω by applying Eq. (2.21): ω¼ 3½ log ðpc =14:70Þ 1 7½Tc =Tb 1 ω¼ 3½ log ð466:9=14:7Þ 1 ¼ 0:2731 7½ð990:67=658Þ 1 (b) Riazi-Daubert (Eq. (2.5)) θ ¼ aTbb γ c exp ½dTb + eγ + fTb γ Applying the above expression and using the appropriate constants yields: M ¼ 581:96ð658Þ0:97476 0:73656:51274 exp 5:4307 104 ð658Þ + 9:53384ð0:7365Þ + 1:11056 103 ð658Þð0:7365Þ M ¼ 96:911 Tc ¼ 10:6443ð658Þ0:810676 0:73650:53691 exp 5:1747 104 ð658Þ 0:54444ð0:7365Þ + 3:5995 104 ð658Þð0:7365Þ Tc ¼ 985:7°R Similarly, pc ¼ 465:83 psia Vc ¼ 0:06257 ft3 =lb The acentric factor and critical compressibility factor can be obtained by applying Eqs. (2.21) and (2.22), respectively. ω¼ 3½ log ðpc =14:70Þ 3½ log ð465:83=14:70Þ 1¼ 1 ¼ 0:2877 7½Tc =Tb 1 7½ð986:7=658Þ 1 Zc ¼ pc Vc M ð465:83Þð0:06257Þð96:911Þ ¼ ¼ 0:2668 RTc 10:73ð986:7Þ (c) Cavett’s Correlations: Step 1. Solve for Tc using Eq. (2.7) with the coefficients i ai bi 0 1 2 3 4 5 6 7 768.0712100000 1.7133693000 0.0010834003 0.0089212579 0.3889058400 106 0.5309492000 105 0.3271160000 107 2.82904060 0.94120109 103 0.30474749 105 0.20876110 104 0.15184103 108 0.11047899 107 0.48271599 107 0.13949619 109 Group 3. The Plus Fraction Characterization Methods 113 Tc ¼ a0 + a1 ðTb Þ + a2 ðTb Þ2 + a3 ðAPIÞðTb Þ + a4 ðTb Þ3 + a5 ðAPIÞðTb Þ2 + a6 ðEÞ2 ðTb Þ2 to give Tc ¼ 978.1°R Step 2. Calculate pc with Eq. (2.8): log ðpc Þ ¼ b0 + b1 ðTb Þ + b2 ðTb Þ2 + b3 ðAPIÞðTb Þ + b4 ðTb Þ3 + b5 ðAPIÞðTb Þ2 + b6 ðAPIÞ2 ðTb Þ + b7 ðAPIÞ2 ðTb Þ2 to give pc ¼ 466.1 psia. Step 3. Solve for the acentric factor by applying the Edmister correlation (Eq. (2.21)): ω¼ 3½ log ð466:1=14:7Þ 1 ¼ 0:3147 7½ð980=658Þ 1 Step 4. Compute the critical compressibility by using Eq. (2.25): Zc ¼ 0:291 0:080ω 0:016ω2 Zc ¼ 0:291 ð0:08Þð0:3147Þ 0:016ð0:3147Þ2 ¼ 0:2642 Step 5. Estimate vc from Eq. (2.22): vc ¼ Zc RTc ð0:2642Þð10:731Þð980Þ ¼ ¼ 5:9495 ft3 =lb mol pc 466:1 Assume a molecular weight of 96; estimate the critical volume: Vc ¼ 5:9495 ¼ 0:06197 ft3 =lb 96 (d) Kesler and Lee Correlations: Step 1. Calculate pc from Eq. (2.9): lnðpc Þ ¼ 8:3634 0:0566=γ 0:24244 + 2:2898=γ + 0:11857=γ 2 103 Tb + 1:4685 + 3:648=γ + 0:47227=γ 2 107 Tb2 0:42019 + 1:6977=γ 2 1010 Tb3 to give pc ¼ 470 psia. Step 2. Solve for Tc by using Eq. (2.10); that is, Tc ¼ 341:7 + 811:1γ + ½0:4244 + 0:1174γ Tb + ½0:4669 3:26238γ 105 Tb to give Tc ¼ 980°R. Step 3. Calculate the molecular weight, M, by using Eq. (2.11): M ¼ 12,272:6 + 9,486:4γ + ½4:6523 3:3287γ Tb + 1 0:77084γ 0:02058γ 2 720:79 107 181:98 1012 1:3437 + 1 0:80882γ 0:02226γ 2 1:8828 Tb Tb Tb Tb3 to give M ¼ 98.7 114 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions Step 4. Compute the Watson characterization factor K and the parameter θ: ð658Þ1=3 ¼ 11:8 0:7365 658 θ¼ ¼ 0:671 980 K¼ Step 5. Solve for acentric factor by applying Eq. (2.13): ω¼ ln h p i 6:09648 + 1:28862 ln ½θ 0:169347θ6 θ ¼ 0:306 15:6875 15:2518 13:4721 ln ½θ + 0:43577θ6 θ c 14:7 5:92714 + Step 6. Estimate for the critical gas compressibility, Zc, by using Eq. (2.26): Zc ¼ 0:2918 0:0928ω Zc ¼ 0:2918 ð0:0928Þð0:306Þ ¼ 0:2634 Step 7. Solve for Vc by applying Eq. (2.22): Vc ¼ Zc RTc ð0:2634Þð10:73Þð980Þ ¼ ¼ 0:0597 ft3 =lb pc M ð470Þð98:7Þ (e) Winn-Sim-Daubert Correlations: Step 1. Estimate pc from Eq. (2.14): γ 2:4853 pc ¼ 3:48242 109 2:3177 Tb pc ¼ 478:6 psia Step 2. Solve for Tc by applying Eq. (2.15): Tc ¼ exp 3:9934718Tb0:08615 γ 0:04614 Tc ¼ 979:2°R Step 3. Calculate M from Eq. (2.16): Tb2:3776 M ¼ 1:4350476 105 0:9371 γ M ¼ 95:93 Step 4. Solve for the acentric factor from Eq. (2.21): ω¼ 3½ log ðpc =14:70Þ 1 7½ðTc =Tb 1Þ ω ¼ 0.3280 Group 3. The Plus Fraction Characterization Methods 115 Solve for Zc by applying Eq. (2.24): Zc ¼ 0:291 0:080ω Zc ¼ 0:291 ð0:08Þð0:3280Þ ¼ 0:2648 Step 5. Calculate the critical volume Vc from Eq. (2.22): Vc ¼ Vc ¼ Zc RTc pc M ð0:2648Þð10:731Þð979:2Þ ¼ 0:06059 ft3 =lb ð478:6Þð95:93Þ (f) Watansiri-Owens-Starling Correlations: Step 1. Because Eqs. (2.17) through (2.19) require the molecular weight, assume M ¼ 96. Step 2. Calculate Tc from Eq. (2.17): lnðTc Þ ¼ 0:0650504 0:00052217Tb + 0:03095 ln ½M + 1:11067 lnðTb Þ h i + M 0:078154γ 1=2 0:061061γ 1=3 0:016943γ Tc ¼ 980:0°R Step 3. Determine the critical volume from Eq. (2.18) to give lnðVc Þ ¼ 76:313887 129:8038γ + 63:1750γ 2 13:175γ 3 + 1:10108 ln ½M + 42:1958 ln ½γ Vc ¼ 0:06548 ft3 =lb Step 4. Solve for the critical pressure of the fraction by applying Eq. (2.19) to produce 0:8 Tc M Tb 0:08843889 8:712 lnðpc Þ ¼ 6:6418853 + 0:01617283 Vc M Tc pc ¼ 426:5 psia Step 5. Calculate the acentric factor from Eq. (2.20) 8 9 5:12316667 104 Tb + 0:281826667ðTb =MÞ + 382:904=M > > > > > > > > 2 > > 5 4 > > ð T =γ Þ 0:12027778 10 ð T Þ ð M Þ + 0:074691 10 > > b b > > > > > > 2 4 > > > > ð Þ ð Þ ð Þ M + 0:001261 γ M + 0:1265 10 < =5T b ω¼ 1=3 ðTb Þ > > 9M 2 4 > > > > > + 0:2016 10 ðγ ÞðMÞ 66:29959 M > > > > > > > > > > > 2=3 > > T > > b > > : 0:00255452 2 ; γ to give ω ¼ 0.2222 116 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions Step 6. Compute the critical compressibility factor by applying Eq. (2.24): Zc ¼ 0:2918 0:0928ω ¼ 0:2918 0:0928ð0:2222Þ ¼ 0:27112 (g) Willman and Teja Correlations: Step 1. Using Microsoft Solver, determine the ECN “n” by matching the given boiling point “Tb” Tb ¼ 434:38878 + 50:125279n 0:9097293n2 + 7:0280657 103 n3 601:85651=n 658 ¼ 434:38878 + 50:125279n 0:9097293n2 + 7:0280657 103 n3 601:85651=n To give n ¼ 7 Step 2. Calculate Tc from Eq. (2.33): h i Tc ¼ Tb 1 + ð1:25127 + 0:137252nÞ0:884540633 h i Tc ¼ Tb 1 + ð1:25127 + 0:137252ð7ÞÞ0:884540633 ¼ 983:7°R Step 3. Calculate pc from Eq. (2.34): Pc ¼ Pc ¼ 339:0416805 + 1184:157759n ½0:873159 + 0:54285n1:9265669 339:0416805 + 1184:157759ð7Þ ½0:873159 + 0:54285ð7Þ1:9265669 ¼ 441:8 psia (h) Eqs. (2.3F) through (2.3J): 0:967105064 Tb Tb + 0:010009512 ¼ 996°R Tc ¼ 1:379242274 + 0:012011487 γ γ Tb pc ¼ 2898:631565 exp ½0:002020245 ðTb =γ Þ 0:00867297 + 0:099138742 γ ¼ 469 psi 1:250922701 Tb Tb ω ¼ 0:000115057 0:92098982 ¼ 0:26 + 0:00068975 γ γ 3 2 Tb vc ¼ 8749:94 106 Tγb 9:75748 106 Tγb + 0:014653 γ 595:7922076 ¼ 6:273 ft3 =lbm mol Tb γ 0:001617123Tb M ¼ 52:11038728exp 132:1481762 ¼ 89 γ 4:60231684 Group 3. The Plus Fraction Characterization Methods 117 The following table summarizes the results for this example. Tc (°R) pc (psia) Vc (ft3/lb) M Method Riazi-Daubert no. 1 990.67 Riazi-Daubert no. 2 986.70 Cavett 978.10 Kesler-Lee 980.00 Winn 979.20 Watansiri 980.00 Willman-Teja 983.7 Eqs. (2.3F)–(2.3J) 996 466.90 465.83 466.10 469.00 478.60 426.50 441.8 469 0.06227 0.06257 0.06197 0.05970 0.06059 0.06548 – 0.0689 96.400 96.911 – 98.700 95.930 – – 89 ω Zc 0.2731 0.26365 0.2877 0.26680 0.3147 0.26420 0.3060 0.26340 0.3280 0.26480 0.2222 0.27112 – – 0.26 – EXAMPLE 2.3 If the molecular weight and specific gravity of the heptanes-plus fraction are 216 and 0.8605, respectively, calculate the critical temperature and pressure by using (a) Rowe’s correlations. (b) Standing’s correlations. (c) Magoulas-Tassios’s correlations. (d) Eqs. (2.3A) and (2.3B). Solution (a) Rowe’s correlations Step 1. Calculate the number of carbon atoms of C7+ from Eq. (2.28) to give n¼ MC7 + 2:0 216 2:0 ¼ ¼ 15:29 14 14 Step 2. Calculate the coefficient a: a ¼ 2:95597 0:090597ðnÞ2=3 a ¼ 2:95597 0:090597ð15:29Þ2=3 ¼ 2:39786 Step 3. Solve for the critical temperature from Eq. (2.27) to yield ðTc ÞC7 + ¼ 1:8½961 10a ðTc ÞC7 + ¼ 1:8 961 102:39786 ¼ 1279:8°R Step 4. Calculate the coefficient Y: Y ¼ 0:0137726826n + 0:6801481651 Y ¼ 0:0137726826ð2:39786Þ + 0:6801481651 ¼ 0:647123 118 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions Step 5. Solve for the critical pressure from Eq. (2.29) to give ðpc ÞC7 + ¼ 10ð4:89165 + Y Þ 10ð4:89165 + 0:647123Þ ¼ ¼ 270 psi ðTc ÞC7 + 1279:8 (b) Standing’s correlations Step 1. Solve for the critical temperature by using Eq. (2.31) to give h i h i ðTc ÞC7 + ¼ 608 + 364 log ðMÞC7 + 71:2 + 2450log ðMÞC7 + 3800 log ðγ ÞC7 + ðTc ÞC7 + ¼ 1269:3°R Step 2. Calculate the critical pressure from Eq. (2.32) to yield h i h h i ðpc ÞC7 + ¼ 1188 431 log ðMÞC7 + 61:1 + 2319 852 log ðMÞC7 + 53:7 h i ðγ ÞC7 + 0:8 ðpc ÞC7 + ¼ 270 psia (c) Magoulas-Tassios correlations Tc ¼ 1247:4 + 0:792M + 1971γ 27,000 707:4 + M γ 27, 000 707:4 + ¼ 1317°R 216 0:8605 227 1:1663 + 1:2702 lnðMÞ lnðpc Þ ¼ 0:01901 0:0048442M + 0:13239γ + M γ Tc ¼ 1247:4 + 0:792ð216Þ + 1971ð0:8605Þ lnðpc Þ ¼ 0:01901 0:0048442ð216Þ + 0:13239ð0:8605Þ 227 1:1663 + 1:2702 ln ð216Þ ¼ 5:6098 216 0:8605 pc ¼ exp ð5:6098Þ ¼ 273 psi + (d) Eqs. (2.3A) and (2.3B) 0:351579776 M M Tc ¼ 231:9733906 0:352767639 233:3891996 ¼ 1294°R γ γ pc ¼ 31829 0:885326626 M M 0:106467189 + 49:62573013 ¼ 262 psi γ γ Summary of results is given below: Method Tc (°R) pc (psia) Rowe Standing Magoulas-Tassios Eqs. (2.3A) and (2.3B) 1279 1269 1317 1294 270 270 273 262 Group 3. The Plus Fraction Characterization Methods 119 EXAMPLE 2.4 Calculate the critical properties and the acentric factor of C7+ with a measured molecular weight of 198.71 and specific gravity of 0.8527. Employ the following methods: (a) Rowe’s correlations. (b) Standing’s correlations. (c) Riazi-Daubert’s correlations. (d) Magoulas-Tassios’s correlations. (e) Eqs. (2.3A) through (2.3E). Solution (a) Rowe’s correlations Step 1. Calculate the number of carbon atoms, n, and the coefficient a of the fraction to give n¼ M 2 198:71 2 ¼ ¼ 14:0507 14 14 a ¼ 2:95597 0:090597n2=3 a ¼ 2:95597 0:090597ð14:0507Þ2=3 ¼ 2:42844 Step 2. Determine Tc from Eq. (2.28) to give ðTc ÞC7 + ¼ 1:8½961 10a ðTc ÞC7 + ¼ 1:8 961 102:42844 ¼ 1247°R Step 3. Calculate the coefficient Y: Y ¼ 0:0137726826n + 0:6801481651 Y ¼ 0:0137726826ð2:42844Þ + 0:6801481651 ¼ 0:6467 Step 4. Compute pc from Eq. (2.29) to yield ðpc ÞC7 + ¼ 10ð4:89165 + 0:6467Þ ¼ 277 psi 1247 Step 5. Determine Tb by applying Eq. (2.30) to give ðTb ÞC7 + ¼ 0:0004347ðTc Þ2C7 + + 265 ¼ 0:0004347ð1247Þ2 + 265 ¼ 941°R Step 6. Solve for the acentric factor by applying Eq. (2.21) to give ω¼ 3½ log ðpc =14:70Þ 1 7½ðTc =Tb 1Þ ω ¼ 0:6123 120 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions (b) Standing’s correlations Step 1. Solve for the critical temperature of C7+ by using Eq. (2.31) to give h i h i ðTc ÞC7 + ¼ 608 + 364 log ðMÞC7 + 71:2 + 2450log ðMÞC7 + 3800 log ðγ ÞC7 + ðTc ÞC7 + ¼ 1247:73°R Step 2. Calculate the critical pressure from Eq. (2.32) to give h i h h i ðpc ÞC7 + ¼ 1188 431 log ðMÞC7 + 61:1 + 2319 852 log ðMÞC7 + 53:7 h i ðγ ÞC7 + 0:8 ðpc ÞC7 + ¼ 291:41 psia (c) Riazi-Daubert correlations θ ¼ aðMÞb γ c exp ½dM + eγ + f γM Tc ¼ 544:2ð198:71Þ0:2998 ð0:8577Þ1:0555 exp 1:3478 104 ð198:71Þ 0:61641ð0:8577Þ ¼ 1294°R pc ¼ 4:5203 104 ð198:71Þ0:8061 ð0:8577Þ1:6015 exp 1:8078 103 ð198:71Þ 0:3084ð0:8577Þ ¼ 264 psi Determine Tb by applying Eq. (2.6) to give Tb ¼ 958.5°R. Solve for the acentric factor from Eq. (2.21) to give ω¼ 3½ log ðpc =14:70Þ 1 7½ðTc =Tb 1Þ ω ¼ 0:5346 (d) Magoulas-Tassios correlations Tc ¼ 1247:4 + 0:792M + 1971γ 27,000 707:4 + M γ Tc ¼ 1247:4 + 0:792ð198:71Þ + 1971ð0:8527Þ lnðpc Þ ¼ 0:01901 0:0048442M + 0:13239γ + 27, 000 707:4 + ¼ 1284°R 198:71 0:8527 227 1:1663 + 1:2702 lnðMÞ M γ lnðpc Þ ¼ 0:01901 0:0048442ð198:71Þ + 0:13239ð0:8527Þ + 1:1663 + 1:2702 ln ð198:71Þ ¼ 5:6656 0:8527 pc ¼ exp ð5:6656Þ ¼ 289 psi 227 198:71 Group 3. The Plus Fraction Characterization Methods 121 4:559 + 0:21699 ln ðMÞ M 4:559 ω ¼ 0:64235 + 0:0014667ð198:71Þ + 0:021876ð0:8527Þ 198:71 + 0:21699 ln ð198:71Þ ¼ 0:531 ω ¼ 0:64235 + 0:0014667M + 0:21876γ (e) Eqs. (2.3A) through (2.3E) 0:351579776 M M Tc ¼ 231:9733906 0:352767639 233:3891996 ¼ 1259°R γ γ 0:885326626 M M + 49:62573013 ¼ 280 psi 0:106467189 γ γ M M ω ¼ 0:279354619 ln + 0:00141973 1:243207091 ¼ 0:61 γ γ pc ¼ 31829 0:984070268 M M + 87:7041 106 vc ¼ 0:054703719 0:001007476 γ γ ¼ 11:717 ft3 =lbm-mol ¼ 0:0589 ft3 =lbm 0:664115311 M M 379:4428973 Tb ¼ 33:72211 1:40066452 ¼ 931°R γ γ ðM=γ Þ The following table summarizes the results for this example. Method Tc (°R) pc (psia) Tb (°R) ω Rowe’s Standing’s Riazi-Daubert Magoulas-Tassios Eqs. (2.3A)–(2.3E) 1247 1247 1294 1284 1259 277 291 264 289 280 941 – 950 – 931 0.612 – 0.534 0.531 0.61 Recommended Plus Fraction Characterizations As discussed in Chapter 5, numerous forms of cubic equations of state that are widely used in the petroleum industry (eg, Peng-Robinson EOS) require the critical properties and acentric factor for each of the defined and undefined petroleum fractions in the hydrocarbon system. The characterizations of the undefined petroleum fractions rely on applying generalized correlations with their parameters, which were developed by regression to match critical and physical properties of pure components. Selected generalized correlations, often with significantly diverging results among themselves, are used in equations of state applications that 122 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions commonly produce erroneous EOS results. As a result, equations of state are not predictive and require adjusting or “tuning” the critical properties of the undefined petroleum fractions by matching the available laboratory PVT data. Recognizing the fact that tuning EOS requires the adjustment of the critical properties of undefined hydrocarbon groups, the need is to select a generalized property correlation (or correlations) that is capable of producing a logical and consistent properties trend of the hydrocarbon groups (eg, Tc and ω are increasing and pc is decreasing with increasing molecular weight of the hydrocarbon group). The Riazi-Daubert critical properties and Kesler-Lee acentric factor generalized correlations provide the suggested required properties trend and, therefore, are recommended for use in EOS applications. As of the determination of the exact composition and critical characteristics of the heavy-end plus fraction (eg, C7+) is a nearly impossible task, a useful type of compositional analysis is to determine the relative amounts of paraffins “P,” naphthenes “N,” and aromatics “A” for pseudofractions distilled by the TBP analysis. Detailed PNA analysis could provide accurate estimation of the physical and critical properties of these petroleum fractions. The crude oil assay testing could include such measurements; however, analysis could be costly and time consuming. This author agrees with Whitson and Brule (2000) that PNA determination has a limited usefulness for improving equations of state fluid characterizations. However, detail of the PNA determination methodologies is summarized below. PNA Determination The vast number of hydrocarbon compounds making up naturally occurring crude oil has been grouped chemically into several series of compounds. Each series consists of those compounds similar in their molecular makeup and characteristics. Within a given series, the compounds range from extremely light, or chemically simple, to heavy, or chemically complex. In general, it is assumed that the heavy (undefined) hydrocarbon fractions are composed of three hydrocarbon groups: n n n paraffins (P) naphthenes (N) aromatics (A) The PNA content of the plus fraction (the undefined hydrocarbon fraction) can be estimated experimentally from distillation or a chromatographic analysis. Both types of analysis provide information valuable for use in characterizing the plus fractions. Bergman et al. (1977) outlined the chromatographic analysis procedure by which distillation cuts are characterized by the density Group 3. The Plus Fraction Characterization Methods 123 and molecular weight as well as by weight average boiling point (WABP). It should be pointed out that when a single boiling point is given for a plus fraction, it is given as its volume average boiling point (VABP). Generally, five methods are used to define the normal boiling point for the plus fraction (eg, C7+) these are 1. VABP — This is defined mathematically by the following expression: X Vi Tbi VABP ¼ (2.36) i where Tbi ¼ boiling point of the distillation cut i, °R Vi ¼ volume fraction of the distillation cut i 2. WABP — This property is defined by the following expression: WABP ¼ X wi Tbi (2.37) i where wi ¼ weight fraction of the distillation cut i 3. Molar average boiling point (MABP) — The MABP is given by the following relationship: MABP ¼ X xi Tbi (2.38) i where xi ¼ mole fraction of the distillation cut i 4. Cubic average boiling point (CABP) — The relationship is defined as " #3 X 1=3 CABP ¼ xi Tbi (2.39) i 5. Mean average boiling point (MeABP) — This is defined by MeABP ¼ MABP + CABP 2 (2.40) The MeABP (Tb) is recommended as the most representative boiling point of the plus fraction because it allows better reproduction of the properties of the entire mixture. As indicated by Edmister and Lee (1984), the above five expressions for calculating normal boiling points result in values that do not differ significantly from one another for narrow boiling petroleum fractions. The three parameters that are commonly employed to estimate the PNA content of the heavy hydrocarbon fraction include: 124 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions n n n molecular weight specific gravity VABP or WABP Hopke and Lin (1974), Erbar (1977), Bergman et al. (1977), and Robinson and Peng (1978) used the above parameters and PNA concept to characterize the undefined hydrocarbon fractions in terms of their critical properties and acentric factors. Two of the most widely used PNA based characterization methods are n n Peng-Robinson Method Bergman’s Method Details of these two methods are summarized below Peng-Robinson’s Method Robinson and Peng (1978) proposed a detailed procedure for characterizing heavy hydrocarbon fractions. The procedure is summarized in the following steps. Step 1. Calculate the PNA content (XP, XN, XA) of the undefined fraction by solving the following three rigorously defined equations: X Xi ¼ 1 (2.41) ½Mi Tbi Xi ¼ ðMÞðWABPÞ (2.42) i¼P, N, A X i¼P, N, A X ½Mi Xi ¼ M (2.43) i¼P, N, A where XP ¼ mole fraction of the paraffinic group in the undefined fraction XN ¼ mole fraction of the naphthenic group in the undefined fraction XA ¼ mole fraction of the aromatic group in the undefined fraction WABP ¼ weight average boiling point of the undefined fraction, °R M ¼ molecular weight of the undefined fraction (Mi) ¼ average molecular weight of each cut (ie, PNA) (Tb)i ¼ boiling point of each cut, °R Eqs. (2.41) through (2.43) can be written in a matrix form as follows: 2 1 1 1 3 2 XP 3 2 1 3 7 7 6 7 6 6 6 ½MTb P ½MTb N ½MTb A 7 6 XN 7 ¼ 6 MðWABPÞ 7 5 5 4 5 4 4 ½M P ½M N ½M A M XA (2.44) Group 3. The Plus Fraction Characterization Methods 125 Robinson and Peng pointed out that it is possible to obtain negative values for the PNA contents. To prevent these negative values, the authors imposed the following constraints: 0 XP 0:90 XN 0:00 XA 0:00 Solving Eq. (2.44) for the PNA content requires the WABP and molecular weight of the cut of the undefined hydrocarbon fraction. Robinson and Peng proposed the following set of working expressions to estimate the boiling point temperature (Tb)P, (Tb)N, and (Tb)A and the molecular waiting (M)P, (M)N, and (M)A: Paraffinic group : ln ðTb ÞP ¼ ln ð1:8Þ + 6 h i X ai ðn 6Þi1 (2.45) i¼1 Naphthenic group : ln ðTb ÞN ¼ ln ð1:8Þ + 6 h i X ai ðn 6Þi1 (2.46) i¼1 Aromatic group : ln ðTb ÞA ¼ ln ð1:8Þ + 6 h i X ai ðn 6Þi1 (2.47) i¼1 Paraffinic group : ðMÞP ¼ 14:026n + 2:016 (2.48) Naphthenic group : ðMÞN ¼ 14:026n 14:026 (2.49) Aromatic group : ðMÞA ¼ 14:026n 20:074 (2.50) where n ¼ number of carbon atoms in the undefined hydrocarbon fraction, eg, 7, 8, …, etc. ai ¼ coefficients of the equations, which are given below Coefficient Paraffin, P Napthene, N Aromatic, A a1 a2 a3 a4 a5 a6 5.83451830 5.85793320 5.86717600 0.84909035 101 0.79805995 101 0.80436947 101 0.52635428 102 0.43098101 102 0.47136506 102 0.21252908 103 0.14783123 103 0.18233365 103 0.44933363 105 0.27095216 105 0.38327239 105 0.37285365 107 0.19907794 107 0.32550576 107 126 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions Step 2. Having obtained the PNA content of the undefined hydrocarbon fraction, as outlined in step 1, calculate the critical pressure of the fraction by applying the following expression: pc ¼ XP ðpc ÞP + XN ðpc ÞN + XA ðpc ÞA (2.51) where pc ¼ critical pressure of the heavy hydrocarbon fraction, psia. The critical pressure for each cut of the heavy fraction is calculated according to the following equations: Paraffinic group : ðpc ÞP ¼ Naphthenic group : ðpc ÞN ¼ 206:126096n + 29:67136 ð0:227n + 0:340Þ2 206:126096n 206:126096 Aromatic group : ðpc ÞA ¼ ð0:227n 0:137Þ2 206:126096n 295:007504 ð0:227n 0:325Þ2 (2.52) (2.53) (2.54) Step 3. Calculate the acentric factor of each cut of the undefined fraction by using the following expressions: Paraffinic group : ðωÞP ¼ 0:432n + 0:0457 (2.55) Naphthenic group : ðωÞN ¼ 0:0432n 0:0880 (2.56) Aromatic group : ðωÞA ¼ 0:0445n 0:0995 (2.57) Step 4. Calculate the critical temperature of the fraction under consideration by using the following relationship: Tc ¼ XP ðTc ÞP + XN ðTc ÞN + XA ðTc ÞA (2.58) where Tc ¼ critical temperature of the fraction, °R. The critical temperatures of the various cuts of the undefined fractions are calculated from the following expressions: ( ) 3 log ðPc ÞP 3:501952 Paraffinic group : ðTc ÞP ¼ S 1 + ð T b ÞP 7 1 + ðωÞP ( Naphthenic group : ðTc ÞN ¼ S1 ( Aromatic group : ðTc ÞA ¼ S1 (2.59) ) 3 log ðPc ÞN 3:501952 1+ ðTb ÞN (2.60) 7 1 + ðωÞP ) 3 log ðPc ÞA 3:501952 1+ ðTb ÞA 7 1 + ðωÞA (2.61) Group 3. The Plus Fraction Characterization Methods 127 where the correction factors S and S1 are defined by the following expressions: S ¼ 0:99670400 + 0:00043155n S1 ¼ 0:99627245 + 0:00043155n Step 5. Calculate the acentric factor of the heavy hydrocarbon fraction (residue) by using the Edmister’s correlation (Eq. (2.21)) to give ω¼ 3½ log ðPc =14:7Þ 1 7½ðTc =Tb Þ 1 (2.62) where ω ¼ acentric factor of the residue heavy fraction Pc ¼ critical pressure of the heavy fraction, psia Tc ¼ critical temperature of the heavy fraction, °R Tb ¼ average weight boiling point, °R EXAMPLE 2.5 Using the Peng and Robinson approach, calculate the critical pressure, critical temperature, and acentric factor of an undefined hydrocarbon fraction that is characterized by the laboratory measurements: n n n molecular weight ¼ 94 specific gravity ¼ 0.726133 WABP ¼ 195°F, ie, 655°R Solution Step 1. Calculate the boiling point of each cut by applying Eqs. (2.45) through (2.47) to give ( ) 6 h i X i1 ¼ 666:58°R ðTb ÞP ¼ exp ln ð1:8Þ + ai ðn 6Þ ) 6 h i X i1 ¼ 630°R ln ð1:8Þ + ai ðn 6Þ ( ) 6 h i X ln ð1:8Þ + ai ðn 6Þi1 ¼ 635:85°R ðTb ÞN ¼ exp ðTb ÞA ¼ exp i¼1 ( i¼1 i¼1 Step 2. Compute the molecular weight of various cuts by using Eqs. (2.48) through (2.50) to give Paraffinic group : ðMÞP ¼ 14:026n + 2:016 ðMÞP ¼ 14:026 7 + 2:016 ¼ 100:198 Naphthenic group : ðMÞN ¼ 14:026n 14:026 ðMÞN ¼ 14:0129 7 14:026 ¼ 84:156 128 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions Aromatic group : ðMÞA ¼ 14:026n 20:074 ðMÞA ¼ 14:026 7 20:074 ¼ 78:180 Step 3. Solve Eq. (2.44) for XP, XN, and XA: 3 3 2 3 2 2 1 1 1 XP 1 7 7 6 7 6 6 6 ½MTb P ½MTb N ½MTb A 7 6 XN 7 ¼ 6 MðWABPÞ 7 5 5 4 5 4 4 ½M P ½M N ½M A M XA 2 1 1 1 3 2 XP 3 2 1 3 7 6 7 7 6 6 6 66689:78 53018:28 49710:753 7 6 XN 7 ¼ 6 61570 7 5 4 5 5 4 4 199:198 84:156 78:180 94 XA to give XP ¼ 0:6313 XN ¼ 0:3262 XA ¼ 0:04250 Step 4. Calculate the critical pressure of each cut in the undefined fraction by applying Eqs. (2.52) and (2.54): ðpc ÞP ¼ ðpc ÞN ¼ ðpc ÞA ¼ 206:126096 7 + 29:67136 ð0:227 7 + 0:340Þ2 ¼ 395:70 psia 206:126096 7 206:126096 ð0:227 7 0:137Þ2 206:126096 7 295:007504 ð0:227 7 0:325Þ2 ¼ 586:61 ¼ 718:46 Step 5. Calculate the critical pressure of the heavy fraction from Eq. (2.51) to give pc ¼ XP ðpc ÞP + XN ðpc ÞN + XA ðpc ÞA pc ¼ 0:6313ð395:70Þ + 0:3262ð586:61Þ + 0:0425ð718:46Þ ¼ 471 psia Step 6. Compute the acentric factor for each cut in the fraction by using Eqs. (2.55) through (2.57) to yield ðωÞP ¼ 0:432 7 + 0:0457 ¼ 0:3481 ðωÞN ¼ 0:0432 7 0:0880 ¼ 0:2144 ðωÞA ¼ 0:0445 7 0:0995 ¼ 0:2120 Group 3. The Plus Fraction Characterization Methods 129 Step 7. Solve for (Tc)P, (Tc)N, and (Tc)A by using Eqs. (2.59) through (2.61) to give S ¼ 0:99670400 + 0:00043155 7 ¼ 0:99972 S1 ¼ 0:99627245 + 0:00043155 7 ¼ 0:99929 3 log ½395:7 3:501952 ðTc ÞP ¼ 0:99972 1 + 666:58 ¼ 969:4°R 7½1 + 0:3481 3 log ½586:61 3:501952 ðTc ÞN ¼ 0:99929 1 + 630 ¼ 974:3°R 7½1 + 0:2144 3 log ½718:46 3:501952 ðTc ÞA ¼ 0:99929 1 + 635:85 ¼ 1014:9°R 7½1 + 0:212 Step 8. Solve for (Tc) of the undefined fraction from Eq. (2.58): Tc ¼ XP ðTc ÞP + XN ðTc ÞN + XA ðTc ÞA ¼ 964:1°R Step 9. Calculate the acentric factor from Eq. (2.62) to give ω¼ 3½ log ð471:7=14:7Þ 1 ¼ 0:3680 7½ð964:1=655Þ 1 Bergman’s Method Bergman et al. (1977) proposed a detailed procedure for characterizing the undefined hydrocarbon fractions based on calculating the PNA content of the fraction under consideration. The proposed procedure originated from analyzing extensive experimental data on lean gases and condensate systems. In developing the correlation, the authors assumed that the paraffinic, naphthenic, and aromatic groups have the same boiling point. The computational procedure is summarized in the following steps. Step 1. Estimate the weight fraction of the aromatic content in the undefined fraction by applying the following expression: wA ¼ 8:47 0:7Kw (2.63) where wA ¼ weight fraction of aromatics Kw ¼ Watson characterization factor, defined mathematically by the following expression Kw ¼ ðTb Þ1=3 =γ γ ¼ specific gravity of the undefined fraction Tb ¼ WABP, °R (2.64) 130 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions Bergman et al. imposed the following constraint on the aromatic content: 0:03 wA 0:35 Step 2. Bergman proposed the following set of expressions for estimating specific gravities of the three groups, ie, γ P, γ N, and γ A γ P ¼ 0:583286 + 0:00069481ðTb 460Þ 0:7572818 106 ðTb 460Þ2 + 0:3207736 109 ðTb 460Þ3 (2.65) γ N ¼ 0:694208 + 0:0004909267ðTb 460Þ 0:659746 106 ðTb 460Þ2 + 0:330966 109 ðTb 460Þ3 (2.66) γ A ¼ 0:916103 0:000250418ðTb 460Þ + 0:357967 106 ðTb 460Þ2 9 0:166318 10 ðTb 460Þ3 ð2:67Þ Bergman suggested a minimum paraffin content of 0.20 was set by Bergman et al. To ensure that this minimum value is met, the estimated aromatic content that results in negative values of wP is increased in increments of 0.03 up to a maximum of 15 times until the paraffin content exceeds 0.20. They pointed out that this procedure gives reasonable results for fractions up to C15. Step 3. With the estimate of the aromatic content as outlined in step 1 and specific gravity of the three PNA groups as calculated in step 2, determine the weight fractions of the paraffinic and naphthenic cuts by solving the following system of linear equations simultaneously: wP + wN ¼ 1 wA (2.68) wP wN 1 wA + ¼ γP γN γ γA (2.69) where wP ¼ weight fraction of the paraffin cut wN ¼ weight fraction of the naphthene cut γ ¼ specific gravity of the undefined fraction γ P, γ N, γ A ¼ specific gravity of the three groups at the WABP of the undefined fraction Combining Eq. (2.68) with (2.69) and solving for weight fraction of the aromatic group gives γ P γ N wA γ P γ N ð1 wA Þγ P γ γA wP γN γP wN ¼ 1 ðwA + wP Þ Group 3. The Plus Fraction Characterization Methods 131 Step 4. Calculate the critical temperature, the critical pressure, and acentric factor of each cut from the following expressions. For paraffins, ðTc ÞP ¼ 735:23 + 1:2061ðTb 460Þ 0:00032984ðTb 460Þ2 (2.70) ðpc ÞP ¼ 573:011 1:13707ðTb 460Þ + 0:00131625ðTb 460Þ2 0:85103 106 ðTb 460Þ3 (2.71) ðωÞP ¼ 0:14 + 0:0009ðTb 460Þ + 0:233 106 ðTb 460Þ2 (2.72) For naphthenes, ðTc ÞN ¼ 616:8906 + 2:6077ðTb 460Þ 0:003801ðTb 460Þ2 + 0:2544 105 ðTb 460Þ3 (2.73) ðpc ÞN ¼ 726:414 1:3275ðTb 460Þ + 0:9846 103 ðTb 460Þ2 0:45169 106 ðTb 460Þ3 (2.74) ðωÞN ¼ ðωÞP 0:075 (2.75) Bergman et al. assigned the following special values of the acentric factor to the C8, C9, and C10 naphthenes: C8 ðωÞN ¼ 0:26 C9 ðωÞN ¼ 0:27 C10 ðωÞN ¼ 0:35 For the aromatics, ðTc ÞA ¼ 749:535 + 1:7017ðTb 460Þ 0:0015843ðTb 460Þ2 + 0:82358 106 ðTb 460Þ3 (2.76) ðpc ÞA ¼ 1184:514 3:44681ðTb 460Þ + 0:0045312ðTb 460Þ2 0:23416 105 ðTb 460Þ3 (2.77) ðωÞA ¼ ðωÞP 0:1 (2.78) Step 5. Calculate the critical pressure, the critical temperature, and acentric factor of the undefined fraction from the following relationships: pc ¼ wP ðpc ÞP + wN ðpc ÞN + wA ðpc ÞA (2.79) Tc ¼ wP ðTc ÞP + wN ðTc ÞN + wA ðTc ÞA (2.80) ω ¼ wP ðωÞP + wN ðωÞN + wA ðωÞA (2.81) Whitson (1984) suggested that the Peng-Robinson and Bergman PNA methods are not recommended for characterizing reservoir fluids containing fractions heavier than C20. 132 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions EXAMPLE 2.6 Using Bergman’s approach, calculate the critical pressure, critical temperature, and acentric factor of an undefined hydrocarbon fraction that is characterized by the laboratory measurements: n n n molecular weight ¼ 94 specific gravity ¼ 0.726133 WABP ¼ 195°F, ie, 655°R Solution Step 1. Calculate Watson characterization factor: Kw ¼ ðTb Þ1=3 ð655Þ1=3 ¼ ¼ 11:96 γ 0:7261 Step 2. Estimate the weight fraction of the aromatic content from wA ¼ 0:847 0:847 ¼ ¼ 0:070819 Kw 11:96 Step 3. Estimate specific gravities of the three groups, ie, γ P, γ N, and γ A, by applying Eqs. (2.65)–(2.67): γ P ¼ 0:583286 + 0:00069481ð195Þ 0:7572818 106 ð195Þ2 + 0:3207736 109 ð195Þ3 ¼ 0:692422 γ N ¼ 0:694208 + 0:0004909267ð195Þ 0:659746 106 ð195Þ2 + 0:330966 109 ð195Þ3 ¼ 0:767306 γ A ¼ 0:916103 0:000250418ð195Þ + 0:357967 106 ð195Þ2 0:166318 109 ð195Þ3 ¼ 0:879652 Step 4. Estimate the weight fraction of the paraffinic and naphthenic contents from γ P γ N wA γ P γ N ð1 wA Þγ P γ γA wP ¼ 0:607933 γN γP wN ¼ 1 ðwA + wP Þ ¼ 1 ð0:079819 + 0:607933Þ ¼ 0:321248 Step 5. Calculate the critical temperature, the critical pressure, and acentric factor of each cut from the following expressions: For paraffins, ðTc ÞP ¼ 735:23 + 1:2061ð195Þ 0:00032984ð195Þ2 ¼ 957:9°R ðpc ÞP ¼ 573:011 1:13707ð159Þ + 0:00131625ð195Þ2 0:85103 106 ð195Þ3 ¼ 395 psi ðωÞP ¼ 0:14 + 0:0009ð195Þ + 0:233 106 ð195Þ2 ¼ 0:3244 Group 3. The Plus Fraction Characterization Methods 133 For naphthenes, ðTc ÞN ¼ 616:8906 + 2:6077ð195Þ 0:003801ð195Þ2 + 0:2544 105 ð195Þ3 ¼ 999:8°R ðpc ÞN ¼ 726:414 1:3275ð195Þ + 0:9846 103 ð195Þ2 0:45169 106 ð195Þ3 ¼ 501:6 psi ðωÞN ¼ ðωÞP 0:075 ¼ 0:2494 For the aromatics, ðTc ÞA ¼ 749:535 + 1:7017ð195Þ 0:0015843ð195Þ2 + 0:82358 106 ð195Þ3 ¼ 1027°R ðpc ÞA ¼ 1184:514 3:44681ð195Þ + 0:0045312ð195Þ2 0:23416 105 ð195Þ3 ¼ 667 psi ðωÞA ¼ ðωÞP 0:1 ¼ 0:2244 Step 6. Calculate the critical pressure, the critical temperature, and acentric factor of the undefined fraction from the following relationships: pc ¼ wP ðpc ÞP + wN ðpc ÞN + wA ðpc ÞA ¼ 449 psi Tc ¼ wP ðTc ÞP + wN ðTc ÞN + wA ðTc ÞA ¼ 976°R ω ¼ wP ðωÞP + wN ðωÞN + wA ðωÞA ¼ 0:2932 Graphical Correlations Several mathematical correlations for determining the physical and critical properties of petroleum fractions have been presented. These correlations are readily adapted to computer applications. However, it is important to present the properties in graphical forms for a better understanding of the behavior and interrelationships of the properties. Boiling Points Numerous graphical correlations have been proposed over the years for determining the physical and critical properties of petroleum fractions. Most of these correlations use the normal boiling point as one of the correlation parameters. As stated previously, five methods are used to define the normal boiling point: 134 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions 1. VABP 2. WABP 3. MABP 4. CABP 5. MeABP. Figure 2.91 schematically illustrates the conversions between the VABP and the other four averaging types of boiling point temperatures. The following steps summarize the procedure of using Fig. 2.9 in determining the desired average boiling point temperature. Step 1. On the basis of ASTM D-86 distillation data, calculate the volumetric average boiling point from the following expressions: VABP ¼ ðt10 + t30 + t50 + t70 + t90 Þ=5 (2.82) where t is the temperature in °F and the subscripts 10, 30, 50, 70, and 90 refer to the volume percent recovered during the distillation. Step 2. Calculate the 10–90% “slope” of the ASTM distillation curve from the following expression: Slope ¼ ðt90 t10 Þ=80 (2.83) Enter the value of the slope in the graph and travel vertically to the appropriate set for the type of boiling point desired. Correction to be added to VABP to obtain other boiling points Slope = 2 1 3 2 t90 – t10 80 4 5 6 7 8 9 0 -2 -4 -6 -8 n FIGURE 2.9 Correction to volumetric average boiling points. 1 The original conversion chart can be found in the API Technical Data Book. Group 3. The Plus Fraction Characterization Methods 135 Step 3. Read from the ordinate a correction factor for the VABP and apply the relationship: Desired boiling point ¼ VABP + correction factor (2.84) The use of the graph can best be illustrated by the following examples. EXAMPLE 2.7 The ASTM distillation data for a 55° API gravity petroleum fraction is given below. Calculate WAPB, MABP, CABP, and MeABP. Cut Distillation % Over Temperature (°F) 1 2 3 4 5 6 7 8 9 10 Residue IBP 10 20 30 40 50 60 70 80 90 EP 159 178 193 209 227 253 282 318 364 410 475 IBP, initial boiling point; EP, end point. Solution Step 1. Calculate VABP from Eq. (2.82): VABP ¼ ð178 + 209 + 253 + 318 + 410Þ=5 ¼ 273°F Step 2. Calculate the distillation curve slope from Eq. (2.83): Slope ¼ ð410 178Þ=80 ¼ 2:9 Step 3. Enter the slope value of 2.9 in Fig. 2.9 and move down to the appropriate set of boiling point curves. Read the corresponding correction factors from the ordinate to give Correction factors for WABP ¼ 6°F Correction factors for CABP ¼ 7°F Correction factors for MeABP ¼ 18°F Correction factors for MABP ¼ 33°F 136 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions Step 4. Calculate the desired boiling point by applying Eq. (2.84): WABP ¼ 273 + 6 ¼ 279°F CABP ¼ 273 7 ¼ 266°F MeABP ¼ 273 18 ¼ 255°F MABP ¼ 273 33 ¼ 240°F Molecular Weight Fig. 2.10 shows a convenient graphical correlation for determining the molecular weight of petroleum fractions from their MeABP and API gravities. The following example illustrates the practical application of the graphical method. n FIGURE 2.10 Relationship between molecular weight, API gravity, and MeABPs. (Courtesy of the Gas Processors Suppliers Association. Published in the GPSA Engineering Data Book, tenth edition, 1987.) Group 3. The Plus Fraction Characterization Methods 137 EXAMPLE 2.8 Calculate the molecular weight of the petroleum fraction with an API gravity and MeABP as given in Example 2.7, ie, API ¼ 55° MeABP ¼ 255°F Solution Enter these values in Fig. 2.10 to give MW ¼ 118. Critical Temperature The critical temperature of a petroleum fraction can be determined by using the graphical correlation shown in Fig. 2.11. The required correlation parameters are the API gravity and the MABP of the undefined fraction. EXAMPLE 2.9 Calculate the critical temperature of the petroleum fraction with physical properties as given in Example 2.7. Solution From Example 2.7, API ¼ 55° MABP ¼ 240°F Enter the above values in Fig. 2.6 to give Tc ¼ 600°F. Critical Pressure Fig. 2.12 is a graphical correlation of the critical pressure of the undefined petroleum fractions as a function of the MeABP and the API gravity. The following example shows the practical use of the graphical correlation. EXAMPLE 2.10 Calculate the critical pressure of the petroleum fraction from Example 2.7. Solution From Example 2.7, API ¼ 55° MeABP ¼ 255°F Determine the critical pressure of the fraction from Fig. 2.12, to give pc ¼ 428 psia. 138 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions n FIGURE 2.11 Critical temperature as a function of API gravity and boiling points. (Courtesy of the Gas Processors Suppliers Association. Published in the GPSA Engineering Data Book, tenth edition, 1987.) Group 3. The Plus Fraction Characterization Methods 139 n FIGURE 2.12 Relationship between critical pressure, API gravity, and MeABPs. (Courtesy of the Gas Processors Suppliers Association. Published in the GPSA Engineering Data Book, tenth edition, 1987.) 140 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions SPLITTING AND LUMPING SCHEMES All forms of equations of state and their modifications require the critical properties and acentric factor of each component in the hydrocarbon mixture. For pure components, these required properties are well defined; however, most hydrocarbon fluids contain hundreds of different components that are difficult to identify and characterize by laboratory separation techniques. These large number of components heavier than C6 are traditionally lumped together and categorized as the “plus fraction”; eg, heptanes-plus “C7+” or undecanes-plus “C11+.” A significant portion of naturally occurring hydrocarbon fluids contain these hydrocarbon-plus fractions, which could create major problems when applying equations of state to model the volumetric behavior and predict the thermodynamic properties of hydrocarbon fluids. These problems result from the difficulty of properly characterizing the plus fractions (heavy ends) in terms of their critical properties and acentric factors. As presented earlier in this chapter, many correlations were developed to generate the critical properties and acentric factor for the plus fraction based on the molecular weight and specific gravity. However, routine laboratory experiments on the plus fraction (eg, C7+) provide description of the plus fraction in terms of its molecular weight and specific gravity, noting that the measured molecular weight of the plus fraction can have an error of as much as 20%. In the absence of detailed analytical TBP distillation or chromatographic analysis data for the plus fraction in a hydrocarbon mixture, erroneous predictions and conclusions can result if the plus fraction is categorized as one component and directly used in EOS phase equilibria calculations. As an example, errors could occur when calculating the saturation pressure for a rich gas condensate sample. The EOS calculation will, at times, predict a bubble point pressure instead of a dew point pressure. These problems associated with treating the plus fraction as a single component can be substantially reduced if the plus fraction is split into a manageable number of pseudofractions. Splitting refers to the process of breaking down the plus fraction into a certain number; an optimum number of SCN fractions. Generally, with a sufficiently large number of pseudocomponents used in characterizing the heavy fraction of a hydrocarbon mixture, a satisfactory prediction of the PVT behavior by the equation of state can be obtained. However, in compositional models, the cost and computing time can increase significantly with the increased number of components in the system. Therefore, strict limitations are set on the maximum number of components that can be used in compositional models and the original components have to be lumped into a smaller number of pseudocomponents for equation-of-state calculations. Splitting and Lumping Schemes 141 The characterization of the fraction (eg, C7+) generally consists of the following three steps: (a) Splitting the plus fraction into pseudocomponents (eg, C7 to C45+) (b) Lumping the generated pseudocomponents into an optimum number of SCN fractions (c) Characterizing the lumped fraction in terms of their critical properties and acentric factors The problem, then, is how to adequately split a C7+ fraction into a number of pseudocomponents characterized by mole fraction, molecular weight, and specific gravity. These characterization properties, when properly combined, should match the measured plus fraction properties; that is, (M)7+ and (γ)7+. Splitting Schemes Splitting schemes refer to the procedures of dividing the heptanes-plus fractions into hydrocarbon groups with a SCN (C7, C8, C9, etc.), described by the same physical properties used for pure components. It should be pointed out that plotting the compositional analysis of any naturally occurring hydrocarbon system will exhibit a discrete compositional distribution for all components lighter than heptanes “C7” and show continuous compositional distribution for all components heavier than hexane fraction, as shown in Fig. 2.13. Mole % 50.000 Discrete distribution Continuous distribution 5.000 0.500 C1 C2 C3 i-C4 n-C4 i-C5 n-C5 C6 C7 C8 C9 C10 C11 C12 C13 C14 C15 C16 C17 C18 C19 n FIGURE 2.13 Discrete and continuous compositional distribution. 142 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions Based on the continuous compositional distribution observation, several authors have proposed different schemes for extending the molar distribution behavior of C7+ in terms of mole fraction as a function of molecular weight or number of carbon atoms. In general, the proposed schemes are based on the observation that lighter systems, such as condensates, usually exhibit exponential molar distribution, while heavier systems often show left-skewed distributions. This behavior is shown schematically in Fig. 2.14. Three important requirements should be satisfied when applying any of the proposed splitting models: 1. The sum of the mole fractions of the individual pseudocomponents is equal to the mole fraction of C7+. 2. The sum of the products of the mole fraction and the molecular weight of the individual pseudocomponents is equal to the product of the mole fraction and molecular weight of C7+. 3. The sum of the product of the mole fraction and molecular weight divided by the specific gravity of each individual component is equal to that of C7+. These requirements can be expressed mathematically by the following relationship: n+ X z n ¼ z7 + (2.85) n¼7 Compositional distribution 6 5 Mole % 4 3 Norm Left-ske w al an d he ed distrib avy c u rude tion oil sy stem Cond Expo s n ensa te an ential dis d ligh tribu tion t hyd roca rbon syste ms 2 1 0 7 8 9 10 11 12 13 14 15 16 Number of carbon atoms n FIGURE 2.14 The exponential and left-skewed distribution functions. 17 18 19 20 21 Splitting and Lumping Schemes 143 n+ X ½zn Mn ¼ z7 + M7 + (2.86) n¼7 n+ X zn M n n¼7 γn ¼ z7 + M7 + γ7 + (2.87) Eqs. (2.86) and (2.87) can be solved for the molecular weight and specific gravity of the last fraction after splitting, as Mn + ¼ γn + ¼ z7 + M7 + Xðn + Þ1 n¼7 ½ zn M n zn + zn + M n + z7 + M7 + Xðn + Þ1 zn Mn n¼7 γ7 + γn where z7+ ¼ mole fraction of C7+ n ¼ number of carbon atoms n+ ¼ last hydrocarbon group in the C7+ with n carbon atoms, such as 20 + zn ¼ mole fraction of pseudocomponent with n carbon atoms M7+, γ 7+ ¼ measured molecular weight and specific gravity of C7+ Mn, γ n ¼ molecular weight and specific gravity of the pseudocomponent with n carbon atoms Several splitting schemes have been proposed. These schemes, as discussed next, are used to predict the compositional distribution of the heavy plus fraction. Modified Katz’s Method Katz (1983) presented an easy-to-use graphical correlation for breaking down into pseudocomponents the C7+ fraction present in condensate systems. The method was originated by plotting the extended compositional analysis of only six condensate systems as a function of the number of carbon atoms on a semilog scale and drawing a curve that best fit the data. The limited amount of data used by Katz were not sufficient to develop a generalized and relatively accurate expression that can be used to extend the compositional analysis of the Heptanes-plus fraction from its mole fraction; ie, z7+. To improve Katz graphical correlation, more experimental 144 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions compositional analysis was added to the Katz data to develop the following expression: zn ¼ 1:269831zC7 + exp ð0:26721nÞ + 0:0060884zC7 + + 10:4275 106 (2.88) where z7+ ¼ mole fraction of C7+ in the condensate system n ¼ number of carbon atoms of the pseudocomponent zn ¼ mole fraction of the pseudocomponent with the number of carbon atoms of n Eq. (2.88) is applied repeatedly until Eq. (2.85) is satisfied. The molecular weight and specific gravity of the last pseudocomponent can be calculated from Eqs. (2.86) and (2.87), respectively. The computational procedure of using Eq. (2.88) is best explained through the following example. EXAMPLE 2.11 A naturally occurring condensate-gas system has the following composition: Component zi C1 C2 C3 i-C4 n-C4 i-C5 n-C5 C6 C7 C8 C9 C10 C11 C12 C13 C14 C15 C16 C17 C18 C19 C20+ 0.7393 0.11938 0.04618 0.0124 0.01544 0.00759 0.00703 0.00996 0.00896 0.00953 0.00593 0.00405 0.00268 0.00189 0.00168 0.00131 0.00107 0.00082 0.00073 0.00063 0.00055 0.00289 Splitting and Lumping Schemes 145 Properties of C7+ and C20+ in terms of their molecular weights and specific gravities are given below: Group Mole % Weight % MW SG Total fluid C7+ C10+ 100.000 4.271 1.830 100.000 22.690 13.346 27.74 147.39 202.31 0.412 0.789 0.832 Using the modified Katz splitting scheme as applied by Eq. (2.88), extend the compositional distribution of C7+ to C20+ and calculate M, γ, Tb, pc, Tc, and ω of C20+. Compare with laboratory extend compositional analysis Solution Applying Eq. (2.88) with z7+ ¼ 0.04271 gives the following results. zn ¼ 1:269831zC7 + exp ð0:26721nÞ + 0:0060884zC7 + + 10:4275 106 zn ¼1:269831ð0:04271Þexp ð0:26721nÞ+0:0060884ð0:04271Þ+10:4275106 Component zi Eq. (2.88) C7 C8 C9 C10 C11 C12 C13 C14 C15 C16 C17 C18 C19 C20+ 0.00896 0.00953 0.00593 0.00405 0.00268 0.00189 0.00168 0.00131 0.00107 0.00082 0.00073 0.00063 0.00055 0.00289 0.008229 0.006478 0.005104 0.004026 0.00318 0.002516 0.001995 0.001587 0.001267 0.001017 0.000822 0.000669 0.00055 0.00467a a This value was obtained by applying Eq. (2.85); ie, 0:04271 X19 z ¼ 0:00467 n¼7 n The procedure for characterizing of C20+ is summarized in the following steps: Step 1. Calculate the molecular weight and specific gravity of C20+ by solving Eqs. (2.86) and (2.87) for these properties. 146 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions n z006E Mn znMn γn znM/γ n 7 8 9 10 11 12 13 14 15 16 17 18 19 20+ P 0.00862 0.00667 0.00517 0.00402 0.00314 0.00247 0.00195 0.00156 0.00126 0.00102 0.00085 0.00071 0.00061 0.00467 96 107 121 134 147 161 175 190 206 222 237 251 263 – 0.827940631 0.713207474 0.625078138 0.538412919 0.461470288 0.397115715 0.341529967 0.295904773 0.258658497 0.227464594 0.200922779 0.17881299 0.160107188 – 5.226625952 0.727 0.749 0.768 0.782 0.793 0.804 0.815 0.826 0.836 0.843 0.851 0.856 0.861 – 1.138845434 0.952212916 0.813903826 0.688507569 0.581929746 0.493925019 0.419055174 0.358238224 0.309400117 0.269827514 0.236101973 0.208893679 0.185954922 – 6.656796111 Solving for the molecular weight of the C20+, gives X19 z7 + M7 + ½Z M n¼7 n n M20 + ¼ Z20 + M20 + ¼ ð0:04271Þð147:39Þ 5:22662 ¼ 228:67 0:00467 With a specific gravity of γ 20 + ¼ γ 20 + ¼ z20 + M20 + z7 + M7 + X19 zn Mn n¼7 γ γ7 + n ð0:00467Þð228:67Þ ¼ 0:808 ½0:04271 147:39=0:789 6:656796 Step 2. Calculate the boiling points, critical pressure, and critical temperature of C20+, using the Riazi-Daubert correlation (Eq. (2.6)), to give θ ¼ aðMÞb γ c exp ½dMeγ + f γM Tc ¼ 544:4ð228:67Þ0:2998 ð0:808Þ1:0555 exp 1:3478 104 ð228:67Þ 0:6164ð0:808Þ ¼ 1305°R pc ¼ 4:5203 104 ð228:67Þ0:8063 ð0:808Þ1:6015 exp 1:8078 103 ð228:67Þ 0:3084ð0:808Þ ¼ 207 psi Splitting and Lumping Schemes 147 Tb ¼ 6:77857ð228:67Þ0:401673 ð0:808Þ1:58262 exp 3:7709 103 ð228:67Þ 2:984036ð0:808Þ 4:25288 103 ð228:67Þð0:808Þ ¼ 1014°R Step 3. Calculate the acentric factor of C20+ by applying the Edmister correlation, to give solve for the acentric factor from Eq. (2.21) to give ω¼ ω¼ 3½ log ðpc =14:70Þ 1 7½ðTc =Tb 1Þ 3 log ð207=14:7Þ 1 ¼ 0:712 7 ð1305=1014Þ 1 To ensure continuity and consistency in characterizing the heavy end (eg, C20+) in terms of critical properties and acentric factor, the reader should consider applying Eqs. (2.3A) through (2.3E), ie, M 228:67 ¼ ¼ 283 γ 0:808 Tc ¼ 231:9733906 ¼ 1355°R 0:351579776 M M 0:352767639 233:3891996 γ γ 0:885326626 M M 0:106467189 + 49:62573013 ¼ 234:3 psi pc ¼ 31829 γ γ ω ¼ 0:279354619 ln M M + 0:00141973 1:243207091 ¼ 0:7357 γ γ Lohrenz’s Method Lohrenz et al. (1964) proposed that the heptanes-plus fraction could be divided into pseudocomponents with carbon number ranges from 7 to 40. They mathematically stated that the mole fraction zn is related to its number of carbon atoms n and the mole fraction of the Hexane fraction z6 by the expression 2 zn ¼ z6 eAðn6Þ + Bðn6Þ (2.89) 148 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions where z6 ¼ mole fraction of the Hexane component n ¼ number of carbon atoms of the pseudocomponent zn ¼ mole fraction of the pseudocomponent with number of carbon atoms of n A and B ¼ correlating parameters Eq. (2.89) assumes that individual C7+ components are distributed from the hexane mole fraction “C6” and tail off to an extremely small quantity of heavy hydrocarbons. Lohrenz’s expression can be linearized to determine the coefficient A and B; however, the method requires a partial extended analysis of the hydrocarbon system (eg, C7 to C11+) to be able to determine the values of the coefficients A and B. In addition, the validity and applicability of the equation could not be verified when testing to extend the molar distribution on several condensate samples. Pedersen’s Method Pedersen et al. (1982) proposed that for naturally occurring hydrocarbon mixtures, an exponential relationship exists between the mole fraction of a component and the corresponding carbon number. They expressed this relationship mathematically in the following exponential form: zn ¼ eðAn + BÞ (2.90) where A and B ¼ constants n ¼ number of carbon atoms, eg, 7, 8, 9, …, etc. For condensates and volatile oils, the authors suggested that A and B can be determined by a least squares fit to the molar distribution of the lighter fractions; that is, the methodology requires partial extended compositional analysis of the plus fraction. Pedersen’s expression can also be expressed in the following linear form to determine the coefficients of the equation, as lnðzn Þ ¼ An + B Eq. (2.90) can then be used to calculate the molar content of each of the heavier fractions by extrapolation. The classical constraints as given by Eqs. (2.85) through (2.87) also are imposed. Splitting and Lumping Schemes 149 EXAMPLE 2.12 Rework Example 2.11 by using the Pedersen splitting correlation. Assume that the partial molar distribution of C7+ is only available from C7 through C10. Extend the analysis to C20+ and compare with the experimental data. Component zi C1 C2 C3 i-C4 n-C4 i-C5 n-C5 C6 C7 C8 C9 C10 C11 C12 C13 C14 C15 C16 C17 C18 C19 C20+ 0.7393 0.11938 0.04618 0.0124 0.01544 0.00759 0.00703 0.00996 0.00896 0.00953 0.00593 0.00405 0.00268 0.00189 0.00168 0.00131 0.00107 0.00082 0.00073 0.00063 0.00055 0.00289 Group Mole % Weight % MW SG Total fluid C7+ C10+ 100.000 4.271 1.830 100.000 22.690 13.346 27.74 147.39 202.31 0.412 0.789 0.832 Solution Step 1. Calculate the coefficients A and B graphically by plotting ln(zn) versus “n” using the compositional analysis from C7 to C10 as shown in Fig. 2.15 to give the slope “A” and intercept “B” as A ¼ 0:252737 B ¼ 2:98744 Step 2. Predict the mole fraction of C10 through C20 by applying Eq. (2.90), as shown below. zn ¼ eðAn + BÞ 150 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions – 4.00000 – 4.20000 – 4.40000 y = −0.252737x − 2.987440 – 4.60000 Ln(zn) – 4.80000 – 5.00000 – 5.20000 – 5.40000 – 5.60000 – 5.80000 – 6.00000 6 7 8 9 10 Number of carbon atoms “n” 11 12 n FIGURE 2.15 Calculating the coefficients A and B for Example 2.12. Component zi Eq. (2.90) C7 C8 C9 C10 C11 C12 C13 C14 C15 C16 C17 C18 C19 C20+ 0.00896 0.00953 0.00593 0.00405 0.00268 0.00189 0.00168 0.00131 0.00107 0.00082 0.00073 0.00063 0.00055 0.00289 0.00896 0.00953 0.00593 0.00405 0.00268 0.00246 0.00191 0.00149 0.00116 0.00090 0.00070 0.00054 0.00042 0.00199 Ahmed’s Method Ahmed et al. (1985) devised a simplified method for splitting the C7+ fraction into pseudocomponents. The method originated from studying the molar behavior of 34 condensate and crude oil systems through detailed laboratory compositional analysis of the heavy fractions. The only required Splitting and Lumping Schemes 151 data for the proposed method are the molecular weight and the total mole fraction of the heptanes-plus fraction. The splitting scheme is based on calculating the mole fraction, zn, at a progressively higher number of carbon atoms. The extraction process continues until the sum of the mole fraction of the pseudocomponents equals the total mole fraction of the heptanes-plus (z7+). z n ¼ zn + Mðn + 1Þ + Mn + Mð n + 1 Þ + M n (2.91) where zn ¼ mole fraction of the pseudocomponent with a number of carbon atoms of n Mn ¼ molecular weight of the hydrocarbon group with n carbon atoms Mn+ ¼ molecular weight of the n+ fraction as calculated by the following expression: Mðn + 1Þ + ¼ M7 + + Sðn 6Þ (2.92) where n is the number of carbon atoms and S is the coefficient of Eq. (2.92) with the values given below. No. of Carbon Atoms Condensate Systems Crude Oil Systems n8 n>8 15.5 17.0 16.5 20.1 The stepwise calculation sequences of the proposed correlation are summarized in the following steps. Step 1. According to the type of hydrocarbon system under investigation (condensate or crude oil), select the appropriate values for the coefficients. Step 2. Knowing the molecular weight of C7+ fraction (M7+), calculate the molecular weight of the octanes-plus fraction (M8+) by applying Eq. (2.92). Step 3. Calculate the mole fraction of the heptane fraction (z7) by using Eq. (2.91). Step 4. Repeat steps 2 and 3 for each component in the system (C8, C9, etc.) until the sum of the calculated mole fractions is equal to the mole fraction of C7+ of the system. The splitting scheme is best explained through the following example. 152 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions EXAMPLE 2.13 Rework Example 2.12 using Ahmed’s splitting method. Component zi C1 C2 C3 i-C4 n-C4 i-C5 n-C5 C6 C7 C8 C9 C10 C11 C12 C13 C14 C15 C16 C17 C18 C19 C20+ 0.7393 0.11938 0.04618 0.0124 0.01544 0.00759 0.00703 0.00996 0.00896 0.00953 0.00593 0.00405 0.00268 0.00189 0.00168 0.00131 0.00107 0.00082 0.00073 0.00063 0.00055 0.00289 Group Mole % Weight % MW SG Total fluid C7+ C10+ 100.000 4.271 1.830 100.000 22.690 13.346 27.74 147.39 202.31 0.412 0.789 0.832 Solution Step 1. Extract the mole fraction of z7; ie, n ¼ 7 Mðn + 1Þ + Mn + zn ¼ zn + Mðn + 1Þ + Mn M8 + M7 + z7 ¼ z 7 + M8 + M7 n Calculate the molecular weight of C8+ by applying Eq. (2.92). n¼7 Mðn + 1Þ + ¼ M7 + + Sðn 6Þ M8 + ¼ 147:39 + 15:5ð7 6Þ ¼ 162:89 Splitting and Lumping Schemes 153 n Solve for the mole fraction of heptane (z7) by applying Eq. (2.91). Mðn + 1Þ + Mn + z n ¼ zn + Mð n + 1 Þ + M n z 7 ¼ z7 + M8 + M7 + 162:89 147:39 ¼ 0:04271 ¼ 0:0099 M8 + M7 162:89 96 Step 2. Extract the mole fraction of z8; ie, n ¼ 8 z n ¼ zn + Mðn + 1Þ + Mn + Mð n + 1 Þ + M n M9 + M8 + z8 ¼ z8 + M9 + M8 n Calculate the molecular weight of C9+ from Eq. (2.92). n¼8 Mðn + 1Þ + ¼ M7 + + Sðn 6Þ M9 + ¼ 147:39 + 15:5ð8 6Þ ¼ 178:39 n Determine the mole fraction of C8+ z8 + ¼ z7 + z7 ¼ 0:04271 0:0099 ¼ 0:0328 n Determine the mole fraction of C8 from Eq. (2.91). z8 ¼ z8 + ½ðM9 + M8 + Þ=ðM9 + M8 Þ z8 ¼ ð0:0328Þ½ð178:39 162:89Þ=ð178:39 107Þ ¼ 0:0074 Step 3. Extract the mole fraction of z9; ie, n ¼ 9 z n ¼ zn + Mðn + 1Þ + Mn + Mð n + 1 Þ + M n M10 + M9 + z9 ¼ z9 + M10 + M9 n Calculate the molecular weight of C9+ from Eq. (2.92). n¼9 Mðn + 1Þ + ¼ M7 + + Sðn 6Þ M10 + ¼ 147:39 + 17ð9 6Þ ¼ 198:39 n Determine the mole fraction of z9+ z9 + ¼ z7 + z7 z8 ¼ 0:04271 0:0099 0074 ¼ 0:0254 154 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions n Determine the mole fraction of C9 from Eq. (2.91). Mðn + 1Þ + Mn + zn ¼ zn + Mðn + 1Þ + Mn M10 + M9 + 198:29 178:39 ¼ 0:0254 ¼ 0:0065 z9 ¼ z9 + M10 + M9 198:29 121 Step 4. Repeat the extracting method outlined in the preceding steps, to give the following results. N Mn+ Mn zn 7 8 9 10 11 12 13 14 15 16 17 18 19 20+ 147.39 162.89 177.39 198.39 215.39 232.39 249.39 266.39 283.39 300.39 317.39 334.39 351.39 – 96 107 121 134 147 161 175 190 206 222 237 251 263 – 0.00990 0.00740 0.00650 0.00395 0.00298 0.00230 0.00180 0.00143 0.00116 0.00094 0.00076 0.00061 0.00048 0.00250 Step 5. The boiling point, critical properties, and the acentric factor of C20+ are determined using the appropriate methods, to give N zn Mn znMn γn znM/γ n 7 8 9 10 11 12 13 14 15 16 17 18 19 20+ P 0.00990 0.00740 0.00650 0.00395 0.00298 0.00230 0.00180 0.00143 0.00116 0.00094 0.00076 0.00061 0.00048 0.00250 96 107 121 134 147 161 175 190 206 222 237 251 263 – 0.9504 0.7918 0.7865 0.529266249 0.437822572 0.371019007 0.315026462 0.272443701 0.239056697 0.209010771 0.179601656 0.152316618 0.126282707 – 5.360546442 0.727 0.749 0.768 0.782 0.793 0.804 0.815 0.826 0.836 0.843 0.851 0.856 0.861 – 1.307290234 1.057142857 1.024088542 0.67681106 0.55210917 0.461466427 0.386535537 0.32983499 0.285952987 0.247936858 0.211047775 0.177939974 0.146669811 – 6.864826221 Splitting and Lumping Schemes 155 Solving for the molecular weight of the C20+, gives z7 + M7 + M20 + ¼ X19 n¼7 ½Zn Mn Z20 + M20 + ¼ ð0:04271Þð147:39Þ 5:36054 ¼ 374 0:00250 With a specific gravity of γ 20 + ¼ γ 20 + ¼ z20 + M20 + z7 + M7 + X19 zn Mn n¼7 γ γ7 + n ð0:0025Þð374Þ ¼ 0:839 ½0:04271x147:39=0:789 6:864826 Apply Eqs. (2.3A) through (2.3E), which gives M 228:67 ¼ ¼ 283 γ 0:808 Tc ¼ 231:9733906 0:351579776 M M 0:352767639 233:3891996 ¼1355°R γ γ 0:885326626 M M pc ¼ 31829 0:106467189 + 49:62573013 ¼ 234 psi γ γ ω ¼ 0:279354619 ln M M + 0:00141973 1:243207091 ¼ 0:736 γ γ Whitson’s Method The most widely used distribution function is the three-parameter gamma probability function as proposed by Whitson (1983). Unlike all previous splitting methods, the gamma function has the flexibility to describe a wider range of distribution by adjusting its variance, which is left as an adjustable parameter. Whitson expressed the function in the following form: pðMÞ ¼ ðM ηÞα1 exp f½M η=βg β α Γ ð ηÞ (2.93) 156 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions where Γ ¼ gamma function η ¼ the minimum molecular weight included in the distribution; eg, MC7 α and β ¼ adjustable parameters that describe the shape of the compositional distribution M ¼ molecular weight The key parameter defines the shape of the distribution “α” with its value usually ranging from 0.5 to 2.5. Fig. 2.16 illustrates Whitson’s model for several values of the parameter α: n n n n For α ¼ 1, the distribution is exponential For α < 1, the distribution model gives accelerated exponential distribution For α > 1, the distribution model gives left-skewed distributions In the application of the gamma distribution to heavy oils, bitumen, and petroleum residues, it is indicated that the upper limit for α is 25–30, which statistically approaches a log-normal distribution Whitson indicates that the parameter η can be physically interpreted as the minimum molecular weight found in the Cn+ fraction. Whitson recommended η ¼ 92 as a good approximation if the C7+ is the plus fraction, for other plus fractions (ie, Cn+) the following approximation is used: η 14n 6 (2.94) MCn + η α (2.95) and β¼ a =1 a = 10 a = 15 a = 25 zi Molecular weight n FIGURE 2.16 Gamma distributions for C7+. Splitting and Lumping Schemes 157 It should be pointed out that when the parameter α ¼ 1, the compositional distribution is exponential and probability function becomes pðMÞ ¼ exp f½M η=βg β Rearranging, gives exp ðη=βÞ pðMÞ ¼ exp ðM=βÞ β (2.96) with β ¼ MC n + η The cumulative frequency of occurrence “fi” for a SCN group, such as the C7 group, C8 group, C9 group, …, etc. having molecular weight boundaries between Mi+1 and Mi is given by fn ¼ ð Mn PðMÞdM ¼ PðMn Þ PðMn1 Þ Mn1 Integrating the above expression, gives fn ¼ exp η Mn Mn1 exp exp β β β (2.97) The mole fraction zn for each SCN group is given by multiplying the cumulative frequency of occurrence fn of the SCN times the mole fraction of the plus fraction, that is, zn ¼ zC7 + fn Whitson and Brule indicated that, when characterizing multiple samples simultaneously, the values of Mn, η, and β must be the same for all samples. Individual sample values of MC7 + and α, however, can be different. The result of this characterization is one set of molecular weights for the C7+ fractions, while each sample has different mole fractions zn (so that their average molecular weights MC7 + are honored). 158 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions EXAMPLE 2.14 Rework Example 2.13 using Whitson’s exponential splitting method Component zi C6 C7 C8 C9 C10 C11 C12 C13 C14 C15 C16 C17 C18 C19 C20+ 0.01 0.009 0.0095 0.0059 0.0041 0.0027 0.0019 0.0017 0.0013 0.0011 0.0008 0.0007 0.0006 0.0006 0.0029 Solution Step 1. Calculate η using η 86 Step 2. Calculate β using β ¼ MCn + η β ¼ 147:39 86 ¼ 61:39 Step 3. Calculate the mole fraction zn for each SCN group: η Mn Mn1 exp exp fn ¼ exp β β β 86 Mn Mn1 exp fn ¼ exp exp 61:39 61:39 61:39 zn ¼ zC7 + fn 86 Mn Mn1 exp zn ¼ zC7 + exp exp 61:39 61:39 61:39 zn ¼ ð0:17335Þ exp Mn Mn1 exp 61:39 61:39 Splitting and Lumping Schemes 159 Step 4. Calculate the mole fraction for each SCN group with values as tabulated below: N Mn zn 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 + 84 96 107 121 134 147 161 175 190 206 222 237 251 263 379 0.0078344 0.0059532 0.0061861 0.0046089 0.0037294 0.0032244 0.0025669 0.0021723 0.0018007 0.0013876 0.0010103 0.0007443 0.0005159 0.0009755 Group Mole % Weight % MW SG C7+ C10+ 4.271 1.83 22.69 13.346 147.39 202.31 0.789 0.832 Ahmed Modified Method The gas rate from unconventional reservoirs typically exhibits a left-skewed distribution; similar to what continues hydrocarbon compositional distribution. Ahmed (2014) adopted the methodology of describing flow-rate methodology to extend compositional analysis of the plus fraction. The methodology, as illustrated by the following steps, requires partial extended analysis of the plus fraction: normalize the Step 1. Given the partial extended compositional analysis “zn,”X composition by calculating and tabulating the ratio zn = zn for each component starting from the C7. Data below illustrates step 1: n zn 7 8 9 0.0583 0.0552 0.0374 X zn 0.0583 0.1135 0.1509 zn = X zn 1 0.486344 0.247846 160 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions 10 11 12 13 14 15 16 0.0338 0.0257 0.0202 0.0202 0.0165 0.0148 0.0116 0.1847 0.2104 0.2306 0.2508 0.2673 0.2821 0.2937 0.182999 0.122148 0.087598 0.080542 0.061728 0.052464 0.039496 X Step 2. Plot the ratio zn = zn versus number of carbon atoms as defined by “n 6” on a log-log scale, as shown in Fig. 2.17. Step 3. Express the best fit of the data by a straight-line fit as described by the power function form: z Xn ¼ aðn 6Þm zn (2.98) And determine the parameters “m” and “a.” The slope “m” must be treated as positive in all calculations. Step 4. Locate the component with the highest mole fraction “zN” in the observed extended molar analysis and “N” designated as its number of carbon atoms. For example, if the observed extended analysis shows that C8 has the highest mole fraction, then set N ¼ 8. 10 1 a = 1.18799 and m = 1.43978 zn / Σzn y = 1.18799x-1.43978 0.1 0.01 1 10 n–6 n FIGURE 2.17 The power function representation. 100 Splitting and Lumping Schemes 161 Step 5. Calculate zn as a function of “n” from the following expressions: zn ¼ zN aðn 6Þm eX ðN 6Þ (2.99) a h i ðn 6Þ1m ðN 6Þ1m 1m (2.100) with X¼ Step 6. Adjust the coefficients “a and m” to match the observed zn; use Microsoft Solver to optimize the two coefficients. Step 7. Using the regressed values of “a and m” from step 5, extend the analysis by assuming different values of n in the above equation and calculate the corresponding mole fraction of the component. EXAMPLE 2.15 Rework Example 2.14 using the modified Ahmed splitting method. Use only the data from C7 to C11. Component zi C7 C8 C9 C10 C11 C12 C13 C14 C15 C16 C17 C18 C19 C20+ 0.009 0.0095 0.0059 0.0041 0.0027 0.0019 0.0017 0.0013 0.0011 0.0008 0.0007 0.0006 0.0006 0.0029 Solution X Step 1. Normalize compositional analysis “zn” by calculating the ratio zn = zn for each component starting from C7 to C11. n zn X zn zn = n26 7 8 9 10 11 0.00896 0.00953 0.00593 0.00405 0.00268 0.00896 0.01849 0.02442 0.02847 0.03115 1 0.515414 0.242834 0.142255 0.086035 1 2 3 4 5 X zn 162 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions 10 y = 1.17556x-1.52272 a = 1.1755 and m = 1.52272 zn / Σzn 1 0.1 0.01 1 10 n–6 n FIGURE 2.18 Dimensionless ratio vs. number of carbon atoms. X Step 2. Plot zn = zn versus number of carbon atoms “n 6” on a log-log scale, as shown in Fig. 2.18 and fit the data with power function to determine the parameters “a” and “m,” to give a ¼ 1.175558 and m ¼ 1.52272 (notice the coefficient “m” must be positive). Step 3. Identify the component C8 as the fraction with the highest mole fraction in the partial analysis and set: zN ¼ 0:00953 and N ¼ 8 Step 4. Apply Eqs. (2.99) and (2.11) and compare with laboratory data: a h i ðn 6Þ1m ðN 6Þ1m X¼ 1m i 1:175558 h ðn 6Þ11:52272 ð8 6Þ11:52272 ¼ 1 1:52272 zn ¼ zN aðn 6Þm eX ðN 6Þ ¼ ð0:00953Þð1:175558Þðn 6Þm eX ð8 6Þ N Lab. zn a 5 1.175558 m 5 1.52272 X zn 7 8 9 10 11 0.00896 0.00953 0.00593 0.00405 0.00268 0.68354 0.00000 0.29897 0.47579 0.59575 0.01131 0.00780 0.00567 0.00437 0.00351 Splitting and Lumping Schemes 163 Step 5. Regress on the coefficients a and m to match composition from C7 through C11 to give the following optimized values: a ¼ 1:988198 and m ¼ 1:992053 N Lab. zn a 5 1.988198 m 5 1.992053 X zn 7 8 9 10 11 0.00896 0.00953 0.00593 0.00405 0.00268 0.99653 0.00000 0.33370 0.50102 0.60161 0.01399 0.00953 0.00593 0.00395 0.00280 Step 6. Extend the molar distribution to C12 and apply the equation through C19. n Lab. zn 7 8 9 10 11 12 13 14 15 16 17 18 19 20 + 0.00896 0.00953 0.00593 0.00405 0.00268 0.00189 0.00168 0.00131 0.00107 0.00082 0.00073 0.00063 0.00055 0.00289 a 5 1.9881981 m 5 1.9920531 X zn 0.66945 0.71789 0.75431 0.78271 0.80547 0.82412 0.83969 0.85289 - 0.00896 0.00953 0.00593 0.00405 0.00268 0.00208 0.00161 0.00128 0.00104 0.00086 0.00073 0.00062 0.00054 0.00281 Results of the methodology are shown graphically in Fig. 2.19. Ahmed (2014) also suggested the use of the logistic growth function (LGF), as given below, for extending the analysis of the C7+: zn ¼ zC7 + m a nm1 ½a + nm 2 Where the coefficients “a and m” are correlating, that must be adjusted to match the observed partial extended analysis. The value of the coefficient “m” will impact the shape of the compositional distribution, that is, exponential, left-skewed, …, etc. 164 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions 0.016 0.014 0.012 0.01 zi 0.008 0.006 0.004 0.002 0 6 8 10 12 14 16 18 20 n n FIGURE 2.19 Example 2.16 using LGF. Reworking Example 2.15 using LGF and matching the partial extended analysis by adjusting the coefficients “a and m”, to give a ¼ 0.2292 and m ¼ 1.9144: zn ¼ zC7 + m a nm1 ½a + nm 2 ¼ ð0:04272Þð1:9144Þð0:2292Þn1:91441 ½0:2292 + n1:9144 2 The extended composition analysis is shown in Fig. 2.20. 0.01200 0.01000 0.00800 zi 0.00600 0.00400 0.00200 0.00000 7 8 9 10 11 12 13 n n FIGURE 2.20 Mole fraction vs. number of carbon atoms. 14 15 16 17 18 19 Splitting and Lumping Schemes 165 Lumping Schemes Equations-of-state calculations frequently are burdened by the large number of components necessary to describe the hydrocarbon mixture for accurate phase-behavior modeling. Often, the problem is either lumping together the many experimentally determined fractions or modeling the hydrocarbon system when the only experimental data available for the C7+ fraction are the molecular weight and specific gravity. The term lumping or pseudoization then denotes the reduction in the number of components used in equations-of-state calculations for reservoir fluids. This reduction is accomplished by employing the concept of the pseudocomponent. The pseudocomponent denotes a group of pure components lumped together and represented by a single component with a SCN. Essentially, two main problems are associated with “regrouping” the original components into a smaller number without losing the predicting power of the equation of state: 1. How to select the groups of pure components to be represented by one pseudocomponent each. 2. What mixing rules should be used for determining the physical properties (eg, pc, Tc, M, γ, and ω) for the new lumped pseudocomponents. Several unique published techniques can be used to address these lumping problems, notably the methods proposed by n n n n n n n Lee et al. (1979) Whitson (1980) Mehra et al. (1980) Montel and Gouel (1984) Schlijper (1984) Behrens and Sandler (1986) Gonzalez et al. (1986) Several of these techniques are presented in the following discussion. Whitson’s Lumping Scheme Whitson (1980) proposed a regrouping scheme whereby the compositional distribution of the C7+ fraction is reduced to only a few multiple carbon number (MCN) groups. Whitson suggested that the number of MCN groups necessary to describe the plus fraction is given by the following empirical rule: Ng ¼ Int½1 + 3:3 log ðN nÞ (2.101) 166 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions where Ng ¼ number of MCN groups Int ¼ integer N ¼ number of carbon atoms of the last component in the hydrocarbon system n ¼ number of carbon atoms of the first component in the plus fraction; that is, n ¼ 7 for C7+ The integer function requires that the real expression evaluated inside the brackets be rounded to the nearest integer. The molecular weights separating each MCN group are calculated from the following expression: MN + I=Ng MI ¼ MC7 MC7 where (2.102) (M)N+ ¼ molecular weight of the last reported component in the extended analysis of the hydrocarbon system MC7 ¼ molecular weight of C7 I ¼ 1, 2, …, Ng For example, components with molecular weight of hydrocarbon groups MI1 to MI falling within the boundaries of these molecular weight values are included in the Ith MCN group. Example 2.16 illustrates the use of Eqs. (2.101) and (2.102). EXAMPLE 2.16 Given the following compositional analysis of the C7+ fraction in a condensate system, determine the appropriate number of pseudocomponents forming in the C7+: Component zi C7 C8 C9 C10 C11 C12 C13 C14 C15 C16+ M16+ 0.00347 0.00268 0.00207 0.001596 0.00123 0.00095 0.00073 0.000566 0.000437 0.001671 259 Splitting and Lumping Schemes 167 Solution Step 1. Determine the molecular weight of each component in the system. Component Zi Mi C7 C8 C9 C10 C11 C12 C13 C14 C15 C16+ 0.00347 0.00268 0.00207 0.001596 0.00123 0.00095 0.00073 0.000566 0.000437 0.001671 96 107 121 134 147 161 175 190 206 259 Step 2. Calculate the number of pseudocomponents from Eq. (2.102). Ng ¼ Int½1 + 3:3 log ð16 7Þ Ng ¼ Int½4:15 Ng ¼ 4 Step 3. Determine the molecular weights separating the hydrocarbon groups by applying Eq. (2.109). MI ¼ 96 259 96 I=4 MI ¼ 96½2:698I=4 where the values for each pseudocomponent, when MI ¼ 96[2.698]I/4, are Pseudocomponent 1 ¼ 123 Pseudocomponent 2 ¼ 158 Pseudocomponent 3 ¼ 202 Pseudocomponent 4 ¼ 259 These are defined as Pseudocomponent 1: The first pseudocomponent includes all components with molecular weight in the range of 96–123. This group then includes C7, C8, and C9. Pseudocomponent 2: The second pseudocomponent contains all components with a molecular weight higher than 123 to a molecular weight of 158. This group includes C10 and C11. Pseudocomponent 3: The third pseudocomponent includes components with a molecular weight higher than 158 to a molecular weight of 202. Therefore, this group includes C12, C13, and C14. Pseudocomponent 4: This pseudocomponent includes all the remaining components, that is, C15 and C16+. 168 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions These are summarized below. Group I Component zi zI 1 C7 C8 C9 C10 C11 C12 C13 C14 C15 C16+ 0.00347 0.00268 0.00207 0.001596 0.00123 0.00095 0.00073 0.000566 0.000437 0.001671 0.00822 2 3 4 0.002826 0.002246 0.002108 It is convenient at this stage to present the mixing rules that can be employed to characterize the pseudocomponent in terms of its pseudophysical and pseudocritical properties. Because the properties of the individual components can be mixed in numerous ways, each giving different properties for the pseudocomponents, the choice of a correct mixing rule is as important as the lumping scheme. Some of these mixing rules are given next. Hong’s Mixing Rules Hong (1982) concluded that the weight fraction average, wi, is the best mixing parameter in characterizing the C7+ fractions by the following mixing rules. Defining the normalized weight fraction of a component, i, within the set of the lumped fraction, that is, I 2 L, as zi M i w∗i ¼ XL ZM i2L i i The proposed mixing rules are Pseudocritical pressure PcL ¼ XL w p i2L i ci XL Pseudocritical temperature TcL ¼ w∗ T i2L i ci XL Pseudocritical volume VcL ¼ w∗ v i2L i ci XL Pseudoacentric factor ωL ¼ w∗ ω i2L i i XL Pseudomolecular weight ML ¼ v∗ M i2L i i XL XL Binary interaction coefficient kkL ¼ 1 w∗ w∗ 1 kij i2L j2L i j Splitting and Lumping Schemes 169 where wi* ¼ normalized weight fraction of component i in the lumped set kkL ¼ binary interaction coefficient between the kth component and the lumped fraction The subscript L in this relationship denotes the lumped fraction. Lee’s Mixing Rules Lee et al. (1979), in their proposed regrouping model, employed Kay’s mixing rules as the characterizing approach for determining the properties of the lumped fractions. Defining the normalized mole fraction of a component, i, within the set of the lumped fraction, that is, I 2 L, as zi z∗i ¼ XL z i2L i The following rules are proposed: ML ¼ XL z∗ M i2L i i XL γ L ¼ ML = VcL ¼ XL pcL ¼ TcL ¼ ωL ¼ i2L (2.103) z∗i Mi =γ i (2.104) z∗i Mi Vci =ML (2.105) i2L XL i2L XL i2L XL i2L z∗i pci (2.106) zi Tci (2.107) z∗i ωi (2.108) EXAMPLE 2.17 Using Lee’s mixing rules, determine the physical and critical properties of the four pseudocomponents in Example 2.16. Solution Step 1. Assign the appropriate physical and critical properties to each component, as shown below. Group Component zi 1 C7 C8 C9 0.00347 0.00268 0.00207 zI 0.00822 Mi γi a Vci a pci a Tci a ωi a 96 0.272 0.06289 453 985 0.280a 107 0.749 0.06264 419 1036 0.312 121 0.768 0.06258 383 1058 0.348 170 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions 2 3 4 C10 C11 C12 C13 C14 C15 C16+ 0.001596 0.002826 134 0.00123 147 0.00095 0.002246 161 0.00073 175 0.000566 190 0.000437 0.002108 206 0.001671 259 0.782 0.793 0.804 0.815 0.826 0.826 0.908 0.06273 0.06291 0.06306 0.06311 0.06316 0.06325 0.0638b 351 1128 325 1166 302 1203 286 1236 270 1270 255 1304 215b 1467 0.385 0.419 0.454 0.484 0.516 0.550 0.68b a From Table 2.2. From Table 2.2. b Step 2. Calculate the physical and critical properties of each group by applying Eqs. (2.103) through (2.108) to give the results below. Group zI ML γL VcL pcL TcL ωL 1 2 3 4 0.00822 0.002826 0.002246 0.002108 105.9 139.7 172.9 248 0.746 0.787 0.814 0.892 0.0627 0.0628 0.0631 0.0637 424 339.7 288 223.3 1020 1144.5 1230.6 1433 0.3076 0.4000 0.4794 0.6531 Behrens and Sandler’s Lumping Scheme Behrens and Sandler (1986) used the semicontinuous thermodynamic distribution theory to model the C7+ fraction for equation-of-state calculations. The authors suggested that the heptanes-plus fraction can be fully described with as few as two pseudocomponents. A semicontinuous fluid mixture is defined as one in which the mole fractions of some components, such as C1 through C6, have discrete values, while the concentrations of others, the unidentifiable components such as C7+, are described as a continuous distribution function, F(I). This continuous distribution function F(I) describes the heavy fractions according to the index I, chosen to be a property of individual components, such as the carbon number, boiling point, or molecular weight. For a hydrocarbon system with k discrete components, the following relationship applies: C6 X zi + z7 + ¼ 1:0 i¼1 The mole fraction of C7+ in this equation is replaced with the selected distribution function, to give ðB C6 X zi + FðIÞdI ¼ 1:0 i¼1 A (2.109) Splitting and Lumping Schemes 171 where A ¼ lower limit of integration (beginning of the continuous distribution, eg, C7) B ¼ upper limit of integration (upper cutoff of the continuous distribution, eg, C45) This molar distribution behavior is shown schematically in Fig. 2.21. The figure shows a semilog plot of the composition zi versus the carbon number n of the individual components in a hydrocarbon system. The parameter A can be determined from the plot or defaulted to C7; that is, A ¼ 7. The value of the second parameter, B, ranges from 50 to infinity; that is, 50 B 1. However, Behrens and Sandler pointed out that the exact choice of the cutoff is not critical. Selecting the index, I, of the distribution function F(I) to be the carbon number, n, Behrens and Sandler proposed the following exponential form of F(I): FðnÞ ¼ DðnÞeαn dn (2.110) with AnB in which the parameter α is given by the following function f(α): 1 ½A BeBα f ðαÞ ¼ cn + A Aα ¼0 α e eBα (2.111) Compositional distribution Discrete distribution Continuous distribution Mole fraction Beginning of the continuous distribution Upper cutoff of the continuous distribution Norm Left-skew al an e d he d distrib avy c u rude tion oil sy stem s Expo Light nential d istrib hydr ocar bon ution syste ms A 7 Number of carbon atoms n FIGURE 2.21 Schematic illustration of the semicontinuous distribution model. B 45 172 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions where cn is the average carbon number as defined by the relationship cn ¼ MC7 + + 4 14 (2.112) Eq. (2.111) can be solved for α iteratively by using the method of successive substitutions or the Newton-Raphson method, with an initial value of α as α ¼ ½1= cn A Substituting Eq. (2.112) into Eq. (2.111) yields ðB C6 X zi + DðnÞeαn dn ¼ 1:0 A i or z7 + ¼ ðB DðnÞeαn dn A By a transformation of variables and changing the range of integration from A and B to 0 and c, the equation becomes z7 + ¼ ðc DðrÞer dr (2.113) 0 where c ¼ ðB AÞα r ¼ dummy variable of integration (2.114) The authors applied the “Gaussian quadrature numerical integration method” with a two-point integration to evaluate Eq. (2.113), resulting in z7 + ¼ 2 X Dðri Þwi ¼ Dðr1 Þw1 + Dðr2 Þw2 (2.115) i¼1 where ri ¼ roots for quadrature of integrals after variable transformation and wi ¼ weighting factor of Gaussian quadrature at point i. The values of r1, r2, w1, and w2 are given in Table 2.9. The computational sequences of the proposed method are summarized in the following steps: Step 1. Find the endpoints A and B of the distribution. As the endpoints are assumed to start and end at the midpoint between the two carbon numbers, the effective endpoints become A ¼ ðstarting carbon numberÞ 0:5 (2.116) B ¼ ðending carbon numberÞ + 0:5 (2.117) Splitting and Lumping Schemes 173 Table 2.9 Behrens and Sandler Roots and Weights for Two-Point Integration c r1 r2 w1 w2 c r1 r2 w1 w2 0.30 0.40 0.50 0.60 0.70 0.80 0.90 1.00 1.10 1.20 1.30 1.40 1.50 1.60 1.70 1.80 1.90 2.00 2.10 2.20 2.30 2.40 2.50 2.60 2.70 2.80 2.90 3.00 3.10 3.20 3.30 3.40 3.50 3.60 3.70 3.80 3.90 4.00 4.10 4.20 4.30 0.0615 0.0795 0.0977 0.1155 0.1326 0.1492 0.1652 0.1808 0.1959 0.2104 0.2245 0.2381 0.2512 0.2639 0.2763 0.2881 0.2996 0.3107 0.3215 0.3318 0.3418 0.3515 0.3608 0.3699 0.3786 0.3870 0.3951 0.4029 0.4104 0.4177 0.4247 0.4315 0.4380 0.4443 0.4504 0.4562 0.4618 0.4672 0.4724 0.4775 0.4823 0.2347 0.3101 0.3857 0.4607 0.5347 0.6082 0.6807 0.7524 0.8233 0.8933 0.9625 1.0307 1.0980 1.1644 1.2299 1.2944 1.3579 1.4204 1.4819 1.5424 1.6018 1.6602 1.7175 1.7738 1.8289 1.8830 1.9360 1.9878 2.0386 2.0882 2.1367 2.1840 2.2303 2.2754 2.3193 2.3621 2.4038 2.4444 2.4838 2.5221 2.5593 0.5324 0.5353 0.5431 0.5518 0.5601 0.5685 0.5767 0.5849 0.5932 0.6011 0.6091 0.6169 0.6245 0.6321 0.6395 0.6468 0.6539 0.6610 0.6678 0.6745 0.6810 0.6874 0.6937 0.6997 0.7056 0.7114 0.7170 0.7224 0.7277 0.7328 0.7378 0.7426 0.7472 0.7517 0.7561 0.7603 0.7644 0.7683 0.7721 0.7757 0.7792 0.4676 0.4647 0.4569 0.4482 0.4399 0.4315 0.4233 0.4151 0.4068 0.3989 0.3909 0.3831 0.3755 0.3679 0.3605 0.3532 0.3461 0.3390 0.3322 0.3255 0.3190 0.3126 0.3063 0.3003 0.2944 0.2886 0.2830 0.2776 0.2723 0.2672 0.2622 0.2574 0.2528 0.2483 0.2439 0.2397 0.2356 0.2317 0.2279 0.2243 0.2208 4.40 4.50 4.60 4.70 4.80 4.90 5.00 5.10 5.20 5.30 5.40 5.50 5.60 5.70 5.80 5.90 6.00 6.20 6.40 6.60 6.80 7.00 7.20 7.40 7.70 8.10 8.50 9.00 10.00 11.00 12.00 14.00 16.00 18.00 20.00 25.00 30.00 40.00 60.00 100.00 1 0.4869 0.4914 0.4957 0.4998 0.5038 0.5076 0.5112 0.5148 0.5181 0.5214 0.5245 0.5274 0.5303 0.5330 0.5356 0.5381 0.5405 0.5450 0.5491 0.5528 0.5562 0.5593 0.5621 0.5646 0.5680 0.5717 0.5748 0.5777 0.5816 0.5836 0.5847 0.5856 0.5857 0.5858 0.5858 0.5858 0.5858 0.5858 0.5858 0.5858 0.5858 2.5954 2.6304 2.6643 2.6971 2.7289 2.7596 2.7893 2.8179 2.8456 2.8722 2.8979 2.9226 2.9464 2.9693 2.9913 3.0124 3.0327 3.0707 3.1056 3.1375 3.1686 3.1930 3.2170 3.2388 3.2674 3.2992 3.3247 3.3494 3.3811 3.3978 3.4063 3.4125 3.4139 3.4141 3.4142 3.4142 3.4142 3.4142 3.4142 3.4142 3.4142 0.7826 0.7858 0.7890 0.7920 0.7949 0.7977 0.8003 0.8029 0.8054 0.8077 0.8100 0.8121 0.8142 0.8162 0.8181 0.8199 0.8216 0.8248 0.8278 0.8305 0.8329 0.8351 0.8371 0.8389 0.8413 0.8439 0.8460 0.8480 0.8507 0.8521 0.8529 0.8534 0.8535 0.8536 0.8536 0.8536 0.8536 0.8536 0.8536 0.8536 0.8536 0.2174 0.2142 0.2110 0.2080 0.2051 0.2023 0.1997 0.1971 0.1946 0.1923 0.1900 0.1879 0.1858 0.1838 0.1819 0.1801 0.1784 0.1754 0.1722 0.1695 0.1671 0.1649 0.1629 0.1611 0.1587 0.1561 0.1540 0.1520 0.1493 0.1479 0.1471 0.1466 0.1465 0.1464 0.1464 0.1464 0.1464 0.1464 0.1464 0.1464 0.1464 174 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions Step 2. Calculate the value of the parameter α by solving Eq. (2.111) iteratively. Step 3. Determine the upper limit of integration c by applying Eq. (2.114). Step 4. Find the integration points r1 and r2 and the weighting factors w1 and w2 from Table 2.9. Step 5. Find the pseudocomponent carbon numbers, ni, and mole fractions, zi, from the following expressions: For the first pseudocomponent, r1 +A α z1 ¼ w1 z7 + n1 ¼ (2.118) For the second pseudocomponent, r2 +A α z2 ¼ w2 z7 + n2 ¼ (2.119) Step 6. Assign the physical and critical properties of the two pseudocomponents from Table 2.2. EXAMPLE 2.18 A heptanes-plus fraction in a crude oil system has a mole fraction of 0.4608 with a molecular weight of 226. Using the Behrens and Sandler lumping scheme, characterize the C7+ by two pseudocomponents and calculate their mole fractions. Solution Step 1. Assuming the starting and ending carbon numbers to be C7 and C50, calculate A and B from Eqs. (2.116) and (2.117): A ¼ ðstarting carbon numberÞ 0:5 A ¼ 7 0:5 ¼ 6:5 B ¼ ðending carbon numberÞ + 0:5 B ¼ 50 + 0:5 ¼ 50:5 Step 2. Calculate cn from Eq. (2.112): MC7 + + 4 14 226 + 4 cn ¼ ¼ 16:43 14 cn ¼ Step 3. Solve Eq. (2.111) iteratively for α, to give 1 ½A BeBα ¼0 cn + A Aα e eBa α 1 ½6:5 50:5e50:5α ¼0 16:43 + 6:5 6:5α e e50:5α α Splitting and Lumping Schemes 175 Solving this expression iteratively for α gives α ¼ 0.0938967. Step 4. Calculate the range of integration c from Eq. (2.114): c ¼ ðB AÞα c ¼ ð50:5 6:5Þ0:0938967 ¼ 4:13 Step 5. Find integration points ri and weights wi from Table 2.9: r1 ¼ 0:4741 r2 ¼ 2:4965 w1 ¼ 0:7733 w2 ¼ 0:2267 Step 6. Find the pseudocomponent carbon numbers ni and mole fractions zi by applying Eqs. (2.118) and (2.119). For the first pseudocomponent, r1 n1 ¼ + A α 0:4741 n1 ¼ + 6:5 ¼ 11:55 0:0938967 z1 ¼ w1 z7 + z1 ¼ ð0:7733Þð0:4608Þ ¼ 0:3563 For the second pseudocomponent, r2 n2 ¼ + A α 2:4965 n2 ¼ + 6:5 ¼ 33:08 0:0938967 z2 ¼ w2 z7 + z2 ¼ ð0:2267Þð0:4608Þ ¼ 0:1045 The C7+ fraction is represented then by the two pseudocomponents below. Pseudocomponent Carbon Number Mole Fraction 1 2 C11.55 C33.08 0.3563 0.1045 Step 7. Assign the physical properties of the two pseudocomponents according to their number of carbon atoms using the Katz and Firoozabadi generalized physical properties as given in Table 2.2 or by calculations from Eq. (2.6). The assigned physical properties for the two fractions are shown below. Pseudocomponent n 1 2 Tb (°R) γ 11.55 848 33.08 1341 M Tc (°R) pc (psia) ω 0.799 154 1185 0.915 426 1629 314 134 0.437 0.921 176 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions Lee’s Lumping Scheme Lee et al. (1979) devised a simple procedure for regrouping the oil fractions into pseudocomponents. Lee and coworkers employed the physical reasoning that crude oil fractions having relatively close physicochemical properties (such as molecular weight and specific gravity) can be accurately represented by a single fraction. Having observed that the closeness of these properties is reflected by the slopes of curves when the properties are plotted against the weight-averaged boiling point of each fraction, Lee et al. used the weighted sum of the slopes of these curves as a criterion for lumping the crude oil fractions. The authors proposed the following computational steps: Step 1. Plot the available physical and chemical properties of each original fraction versus its weight-averaged boiling point. Step 2. Calculate numerically the slope mij for each fraction at each WABP, where mij ¼ slope of the property curve versus boiling point i ¼ l, …, nf j ¼ l, …, np nf ¼ number of original oil fractions np ¼ number of available physicochemical properties ij as defined: Step 3. Compute the normalized absolute slope m ij ¼ m mij max i¼1, …, nf mij (2.120) i for each fraction as follows: Step 4. Compute the weighted sum of slopes M Xnp i ¼ M j¼1 np ij m (2.121) i represents the averaged change of physicochemical where M properties of the crude oil fractions along the boiling point axis. i for each fraction, group those Step 5. Judging the numerical values of M i values. fractions that have similar M Step 6. Using the mixing rules given by Eqs. (2.103) through (2.108), calculate the physical properties of pseudocomponents. EXAMPLE 2.19 The data in Table 2.10, as given by Hariu and Sage (1969), represent the average boiling point molecular weight, specific gravity, and molar distribution of 15 pseudocomponents in a crude oil system. Lump the given data into an optimum number of pseudocomponents using the Lee method. Splitting and Lumping Schemes 177 Table 2.10 Characteristics of Pseudocomponents in Example 2.18 Pseudocomponent Tb (°R) M γ zi 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 600 654 698 732 770 808 851 895 938 983 1052 1154 1257 1382 1540 95 101 108 116 126 139 154 173 191 215 248 322 415 540 700 0.680 0.710 0.732 0.750 0.767 0.781 0.793 0.800 0.826 0.836 0.850 0.883 0.910 0.940 0.975 0.0681 0.0686 0.0662 0.0631 0.0743 0.0686 0.0628 0.0564 0.0528 0.0474 0.0836 0.0669 0.0535 0.0425 0.0340 Solution Step 1. Plot the molecular weights and specific gravities versus the boiling points and calculate the slope mij for each pseudocomponent as in the following table. Pseudocomponent Tb mi1 5 @M/@Tb mi2 5 @γ/@Tb 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 600 654 698 732 770 808 851 895 938 983 1052 1154 1257 1382 1540 0.1111 0.1327 0.1923 0.2500 0.3026 0.3457 0.3908 0.4253 0.4773 0.5000 0.6257 0.8146 0.9561 1.0071 1.0127 0.00056 0.00053 0.00051 0.00049 0.00041 0.00032 0.00022 0.00038 0.00041 0.00021 0.00027 0.00029 0.00025 0.00023 0.00022 178 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions i Table 2.11 Absolute Slope mij and Weighted Slopes M Pseudocomponent i1 ¼ mij =1:0127 m i2 ¼ mij =0:00056 m i ¼ ðm i1 + m i2 Þ=2 M 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 0.1097 0.1310 0.1899 0.2469 0.2988 0.3414 0.3859 0.4200 0.4713 0.4937 0.6179 0.8044 0.9441 0.9945 1.0000 1.0000 0.9464 0.9107 0.8750 0.7321 0.5714 0.3929 0.6786 0.7321 0.3750 0.4821 0.5179 0.4464 0.4107 0.3929 0.55485 0.53870 0.55030 0.56100 0.51550 0.45640 0.38940 0.54930 0.60170 0.43440 0.55000 0.66120 0.69530 0.70260 0.69650 Step 2. Calculate the normalized absolute slope mij and the weighted sum of i i using Eqs. (2.121) and (2.122), respectively. This is shown in slopes M Table 2.11. Note that the maximum value of mi1 is 1.0127 and for mi2 is 0.00056. i , the pseudocomponents can be lumped into Examining the values of M three groups: Group 1: Combine fractions 1–5 with a total mole fraction of 0.3403. Group 2: Combine fractions 6–10 with a total mole fraction of 0.2880. Group 3: Combine fractions 11–15 with a total mole fraction of 0.2805. Step 3. Calculate the physical properties of each group. This can be achieved by computing M and γ of each group by applying Eqs. (2.103) and (2.104), respectively, followed by employing the Riazi-Daubert correlation (Eq. (2.6)) to characterize each group. Results of the calculations are shown below. Group I Mole Fraction (zi) M γ Tb Tc pc Vc 1 2 3 0.3403 0.2880 0.2805 109.4 171.0 396.5 0.7299 0.8073 0.9115 694 891 1383 1019 1224 1656 404 287 137 0.0637 0.0634 0.0656 Splitting and Lumping Schemes 179 The physical properties M, γ, and Tb reflect the chemical makeup of petroleum fractions. Some of the methods used to characterize the pseudocomponents in terms of their boiling point and specific gravity assume that a particular factor, called the characterization factor, Cf, is constant for all the fractions that constitute the C7+. Soreide (1989) developed an accurate correlation for determining the specific gravity based on the analysis of 843 TBP fractions from 68 reservoir C7+ samples and introduced the characterization factor, Cf, into the correlation, as given by the following relationship: γ i ¼ 0:2855 + Cf ðMi 66Þ0:13 (2.122) Characterization factor Cf is adjusted to satisfy the following relationship: zC M C γ C7 + exp ¼ C 7 + 7 + (2.123) N+ X zi Mi i¼C7 γi where γ C7 + exp is the measured specific gravity of the C7+. Combining Eq. (2.122) with Eq. (2.123) gives zC7 + MC7 + # γ C7 + exp ¼ C " N+ X zi Mi 0:2855 + Cf ðMi 66Þ0:13 i¼C7 Rearranging, f ðCf Þ ¼ CN + X " # zi Mi 0:13 i¼C7 0:2855 + Cf ðMi 66Þ zC MC 7+ 7+ ¼ 0 γ C7 + exp (2.124) The optimum value of the characterization factor can be determined iteratively by solving this expression for Cf. The Newton-Raphson method is an efficient numerical technique that can be used conveniently to solve the preceding nonlinear equation by employing the relationship f Cnf n+1 n Cf ¼ Cf @f Cnf =@Cnf with 2 CN + X6 @f Cnf ¼ 4 @Cnf i¼C7 3 ðMi 66Þ0:13 0:2855 + Cnf ðMi 66Þ 0:13 7 2 5 The method is based on assuming a starting value Cnf and calculating an improved value Cnf + 1 that can be used in the second iteration. The iteration process continues until the difference between the two, Cnf + 1 and Cnf , is small, say, 106. The calculated value of Cf then can be used in Eq. (2.122) to determine specific gravity of any pseudocomponent γ i. Soreide also developed the following boiling point temperature, Tb, correlation, which is a function of the molecular weight and specific gravity of the fraction 180 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions Tb ¼ 1928:3 1:695 105 Aγ 3:266 M0:03522 with A ¼ exp ½0:003462Mγ 0:004922M 4:7685γ where Tb ¼ boiling point temperature, °R M ¼ molecular weight EXAMPLE 2.20 The molar distribution of a heptanes-plus, as given by Whitson and Brule, is listed below, where MC7 + ¼ 143 and γ C7 + exp ¼ 0:795. Calculate the specific gravity of the five pseudocomponents. C7+ Fraction (i) Mole Fraction (zi) Molecular Weight (Mi) 1 2 3 4 5 2.4228 2.8921 1.2852 0.2367 0.0132 95.55 135.84 296.65 319.83 500.00 Solution Step 1. Solve Eq. (2.124) for the characterization factor, Cf, by trial and error or Newton-Raphson method, to give Cf ¼ 0:26927 Step 2. Using Eq. (2.122), calculate the specific gravity of the five, shown in the following table, where γ i ¼ 0:2855 + Cf ðMi 66Þ0:13 γ i ¼ 0:2855 + 0:26927ðMi 66Þ0:13 C7+ Fraction (i) Mole Fraction (zi) Molecular Weight (Mi) Specific gravity (γ i 5 0.2855 + 0.26927(Mi 2 66)0.13) 1 2 3 4 5 2.4228 2.8921 1.2852 0.2367 0.0132 95.55 135.84 296.65 319.83 500.00 0.7407 0.7879 0.8358 0.8796 0.9226 Splitting and Lumping Schemes 181 Characterization of Multiple Hydrocarbon Samples When characterizing multiple samples simultaneously for use in the equation of state applications, the following procedure is recommended: 1. Identify and select the main fluid sample among multiple samples from the same reservoir. In general, the sample that dominates the simulation process or has extensive laboratory data is chosen as the main sample. 2. Split the plus fractions of the main sample into several components using Whitson’s method and determine the parameters MN, η, and β*, as outlined by Eqs. (2.93) and (2.94). 3. Calculate the characterization factor, Cf, by solving Eq. (2.124) and calculate specific gravities of the C7+ fractions by using Eq. (2.122). 4. Lump the produced splitting fractions of the main sample into a number of pseudocomponents characterized by the physical properties pc, Tc, M, γ, and ω. 5. Assign the gamma function parameters (MN, η, and β) of the main sample to all the remaining samples; in addition, the characterization factor, Cf, must be the same for all mixtures. However, each sample can have a different MC7 + and α. 6. Using the same gamma function parameters as the main sample, split each of the remaining samples into a number of components. 7. For each of the remaining samples, group or lump the fractions into pseudocomponents with the condition that these pseudocomponents are lumped in a way to give similar or the same molecular weights to those of the lumped fractions of the main sample. 8. The results of this characterization procedure are n n One set of molecular weights for the C7+ fractions Each pseudocomponent has the same properties (ie, M, γ, Tb, pc, Tc, and ω) as those of the main sample but with a different mole fraction It should be pointed out that, when characterizing multiple samples, the outlier samples must be identified and treated separately. Assuming that these multiple samples are obtained from different depths and each is described by laboratory PVT measurements, the outlier samples can be identified by making the following plots: n n n n Saturation pressure versus depth C1 mole % versus depth C7+ mole % versus depth Molecular weight of C7+ versus depth Data deviating significantly from the general trend, as shown in Figs. 2.22– 2.25, should not be used in the multiple samples characterization procedure. Takahashi et al. (2002) point out that a single EOS model never 182 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions Saturation pressure X X True vertical depth X X X X X Outlier (removed) data n FIGURE 2.22 Measured saturation pressure versus depth. Mole % of methane X True vertical depth X X X X X X Outlier (removed) data n FIGURE 2.23 Measured methane content versus depth. Splitting and Lumping Schemes 183 Molecular weight of heptanes-plus fraction True vertical depth X X X X Outlier (removed) data n FIGURE 2.24 Measured C7+ molecular weight versus depth. Mole % of heptanes-plus fraction X True vertical depth X X X X n FIGURE 2.25 Measured mole % of C7+ versus depth. X Outlier (removed) data 184 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions predicts such outlier behavior using a common characterization. Also note that systematic deviation from a general trend may indicate a separate fluid system, requiring a separate equation-of-state model, while randomlike deviations from clear trends often are due to experimental data error or inconsistencies in reported compositional data as compared with the actual samples used in laboratory tests. PROBLEMS 1. A heptanes-plus fraction with a molecular weight of 198 and a specific gravity of 0.8135 presents a naturally occurring condensate system. The reported mole fraction of the C7+ is 0.1145. Predict the molar distribution of the plus fraction using (a) Katz’s correlation (b) Ahmed’s correlation Characterize the last fraction in the predicted extended analysis in terms of its physical and critical properties. 2. A naturally occurring crude oil system has a heptanes-plus fraction with the following properties: M7+ ¼ 213.0000 γ 7+ ¼ 0.8405 x7+ ¼ 0.3497 Extend the molar distribution of the plus fraction and determine the critical properties and acentric factor of the last component. 3. A crude oil system has the following composition: s xi C1 C2 C3 C4 C5 C6 C7 C8 C9 C10 C11 C12 C13+ 0.3100 0.1042 0.1187 0.0732 0.0441 0.0255 0.0571 0.0472 0.0246 0.0233 0.0212 0.0169 0.1340 The molecular weight and specific gravity of C13+ are 325 and 0.842, respectively. Calculate the appropriate number of pseudocomponents necessary to adequately represent these components using Problems 185 (a) Whitson’s lumping method (b) Behrens-Sandler’s method Characterize the resulting pseudocomponents in terms of their critical properties. 4. If a petroleum fraction has a measured molecular weight of 190 and a specific gravity of 0.8762, characterize this fraction by calculating the boiling point, critical temperature, critical pressure, and critical volume of the fraction. Use Riazi-Daubert’s correlation. 5. Calculate the acentric factor and critical compressibility factor of the component in problem 4. 6. A petroleum fraction has the following physical properties: API ¼ 50° Tb ¼ 400°F M ¼ 165 γ ¼ 0.79 Calculate pc, Tc, Vc, ω, and Zc using the following correlations: (a) Cavett (b) Kesler-Lee (c) Winn-Sim-Daubert (d) Watansiri-Owens-Starling 7. An undefined petroleum fraction with 10 carbon atoms has a measured average boiling point of 791°R and a molecular weight of 134. If the specific gravity of the fraction is 0.78, determine the critical pressure, critical temperature, and acentric factor of the fraction using (a) Robinson-Peng’s PNA method (b) Bergman PNA’s method (c) Riazi-Daubert’s method (d) Cavett’s correlation (e) Kesler-Lee’s correlation (f) Willman-Teja’s correlation 8. A heptanes-plus fraction is characterized by a molecular weight of 200 and specific gravity of 0.810. Calculate pc, Tc, Tb, and acentric factor of the plus fraction using (a) Riazi-Daubert’s method (b) Rowe’s correlation (c) Standing’s correlation 9. Using the data given in problem 8 and the boiling point as calculated by Riazi-Daubert’s correlation, determine the critical properties and acentric factor employing (a) Cavett’s correlation (b) Kesler-Lee’s correlation Compare the results with those obtained in problem 8. 186 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions REFERENCES Ahmed, T., 1985. Composition modeling of Tyler and Mission canyon formation oils with CO2 and lean gasses. Final report submitted to Montanans on a New Track for Science (MONTS), Montana National Science Foundation Grant Program. Ahmed, T., 2014. Equations of State and PVT Analysis. Course Material. Ahmed, T., Cady, G., Story, A., 1985. A generalized correlation for characterizing the hydrocarbon heavy fractions. Paper SPE 14266, In: Presented at the 60th Annual Technical Conference of the Society of Petroleum Engineers, Las Vegas, September 22–25. Austad, T., 1983. Practical aspects of characterizing petroleum fluids. In: Paper, Presented at the North Sea Condensate Reserves, London, May 24–25. Behrens, R., Sandler, S., 1986. The use of semi-continuous description to model the C7+ fraction in equation-of-state calculation. Paper SPE/DOE 14925, In: Presented at the Fifth Annual Symposium on EOR, Held in Tulsa, OK, April 20–23. Bergman, D.F., Tek, M.R., Katz, D.L., 1977. Retrograde condensation in natural gas pipelines. Project PR 2-29 of Pipelines Research Committee, AGA, January. Cavett, R.H., 1962. Physical data for distillation calculations — vapor-liquid equilibrium. In: Proceedings of the 27th Meeting, API, San Francisco, pp. 351–366. Edmister, W.C., 1958. Applied hydrocarbon thermodynamics, part 4, compressibility factors and equations of state. Petroleum Refiner. 37 (April), 173–179. Edmister, W.C., Lee, B.I., 1984. Applied Hydrocarbon Thermodynamics, vol. 1 Gulf Publishing Company, Houston, TX. Erbar, J.H., 1977. Prediction of absorber oil K-values and enthalpies. Research Report 13. GPA, Tulsa, OK. Gonzalez, E., Colonomos, P., Rusinek, I., 1986. A new approach for characterizing oil fractions and for selecting pseudo-components of hydrocarbons. Can. J. Petrol. Tech. (March–April), 78–84. Hall, K.R., Yarborough, L., 1971. New simple correlation for predicting critical volume. Chem. Eng. (November), 76. Hariu, O., Sage, R., 1969. Crude split figured by computer. Hydrocarb. Process. (April), 143–148. Haugen, O.A., Watson, K.M., Ragatz, R.A., 1959. Chemical Process Principles, second ed. Wiley, New York. p. 577. Hong, K.S., 1982. Lumped-component characterization of crude oils for compositional simulation. Paper SPE/DOE 10691, In: Presented at the 3rd Joint Symposium On EOR, Tulsa, OK, April 4–7. Hopke, S.W., Lin, C.J., 1974. Application of BWRS equation to absorber oil systems. In: Proceedings 53rd Annual Convention GPA, Denver, CO, pp. 63–71. Katz, D., 1983. Overview of phase behavior of oil and gas production. J. Petrol. Tech. (June), 1205–1214. Katz, D.L., Firoozabadi, A., 1978. Predicting phase behavior of condensate/crude-oil systems using methane interaction coefficients. J. Petrol. Tech. (November), 1649–1655. Kesler, M.G., Lee, B.I., 1976. Improve prediction of enthalpy of fractions. Hydrocarb. Process. (March), 153–158. Lee, S., et al., 1979. Experimental and theoretical studies on the fluid properties required for simulation of thermal processes. Paper SPE 8393, In: Presented at the 54th Annual Technical Conference of the Society of Petroleum Engineers, Las Vegas, September 23–26. References 187 Lohrenz, J., Bray, B.G., Clark, C.R., 1964. Calculating viscosities of reservoir fluids from their compositions. J. Petrol. Tech. (October), 1171–1176. Magoulas, S., Tassios, D., 1990. Predictions of phase behavior of HT-HP reservoir fluids. Paper SPE 37294, Society of Petroleum Engineers, Richardson, TX. Matthews, T., Roland, C., Katz, D., 1942. High pressure gas measurement. In: Proceedings of the Natural Gas Association of America (NGAA). Mehra, R.K., Heidemann, R., Aziz, K., 1980. Computation of multi-phase equilibrium for compositional simulators. Soc. Petrol. Eng. J. (February). Miquel, J., Hernandez, J., Castells, F., 1992. A new method for petroleum fractions and crude oil characterization. SPE Reserv. Eng. 265(May). Montel, F., Gouel, P., 1984. A new lumping scheme of analytical data for composition studies. Paper SPE 13119, In: Presented at the 59th Annual Society of Petroleum Engineers Technical Conference, Houston, TX, September 16–19. Mudgal, P., 2014. (Master of engineering thesis) Dalghouse University, Halifax, Nova Scotia. Nath, J., 1985. Acentric factor and the critical volumes for normal fluids. Ind. Eng. Chem. Fundam. 21 (3), 325–326. Pedersen, K., Thomassen, P., Fredenslund, A., 1982. Phase equilibria and separation process. Report SEP 8207. Institute for Kemiteknik, Denmark Tekniske Hojskole, July. Reid, R., Prausnitz, J.M., Sherwood, T., 1977. The Properties of Gases and Liquids, third ed. McGraw-Hill, New York. p. 21. Riazi, M.R., Daubert, T.E., 1980. Simplify property predictions. Hydrocarb. Process. (March), 115–116. Riazi, M.R., Daubert, T.E., 1987. Characterization parameters for petroleum fractions. Ind. Eng. Chem. Fundam. 26 (24), 755–759. Robinson, D.B., Peng, D.Y., 1978. The characterization of the heptanes and heavier fractions. Research Report 28. GPA, Tulsa, OK. Rowe, A.M., 1978. Internally consistent correlations for predicting phase compositions for use in reservoir compositional simulators. Paper SPE 7475, In: Presented at the 53rd Annual Society of Petroleum Engineers Fall Technical Conference and Exhibition. Salerno, S., et al., 1985. Prediction of vapor pressures and saturated volumes. Fluid Phase Equilib. 27, 15–34. Sancet, J., Heavy Faction Characterization. In: SPE 113025, 2007 SPE Annual Conference, November 11–14, Anaheim, CA 2007. Schlijper, A.G., 1984. Simulation of compositional process: the use of pseudo-components in equation of state calculations. Paper SPE/DOE 12633, In: Presented at the SPE/DOE Fourth Symposium on EOR, Tulsa, OK, April 15–18. Silva, M.B., Rodriguez, F., 1992. Automatic fitting of equations of state for phase behavior matching. Paper SPE 23703, Society of Petroleum Engineers, Richardson, TX. Sim, W.J., Daubert, T.E., 1980. Prediction of vapor-liquid equilibria of undefined mixtures. Ind. Eng. Chem. Process. Des. Dev. 19 (3), 380–393. Soreide, I., 1989. Improved Phase Behavior Predictions of Petroleum Reservoir Fluids From a Cubic Equation of State. (Doctor of engineering dissertation). Norwegian Institute of Technology, Trondheim. Standing, M.B., 1977. Volumetric and Phase Behavior of Oil Field Hydrocarbon Systems. Society of Petroleum Engineers, Dallas, TX. Takahashi, K., et al., 2002. Fluid characterization for gas injection study using equilibrium contact mixing. Paper SPE 78483, Society of Petroleum Engineers, Richardson, TX. 188 CHAPTER 2 Characterizing Hydrocarbon-Plus Fractions Twu, C., 1984. An internally consistent correlation for predicting the critical properties and molecular weight of petroleum and coal-tar liquids. Fluid Phase Equilib. 16, 137. Watansiri, S., Owens, V.H., Starling, K.E., 1985. Correlations for estimating critical constants, acentric factor, and dipole moment for undefined coal-fluid fractions. Ind. Eng. Chem. Process. Des. Dev. 24, 294–296. Watson, K.M., Nelson, E.F., Murphy, G.B., 1935. Characterization of petroleum fractions. Ind. Eng. Chem. 27, 1460. Whitson, C., 1980. Characterizing hydrocarbon-plus fractions. Paper EUR 183, In: Presented at the European Offshore Petroleum Conference, London, October 21–24. Whitson, C.H., 1983. Characterizing hydrocarbon-plus fractions. Soc. Petrol. Eng. J. 275 (August), 683. AIME, 275. Whitson, C.H., 1984. Effect of physical properties estimation on equation-of-state predictions. Soc. Petrol. Eng. J. (December), 685–696. Whitson, C.H., Brule, M.R., 2000. Phase Behavior. Society of Petroleum Engineers, Richardson, TX. Willman, B., Teja, A., 1987. Prediction of dew points of semicontinuous natural gas and petroleum mixtures. Ind. Eng. Chem. Res. 226 (5), 948–952. Winn, F.W., 1957. Simplified nomographic presentation, characterization of petroleum fractions. Petroleum Refiner. 36 (2), 157. Chapter 3 Natural Gas Properties LAWS THAT DESCRIBE the behavior of reservoir gases in terms of pressure, p, volume, V, and temperature, T, have been known for many years. These laws are relatively simple for a hypothetical fluid known as a perfect or ideal gas. This chapter reviews these laws and how they can be modified to describe the behavior of reservoir gases, which may deviate significantly from these simple laws. Gas is defined as a homogeneous fluid of low viscosity and density, which has no definite volume but expands to completely fill the vessel in which it is placed. Generally, the natural gas is a mixture of hydrocarbon and nonhydrocarbon gases. The hydrocarbon gases normally found in a natural gas are methane, ethane, propane, butanes, pentanes, and small amounts of hexanes and heavier components. The nonhydrocarbon gases, that is, impurities, include carbon dioxide, hydrogen sulfide, and nitrogen. Knowledge of pressure-volume-temperature (PVT) relationships and other physical and chemical properties of gases are essential for solving problems in natural gas reservoir engineering. The properties of interest include n n n n n n n n n Apparent molecular weight, Ma. Specific gravity, γ g. Compressibility factor, Z. Gas density, ρg. Specific volume, υ. Isothermal gas compressibility coefficient, cg. Gas formation volume factor, Bg. Gas expansion factor, Eg. Viscosity, μg. These gas properties may be obtained from direct laboratory measurements or by prediction from generalized mathematical expressions. This chapter reviews laws that describe the volumetric behavior of gases in terms of pressure and temperature and documents the mathematical correlations widely used in determining the physical properties of natural gases. Equations of State and PVT Analysis. http://dx.doi.org/10.1016/B978-0-12-801570-4.00003-9 Copyright # 2016 Elsevier Inc. All rights reserved. 189 190 CHAPTER 3 Natural Gas Properties BEHAVIOR OF IDEAL GASES The kinetic theory of gases postulates that the gas is composed of a very large number of particles called molecules. For an ideal gas, the volume of these molecules is insignificant compared with the total volume occupied by the gas. It is also assumed that these molecules have no attractive or repulsive forces between them, and it is assumed that all collisions of molecules are perfectly elastic. Based on this kinetic theory of gases, a mathematical equation, called equation of state, can be derived to express the relationship existing between pressure, p, volume, V, and temperature, T, for a given quantity of moles of gas, n. This relationship for perfect gases, called the ideal gas law, is expressed mathematically by the following equation: pV ¼ nRT (3.1) where p ¼ absolute pressure, psia V ¼ volume, ft3 T ¼ absolute temperature, °R n ¼ number of moles of gas, lb-mol R ¼ the universal gas constant that, for these units, has the value 10.73 psia ft3/lb-mol °R The number of pound-moles of gas, that is, n, is defined as the weight of the gas, m, divided by the molecular weight, M, or n¼ m M (3.2) Combining Eq. (3.1) with Eq. (3.2) gives pV ¼ m M where RT (3.3) m ¼ weight of gas, lb M ¼ molecular weight, lb/lb-mol Because the density is defined as the mass per unit volume of the substance, Eq. (3.3) can be rearranged to estimate the gas density at any pressure and temperature: ρg ¼ m pM ¼ V RT (3.4) where ρg ¼ density of the gas, lb/ft3. It should be pointed out that lb refers to pounds mass in all the subsequent discussions of density in this text. Behavior of Ideal Gases 191 EXAMPLE 3.1 Three pounds of n-butane are placed in a vessel at 120°F and 60 psia. Calculate the volume of the gas assuming an ideal gas behavior. Solution Step 1. Determine the molecular weight of n-butane from Table 1.1 to give M ¼ 58.123 Step 2. Solve Eq. (3.3) for the volume of gas: m RT V¼ M p 3 ð10:73Þð120 + 460Þ V¼ ¼ 5:35ft3 58:123 60 EXAMPLE 3.2 Using the data given in the preceding example, calculate the density of n-butane. Solution Solve for the density by applying Eq. (3.4): m pM ¼ V RT ð60Þð58:123Þ ρg ¼ ¼ 0:56lb=ft3 ð10:73Þð580Þ ρg ¼ Petroleum engineers usually are interested in the behavior of mixtures and rarely deal with pure component gases. Because natural gas is a mixture of hydrocarbon components, the overall physical and chemical properties can be determined from the physical properties of the individual components in the mixture by using appropriate mixing rules. The basic properties of gases commonly are expressed in terms of the apparent molecular weight, standard volume, density, specific volume, and specific gravity. These properties are defined in the following sections. Apparent Molecular Weight, Ma One of the main gas properties frequently of interest to engineers is the apparent molecular weight. If yi represents the mole fraction of the ith component in a gas mixture, the apparent molecular weight is defined mathematically by the following equation: Ma ¼ X yi Mi i¼1 (3.5) 192 CHAPTER 3 Natural Gas Properties where Ma ¼ apparent molecular weight of a gas mixture Mi ¼ molecular weight of the ith component in the mixture yi ¼ mole fraction of component i in the mixture Conventionally, natural gas compositions can be expressed in three different forms: mole fraction, yi, weight fraction, wi, and volume fraction, υi. The mole fraction of a particular component, i, is defined as the number of moles of that component, ni, divided by the total number of moles, n, of all the components in the mixture: yi ¼ ni ni ¼X n ni i The weight fraction of a particular component, i, is defined as the weight of that component, mi, divided by the total weight of m, the mixture: wi ¼ mi mi ¼X m mi i Similarly, the volume fraction of a particular component, vi, is defined as the volume of that component, Vi, divided by the total volume of V, the mixture: vi ¼ Vi Vi ¼X V Vi i It is convenient in many engineering calculations to convert from mole fraction to weight fraction and vice versa. The procedure is given in the following steps. 1. As the composition is one of the intensive properties and independent of the quantity of the system, assume that the total number of gas is 1; that is, n ¼ 1. 2. From the definitions of mole fraction and number of moles (see Eq. (3.2)), ni ni ¼ ¼ ni n 1 mi ¼ ni Mi ¼ yi Mi yi ¼ 3. From the above two expressions, calculate the weight fraction to give wi ¼ mi mi yi Mi yi Mi ¼X ¼X ¼ m mi yi Mi Ma i 4. Similarly, i wi =Mi yi ¼ X w =Mi i i Behavior of Ideal Gases 193 Standard Volume, Vsc In many natural gas engineering calculations, it is convenient to measure the volume occupied by l lb-mol of gas at a reference pressure and temperature. These reference conditions are usually 14.7 psia and 60°F and are commonly referred to as standard conditions. The standard volume then is defined as the volume of gas occupied by 1 lb-mol of gas at standard conditions. Applying these conditions to Eq. (3.1) and solving for the volume, that is, the standard volume, gives Vsc ¼ ð1ÞRTsc ð1Þð10:73Þð520Þ ¼ psc 14:7 (3.6) Vsc ¼ 379:4scf=1b mol where Vsc ¼ standard volume, scf/lb-mol scf ¼ standard cubic feet Tsc ¼ standard temperature, °R psc ¼ standard pressure, psia Gas Density, ρg The density of an ideal gas mixture is calculated by simply replacing the molecular weight, M, of the pure component in Eq. (3.4) with the apparent molecular weight, Ma, of the gas mixture to give ρg ¼ pMa RT (3.7) where ρg ¼ density of the gas mixture, lb/ft3, and Ma ¼ apparent molecular weight. Specific Volume, υ The specific volume is defined as the volume occupied by a unit mass of the gas. For an ideal gas, this property can be calculated by applying Eq. (3.3): υ¼ V 1 ¼ m ρg Combining with Eq. (3.7) gives: υ¼ RT pMa where υ ¼ specific volume, ft3/lb, and ρg ¼ gas density, lb/ft3. (3.8) 194 CHAPTER 3 Natural Gas Properties Specific Gravity, γ g The specific gravity is defined as the ratio of the gas density to that of the air. Both densities are measured or expressed at the same pressure and temperature. Commonly, the standard pressure, psc, and standard temperature, Tsc, are used in defining the gas specific gravity. γg ¼ gas density @ 14:7 and 60° ρg ¼ air density @ 14:7 and 60° ρair (3.9) Assuming that the behavior of both the gas mixture and the air is described by the ideal gas equation, the specific gravity can be then expressed as psc Ma RTsc γg ¼ psc Mair RTsc or γg ¼ Ma Ma ¼ Mair 28:96 (3.10) where γ g ¼ gas specific gravity, 60°/60° ρair ¼ density of the air Mair ¼ apparent molecular weight of the air ¼ 28.96 Ma ¼ apparent molecular weight of the gas psc ¼ standard pressure, psia Tsc ¼ standard temperature, °R EXAMPLE 3.3 A gas well is producing gas with a specific gravity of 0.65 at a rate of 1.1 MMscf/ day. The average reservoir pressure and temperature are 1500 psi and 150°F. Calculate 1. The apparent molecular weight of the gas. 2. The gas density at reservoir conditions. 3. The flow rate in lb/day. Solution From Eq. (3.10), solve for the apparent molecular weight: Ma ¼ 28:96γ g Ma ¼ ð28:96Þð0:65Þ ¼ 18:82 Apply Eq. (3.7) to determine gas density: ρg ¼ pMa RT Behavior of Ideal Gases 195 ρg ¼ ð1500Þð18:82Þ ¼ 4:311b=ft3 ð10:73Þð610Þ The following steps are used to calculate the flow rate. Step 1. Because 1 lb-mol of any gas occupies 379.4 scf at standard conditions, then the daily number of moles, n, that the gas well is producing is n¼ ð1:1Þð10Þ6 ¼ 28991b mol 379:4 Step 2. Determine the daily mass, m, of the gas produced from Eq. (2.2): m ¼ nMa m ¼ 2899 ð18:82Þ ¼ 54,5591b=day EXAMPLE 3.4 A gas well is producing a natural gas with the following composition: Component yi CO2 C1 C2 C3 0.05 0.90 0.03 0.02 Assuming an ideal gas behavior, calculate the apparent molecular weight, specific gravity, gas density at 2000 psia and 150°F, and specific volume at 2000 psia and 150°F. Solution The table below describes the gas composition. Component yi Mi yiMi CO2 C1 C2 C3 0.05 0.90 0.03 0.02 44.01 16.04 30.07 44.11 2.200 14.436 0.902 0.882 Ma ¼ 18.42 Apply Eq. (3.5) to calculate the apparent molecular weight: X yi Mi Ma ¼ i¼1 Ma ¼ 18:42 Calculate the specific gravity by using Eq. (3.10): γ g ¼ Ma =28:96 ¼ 18:42=28:96 ¼ 0:636 196 CHAPTER 3 Natural Gas Properties Solve for the density by applying Eq. (3.7): ρg ¼ ð2000Þð18:42Þ ¼ 5:6281b=ft3 ð10:73Þð610Þ Determine the specific volume from Eq. (3.8): υ¼ 1 1 ¼ ¼ 0:178ft3 =1b ρg 5:628 BEHAVIOR OF REAL GASES In dealing with gases at a very low pressure, the ideal gas relationship is a convenient and generally satisfactory tool. At higher pressures, the use of the ideal gas equation of state may lead to errors as great as 500%, as compared to errors of 2–3% at atmospheric pressure. Basically, the magnitude of deviations of real gases from the conditions of the ideal gas law increases with increasing pressure and temperature and varies widely with the composition of the gas. Real gases behave differently than ideal gases. The reason for this is that the perfect gas law was derived under the assumption that the volume of molecules is insignificant and neither molecular attraction nor repulsion exists between them. This is not the case for real gases. Numerous equations of state have been developed in the attempt to correlate the PVT variables for real gases with experimental data. To express a more exact relationship between the variables p, V, and T, a correction factor, called the gas compressibility factor, gas deviation factor, or simply the Z-factor, must be introduced into Eq. (3.1) to account for the departure of gases from ideality. The equation has the following form: pV ¼ ZnRT (3.11) where the gas compressibility factor, Z, is a dimensionless quantity, defined as the ratio of the actual volume of n-moles of gas at T and p to the ideal volume of the same number of moles at the same T and p: Z¼ Vactual V ¼ Videal ðnRT Þ=p Studies of the gas compressibility factors for natural gases of various compositions show that the compressibility factor can be generalized with sufficient accuracy for most engineering purposes when they are expressed in terms of the following two dimensionless properties: pseudoreduced Behavior of Real Gases 197 pressure, ppr, and pseudoreduced temperature, Tpr. These dimensionless terms are defined by the following expressions: ppr ¼ p ppc (3.12) Tpr ¼ T Tpc (3.13) where p ¼ system pressure, psia ppr ¼ pseudoreduced pressure, dimensionless T ¼ system temperature, °R Tpr ¼ pseudoreduced temperature, dimensionless ppc, Tpc ¼ pseudocritical pressure and temperature, respectively, defined by the following relationships: Ppc ¼ X yi pci (3.14) X yi Tci (3.15) i¼1 Tpc ¼ i¼1 Matthews et al. (1942) correlated the critical properties of the C7+ fraction as a function of the molecular weight and specific gravity: ðpc ÞC7 + ¼ 1188 431 log ðMC7 + 61:1Þ + ½2319 852log ðMC7 + 53:7Þ γ C7 + 0:8 ðTc ÞC7 + ¼ 608 + 364log ðMC7 + 71:2Þ + ½2450log ðMC7 + Þ 3800log γ C7 + It should be pointed out that these pseudocritical properties, that is, ppc and Tpc, do not represent the actual critical properties of the gas mixture. These pseudoproperties are used as correlating parameters in generating gas properties. Based on the concept of pseudoreduced properties, Standing and Katz (1942) presented a generalized gas compressibility factor chart as shown in Fig. 3.1. The chart represents the compressibility factors of sweet natural gas as a function of ppr and Tpr. This chart is generally reliable for natural gas with a minor amount of nonhydrocarbons. It is one of the most widely accepted correlations in the oil and gas industry. 198 CHAPTER 3 Natural Gas Properties n FIGURE 3.1 Standing and Katz compressibility factors chart. (Courtesy of the Gas Processors Suppliers Association, 1987. GPSA Engineering Data Book, 10th ed. Gas Processors Suppliers Association, Tulsa, OK.) EXAMPLE 3.5 A gas reservoir has the following gas composition: Component yi CO2 N2 C1 0.02 0.01 0.85 Behavior of Real Gases 199 C2 C3 i-C4 n-C4 0.04 0.03 0.03 0.02 The initial reservoir pressure and temperature are 3000 psia and 180°F, respectively. Calculate the gas compressibility factor under initial reservoir conditions. Solution The table below describes the gas composition. Component yi Tci (°R) yiTci pci yipci CO2 N2 C1 C2 C3 i-C4 n-C4 0.02 0.01 0.85 0.04 0.03 0.03 0.02 547.91 227.49 343.33 549.92 666.06 734.46 765.62 10.96 2.27 291.83 22.00 19.98 22.03 15.31 Tpc ¼ 383.38 1071 493.1 666.4 706.5 616.40 527.9 550.6 21.42 4.93 566.44 28.26 18.48 15.84 11.01 ppc ¼ 666.38 Step 1. Determine the pseudocritical pressure from Eq. (3.14): X ppc ¼ yi pci ¼ 666:38psi Step 2. Calculate the pseudocritical temperature from Eq. (3.15): Tpc ¼ X yi Tci ¼ 383:38°R Step 3. Calculate the pseudoreduced pressure and temperature by applying Eqs. (3.12), (3.13), respectively: ppr ¼ p 3000 ¼ ¼ 4:50 ppc 666:38 Tpc ¼ T 640 ¼ ¼ 1:67 Tpc 383:38 Step 4. Determine the Z-factor from Fig. 3.1, using the calculate values of ppr and Tpr to give Z ¼ 0.85. Eq. (3.11) can be written in terms of the apparent molecular weight, Ma, and the weight of the gas, m: m RT pV ¼ Z Ma 200 CHAPTER 3 Natural Gas Properties Solving this relationship for the gas’s specific volume and density give V ZRT ¼ m pMa (3.16) 1 pMa ρg ¼ ¼ υ ZRT (3.17) υ¼ where υ ¼ specific volume, ft3/lb ρg ¼ density, lb/ft3 EXAMPLE 3.6 Using the data in Example 3.5 and assuming real gas behavior, calculate the density of the gas phase under initial reservoir conditions. Compare the results with that of ideal gas behavior. Solution The table below describes the gas composition. Component yi Mi CO2 N2 C1 C2 C3 i-C4 n-C4 44.01 0.88 547.91 28.01 0.28 227.49 16.04 13.63 343.33 30.1 1.20 549.92 44.1 1.32 666.06 58.1 1.74 734.46 58.1 1.16 765.62 Ma ¼ 20.23 0.02 0.01 0.85 0.04 0.03 0.03 0.02 yiMi Tci (°R) yiTci pci yipci 10.96 1071 21.42 2.27 493.1 4.93 291.83 666.4 566.44 22.00 706.5 28.26 19.98 616.40 18.48 22.03 527.9 15.84 15.31 550.6 11.01 Tpc ¼ 383.38 ppc ¼ 666.38 Step 1. Calculate the apparent molecular weight from Eq. (3.5): X Ma ¼ yi Mi ¼ 20:23 Step 2. Determine the pseudocritical pressure from Eq. (3.14): X yi pc i ¼ 666:38psi ppc ¼ Step 3. Calculate the pseudocritical temperature from Eq. (3.15): X Tpc ¼ yi Tci ¼ 383:38°R Step 4. Calculate the pseudoreduced pressure and temperature by applying Eqs. (3.12), (3.13), respectively: Behavior of Real Gases 201 p 3000 ¼ ¼ 450 ppc 666:38 T 640 ¼ ¼ 1:67 Tpr ¼ Tpc 383:38 ppr ¼ Step 5. Determine the Z-factor from Fig. 3.1 using the calculated values of ppr and Tpr to give Z ¼ 0.85 Step 6. Calculate the density from Eq. (3.17). pMa ZRT ð3000Þð20:23Þ ¼ 10:41b=ft3 ρg ¼ ð0:85Þð10:73Þð640Þ ρg ¼ Step 7. Calculate the density of the gas, assuming ideal gas behavior, from Eq. (3.7): ð3000Þð20:23Þ ¼ 8:841b=ft3 ρg ¼ ð10:73Þð640Þ The results of this example show that the ideal gas equation estimated the gas density with an absolute error of 15% when compared with the density value as predicted with the real gas equation. In cases where the composition of a natural gas is not available, the pseudocritical properties, ppc and Tpc, can be predicted solely from the specific gravity of the gas. Brown et al. (1948) presented a graphical method for a convenient approximation of the pseudocritical pressure and pseudocritical temperature of gases when only the specific gravity of the gas is available. Standing (1977) expressed this graphical correlation in the following mathematical forms. For Case 1, natural gas systems, Tpc ¼ 168 + 325γ g 12:5γ 2g (3.18) ppc ¼ 677 + 15:0γ 37:5γ 2g (3.19) For Case 2, wet gas systems, Tpc ¼ 187 + 330γ 71:5γ 2g (3.20) ppc ¼ 706 51:7γ g 11:1γ 2g (3.21) where Tpc ¼ pseudocritical temperature, °R ppc ¼ pseudocritical pressure, psia γ g ¼ specific gravity of the gas mixture 202 CHAPTER 3 Natural Gas Properties EXAMPLE 3.7 Using Eqs. (3.18), (3.19), calculate the pseudocritical properties and rework Example 3.5. Solution Step 1. Calculate the specific gravity of the gas: γg ¼ Ma 20:23 ¼ ¼ 0:699 28:96 28:96 Step 2. Solve for the pseudocritical properties by applying Eqs. (3.18), (3.19): 2 Tpc ¼ 168 + 325γ g 12:5 γ g Tpc ¼ 168 + 325ð0:699Þ 12:5ð0:699Þ2 ¼ 389:1°R 2 ppc ¼ 677 + 15γ g 37:5 γ g ppc ¼ 677 + 15ð0:699Þ 37:5ð0:699Þ2 ¼ 669:2psi Step 3. Calculate ppr and Tpr: 3000 ¼ 4:48 669:2 640 ¼ 1:64 Tpr ¼ 389:1 ppr ¼ Step 4. Determine the gas compressibility factor from Fig. 3.1: Z ¼ 0:824 Step 5. Calculate the density from Eq. (3.17): pMa ZRT ð3000Þð20:23Þ ¼ 10:461b=ft3 ρg ¼ ð0:845Þð10:73Þð640Þ ρg ¼ Effect of Nonhydrocarbon Components on the Z-Factor Natural gases frequently contain materials other than hydrocarbon components, such as nitrogen, carbon dioxide, and hydrogen sulfide. Hydrocarbon gases are classified as sweet or sour depending on the hydrogen sulfide content. Both sweet and sour gases may contain nitrogen, carbon dioxide, or both. A hydrocarbon gas is termed a sour gas if it contains 1 grain of H2S per 100 ft3. The common occurrence of small percentages of nitrogen and carbon dioxide, in part, is considered in the correlations previously cited. Concentrations of up to 5% of these nonhydrocarbon components do not seriously Nonhydrocarbon Adjustment Methods 203 affect accuracy. Errors in compressibility factor calculations as large as 10% may occur in higher concentrations of nonhydrocarbon components in gas mixtures. NONHYDROCARBON ADJUSTMENT METHODS Two methods were developed to adjust the pseudocritical properties of the gases to account for the presence of the n-hydrocarbon components: the Wichert-Aziz method and the Carr-Kobayashi-Burrows method. Wichert-Aziz's Correction Method Natural gases that contain H2S and/or CO2 frequently exhibit different compressibility factors behavior than sweet gases. Wichert and Aziz (1972) developed a simple, easy-to-use calculation to account for these differences. This method permits the use of the Standing-Katz Z-factor chart (Fig. 3.1), by using a pseudocritical temperature adjustment factor, ε, which is a function of the concentration of CO2 and H2S in the sour gas. This correction factor then is used to adjust the pseudocritical temperature and pressure according to the following expressions: 0 Tpc ¼ Tpc ε p0pc ¼ 0 ppc Tpc Tpc + Bð1 BÞε (3.22) (3.23) where Tpc ¼ pseudocritical temperature, °R ppc ¼ pseudocritical pressure, psia Tpc0 ¼ corrected pseudocritical temperature, °R ppc0 ¼ corrected pseudocritical pressure, psia B ¼ mole fraction of H2S in the gas mixture ε ¼ pseudocritical temperature adjustment factor, defined mathematically by the following expression: ε ¼ 120 A0:9 A1:6 + 15 B0:5 B4:0 (3.24) where the coefficient A is the sum of the mole fraction H2S and CO2 in the gas mixture: A ¼ yH2 S + yCO2 The computational steps of incorporating the adjustment factor ε into the Z-factor calculations are summarized next. 204 CHAPTER 3 Natural Gas Properties Step 1. Calculate the pseudocritical properties of the whole gas mixture by applying Eqs. (3.18), (3.19) or Eqs. (3.20), (3.21). Step 2. Calculate the adjustment factor, ε, from Eq. (3.24). Step 3. Adjust the calculated ppc and Tpc (as computed in step 1) by applying Eqs. (3.22), (3.23). Step 4. Calculate the pseudoreduced properties, ppr and Tpr, from Eqs. (3.11), (3.12). Step 5. Read the compressibility factor from Fig. 3.1. EXAMPLE 3.8 A sour natural gas has a specific gravity of 0.70. The compositional analysis of the gas shows that it contains 5% CO2 and 10% H2S. Calculate the density of the gas at 3500 psia and 160°F. Solution Step 1. Calculate the uncorrected pseudocritical properties of the gas from Eqs. (3.18), (3.19): 2 Tpc ¼ 168 + 325γ g 12:5 γ g Tpc ¼ 168 + 325ð0:7Þ 12:5ð0:7Þ2 ¼ 389:38°R 2 ppc ¼ 677 + 15γ g 37:5 γ g ppc ¼ 677 + 15ð0:7Þ 37:5ð0:7Þ2 ¼ 669:1psi Step 2. Calculate the pseudocritical temperature adjustment factor from Eq. (3.24): ε ¼ 120 0:150:9 0:151:6 + 15 0:50:5 0:14 ¼ 20:735 Step 3. Calculate the corrected pseudocritical temperature by applying Eq. (3.22): 0 Tpc ¼ 389:38 20:735 ¼ 368:64 Step 4. Adjust the pseudocritical pressure ppc by applying Eq. (3.23): p0pc ¼ ð669:1Þð368:64Þ ¼ 630:44psia 389:38 + 0:1ð1 0:1Þð20:635Þ Step 5. Calculate ppr and Tpr: ppr ¼ 3500 ¼ 5:55 630:44 Tpr ¼ 160 + 460 ¼ 1:68 368:64 Nonhydrocarbon Adjustment Methods 205 Step 6. Determine the Z-factor from Fig. 3.1: Z ¼ 0:89 Step 7. Calculate the apparent molecular weight of the gas from Eq. (3.10): Ma ¼ 28:96γ g ¼ 28:96ð0:7Þ 20:27 Step 8. Solve for the gas density by applying Eq. (3.17): pMa ZRT ð3500Þð20:27Þ ¼ 11:98lb=ft3 ρg ¼ ð0:89Þð10:73Þð620Þ ρg ¼ Whitson and Brule (2000) point out that, when only the gas specific gravity and nonhydrocarbon content are known (including nitrogen, yN 2 ), the following procedure is recommended: n Calculate hydrocarbon specific gravity γ gHC (excluding the nonhydrocarbon components) from the following relationship: γ gHC ¼ n 28:96γ g ðyN2 MN2 + yCO2 + MCO2 + yH2 S MH2 S Þ 28:96ð1 yN2 yCO2 yH2 S Þ Using the calculated hydrocarbon specific gravity, γ gHC, determine the pseudocritical properties from Eqs. (3.18), (3.19): T ¼ 168 + 325γ gHC 12:5γ 2gHC pc HC ppc HC ¼ 677 + 150γ gHC 37:5γ 2gHC n Adjust these two values to account for the nonhydrocarbon components by applying the following relationships: ppc ¼ ð1 yN2 yCO2 yH2 S Þ ppc HC + yN2 ð pc ÞN2 + yCO2 ð pc ÞCO2 + yH2 S ðpc ÞH2 S Tpc ¼ ð1 yN2 yCO2 yH2 S Þ Tpc HC + yN2 ðTc ÞN2 + yCO2 ðTc ÞCO2 + yH2 S ðTc ÞH2 S n Use the above calculated pseudocritical properties in Eqs. (3.22), (3.23) to obtain the adjusted properties for the Wichert-Aziz method of calculating the Z-factor. Carr-Kobayashi-Burrows's Correction Method Carr et al. (1954) proposed a simplified procedure to adjust the pseudocritical properties of natural gases when nonhydrocarbon components are 206 CHAPTER 3 Natural Gas Properties present. The method can be used when the composition of the natural gas is not available. The proposed procedure is summarized in the following steps. Step 1. Knowing the specific gravity of the natural gas, calculate the pseudocritical temperature and pressure by applying Eqs. (3.18), (3.19). Step 2. Adjust the estimated pseudocritical properties by using the following two expressions: 0 Tpc ¼ Tpc 80yCO2 + 130yH2 S 250yN2 (3.25) p0pc ¼ ppc + 440yCO2 + 600yH2 S 170yN2 (3.26) where Tpc0 ¼ the adjusted pseudocritical temperature, °R Tpc ¼ the unadjusted pseudocritical temperature, °R yCO 2 ¼ mole fraction of CO2 yH 2 S ¼ mole fraction of H2S in the gas mixture yN2 ¼ mole fraction of nitrogen ppc0 ¼ the adjusted pseudocritical pressure, psia ppc ¼ the unadjusted pseudocritical pressure, psia Step 3. Use the adjusted pseudocritical temperature and pressure to calculate the pseudoreduced properties. Step 4. Calculate the Z-factor from Fig. 3.1. EXAMPLE 3.9 Using the data in Example 3.8, calculate the density by employing the preceding correction procedure. Solution Step 1. Determine the corrected pseudocritical properties from Eqs. (3.25), (3.26). 0 Tpc ¼ 389:38 80ð0:05Þ + 130ð0:10Þ 250ð0Þ ¼ 398:38°R 0 ppc ¼ 669:1 + 440ð0:05Þ + 600ð0:01Þ 170ð0Þ ¼ 751:1psia Step 2. Calculate ppr and Tpr: 3500 ¼ 4:56 751:1 620 ¼ 1:56 Tpr ¼ 398:38 ppr ¼ Correction for High-Molecular Weight Gases 207 Step 3. Determine the gas compressibility factor from Fig. 3.1: Z ¼ 0:820 Step 4. Calculate the gas density: ρg ¼ pMa ZRT γg ¼ ð3500Þð20:27Þ ¼ 13:0 lb=ft3 ð0:82Þð10:73Þð620Þ The gas density as calculated in Example 3.8 with a Z-factor of 0.89 is ρg ¼ ð3500Þð20:27Þ ¼ 11:98lb=ft3 ð0:89Þð10:73Þð620Þ CORRECTION FOR HIGH-MOLECULAR WEIGHT GASES It should be noted that the Standing and Katz Z-factor chart (Fig. 3.1) was prepared from data on binary mixtures of methane with propane, ethane, and butane and on natural gases, thus covering a wide range in composition of hydrocarbon mixtures containing methane. No mixtures having molecular weights in excess of 40 were included in preparing this plot. Sutton (1985) evaluated the accuracy of the Standing-Katz compressibility factor chart using laboratory-measured gas compositions and Z-factors and found that the chart provides satisfactory accuracy for engineering calculations. However, Kay’s mixing rules, that is, Eqs. (3.13), (3.14), or comparable gravity relationships for calculating pseudocritical pressure and temperature, result in unsatisfactory Z-factors for high-molecular weight reservoir gases. Sutton observed that large deviations occur to gases with high heptanes-plus concentrations. He pointed out that Kay’s mixing rules should not be used to determine the pseudocritical pressure and temperature for reservoir gases with specific gravities greater than about 0.75. Sutton proposed that this deviation can be minimized by utilizing the mixing rules developed by Stewart et al. (1959), together with newly introduced empirical adjustment factors (FJ, EJ, and EK) to account for the presence of the heptanes-plus fraction, C7+, in the gas mixture. The proposed approach is outlined in the following steps. 208 CHAPTER 3 Natural Gas Properties Step 1. Calculate the parameters J and K from the following relationships: " " # #2 1 X Tci 2 X Tci 0:5 + J¼ yi yi pci pci 3 i 3 i K¼ X yi Tci pffiffiffiffiffiffi pci i (3.27) (3.28) where J ¼ Stewart-Burkhardt-Voo correlating parameter, °R/psia K ¼ Stewart-Burkhardt-Voo correlating parameter, °R/psia yi ¼ mole fraction of component i in the gas mixture Step 2. Calculate the adjustment parameters FJ, EJ, and EK from the following expressions: sffiffiffiffiffi!#2 " 1 Tc 2 Tc + FJ ¼ y y pc pc 3 3 (3.29) C7 + EJ ¼ 0:6081FJ + 1:1325F2J 14:004FJ y2C7 + h i T 0:3129yC7 + 4:8156ðyC7 + Þ2 + 27:3751ðyC7 + Þ3 EK ¼ pffiffiffiffiffi p c C7 + (3.30) (3.31) where yC7 + ¼ mole fraction of the heptanes-plus component ðTc ÞC7 + ¼ critical temperature of the C7+ ðpc ÞC7 + ¼ critical pressure of the C7+ Step 3. Adjust the parameters J and K by applying the adjustment factors EJ and EK, according to these relationships: J 0 ¼ J EJ 0 K ¼ K EK (3.32) (3.33) where J, K are calculated from Eqs. (3.27), (3.28) EJ, EK are calculated from Eqs. (3.30), (3.31) Step 4. Calculate the adjusted pseudocritical temperature and pressure from the expressions 0 Tpc ¼ ðK 0 Þ2 J0 (3.34) Correction for High-Molecular Weight Gases 209 p0pc ¼ 0 Tpc J0 (3.35) Step 5. Having calculated the adjusted Tpc and ppc, the regular procedure of calculating the compressibility factor from the Standing and Katz chart is followed. Sutton’s proposed mixing rules for calculating the pseudocritical properties of high-molecular weight reservoir gases, γ g > 0.75, should significantly improve the accuracy of the calculated Z-factor. EXAMPLE 3.10 A hydrocarbon gas system has the following composition: Component yi C1 C2 C3 n-C4 n-C5 C6 C7+ 0.83 0.06 0.03 0.02 0.02 0.01 0.03 The heptanes-plus fraction is characterized by a molecular weight and specific gravity of 161 and 0.81, respectively. Using Sutton’s methodology, calculate the density of the gas 2000 psi and 150°F. Then, recalculate the gas density without adjusting the pseudocritical properties. Solution using Sutton's methodology Step 1. Calculate the critical properties of the heptanes-plus fraction by the Riazi and Daubert correlation [Chapter 2, Eq. (2.6)] and construct the table as shown below: θ ¼ aðMÞb γ c exp½d ðMÞ + eγ + f ðMÞγ ðTc ÞC7 + ¼ ð544:2Þ ð161Þ0:2998 ð0:81Þ1:0555 exp h i 1:3478ð10Þ4 ð150Þ 0:61641ð0:81Þ ¼ 1189°R ðpc ÞC7 + ¼ 4:5203ð10Þ4 1610:8063 0:811:6015 exp h i 1:8078ð10Þ3 ð150Þ 0:3084ð0:81Þ ¼ 318:4psia 210 CHAPTER 3 Natural Gas Properties pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ðTc =pc Þi Component yi Mi Tci pci yiMi yi (Tci/pci) yi C1 C2 C3 n-C4 n-C5 C6 C7+ Total 0.83 0.06 0.03 0.02 0.02 0.01 0.03 16.0 30.1 44.1 58.1 72.2 84.0 161.0 343.33 549.92 666.06 765.62 845.60 923.00 1189.0 666.4 706.5 616.4 550.6 488.6 483.0 318.4 13.31 1.81 1.32 1.16 1.45 0.84 4.83 24.72 0.427 0.047 0.032 0.028 0.035 0.019 0.112 0.700 0.596 0.053 0.031 0.024 0.026 0.014 0.058 0.802 pffiffiffiffiffi yi Tc = pc i 11.039 1.241 0.805 0.653 0.765 0.420 1.999 16.922 Step 2. Calculate the parameters J and K from Eqs. (3.27), (3.28) and using the values from the above table. 1 2 J ¼ ð0:700Þ + ð0:802Þ2 ¼ 0:662 3 3 " " # 0:5 #2 1 X Tci 2 X Tci + yi yi J¼ pci pci 3 i 3 i K¼ X yi Tci pffiffiffiffiffiffi ¼ 16:922 pci i Step 3. To account for the presence of the C7+ fraction, calculate the adjustment factors FJ, EJ, and EK by applying Eqs. (3.29)–(3.31): sffiffiffiffiffi!#2 " 1 Tc 2 Tc FJ ¼ y + pc pc 3 3 C7 + 1 2 FJ ¼ ½0:112 + ½0:0582 ¼ 0:0396 3 3 EJ ¼ 0:6081FJ + 1:1325F2J 14:004FJ yC7 + + 64:434FJ y2C7 + EJ ¼ 0:6081ð0:04Þ + 1:0352ð0:04Þ2 14:004ð0:04Þð0:03Þ + 64:434ð0:04Þð0:3Þ2 ¼ 0:012 h i h i ffi 0:3129yC7 + 4:8156ðyC7 + Þ2 + 27:2751ð0:03Þ3 EK ¼ pTffiffiffi pc C7 + h i EK ¼ 66:634 0:3129ð0:03Þ 4:8156ð0:03Þ2 + 27:3751ð0:03Þ3 ¼ 0:386 Step 4. Calculate the parameters J0 and K0 from Eqs. (3.34), (3.35): J 0 ¼ J EJ ¼ 0:662 0:012 ¼ 0:650 K 0 ¼ K EK ¼ 16:922 0:386 ¼ 16:536 Correction for High-Molecular Weight Gases 211 Step 5. Determine the adjusted pseudocritical properties from Eqs. (3.34), (3.35): ðK 0 Þ2 ð16:536Þ2 ¼ ¼ 420:7 J0 0:65 0 Tpc 420:7 p0pc ¼ 0 ¼ ¼ 647:2 J 0:65 0 ¼ Tpc Step 6. Calculate the pseudoreduced properties of the gas by applying Eqs. (3.11), (3.12) to give 2000 ¼ 3:09 647:2 610 ¼ 1:45 Tpr ¼ 420:7 ppr ¼ Step 7. Calculate the Z-factor from Fig. 3.1 to give Z ¼ 0:745 Step 8. From Eq. (3.17), calculate the density of the gas: pMa ZRT ð2000Þð24:73Þ ¼ 10:14lb=ft3 ρg ¼ ð10:73Þð610Þð0:745Þ ρg ¼ Solution without adjusting the pseudocritical properties Step 1. Calculate the specific gravity of the gas: γg ¼ Ma 24:73 ¼ ¼ 0:854 28:96 28:96 Step 2. Solve for the pseudocritical properties by applying Eqs. (3.18), (3.19): Tpc ¼ 168 + 325ð0:84Þ 12:5ð0:854Þ2 ¼ 436:4°R ppc ¼ 677 + 15ð0:854Þ 37:5ð0:854Þ2 ¼ 662:5psia Step 3. Calculate ppr and Tpr: 2000 ¼ 3:02 662:5 610 ¼ 1:40 Tpr ¼ 436:4 ppr ¼ Step 4. Calculate the Z-factor from Fig. 3.1 to give Z ¼ 0:710 Step 5. From Eq. (3.17), calculate the density of the gas: pMa ZRT ð2000Þð24:73Þ ¼ 10:64lb=ft3 ρg ¼ ð10:73Þð610Þð0:710Þ ρg ¼ 212 CHAPTER 3 Natural Gas Properties Direct Calculation of Compressibility Factors After four decades of existence, the Standing-Katz Z-factor chart is still widely used as a practical source of natural gas compressibility factors. As a result, there was an apparent need for a simple mathematical description of that chart. Several empirical correlations for calculating Z-factors have been developed over the years. Papay (1985) proposed a simple expression for calculating the gas compressibility factor explicitly. Papay correlated the Z-factor with pseudoreduced pressure ppr and temperature Tpr as expressed next: Z ¼1 0:274p2pr 3:53ppr + 0:8157T 0:9813Tpr 10 10 Tpr For example, at ppr ¼ 3 and Tpr ¼ 2, the Z-factor from the preceding equation is Z ¼1 0:274p2pr 3:53ppr 3:53ð3Þ 0:274ð3Þ2 + 0:8157Tpr ¼ 1 0:9813ð2Þ + 0:8157ð2Þ ¼ 0:9422 0:9813Tpr 10 10 10 10 as compared with value obtained from the Standing and Katz chart of 0.954. Numerous rigorous mathematical expressions have been proposed to accurately reproduce the Standing and Katz Z-factor chart. Most of these expressions are designed to solve for the gas compressibility factor at any ppr and Tpr iteratively. Three of these empirical correlations are described next: Hall-Yarborough, Dranchuk-Abu-Kassem, and Dranchuk-PurvisRobinson. HALL-YARBOROUGH'S METHOD Hall and Yarborough (1973) presented an equation of state that accurately represents the Standing and Katz Z-factor chart. The proposed expression is based on the Starling-Carnahan equation of state. The coefficients of the correlation were determined by fitting them to data taken from the Standing and Katz Z-factor chart. Hall and Yarborough proposed the following mathematical form: Z¼ h i 0:06125tppr exp 1:2ð1 tÞ2 Y (3.36) where ppr ¼ pseudoreduced pressure t ¼ reciprocal of the pseudoreduced temperature (ie, Tpc/T) Y ¼ the reduced density, which can be obtained as the solution of the following equation: Hall-Yarborough's Method 213 FðY Þ ¼ X1 + where Y + Y2 + Y3 Y4 ðX2 ÞY 2 + ðX3 ÞY X4 ¼ 0 ð1 Y Þ (3.37) h i X1 ¼ 0:06125ppr texp 1:2ð1 tÞ2 X2 ¼ 14:76t 9:76t2 + 4:58t3 X3 ¼ 90:7t 242:2t2 + 42:4t3 X4 ¼ ð2:18 + 2:82tÞ Eq. (3.37) is a nonlinear equation and can be solved conveniently for the reduced density Y by using the Newton-Raphson iteration technique. The computational procedure of solving Eq. (3.37) at any specified pseudoreduced pressure, ppr, and temperature, Tpr, is summarized in the following steps. Step 1. Make an initial guess of the unknown parameter, Yk, where k is an iteration counter. An appropriate initial guess of Y is given by the following relationship: h i Y k ¼ 0:0125ppr texp 1:2ð1 tÞ2 Step 2. Substitute this initial value in Eq. (3.37) and evaluate the nonlinear function. Unless the correct value of Y has been initially selected, Eq. (3.37) will have a nonzero value of f(Y). Step 3. A new improved estimate of Y, that is, Yk+1, is calculated from the following expression: f Yk Yk + 1 ¼ Yk 0 k f ðY Þ (3.38) where f 0 (Yk) is obtained by evaluating the derivative of Eq. (3.37) at Yk, or f 0 ðY Þ ¼ 1 + 4Y + 4Y 2 4Y 3 + Y 4 ð1 Y Þ4 2ðX2 ÞY + ðX3 ÞðX4 ÞY ðX4 1Þ (3.39) Step 4. Steps 2 and 3 are repeated n times until the error; that is, abs(Yk Yk+1) becomes smaller than a preset tolerance, say 1012. Step 5. The correct value of Y is then used to evaluate Eq. (3.36) for the compressibility factor: 214 CHAPTER 3 Natural Gas Properties Z¼ h i 0:06125tppr exp 1:2ð1 tÞ2 Y Hall and Yarborough pointed out that the method is not recommended for application if the pseudoreduced temperature is less than 1. DRANCHUK AND ABU-KASSEM'S METHOD Dranchuk and Abu-Kassem (1975) derived an analytical expression for calculating the reduced gas density that can be used to estimate the gas compressibility factor. The reduced gas density ρr is defined as the ratio of the gas density at a specified pressure and temperature to that of the gas at its critical pressure or temperature: ρr ¼ ρ ½pMa =ðZRT Þ ½p=ðZT Þ ¼ ¼ ρc ½pc Ma =ðZc RTc Þ ½pc =ðZc Tc Þ The critical gas compressibility factor Zc is approximately 0.27, which leads to the following simplified expression for the reduced gas density as expressed in terms of the reduced temperature Tr and reduced pressure pr: ρr ¼ 0:27ppr ZTpr (3.40) The authors proposed the following 11-constant equation of state for calculating the reduced gas density: f ðρr Þ ¼ ðR1 Þρr R2 + ðR3 Þρ2r ðR4 Þρ5r + ðR5 Þ 1 + A11 ρ2r ρ2r exp A11 ρ2r + 1 ¼ 0 ρr (3.41) with the coefficients R1 through R5 as defined by the following relations: R1 ¼ A1 + A2 A3 Ar At + 3 + 4 + 5 Tpr Tpr Tpr Tpr 0:27ppr Tpr A7 Ag R3 ¼ A6 + + 2 Tpr Tpr " # A7 A8 R4 ¼ A9 + 2 Tpr Tpr " # A10 R5 ¼ 3 Tpr R2 ¼ Dranchuk and Abu-Kassem's Method 215 The constants A1 through A11 were determined by fitting the equation, using nonlinear regression models, to 1500 data points from the Standing and Katz Z-factor chart. The coefficients have the following values: A1 ¼ 0:3265 A4 ¼ 0:01569 A7 ¼ 0:7361 A10 ¼ 0:6134 A2 ¼ 1:0700 A3 ¼ 0:5339 A5 ¼ 0:05165 A6 ¼ 0:5475 A8 ¼ 0:1844 A9 ¼ 0:1056 A11 ¼ 0:7210 Eq. (3.41) can be solved for the reduced gas density ρr by applying the Newton-Raphson iteration technique as summarized in the following steps. Step 1. Make an initial guess of the unknown parameter, ρkr , where k is an iteration counter. An appropriate initial guess of ρkr is given by the following relationship: ρr ¼ 0:27ppr Tpr Step 2. Substitute this initial value in Eq. (3.41) and evaluate the nonlinear function. Unless the correct value of ρkr has been initially selected, Eq. (3.41) will have a nonzero value for the function f( ρkr ). Step 3. A new improved estimate of ρr, that is, ρkr + 1 , is calculated from the following expression: f ρk ρkr + 1 ¼ ρkr 0 rk f ρr where R2 + 2ðR3 Þγ r 5ðR4 Þρ4r ρ2r + 2ðR5 Þρr exp A11 ρ2r 1 + 2A11 ρ3r A11 ρ2r 1 + A11 ρ2r f 0 ðρr Þ ¼ ðR1 Þ + Step 4. Steps 2 and 3 are repeated n times, until the error, that is, abs ρkr ρkr + 1 , becomes smaller than a preset tolerance, say, 1012. Step 5. The correct value of ρr is then used to evaluate Eq. (3.40) for the compressibility factor: Z¼ 0:27ppr ρr Tpr The proposed correlation was reported to duplicate compressibility factors from the Standing and Katz chart with an average absolute error of 0.585% and is applicable over the ranges 0:2 ppr < 30 1:0 < Tpr 3:0 216 CHAPTER 3 Natural Gas Properties DRANCHUK-PURVIS-ROBINSON METHOD Dranchuk et al. (1974) developed a correlation based on the BenedictWebb-Rubin type of equation of state. Fitting the equation to 1500 data points from the Standing and Katz Z-factor chart optimized the eight coefficients of the proposed equations. The equation has the following form: T5 1 + T1 ρr + T2 ρ2r + T3 ρ5r + T4 ρ2r 1 + A8 ρ2r exp A8 ρ2r ¼ 0 ρr (3.42) with A2 A3 + 3 Tpr Tpr A5 T2 ¼ A4 + Tpr A5 A6 T3 ¼ Tpr A7 T4 ¼ 3 Tpr 0:27ppr T5 ¼ Tpr T1 ¼ A1 + where ρr is defined by Eq. (3.41) and the coefficients A1 through A8 have the following values: A1 ¼ 0:31506237 A5 ¼ 0:31506237 A2 ¼ 1:0467099 A6 ¼ 1:0467099 A3 ¼ 0:57832720 A7 ¼ 0:57832720 A4 ¼ 0:53530771 A8 ¼ 0:53530771 The solution procedure of Eq. (3.43) is similar to that of Dranchuk and AbuKassem. The method is valid within the following ranges of pseudoreduced temperature and pressure: 1:05 Tpr < 3:0 0:2 ppr 3:0 Compressibility of Natural Gases Knowledge of the variability of fluid compressibility with pressure and temperature is essential in performing many reservoir engineering calculations. For a liquid phase, the compressibility is small and usually assumed to Dranchuk-Purvis-Robinson Method 217 be constant. For a gas phase, the compressibility is neither small nor constant. By definition, the isothermal gas compressibility is the change in volume per unit volume for a unit change in pressure, or in equation form, cg ¼ 1 @V V @p T (3.43) where cg ¼ isothermal gas compressibility, 1/psi. From the real gas equation of state, V¼ nRTZ p Differentiating this equation with respect to pressure at constant temperature T gives @V 1 @Z Z ¼ nRT 2 @p p @p p Substituting into Eq. (3.43) produces the following generalized relationship: 1 1 @Z cg ¼ p Z @p T (3.44) For an ideal gas, Z ¼ 1 and (@Z/@p)T ¼ 0; therefore, cg ¼ 1 p (3.45) It should be pointed out that Eq. (3.45) is useful in determining the expected order of magnitude of the isothermal gas compressibility. Eq. (3.44) can be conveniently expressed in terms of the pseudoreduced pressure and temperature by simply replacing p with ( ppr ppc): " # 1 1 @Z cg ¼ ppr ppc Z @ ppr ppc Tpr Multiplying this equation by ppc yields cg ppc ¼ cpr ¼ 1 1 @Z ppr Z @ppr Tpr (3.46) 218 CHAPTER 3 Natural Gas Properties The term cpr is called the isothermal pseudoreduced compressibility, defined by the relationship cpr ¼ cg ppc (3.47) where cpr ¼ isothermal pseudoreduced compressibility cg ¼ isothermal gas compressibility, psi1 ppc ¼ pseudoreduced pressure, psi Values of (@z/@ppr)Tpr can be calculated from the slope of the Tpr isotherm on the Standing and Katz Z-factor chart. EXAMPLE 3.11 A hydrocarbon gas mixture has a specific gravity of 0.72. Calculate the isothermal gas compressibility coefficient at 2000 psia and 140°F assuming, first, an ideal gas behavior, then a real gas behavior. Solution Assuming an ideal gas behavior, determine cg by applying Eq. (3.44) cg ¼ 1 ¼ 500 106 psi1 2000 Assuming a real gas behavior, use the following steps. Step 1. Calculate Tpc and ppc by applying Eqs. (3.17), (3.18): Tpc ¼ 168 + 325ð0:72Þ 12:5ð0:72Þ2 ¼ 395:5°R ppc ¼ 677 + 15ð0:72Þ 37:5ð0:72Þ2 ¼ 668:4 psia Step 2. Compute ppr and Tpr from Eqs. (3.11), (3.12): ppr ¼ 2000 ¼ 2:99 668:4 Tpr ¼ 600 ¼ 1:52 395:5 Step 3. Determine the Z-factor from Fig. 3.1: Z ¼ 0:78 Step 4. Calculate the slope [dZ/dppr]Tpr ¼ 1.52: @Z ¼ 0:022 @ppr Tpr Dranchuk-Purvis-Robinson Method 219 Step 5. Solve for cpr by applying Eq. (3.47): cpr ¼ 1 1 ½0:022 ¼ 0:3627 2:99 0:78 Step 6. Calculate cg from Eq. (3.48): cpr ¼ cg ppc cg ¼ cpr 0:327 ¼ ¼ 543 106 psi1 ppc 668:4 Trube (1957a,b) presented graphs from which the isothermal compressibility of natural gases may be obtained. The graphs, Figs. 3.2 and 3.3, give the isothermal pseudoreduced compressibility as a function of pseudoreduced pressure and temperature. n FIGURE 3.2 Trube's pseudoreduced compressibility for natural gases. (Permission to publish from the Society of Petroleum Engineers of the AIME. # SPE-AIME.) 220 CHAPTER 3 Natural Gas Properties n FIGURE 3.3 Trube's pseudoreduced compressibility for natural gases. (Permission to publish from the Society of Petroleum Engineers of the AIME. # SPE-AIME.) EXAMPLE 3.12 Using Trube’s generalized charts, rework Example 3.11. Solution Step 1. From Fig. 3.3, find cpr to give cpr ¼ 0.36 Step 2. Solve for cg by applying Eq. (3.49): cg ¼ 0:36 ¼ 539 106 psi1 668:4 Mattar et al. (1975) presented an analytical technique for calculating the isothermal gas compressibility. The authors expressed cpr as a function of @Z/ @ρr rather than @Z/@ppr to give 2 3 @Z 0:27 6 ð@Z=@ρr ÞTpr 7 ¼ 5 4 @ppr Z2 Tpr 1 + ρr ð@Z=@ρ Þ r Tpr Z (3.48) Dranchuk-Purvis-Robinson Method 221 Eq. (3.48) may be substituted into Eq. (3.46) to express the pseudoreduced compressibility as 3 2 1 0:27 6 ð@Z=@ρr ÞTpr 7 cpr ¼ 2 4 ρ 5 pr Z Tpr 1 + r ð@Z=@ρ Þ r Tpr Z (3.49) where ρr ¼ pseudoreduced gas density. The partial derivative appearing in Eq. (3.49) is obtained from Eq. (3.42) as @Z ¼ T1 + 2T2 ρr + 5T3 ρ4r + 2T4 ρr 1 + A8 ρ2r A28 ρ4r exp A8 ρ2r @ρr Tpr (3.50) where the coefficients T1 through T4 and A1 through A8 are as defined previously by Eq. (3.42). Gas Formation Volume Factor The gas formation volume factor is used to relate the volume of gas, as measured at reservoir conditions, to the volume of the gas as measured at standard conditions; that is, 60°F and 14.7 psia. This gas property is then defined as the actual volume occupied by a certain amount of gas at a specified pressure and temperature, divided by the volume occupied by the same amount of gas at standard conditions. In equation form, the relationship is expressed as Bg ¼ ðV Þp, T Vsc (3.51) where Bg ¼ gas formation volume factor, ft3/scf Vp,T ¼ volume of gas at pressure p and temperature T, ft3 Vsc ¼ volume of gas at standard conditions Applying the real gas equation of state, Eq. (3.11), and substituting for the volume V, gives ZnRT psc ZT p Bg ¼ ¼ Zsc nRTsc Tsc p psc where Zsc ¼ Z-factor at standard conditions ¼ 1.0 psc, Tsc ¼ standard pressure and temperature 222 CHAPTER 3 Natural Gas Properties Assuming that the standard conditions are represented by psc ¼ 14.7 psia and Tsc ¼ 520, the preceding expression can be reduced to the following relationship: Bg ¼ 0:02827 ZT p (3.52) where Bg ¼ gas formation volume factor, ft3/scf Z ¼ gas compressibility factor T ¼ temperature, °R In other field units, the gas formation volume factor can be expressed in bbl/ scf, to give Bg ¼ 0:005035 ZT p (3.53) Gas Expansion Factor The reciprocal of the gas formation volume factor, called the gas expansion factor, is designated by the symbol Eg: Eg ¼ 1 Bg In terms of scf/ft3, the gas expansion factor is p , scf=ft3 ZT (3.54) p , scf=bbl ZT (3.55) Eg ¼ 35:37 In other units, Eg ¼ 198:6 EXAMPLE 3.13 A gas well is producing at a rate of 15,000 ft3/day from a gas reservoir at an average pressure of 2000 psia and a temperature of 120°F. The specific gravity is 0.72. Calculate the gas flow rate in scf/day. Solution Step 1. Calculate the pseudocritical properties from Eqs. (3.17), (3.18), to give Tpc ¼ 168 + 325γ g 12:5γ 2g ¼ 395:5°R ppc ¼ 677 + 15:0γ g 37:5γ 2g ¼ 668:4psia Dranchuk-Purvis-Robinson Method 223 Step 2. Calculate the ppr and Tpr: ppr ¼ p 2000 ¼ ¼ 2:99 ppc 668:4 Tpr ¼ T 600 ¼ ¼ 1:52 Tpc 395:5 Step 3. Determine the Z-factor from Fig. 3.1: Z ¼ 0:78 Step 4. Calculate the gas expansion factor from Eq. (3.54): Eg ¼ 35:37 p 2000 ¼ 35:37 ¼ 151:15scf=ft3 ZT ð0:78Þð600Þ Step 5. Calculate the gas flow rate in scf/day by multiplying the gas flow rate (in ft3/day) by the gas expansion factor, Eg, as expressed in scf/ft3: Gas flow rate ¼ ð151:15Þð15, 000Þ ¼ 2:267MMscf=day It is also convenient in many engineering calculations to express the gas density in terms of the Bg or Eg, from the definitions of gas density, gas formation volume factor, and the gas expansion factor as given by Ma p R ZT psc ZT Bg ¼ Tsc P Tsc p Eg ¼ psc ZT ρg ¼ Combining the gas density equation with Bg and Eg equations gives ρg ¼ psc Ma 1 0:002635Ma ¼ Tsc R Bg Bg ρg ¼ psc Ma Eg ¼ 0:002635Ma Eg Tsc R Gas Viscosity The viscosity of a fluid is a measure of the internal fluid friction (resistance) to flow. If the friction between layers of the fluid is small, that is, low viscosity, an applied shearing force will result in a large velocity gradient. As the viscosity increases, each fluid layer exerts a larger frictional drag on the adjacent layers and velocity gradient decreases. 224 CHAPTER 3 Natural Gas Properties The viscosity of a fluid is generally defined as the ratio of the shear force per unit area to the local velocity gradient. Viscosities are expressed in terms of poises, centipoises, or micropoises. One poise equals a viscosity of 1 dyn-s/cm2 and can be converted to other field units by the following relationships: 1poise ¼ 100centipoises ¼ 1 106 micropoises ¼ 6:72 102 lbmass=ft s ¼ 20:9 103 lbf s=ft2 The gas viscosity is not commonly measured in the laboratory because it can be estimated precisely from empirical correlations. Like all intensive properties, viscosity of a natural gas is completely described by the following function: μg ¼ ðp, T, yi Þ where μg ¼ the viscosity of the gas phase. This relationship simply states that the viscosity is a function of pressure, temperature, and composition. Many of the widely used gas viscosity correlations may be viewed as modifications of that expression. Two popular methods that are commonly used in the petroleum industry are the Carr-Kobayashi-Burrows correlation and the Lee-Gonzalez-Eakin method, which are described next. CARR-KOBAYASHI-BURROWS'S METHOD Carr et al. (1954) developed graphical correlations for estimating the viscosity of natural gas as a function of temperature, pressure, and gas gravity. The computational procedure of applying the proposed correlations is summarized in the following steps. Step 1. Calculate the pseudocritical pressure, pseudocritical temperature, and apparent molecular weight from the specific gravity or the composition of the natural gas. Corrections to these pseudocritical properties for the presence of the nonhydrocarbon gases (CO2, N2, and H2S) should be made if they are present in concentration greater than 5 mol%. Step 2. Obtain the viscosity of the natural gas at one atmosphere and the temperature of interest from Fig. 3.4. This viscosity, as denoted by μ1, must be corrected for the presence of nonhydrocarbon components using the inserts of Fig. 3.4. The nonhydrocarbon fractions tend to increase the viscosity of the gas phase. The effect Carr-Kobayashi-Burrows's Method 225 n FIGURE 3.4 Carr et al.'s atmospheric gas viscosity correlation. (Carr, N., Kobayashi, R., Burrows, D., 1954. Viscosity of hydrocarbon gases under pressure. Trans. AIME 201, 270–275. Permission to publish from the Society of Petroleum Engineers of the AIME. # SPE-AIME.) 226 CHAPTER 3 Natural Gas Properties of nonhydrocarbon components on the viscosity of the natural gas can be expressed mathematically by the following relationship: μ1 ¼ ðμ1 Þuncorrected + ðΔμÞN2 + ðΔμÞCO2 + ðΔμÞH2 S (3.56) where μ1 ¼ “corrected” gas viscosity at 1 atm and reservoir temperature, cp ðΔμÞN2 ¼ viscosity corrections due to the presence of N2 ðΔμÞCO2 ¼ viscosity corrections due to the presence of CO2 ðΔμÞH2 S ¼ viscosity corrections due to the presence of H2S (μ1)uncorrected ¼ uncorrected gas viscosity, cp Step 3. Calculate the pseudoreduced pressure and temperature. Step 4. From the pseudoreduced temperature and pressure, obtain the viscosity ratio (μg/μ1) from Fig. 3.5. The term μg represents the viscosity of the gas at the required conditions. n FIGURE 3.5 Carr et al.'s viscosity ratio correlation. (Carr, N., Kobayashi, R., Burrows, D., 1954. Viscosity of hydrocarbon gases under pressure, Trans. AIME 201, 270–275. Permission to publish from the Society of Petroleum Engineers of the AIME. # SPE-AIME.) Carr-Kobayashi-Burrows's Method 227 Step 5. The gas viscosity, μg, at the pressure and temperature of interest, is calculated by multiplying the viscosity at 1 atm and system temperature, μ1, by the viscosity ratio. The following examples illustrate the use of the proposed graphical correlations. EXAMPLE 3.14 Using the data given in Example 3.13, calculate the viscosity of the gas. Solution Step 1. Calculate the apparent molecular weight of the gas: Ma ¼ ð0:72Þð28:96Þ ¼ 20:85 Step 2. Determine the viscosity of the gas at 1 atm and 140°F from Fig. 3.4: μ1 ¼ 0:0113 Step 3. Calculate ppr and Tpr: ppr ¼ 2:99 Tpr ¼ 1:52 Step 4. Determine the viscosity rates from Fig. 3.5: μg ¼1:5 μ1 Step 5. Solve for the viscosity of the natural gas: μg μg ¼ ðμ1 Þ ¼ ð1:5Þð0:0113Þ ¼ 0:01695cp μ1 Standing (1977) proposed a convenient mathematical expression for calculating the viscosity of the natural gas at atmospheric pressure and reservoir temperature, μ1. Standing also presented equations for describing the effects of N2, CO2, and H2S on μ1. The proposed relationships are μ1 ¼ ðμ1 Þuncorrected + ðΔμÞN2 + ðΔμÞCO2 + ðΔμÞH2 S ðμ1 Þuncorrected ¼ 8:118 103 6:15 103 log γ g + 1:709 105 2:062 106 γ g ðT 460Þ ðΔμÞN2 ¼ yN2 8:48 103 log γ g + 9:59 103 ðΔμÞCO2 ¼ yCO2 9:08 103 log γ g + 6:24 103 ðΔμÞH2 S ¼ yH2 S 8:49 103 log γ g + 3:73 103 (3.57) (3.58) (3.59) (3.60) (3.61) 228 CHAPTER 3 Natural Gas Properties where μ1 ¼ viscosity of the gas at atmospheric pressure and reservoir temperature, cp T ¼ reservoir temperature, °R γ g ¼ gas gravity; yN2 , yCO2 , yH2 S ¼ mole fraction of N2, CO2, and H2S, respectively Dempsey (1965) expressed the viscosity ratio μg/μ1 by the following relationship: ln Tpr μg ¼ a0 + a1 ppr + a2 p2pr + a3 p3pr + Tpr a1 + a5 ppr + a6 p2pr + a7 p3pr μ1 3 2 a8 + a9 ppr + a10 p2pr + a11 p3pr + Tpr a12 + a13 ppr + a14 p2pr + a15 p3pr + Tpr where Tpr ¼ pseudoreduced temperature of the gas mixture ppr ¼ pseudoreduced pressure of the gas mixture a0, …, a17 ¼ coefficients of the equations, as follows: a8 ¼ 7:93385648 101 a0 ¼ 2:46211820 a1 ¼ 2:970547414 1 a2 ¼ 2:86264054 10 a3 ¼ 8:05420522 103 a4 ¼ 2:80860949 a5 ¼ 3:49803305 a6 ¼ 3:60373020 101 a7 ¼ 1:044324 102 a9 ¼ 1:39643306 a10 ¼ 1:47144925 101 a11 ¼ 4:41015512 103 a12 ¼ 8:39387178 102 a13 ¼ 1:86408848 101 a14 ¼ 2:03367881 102 a15 ¼ 6:09579263 104 LEE-GONZALEZ-EAKIN'S METHOD Lee et al. (1966) presented a semiempirical relationship for calculating the viscosity of natural gases. The authors expressed the gas viscosity in terms of the reservoir temperature, gas density, and the molecular weight of the gas. Their proposed equation is given by μg ¼ 104 K exp X ρ Y g 62:4 (3.62) Lee-Gonzalez-Eakin's Method 229 where K¼ ð9:4 + 0:02Ma ÞT 1:5 209 + 19Ma + T (3.63) 986 + 0:01Ma T (3.64) X ¼ 3:5 + Y ¼ 2:4 0:2X (3.65) ρ ¼ gas density at reservoir pressure and temperature, lb/ft3 T ¼ reservoir temperature, °R Ma ¼ apparent molecular weight of the gas mixture The proposed correlation can predict viscosity values with a standard deviation of 2.7% and a maximum deviation of 8.99%. The correlation is less accurate for gases with higher specific gravities. The authors pointed out that the method cannot be used for sour gases. EXAMPLE 3.15 Rework Example 3.14 and calculate the gas viscosity by using the LeeGonzales-Eakin method. Solution Step 1. Calculate the gas density from Eq. (3.17): pMa ZRT ð2000Þð20:85Þ ¼ 8:3 lb=ft3 ρg ¼ ð10:73Þð600Þð0:78Þ ρg ¼ Step 2. Solve for the parameters K, X, and Y using Eqs. (3.63), (3.64), (3.65), respectively: K¼ ð9:4 + 0:02Ma ÞT 1:5 209 + 19Ma + T ½9:4 + 0:02ð20:85Þð600Þ1:5 ¼ 119:72 209 + 19ð20:85Þ + 600 986 + 0:01Ma X ¼ 3:5 + T 986 X ¼ 3:5 + + 0:01ð20:85Þ ¼ 5:35 600 Y ¼ 2:4 0:2X K¼ Y ¼ 2:4 0:2ð5:35Þ ¼ 1:33 230 CHAPTER 3 Natural Gas Properties Step 3. Calculate the viscosity from Eq. (3.62): μg ¼ 104 K exp X ρ Y g 62:4 " # 8:3 1:33 ¼ 0:0173cp μg ¼ 10 ð119:72Þexp 5:35 62:4 4 Specific Gravity Wet Gases The specific gravity of a wet gas, γ g, is described by the weighted average of the specific gravities of the separated gas from each separator. This weighted average approach is based on the separator gas/oil ratio: Xn γg ¼ γ sep i + Rst γ st Rsep i + Rst i¼1 R sep i¼1 Xn i (3.66) where n ¼ number of separators Rsep ¼ separator gas/oil ratio, scf/STB γ sep ¼ separator gas gravity Rst ¼ gas/oil ratio from the stock tank, scf/STB γ st ¼ gas gravity from the stock tank For wet gas reservoirs that produce liquid (condensate) at separator conditions, the produced gas mixtures normally exist as a “single” gas phase in the reservoir and production tubing. To determine the well-stream specific gravity, the produced gas and condensate (liquid) must be recombined in the correct ratio to find the average specific gravity of the “single-phase” gas reservoir. To calculate the specific gravity of the well-stream gas, let γ w ¼ well-stream gas gravity γ o ¼ condensate (oil) stock-tank gravity, 60°/60° γ g ¼ average surface gas gravity as defined by Eq. (3.66) Mo ¼ molecular weight of the stock-tank condensate (oil) rp ¼ producing oil/gas ratio (reciprocal of the gas/oil ratio, Rs), STB/scf The average specific gravity of the well stream is given by γw ¼ γ g + 4580rp γ o 1 + 133, 000rp ðγ o =Mo Þ (3.67) Lee-Gonzalez-Eakin's Method 231 In terms of gas/oil ratio, Rs, Eq. (3.67) can be expressed as γw ¼ γ g Rs + 4580γ o Rs + 133, 000γ o =Mo (3.68) Standing (1974) suggested the following correlation for estimating the molecular weight of the stock-tank condensate: Mo ¼ 6084 API 5:9 (3.69) where API is the API gravity of the liquid as given by API ¼ 141:5 131:5 γo Eq. (3.70) should be used only in the range 45° < API < 60°. Eilerts (1947) proposed an expression for the ratio γ o/Mo as a function of the condensate stock-tank API gravity: γ o =Mo ¼ 0:001892 + 7:35 105 API 4:52 108 ðAPIÞ2 (3.70) In retrograde and wet gas reservoirs calculations, it is convenient to express the produced separated gas as a fraction of the total system produced. This fraction fg can be expressed in terms of the separated moles of gas and liquid as fg ¼ ng ng ¼ nt ng + nl (3.71) where fg ¼ fraction of the separated gas produced in the entire system ng ¼ number of moles of the separated gas nl ¼ number of moles of the separated liquid nt ¼ total number of moles of the well stream For a total producing gas/oil ratio of Rs scf/STB, the equivalent number of moles of gas as described by Eq. (3.6) is ng ¼ Rs =379:4 (3.72) The number of moles of 1 STB of the separated condensate is given by no ¼ mass ðvolumeÞðdensityÞ ¼ molecular weight Mo or no ¼ ð1Þð5:615Þð62:4Þγ o 350:4γ o ¼ Mo Mo (3.73) 232 CHAPTER 3 Natural Gas Properties Substituting Eqs. (3.72), (3.73) into Eq. (3.71) gives fg ¼ Rs Rs + 133, 000ðγ o =Mo Þ (3.74) When applying the material balance equation for a gas reservoir, it assumes that a volume of gas in the reservoir will remain as a gas at surface conditions. When liquid is separated, the cumulative liquid volume must be converted into an equivalent gas volume, Veq, and added to the cumulative gas production for use in the material balance equation. If Np STB of liquid (condensate) has been produced, the equivalent number of moles of liquid is given by Eq. (3.73) as Np ð5:615Þð62:4Þγ o 350:4γ o Np no ¼ ¼ Mo Mo Expressing this number of moles of liquid as an equivalent gas volume at standard conditions by applying the ideal gas equation of state gives Veq ¼ 350:4γ o Np ð10:73Þð520Þ no RTsc ¼ psc Mo 14:7 Simplifying the above expression to give Veq, γ Np Veq ¼ 133,000 o Mo More conveniently, the equivalent gas volume can be expressed in scf/STB as Veq ¼ 133, 000 γo Mo where Veq ¼ equivalent gas volume, scf/STB Np ¼ cumulative, or daily, liquid volume, STB γ o ¼ specific gravity of the liquid, 60°/60° Mo ¼ molecular weight of the liquid EXAMPLE 3.16 The following data is available on a wet gas reservoir: n n n n n Initial reservoir pressure pi ¼ 3200 psia. Reservoir temperature T ¼ 200°F. Average porosity φ ¼ 18%. Average connate water saturation Swi ¼ 30%. Condensate daily flow rate Qo ¼ 400 STB/day. (3.75) Lee-Gonzalez-Eakin's Method 233 n n n n n API gravity of the condensate ¼ 50°. Daily separator gas rate Qgsep ¼ 4.20 MMscf/day. Separator gas specific gravity γ sep ¼ 0.65. Daily stock-tank gas rate Qgst ¼ 0.15 MMscf/day. Stock-tank gas specific gravity γ gst ¼ 1.05. Based on 1 ac-ft sand volume, calculate the initial oil (condensate) and gas in place and the daily well-stream gas flow rate in scf/day. Solution for initial oil and gas in place Step 1. Use Eq. (3.66) to calculate the specific gravity of the separated gas: Xn¼1 γg ¼ γg ¼ Rsep i γ sep i + Rst γ st Rsep 1 γ sep 1 + Rst γ st ¼ Xn¼1 Rsep 1 + Rst R + R sep st i i¼1 i¼1 ð4:2Þð0:65Þ + ð0:15Þð1:05Þ ¼ 0:664 4:2 + 0:15 Step 2. Calculate liquid specific gravity: γo ¼ 141:5 141:5 ¼ ¼ 0:780 API + 131:5 50 + 131:5 Step 3. Estimate the molecular weight of the liquid by applying Eq. (3.69): 6084 API 5:9 6084 Mo ¼ ¼ 138lb=lb mol 50 5:9 Mo ¼ Step 4. Calculate the producing gas/oil ratio: Qg ð4:2Þ 106 + ð0:15Þ 106 ¼ 10, 875scf=STB Rs ¼ ¼ Qo 400 Step 5. Calculate the well-stream specific gravity from Eq. (3.68): γw ¼ ð0:664Þð10, 875Þ + ð4, 580Þð0:78Þ ¼ 0:928 10, 875 + ð133, 000Þð0:78=138Þ Step 6. Solve for the pseudocritical properties of this wet gas using Eqs. (3.20), (3.21): Tpc ¼ 187 + 330ð0:928Þ 71:5ð0:928Þ2 ¼ 432°R ppc ¼ 706 51:7ð0:928Þ 11:1ð0:928Þ2 ¼ 648 psi Step 7. Calculate the pseudoreduced properties by applying Eqs. (3.12), (3.13): 3200 ¼ 4:9 648 200 + 460 ¼ 1:53 Tpr ¼ 432 ppr ¼ 234 CHAPTER 3 Natural Gas Properties Step 8. Determine the Z-factor from Fig. 3.1 to give Z ¼ 0:81 Step 9. Calculate the hydrocarbon space per acre-foot at reservoir conditions: VHy ¼ 43, 560ðAbÞφð1 Swi Þ, ft3 VHy ¼ 43, 560ð1Þð0:18Þð1 0:2Þ ¼ 6273ft3 =ac ft Step 10. Calculate total moles of gas per acre-foot at reservoir conditions by applying the real gas equation of state, Eq. (3.11): n¼ PV ð3200Þð6273Þ ¼ ¼ 3499mol=ac ft ZRT ð0:81Þð10:73Þð660Þ Step 11. Because 1 mol of gas occupies 379.4 scf (see Eq.( 3.6)) calculate the total gas initially in place per acre-foot, to give G ¼ ð379:4Þð3499Þ ¼ 1328 Mscf=ac ft If all this gas quantity is produced at the surface, a portion of it will be produced as stock-tank condensate. Step 12. Calculate the fraction of the total that is produced as gas at the surface from Eq. (3.75): fg ¼ 10,875 ¼ 0:935 10,875 + 133, 000ð0:78=138Þ Step 13. Calculate the initial gas and oil in place: Initial gas in place ¼ Gfg ¼ ð1328Þð0:935Þ ¼ 1242Mscf=ac ft G 1328 103 ¼ 122STB=ac ft Initial oil in place ¼ ¼ 10, 875 Rs Calculating Daily Well-Stream Gas Rate: Step 1. Calculate the equivalent gas volume from Eq. (3.75): γo Mo 0:78 Veq ¼ 133, 000 ¼ 752 scf=STB 138 Veq ¼ 133, 000 Step 2. Calculate the total well-stream gas flow rate: Qg ¼ Total Qg sep + Veq Qo Qg ¼ ð4:2 + 0:15Þ106 + ð752Þð400Þ ¼ 4:651MMscf=day Problems 235 PROBLEMS 1. Assuming an ideal gas behavior, calculate the density of n-butane at 200°F and 50 psia. Recalculate the density assuming real gas behavior. 2. Show that wi =Mi yi ¼ X ðwi =Mi Þ i 3. Given the following weight fractions of a gas: Component yi C1 C2 C3 n-C4 n-C5 0.60 0.17 0.13 0.06 0.04 Calculate a. Mole fraction of the gas. b. Apparent molecular weight. c. Specific gravity, and specific volume at 300 psia and 130°F, assuming an ideal gas behavior. d. Density at 300 psia and 130°F, assuming real gas behavior. e. Gas viscosity at 300 psia and 130°F. 4. An ideal gas mixture has a density of 2.15 lb/ft3 at 600 psia and 100°F. Calculate the apparent molecular weight of the gas mixture. 5. Using the gas composition as given in problem 3 and assuming real gas behavior, calculate a. Gas density at 2500 psia and 160°F. b. Specific volume at 2500 psia and 160°F. c. Gas formation volume factor in scf/ft3. d. Gas viscosity. 6. A natural gas with a specific gravity of 0.75 has a measured gas formation volume factor of 0.00529 ft3/scf at the current reservoir pressure and temperature. Calculate the density of the gas. 7. A natural gas has the following composition: Component yi C1 C2 C3 i-C4 n-C4 i-C5 n-C5 0.80 0.07 0.03 0.04 0.03 0.02 0.01 236 CHAPTER 3 Natural Gas Properties Reservoir conditions are 3500 psia and 200°F. Using, first, the CarrKobayashi-Burrows method, then the Lee-Gonzales-Eakin method, calculate a. Isothermal gas compressibility coefficient. b. Gas viscosity. 8. Given the following gas composition: Component yi CO2 N2 C1 C2 C3 n-C4 n-C5 0.06 0.003 0.80 0.05 0.03 0.02 0.01 If the reservoir pressure and temperature are 2500 psia and 175°F, respectively, calculate a. Gas density by accounting for the presence of nonhydrocarbon components using the Wichert-Aziz method, then the CarrKobayashi-Burrows method. b. Isothermal gas compressibility coefficient. c. Gas viscosity using the Carr-Kobayashi-Burrows method, then the Lee-Gonzales-Eakin method. 9. A high-pressure cell has a volume of 0.33 ft3 and contains gas at 2500 psia and 130°F, at which conditions its Z-factor is 0.75. When 43.6 scf of the gas are bled from the cell, the pressure drops to 1000 psia while the temperature remains at 130°F. What is the gas deviation factor at 1000 psia and 130°F? 10. A hydrocarbon gas mixture with a specific gravity of 0.65 has a density of 9 lb/ft3 at the prevailing reservoir pressure and temperature. Calculate the gas formation volume factor in bbl/scf. 11. A gas reservoir exists at a 150°F. The gas has the following composition: C1 ¼ 89mol% C2 ¼ 7mol% C3 ¼ 4mol% The gas expansion factory, Eg, was calculated as 210 scf/ft3 at the existing reservoir pressure and temperature. Calculate the viscosity of the gas. 12. A 20 ft3 tank at a pressure of 2500 psia and 200°F contains ethane gas. How many pounds of ethane are in the tank? References 237 13. A gas well is producing dry gas with a specific gravity of 0.60. The gas flow rate is 1.2 MMft3/day at a bottom-hole flowing pressure of 1200 psi and temperature of 140°F. Calculate the gas flow rate in MMscf/day. 14. The following data are available on a wet gas reservoir: Initial reservoir pressure, Pi ¼ 40,000 psia Reservoir temperature, T ¼ 150°F Average porosity, φ ¼ 15% Average connate water saturation, Swi ¼ 25% Condensate daily flow rate, Qo ¼ 350 STB/day API gravity of the condensate ¼ 52° Daily separator gas rate, Qgsep ¼ 3.80 MMscf/day Separator gas specific gravity, γ sep ¼ 0.71 Daily stock-tank gas rate, Qgst ¼ 0.21 MMscf/day Stock-tank gas specific gravity, γ gst ¼ 1.02 Based on 1 ac-ft sand volume, calculate a. Initial oil (condensate) and gas in place. b. Daily well-stream gas flow rate in scf/day. REFERENCES Brown, G.G., et al., 1948. Natural Gasoline and the Volatile Hydrocarbons. National Gas Association of America, Tulsa, OK. Carr, N., Kobayashi, R., Burrows, D., 1954. Viscosity of hydrocarbon gases under pressure. Trans. AIME 201, 270–275. Dempsey, J.R., 1965. Computer routine treats gas viscosity as a variable. Oil Gas J. 141–143. Dranchuk, P.M., Abu-Kassem, J.H., 1975. Calculate of Z-factors for natural gases using equations-of-state. J. Can. Pet. Technol. 34–36. Dranchuk, P.M., Purvis, R.A., Robinson, D.B., 1974. Computer Calculations of Natural Gas Compressibility Factors Using the Standing and Katz Correlation. Technical Series, no. IP 74–008. Institute of Petroleum, Alberta, Canada. Eilerts, C., 1947. The reservoir fluid, its composition and phase behavior. Oil Gas J. Hall, K.R., Yarborough, L., 1973. A new equation of state for Z-factor calculations. Oil Gas J. 82–92. Lee, A.L., Gonzalez, M.H., Eakin, B.E., 1966. The viscosity of natural gases. J. Petrol. Technol. 997–1000. Mattar, L.G., Brar, S., Aziz, K., 1975. Compressibility of natural gases. J. Can. Pet. Technol. 77–80. Matthews, T., Roland, C., Katz, D., 1942. High pressure gas measurement. Proc. Nat. Gas Assoc. AIME. Papay, J.A., 1985. Termelestechnologiai Parameterek Valtozasa a Gazlelepk Muvelese Soran. OGIL MUSZ, Tud, Kuzl. [Budapest], pp. 267–273. Standing, M., 1974. Petroleum Engineering Data Book. Norwegian Institute of Technology, Trondheim. Standing, M.B., 1977. Volumetric and Phase Behavior of Oil Field Hydrocarbon Systems. Society of Petroleum Engineers, Dallas. pp. 125–126. 238 CHAPTER 3 Natural Gas Properties Standing, M.B., Katz, D.L., 1942. Density of natural gases. Trans. AIME 146, 140–149. Stewart, W.F., Burkhard, S.F., Voo, D., 1959. Prediction of Pseudo-Critical Parameters for Mixtures. Paper Presented at the AIChE Meeting, Kansas City, MO. Sutton, R.P.M., 1985. Compressibility factors for high-molecular-weight reservoir gases. In: Paper SPE 14265, Presented at the 60th Annual Technical Conference and Exhibition of the Society of Petroleum Engineers, Las Vegas. Trube, A.S., 1957a. Compressibility of undersaturated hydrocarbon reservoir fluids. Trans. AIME 210, 341–344. Trube, A.S., 1957b. Compressibility of natural gases. Trans. AIME 210, 355–357. Whitson, C.H., Brule, M.R., 2000. Phase Behavior. Society of Petroleum Engineers, Richardson, TX. Wichert, E., Aziz, K., 1972. Calculation of Z’s for sour gases. Hydrocarb. Process. 51 (5), 119–122. Chapter 4 PVT Properties of Crude Oils PETROLEUM (an equivalent term is crude oil) is a complex mixture consisting predominantly of hydrocarbons and containing sulfur, nitrogen, oxygen, and helium as minor constituents. The physical and chemical properties of crude oils vary considerably and depend on the concentration of the various types of hydrocarbons and minor constituents present. An accurate description of physical properties of crude oils is of a considerable importance in the fields of both applied and theoretical science and especially in the solution of petroleum reservoir engineering problems. Physical properties of primary interest in petroleum engineering studies include 1. Fluid gravity. 2. Specific gravity of the solution gas. 3. Oil density. 4. Gas solubility. 5. Bubble point pressure. 6. Oil formation volume factor (FVF). 7. Isothermal compressibility coefficient of undersaturated crude oils. 8. Undersaturated oil properties. 9. Total FVF. 10. Crude oil viscosity. 11. Surface tension. Data on most of these fluid properties is usually determined by laboratory experiments performed on samples of actual reservoir fluids. In the absence of experimentally measured properties of crude oils, it is necessary for the petroleum engineer to determine the properties from empirically derived correlations. CRUDE OIL API GRAVITY The crude oil density is defined as the mass of a unit volume of the crude at a specified pressure and temperature. It is usually expressed in pounds per cubic foot. The specific gravity of a crude oil is defined as the ratio of Equations of State and PVT Analysis. http://dx.doi.org/10.1016/B978-0-12-801570-4.00004-0 Copyright # 2016 Elsevier Inc. All rights reserved. 239 240 CHAPTER 4 PVT Properties of Crude Oils the density of the oil to that of water. Both densities are measured at 60°F and atmospheric pressure: γo ¼ ρo ρw (4.1) where γ o ¼ specific gravity of the oil ρo ¼ density of the crude oil, lb/ft3 ρw ¼ density of the water, lb/ft3 It should be pointed out that the liquid specific gravity is dimensionless but traditionally is given the units 60°/60° to emphasize that both densities are measured at standard conditions (SC). The density of the water is approximately 62.4 lb/ft3, or γo ¼ ρo , 60°=60° 62:4 Although the density and specific gravity are used extensively in the petroleum industry, the API gravity is the preferred gravity scale. This gravity scale is precisely related to the specific gravity by the following expression: API ¼ 141:5 131:5 γo (4.2) This equation can be also rearranged to express the oil specific gravity in terms of the API gravity: γo ¼ 141:5 131:5 + API The API gravities of crude oils usually range from 47°API for the lighter crude oils to 10°API for the heavier asphaltic crude oils. EXAMPLE 4.1 Calculate the specific gravity and the API gravity of a crude oil system with a measured density of 53 lb/ft3 at SC. Solution Step 1. Calculate the specific gravity from Eq. (4.1): ρ γo ¼ o ρw γo ¼ 53 ¼ 0:849 62:4 Specific Gravity of the Solution Gas “γ g” 241 Step 2. Solve for the API gravity: API ¼ 141:5 131:5 γo API ¼ 141:5 131:5 ¼ 35:2°API 0:849 SPECIFIC GRAVITY OF THE SOLUTION GAS “γg” The specific gravity of the solution gas, γ g, is described by the weighted average of the specific gravities of the separated gas from each separator. This weighted average approach is based on the separator gas-oil ratio (GOR): Xn Rsep i γ sep i + Rst γ st i¼1 Xn γg ¼ Rsep i + Rst i¼1 (4.3) where n ¼ number of separators Rsep ¼ separator GOR, scf/STB γ sep ¼ separator gas gravity Rst ¼ GOR from the stock-tank, scf/STB γ st ¼ gas gravity from the stock-tank EXAMPLE 4.2 Separator tests were conducted on a crude oil sample. Results of the test in terms of the separator GOR and specific gravity of the separated gas are given in the table below. Calculate the specific gravity of the separated gas. Separator Pressure (psig) Temperature (°F) Gas-Oil Ratio (scf/STB) Gas Specific Gravity Primary Intermediate Stock-tank 660 75 0 150 110 60 724 202 58 0.743 0.956 1.296 Solution Calculate the solution specific gravity by using Eq. (4.3): Xn¼2 γg ¼ γg ¼ Rsep i γ sep i + Rst γ st Xn¼2 Rsep i + Rst i¼1 i¼1 ð724Þð0:743Þ + ð202Þð0:956Þ + ð58Þð1:296Þ ¼ 0:819 724 + 202 + 58 242 CHAPTER 4 PVT Properties of Crude Oils Ostermann et al. (1987) proposed a correlation to account for the increase of the solution gas gravity in a gas drive reservoir with decreasing the reservoir pressure, p, below the bubble point pressure, pb. The authors expressed the correlation in the following form: ( γ g ¼ γ gi a½1 ðp=pb Þb 1+ ðp=pb Þ ) Parameters a and b are functions of the bubble point pressure, given by the following two polynomials: a ¼ 0:12087 0:06897pb + 0:01461ðpb Þ2 b ¼ 1:04223 + 2:17073pb 0:68851ðpb Þ2 γ g ¼ specific gravity of the solution gas at pressure p γ gi ¼ initial specific gravity of the solution gas at the bubble point pressure pb pb ¼ bubble point pressure, psia p ¼ reservoir pressure, psia CRUDE OIL DENSITY “ρO” The crude oil density is defined as the mass of a unit volume of the crude at a specified pressure and temperature, mass/volume. The density usually is expressed in pounds per cubic foot and it varies from 30 lb/ft3 for light volatile oil to 60 lb/ft3 for heavy crude oil with little or no gas solubility. It is one of the most important oil properties, because its value substantially affects crude oil volume calculations. This vital oil property is traditionally measured in the laboratory as part of routine pressure-volume-temperature (PVT) tests. When laboratory crude oil density measurement is not available, correlations can be used to generate the required density data under reservoir pressure and temperature. Numerous empirical correlations for calculating the density of liquids have been proposed over the years. Based on the available limited measured data on the crude, the correlations can be divided into the following two categories: n n Correlations that use the crude oil composition to determine the saturated oil density and temperature. Correlations that use limited PVT data, such as gas gravity, oil gravity, and GOR, as correlating parameters. Crude Oil Density “ρo” 243 Density Correlations Based on the Oil Composition Several reliable methods are available for determining the density of saturated crude oil mixtures from their compositions. The best known and most widely used calculation methods are the following two: n n Standing-Katz (1942) Alani-Kennedy (1960) These two methods were developed to calculate the saturated oil density. Standing-Katz’s Method Standing and Katz (1942) proposed a graphical correlation for determining the density of hydrocarbon liquid mixtures. The authors developed the correlation from evaluating experimental, compositional, and density data on 15 crude oil samples containing up to 60 mol% methane. The proposed method yielded an average error of 1.2% and maximum error of 4% for the data on these crude oils. The original correlation had no procedure for handling significant amounts of nonhydrocarbons. The authors expressed the density of hydrocarbon liquid mixtures as a function of pressure and temperature by the following relationship: ρo ¼ ρsc + Δρp ΔρT (4.4) where ρo ¼ crude oil density at p and T, lb/ft3 ρsc ¼ crude oil density (with all the dissolved solution gas) at SC, that is, 14.7 psia and 60°F, lb/ft3 Δρp ¼ density correction for compressibility of oils, lb/ft3 ΔρT ¼ density correction for thermal expansion of oils, lb/ft3 Standing and Katz correlated graphically the liquid density at SC ρsc with the density of the propane-plus fraction, ρC03 + ; the weight percent of methane in the entire system, ðmC1 ÞC1 + ; and the weight percent of ethane in the ethane-plus, ðmC2 ÞC2 + . This graphical correlation is shown in Fig. 4.1. The proposed calculation procedure is performed on the basis of 1 lb-mol; that is, nt ¼ 1, of the hydrocarbon system. The following are the specific steps in the Standing and Katz procedure of calculating the liquid density at a specified pressure and temperature. Step 1. Calculate the total weight and the weight of each component in 1 lb-mol of the hydrocarbon mixture by applying the following relationships: 244 CHAPTER 4 PVT Properties of Crude Oils 70 20 30 10 e pl ane in e than 60 70 cen t eth per 0 40 60 10 e ur ixt 30 m al 50 t ne ha in to 20 et tm en 40 c er p ht g ei W 30 30 20 Density of system including methane and ethane (lb/ft3) 50 We ight Density of propane-plus (lb/ft3) us 40 0 50 10 n FIGURE 4.1 Standing and Katz density correlation. (Courtesy of the Gas Processors Suppliers Association, 1987. Engineering Data Book, 10th ed.) mi ¼ X xi Mi mt ¼ x i Mi where mi ¼ weight of component i in the mixture, lb/lb-mol xi ¼ mole fraction of component i in the mixture Mi ¼ molecular weight of component i mt ¼ total weight of 1 lb-mol of the mixture, lb/lb-mol Crude Oil Density “ρo” 245 Step 2. Calculate the weight percent of methane in the entire system and the weight percent of ethane in the ethane-plus from the following expressions: " # xC MC m C1 100 ðmC1 ÞC1 + ¼ Xn1 1 100 ¼ mt xM i¼1 i i (4.5) xC2 MC2 mC2 100 ¼ 100 mC2 + mt mC1 (4.6) and ðmC2 ÞC2 + ¼ where ðmC1 ÞC1 + ¼ weight percent of methane in the entire system mC1 ¼ weight of methane in 1 lb-mol of the mixture, that is, xC1 MC1 ðmC2 ÞC2 + ¼ weight percent of ethane in ethane-plus mC2 ¼ weight of ethane in 1 lb-mol of the mixture, that is, xC21 MC2 MC1 ¼ molecular weight of methane MC2 ¼ molecular weight of ethane Step 3. Calculate the density of the propane-plus fraction at SC by using the following equations: Xn xi Mi mC3 + mt mC1 mC2 ¼ Xn ρC3 + ¼ ¼ Xn i¼ C3 VC3 + ððxi Mi Þ=ρoi Þ ððxi Mi Þ=ρoi Þ i¼C3 (4.7) i¼C3 with mC3 + ¼ X xi Mi (4.8) i¼C3 VC3 + ¼ X i¼ C3 Vi ¼ X mi ρ i¼ C3 oi (4.9) where ρC3 + ¼ density of the propane and heavier components, lb/ft3 mC3 + ¼ weight of the propane and heavier components, lb/ft3 VC3 + ¼ volume of the propane-plus fraction, ft3/lb-mol Vi ¼ volume of component i in 1 lb-mol of the mixture mi ¼ weight of component i; that is, xiMi, lb/lb-mol ρoi ¼ Density of component i at SC, lb/ft3 (density values for pure components are tabulated in Table 1.1 in Chapter 1, but the density of the plus fraction must be measured) 246 CHAPTER 4 PVT Properties of Crude Oils Step 4. Using Fig. 4.1, enter the ρC3 + value into the left ordinate of the chart and move horizontally to the line representing ðmC2 ÞC0 then 2+ drop vertically to the line representing ðmC1 ÞC1 + . The density of the oil at SC is read on the right side of the chart. Standing (1977) expressed the graphical correlation in the following mathematical form: h i ρsc ¼ ρC2 + 1 0:012ðmC1 ÞC1 + 0:000158ðmC1 Þ2C1 + + 0:0133ðmC1 ÞC1 + + 0:00058ðmC1 Þ2C2 + (4.10) with h i ρC2 + ¼ ρC3 + 1 0:01386ðmC2 ÞC2 + 0:000082ðmC2 Þ2C2 + + 0:379ðmC2 ÞC2 + + 0:0042ðmC2 Þ2C2 + (4.11) where ρC2 + ¼ density of the ethane-plus fraction. Step 5. Correct the density at SC to the actual pressure by reading the additive pressure correction factor, Δρp, from Fig. 4.2: ρp, 60 ¼ ρsc + Δρp The pressure correction term Δρp can be expressed mathematically by Δρp ¼ 0:000167 + ð0:016181Þ100:0425ρsc p 8 10 0:299 + ð263Þ100:0603ρsc p2 (4.12) Step 6. Correct the density at 60°F and pressure to the actual temperature by reading the thermal expansion correction term, ΔρT, from Fig. 4.3: ρo ¼ ρp, 60 ΔρT The thermal expansion correction term, ΔρT, can be expressed mathematically by h 2:45 i ΔρT ¼ ðT 520Þ 0:0133 + 152:4 ρsc + Δρp h i ðT 520Þ2 8:1 106 ð0:0622Þ100:0764ðρsc + Δρp Þ (4.13) where T is the system temperature in °R. Crude Oil Density “ρo” 247 Density correction for compressibility 9 ia) s (p re su 00 es Pr 15,0 8 7 0 ,00 10 00 80 6 00 60 5 00 50 4 00 40 Density at pressure minus density at 60°F and 14.7 psia (lb/ft3) 10 3 30 00 2 20 00 1 0 25 1000 30 35 40 45 50 55 60 Density at 60°F and 14.7 psia (lb/ft3) n FIGURE 4.2 Density correction for compressibility of crude oils. (Courtesy of the Gas Processors Suppliers Association, 1987. Engineering Data Book, 10th ed.) EXAMPLE 4.3 A crude oil system has the following composition: Component xi C1 C2 C3 C4 C5 C6 C7+ 0.45 0.05 0.05 0.03 0.01 0.01 0.40 65 248 CHAPTER 4 PVT Properties of Crude Oils Density correction for thermal expansion 16 16 250 Density at 60°F minus density at temperature (lb/ft3) 14 12 14 600°F 200 12 550 10 10 180 500 450 8 8 160 400 350 6 140 6 300 120 4 250 100 4 200 2 2 80 100 60°F 0 60 0 40 20 40 -2 20 -60 -80 0 -4 -100 -20 -6 0°F -20 -40 -40 25 30 35 40 45 50 55 60 -2 -4 -6 Density at 60°F and pressure (lb/ft3) n FIGURE 4.3 Density correction for isothermal expansion of crude oils. (Courtesy of the Gas Processors Suppliers Association, 1987. Engineering Data Book, 10th ed.) If the molecular weight and specific gravity of the C7+ fractions are 215 and 0.87, respectively, calculate the density of the crude oil at 4000 psia and 160°F using the Standing and Katz method. Solution The table below shows the values of the components. mi 5 xiMi Component xi Mi C1 C2 C3 C4 C5 C6 C7+ 0.45 0.05 0.05 0.03 0.01 0.01 0.40 16.04 7.218 30.07 1.5035 44.09 2.2045 58.12 1.7436 72.15 0.7215 86.17 0.8617 215.0 86.00 P ¼ mt ¼ 100.253 a From Table 1.1. ρC7 + ¼ (0.87)(62.4) ¼ 54.288. b ρoi (lb/ft3)a – – 31.64 35.71 39.08 41.36 54.288b Vi 5 mi/ρoi – – 0.0697 0.0488 0.0185 0.0208 1.586 P ¼ VC3 + ¼ 1:7418 Crude Oil Density “ρo” 249 Step 1. Calculate the weight percent of C1 in the entire system and the weight percent of C2 in the ethane-plus fraction: " # xC1 MC1 m C1 100 ¼ 100 ðmC1 ÞC1 + ¼ Xn mt xM i¼1 i i 7:218 100 ¼ 7:2% 100:253 xC2 MC2 mC2 100 ¼ 100 ðmC2 ÞC2 + ¼ mC2 + mt mC1 1:5035 ðmC2 ÞC2 + ¼ 100 ¼ 1:616% 100:253 7:218 ðmC1 ÞC1 + ¼ Step 2. Calculate the density of the propane and heavier components: ρC3 + ¼ mC3 + mt mC1 mC2 ¼ Xn VC3 + ððxi Mi Þ=ρoi Þ i¼C3 ρC3 + ¼ 100:253 7:218 1:5035 ¼ 52:55lb=ft3 1:7418 Step 3. Determine the density of the oil at SC from Fig. 4.1: ρsc ¼ 47:5lb=ft3 Step 4. Correct for the pressure using Fig. 4.2: Δρp ¼ 1:18lb=ft3 Density of the oil at 4000 psia and 60°F is then calculated by the expression ρp, 60 ¼ ρsc + Δρp ¼ 47:5 + 1:18 ¼ 48:68lb=ft3 Step 5. From Fig. 4.3, determine the thermal expansion correction factor: ΔρT ¼ 2:45lb=ft3 Step 6. The required density at 4000 psia and 160°F is ρo ¼ ρp, 60 ΔρT ρo ¼ 48:68 2:45 ¼ 46:23lb=ft3 250 CHAPTER 4 PVT Properties of Crude Oils Alani-Kennedy’s Method Alani and Kennedy (1960) developed an equation to determine the molar liquid volume, Vm, of pure hydrocarbons over a wide range of temperatures and pressures. The equation then was adopted to apply to crude oils with heavy hydrocarbons expressed as a heptanes-plus fraction; that is, C7+. The Alani-Kennedy equation is similar in form to the Van der Waals equation, which takes the following form: Vm3 RT aVm ab ¼0 + b Vm2 + p p p (4.14) where R ¼ gas constant, 10.73 psia ft3/lb-mol, °R T ¼ temperature, °R p ¼ pressure, psia Vm ¼ molar volume, ft3/lb-mol a, b ¼ constants for pure substances Alani and Kennedy considered the constants a and b functions of temperature and proposed the expressions for calculating the two parameters: a ¼ Ken=T (4.15) b ¼ mT + c (4.16) where K, n, m, and c are constants for each pure component in the mixture and are tabulated in Table 4.1. The table does not contain constants from which the values of the parameters a and b for heptanes-plus can be calculated. Therefore, Alani and Kennedy proposed the following equations for determining a and b of C7+: M ln aC7 + ¼ 3:8405985 103 ðMÞC7 + 9:5638281 104 γ + C7 + 261:80818 T + 7:3104464 106 ðMÞ2C7 + + 10:753517 bC7 + ¼ 0:03499274ðMÞC7 + 7:2725403ðγ ÞC7 + + 2:232395 104 T M 0:016322572 γ + 6:2256545 C7 + Crude Oil Density “ρo” 251 Table 4.1 Alani and Kennedy Coefficients Components Temperature Range (°F) K n m × 104 C1 C1 C2 C2 C3 i-C4 n-C4 i-C5 n-C5 n-C6 H2Sa N2a CO2a 70–300 301–460 100–249 250–460 9160.6413 147.47333 46,709.57 17,495.34 20,247.76 32,204.42 33,016.21 37,046.23 37,046.23 52,093.01 13,200.00 4300 8166 61.893223 3247.4533 404.48844 34.163551 190.2442 131.63171 146.15445 299.6263 299.6263 254.56097 0 2.293 126 3.3162472 14.072637 5.1520981 2.8201736 2.1586448 3.3862284 2.902157 2.1954785 2.1954785 3.6961858 17.9 4.49 1.818 a Values for nonhydrocarbon components as proposed by Lohrenz et al. (1964). where MC7 + ¼ molecular weight of C7+ γ C7 + ¼ specific gravity of C7+ aC7 + , bC7 + ¼ constants of the heptanes-plus fraction T ¼ temperature in °R For hydrocarbon mixtures, the values of a and b of the mixture are calculated using the following mixing rules: am ¼ C7 + X ai xi (4.17) i¼1 bm ¼ C7 + X bi xi (4.18) i¼1 where the coefficients ai and bi refer to the values of pure hydrocarbon, i, as calculated from Eqs. (4.15), (4.16), at the existing temperature. xi is the mole fraction of component i in the liquid phase. The values am and bm are then used in Eq. (4.14) to solve for the molar volume Vm. The density of the mixture at the pressure and temperature of interest is determined from the following relationship: ρo ¼ Ma Vm (4.19) c 0.508743 1.832666 0.522397 0.623099 0.908325 1.101383 1.116814 1.436429 1.436429 1.592941 0.3945 0.3853 0.3872 252 CHAPTER 4 PVT Properties of Crude Oils where ρo ¼ density of the crude oil, lb/ft3 P Ma ¼ apparent molecular weight; that is, Ma ¼ xiMi Vm ¼ molar volume, ft3/lb-mol The Alani and Kennedy method for calculating the density of liquids is summarized in the following steps. Step 1. Calculate the constants a and b for each pure component from Eqs. (4.15), (4.16): a ¼ Ken=T b ¼ mT + c Step 2. Determine the C7+ parameters aC7 + and bC7 + . Step 3. Calculate values of the mixture coefficients am and bm from Eqs. (4.17), (4.18). Step 4. Calculate the molar volume Vm by solving Eq. (4.14) for the smallest real root. The equation can be solved iteratively by using the Newton-Raphson iterative method. In applying this technique, assume a starting value of Vm ¼ 2 and evaluate Eq. (4.14) using the assumed value: f ðVm Þ ¼ Vm3 RT am Vm am bm + bm Vm2 + p p p If the absolute value of this function is smaller than a preset tolerance, say, 1010, then the assumed value is the desired volume. If not, a new assumed value of (Vm)new is used to evaluate the function. The new value can be calculated from the following expression: f ðVm Þold ðVm Þnew ¼ ðVm Þold 0 f ðVm Þold where the derivative f´(Vm) is given by RT am f 0 ðVm Þ ¼ 3Vm2 2 + bm Vm + p p Step 5. Compute the apparent molecular weight Ma from Ma ¼ X xi Mi Step 6. Determine the density of the crude oil from ρo ¼ Ma Vm Crude Oil Density “ρo” 253 EXAMPLE 4.4 A crude oil system has the composition: Component xi CO2 N2 C1 C2 C3 i-C4 n-C4 i-C5 n-C5 C6 C7+ 0.0008 0.0164 0.2840 0.0716 0.1048 0.0420 0.0420 0.0191 0.0191 0.0405 0.3597 Given the following additional data: MC7 + ¼ 252 γ C7 + ¼ 0:8424 Pressure ¼ 1708.7 psia Temperature ¼ 591°R Calculate the density of the crude oil. Solution Step 1. Calculate the parameters aC7 + and bC7 + : M lnðaC7 + Þ ¼ 3:8405985 103 MC7 + 9:5638281 104 γ C7 + 261:80818 + 7:3104464 106 ðMC7 + Þ2 + 10:753517 + T 252 261:80818 + lnðaC7 + Þ ¼ 3:8405985 103 ð252Þ 9:5638281 104 0:8424 C7 + 591 6 2 + 7:3104464 10 ð252Þ + 10:753517 ¼ 12:34266 aC7 + ¼ exp ð12:34266Þ ¼ 229269:9 and bC7 + ¼ 0:03499274MC7 + 7:2725403ðγ ÞC7 + 4 M + 2:232395 10 T 0:016322572 γ + 6:2256545 C7 + bC7 + ¼ 0:03499274ð252Þ 7:2725403ð0:8424Þ 252 + 6:2256545 ¼ 4:165811 + 2:232395 104 591 0:016322572 0:8424 254 CHAPTER 4 PVT Properties of Crude Oils Step 2. Calculate the mixture parameters am and bm: am ¼ C7 + X ai xi ¼ 99111:71 i¼1 C7 + X bm ¼ bi xi ¼ 2:119383 i¼1 Step 3. Solve Eq. (4.14) iteratively for the molar volume: RT am Vm am bm Vm3 ¼0 + bm Vm2 + p p p 10:73ð591Þ 99111:71Vm 99111:71ð2:119383Þ Vm3 + 2:119383 Vm2 + ¼0 1708:7 1708:7 1708:7 Vm3 5:83064Vm2 + 58:004Vm 122:933 ¼ 0 Solving this expression iteratively for Vm as outlined previously gives Vm ¼ 2:528417 Step 4. Determine the apparent molecular weight of this mixture: X xi Mi ¼ 113:5102 Ma ¼ Step 5. Compute the density of the oil system: Ma Vm 113:5102 ρo ¼ ¼ 44:896lb=ft3 2:528417 ρo ¼ Density Correlations Based on Limited PVT Data Several empirical correlations for calculating the density of liquids of unknown compositional analysis have been proposed. The correlations employ limited PVT data such as gas gravity, oil gravity, and gas solubility as correlating parameters to estimate liquid density at the prevailing reservoir pressure and temperature. Two methods are presented as representatives of this category: n n Katz’s method Standing’s method Katz’s Method Density, in general, can be defined as the mass of a unit volume of material at a specified temperature and pressure. Accordingly, the density of a saturated crude oil (ie, with gas in solution) at SC can be defined mathematically by the following relationship: Crude Oil Density “ρo” 255 ρsc ¼ weight of stock tank oil + weight of solution gas volume of stock tank oil + increase in stock tank volume due to solution gas or ρsc ¼ mo + mg ðVo Þsc + ðΔVo Þsc where ρsc ¼ density of the oil at SC, lb/ft3 (Vo)sc ¼ volume of oil at SC, ft3/STB mo ¼ total weight of one stock-tank barrel of oil, lb/STB mg ¼ weight of the solution gas, lb/STB (ΔVo)sc ¼ increase in stock-tank oil volume due to solution gas, ft3/STB The procedure of calculating the density at SC is illustrated schematically in Fig. 4.4. Katz (1942) expressed the increase in the volume of oil, (ΔVo)sc, relative to the solution gas by introducing the apparent liquid density of the dissolved gas density, ρga, to the preceding expression as ðΔVo Þsc ¼ mg Ega Combining the preceding two expressions gives ρsc ¼ mo + mg mg ðVo Þsc + ρga where ρga, as introduced by Katz, is called the apparent density of the liquefied dissolved gas at 60°F and 14.7 psia. 14.7 psia 60°F (ΔVo)sc is the volume of the liquified Rs scf of solution gas 14.7 psia 60°F Rs, scf (VO)sc = 1 STB (ΔVo)sc (VO)sc = 1 STB n FIGURE 4.4 Schematic illustration of Katz’s density model at standard conditions. 256 CHAPTER 4 PVT Properties of Crude Oils Katz correlated the apparent gas density, in lb/ft3, graphically with gas specific gravity, γ g; gas solubility (solution GOR), Rs; and specific gravity (or the API gravity) of the stock-tank oil, γ o. This graphical correlation is presented in Fig. 4.5. The proposed method does not require the composition of the crude oil. Referring to the equation representing the surface density, that is, ρsc ¼ mo + mg mg ðVo Þsc + ρga the weights of the solution gas and the stock-tank can be determined in terms of the previous oil properties by the following relationships: mg ¼ Rs ð28:96Þ γ g , lb of solution gas=STB 379:4 mo ¼ ð5:615Þð62:4Þðγ o Þ, lb of oil=STB Substituting these terms into Katz’s equation gives Rs ð5:615Þð62:4Þðγ o Þ + ð28:96Þ γ g 379:4 ρsc ¼ Rs 5:615 + ð28:96Þ γ g =ρga 379:4 Apparent liquid density of dissolved gas At standard conditions (lb/ft3) 45 40 35 30 25 20 15 0.6 0.7 0.8 0.9 1.0 1.1 Gas gravity n FIGURE 4.5 Apparent liquid density of natural gas. 1.2 1.3 1.4 Crude Oil Density “ρo” 257 or Rs γ g 350:376γ o + 13:1 ! ρsc ¼ Rs γ g 5:615 + 13:1ρga (4.20) Standing (1981) showed that the apparent liquid density of the dissolved gas as represented by Katz’s chart can be closely approximated by the following relationship: ρga ¼ ð38:52Þ100:00326API + ½94:75 33:93log ðAPIÞ log γ g The pressure correction adjustment, Δρp, and the thermal expansion adjustment, ΔρT, for the calculated ρsc can be made using Figs. 4.2 and 4.3, respectively. EXAMPLE 4.5 A saturated crude oil exists at its bubble point pressure of 4000 psia and a reservoir temperature of 180°F. Given APIgravity ¼ 50° Rs ¼ 650scf=STB γ g ¼ 0:7 calculate the oil density at the specified pressure and temperature using Katz’s method. Solution Step 1. From Fig. 4.5 or the preceding equation, determine the apparent liquid density of dissolved gas: ρga ¼ ð38:52Þ100:00326API + ½94:75 33:93log ðAPIÞ log γ g ρga ¼ ð38:52Þ100:00326ð50ÞI + ½94:75 33:93 log ð50Þ log ð0:7Þ ¼ 20:7lb=ft3 Step 2. Calculate the stock-tank liquid gravity from Eq. (4.2): γo ¼ 141:5 141:5 ¼ ¼ 0:7796 API + 131:5 50 + 131:5 Step 3. Apply Eq. (4.20) to calculate the liquid density at SC: Rs γ g 350:376γ o + 13:1 ! ρsc ¼ Rs γ g 5:615 + 13:1ρga ð650Þð0:7Þ ð350:376Þð0:7796Þ + 13:1 ¼ 42:12 lb=ft3 ρsc ¼ ð650Þð0:7Þ 5:615 + ð13:1Þð20:7Þ 258 CHAPTER 4 PVT Properties of Crude Oils Step 4. Determine the pressure correction factor from Fig. 4.2: Δρp ¼ 1:55lb=ft3 Step 5. Adjust the oil density, as calculated at SC, to reservoir pressure: ρp, 60° ¼ ρsc + Δρp ρp, 60°F ¼ 42:12 + 1:55 ¼ 43:67lb=ft3 Step 6. Determine the isothermal adjustment factor from Fig. 4.3: ΔρT ¼ 3:25lb=ft3 Step 7. Calculate the oil density at 4000 psia and 180°F: ρo ¼ ρp, 60 ΔρT ρo ¼ 43:67 3:25 ¼ 40:42lb=ft3 Standing’s Method Standing (1981) proposed an empirical correlation for estimating the oil FVF as a function of the gas solubility, Rs; the specific gravity of stock-tank oil, γ o; the specific gravity of solution gas, γ g; and the system temperature, T. By coupling the mathematical definition of the oil FVF (as discussed in a later section) with Standing’s correlation, the density of a crude oil at a specified pressure and temperature can be calculated from the following expression: ρo ¼ 62:4γ o + 0:0136Rs γ g 1:175 γ g 0:5 0:972 + 0:000147 Rs γ + 1:25ðT 460Þ (4.21) o where T ¼ system temperature, °R γ o ¼ specific gravity of the stock-tank oil, 60°/60° γ g ¼ specific gravity of the gas Rs ¼ gas solubility, scf/STB ρo ¼ oil density, lb/ft3 EXAMPLE 4.6 Rework Example 4.5 and solve for the density using Standing’s correlation. Solution ρo ¼ 62:4γ o + 0:0136Rs γ g 1:175 γ g 0:5 0:972 + 0:000147 Rs γ + 1:25ðT 460Þ o 62:4ð0:7796Þ + 0:0136ð650Þð0:7Þ 3 ρo ¼ h i1:175 ¼ 39:92lb=ft 0:7 0:5 + 1:25ð180Þ 0:972 + 0:000147 650 0:7796 Crude Oil Density “ρo” 259 The obvious advantage of using Standing’s correlation is that it gives the density of the oil at the temperature and pressure at which the gas solubility is measured, and therefore, no further corrections are needed; that is, Δρp and ΔρT. Ahmed (1988) proposed another approach for estimating the crude oil density at SC based on the apparent molecular weight of the stock-tank oil. The correlation calculates the apparent molecular weight of the oil from the readily available PVT on the hydrocarbon system. Ahmed expressed the apparent molecular weight of the crude oil with its dissolved solution gas by the following relationship: Ma ¼ 0:0763Rs γ g Mst + 350:376γ o Mst 0:0026537Rs Mst + 350:376γ o where Ma ¼ apparent molecular weight of the oil Mst ¼ molecular weight of the stock-tank oil and can be taken as the molecular weight of the heptanes-plus fraction γ o ¼ specific gravity of the stock-tank oil or the C7+ fraction, 60°/60° The density of the oil at SC can then be determined from the expression ρsc ¼ 0:0763Rs γ g + 350:376γ o 199:71432 0:0026537Rs + γ o 5:615 + Mst If the molecular weight of the stock-tank oil is not available, the density of the oil with its dissolved solution gas at SC can be estimated from the following equation: ρsc ¼ 0:0763Rs γ g + 350:4γ o 0:0027Rs + 2:4893γ o + 3:491 The proposed approach requires the two additional steps of accounting for the effect of reservoir pressure and temperature. EXAMPLE 4.7 Using the data given in Example 4.5, calculate the density of the crude oil by using the preceding simplified expression. Solution Step 1. Calculate the density of the crude oil at SC from ρsc ¼ 0:0763Rs γ g + 350:4γ o 0:0027Rs + 2:4893γ o + 3:491 ρsc ¼ 0:0763ð650Þð0:7Þ + 350:4ð0:7796Þ ¼ 42:8lb=ft3 0:0027ð650Þ + 2:4893ð0:7796Þ + 3:491 Step 2. Determine Δρp from Fig. 4.2: Δρp ¼ 1:5lb=ft3 260 CHAPTER 4 PVT Properties of Crude Oils Step 3. Determine ΔρT from Fig. 4.3: ΔρT ¼ 3:6lb=ft3 Step 4. Calculate po at 4000 psia and 180°F: ρo ¼ ρsc + Δρp ΔρT ρo ¼ 42:8 + 1:5 3:6 ¼ 40:7lb=ft3 The oil density can also be calculated rigorously from the experimental measured PVT data at any specified pressure and temperature. The following expression (as derived later in this chapter) relates the oil density, ρo, to the gas solubility, Rs, specific gravity of the oil, gas gravity, and the oil FVF: ρo ¼ 62:4γ o + 0:0136Rs γ g Bo (4.22) where γ o ¼ specific gravity of the stock-tank oil, 60°/60° Rs ¼ gas solubility, scf/STB ρo ¼ oil density at pressure and temperature, lb/ft3 GAS SOLUBILITY “RS” The gas solubility, Rs, is defined as the number of standard cubic feet of gas that dissolve in one stock-tank barrel of crude oil at certain pressure and temperature. The solubility of a natural gas in a crude oil is a strong function of the pressure, the temperature, the API gravity, and the gas gravity. For a particular gas and crude oil to exist at a constant temperature, the solubility increases with pressure until the saturation pressure is reached. At the saturation pressure (bubble point pressure), all the available gases are dissolved in the oil and the gas solubility reaches its maximum value. Rather than measuring the amount of gas that dissolves in a given stocktank crude oil as the pressure is increased, it is customary to determine the amount of gas that comes out of a sample of reservoir crude oil as pressure decreases. A typical gas solubility curve, as a function of pressure for an undersaturated crude oil, is shown in Fig. 4.6. As the pressure is reduced from the initial reservoir pressure, pi, to the bubble point pressure, pb, no gas evolves from the oil and consequently the gas solubility remains constant at its maximum value of Rsb. Below the bubble point pressure, the solution gas is liberated and the value of Rs decreases with pressure. In the absence of experimentally measured gas solubility on a crude oil system, it is necessary to determine this property from empirically derived correlations. Five empirical correlations for estimating the gas solubility are Gas Solubility “Rs” 261 Rs = g g p 18.2 1.2048 + 1.4 10X X = 0.0125 API − 0.00091(T−460) Rsb Rs 0 pb 0 pi Pressure n FIGURE 4.6 Typical gas solubility-pressure relationship. n n n n n Standing’s correlation Vasquez-Beggs’s correlation Glaso’s correlation Marhoun’s correlation Petrosky-Farshad’s correlation Standing’s Correlation Standing (1947) proposed a graphical correlation for determining the gas solubility as a function of pressure, gas specific gravity, API gravity, and system temperature. The correlation was developed from a total of 105 experimentally determined data points on 22 hydrocarbon mixtures from California crude oils and natural gases. The proposed correlation has an average error of 4.8%. In a mathematical form, Standing (1981) expressed his proposed graphical correlation in the following more convenient mathematical form: Rs ¼ γ g h i1:2048 p + 1:4 10x 18:2 with x ¼ 0:0125API 0:00091ðT 460Þ where Rs ¼ gas solubility, scf/STB T ¼ temperature, °R (4.23) 262 CHAPTER 4 PVT Properties of Crude Oils p ¼ system pressure, psia γ g ¼ solution gas specific gravity API ¼ oil gravity, °API It should be noted that Standing’s equation is valid for applications at and below the bubble point pressure of the crude oil. EXAMPLE 4.8 The following table shows experimental PVT data on six crude oil systems. Results are based on two-stage surface separation. Using Standing’s correlation, estimate the gas solubility at the bubble point pressure and compare with the experimental value in terms of the absolute average error (AAE). T Oil (°F) pb Rs (scf/STB) Bo 1 250 2377 751 2 220 2620 768 3 260 2051 693 4 237 2884 968 5 218 3045 943 6 180 4239 807 ρo (lb/ft3) co at p > pb 6 1.528 38.13 22.14 10 at 2689 psi 0.474 40.95 18.75 106 at 2810 psi 1.529 37.37 22.69 106 at 2526 psi 1.619 38.92 21.51 106 at 2942 psi 1.570 37.70 24.16 106 at 3273 psi 1.385 46.79 11.45 106 at 4370 psi psep Tsep API γ g 150 60 47.1 0.851 100 75 40.7 0.855 100 72 48.6 0.911 60 120 40.5 0.898 200 60 44.2 0.781 85 173 27.3 0.848 T ¼ reservoir temperature, °F. pb ¼ bubble point pressure, psig. Bo ¼ oil FVF, bbl/STB. psep ¼ separator pressure, psig. Tsep ¼ separator temperature, °F. co ¼ isothermal compressibility coefficient of the oil at a specified pressure, psi1. Solution Apply Eq. (4.23) to determine the gas solubility: h p i1:2048 Rs ¼ γ g + 1:4 10x 18:2 with x ¼ 0:0125API 0:00091ðT 460Þ Results of the calculations are given in the table below. Gas Solubility “Rs” 263 Oil T (°F) API x 10x Predicted Rs, Eq. (4.23) Measured Rs % Error 1 2 3 4 5 6 250 220 260 237 218 180 47.1 40.7 48.6 40.5 44.2 27.3 0.361 0.309 0.371 0.291 0.322 0.177 2.297 2.035 2.349 1.952 2.097 1.505 838 817 774 914 1012 998 751 768 693 968 943 807 AAE 11.6 6.3 11.7 5.5 7.3 23.7 9.7% x ¼ 0.0125API 460). 1:2048 p 0.00091(T + 1:4 10x Rs ¼ γ g 18:2 . Vasquez-Beggs’s Correlation Vasquez and Beggs (1980) presented an improved empirical correlation for estimating Rs. The correlation was obtained by regression analysis using 5008 measured gas solubility data points. Based on oil gravity, the measured data were divided into two groups. This division was made at a value of oil gravity of 30°API. The proposed equation has the following form: API Rs ¼ C1 γ gs pC2 exp C3 T (4.24) Values for the coefficients are as follows: Coefficients API 30 API > 30 C1 C2 C3 0.362 1.0937 25.7240 0.0178 1.1870 23.931 Realizing that the value of the specific gravity of the gas depends on the conditions under which it is separated from the oil, Vasquez and Beggs proposed that the value of the gas specific gravity as obtained from a separator pressure of 100 psig be used in this equation. The reference pressure was chosen because it represents the average field separator conditions. The authors proposed the following relationship for adjustment of the gas gravity, γ g, to the reference separator pressure: h γ gs ¼ γ g 1 + 5:912 105 ðAPIÞ Tsep 460 log psep i 114:7 (4.25) 264 CHAPTER 4 PVT Properties of Crude Oils where γ gs ¼ gas gravity at the reference separator pressure γ g ¼ gas gravity at the actual separator conditions of psep and Tsep psep ¼ actual separator pressure, psia Tsep ¼ actual separator temperature, °R The gas gravity used to develop all the correlations reported by the authors was that which would result from a two-stage separation. The first-stage pressure was chosen as 100 psig and the second stage was the stock-tank. If the separator conditions are unknown, the unadjusted gas gravity may be used in Eq. (4.24). An independent evaluation of this correlation by Sutton and Farshad (1984) shows that the correlation is capable of predicting gas solubilities with an average absolute error of 12.7%. EXAMPLE 4.9 Using the PVT of the six crude oil systems of Example 4.8, solve for the gas solubility using the Vasquez and Beggs method. Solution The results are shown in the following table. Oil γ gs, Eq. (4.25) Predicted Rs, Eq. (4.24) Measured Rs % Error 1 2 3 4 5 6 0.8731 0.855 0.911 0.850 0.814 0.834 779 733 702 820 947 841 751 768 693 968 943 807 AAE 3.76 4.58 1.36 5.2 0.43 4.30 4.9% h psep i γ gs ¼ γ g 1 + 5:912 105 ðAPI Þ Tsep 460 log 114:7 N. API Rs ¼ C1 γ gs pC2 exp C3 . T Glaso’s Correlation Glaso (1980) proposed a correlation for estimating the gas solubility as a function of the API gravity, the pressure, the temperature, and the gas specific gravity. The correlation was developed from studying 45 North Sea crude oil samples. Glaso reported an average error of 1.28% with a standard deviation of 6.98%. The proposed relationship has the following form: Gas Solubility “Rs” 265 (" Rs ¼ γ g API0:989 ðT 460Þ # )1:2255 ðAÞ 0:172 (4.26) The parameter A is a pressure-dependent coefficient defined by the following expression: A ¼ 10 X with the exponent X as given by X ¼ 2:8869 ½14:1811 3:3093 log ðpÞ0:5 EXAMPLE 4.10 Rework Example 4.8 and solve for the gas solubility using Glaso’s correlation. Solution The results are shown in the table below. Oil p X A Predicted Rs, Eq. (4.26) Measured Rs % Error 1 2 3 4 5 6 2377 2620 2051 5884 3045 4239 1.155 1.196 1.095 1.237 1.260 1.413 14.286 15.687 12.450 17.243 18.210 25.883 737 714 686 843 868 842 751 768 693 968 943 807 AAE 1.84 6.92 0.90 12.92 7.94 4.34 5.8% X ¼ 2:8869 ½14:1811 3:3093 log ðpÞ0:5 N. A ¼ 10X . h i1:2255 API0:989 Rs ¼ γ g ðT 460 ð AÞ . Þ0:172 Marhoun’s Correlation Marhoun (1988) developed an expression for estimating the saturation pressure of the Middle Eastern crude oil systems. The correlation originates from 160 experimental saturation pressure data. The proposed correlation can be rearranged and solved for the gas solubility to give h ie Rs ¼ aγ bg γ co T d p where γ g ¼ gas specific gravity γ o ¼ stock-tank oil gravity (4.27) 266 CHAPTER 4 PVT Properties of Crude Oils T ¼ temperature, °R a–e ¼ coefficients of the above equation having these values a ¼ 185.843208 b ¼ 1.877840 c ¼ 3.1437 d ¼ 1.32657 e ¼ 1.39844 EXAMPLE 4.11 Resolve Example 4.8 using Marhoun’s correlation. Solution The results are shown in the table below. Oil T (°F) p Predicted Rs, Eq. (4.27) 1 2 3 4 5 6 250 220 260 237 218 180 2377 2620 2051 2884 3045 4239 740 792 729 1041 845 1186 h ie Rs ¼ aγ bg γ co T d p . Measured Rs % Error 751 768 693 968 943 807 AAE 1.43 3.09 5.21 7.55 10.37 47.03 12.4% a ¼ 185:843208. b ¼ 1:877840. c ¼ 3:1437. d ¼ 1:32657. e ¼ 1:39844. Petrosky-Farshad’s Correlation Petrosky and Farshad (1993) used nonlinear multiple regression software to develop a gas solubility correlation. The authors constructed a PVT database from 81 laboratory analyses from the Gulf of Mexico (GOM) crude oil system. Petrosky and Farshad proposed the following expression: Rs ¼ h i1:73184 p 10x + 12:340 γ 0:8439 g 112:727 with x ¼ 7:916 104 ðAPIÞ1:5410 4:561 105 ðT 460Þ1:3911 (4.28) Gas Solubility “Rs” 267 where Rs ¼ gas solubility, scf/STB T ¼ temperature, °R p ¼ system pressure, psia γ g ¼ solution gas specific gravity API ¼ oil gravity, °API EXAMPLE 4.12 Test the predictive capability of the Petrosky and Farshad equation by resolving Example 4.8. Solution The results are shown in the table below. Oil API T (°F) x Predicted Rs, Eq. (4.28) 1 2 3 4 5 6 47.1 40.7 48.6 40.5 44.2 27.3 250 220 260 237 218 180 0.2008 0.1566 0.2101 0.1579 0.1900 0.0667 772 726 758 875 865 900 x ¼ 7:916 104 ðAPIÞ1:5410 4:561 105 ðT 460Þ1:3911 . h i 1:73184 p Rs ¼ 112:727 + 12:340 γ 0:8439 10x . g Measured Rs % Error 751 768 693 968 943 807 AAE 2.86 5.46 9.32 9.57 8.28 11.57 7.84% As derived later in this chapter from the material balance approach, the gas solubility can also be calculated rigorously from the experimental, measured PVT data at the specified pressure and temperature. The following expression relates the gas solubility, Rs, to oil density, specific gravity of the oil, gas gravity, and the oil FVF: Rs ¼ Bo ρo 62:4γ o 0:0136γ g (4.29) where ρo ¼ oil density at pressure and temperature, lb/ft3 Bo ¼ oil FVF, bbl/STB γ o ¼ specific gravity of the stock-tank oil, 60°/60° γ g ¼ specific gravity of the solution gas McCain (1991) pointed out that the weight average of separator and stock-tank gas specific gravities should be used for γ g. The error in calculating Rs by using the preceding equation depends on the accuracy of only the available PVT data. 268 CHAPTER 4 PVT Properties of Crude Oils EXAMPLE 4.13 Using the data of Example 4.8, estimate Rs by applying Eq. (4.29). Solution The results are shown in the table below. Oil pb Bo ρo API γo γg Predicted Rs, Eq. (4.29) 1 2 3 4 5 6 2377 2620 2051 2884 3045 4239 1.528 1.474 1.529 1.619 1.570 1.385 38.13 40.95 37.37 38.92 37.70 46.79 47.1 40.7 48.6 40.5 44.2 27.3 0.792 0.822 0.786 0.823 0.805 0.891 0.851 0.855 0.911 0.898 0.781 0.848 762 781 655 956 843 798 Measured Rs % Error 751 768 693 968 943 807 AAE 1.53 1.73 5.51 1.23 10.57 1.13 3.61% 141:5 γ o ¼ 131:5 + API. ρo 62:4γ o Rs ¼ Bo0:0136γ . g BUBBLE POINT PRESSURE “pb” The bubble point pressure, pb, of a hydrocarbon system is defined as the highest pressure at which a bubble of gas is first liberated from the oil. This important property can be measured experimentally for a crude oil system by conducting a constant composition expansion (CCE) test. In the absence of the experimentally measured bubble point pressure, it is necessary for the engineer to make an estimate of this crude oil property from the readily available measured producing parameters. Several graphical and mathematical correlations for determining pb have been proposed during the last four decades. These correlations are essentially based on the assumption that the bubble point pressure is a strong function of gas solubility, Rs; gas gravity, γ g; oil gravity, API; and temperature “T”: pb ¼ f Rs , γ g , API, T Techniques of combining these parameters in a graphical form or a mathematical expression were proposed by several authors, including: n n n n n Standing Vasquez and Beggs Glaso Marhoun Petrosky and Farshad Bubble Point Pressure “pb” 269 The empirical correlations for estimating the bubble point pressure proposed by these authors are detailed below. Standing’s Correlation Based on 105 experimentally measured bubble point pressures on 22 hydrocarbon systems from California oil fields, Standing (1947) proposed a graphical correlation for determining the bubble point pressure of crude oil systems. The correlating parameters in the proposed correlation are the gas solubility, Rs; gas gravity, γ g; oil API gravity; and the system temperature. The reported average error is 4.8%. In a mathematical form, Standing (1981) expressed the graphical correlation by the following expression: h i 0:83 ð10Þa 1:4 pb ¼ 18:2 Rs =γ g (4.30) a ¼ 0:00091ðT 460Þ 0:0125ðAPIÞ (4.31) with where Rs ¼ gas solubility, scf/STB pb ¼ bubble point pressure, psia T ¼ system temperature, °R Standing’s correlation should be used with caution if nonhydrocarbon components are known to be present in the system. Lasater (1958) used a different approach in predicting the bubble point pressure by introducing and using the mole fraction of the solution gas, ygas, in the crude oil as a correlating parameter, defined by ygas ¼ Mo R s Mo Rs + 133, 000γ o where Mo ¼ molecular weight of the stock-tank oil and γ o ¼ specific gravity of the stock-tank oil, 60°/60°. If the molecular weight is not available, it can be estimated from Cragoe (1997) as Mo ¼ 6084 API 5:9 The bubble point pressure as proposed by Lasater is given by the following relationship: pb ¼ ! T A γg 270 CHAPTER 4 PVT Properties of Crude Oils where T is the temperature in °R and A is a graphical correlating parameter that is a function of the mole fraction of the solution gas, ygas. Whitson and Brule (2000) described this correlating parameter by A ¼ 0:83918 101:17664ygas y0:57246 , when ygas 0:6 gas 1:08000y 0:31109 gas A ¼ 0:83918 10 ygas , when ygas > 0:6 EXAMPLE 4.14 The experimental data given in Example 4.8 are repeated below for convenience. Predict the bubble point pressure using Standing’s correlation. Oil T (°F) pb Rs (scf/STB) Bo ρo (lb/ft3) co at p > pb psep Tsep API γg 1 250 2377 751 1.528 38.13 150 60 47.1 0.851 2 220 2620 768 0.474 40.95 100 75 40.7 0.855 3 260 2051 693 1.529 37.37 100 72 48.6 0.911 4 237 2884 968 1.619 38.92 60 120 40.5 0.898 5 218 3045 943 1.570 37.70 200 60 44.2 0.781 6 180 4239 807 1.385 46.79 22.14 106 at 2689 psi 18.75 106 at 2810 psi 22.69 106 at 2526 psi 21.51 106 at 2942 psi 24.16 106 at 3273 psi 11.45 106 at 4370 psi 85 173 27.3 0.848 Solution The results are shown in the following table. Oil API γg Coefficient a, Eq. (4.31) Predicted pb, Eq. (4.30) Measured pb % Error 1 2 3 4 5 6 47.1 40.7 48.6 40.5 44.2 27.3 0.851 0.855 0.911 0.898 0.781 0.848 0.3613 0.3086 0.3709 0.3115 0.3541 0.1775 2181 2503 1883 2896 2884 3561 2392 2635 2066 2899 3060 4254 AAE 8.8 5.0 8.8 0.1 5.7 16.3 7.4% a ¼ 0.00091(T 460) 0.0125(API). pb ¼ 18.2[(Rs /γ g)0.83(10)a 1.4]. McCain (1991) suggested that replacing the specific gravity of the gas in Eq. (4.30) with that of the separator gas, that is, excluding the gas from the stocktank would improve the accuracy of the equation. Bubble Point Pressure “pb” 271 EXAMPLE 4.15 Using the data from Example 4.14 and given the following first separator gas gravities as observed by McCain, estimate the bubble point pressure by applying Standing’s correlation. Oil First Separator Gas Gravity 1 2 3 4 5 6 0.755 0.786 0.801 0.888 0.705 0.813 Solution The results are shown in the table below. First Separation Coefficient a, Predicted pb, Measured % Oil API γ g Eq. (4.31) Eq. (4.30) pb Error 1 2 3 4 5 6 47.1 0.755 40.7 0.786 48.6 0.801 40.5 0.888 44.2 0.705 27.3 0.813 0.3613 0.3086 0.3709 0.3115 0.3541 0.1775 2411 2686 2098 2923 3143 3689 2392 2635 2066 2899 3060 4254 AAE 0.83 1.93 1.53 0.84 2.70 13.27 3.5% a ¼ 0.00091(T 460) 0.0125(API). pb ¼ 18.2[(Rs /γ g)0.83(10)a 1.4]. Vasquez-Beggs’s Correlation Vasquez and Beggs’s gas solubility correlation as presented by Eq. (4.24) can be solved for the bubble point pressure, pb, to give " pb ¼ ! # C2 Rs a ð10Þ C1 γ gs with exponent a as given by a ¼ C3 (4.32) API T where the temperature, T, is in °R. The coefficients C1, C2, and C3 of Eq. (4.32) have the following values: 272 CHAPTER 4 PVT Properties of Crude Oils Coefficient API 30 API > 30 C1 C2 C3 27.624 10.914328 11.172 56.18 0.84246 10.393 The gas specific gravity γ gs at the reference separator pressure is defined by Eq. (4.25) as h γ gs ¼ γ g 1 + 5:912 105 ðAPIÞ Tsep 460 log psep i 114:7 EXAMPLE 4.16 Rework Example 4.14 by applying Eq. (4.32). Solution The results are shown in the table below. Oil T Rs API γ gs, Eq. (4.25) a Predicted pb Measured pb % Error 1 2 3 4 5 6 250 220 260 237 218 180 751 768 693 968 943 807 47.1 40.7 48.6 40.5 44.2 27.3 0.873 0.855 0.911 0.850 0.814 0.834 0.689 0.622 0.702 0.625 0.678 0.477 2319 2741 2043 3331 3049 4093 2392 2635 2066 2899 3060 4254 AAE 3.07 4.03 1.14 14.91 0.36 3.78 4.5% a ¼ C3 pb ¼ h API T . C1 γRs ð10Þa gs iC2 . Glaso’s Correlation Glaso (1980) used 45 oil samples, mostly from the North Sea hydrocarbon system, to develop an accurate correlation for bubble point pressure prediction. Glaso proposed the following expression: log ðpb Þ ¼ 1:7669 + 1:7447log ðAÞ 0:30218½ log ðAÞ2 (4.33) The correlating parameter A in the equation is defined by the following expression: Bubble Point Pressure “pb” 273 A¼ Rs γg !0:816 ðT 460Þ0:172 (4.34) ðAPIÞ0:989 where Rs ¼ gas solubility, scf/STB T ¼ system temperature, °R γ g ¼ average specific gravity of the total surface gases For volatile oils, Glaso recommends that the temperature exponent of Eq. (4.34) be slightly changed from 0.172 to the value of 0.130, to give A¼ Rs γg !0:816 ðT 460Þ0:1302 ðAPIÞ0:989 EXAMPLE 4.17 Resolve Example 4.14 using Glaso’s correlation. Solution The results are shown in the table below. Oil T Rs API A, Eq. (3.34) pb, Eq. (4.33) Measured pb % Error 1 2 3 4 5 6 250 220 260 237 218 180 751 768 693 968 943 807 47.1 40.7 48.6 40.5 44.2 27.3 14.51 16.63 12.54 19.30 19.48 25.00 2431 2797 2083 3240 3269 4125 2392 2635 2066 2899 3060 4254 AAE 1.62 6.14 0.82 11.75 6.83 3.04 5.03% 0:816 ðT 460Þ0:172 ðAPIÞ0:989 . A¼ Rs γg log ðpb Þ ¼ 1:7669 + 1:7447 log ðAÞ 0:30218½ log ðAÞ2 . Marhoun’s Correlation Marhoun (1988) used 160 experimentally determined bubble point pressures from the PVT analysis of 69 Middle Eastern hydrocarbon mixtures to develop a correlation for estimating pb. The author correlated the bubble point pressure with the gas solubility, Rs, the temperature, T, and the specific gravity of the oil and the gas. Marhoun proposed the following expression: pb ¼ aRbs γ cg γ do T e (4.35) 274 CHAPTER 4 PVT Properties of Crude Oils where T ¼ temperature, °R Rs ¼ gas solubility, scf/STB γ o ¼ stock-tank oil specific gravity γ g ¼ gas specific gravity a–e ¼ coefficients of the correlation having the following values a ¼ 5.38088 103 b ¼ 0.715082 c ¼ 1.87784 d ¼ 3.1437 e ¼ 1.32657 The reported average absolute relative error for the correlation is 3.66% when compared with the experimental data used to develop the correlation. EXAMPLE 4.18 Using Eq. (4.35), rework Example 4.13. Solution The results are shown in the table below. Oil Predicted pb Measured pb % Error 1 2 3 4 5 6 2417 2578 1992 2752 3309 3229 2392 2635 2066 2899 3060 4254 1.03 2.16 3.57 5.07 8.14 24.09 AAE ¼ 7.3% pb ¼ 0:00538033R0:715082 γ 1:87784 γ 3:1437 T 1:32657 . s g o Petrosky-Farshad’s Correlation Petrosky and Farshad’s gas solubility equation, Eq. (4.28), can be solved for the bubble point pressure to give " # 112:727R0:577421 s 1391:051 pb ¼ γ 0:8439 ð10Þx g (4.36) where the correlating parameter, x, is as previously defined by the following expression: x ¼ 7:916 104 ðAPIÞ1:5410 4:561 105 ðT 460Þ1:3911 Oil FVF 275 where Rs ¼ gas solubility, scf/STB T ¼ temperature, °R p ¼ system pressure, psia γ g ¼ solution gas specific gravity API ¼ oil gravity, °API The authors concluded that the correlation predicts measured bubble point pressures with an average absolute error of 3.28%. EXAMPLE 4.19 Use the Petrosky and Farshad correlation to predict the bubble point pressure data given in Example 4.14. Solution The results are shown in the table below. Oil T Rs API x pb, Eq. (4.36) Measured pb % Error 1 2 3 4 5 6 250 220 260 237 218 180 751 768 693 968 943 807 47.1 40.7 48.6 40.5 44.2 27.3 0.2008 0.1566 0.2101 0.1579 0.1900 0.0667 2331 2768 1893 3156 3288 3908 2392 2635 2066 2899 3060 4254 AAE 2.55 5.04 8.39 8.86 7.44 8.13 6.74% x ¼ 7:916 104 ðAPIÞ1:5410 4:561 105 ðT 460Þ1:3911 . " # 112:727R0:577421 s 1391:051. pb ¼ x γ 0:8439 ð10Þ g Using the neural network approach, Al-Shammasi (1999) proposed the following expression that estimated the bubble point pressure with an average absolute error of 18%: 0:783716 e1:841408½γo γg Rs Tγ g pb ¼ γ 5:527215 o where T is the temperature in °R. OIL FVF The oil FVF, Bo, is defined as the ratio of the volume of oil (plus the gas in solution) at the prevailing reservoir temperature and pressure to the volume 276 CHAPTER 4 PVT Properties of Crude Oils of oil at SC. Evidently, Bo always is greater than or equal to unity. The oil FVF can be expressed mathematically as Bo ¼ ðVo Þp, T ðVo Þsc (4.37) where Bo ¼ oil FVF, bbl/STB (Vo)p,T ¼ volume of oil, in bbl, under reservoir pressure, p, and temperature, T (Vo)sc ¼ volume of oil is measured under SC, STB A typical oil formation factor curve, as a function of pressure for an undersaturated crude oil (pi > pb), is shown in Fig. 4.7. As the pressure is reduced below the initial reservoir pressure, pi, the oil volume increases due to the oil expansion. This behavior results in an increase in the oil FVF and continues until the bubble point pressure is reached. At pb, the oil reaches its maximum expansion and consequently attains a maximum value of Bob for the oil FVF. As the pressure is reduced below pb, volume of the oil and Bo are decreased as the solution gas is liberated. When the pressure is reduced to atmospheric pressure and the temperature to 60°F, the value of Bo is equal to 1. Most of the published empirical Bo correlations utilize the following generalized relationship: Bo ¼ f Rs , γ g , γ o , T 1.2 0.5 Bo 0.9759 0.000120 Rs g 1.25 T 460 o Bob Bo pb 1 0 Pressure n FIGURE 4.7 Typical oil formation volume factor-pressure relationship. Oil FVF 277 Six methods of predicting the oil FVF are presented, including: n n n n n n Standing’s correlation Vasquez and Beggs’s correlation Glaso’s correlation Marhoun’s correlation Petrosky and Farshad’s correlation The material balance equation It should be noted that all the correlations could be used for any pressure equal to or below the bubble point pressure. Standing’s Correlation Standing (1947) presented a graphical correlation for estimating the oil FVF with the gas solubility, gas gravity, oil gravity, and reservoir temperature as the correlating parameters. This graphical correlation originated from examining 105 experimental data points on 22 California hydrocarbon systems. An average error of 1.2% was reported for the correlation. Standing (1981) showed that the oil FVF can be expressed more conveniently in a mathematical form by the following equation: " γg γo #1:2 0:5 Bo ¼ 0:9759 + 0:000120 Rs + 1:25ðT 460Þ (4.38) where T ¼ temperature, °R γ o ¼ specific gravity of the stock-tank oil, 60°/60° γ g ¼ specific gravity of the solution gas Vasquez-Beggs’s Correlation Vasquez and Beggs (1980) developed a relationship for determining Bo as a function of Rs, γ o, γ g, and T. The proposed correlation was based on 6000 measurements of Bo at various pressures. Using the regression analysis technique, Vasquez and Beggs found the following equation to be the best form to reproduce the measured data: ! API Bo ¼ 1:0 + C1 Rs + ðT 520Þ ½C2 + C3 Rs γ gs (4.39) 278 CHAPTER 4 PVT Properties of Crude Oils where R ¼ gas solubility, scf/STB T ¼ temperature, °R γ gs ¼ gas specific gravity as defined by Eq. (4.25): h γ gs ¼ γ g 1 + 5:912 105 ðAPIÞ Tsep 460 log psep i 114:7 Values for the coefficients C1, C2, and C3 of Eq. (4.39) follow: Coefficient API 30 API > 30 4 4.677 10 1.751 105 1.811 108 C1 C2 C3 4.670 104 1.100 105 1.337 109 Vasquez and Beggs reported an average error of 4.7% for the proposed correlation. Glaso’s Correlation Glaso (1980) proposed the following expressions for calculating the oil FVF: Bo ¼ 1:0 + 10A (4.40) 2 A ¼ 6:58511 + 2:91329log B∗ob 0:27683 log B∗ob (4.41) where B∗ob is a “correlating number,” defined by the following equation: B∗ob ¼ Rs γg γo 0:526 + 0:968ðT 460Þ (4.42) where T ¼ temperature, °R, and γ o ¼ specific gravity of the stock-tank oil, 60°/60°. These correlations were originated from studying PVT data on 45 oil samples. The average error of the correlation was reported at 0.43% with a standard deviation of 2.18%. Sutton and Farshad (1984) concluded that Glaso’s correlation offers the best accuracy when compared with Standing’s and Vasquez-Beggs’s correlations. In general, Glaso’s correlation underpredicts FVF, Standing’s expression tends to overpredict oil FVFs, while Vasquez-Beggs’s correlation typically overpredicts the oil FVF. Oil FVF 279 Marhoun’s Correlation Marhoun (1988) developed a correlation for determining the oil FVF as a function of the gas solubility, stock-tank oil gravity, gas gravity, and temperature. The empirical equation was developed by use of the nonlinear multiple regression analysis on 160 experimental data points. The experimental data were obtained from 69 Middle Eastern oil reserves. The author proposed the following expression: Bo ¼ 0:497069 + 0:000862963T + 0:00182594F + 0:00000318099F2 (4.43) with the correlating parameter F as defined by the following equation: F ¼ Ras γ bg γ co (4.44) where T is the system temperature in °R and the coefficients a, b, and c have the following values: a ¼ 0:742390 b ¼ 0:323294 c ¼ 1:202040 Petrosky-Farshad’s Correlation Petrosky and Farshad (1993) proposed a new expression for estimating Bo. The proposed relationship is similar to the equation developed by Standing; however, the equation introduces three additional fitting parameters to increase the accuracy of the correlation. The authors used a nonlinear regression model to match experimental crude oil from the GOM hydrocarbon system. Their correlation has the following form: Bo ¼ 1:0113 + 7:2046 105 A (4.45) with the term A as given by " A¼ R0:3738 s γ 0:2914 g γ 0:6265 o ! #3:0936 + 0:24626ðT 460Þ 0:5371 where T ¼ temperature, °R, and γ o ¼ specific gravity of the stock-tank oil, 60°/60°. 280 CHAPTER 4 PVT Properties of Crude Oils Material Balance Equation From the definition of Bo as expressed mathematically by Eq. (4.37), Bo ¼ ðVo Þp, T ðVo Þsc The oil volume under p and T can be replaced with total weight of the hydrocarbon system divided by the density at the prevailing pressure and temperature: mt ρo Bo ¼ ðVo Þsc where the total weight of the hydrocarbon system is equal to the sum of the stock-tank oil plus the weight of the solution gas: mt ¼ mo + mg or Bo ¼ mo + mg ρo ðVo Þsc Given the gas solubility, Rs, per barrel of the stock-tank oil and the specific gravity of the solution gas, the weight of Rs scf of the gas is calculated as mg ¼ Rs ð28:96Þ γ g 379:4 where mg ¼ weight of solution gas, lb of solution gas/STB. The weight of one barrel of the stock-tank oil is calculated from its specific gravity by the following relationship: mo ¼ ð5:615Þð62:4Þðγ o Þ Substituting for mo and mg, Bo ¼ Rs ð28:96Þγ g 379:4 5:615ρo ð5:615Þð62:4Þγ o + or Bo ¼ 62:4Eo + 0:0136Rs γ g ρo (4.46) where ρo ¼ density of the oil at the specified pressure and temperature, lb/ft3. The error in calculating Bo by using Eq. (4.46) depends on the accuracy of only the input variables (Rs, γ g, and γ o) and the method of calculating ρo. Oil FVF 281 EXAMPLE 4.20 The table below shows experimental PVT data on six crude oil systems. Results are based on two-stage surface separation. Calculate the oil FVF at the bubble point pressure using the preceding six different correlations. Compare the results with the experimental values and calculate the AAE. Oil 1 2 3 4 5 6 T pb 250 220 260 237 218 180 Rs 2377 2620 2051 2884 3065 4239 ρo Bo 751 768 693 968 943 807 1.528 1.474 0.529 1.619 0.570 0.385 co at p > pb 6 38.13 40.95 37.37 38.92 37.70 46.79 22.14 10 at 2689 18.75 106 at 2810 22.69 106 at 2526 21.51 106 at 2942 24.16 106 at 3273 11.65 106 at 4370 0:5 #1:2 Solution For convenience, these six correlations follow: " Method 1 γg Bo ¼ 0:9759 + 0:000120 Rs γo + 1:25ðT 460Þ ! API ½C2 + C3 Rs Bo ¼ 1:0 + C1 Rs + ðT 520Þ γ gs Method 2 Method 3 Bo ¼ 1:0 + 10A Method 4 Bo ¼ 0:497069 + 0:000862963T + 0:00182594F + 0:00000318099F2 Bo ¼ 1:0113 + 7:2046 105 A Method 5 Method 6 Bo ¼ 62:4γ o + 0:0136Rs γ g ρo Results of applying these correlations for calculating Bo are tabulated in the table below. Method Crude Oil 1 2 3 4 5 Bo 1 2 3 4 5 6 1.528 1.474 1.529 1.619 1.570 1.506 1.487 1.495 1.618 1.571 1.474 1.450 1.451 1.542 1.546 1.473 1.459 1.461 1.589 1.541 1.516 1.477 1.511 1.575 1.554 1.552 1.508 1.556 1.632 1.584 1.525 1.470 1.542 1.623 1.599 psep Tsep API γg 150 100 100 60 200 85 60 75 72 120 60 173 47.1 40.7 48.6 40.5 44.2 27.3 0.851 0.855 0.911 0.898 0.781 0.848 282 CHAPTER 4 PVT Properties of Crude Oils 6 %AAE 1.385 – 1.461 1.7 1.389 2.8 1.438 2.8 1.414 1.3 1.433 1.8 1.387 0.6 Method 1 ¼ Standing’s correlation. Method 2 ¼ Vasquez and Beggs’s correlation. Method 3 ¼ Glaso’s correlation. Method 4 ¼ Marhoun’s correlation. Method 5 ¼ Petrosky and Farshad’s correlation. Method 6 ¼ Material balance equation. Al-Shammasi (1999) used the neural network approach to generate an expression for predicting Bo. The relationship, as given below, gives an average absolute error of 1.81%: Bo ¼ 1 + 5:53 107 ðT 520ÞRs + 0:000181ðRs =γ o Þ + ½0:000449ðT 520Þ=γ o + 0:000206Rs γ g =γ o where T is the temperature in °R. ISOTHERMAL COMPRESSIBILITY COEFFICIENT OF CRUDE OIL “co” The isothermal compressibility coefficient is defined as the rate of change in volume with respect to pressure increase per unit volume, all variables other than pressure being constant, including temperature. Mathematically, the isothermal compressibility, c, of a substance is defined by the following expression: c¼ 1 @V V @p T Isothermal compressibility coefficients are required in solving many reservoir engineering problems, including transient fluid flow problems; also, they are required in the determination of the physical properties of the undersaturated crude oil. For a crude oil system, the isothermal compressibility coefficient of the oil phase, co, is categorized into the following two types based on reservoir pressure: n At reservoir pressures that are greater than or equal to the bubble point pressure (p pb), the crude oil exists as a single phase with all its dissolved gas still in solution. The isothermal compressibility coefficient of the oil phase, co, above the bubble point reflects the changes in the volume associated with oil expansion or compression of the single-phase oil with changing the reservoir pressure. The oil compressibility in this case is termed undersaturated isothermal compressibility coefficient. Isothermal Compressibility Coefficient of Crude Oil “co” 283 n Below the bubble point pressure, the solution gas is liberated with decreasing reservoir pressure or redissolved with increasing the pressure. The changes of the oil volume as the result of changing the gas solubility must be considered when determining the isothermal compressibility coefficient. The oil compressibility in this case is termed saturated isothermal compressibility coefficient. Undersaturated Isothermal Compressibility Coefficient Generally, isothermal compressibility coefficients of an undersaturated crude oil are determined from a laboratory PVT study. A sample of the crude oil is placed in a PVT cell at the reservoir temperature and a pressure greater than the bubble point pressure of the crude oil. At these initial conditions, the reservoir fluid exists as a single-phase liquid. The volume of the oil is allowed to expand as its pressure declines. This volume is recorded and plotted as a function of pressure. If the experimental pressure-volume diagram for the oil is available, the instantaneous compressibility coefficient, co, at any pressure can be calculated by graphically determining the volume, V, and the corresponding slope, (@V/@p)T, at this pressure. At pressures above the bubble point, any of following equivalent expressions is valid in defining co: 1 @V co ¼ V @p T 1 @Bo co ¼ Bo @p T 1 @ρo co ¼ ρo @p T (4.47) where co ¼ isothermal compressibility of the crude oil, psi1 @V @p ¼ slope of the isothermal pressure=volume curve atp T ρo ¼ oil density, lb/ft3 Bo ¼ oil FVF, bbl/STB Craft and Hawkins (1959) introduced the cumulative or average isothermal compressibility coefficient, which defines the compressibility from the initial reservoir pressure to current reservoir pressure. Craft and Hawkins’s 284 CHAPTER 4 PVT Properties of Crude Oils average compressibility, as defined here, is used in all material balance calculations: ð pi V co ¼ co dp p V ½pi p ¼ V Vi V ½pi p where the subscript i represents initial condition. Equivalently, the average compressibility can be expressed in terms of Bo and ρo by the expressions Bo Boi Bo ½pi p ρ ρo co ¼ oi ρo ½pi p co ¼ The following four correlations for estimating the undersaturated oil compressibility are presented below: n n n n Trube’s correlation Vasquez-Beggs’s correlation Petrosky-Farshad’s correlation Standing’s correlation Trube’s Correlation Trube (1957) introduced the concept of the isothermal pseudoreduced compressibility, cr, of undersaturated crude oils as defined by the relationship cr ¼ co ppc (4.48) Trube correlated this property graphically with the pseudoreduced pressure and temperature, ppr and Tpr, as shown in Fig. 4.8. Additionally, Trube presented two graphical correlations, as shown in Figs. 4.9 and 4.10, to estimate the pseudocritical properties of crude oils. The calculation procedure of the proposed method is summarized in the following steps. Step 1. From the bottom-hole pressure measurements and pressure gradient data, calculate the average density of the undersaturated reservoir oil, in g/cm3, from the following expression: ðρo ÞT ¼ dp=dh 0:433 Isothermal Compressibility Coefficient of Crude Oil “co” 285 Ps eu do re du ce d 10 1. Pseudoreduced compressibility 0.1 0.01 1. 00 0. 95 te m pe ra tu re 0. 90 0. 80 0. 70 0. 60 0. 50 0.001 0.4 0 1 2 3 4 5 10 20 30 40 50 100 Pseudoreduced pressure n FIGURE 4.8 Trube’s pseudoreduced compressibility of undersaturated crude oil. (Permission to publish by the Society of Petroleum Engineers of AIME. Copyright SPE-AIME.) where (ρo)T ¼ oil density at reservoir pressure and temperature T, g/ cm3, and dp/dh ¼ pressure gradient as obtained from a pressure buildup test. Step 2. Adjust the calculated undersaturated oil density to its value at 60°F using the following equation: ðρo Þ60 ¼ ðρo ÞT ¼ 0:00046ðT 520Þ (4.49) where (ρo)60 ¼ adjusted undersaturated oil density to 60°F, g/cm3, and T ¼ reservoir temperature, °R. 286 CHAPTER 4 PVT Properties of Crude Oils 6000 5000 4000 0. 2000 66 0. 0.7 68 70 2 Specific gravity of reservoir liquid at 60°F. 0. 6 0.8 4 0.8 2 1000 900 800 700 600 0.8 0 0.8 8 0.7 6 0.7 4 0.7 Bubblepoint pressure at 60°F (psia) 3000 500 400 300 200 100 700 800 900 1000 1100 Pseudocritical temperature (°R) n FIGURE 4.9 Trube’s pseudocritical temperature correlation. (Permission to publish by the Society of Petroleum Engineers of AIME. Copyright SPE-AIME.) Step 3. Determine the bubble point pressure, pb, of the crude oil at reservoir temperature. If the bubble point pressure is not known, it can be estimated from Standing (1981), Eq. (4.30), written in a compacted form as 2 Rs pb ¼ 18:24 γg !0:83 100:00091ðT460Þ ° 100:0125 API 3 1:45 Step 4. Correct the calculated bubble point pressure, pb, at reservoir temperature to its value at 60°F using the following equation as proposed by Standing and Katz (1942): ðpb Þ60 ¼ 1:134pb 100:00091ðT460Þ (4.50) Isothermal Compressibility Coefficient of Crude Oil “co” 287 1300 1200 1000 Pseudocritical temperature 900 800 700 600 Pseudocritical pressure 500 Pseudocritical pressure Pseudocritical temperature (°R) 1100 400 300 200 0.62 0.64 0.66 0.68 0.70 0.72 0.74 0.76 0.78 0.80 0.82 0.84 0.86 0.88 Specific gravity of undersaturated crude oil at reservoir pressure corrected to 60°F n FIGURE 4.10 Trube’s pseudocritical properties correlation. (Permission to publish by the Society of Petroleum Engineers of AIME. Copyright SPE-AIME.) where (pb)60 ¼ bubble point pressure at 60°F, psi pb ¼ bubble point pressure at reservoir temperature, psia T ¼ reservoir temperature, °R Step 5. Enter in Fig. 4.9 the values of (pb)60 and (ρo)60 and determine the pseudocritical temperature, Tpc, of the crude. Step 6. Enter the value of Tpc in Fig. 4.10 and determine the pseudocritical pressure, ppc, of the crude. Step 7. Calculate the pseudoreduced pressure, ppr, and temperature, Tpr, from the following relationships: T Tpc p ppr ¼ ppc Tpr ¼ 288 CHAPTER 4 PVT Properties of Crude Oils Step 8. Determine cr by entering into Fig. 4.8 the values of Tpr and ppr. Step 9. Calculate co from Eq. (4.48): co ¼ cr ppc Trube did not specify the data used to develop the correlation nor did he allude to their accuracy, although the examples presented in his paper showed an average absolute error of 7.9% between calculated and measured values. Trube’s correlation can be best illustrated through the following example. EXAMPLE 4.21 Given the following data: Oil gravity ¼ 45° Gas solubility ¼ 600 scf/STB Solution gas gravity ¼ 0.8 Reservoir temperature ¼ 212°F Reservoir pressure ¼ 2000 psia Pressure gradient of reservoir liquid at 2000 psia and 212°F ¼ 0.032 psi/ft find co at 2000, 3000, and 4000 psia. Solution Step 1. Determine (ρo)T: ðρo ÞT ¼ dp=dh 0:32 ¼ ¼ 0:739g=cm3 0:433 0:433 Step 2. Correct the calculated oil specific gravity to its value at 60°F by applying Eq. (4.49): ðρo Þ60 ¼ ðρo ÞT + 0:00046ðT 520Þ ðρo Þ60 ¼ 0:739 + 0:00046ð152Þ ¼ 0:8089 Step 3. Calculate the bubble point pressure from Standing’s correlation (Eq. (4.40)): 2 Rs pb ¼ 18:24 γg 3 !0:83 100:00091ðT460Þ 1:45 100:0125°API " 600 pb ¼ 18:2 0:8 0:83 # 100:0091ðT212Þ 1:4 ¼ 1866psia 100:0125ð45Þ Isothermal Compressibility Coefficient of Crude Oil “co” 289 Step 4. Adjust pb to its value at 60°F by applying Eq. (4.50): ðpb Þ60 ¼ 1:134pb 100:00091 ðT 460Þ ðpb Þ60 ¼ 1:134ð1866Þ ¼ 1357:1psia 100:00091ð212Þ Step 5. Estimate the pseudocritical temperature, Tpc, of the crude oil from Fig. 4.9, to give Tpc ¼ 840°R Step 6. Estimate the pseudocritical pressure, ppc, of the crude oil from Fig. 4.10, to give ppc ¼ 500psia Step 7. Calculate the pseudoreduced temperature: Tpr ¼ 672 ¼ 0:8 840 Step 8. Calculate the pseudoreduced pressure, ppr, the pseudoreduced isothermal compressibility coefficient, cr, and tabulate values of the corresponding isothermal compressibility coefficient as shown in the table below. p ppr 5 p/500 cr co 5 cr/ppr1026 psia21 2000 3000 4000 4 6 8 0.0125 0.0089 0.0065 25.0 17.8 13.0 Vasquez-Beggs’s Correlation From a total of 4036 experimental data points used in a linear regression model, Vasquez and Beggs (1980) correlated the isothermal oil compressibility coefficients with Rs, T, °API, γ g, and p. They proposed the following expression: co ¼ 1433 + 5Rsb + 17:2ðT 460Þ 1180γ gs + 12:61°API 105 p where T ¼ temperature, °R p ¼ pressure above the bubble point pressure, psia Rsb ¼ gas solubility at the bubble point pressure γ gs ¼ corrected gas gravity as defined by Eq. (4.3) (4.51) 290 CHAPTER 4 PVT Properties of Crude Oils Petrosky-Farshad’s Correlation Petrosky and Farshad (1993) proposed a relationship for determining the oil compressibility for undersaturated hydrocarbon systems. The equation has the following form: co ¼ 1:705 107 R0:69357 γ 0:1885 API0:3272 ðT 460Þ0:6729 p0:5906 sb g (4.52) where T ¼ temperature, °R, and Rsb ¼ gas solubility at the bubble point pressure, scf/STB. Standing’s Correlation Standing (1974) proposed a graphical correlation for determining the oil compressibility for undersaturated hydrocarbon systems. Whitson and Brule expressed his relationship in the following mathematical form: 6 co ¼ 10 ρob + 0:004347ðp pb Þ 79:1 exp 0:0007141ðp pb Þ 12:938 (4.53) where ρob ¼ oil density at the bubble point pressure, lb/ft3 pb ¼ bubble point pressure, psia co ¼ oil compressibility, psia1 EXAMPLE 4.22 Using the experimental data given in Example 4.20, estimate the undersaturated oil compressibility coefficient using the Vasquez-Beggs, Petrosky-Farshad, and Standing correlations. The data given in Example 4.20 is reported in the following table for convenience. Oil 1 2 3 4 5 6 T 250 220 260 237 218 180 pb 2377 2620 2051 2884 3065 4239 Rs 751 768 693 968 943 807 ρo Bo 1.528 1.474 0.529 1.619 0.570 0.385 38.13 40.95 37.37 38.92 37.70 46.79 co at p > pb 6 22.14 10 at 2689 18.75 106 at 2810 22.69 106 at 2526 21.51 106 at 2942 24.16 106 at 3273 11.65 106 at 4370 Solution The results are shown in the table below. psep Tsep API γg 150 100 100 60 200 85 60 75 72 120 60 173 47.1 40.7 48.6 40.5 44.2 27.3 0.851 0.855 0.911 0.898 0.781 0.848 Isothermal Compressibility Coefficient of Crude Oil “co” 291 Measured co Standing’s Vasquez-Beggs Petrosky-Farshad Oil Pressure 1026 psi 1026 psi 1026 psi 1026 psi 1 2 3 4 5 6 2689 2810 2526 2942 3273 4370 AAE 22.14 18.75 22.60 21.51 24.16 11.45 2.0% Vasquez-Beggs : co ¼ 22.54 18.46 23.30 22.11 23.72 11.84 6.18% 22.88 20.16 23.78 22.31 20.16 11.54 4.05% 22.24 19.27 22.92 21.78 20.39 11.77 1433 + 5Rsb + 17:2ðT 460Þ 1180γ gs + 12:61°API 105 p : Petrosky-Farshad : co ¼ 1:705 107 R0:69357 γ 0:1885 API0:3272 ðT 460Þ0:6729 p0:5906 : g sb Standing’s : co ¼ 106 exp ρob + 0:004347ðp pb Þ 79:1 : 0:0007141ðp pb Þ 12:938 Saturated Isothermal Compressibility Coefficient At pressures below the bubble point pressure, the isothermal compressibility coefficient of the oil must be modified to account for the shrinkage associated with the liberation of the solution gas with decreasing reservoir pressure or swelling of the oil with repressurizing the reservoir and redissolving the solution gas. These changes in the oil volume as the result of changing the gas solubility must be considered when determining the isothermal compressibility coefficient of the oil compressibility. Below the bubble point pressure, co, is defined by the following expression: co ¼ 1@Bo Bg @Rs + Bo @p Bo @p (4.54) where Bg ¼ gas FVF, bbl/scf. Analytically, Standing’s correlations for Rs (Eq. (4.23)) and Bo (Eq. (4.38)) can be differentiated with respect to the pressure, p, to give @Rs Rs ¼ @p 0:83p + 21:75 γg @Bo 0:000144Rs ¼ @p 0:83p + 21:75 γ o 0:5 " γg Rs γo #0:2 0:5 + 1:25ðT 46Þ 292 CHAPTER 4 PVT Properties of Crude Oils These two expressions can be substituted into Eq. (4.54) to give the following relationship: ( ) 0:2 rffiffiffiffiffi rffiffiffiffiffi γg γg Rs 0:000144 Rs + 1:25ðT 460Þ co ¼ Bg Bo ð0:83p + 21:75Þ γo γo (4.55) where p ¼ pressure, psia T ¼ temperature, °R Bg ¼ gas FVF at pressure p, bbl/scf Rs ¼ gas solubility at pressure p, scf/STB Bo ¼ oil FVF at p, bbl/STB γ o ¼ specific gravity of the stock-tank oil γ g ¼ specific gravity of the solution gas McCain and coauthors (1988) correlated the oil compressibility with pressure, p, psia; oil API gravity; gas solubility at the bubble point, Rsb, in scf/ STB; and temperature, T, in °R. Their proposed relationship has the following form: co ¼ exp ðAÞ (4.56) where the correlating parameter A is given by the following expression: A ¼ 7:633 1:497 ln ðpÞ + 1:115 ln ðT Þ + 0:533 ln ðAPIÞ + 0:184 ln Rsp (4.57) The authors suggested that the accuracy of Eq. (4.56) can be substantially improved if the bubble point pressure is known. They improved correlating parameter, A, by including the bubble point pressure, pb, as one of the parameters in the preceding equation, to give A ¼ 7:573 1:45 ln ðpÞ 0:383 ln ðpb Þ + 1:402 ln ðT Þ + 0:256 ln ðAPIÞ + 0:449 ln Rsp (4.58) EXAMPLE 4.23 A crude oil system exists at 1650 psi and a temperature of 250°F. The system has the following PVT properties: API ¼ 47.1 pb ¼ 2377 psi γ g ¼ 0.851 γ gs ¼ 0.873 Rsb ¼ 751 scf/STB Bob ¼ 1.528 bbl/STB Isothermal Compressibility Coefficient of Crude Oil “co” 293 The laboratory-measured oil PVT data at 1650 psig follow: Bo ¼ 1.393 bbl/STB Rs ¼ 515 scf/STB Bg ¼ 0.001936 bbl/scf co 324.8 106 psi1 Estimate the oil compressibility using McCain’s correlation, then Eq. (4.55). Solution Calculate the correlating parameter A by applying Eq. (4.58): A ¼ 7:573 1:45 ln ðpÞ 0:383 ln ðpb Þ + 1:402 ln ðT Þ + 0:256 ln ðAPIÞ + 0:449 ln ðRsb ÞA ¼ 7:573 1:45 ln ð1665Þ 0:383 ln ð2392Þ + 1:402 ln ð710Þ + 0:256 ln ð47:1Þ + 0:449 ln ð451Þ ¼ 8:1445 Solve for co using Eq. (4.56): co ¼ exp ðAÞ co ¼ exp ð8:1445Þ ¼ 290:3 106 psi1 Solve for co using Eq. (4.55): ( ) 0:2 rffiffiffiffiffi rffiffiffiffiffi γg γg Rs co ¼ Bg 0:000144 Rs + 1:25ðT 460Þ Bo ð0:83p + 21:75Þ γo γo co ¼ 515 1:393ð0:83ð1665Þ + 21:75Þ 8 9 #0:2 rffiffiffiffiffiffiffiffiffiffiffi" rffiffiffiffiffiffiffiffiffiffiffi < = 0:851 0:851 0:001936 515 + 1:25ð250Þ 0:000144 : ; 0:792 0:792 co ¼ 358 106 psi1 It should be pointed out that, when it is necessary to establish PVT relationships for the hydrocarbon system through correlations or by extrapolation below the bubble point pressure, care should be exercised to see that the PVT functions are consistent. This consistency is assured if the increase in oil volume with increasing pressure is less than the decrease in volume associated with the gas going into solution. Because the oil compressibility coefficient, co, as expressed by Eq. (4.54) must be positive, that leads to the following consistency criteria: @Bo @Rs < Bg @p @p This consistency can easily be checked in the tabular form of PVT data. The PVT consistency errors most frequently occur at higher pressures, where the gas FVF, Bg, assumes relatively small values. 294 CHAPTER 4 PVT Properties of Crude Oils UNDERSATURATED OIL PROPERTIES Fig. 4.11 shows the volumetric behavior of the gas solubility, Rs, oil FVF, Bo, and oil density, ρo, as a function of pressure. As previously defined, the gas solubility, Rs, measures the tendency of the gas to dissolve in the oil at a particular temperature with increasing pressure. This solubility tendency, as measured in standard cubic feet of gas per one stock-tank barrel of oil, increases with pressure until the bubble point pressure is reached. With increasing of the gas solubility in the crude oil, the oil swells with the resulting increase in the oil FVF and an associated decrease in both its density and viscosity, as shown in Fig. 4.11. With increasing the pressure above the bubble point pressure, pb, the gas solubility remains constant with a maximum value that is denoted by Rsb (gas solubility at the bubble point pressure). However, as the pressure increases above pb, the crude oil system experiences a reduction in its volume (Vo)p,T during this isothermal compression. From the mathematical definition of Bo and ρo, ðVo Þp, T ðVo Þsc mass ρo ¼ ðVo Þp, T Bo ¼ These two expressions indicate that, with increasing the pressure above pb, the associated reduction in the volume of oil (Vo)p,T results in a decrease in Bo and an increase in ρo. Rs, Bo, ro, mo Rs Bo Bo = B ob exp [− mo ro pb 0 Pressure n FIGURE 4.11 Undersaturated fluid properties vs. pressure. co (p − pb )] Undersaturated Oil Properties 295 The magnitude of the changes in the oil volume above the bubble point pressure is a function of its isothermal compressibility coefficient co. To account for the effect of oil compression on the oil FVF and oil density, these two properties are first calculated at the bubble point pressure as Bob and ρob using any of the methods previously described and then adjusted to reflect the increase in p over pb. Undersaturated Oil FVF The isothermal compressibility coefficient [as expressed mathematically by Eq. (4.47)] can be equivalently written in terms of the oil FVF: co ¼ 1 @Vo Vo @p This relationship can be equivalently expressed in terms of Bo as 1 @ Vo =ðVo Þsc 1 @Bo co ¼ ¼ @p Vo =ðVo Þsc Bo @p (4.59) This relationship can be rearranged and integrated to produce ðp pb co dp ¼ ð Bo 1 Bob Bo dBo (4.60) Evaluating co at the arithmetic average pressure and concluding the integration procedure to give Bo ¼ Bob exp ½co ðp pb Þ (4.61) where Bo ¼ oil FVF at the pressure of interest, bbl/STB Bob ¼ oil FVF at the bubble point pressure, bbl/STB p ¼ pressure of interest, psia pb ¼ bubble point pressure, psia Replacing the isothermal compressibility coefficient in Eq. (4.60) with the Vasquez-Beggs’s co expression, Eq. (4.51), and integrating the resulting equation gives p Bo ¼ Bob exp A ln pb where A ¼ 105 1433 + 5Rsb + 17:2ðT 460Þ 1180γ gs + 12:61API (4.62) 296 CHAPTER 4 PVT Properties of Crude Oils Similarly, replacing co in Eq. (4.60) with the Petrosky and Farshad expression (Eq. (4.52)) and integrating gives Bo ¼ Bob exp A p0:4094 p0:4094 b (4.63) with the correlating parameter A as defined by A ¼ 4:1646 107 R0:69357 γ 0:1885 ðAPIÞ0:3272 ðT 460Þ0:6729 sb g (4.64) where T ¼ temperature, °R p ¼ pressure, psia Rsb ¼ gas solubility at the bubble point pressure EXAMPLE 4.24 Using the PVT data given in Example 4.23, calculate the oil FVF at 5000 psig using Eq. (4.62) then Eq. (4.64). The experimental measured Bo is 1.457 bbl/STB. Solution Using Eq. (4.62) Step 1. Calculate parameter A of Eq. (4.62): h i A ¼ 105 1433 + 5Rsb + 17:2ðT 460Þ 1180γ gs + 12:61API A ¼ 105 ½1433 + 5ð751Þ + 17:2ð250Þ 1180ð0:873Þ + 12:61ð47:1Þ ¼ 0:061858 Step 2. Apply Eq. (4.62): p Bo ¼ Bob exp A ln pb 5015 Bo ¼ 1:528 exp 0:061858 ln ¼ 1:459bbl=STB 2392 Solution Using Eq. (4.64) Step 1. Calculate the correlating parameter, A, from Eq. (4.64): γ 0:1885 ðAPIÞ0:3272 ðT 460Þ0:6729 A ¼ 4:1646 107 R0:69357 sb g A ¼ 4:1646 107 ð751Þ0:69357 ð0:851Þ0:1885 ð47:1Þ0:3272 ð250Þ0:6729 ¼ 0:005778 Step 2. Solve for Bo by applying Eq. (4.63): Bo ¼ Bob exp A p0:4094 p0:4094 b Bo ¼ ð1:528Þ exp 0:005778 50150:4094 23920:4096 ¼ 1:453bbl=STB Undersaturated Oil Density Oil density ρo is defined as mass, m, over volume, Vo, at any specified pressure and temperature, or the volume can be expressed as Undersaturated Oil Properties 297 Vo ¼ m ρo Differentiating density with respect to pressure gives @Vo @p ¼ T m @ρo ρ2o @p Substituting these two relations into Eq. (4.47) gives co ¼ 1 @Vo 1 m @ρo ¼ Vo @p ðm=ρo Þ ρ2o @p or, in terms of density, co ¼ 1 @ρo ρo @p Rearranging and integrating yields ðp ðp dρo pb pb ρo ρo co ðp pb Þ ¼ ln ρob co dp ¼ Solving for the density of the oil at pressures above the bubble point pressure, ρo ¼ ρob exp ½co ðp pb Þ (4.65) where ρo ¼ density of the oil at pressure p, lb/ft3 ρob ¼ density of the oil at the bubble point pressure, lb/ft3 co ¼ isothermal compressibility coefficient at average pressure, psi1 Vasquez-Beggs’s oil compressibility correlation and Petrosky-Farshad’s co expression can be incorporated in Eq. (4.65) to give the following. For Vasquez-Beggs’s co equation, p ρo ¼ ρob exp A ln pb (4.66) where A ¼ 105 1433 + 5Rsb + 17:2ðT 460Þ 1180γ gs + 12:61°API For Petrosky-Farshad’s co expression, ρo ¼ ρob exp A p0:4094 p0:4094 b with the correlating parameter A given by Eq. (4.64): A ¼ 4:1646 107 R0:69357 γ 0:1885 ðAPIÞ0:3272 ðT 460Þ0:6729 sb g (4.67) 298 CHAPTER 4 PVT Properties of Crude Oils TOTAL FVF “Bt” To describe the pressure-volume relationship of hydrocarbon systems below their bubble point pressure, it is convenient to express this relationship in terms of the total FVF as a function of pressure. This property defines the total volume of a system regardless of the number of phases present. The total FVF, denoted Bt, is defined as the ratio of the total volume of the hydrocarbon mixture; that is, oil and gas, if present, at the prevailing pressure and temperature per unit volume of the stock-tank oil. Because naturally occurring hydrocarbon systems usually exist in either one or two phases, the term two-phase FVF has become synonymous with the totalformation volume. Mathematically, Bt is defined by the following relationship: Bt ¼ ðVo Þp, T + Vg p, T ðVo Þsc (4.68) where Bt ¼ total FVF, bbl/STB (Vo)p,T ¼ volume of the oil at p and T, bbl (Vg)p,T ¼ volume of the liberated gas at p and T, bbl (Vo)sc ¼ volume of the oil at SC, STB Note that, when reservoir pressures are greater than or equal to the bubble point pressure, pb, no free gas will exist in the reservoir and, therefore, (Vg)p,T ¼ 0. At these conditions, Eq. (4.68) is reduced to the equation that describes the oil FVF: Bt ¼ ðVo Þp, T + 0 ðVo Þp, T ¼ ¼ Bo ðVo Þsc ðVo Þsc A typical plot of Bt as a function of pressure for an undersaturated crude oil is shown in Fig. 4.12. The oil FVF curve is also included in the illustration. As pointed out, Bo and Bt are identical at pressures above or equal to the bubble point pressure because only one phase, the oil phase, exists at these pressures. It should also be noted that at pressures below the bubble point pressure, the difference in the values of the two oil properties represents the volume of the evolved solution gas as measured at system conditions per stock-tank barrel of oil. Consider a crude oil sample placed in a PVT cell at its bubble point pressure, pb, and reservoir temperature, as shown schematically in Fig. 4.13. Assume that the volume of the oil sample is sufficient to yield one stock-tank barrel of oil at SC. Let Rsb represent the gas solubility at pb. By lowering the cell Total FVF “Bt” 299 Bo or Bt (Rsb – Rs) Bg Bt Bt = B o=B e ob xp [− co (p − pb )] Bo pb 0 Pressure n FIGURE 4.12 Bo and Bt vs. pressure. Constant composition expansion pb First bubble of gas p Evolved solution gas: (Vg)p,T = (Rsb–Rs) Bg Oil Rsb , Bob 14.7 psi Total volume of evolved gas at ST Rsb scf Bt Oil Rs , Bo Hg n FIGURE 4.13 The concept of the two-phase formation volume factor. Hg ST oil Bo= 1 STB Hg 300 CHAPTER 4 PVT Properties of Crude Oils pressure to p, a portion of the solution gas is evolved and occupies a certain volume of the PVT cell. Let Rs and Bo represent the corresponding gas solubility and oil FVF at p. Obviously, the term (Rsb Rs) represents the volume of the free gas as measured in scf per stock-tank barrel of the oil. The volume of the free gas at the cell conditions is then Vg p, T ¼ ðRsb Rs ÞBg where (Vg)p,T ¼ volume of the free gas at p and T, bbl of gas/STB of oil, and Bg ¼ gas FVF, bbl/scf. The volume of the remaining oil at the cell condition is ðV Þp, T ¼ Bo From the definition of the two-phase FVF, Bt ¼ Bo + ðRsb Rs ÞBg (4.69) where Rsb ¼ gas solubility at the bubble point pressure, scf/STB Rs ¼ gas solubility at any pressure, scf/STB Bo ¼ oil FVF at any pressure, bbl/STB Bg ¼ gas FVF, bbl/scf It is important to note from the previously described laboratory procedure that no gas or oil has been removed from the PVT cell with declining pressure; that is, no changes in the overall composition of the hydrocarbon mixture. As described later in this chapter, this is a property determined from the constant composition expansion test. It should be pointed out that the accuracy of estimating Bt by Eq. (4.69) depends largely on the accuracy of the other PVT parameters used in the equation: Bo, Rsb, Rs, and Bg. Several correlations can be used to estimate the two-phase FVF when the experimental data are not available; three of these methods are detailed below: n n n Standing’s correlation Glaso’s correlation Marhoun’s correlation Standing’s Correlation Standing (1947) used 387 experimental data points to develop a graphical correlation for predicting the two-phase FVF with a reported average error of 5%. The proposed correlation uses the following parameters for estimating the two-phase FVF: Total FVF “Bt” 301 n n n n n The gas solubility at pressure of interest, Rs. Solution gas gravity, γ g. Oil gravity, γ o, 60°/60°. Reservoir temperature, T. Pressure of interest, p. In developing his graphical correlation, Standing used a combined correlating parameter that is given by " log ðA∗Þ ¼ log Rs # ðT 460Þ0:5 ðγ o ÞC 96:8 10:1 0:3 6:604 + log ðpÞ γg (4.70) with the exponent C as defined by C ¼ ð2:9Þ100:00027Rs (4.71) Whitson and Brule (2000) expressed Standing’s graphical correlation by the following mathematical form: log ðBt Þ ¼ 5:223 47:4 log ðA∗ Þ 12:22 (4.72) Glaso’s Correlation Experimental data on 45 crude oil samples from the North Sea were used by Glaso (1980) in developing a generalized correlation for estimating Bt. Glaso modified Standing’s correlating parameter A* as given by Eq. (4.70) and used it in a regression analysis model to develop the following expression for Bt: log ðBt Þ ¼ 0:080135 + 0:47257log ðA∗Þ + 0:17351½ log ðA∗Þ2 (4.73) The author included the pressure in Standing’s correlating parameter A* to give: " # Rs ðT 460Þ0:5 ðγ o ÞC 1:1089 p A∗ ¼ 0:3 γg (4.74) with the exponent C as given by Eq. (4.71): C ¼ ð2:9Þ100:00027Rs Glaso reported a standard deviation of 6.54% for the total FVF correlation. 302 CHAPTER 4 PVT Properties of Crude Oils Marhoun’s Correlation Based on 1556 experimentally determined total FVFs, Marhoun (1988) used a nonlinear multiple regression model to develop a mathematical expression for Bt. The empirical equation has the following form: Bt ¼ 0:314693 + 0:106253 104 F + 0:18883 1010 F2 (4.75) with the correlating parameter F as given by F ¼ Ras γ bg γ co T d pe where T is the temperature in °R and the coefficients a–e as given below: a ¼ 0.644516 b ¼ 1.079340 c ¼ 0.724874 d ¼ 2.006210 e ¼ 0.761910 Marhoun reported an average absolute error of 4.11% with a standard deviation of 4.94% for the correlation. EXAMPLE 4.25 Given the following PVT data: pb ¼ 2744 psia T ¼ 600°R γ g ¼ 0.6744 Rs ¼ 444 scf/STB Rsb ¼ 603 scf/STB γ sb ¼ 0.843 60°/60° p ¼ 2000 psia Bo ¼ 1.1752 bbl/STB calculate Bt at 2000.7 psia using Eq. (4.69), Standing’s correlation, Glaso’s correlation, and Marhoun’s correlation. Solution Using Eq. (4.69) Step 1. Calculate Tpc and ppc of the solution gas from its specific gravity by applying Eqs. (3.18), (3.19): 2 Tpc ¼ 168 + 325γ g 12:5 γ g Tpc ¼ 168 + 325ð0:6744Þ 12:5ð0:6744Þ2 ¼ 381:49°R 2 ppc ¼ 677 + 15γ g 37:5 γ g ¼ 670:06psia Step 2. Calculate ppr and Tpr: ppr ¼ p 2000 ¼ ¼ 2:986 ppc 670:00 Tpr ¼ T 600 ¼ ¼ 1:57 Tpc 381:49 Total FVF “Bt” 303 Step 3. Determine the gas compressibility factor from Fig. 3.1: Z ¼ 0:81 Step 4. Calculate Bg from Eq. (3.54): Bg ¼ 0:00504 ð0:81Þð600Þ ¼ 0:001225bbl=scf 2000 Step 5. Solve for Bt by applying Eq. (4.69): Bt ¼ Bo + ðRsb Rs ÞBg Bt ¼ 1:1752 + 0:0001225ð603 444Þ ¼ 1:195bbl=STB Solution Using Standing’s Correlation Calculate the correlating parameters C and A* by applying Eqs. (4.71), (4.70), respectively: C ¼ ð2:9Þ100:00027Rs C ¼ ð2:9Þ100:00027ð444Þ ¼ 2:20 2 3 96:8 6 ðT 460Þ ðγ o Þ 7 log ðA∗Þ ¼ log 4Rs 10:1 5 0:3 6:604 + log ðpÞ γg " # ð140Þ0:5 ð0:843Þ2:2 96:8 ¼ 3:281 log ðA∗Þ ¼ log ð444Þ 10:1 0:3 6:604 + log ð2000Þ ð0:6744Þ 0:5 C Estimate Bt from Eq. (4.72): log ðBt Þ ¼ 5:223 47:4 log ðA∗ Þ 12:22 log ðBt Þ ¼ 5:223 47:4 ¼ 0:0792 3:281 12:22 to give Bt ¼ 100:0792 ¼ 1:200bbl=STB Solution Using Glaso’s Correlation Step 1. Determine the coefficient C from Eq. (4.71): C ¼ ð2:9Þ100:00027ð444Þ ¼ 2:2 Step 2. Calculate the correlating parameter A* from Eq. (4.74): " # Rs ðT 460Þ0:5 ðγ o ÞC 1:1089 p A∗ ¼ 0:3 γg " # ð444Þð140Þ0:5 ð0:843Þ2:2 20001:1089 ¼ 0:8873 A∗ ¼ ð0:6744Þ0:3 Step 3. Solve for Bt by applying Eq. (4.73) to yield 2 log ðBt Þ ¼ 0:080135 + 0:47257log A∗ + 0:17351 log A∗ log ðBt Þ ¼ 0:080135 + 0:47257log ð0:8873Þ + 0:17351½ log ð0:8873Þ2 ¼ 0:0561 304 CHAPTER 4 PVT Properties of Crude Oils to give Bt ¼ 100:0561 ¼ 1:138 Solution Using Marhoun’s Correlation Step 1. Determine the correlating parameter F of Eq. (4.75) to give F ¼ Ras γ bg γ co T d pe F ¼ 4440:644516 0:67441:07934 0:8430:724874 6002:00621 20000:76191 ¼ 78590:6789 Step 2. Solve for Bt from Eq. (4.75): Bt ¼ 0:314693 + 0:106253 104 F + 0:18883 1010 F2 Bt ¼ 0:314693 + 0:106253 104 ð78590:6789Þ + 0:18883 1010 ð78590:6789Þ2 Bt ¼ 1:2664bbl=STB CRUDE OIL VISCOSITY “μO” Crude oil viscosity is an important physical property that controls the flow of oil through porous media and pipes. The viscosity, in general, is defined as the internal resistance of the fluid to flow. It ranges from 0.1 cp for near critical to over 100 cp for heavy oil. It is considered the most difficult oil property to calculate with a reasonable accuracy from correlations. Oil’s viscosity is a strong function of the temperature, pressure, oil gravity, gas gravity, gas solubility, and composition of the crude oil. Whenever possible, oil viscosity should be determined by laboratory measurements at reservoir temperature and pressure. The viscosity usually is reported in standard PVT analyses. If such laboratory data are not available, engineers may refer to published correlations, which usually vary in complexity and accuracy, depending on the available data on the crude oil. Based on the available data on the oil mixture, correlations can be divided into the following two types: correlations based on other measured PVT data, such as API or Rs, and correlations based on oil composition. Depending on the pressure, p, the viscosity of crude oils can be classified into three categories: n n Dead oil viscosity, μod. The dead oil viscosity (oil with no gas in the solution) is defined as the viscosity of crude oil at atmospheric pressure and system temperature, T. Saturated oil viscosity, μob. The saturated (bubble point) oil viscosity is defined as the viscosity of the crude oil at any pressure less than or equal to the bubble point pressure. Crude Oil Viscosity “μo” 305 n Undersaturated oil viscosity, μo. The undersaturated oil viscosity is defined as the viscosity of the crude oil at a pressure above the bubble point and reservoir temperature. The definition of the three categories of oil viscosity is illustrated conceptually in Fig. 4.14. At atmospheric pressure and reservoir temperature, there is no dissolved gas in the oil (ie, Rs ¼ 0) and therefore the oil has its highest viscosity value of μod. As the pressure increases, the solubility of the gas increases accordingly, with the resulting decrease in the oil viscosity. The oil viscosity at any pressure pb is considered “saturated oil at this p.” As the pressure reaches the bubble point pressure, the amount of gas in solution reaches its maximum at Rsb and the oil viscosity at its minimum of μob. With increasing the pressure above pb, the viscosity of the undersaturated crude oil μo increases with pressure due to the compression of the oil. Based on these three categories, predicting the oil’s viscosity follows a similar three-step procedure: Step 1. Calculate the dead oil viscosity, μod, at the specified reservoir temperature and atmospheric pressure without dissolved gas: Rs ¼ 0. Step 2. Adjust the dead oil viscosity to any specified reservoir pressure (p pb) according to the gas solubility at p. Step 3. For pressures above the bubble point pressure, a further adjustment is made to μob to account for the compression of the oil above pb. Crude oil viscosity mod mod = 0.32 + A 1.8(107) 360 API1.53 T − 260 0.42 + A 10 8.33 ( API ) Rs Rs and mo mob ob a b od a = 10.715 (Rs + 100)–0.515 b = 5.44 (Rs + 150)–0.338 pb 0 Pressure n FIGURE 4.14 Oil viscosity as a function of Rs and p. 306 CHAPTER 4 PVT Properties of Crude Oils Viscosity Correlations as Correlated with PVT Data Dead Oil Correlations Several empirical methods are proposed to estimate the viscosity of the dead oil that are based on the API gravity of the crude oil and system temperature. These correlations, as listed below, calculate the viscosity of the dead oil (ie, no dissolved gas) at reservoir temperature “T ”: n n n Beal’s correlation Beggs-Robinson’s correlation Glaso’s correlation Discussion of the above listed three methods follows. Beal’s Correlation From a total of 753 values for dead oil viscosity at and above 100°F, Beal (1946) developed a graphical correlation for determining the viscosity of the dead oil as a function of temperature and the API gravity of the crude. Standing (1981) expressed the proposed graphical correlation in a mathematical relationship as follows: 18 107 360 μod ¼ 0:32 + API4:53 T 260 A (4.76) with A ¼ 100:42 + ð API Þ 8:33 where μod ¼ viscosity of the dead oil as measured at 14.7 psia and reservoir temperature, cp, and T ¼ temperature, °R. Beggs-Robinson’s Correlation Beggs and Robinson (1975) developed an empirical correlation for determining the viscosity of the dead oil. The correlation originated from analyzing 460 dead oil viscosity measurements. The proposed relationship is expressed mathematically as follows: μod ¼ 10AðT460Þ 1:163 1:0 (4.77) where A ¼ 103.03240.02023 API T ¼ temperature in °R An average error of 0.64% with a standard deviation of 13.53% was reported for the correlation when tested against the data used for its development. However, Sutton and Farshad (1984) reported an error of 114.3% when the correlation was tested against 93 cases from the literature. Crude Oil Viscosity “μo” 307 Glaso’s Correlation Glaso (1980) proposed a generalized mathematical relationship for computing the dead oil viscosity. The relationship was developed from experimental measurements on 26 crude oil samples. The correlation has the following form: μod ¼ 3:141 1010 ðT 460Þ3:444 ½ log ðAPIÞA (4.78) The temperature T is expressed in °R and the coefficient A is given by A ¼ 10:313½ log ðT 460Þ 36:447 This expression can be used within the range of 50–300°F for the system temperature and 20–48° for the API gravity of the crude. Sutton and Farshad (1984) concluded that Glaso’s correlation showed the best accuracy of the three previous correlations. Saturated Oil Viscosity Having calculated the viscosity of the dead crude oil at reservoir temperature using one of the three above correlations, the dead oil viscosity must be adjusted to account for the solution gas (ie, “Rs”). The following three empirical methods are proposed to estimate the viscosity of the saturated oil: n n n Chew-Connally correlation Beggs-Robinson correlation Abu-Khamsin and Al-Marhoun These correlations are discussed next. Chew-Connally Correlation Chew and Connally (1959) presented a graphical correlation to account for the reduction of the dead oil viscosity due to gas solubility. The proposed correlation was developed from 457 crude oil samples. Standing (1981) expressed the correlation in a mathematical form as follows: μob ¼ ð10a Þðμod Þb with a ¼ Rs 2:2 107 Rs 7:4 104 b ¼ 0:68ð10Þc + 0:25ð10Þd + 0:062ð10Þe c ¼ 0:0000862Rs d ¼ 0:0011Rs e ¼ 0:00374Rs (4.79) 308 CHAPTER 4 PVT Properties of Crude Oils where μob ¼ viscosity of the oil at the bubble point pressure, cp, and μod ¼ viscosity of the dead oil at 147.7 psia and reservoir temperature, cp. The experimental data used by Chew and Connally to develop their correlation encompassed the following ranges of values for the independent variables: Pressure: 132–5645 psia Temperature: 72–292°F Gas solubility: 51–3544 scf/STB Dead oil viscosity: 0.377–50 cp Beggs-Robinson Correlation From 2073 saturated oil viscosity measurements, Beggs and Robinson (1975) proposed an empirical correlation for estimating the saturated oil viscosity. The proposed mathematical expression has the following form: μob ¼ aðμod Þb (4.80) where a ¼ 10.715(Rs + 100)0515 b ¼ 5.44(Rs + 150)0338 The reported accuracy of the correlation is 1.83% with a standard deviation of 27.25%. The ranges of the data used to develop Beggs and Robinson’s equation are: Pressure: 132–5265 psia Temperature: 70–295°F API gravity: 16–58 Gas solubility: 20–2070 scf/STB Abu-Khamsin and Al-Marhoun An observation by Abu-Khamsin and Al-Marhoun (1991) suggests that the saturated oil viscosity, μob, correlates very well with the saturated oil density, ρob, as given by lnðμob Þ ¼ 8:484462ρ4ob 2:652294 where the saturated oil density, ρob, is expressed in g/cm3; that is, ρob/62.4. Viscosity of the Undersaturated Oil Oil viscosity at pressures above the bubble point is estimated by first calculating the oil viscosity at its bubble point pressure and adjusted to higher pressures. Three methods for estimating the oil viscosity at pressures above saturation pressure are presented next: Crude Oil Viscosity “μo” 309 n n n Beal’s correlation Khan’s correlation Vasquez-Beggs correlation Beal’s Correlation Based on 52 viscosity measurements, Beal (1946) presented a graphical correlation for the variation of the undersaturated oil viscosity with pressure where it has been curve fit by Standing (1981) by 0:56 μo ¼ μob + 0:001ðp pb Þ 0:024μ1:6 ob + 0:038μob (4.81) where μo ¼ undersaturated oil viscosity at pressure p and μob ¼ oil viscosity at the bubble point pressure, cp. The reported average error for Beal’s expression is 2.7%. Khan’s Correlation From 1500 experimental viscosity data points on Saudi Arabian crude oil systems, Khan et al. (1987) developed the following equation with a reported absolute average relative error of 2%: μo ¼ μob exp 9:6 106 ðp pb Þ (4.82) Vasquez-Beggs’s Correlation Vasquez and Beggs proposed a simple mathematical expression for estimating the viscosity of the oil above the bubble point pressure. From 3593 data points, Vasquez and Beggs (1980) proposed the following expression for estimating the viscosity of undersaturated crude oil: μo ¼ μob where p pb m (4.83) A m ¼ 2:6p1:187 10 A ¼ 3:9 105 p 5 The data used in developing the above correlation have the following ranges: Pressure: 141–9151 psia Gas solubility: 9.3–2199 scf/STB Viscosity: 0.117–148 cp Gas gravity: 0.511–1.351 API gravity: 15.3–59.5 The average error of the viscosity correlation is reported as 7.54%. 310 CHAPTER 4 PVT Properties of Crude Oils EXAMPLE 4.26 The experimental PVT data given in Example 4.22 are repeated below. In addition to this information, viscosity data are presented in the next table. Using all the oil viscosity correlations discussed in this chapter, calculate μod, μob, and the viscosity of the undersaturated oil. Oil 1 2 3 4 5 6 T 250 220 260 237 218 180 pb 2377 2620 2051 2884 3065 4239 Rs 751 768 693 968 943 807 ρo Bo 1.528 1.474 0.529 1.619 0.57 0.385 co at p > pb 6 22.14 10 at 2689 18.75 106 at 2810 22.69 106 at 2526 21.51 106 at 2942 24.16 106 at 3273 11.65 106 at 4370 38.13 40.95 37.37 38.92 37.7 46.79 psep Tsep API γg 150 100 100 60 200 85 60 75 72 120 60 173 47.1 40.7 48.6 40.5 44.2 27.3 0.851 0.855 0.911 0.898 0.781 0.848 Oil Dead Oil, μod, @ T Saturated Oil, μob, cp Undersaturated Oil, μo @ p 1 2 3 4 5 6 0.765 @ 250°F 1.286 @ 220°F 0.686 @ 260°F 1.014 @ 237°F 1.009 @ 218°F 4.166 @ 180°F 0.224 0.373 0.221 0.377 0.305 0.950 0.281 @ 5000 psi 0.450 @ 5000 psi 0.292 @ 5000 psi 0.414 @ 6000 psi 0.394 @ 6000 psi 1.008 @ 5000 psi Solution Dead Oil Viscosity The results are shown in the table below. Oil Measured μod Beal Beggs-Robinson Glaso 1 2 3 4 5 6 AAE 0.765 1.286 0.686 1.014 1.009 4.166 0.322 0.638 0.275 0.545 0.512 4.425 44.9% 0.568 1.020 0.493 0.917 0.829 4.246 17.32% 0.417 0.775 0.363 0.714 0.598 4.536 35.26% Beal : μod ¼ 0:32 + 1:8 107 360 T 260 API4:53 1:163 A : Beggs-Robinson : μod ¼ 10AðT460Þ 1:0: 10 3:444 Glaso : μod ¼ 3:141 10 ðT 460Þ ½ log ðAPIÞA : Crude Oil Viscosity “μo” 311 Saturated Oil Viscosity The results are shown in the table below. Oil 1 2 3 4 5 6 AAE Measured ρob (g/cm3) Measured μob ChewConnally BeggsRobinson AbuKhamsin and Al-Marhoun 0.6111 0.6563 0.5989 0.6237 0.6042 0.7498 0.224 0.373 0.221 0.377 0.305 0.950 0.313a 0.426 0.308 0.311 0.316 0.842 21% 0.287a 0.377 0.279 0.297 0.300 0.689 17% 0.230 0.340 0.210 0.255 0.218 1.030 14% Chew-Connally: μob ¼ (10)a (μod)b. Beggs-Robinson: μob ¼ a(μod)b. Abu-Khamsin and Al-Marhoun: lnðμob Þ ¼ 8:484462ρ4ob 2:652294. a Using the measured μod. Undersaturated Oil Viscosity The results are shown in the table below. Oil 1 2 3 4 5 6 AAE Measured μo Beal Vasquez-Beggs a 0.281 0.45 0.292 0.414 0.396 1.008 0.273 0.437 0.275 0.434 0.373 0.945 3.8% 0.303a 0.485 0.318 0.472 0.417 1.016 7.5% 0:56 Beals: μo ¼ μob + 0:001ðp pb Þ 0:024μ1:6 ob + 0:038μob . Vasquez-Beggs: μo ¼ μob ppb m . Using the measured μob. a Viscosity Correlations From Oil Composition Like all intensive properties, viscosity is a strong function of the pressure, temperature, and composition of the oil, xi. In general, this relationship can be expressed mathematically by the following function: μo ¼ f ðp, T, x1 , …, xn Þ In compositional reservoir simulation of miscible gas injection, the compositions of the reservoir gases and oils are calculated based on the equationof-state approach and used to perform the necessary volumetric behavior of 312 CHAPTER 4 PVT Properties of Crude Oils the reservoir hydrocarbon system. The calculation of the viscosities of these fluids from their compositions is required for a true and complete compositional material balance. Two empirical correlations of calculating the viscosity of crude oils from their compositions are widely used: n n Lohrenz-Bray-Clark correlation Little-Kennedy correlation Lohrenz-Bray-Clark’s Correlation Lohrenz, Bray, and Clark (1964) (LBC) developed an empirical correlation for determining the viscosity of the saturated oil from its composition. The proposed correlation has been enjoying great acceptance and application by engineers in the petroleum industry. The authors proposed the following generalized form of the equation: a1 + a2 ρr + a3 ρ2r + a4 ρ3r + a5 ρ4r μob ¼ μo + ξm 4 0:0001 (4.84) Coefficients a1 through a5 follow: a1 ¼ 0.1023 a2 ¼ 0.023364 a3 ¼ 0.058533 a4 ¼ 0.040758 a5 ¼ 0.0093324 with the mixture viscosity parameter, ξm, the viscosity parameter, μo, and the reduced density, ρr, given mathematically by 1=6 5:4402 Tpc ξm ¼ pffiffiffiffiffiffi 2=3 Ma ppc Xn μo ¼ pffiffiffiffiffiffi xi μi Mi pffiffiffiffiffiffi xi Mi i¼ 1 i¼ 1 Xn (4.86) 0 where 1 Xn @ i ¼ 1 ½ðxi Mi Vci Þ + xC7 + VC7 + Aρo i62C7 + ρr ¼ Ma (4.85) μo ¼ oil viscosity parameter, cp μi ¼ viscosity of component i, cp Tpc ¼ pseudocritical temperatures of the crude, °R (4.87) Crude Oil Viscosity “μo” 313 ppc ¼ pseudocritical pressure of the crude, psia Ma ¼ apparent molecular weight of the mixture ρo ¼ oil density at the prevailing system condition, lb/ft3 xi ¼ mole fraction of component i Mi ¼ molecular weight of component i Vci ¼ critical volume of component i, ft3/lb xC7 + ¼mole fraction of C7+ VC7 + ¼critical volume of C7+, ft3/lb-mol n ¼ number of components in the mixture The oil viscosity parameter, μo, essentially represents the oil viscosity at reservoir temperature and atmospheric pressure. The viscosity of the individual component, μi, in the mixture is calculated from the following relationships: 34 105 ðTri Þ0:94 ξi 5 17:78 10 ½4:58Tri 1:670:625 when Tri 1:5 : μi ¼ ξi when Tri 1:5 : μi ¼ (4.88) (4.89) where Tri ¼ reduced temperature of component i; T=Tci 5:4402ðTci Þ1=6 (4.90) ξi ¼ viscosity parameter of component i, given by ξi ¼ pffiffiffiffiffiffi Mi ðpci Þ2=3 It should be noted that, when applying Eq. (4.86), the viscosity of the nth component, C7+, must be calculated by any of the dead oil viscosity correlations discussed before. Lohrenz and coworkers proposed the following expression for calculating VC7 + : VC7 + ¼ a1+ a2 MC7 + + a3 γ C7 + + a4 MC7 + γ C7 + (4.91) where MC7 + ¼ molecular weight of C7+ γ C7 + ¼ specific gravity of C7+ a1 ¼ 21.572 a2 ¼ 0.015122 a3 ¼ 27.656 a4 ¼ 0.070615 Experience with the Lohrenz et al. equation has shown that the correlation is extremely sensitive to the density of the oil and the critical volume of C7+. When observed viscosity data are available, values of the coefficients a1–a5 314 CHAPTER 4 PVT Properties of Crude Oils in Eq. (4.84) and the critical volume of C7+ usually are adjusted and used as tuning parameters until a match with the experimental data is achieved. Little-Kennedy’s Correlation Little and Kennedy (1968) proposed an empirical equation for predicting the viscosity of the saturated crude oil. The correlation was originated from studying the behavior of 828 distinct crude oil systems representing 3349 viscosity measurements. The equation is similar in form to van der Waals’s equation of state (EOS). The authors expressed the equation in the following form: μ3ob p 2 am am bm μ bm + μob + ¼0 T ob T T (4.92) where μob ¼ viscosity of the saturated crude oil, cp p ¼ system pressure, psia T ¼ system temperature, °R am, bm ¼ mixture coefficient parameters Little and Kennedy proposed the following relationships for calculating the parameters am and bm: am ¼ exp ðAÞ (4.93) bm ¼ exp ðBÞ (4.94) The authors correlated the coefficients A and B with Temperature, T. Molecular weight of the C7+. Specific gravity of the C7+. Apparent molecular weight of the crude oil, Ma. Density of the oil, ρo, at p and T. The coefficients are given by the following equations: A ¼ Ao + A1 A3 MC7 + A4 ρo A5 ρ2o + + A2 MC7 + + + 2 + A6 Ma + A7 Ma3 T γ C7 + T T + A8 Ma ρo + A9 ðMa ρo Þ 3 (4.95) + A10 E2o B5 M 4 B1 B2 B6 ρ4 + 4 + B3 γ 3C7 + + B4 γ 4C7 + + 4 C7 + + 4 o + B7 Ma + B8 Ma ρo T T T γ C7 + + B9 ðMa ρo Þ4 + B10 ρ3o + B11 ρ4o (4.96) B ¼ Bo + Surface/Interfacial Tension “σ” 315 where MC7 + ¼ molecular weight of C7+ γ C7 + ¼ specific gravity of C7+ ρo ¼ density of the saturated crude oil, lb/ft3 T ¼ temperature, °R Ma ¼ apparent molecular weight of the oil A0–A10, B0–B11 ¼ coefficients of Eqs. (4.95), (4.96) are tabulated below. A0 ¼ 21:918581 B0 ¼ 2:6941621 A1 ¼ 16,815:621 B1 ¼ 3757:4919 A2 ¼ 0:0233159830 B2 ¼ 0:31409829 1012 A3 ¼ 0:0192189510 B3 ¼ 33:744827 A4 ¼ 479:783669 B4 ¼ 31:333913 A5 ¼ 719:808848 A6 ¼ 0:0968584490 A7 ¼ 0:5432455400 106 B5 ¼ 0:24400196 1010 B6 ¼ 4:632634 B7 ¼ 0:0370221950 A8 ¼ 0:0021040196 B8 ¼ 0:0011348044 A9 ¼ 0:4332274341 1011 B9 ¼ 0:0547665355 1015 3 A10 ¼ 0:0081362043 B10 ¼ 0:0893548761 10 B11 ¼ 2:0501808400 106 Eq. (4.92) can be solved for the viscosity of the saturated crude oil and the pressure and temperature of interest by extracting the minimum real root of the equation. An iterative technique, such as the Newton-Raphson iterative method, can be employed to solve the proposed equation. SURFACE/INTERFACIAL TENSION “σ” The surface tension is defined as the force exerted on the boundary layer between a liquid phase and a vapor phase per unit length. This force is caused by differences between the molecular forces in the vapor phase and those in the liquid phase and also by the imbalance of these forces at the interface. The surface can be measured in the laboratory and usually is expressed in dynes per centimeter. The surface tension is an important property in reservoir engineering calculations and designing enhanced oil recovery (EOR) projects. Sugden (1924) suggested a relationship that correlates the surface tension of a pure liquid in equilibrium with its own vapor. The correlating parameters of the proposed relationship are the molecular weight, M, of the pure component; the densities of both phases; and a newly introduced temperature 316 CHAPTER 4 PVT Properties of Crude Oils independent parameter, Pch. The relationship is expressed mathematically in the following form: σ¼ Pch ðρL ρv Þ 4 M (4.97) where σ is the surface tension and Pch is a temperature independent parameter, called the parachor. The parachor is a dimensionless constant characteristic of a pure compound, calculated by imposing experimentally measured surface tension and density data on Eq. (4.97) and solving for Pch. The parachor values for a selected number of pure compounds are given in the table below, as reported by Weinaug and Katz (1943). Component Parachor Component Parachor CO2 N2 C1 C2 C3 i-C4 49.0 41.0 77.0 108.0 150.3 181.5 n-C4 i-C5 n-C5 n-C6 n-C7 n-C8 189.9 225.0 231.5 271.0 312.5 351.5 Fanchi (1985) correlated the parachor with molecular weight with a simple linear equation. This linear equation is valid only for components heavier than methane. Fanchi’s linear equation has the following form: ðPch Þi ¼ 69:9 + 2:3Mi (4.98) where Mi ¼ molecular weight of component i and (Pch)i ¼ parachor of component i. Firoozabadi et al. (1988) developed a correlation that can be used to approximate the parachor of pure hydrocarbon fractions from C1 through C6 and for C7+ fractions: ðPch Þi ¼ 11:4 + 3:23Mi 0:0022ðMi Þ2 Katz and Saltman (1939) suggested the following expression for estimating the parachor for C7+ fractions as ðPch ÞC7 + ¼ 25:2 + 2:86MC7 + For a complex hydrocarbon mixture, Weinaug and Katz (1943) employed the Sugden correlation for mixtures by introducing the compositions of the two phases into Eq. (4.97). The modified expression has the following form: σ 1=4 ¼ n X ðPch Þi ðAxi Byi Þ i¼1 (4.99) Surface/Interfacial Tension “σ” 317 with the parameters A and B defined by ρo 62:4Mo ρg B¼ 62:4Mg A¼ where ρo ¼ density of the oil phase, lb/ft3 Mo ¼ apparent molecular weight of the oil phase ρg ¼ density of the gas phase, lb/ft3 Mg ¼ apparent molecular weight of the gas phase xi ¼ mole fraction of component i in the oil phase yi ¼ mole fraction of component i in the gas phase n ¼ total number of components in the system The interfacial tension varies between 72 dynes/cm for a water/gas system to 20–30 dynes/cm for a water/oil system. Ramey (1973) proposed a graphical correlation for estimating water/hydrocarbon interfacial tension that was curve fit by Whitson and Brule (2000) by the following expression: σ wb ¼ 20 + 0:57692ðρw ρb Þ where σ w-h ¼ interfacial tension, dynes/cm ρw ¼ density of the water phase, lb/ft3 ρh ¼ density of the hydrocarbon phase, lb/ft3 EXAMPLE 4.27 The composition of a crude oil and the associated equilibrium gas follows. The reservoir pressure and temperature are 4000 psia and 160°F, respectively. Component xi yi C1 C2 C3 n-C4 n-C5 C6 C7+ 0.45 0.05 0.05 0.03 0.01 0.01 0.40 0.77 0.08 0.06 0.04 0.02 0.02 0.01 The following additional PVT data are available: Oil density, ρo ¼ 46.23 lb/ft3 Gas density, ρg ¼ 18.21 lb/ft3 Molecular weight of C7+ ¼ 215 Calculate the surface tension. 318 CHAPTER 4 PVT Properties of Crude Oils Solution Step 1. Calculate the apparent molecular weight of the liquid and gas phases: X xi Mi ¼ 100:253 Mo ¼ X Mg ¼ yi Mi ¼ 24:99 Step 2. Calculate the coefficients A and B: A¼ ρo 46:23 ¼ ¼ 0:00739 62:4Mo ð62:4Þð100:253Þ B¼ ρg 18:21 ¼ ¼ 0:01168 62:4Mg ð62:6Þð24:99Þ Step 3. Calculate the parachor of C7+ from Eq. (4.98): ðPch Þi ¼ 69:9 + 2:3Mi ðPch ÞC7 + ¼ 69:9 + ð2:3Þð215Þ ¼ 564:4 Step 4. Construct a working table as shown in the table below. Component Pch Axi 5 0.00739 xi Byi 5 0.01169 yi Pch(Axi 2 Byi) C1 C2 C3 n-C4 n-C5 C6 C7+ 77.0 108.0 150.3 189.9 231.5 271.0 564.4 0.00333 0.00037 0.00037 0.00022 0.00007 0.000074 0.00296 0.0090 0.00093 0.00070 0.00047 0.00023 0.00023 0.000117 0.4361 0.0605 0.0497 0.0475 0.0370 0.0423 P1.6046 ¼ 0.9315 Step 5. Calculate the surface tension from Eq. (4.99): σ 1=4 ¼ n X ðPch Þi ðAxi Byi Þ ¼ 0:9315 i¼1 σ ¼ ð0:9315Þ4 ¼ 0:753dynes=cm PVT CORRELATIONS FOR GOM OIL Dindoruk and Christman (2004) developed a set of PVT correlations for GOM crude oil systems. These correlations have been developed for the following properties: n n n Bubble point pressure. Solution GOR at bubble point pressure. Oil FVF at bubble point pressure. PVT Correlations for GOM Oil 319 n n n n Undersaturated isothermal oil compressibility. Dead oil viscosity. Saturated oil viscosity. Undersaturated oil viscosity. In developing their correlations, the authors used the solver tool built into Microsoft Excel to match data on more than 100 PVT reports from the GOM. Dindoruk and Christman suggest that the proposed correlations can be tuned for other basins and areas or certain classes of oils. The authors’ correlations, as briefly summarized here, employ the following field units: Temperature, T, in 0176°R. Pressure, p, in psia. FVF, Bo, in bbl/STB. Gas solubility, Rs, in scf/STB. Viscosity, μo, in cp. For the bubble point pressure correlation, a9 R pb ¼ a8 as10 10A + a11 γg The correlating parameter A is given by a1 T a2 + a3 APIa4 A¼ 2Ra6 2 as + as7 γg The pb correlation coefficients a1–a11 are a1 ¼ 1.42828(1010) a2 ¼ 2.844591797 a3 ¼ 6.74896(104) a4 ¼ 1.225226436 a5 ¼ 0.033383304 a6 ¼ 0.272945957 a7 ¼ 0.084226069 a8 ¼ 1.869979257 a9 ¼ 1.221486524 a10 ¼ 1.370508349 a11 ¼ 0.011688308 For the gas solubility correlation, Rsb, Rsb ¼ a11 pb + a9 γ ag10 10A a8 320 CHAPTER 4 PVT Properties of Crude Oils with A¼ a1 APIa2 + a3 T a4 a 6 a5 + 2API pa7 2 b The coefficients (Rsb correlation) a1–a11 are a1 ¼ 4.86996(106) a2 ¼ 5.730982539 a3 ¼ 9.92510(103) a4 ¼ 1.776179364 a5 ¼ 44.25002680 a6 ¼ 2.702889206 a7 ¼ 0.744335673 a8 ¼ 3.359754970 a9 ¼ 28.10133245 a10 ¼ 1.579050160 a11 ¼ 0.928131344 For the oil FVF (Bob), Bob ¼ a11 + a12 A + a13 A2 + a14 ðT 60Þ API γg with the parameter A given by h a1 a1 A¼ Rs γ g a γ o3 + a4 ðT 60Þa5 + a6 Rs h i2 a9 s a8 + 2R ðT 60Þ a γ 10 ia7 g The coefficients a1–a14 for the proposed Bob correlation are a1 ¼ 2.510755(100) a2 ¼ 4.852538(100) a3 ¼ 1.183500(101) a4 ¼ 1.365428(105) a5 ¼ 2.252880(100) a6 ¼ 1.007190(101) a7 ¼ 4.450849(101) a8 ¼ 5.352624(100) a9 ¼ 6.308052(101) a10 ¼ 9.000749(101) a11 ¼ 9.871766(107) a12 ¼ 7.865146(104) a13 ¼ 2.689173(106) a14 ¼ 1.100001(105) PVT Correlations for GOM Oil 321 When the pressure is above the bubble point pressure, the authors proposed the following expression for undersaturated isothermal oil compressibility: co ¼ a11 + a12 A + a13 A2 106 h a1 a2 with A¼ Rs γ g a γ o3 + a4 ðT 60Þa5 + a6 Rs 2 a9 a8 + 2Ra10s ðT 60Þ ia7 γg The coefficients a1–a13 of the proposed co correlation are a1 ¼ 0.980922372 a2 ¼ 0.021003077 a3 ¼ 0.338486128 a4 ¼ 20.00006368 a5 ¼ 0.300001059 a6 ¼ 0.876813622 a7 ¼ 1.759732076 a8 ¼ 2.749114986 a9 ¼ 1.713572145 a10 ¼ 9.999932841 a11 ¼ 4.487462368 a12 ¼ 0.005197040 a13 ¼ 0.000012580 Normally, the dead oil viscosity (μod) at reservoir temperature is expressed in terms of reservoir temperature, T, and the oil API gravity. However, the authors introduced two additional parameters: the bubble point pressure, pb, and gas solubility, Rsb, at the bubble point pressure. Dindoruk and Christman justify this approach because the same amount of solution gas can cause different levels of bubble point pressures for paraffinic and aromatic oils. The authors point out that using this methodology will capture some information about the oil type without requiring additional data. The proposed correlation has the following form: μod ¼ a3 Ta4 ð log APIÞA a5 pab6 + a7 Rasb8 with A ¼ a1 log ðT Þ + a2 The coefficients (μod correlation) a1–a8 for the dead oil viscosity are a1 ¼ 14.505357625 a2 ¼ 44.868655416 322 CHAPTER 4 PVT Properties of Crude Oils a3 ¼ 9.36579(109) a4 ¼ 4.194017808 a5 ¼ 3.1461171(109) a6 ¼ 1.517652716 a7 ¼ 0.010433654 a8 ¼ 0.000776880 For the saturated oil viscosity (μob), μob ¼ Aðμod ÞB with the parameters A and B given by a1 a3 Ras 4 + exp ða2 Rs Þ exp ða5 Rs Þ a6 a8 Ras 9 B¼ + exp ða7 Rs Þ exp ða10 Rs Þ A¼ with the following coefficients a1–a10 of the proposed saturated oil viscosity correlation: a1 ¼ 1.000000(100) a2 ¼ 4.740729(104) a3 ¼ 1.023451(102) a4 ¼ 6.600358(101) a5 ¼ 1.075080(103) a6 ¼ 1.000000(100) a7 ¼ 2.191172(105) a8 ¼ 1.660981(102) a9 ¼ 4.233179(101) a10 ¼ 2.273945(104) For the undersaturated oil viscosity (μo), μo ¼ μob + a6 ðp pb Þ10A with A ¼ a1 + a2 logμob + a3 log ðRs Þ + a4 μob log ðRs Þ + a5 ðp pb Þ The coefficients (μo correlation) a1–a6 are a1 ¼ 0.776644115 a2 ¼ 0.987658646 a3 ¼ 0.190564677 a4 ¼ 0.009147711 a5 ¼ 0.0000191111 a6 ¼ 0.000063340 Properties of Reservoir Water 323 PROPERTIES OF RESERVOIR WATER Like hydrocarbon systems, the properties of the formation water depend on the pressure, temperature, and composition of the water. The compressibility of the connate water contributes materially in some cases to the production of undersaturated volumetric reservoirs and accounts for much of the water influx in water drive reservoir. The formation water usually contains salts, mainly sodium chloride (NaCl), and dissolved gas that affects its properties. Generally, formation water contains a much higher concentration of NaCl, ranging from 10,000 to 300,000 ppm, than the seawater concentration of approximately 30,000 ppm. Haas (1976) points out that there is a limit on the maximum concentration of salt in the water that is given by ðCsw Þmax ¼ 262:18 + 72:0T + 1:06T 2 where (Csw)max ¼ maximum salt concentration, ppm by weight T ¼ temperature, °C In addition to describing the salinity of the water in terms of ppm by weight, it can be also expressed in terms of the weight fraction, ws, or as ppm by volume, Csv. The conversion rules between these expressions are ρw Csw 62:4 Csw ¼ 106 ws Csv ¼ where ρw is the density of the brine at SC in lb/ft3 and can be estimated from the Rowe and Chou (1970) correlation as ρw ¼ 1 0:01604 0:011401ws + 0:0041755w2s The solubility of gas in water is normally less than 30 scf/STB; however, when combined with the concentration of salt, it can significantly affect the properties of the formation water. Methods of estimating the PVT properties, as documented next, are based on reservoir pressure, p, reservoir temperature, T, and salt concentration, Csw. Water FVF The water FVF can be calculated by the following mathematical expression (Hewlett-Packard, Petroleum Fluids PAC Manual, H.P. 41C, 1982): Bw ¼ A1 + A2 p + A3 p2 (4.100) where the coefficients A1–A3 are given by the following expression: Ai ¼ a1 + a2 ðT 460Þ + a3 ðT 460Þ2 324 CHAPTER 4 PVT Properties of Crude Oils with a1–a3 for gas-free and gas-saturated water given in the table below. The temperature T in Eq. (4.100) is in °R. Ai a1 Gas-free water 0.9947 A1 4.228(106) A2 1.3(1010) A3 Gas-saturated water 0.9911 A1 1.093(106) A2 5.0(1011) A3 a2 a3 5.8(106) 1.8376(108) 1.3855(1012) 1.02(106) 6.77(1011) 4.285(1015) 6.35(105) 3.497(109) 6.429(1013) 8.5(107) 4.57(1012) 1.43(1015) Source: Hewlett-Packard, Petroleum Fluids PAC Manual, H.P. 41C, 1982. McCain et al. (1988) developed the following correlation for calculating Bw: Bw ¼ ð1 + ΔVwt Þ 1 + ΔVwp with ΔVwt ¼ 0:01 + 1:33391 104 TF + 5:50654 107 TF2 ΔVwp ¼ pTF 1:95301 109 1:72834 1013 p p 3:58922 107 + 2:25341 1010 p where TF ¼ temperature, °F p ¼ pressure, psi The correlation does not account for the salinity of the water; however, McCain et al. observed that variations in salinity caused offsetting errors in terms ΔVwt and ΔVwp. Water Viscosity The viscosity of the water is a function of the pressure, temperature, and salinity. Ahmed (2014) proposed a pure water viscosity correlation that accounts for both the effects of temperature and pressure. Ahmed proposed the following expression for calculating pure water viscosity (salinity ¼ 0) at 1 atm and temperature “TF”: 0:763526623 μpure + w, 1 atm ¼ 30:86543156TF 0:07909888 0:234919042 TF2 To account for the impact of including the salinity on pure water viscosity, the following simplified expression is proposed: Properties of Reservoir Water 325 μbrine w, 1 atm ¼ 5167847:4μpure w, 1 atm 0:234919042wS 517818:4 wS where (μw,1atm)pure ¼ pure water viscosity at 1 atm and TF, cp (μw,1atm)brine ¼ brine water viscosity at 1 atm and TF, cp wS ¼ concentration of NaCl as expressed in parts per million “ppm” The water viscosity at pressure “p” is calculated from the following relationship: μw ¼ μbrine w, 1 atm exp ½0:000055p (4.101) with μwD ¼ A + B=TF A ¼ 4:518 102 + 9:313 107 Y 3:93 1012 Y 2 B ¼ 70:634 + 9:576 1010 Y 2 where μw ¼ water viscosity at p and T, cp μw,1atm ¼ water viscosity at p ¼ 14.7 and reservoir temperature in °F, cp μwD ¼ dead water viscosity, cp p ¼ pressure of interest, psia TF ¼ temperature of interest, °F Brill and Beggs (1973) presented a simpler equation, which considers only the temperature effects: μw ¼ exp 1:003 0:01479TF + 1:982 105 TF2 (4.102) McCain et al. (1988) accounted for the salinity and developed the following correlation for estimating the water viscosity at atmospheric pressure, reservoir temperature, and water salinity: μw1 ¼ ATFB with A ¼ 109:574 0:0840564ws + 0:313314ðws Þ2 + 0:00872213ðws Þ3 B ¼ 1:12166 + 0:0263951ws 0:000679461ðws Þ2 0:0000547119ðws Þ3 + 1:55586 106 ðws Þ4 where TF ¼ temperature, °F μwl ¼ water viscosity at 1 atm and reservoir temperature, cp ws ¼ water salinity, percent by weight, solids 326 CHAPTER 4 PVT Properties of Crude Oils McCain et al. adjusted μw1 to account for the increase of the pressure from 1 atm to p by the following correlation: μw ¼ 0:9994 + 0:0000450295p + 3:1062 109 p2 μw1 Gas Solubility in Water The gas solubility in pure saltwater (Rsw)pure, as expressed in scf/STB, can be estimated from the following correlation: ðRsw Þpure ¼ A + Bp + Cp2 (4.103) where A ¼ 2:12 + 3:45 103 TF 3:59 105 TF2 5 7 2 B ¼ 0:0107 5:26 10 TF + 1:48 10 TF C ¼ 8:75 107 + 3:9 109 TF 1:02 1011 TF2 p ¼ pressure, psi To account for the salinity of the water, which reduces the tendency of the gas to dissolve in the water, McKetta and Wehe (1962) proposed the following adjustment: Rsw ¼ D ðRsw Þpure where 0:285854 D ¼ ws 100:0840655TF Water Isothermal Compressibility Brill and Beggs (1973) proposed the following equation for estimating water isothermal compressibility, ignoring the corrections for dissolved gas and solids: cw ¼ C1 + C2 T + C3 TF2 106 where C1 ¼ 3.8546 0.000134p C2 ¼ 0.01052 + 4.77 107p C3 ¼ 3.9267 105 8.8 1010p p ¼ pressure in psia cw ¼ water compressibility coefficient in psi1 (4.104) Laboratory Analysis of Reservoir Fluids 327 McCain (1991) suggested the following approximation for water isothermal compressibility: cw ¼ 1 7:033p + 541:55Csw 537:0TF + 403, 300 where Csw is the salinity of the water in mg/L. LABORATORY ANALYSIS OF RESERVOIR FLUIDS Accurate laboratory studies of PVT and phase-equilibria behavior of reservoir fluids are necessary for characterizing these fluids and evaluating their volumetric performance at various pressure levels by conducting laboratory tests on a reservoir fluid sample. The amount of data desired determines the number of tests performed in the laboratory; however, for a successful PVT analysis it requires fluid samples that represent the original hydrocarbon in the reservoir. Fluid sampling for laboratory PVT analysis must be collected immediately after the exploration wells are drilled to properly characterize the original hydrocarbon in place. Sampling techniques depend on the reservoir. This chapter focuses on discussing well conditioning procedures and fluid sampling methods, reviewing the PVT laboratory tests and the proper use of the information contained in PVT reports, and describing details of several of the routine and special laboratory tests. Well Conditioning and Fluid Sampling Improved estimates of the PVT properties of the reservoir fluids at measure in the laboratory can be made by obtaining samples that are representative of the reservoir fluids. The proper sampling of the fluids is of the greatest importance in securing accurate data. A prerequisite step for obtaining a representative fluid sample is a proper well conditioning. Well Conditioning Proper well conditioning is essential to obtain representative samples from the reservoir. 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 bottom-hole pressure during the conditioning period is desirable and the produced gas-liquid ratio should remain constant (within about 2%) for several days; less-permeable reservoirs require longer periods. The farther the well deviates from the constant produced gas/liquid ratio, the greater the likelihood that the samples will not be representative. 328 CHAPTER 4 PVT Properties of Crude Oils The following procedure is recommended in conditioning of an oil well for subsurface sampling. Before collecting fluid samples, the following guidelines and well condition consecutive steps must be considered: 1. The tested well should be allowed to produce for a sufficient amount of time to remove the drilling fluids, acids, and other well stimulation materials. 2. After the cleaning period, the flow rate should be reduced to one-half the flow rate used during the cleaning period. 3. The well should be allowed to flow at this reduced rate for at least 24 h. However, for tight formation or viscous oil, the reduced flow-rate period must be extended to 48 h or higher to stabilize the various parameters monitored. 4. The pressure drawdown should be controlled and remain low to ensure that bottom-hole pressure does not fall below the saturation pressure. 5. During the reduced flow-rate period, the well should be closely monitored to establish when the wellhead pressure, production rate, and GOR have stabilized. Fluid Sampling The main objective of a successful sampling process is to obtain representative fluid samples for determining PVT properties. For proper definition of the type of reservoir fluid and to perform a proper fluid study, the collection of reservoir fluid samples must occur before the reservoir pressure is allowed to deplete below the saturation pressure of the reservoir fluid. Therefore, it is essential that the reservoir fluid samples be collected immediately after the hydrocarbon discovery, and the only production before sampling be as a result of well cleanout and to remove contamination. The critical steps in any successful sampling program include n n n n Avoiding two-phase flow in the reservoir; that is, bottom-hole pressure below the saturation pressure. Minimizing fluid contamination introduced by drilling and completion fluids. Obtaining adequate volumes. Preserving sample integrity. It should be pointed out that the resulting definition of the type of reservoir fluid, the overall quality of the study, and the subsequent engineering calculations based on that study are no better than the quality of the fluid samples originally collected during the field sampling process. Laboratory Analysis of Reservoir Fluids 329 There are three basic methods of sample collection: n n n Subsurface (bottom-hole) sampling Surface (separator) sampling Wellhead sampling Surface sampling from the wellhead, separator, and stock-tank is performed routinely during most well tests and is occasionally required from production process lines. A thorough review of the equipment and techniques used to obtain these different types of fluid samples is outside the scope of this book. Individuals who are interested in more extensive information on sampling procedures should refer to information available from equipment vendors and service companies that specialize in sampling; for example, Schlumberger, Haliburton, Baker Hughes, and others. However, a brief review of the most common sampling techniques will be useful to establish the fundamental principles that will be discussed later in this chapter. Subsurface Samples Subsurface sampling uses a sampler tool (eg, Schlumberger Repeat Formation Testing “RFT” or Modular Dynamic Testing tool “MDT”) that is run on a wireline to the reservoir depth. After lowering the sampling chamber to the top of the producing zone, the chamber is opened hydraulically or via an electronic signal from the surface. The crude oil is allowed to flow at a very low flow rate into the tool at constant pressure to avoid the liberation of the solution gas. A piston seals the chamber and the tool is brought to the surface. It should be noted that sampling of saturated reservoirs with this technique requires special care to obtain a representative sample because when the flowing bottom-hole pressure is lower than the bubble point, the validity of the sample remains doubtful. Multiple subsurface samples usually are taken by running sample chambers in cycle or performing repeat runs. The samples are checked for consistency by measuring their bubble point pressure in surface temperatures. Samples whose bubble point lies within 2% of each other are considered reliable and may be sent to the laboratory for PVT analysis. To obtain a bottom-hole sample, a period of reduced flow rate that generally lasts from 2 to 4 days is required to remove drilling fluids contamination and mud filtrates before sampling. After the reduced flowrate period, the well would be shut-in and allowed to reach static pressure. This shut-in period generally lasts from a day up to a week of more, based primarily on the permeability. The following guidelines are suggested: 330 CHAPTER 4 PVT Properties of Crude Oils 1. Bottom-hole sampling is recommended for all oil reservoirs, and may be successful with rich gas-condensate fluids. 2. A minimum of three fluid samples of approximately 600 cm3 each should be collected to ensure that at least one sample is valid. 3. Subsurface sampling generally is not recommended for gas-condensate reservoirs or oil reservoirs producing substantial quantities of water. 4. A large water column in the tubing of a shut-in oil well prevents sampling at the proper depth and usually creates a situation where the collection of representative subsurface fluid is impossible. 5. A static pressure gradient should be run and interpreted to determine the gas-oil level and oil-water level in the tubing. The sampler should not be lowered below the oil/water interface. 6. A surface sample must be collected in the case of downhole twophase fluid. It should be pointed out that wellbore frequently contains water even in wells that do not experience water production. Age water column in the tubing of a shut-in well might prevent sampling and obtaining a representative sample at the proper depth. For this reason, a static pressure gradient should be performed and interpreted to determine levels of the gas-oil and oil-water in the tubing, which is normally allowed to reach static pressure, as shown in Fig. 4.15. Pressure (psig) 2500 0 3000 3500 4000 4500 2000 4000 Depth Oil level 6000 8000 Water level 10,000 12,000 n FIGURE 4.15 Static pressure gradient. 5000 Pressure Depth Gradient (psig) (ft) (psi/ft) 2670 0 2720 2000 0.025 2770 4000 0.025 2830 5000 0.06 3165 6000 0.335 3497 7000 0.332 3827 8000 0.33 4157 9000 0.33 4291 9400 0.335 4370 9600 0.395 4458 9800 0.44 4547 10,000 0.445 Laboratory Analysis of Reservoir Fluids 331 Separator Samples Separator recombination samples are often the only available representative samples. In these circumstances, accurate measurement of hydrocarbon gas and liquid production rates and stability of separation conditions are critical because the laboratory tests will 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. Surface sampling involves taking samples of the gas and liquid flowing through the surface separator, as conceptually shown in Fig. 4.16, and recombining the two fluids in an appropriate ratio in such a way that the recombined sample is representative of the overall reservoir fluid stream. 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 the observed hydrocarbon gas/liquid ratio. GOR, scf/STB Must be converted to: GOR, scf/bbl psep psep Gas meter Separator liquid sample cylinder Separator gas Gas sample sample cylinder Liquid sample First stage separator n FIGURE 4.16 Separator samples. 332 CHAPTER 4 PVT Properties of Crude Oils The oil and gas samples are taken from the appropriate flow lines of the same separator, whose pressure, temperature, and flow rate must be carefully recorded to allow the recombination ratios to be calculated. The oil and gas samples are sent separately to the laboratory, where they are recombined before the PVT analysis is performed. Recombining separator samples requires thorough understanding of the relationship between gas solubility “Rs” and GOR. The surface GOR is traditionally defined as the ratio of the instantaneous gas rate to that of the oil rate, or GOR ¼ total gas rate liberated solution gas rate + free gas rate ¼ oil rate oil rate GOR ¼ Qo Rs + Qg Qg ¼ Rs + Qo Qo Darcy’s equation for gas and oil flow rates is given by Qo ¼ Qg ¼ 0:00708hkkrog ðpe pwf Þ re μo Bo ln rw 0:00708hkkrg ðpe pwf Þ re μg Bg ln rw To give the instantaneous GOR, " # krg Bo μo GOR ¼ Rs + kro Bg μg The above set of mathematical relationships describing the relationship between GOR and gas solubility “Rs” suggest that a representative sample, surface or bottom-hole sample, can be obtained if The well is producing at a STABILIZED GOR; that is, GOR ¼ constant (b) The reservoir pressure and pwf are both pb, which suggests a single phase is flowing; that is, (a) " # krg Bo μo ¼0 kro Bg μg The above equality specifies the required setting for a single-phase flow, which consequently satisfies the condition of representative samples. Figs. 4.17 and 4.18 illustrate the impact and difference and between representative and unrepresentative samples. Qg, scf/day Qg, scf/day GOR = Rs GOR>>>Rs Qo, STB/day Qo, STB/day Free gas Oil “Rs” Oil Oil “Rs” Representative samples Observed GOR Unrepresentative samples Stabilized GOR ≈ Rsb Rsb GORrecombine Time n FIGURE 4.17 Classification of surface samples. GO R (GOR)current Rsb krg mo Bo GOR = Rs + kro mg Bg (Rs)current (GOR)current = (Rs)current + Rs krg mo Bo kro mg Bg current pcurrent (GOR)current >> (Rs)current 0 (pb)original Pressure 0 Time n FIGURE 4.18 Impact of recombining samples based on GOR below pb. 334 CHAPTER 4 PVT Properties of Crude Oils A quality check (QC) on the sampling technique is that the bubble point of the recombined sample at the temperature of the separator from which the samples were taken should be equal to the separator pressure. The advantages of surface sampling and recombination include: 1. Large gas and oil volumes can be collected. 2. Access to surface samples is readily available. 3. Stabilized conditions can be established over a number of hours prior to sampling. 4. Cost of collecting surface samples is much lower than bottom-hole samples. 5. The subsurface sampling requirements also apply to surface sampling: If pwf is below pb, then it is probable that an excess volume of gas will enter the wellbore and even good surface sampling practice will not obtain a true reservoir fluid sample. 6. There is virtually no interruption of production during the sampling period (although conditioning of the well prior to sampling is also required). The following guidelines for surface sampling are suggested: 1. Should be used routinely for all well tests where fluids are produced into a separator. 2. It can be used as the principal method or as a backup or cross-check for other methods. 3. Minimum of two gas samples of 20 L volume plus two liquid samples of 500 cm3 each are required. 4. The well must be producing under a constant GOR and A MINIMUM 24 h of stabilized GOR is recommended before obtaining a representative sample. 5. Drawdown must be as low as possible; that is, low flow rate. 6. The produced GOR must be representative of the actual gas solubility Rs; that is, GOR ¼ Rs. Separator Shrinkage Factor “S”. The separator shrinkage factor, expressed in STB/bbl, is obtained by flashing a known volume of separator oil at separator conditions to that of stock-tank conditions. Mathematically, the separator shrinkage factor is defined as the ratio of volume of oil at stock-tank condition to that of separator conditions; that is, S¼ ðvolume of oilÞST ; STB=bbl ðvolume of oilÞsep This shrinkage factor is then used to convert the known produced GOR (expressed in scf/STB) to separator GOR (expressed in scf/bbl). Using Laboratory Analysis of Reservoir Fluids 335 the separator GOR, the separator fluids are recombined in a PVT cell for compositional and fluid analysis based on the volume ratio ðGORÞscf=bbl ¼ ðSÞSTB=bbl ðGORÞscf=STB Quality Checks of Separator Oil and Gas Samples. It should be pointed out that when the opening pressures of the gas samples and bubble point pressures of the liquid samples are similar to the reported field separator pressure, the indication is that the reported separator is reasonably correct and samples can be used for recombination and further PVT analysis. On the other hand, if the bubble point pressures of the liquid samples are significantly lower than the first-stage separator pressure or lower than the opening pressure of the separator gas samples it might indicate (a) Leakage in the liquid cylinder (b) The separator liquid sample had been collected from a secondary separator, while the gas sample had been collected from the highpressure first-stage separator Under the above two scenarios, this combination of separator products cannot be used to combine the two samples. In addition, it is critical that each separator cylinder be checked for its integrity upon arriving in the laboratory. The “quality” checks of the various gas and liquid samples include the following: (A) Separator gas samples n Opening pressure of each gas sample cylinder. All gas samples should be checked for opening pressure to verify the integrity of the sample. Gas sample cylinders should open at a pressure near the pressure of the separator in the field. The cylinder(s) found to have such pressure are heated to a temperature slightly higher than that of the separator in the field. If the gas cylinder opening temperature Topen is different from the separator temperature Tsep, the cylinder opening pressure should be close to popen ¼ psep n Topen Tsep Zsep Zopen If the opening pressure is much lower than the separator pressure, it is highly probable that a leakage had occurred and the sample should not be approved for PVT analysis. Gas sample conditioning. Before performing compositional analysis on a gas sample, the sample is conditioned and held at a temperature higher than the 336 CHAPTER 4 PVT Properties of Crude Oils (B) sampling temperature for at least 24 h. The sample should be rotated from time to time to evaporate any liquid present in the cylinder. If liquid is present after the condition process, this could be water of heavy-end hydrocarbons “carry over” during the separator sampling process. n Existence of condensed liquid at separator temperature. The gas sample cylinders that are found to contain no condensed hydrocarbon liquids are selected for possible use in the recombination and reservoir fluid study. n Air content determination. The gas sample cylinder with the highest opening pressure not containing liquid phase is analyzed for air. If the air content is found to be less than 0.2 mol%, the sample is selected for recombination of separator products and reservoir fluid study. Separator liquid samples n Bubble point pressure. The validity of the separator liquid cylinder samples is determined by measuring their saturation pressures at ambient temperature; all samples should show similar saturation pressures and be compared to that of the field separator pressure. The selected liquid sample is heated to slightly above the reservoir temperature to ensure all waxes and resins are dissolved in solution. n Presence and quantity of water. Once each separator liquid sample cylinder’s opening pressure is measured, the bottom cylinder valve is opened slightly to check for the presence of water. If water is found to be present, it is drained and its volume measured. Often, the water recovered from the cylinder is subjected to a water analysis to determine its origin. Quality Checks (1) If the opening pressures of the gas samples and bubble point pressures of the liquid samples are similar to the reported field separator pressure, the indication is that the reported separator is reasonably correct and samples can be used for recombination and further PVT analysis. As shown in Table 4.2, samples with higher pressure are usually selected for recombination. (2) At times, there are QC results that indicate first-stage separator gas had been collected with the separator liquid sample had been collected from a secondary separator. In this situation, the separator gas samples would open at a pressure near that of the first-stage separator, and the bubble point pressures of the liquid samples would Laboratory Analysis of Reservoir Fluids 337 Table 4.2 Samples Selection for Recombination Summary of Samples Received in the Laboratory Separator Gas Sampling Conditions Laboratory Opening Conditions Cylinder Number Pressure (psig) Temperature (°F) Pressure (psig) Temperature (°F) 1357a 3571 1050 1050 90 90 1045 Acceptable samples pb within 2% 1043 90 90 Separator Liquid Sampling Conditions Laboratory Bubble Point Pressure Cylinder Number Pressure (psig) Temperature (°F) Pressure (psig) Temperature (°F) 2468a 4682 1050 1050 90 90 937 Acceptable samples 929 70 70 a Samples with higher pressure are selected for compositional and analysis and recombination. #2006 Tarek Ahmed & Associates, Ltd. All Rights Reserved. be significantly lower than the first-stage separator pressure. Under no circumstances should this combination of separator products be used to combine the two samples. Wellhead Sampling Wellhead sampling involves the collection of a fluid sample at the surface from the wellhead itself or from the flow line or the upstream side of the choke manifold, provided that the fluid is still in one-phase condition. This option is restricted to wells producing dry gas, very low GOR oils (ie, heavy oils), and highly undersaturated reservoir fluids. Dry gas wellhead samples can be collected as for gas sampling from a separator, whereas wellhead of other or unknown fluids should be performed as for separator liquids. The following wellhead sampling applicability and guidelines are recommended: 1. For dry gases or highly undersaturated fluids wells. 2. This method can avoid bottom-hole sampling costs. 3. A backup surface samples should be taken as the state of the fluid is not usually known with certainty; that is, in the case of two-phase flow at the wellhead. 338 CHAPTER 4 PVT Properties of Crude Oils 4. All equipment and fluid sampling cylinders must be compatible with maximum wellhead pressure. 5. Minimum of two fluid samples of approximately 600 cm3 for undersaturated crude oils. 6. For a highly undersaturated retrograded or rich gas fluids, standard 20 L gas bottles can be used. Surface Samples Recombination Procedure When p < pb For a depleted reservoir (ie, p < pb), determining the original or current oil composition from a current surface sample poses a special problem. As shown in Fig. 4.19, when the reservoir or bottom-hole pressure is below the bubble point pressure, the GOR is greater than Rs and Rsb. The following procedure can be used to estimate the current or original oil composition: psep Tsep Gas (GOR) current Step 1 Oil (pb) recombined GOR (GOR)current >>>Rsb GORcurrent Rsb = GOR (Rs)current R s krg o Bo kro g Bg Step 2 Reduce pressure to pcurrent or (pb) original Remove all gas Step 3 pi Original (pb) pcurrent Pressure Representative sample n FIGURE 4.19 Illustration of recombining surface samples when p < pb. Step 4 Laboratory Analysis of Reservoir Fluids 339 (a) Place the separator gas and liquid samples in a PVT cell in a ratio that is equivalent to that of the separator test GOR. It should be pointed out that current (test) GOR > Rs. (b) PVT cell is heated to reservoir temperature and the pressure is increased to dissolve the gas to obtain the bubble point of the recombined sample; that is, (pb)recombined ≫ (pb)original. (c) The pressure in the PVT cell is slowly reduced to the current reservoir pressure “pcurrent” and/or to original bubble point pressure “pb.” The sample is then stabilized at that pressure for approximately 6 h. (d) After establishing complete equilibrium at the desired reservoir pressure, the excess equilibrium gas phase is removed leaving behind a saturated liquid that closely represents the composition at (pb)original or pcurrent. Fevang and Whitson (1994) provided an alternative approach to the above methodology to estimate the original crude oil composition in a depleted reservoir. The method is summarized in the following steps: (a) Place the separator gas and liquid samples in a PVT cell in a ratio that is equivalent to that of the separator test GOR. (b) PVT cell is heated to reservoir temperature and brought to equilibrium at the current average reservoir pressure “pcurrent.” (c) After establishing complete equilibrium, remove ALL equilibrium gas phase from the cell at constant pressure and transfer to a container, leaving the equilibrium oil in the PVT cell. (d) The removed gas is reinjected incrementally back to the PVT cell. After each injection, the bubble point is measured and continues until reaching the desired bubble point pressure. The mixture is considered a good approximation of composition at the original bubble point pressure. Representative and Unrepresentative Samples There are commonly three types of fluid sample classifications: (a) Representative sample (b) Unrepresentative sample (c) In situ representative sample The traditional definition of a "representative sample" is defined as any sample produced from a reservoir where its measured composition and PVT properties describe the original “undersaturated or saturated” hydrocarbon system; otherwise, the sample is traditionally labeled as “unrepresentative.” 340 CHAPTER 4 PVT Properties of Crude Oils As defined by Fevang and Whitson (1994), an "in situ representative sample" is a special case where the sample represents an in situ reservoir composition at a specific depth (or an average composition for a depth interval). Using or accepting experimental PVT data that are performed on a hydrocarbon sample, regardless of its classification as a representative or unrepresentative sample, will depend entirely on the type of numerical simulator used to model a reservoir with the most commonly used simulators being the black oil and compositional simulators. Black Oil Simulator In the black oil formulation, it is assumed that the hydrocarbon fluids may be sufficiently described by two components: a (pseudo) oil component and a (pseudo) gas component. The main assumption when using a black oil simulator is that the overall hydrocarbon fluid composition will remain constant during the simulation with all fluid PVT properties determined as a function of the oil pressure. All mass transfer between the two components (ie, solubility and liberation of gas) is described by the gas solubility “Rs.” The behavior of these two phases can then be sufficiently described in the black oil simulator by a set of tabulated PVT data (ie, Bo, Rs, μo, etc.). These PVT properties must be generated from a laboratory data on a “representative hydrocarbon” sample with an acceptable and well-defined saturation pressure. It should be pointed out that the “black oil" approach uses a simple interpolation of PVT properties as a function of pressure when modeling the performance of a reservoir under different development scenarios, including the natural depletion process of the reservoir, secondary recovery process, and gas and water coning, among other development scenarios. However, black oil simulators fail when they are used to model processes such as the vaporization process by high-pressure lean gas/CO2 injection, varying fluid volatility associated compositional gradient, retrograde gas systems. It should be noted the impact of accurate bubble point pressure can be significant when performing oil recovery calculations using a black oil simulator or the material balance approach. Fig. 4.20 shows the oil recovery prediction results from material balance calculations using a representative sample with a bubble point pressure of 2620 psig as compared with a bubble point pressure of 2220 psig from an unrepresentative sample. Fig. 4.20 illustrates the impact of the saturation pressure on oil recovery. The impact is clearly evident when comparing cumulative oil production using the PVT characteristics of both samples. As the reservoir approaches the bubble point Laboratory Analysis of Reservoir Fluids 341 5000 LEGEND Measured fluid data — sample is representative. Measured fluid data — sample bubblelpoint is 400 psi low. Reservoir pressure 4000 Oil in place - 43.3 MMSTB Closed reservoir solution 68s drive recovery calculations 3000 Re pre oil senta sam tiv ple e 2000 Unr epr e oil s senta t am ple ive 1000 4.5% Low 0 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 Cumulative oil (MMSTB ) n FIGURE 4.20 Oil recovery using representative and unrepresentative oil samples. of the unrepresentative sample (ie, 2220 psig) the cumulative oil from the representative sample case is 3.05 MMSTB as compared with 1.7 MMSTB from the unrepresentative sample case. Compositional Modeling In cases where compositional effects are important, reliable simulation results can only be obtained with a compositional simulator that is based on a thermodynamically consistent model such as a cubic EOS with applications linked to n n n The reservoir type; for example, retrograde gas reservoir. The type of the hydrocarbon system; for example, volatile oil. The current and future field development processes; for example, miscible gas injection. Unlike the black oil simulator, a compositional simulator does not require a tabulated PVT data, it only requires the optimized parameters of the selected EOS; for example, Peng-Robinson EOS. These optimized parameters, as detailed later in Chapter 5, result from matching a wide ranges of specialized laboratory data. A compositional simulator is used in a variety of 4.0 4.5 5.0 342 CHAPTER 4 PVT Properties of Crude Oils applications in which a black oil simulator does not adequately describe the fluid behavior experiencing compositional changes that can be attributed to (1) (2) (3) (4) (5) (6) (7) Natural depletion behavior of the reservoir below saturation pressure Mixing of the gas-cap gas with oil rim Miscible or immiscible gas injection Significant compositional gradient Gas-condensate systems Compositional changes due to crossflow, fluid mixing, and comingling Other sources To accurately simulate varying fluid combinations listed above and to account for the dynamic compositional behavior of the existing reservoir fluids will require special laboratory tests performed on representative and unrepresentative hydrocarbon samples. For example, assuming two of the collected crude oil samples from a reservoir were classified as “unrepresentative” due to n n The first sample collected below the saturation pressure. The second sample was considered to be “contaminated” with gas from the gas-cap. Disregarding these two samples without performing PVT analysis is greatly discouraged. These samples represent a natural reservoir depletion behavior that should be accounted for by tuning EOS to match all experimental PVT tests performed on all samples produced from the reservoir; independent of whether samples are representative or unrepresentative. For an accurate application of a compositional EOS simulator for field development and forecast, any reservoir sample should be automatically considered as a representative of the reservoir fluid. The final EOS parameters characterizations should be optimized to match all accurate PVT measurements of all samples produced from the reservoir. PVT MEASUREMENTS Black oil and compositional simulators are commonly used in the industry when conducting quantitative simulation studies for field development and optimization. Both simulators require an extensive collections of PVT data that describe the volumetric phase behavior of the reservoir existing phases. The “black oil” approach is based on a simple interpolation of PVT properties as a function of pressure. With the black oil approach, calculations can be done quite simply, as the fluid behavior is less complex than that with a compositional model. The compositional approach employs a thermodynamically consistent model such as a cubic EOS. Selection of a PVT Measurements 343 compositional model is preferable where compositional effects are important, for example, in cases where the fluid composition varies significantly in the reservoir, and/or the injected fluids are very different from the fluids already present in the reservoir. In this case, reliable simulation results can only be obtained with a compositional model. The compositional approach is a more complex model and requires tuning EOS to match a variety of PVT properties of oil and gas phases. The appropriate reservoir applications and PVT data requirements of these two approaches are briefly outlined below. (a) The black oil approach: The black oil approach is based on a simple interpolation of PVT properties as a function of pressure. These PVT properties are converted internally to a two-component model, defined as surface oil and surface gas. The model is structured to use laboratory-generated PVT data to describe the properties of each of the existing reservoir fluids as a function of pressure and temperature. These PVT data are generated by conducting a series of laboratory tests to provide the required model with the required PVT data, including n Oil, gas, and water FVFs “Bo, Bg, Bw” n Oil and water gas solubility “Rs, Rsw” n Oil, gas, and water viscosities “μo, μg, μw” n Solution oil-gas ratio “rs” if the reservoir gas contains unsubstantial amounts of condensate in solution, with solution oil-gas ratio rs 0 However, the black oil modeling lacks the ability to explicitly model the volumetric fluid behavior associated with phase-equilibria calculations. However, the black oil simulator does not consider changes in composition of the hydrocarbons with the decline in the reservoir pressure. Therefore, black oil simulators are used in situations where recovery processes are insensitive to compositional changes in the reservoir fluids. (b) The compositional approach: Compositional simulators are used when the reservoir depletion and development processes are sensitive to compositional changes. The approach is the natural choice for applications where complex phase behavior and compositionally dependent properties are critical in predicting field recovery and performance. The compositional approach employs a thermodynamically consistent model such as a cubic EOS. An EOS has the ability to provide accurate descriptions of the fluid thermodynamic properties and volumetric behavior of hydrocarbon mixtures in all type of reservoir classifications and developments, including 344 CHAPTER 4 PVT Properties of Crude Oils Gas condensate n Volatile oil systems n Miscible gas injection An EOS requires a detailed knowledge of the molar composition of the mixture as well as the critical properties and the acentric factor of all the components. However, reservoir fluids contain several hundred components that are impossible to identify to appropriately evaluate the hydrocarbon fluid behavior using a thermodynamic-based EOS model. As a result, a wide range of laboratory investigations are performed to study the volumetric behavior of hydrocarbon reservoir samples and to use the experimental results to calibrate/tune the thermodynamic-based EOS model. These laboratory studies are designed and performed on reservoir fluid samples to determine their n Compositions n Volumetric behavior as a function of pressure and temperature n Physical properties, such as density and viscosity n PVT tests are conducted in a “PVT CELL” and designed to study and quantify the phase behavior and properties of a reservoir fluid for the use in a numerical simulator or performing material balance applications. Routine laboratory tests are described as depletion experiments where the pressure on a hydrocarbon sample is reduced in successive steps. Regardless of the type of simulation approach, all aspects of PVT studies are carried out on a complete range of reservoir fluid types, including n n n gas-condensate systems crude oils critical fluids There are simple field (on-site) routine tests for reservoir fluid identification and other data; Including the following measurements: n n n n n n Crude oil API gravity Separator GOR Tubing head pressure Separator operating conditions H2S content Water analysis In general, there are two types of laboratory tests that are traditionally performed on hydrocarbon samples: 1. Routine laboratory tests: Several laboratory tests are routinely conducted to characterize the reservoir hydrocarbon fluid, including PVT Measurements 345 Compositional analysis of the hydrocarbon system. n Constant composition expansion. n Differential liberation. n Separator tests. n Constant volume depletion. 2. Special laboratory PVT tests: These types of tests are performed for very specific applications. If a reservoir is to be depleted under miscible gas injection or gas cycling scheme, the following tests may be performed: n Slim-tube test n Swelling test n Multiple forward contact experiment n Multiple backward contact experiment n Flow assurance tests n Routine Laboratory PVT Tests Compositional Analysis of the Reservoir Fluid An important test on all reservoir samples is the determination of the fluid composition. Compositional analysis generally refers to the measurement of the distribution of hydrocarbons and other components present in oil and gas samples. The components of petroleum and petroleum products number in the tens of thousands. They range in molecular weight from methane (16) to very large uncharacterized components with molecular weight in the thousands. Analysis of the hydrocarbon mixture is complicated not only by the enormous number of components, but also by the large variation in compound concentrations. Using modern chromatography techniques, samples are analyzed to determine the breakdown of the components in the sample. Gas chromatography (GC) is a technique for separating and identifying components of hydrocarbon mixtures. GC, as shown schematically in Fig. 4.21, is a common type of chromatography used in analyzing hydrocarbon mixtures that can be vaporized without decomposition. A small hydrocarbon sample of the hydrocarbon mixture is injected into a heated entrance “port” of the GC by a microsyringe. The heated port is set to a temperature higher than the components’ boiling point to vaporize the injected sample inside the port. A carrier gas, such as helium, flows through the port and displaces the gaseous components of the sample into a capillary column. It is within the column that separation of the components takes place. The column is packed with a filling called the “stationary phase,” which has the ability to adsorb and release the various components successively after certain retention time. Various released components travel through the stationary phase at different speeds, which causes them to separate. Because different components emerge from the column at 346 CHAPTER 4 PVT Properties of Crude Oils Injector port Flow controller Recorder Column Detector Column oven Carrier gas n FIGURE 4.21 Schematic of GC main components. different times, they can be identified by a detector at the outlet of the column. The detector is used to monitor the outlet stream from the column; thus, the time at which each component reaches the outlet and the amount of that component can be determined. The detector sends a signal to a chart recorder that results in a peak on the chart paper. Each peak corresponds to a different component, and the area under these peaks can be used to quantify the mole fraction of each component. To quantify the composition in terms of mole fractions, a calibration curve (peak area vs. concentration) is obtained by performing GC analysis of samples of known concentrations. With the development of more sophisticated equations of state to calculate fluid properties, it is recommended that compositional analyses of the reservoir fluid include a separation of components through C20+ as a minimum. Table 4.3 shows a chromatographic compositional analysis of surface separator oil and gas samples that were collected from the Big Butte Field. A physical recombination is required to produce a representative reservoir fluid. Following the compositional analyses of the two samples, separator gas and separator liquid were physically recombined in a PVT cell at a stabilized producing GOR of 1597 scf/sep bbl. The recombined reservoir fluid sample exhibited a bubble point of 3870 psia at 218°F. Quality Assessment of the Laboratory Compositional Analysis Obtaining an accurate fluid composition is perhaps the single most important laboratory measurement as compared with other PVT experimental analyses. This is due to the fact that determining the compositional analysis of the hydrocarbon sample precedes all subsequent PVT tests. Before PVT Measurements 347 Table 4.3 Hydrocarbon Analysis of Recombined Reservoir Fluid Samples Mathematical Recombination of Separator Products (Basis of Recombination: 1597 scf Separator Gas/bbl Separator Liquid) Component (Symbol) N2 CO2 H2S C1 C2 C3 i-C4 n-C4 i-C5 n-C5 C6 C7 C8 C9 C10 C11 C12 C13 C14 C15 C16 C17 C18 C19 C20 C21 C22 C23 C24 C25 C26 C27 C28 C29 C30+ Separator Gas (mol%) Separator Liquid (mol%) Separator Liquid (wt.%) 0.096 1.376 0 77.234 14.157 4.58 0.74 1.036 0.281 0.223 0.155 0.074 0.037 0.009 0.002 0.016 0.53 0 14.666 11.076 9.218 2.702 5.103 2.998 3.631 5.538 6.109 7.419 5.413 4.022 2.922 2.257 2.187 1.859 1.631 1.282 1.183 1.066 0.983 0.755 0.645 0.592 0.503 0.434 0.453 0.312 0.296 0.26 0.223 1.717 0.004 0.228 0 2.297 3.254 3.968 1.533 2.896 2.111 2.558 4.659 5.67 7.747 6.276 5.266 4.193 3.548 3.736 3.448 3.279 2.779 2.737 2.613 2.523 2.026 1.832 1.764 1.563 1.401 1.526 1.092 1.079 0.983 0.875 8.54 Molecular Weight Specific Gravity (Water 5 1.0) Reservoir Fluid (mol%) Reservoir Fluid (wt.%) 28.01 44.01 34.08 16.04 30.07 44.1 58.12 58.12 72.15 72.15 86.18 95.05 106.95 118.76 134.14 147 161 175 190 206 222 237 251 263 275 291 305 318 331 345 359 374 388 402 509.6 0.809 0.818 0.801 0.3 0.356 0.507 0.563 0.584 0.624 0.631 0.664 0.71 0.74 0.773 0.779 0.79 0.801 0.812 0.823 0.833 0.84 0.848 0.853 08.58 0.863 0.868 0.873 0.878 0.882 0.886 0.89 0.894 0.897 0.9 0.922 0.067 1.066 0 54.289 13.027 6.281 1.459 2.528 1.277 1.473 2.129 2.287 2.744 1.991 1.476 1.071 0.828 0.802 0.682 0.598 0.47 0.434 0.391 0.36 0.277 0.237 0.217 0.185 0.159 0.166 0.114 0.108 0.095 0.082 0.63 0.037 0.923 0 17.133 7.706 5.448 1.669 2.89 1.813 2.091 3.609 4.276 5.773 4.651 3.89 3.098 2.622 2.761 2.548 2.423 2.054 2.023 1.931 1.864 1.497 1.354 1.304 1.155 1.036 1.128 0.807 0.798 0.727 0.647 6.311 (Continued) 348 CHAPTER 4 PVT Properties of Crude Oils Table 4.3 Hydrocarbon Analysis of Recombined Reservoir Fluid Samples Continued Mathematical Recombination of Separator Products (Basis of Recombination: 1597 scf Separator Gas/bbl Separator Liquid) Component (Symbol) Total Molecular weight Density, lb/ft3 Separator Gas (mol%) Separator Liquid (mol%) Separator Liquid (wt.%) 100 20.95 100 102.44 100 Molecular Weight Specific Gravity (Water 5 1.0) Reservoir Fluid (mol%) Reservoir Fluid (wt.%) 100 50.80 100 43.0248 Compositional Groupings of Reservoir Fluid Group mol% wt.% MW SG Total fluid C7+ C10+ C20+ C30+ 100 16.404 9.382 2.27 0.63 100 56.682 41.981 16.763 6.311 50.83 175.65 227.46 375.43 509.6 0.557 0.816 0.846 0.895 0.922 instructing a commercial laboratory to proceed with other type of experiments, engineers should ensure the quality and reliability of the representative or recombined hydrocarbon samples. It should be pointed out that the degree of accuracy of results from applying equations of state is strictly linked to the accuracy of the original hydrocarbon composition as determined in the laboratory. There are four main QCs on the consistency of the reported hydrocarbon compositions that should be performed: n n n n Trend of equilibrium ratios “K-values” Hoffman plot Material balance test on overall composition Material balance test on the GOR Equilibrium Ratios Trend. As discussed in detail in Chapter 5, the equilibrium ratio, Ki, of a given component in a hydrocarbon mixture is defined as the ratio of the mole fraction of the component in the gas phase, yi, to that of the mole fraction of the component in the liquid phase, xi. Mathematically, the relationship is expressed as Ki ¼ yi xi PVT Measurements 349 where Ki ¼ equilibrium ratio of component i yi ¼ mole fraction of component i in the gas phase xi ¼ mole fraction of component i in the liquid phase The K-value measures the tendency of the component (ie, component “i”), to remain or escape to the gas phase. It is essentially a property that measures the volatility of the component at a pressure and temperature. This tendency suggests that in a multicomponent system, components’ Ki-value should follow the trend of their decreasing K-values with the increasing order of their normal boiling points; that is, KN2 > KCl > KCO2 > KC2 > KH2 S > KC3 > Ki-C4 > Kn-C4 > Ki-C5 > Kn-C5 > ⋯ To check the consistency of the laboratory recombined gas and liquid samples listed in Table 4.3 by applying the above K-value criteria, the equilibrium ratio for each component was calculated as shown in Fig. 4.22, which indicated a consistency in the laboratory-reported compositional analysis. Hoffman Plot. The Hoffman plot is a standard technique for evaluating the consistency of the recombined fluid samples through a graphical technique. The Hoffman plot is created by plotting the log(Kipsep) or alternately log(Ki) versus the component characteristic factor “Fi” as defined by the following relationship: 3 1 1 7 pci 6 6Tbi Tsep 7 Fi ¼ log 1 5 14:7 4 1 Tbi Tci 2 where psep ¼ separator pressure, psia Tsep ¼ separator temperature, °R pci ¼ critical pressure of component “i,” psia Tci ¼ critical temperature of component “i,” °R Tbi ¼ boiling point temperature of component “i,” °R When log(Kipsep) is plotted against Fi for each component on a Cartesian scale, the resulting plot should be a nearly straight-line for the light hydrocarbon components; specifically for components C1–C6, components. However, some downward curvature could occur when plotting the heavy-end components. A resulting straight line indicates that the liquid and vapor samples are reasonably in equilibrium at the separator conditions, and the measurement of the liquid and vapor compositions generally error free. 350 CHAPTER 4 PVT Properties of Crude Oils Nitrogen Methane Carbon dioxide Ethane Propane i-Butane n-Butane i-Pentane n-Pentane Hexane Heptane Octane Nonane Separator Liquid “xi” (mol%) Separator Gas “yi” (mol%) Ki = yi/xi 0.016 14.666 0.53 11.076 9.218 2.702 5.103 2.998 3.631 5.538 6.109 7.419 5.413 0.096 77.234 1.376 14.157 4.58 0.74 1.036 0.281 0.223 0.155 0.074 0.037 0.009 6 5.2661939 2.5962264 1.278169 0.496854 0.2738712 0.2030178 0.0937292 0.0614156 0.0279884 0.0121133 0.0049872 0.0016627 The K-value plot 7 6 5 Ki 4 3 2 1 ta ne i-P en ta ne nPe nt an e H ex an e H ep ta ne O ct an e N on an e e Bu an n- ut i-B pa ne ne de ha Pr o Et di ox i ne ha bo n C ar M et N itr og en 0 n FIGURE 4.22 The K-value consistency checks. Usually the plus fraction (eg, heptane-plus (C7+)), is not included in Hoffman consideration because it is generally not measured accurately. Unless nonhydrocarbon components (particularly the nitrogen component) they are usually excluded from the Hoffman plot. To further QC the laboratory recombined compositions listed in Table 4.3, the compositions were used to generate the Hoffman plot as shown in Fig. 4.23, which indicates consistency in the laboratory measurements. The Material Balance Consistency Test on the Overall Composition. The consistency test is derived from the component molal balance criteria as given by the following relationship: zi nt ¼ xi nL + yi nv PVT Measurements 351 Methane Ethane Propane i-Butane n-Butane i-Pentane n-Pentane Hexane Heptane Octane Nonane Separator Liquid (mol%) 14.666 11.076 9.218 2.702 5.103 2.998 3.631 5.538 6.109 7.419 5.413 Separator Pressure: 625 psia Separator Temperature: 104°F Separator Gas (mol%) 77.234 14.157 4.58 0.74 1.036 0.281 0.223 0.155 0.074 0.037 0.009 K = yi /xi Tb (°R) 200.95 332.21 415.94 470.45 490.75 541.76 556.56 615.37 668.74 717.84 763.07 log(K*P) 3.517377 2.902468 2.492109 2.233426 2.103414 1.767755 1.584159 1.242859 0.879142 0.493736 0.016684 5.266193918 1.278169014 0.496853981 0.273871207 0.203017833 0.093729153 0.061415588 0.027988443 0.012113275 0.004987195 0.001662664 Tc (°R) 343.01 549.74 665.59 734.08 765.18 828.63 845.37 911.47 970.57 1023.17 1070.47 pc (psia) 667 707.8 615 527.9 548.8 490.4 488.1 439.5 397.4 361.1 330.7 b 803.8875 1412.635 1798.201 2037.314 2151.155 2383.685 2478.166 2795.253 3079.198 3344.396 3592.934 b(1/Tb−1/T) 2.574268643 1.7460969 1.133056126 0.716190528 0.567065255 0.171024903 0.056164664 −0.416629931 −0.858290198 −1.274280291 −1.665655175 4.0 log(Ki psep) or log(Ki) C1 3.0 C2 i-C4 2.0 i-C5 n-C5 C3 n-C4 C6 1.0 C7 C8 0.0 –2.0 C9 –1.0 0.0 1.0 bi (1/Tbi − 1/Tsep) n FIGURE 4.23 Hoffman plot consistency check. where zi ¼ mole fraction of component i in the entire hydrocarbon mixture xi ¼ mole fraction of component i in the liquid phase yi ¼ mole fraction of component i in the gas phase nt ¼ total moles of the hydrocarbon mixture, lb-mol nL ¼ total moles of the liquid phase nv ¼ total moles of the gas phase zint ¼ total number of moles of component i in the system xinL ¼ total number of moles of component i in the liquid phase yinv ¼ total number of moles of component i in the vapor phase If the laboratory-reported overall composition “zi” was recombined mathematically based on a separator GOR expressed in scf/bbl, the martial balance can be used to identify discrepancies in laboratory recombined overall composition “zi.” This quality assessment can be performed by applying the following mathematical expression: zi ¼ 2130:331ρo xi + ML ðGORÞyi 2130:331ρo + ML ðGORÞ 2.0 3.0 352 CHAPTER 4 PVT Properties of Crude Oils where ML ¼ molecular weight of the separator liquid sample ρo ¼ density of separator liquid sample at separator pressure and temperature, lb/ft3 GOR ¼ recombined GOR, scf/bbl xi ¼ mole fraction of component i in the separator liquid phase sample yi ¼ mole fraction of component i in the separator gas phase sample The laboratory gas and liquid samples listed in Table 4.3 were recombined at GOR of 1597 scf/bbl. The separator liquid sample exhibits a molecular weight “ML” of 102.44 and liquid density “ρo” of 43.0248 lb/ft3 at separator pressure and temperature. To assess the quality of the reported compositional analysis given in Table 4.3, the molal material balance is applied to give zi ¼ zi ¼ 2130:331ρo xi + ML ðGORÞyi 2130:331ρo + ML ðGORÞ 2130:331ð43:0248Þxi + ð102:44Þð1597Þyi 2130:331ð43:0248Þ + ð102:44Þð1597Þ Applying the above relationship to assess the quality of mathematically recombined overall composition given in Table 4.3, results in a very good agreement with the laboratory data as shown in Table 4.4. The Material Balance Test Consistency on the GOR. As shown in applying the component material balance (MB) to test and assess the quality of the recombination of gas and liquid samples, a concise QC test on the laboratory-reported GOR can also be performed by solving the component MB relationship for the ratio “nL/nv” as follows: zi nt ¼ xi nL + yi nv yi nt nL xi ¼ zi nv nv zi The above form of the component material balance suggests plotting “yi/ zi” versus “xi/zi” would produce a straight line with a negative slope value of “nL/nv.” The GOR can then be calculated from the following expression: GOR ¼ 2130:331ðρo Þpsep , Tsep ðMWÞLiquid ðnL =nv Þ ; scf=bbl Table 4.4 Material Balance Quality Test Component (Symbol) Separator Liquid (mol%) Separator Liquid (wt.%) 0.096 1.376 77.234 14.157 4.58 0.74 1.036 0.281 0.223 0.155 0.074 0.037 0.009 0.002 0.016 0.53 14.666 11.076 9.218 2.702 5.103 2.998 3.631 5.538 6.109 7.419 5.413 4.022 2.922 2.257 2.187 1.859 1.631 1.282 1.183 1.066 0.983 0.755 0.645 0.592 0.503 0.004 0.228 2.297 3.251 3.968 1.533 2.896 2.111 2.558 4.659 5.67 7.747 6.276 5.266 4.193 3.548 3.736 3.448 3.279 2.779 2.737 2.613 2.523 2.026 1.832 1.764 1.563 Molecular Weight Specific Gravity (Water 5 1.0) Reservoir Fluid (mol%) Reservoir Fluid (wt.%) Calculated zj 28.01 44.01 16.04 30.07 44.1 58.12 58.12 72.15 72.15 86.18 95.05 106.95 118.76 134.14 147 161 175 190 206 222 237 251 263 275 291 305 318 0.809 0.818 0.3 0.356 0.507 0.563 0.584 0.624 0.631 0.664 0.71 0.74 0.773 0.779 0.79 0.801 0.812 0.823 0.833 084 0.848 0.853 0.858 0.863 0.868 0.873 0.878 0.067 1.066 54.289 13.027 6.281 1.459 2.528 1.277 1.473 2.129 2.287 2.744 1.991 1.476 1.074 0.828 0.082 0.682 0.598 0.47 0.434 0.391 0.36 0.277 0.237 0.217 0.185 0.037 0.923 17.133 7.706 5.448 1.669 2.89 1.813 2.091 3.609 4.276 5.773 4.651 3.895 3.098 2.622 2.761 2.548 2.423 2.054 2.023 1.931 1.864 1.497 1.354 1.304 1.155 0.000664 0.010631 0.540952 0.130176 0.062952 0.014656 0.0254 0.012858 0.014833 0.021457 0.023059 0.02767 0.020075 0.014887 0.010806 0.008347 0.008088 0.006875 0.006032 0.004741 0.004375 0.003942 0.003635 0.002792 0.002385 0.002189 0.00186 (Continued) PVT Measurements 353 N2 CO2 C1 C2 C3 i-C4 n-C4 i-C5 n-C5 C6 C7 C8 C9 C10 C11 C12 C13 C14 C15 C16 C17 C18 C19 C20 C21 C22 C23 Separator Gas (mol%) Component (Symbol) C24 C25 C26 C27 C28 C29 C30+ Total Molecular weight Density, lb/ft3 Separator Gas (mol%) Separator Liquid (mol%) Separator Liquid (wt.%) 100 20.95 0.434 0.453 0.312 0.296 0.26 0.223 1.717 100 102.44 1.401 1.526 1.092 1.079 0.983 0.875 8.54 100 43.0248 Molecular Weight 331 345 359 374 388 402 509.6 Specific Gravity (Water 5 1.0) 0.882 0.886 0.89 0.894 0.897 0.9 0.922 Reservoir Fluid (mol%) Reservoir Fluid (wt.%) 0.159 0.166 0.114 0.108 0.095 0.082 0.63 100 50.83 1.036 1.128 0.807 0.798 0.727 0.647 6.311 100 Calculated zj 0.001605 0.001675 0.001154 0.001095 0.000962 0.000825 0.00635 354 CHAPTER 4 PVT Properties of Crude Oils Table 4.4 Material Balance Quality Test Continued PVT Measurements 355 1.6 1.4 Component N2 CO2 C1 C2 C3 i-C4 n-C4 i-C5 n-C5 C6 C7 C8 C9 1.2 yi /zi = –0.57797168(xi /zi ) + 1.57653829 yi /zi 1 0.8 Slope− = nL/nv = 0.57797 0.6 0.4 0.2 0 0 0.5 1 1.5 xi /zi 2 2.5 n FIGURE 4.24 Recombined GOR consistency check. It should be noted that the slope value of “nL/nv” must be treated as positive in the above equation. Fig. 4.24 shows the straight line resulting from plotting “yi/zi” versus “xi/zi” using the laboratory gas and liquid samples listed in Table 4.3. The slope of the line in the figure is 0.57797, which is converted into GOR to give GOR ¼ GOR ¼ 2130:331ðρo Þpsep , Tsep ðMWÞLiquid ðnL =nv Þ 2130:331ð43:0248Þ ¼ 1548scf=bbl ð102:44Þð0:57797Þ The calculated GOR value is about 3% lower than the reported laboratory value of 1597 scf/bbl. The error is slightly higher than expected, usually around 1%; however, considering the other three quality assessment tests that were performed on the samples with good matches, the reported recombined GOR appears to be reasonable. CCE Tests CCE tests are conventionally performed on all types of reservoir fluids to obtain a better understanding of the pressure-volume relationship of these hydrocarbon systems. The CCE test can also be referred to by different names, including pressure-volume relations, constant mass expansion, flash liberation, flash vaporization, or flash expansion. The CCE test is performed on crude oil systems for the purposes of determining 3 xi /zi yi /zi 0.23881 0.49719 0.27015 0.85023 1.46760 1.85195 2.01859 2.34769 2.46504 2.60122 2.67118 2.70372 2.71873 1.43284 1.29081 1.42265 1.08674 0.72918 0.50720 0.40981 0.22005 0.15139 0.07280 0.03236 0.01348 0.00452 356 CHAPTER 4 PVT Properties of Crude Oils n n n n n n n n Saturation pressure Relative volume “Vrel” Isothermal compressibility coefficient “Co” above pb Oil density “ρo” at and above pb The traditional Y-function The extended Y-function “YEXT” Oil FVF “Bo” above pb Two-phase formation factor Bt below pb The experimental procedure involves placing a hydrocarbon fluid sample in a visual PVT cell at reservoir temperature that is held constant during the experiment. To ensure that the hydrocarbon sample exits in a single phase, the sample is pressurized to a much higher pressure than the initial reservoir pressure. The pressure is reduced isothermally in steps by removing mercury from the PVT cell and the total hydrocarbon volume “Vt” is measured at each pressure. The combination of successive reductions in the pressure and the measurement of total fluid volume “Vt” are performed at a CCE; that is, no gas or liquid is removed from the PVT cell at any time throughout the experiment. A schematic of CCE experiment is illustrated in Fig. 4.25. As the pressure approaches the saturation pressure, the total volume continuously increases due to the expansion of the single-phase crude oil. Reaching the bubble point pressure marks the appearance and start of liberation of the solution gas. The saturation pressure value is identified by visual inspection and by the point of discontinuity on the plot “pressure-volume” curve. As shown in Fig. 4.25, the saturation pressure and corresponding Voil = 100% 100% Oil p3= pb p2> > pb p1> > pb Vt1 oil Voil < 100% Voil ≈ 100% Voil = 100% Vt2 oil Voil <<100% p5<<<pb p4<<pb gas gas Vsat oil Vt4 oil Hg (a) (b) n FIGURE 4.25 Constant composition expansion test. (c) (d) (e) Vt5 PVT Measurements 357 volume are recorded and designated as “psat” and “Vsat,” respectively. The volume at the saturation pressure “Vsat” is used as a reference volume and the total volume “Vt” as a function of pressure is reported relative to Vsat as a ratio of “Vt/Vsat.” This volume is termed the relative volume and is expressed mathematically by the following equation: Vrel ¼ Vt Vsat (4.105) where Vrel ¼ relative volume Vt ¼ total hydrocarbon volume Vsat ¼ volume at the saturation pressure Table 4.5 shows the results of performing a CCE test on a recombined surface sample taken from the Big Butte crude oil system. The bubble point pressure of the hydrocarbon system is 1930 psi at 247°F. The tabulated data includes n n n n Relative volume Oil density at and above the saturation pressure Average single-phase compressibility; that is, Co at p pb The Y-function Table 4.5 Constant Composition Expansion Data (Pressure/Volume Relations at 247°F) Pressure (psig) Relative Volume 6500 6000 5500 5000 4500 4000 3500 3000 2500 2400 2300 2200 2100 2000 0.9371 0.9422 0.9475 0.9532 0.9592 0.9657 0.9728 0.9805 0.989 0.9909 0.9927 0.9947 0.9966 0.9987 Y-Function Density (g/cm3) 0.6919 0.6882 0.6843 0.6802 0.6760 0.6714 0.6665 0.6613 0.6556 0.6544 0.6532 0.6519 0.6506 0.6492 (Continued) 358 CHAPTER 4 PVT Properties of Crude Oils Table 4.5 Constant Composition Expansion Data (Pressure/Volume Relations at 247°F) Continued Pressure (psig) Relative Volume pb ≥ 1936 1930 1928 1923 1918 1911 1878 1808 1709 1600 1467 1313 1161 1035 782 600 437 1 1.0014 1.0018 1.003 1.0042 1.0058 1.0139 1.0324 1.0625 1.1018 1.0611 1.2504 1.3694 1.502 1.9283 2.496 3.4464 Pressure Range (psig) Y-Function 0.6484 2.108 2.044 1.965 1.874 1.784 1.71 1.56 1.453 1.356 Single-Phase Compressibility (psi21) 6500–6000 10.73(106) 6000–5500 11.31(106) 5500–5000 11.96(106) 5000–4500 12.70(106) 4500–4000 13.57(106) 4000–3500 14.61(106) 3500–3000 15.86(106) 3000–2500 17.43(106) 2500–2000 19.47(106) 2000–1936 20.79(106) Saturation pressure (psat) ¼ 1936 psig Vrel ¼ ρ¼ Vt : Vsat ρsat : Vrel 1 ðVrel Þ1 ðVrel Þ2 : ðVrel Þ1 + ðVrel Þ2 p1 p2 2 psat p : Y¼ pðVrel 1Þ ðco Þaverage ¼ Density (g/cm3) PVT Measurements 359 Traditionally, measured volumes from the CCE are expressed in terms of the relative volume and plotted a function of pressure. The plot shows clearly the transition from the single-phase system to the two-phase system. The pressure at which the slope changes represents the bubble point pressure, as shown in Fig. 4.26. It should be pointed out the change in the slope is related and attributed to the fluid compressibility resulting from the transition from the single phase to the two-phase state; that is, n n In the single liquid phase region, the liquid compressibility is relatively small resulting in a steep decline in the pressure with a small increase in the liquid volume. As the pressure approaches the bubble point pressure and initiation of liberation solution gas, the evolved gas is characterized by a high compressibility that causes less reduction in the pressure resulting in a sharp break in the slope between the single- and two-phase regions. Single-phase expansion 4000 3500 3000 Pressure 2500 pb=1936 psi 2000 1500 Twophas e exp ansio n 1000 500 0 0 0.5 1 1.5 2 Relative volume n FIGURE 4.26 Pressure-volume relations from the CCE test. 2.5 3 3.5 4 p 6500 6000 5500 5000 4500 4000 3500 3000 2500 2400 2300 2200 2100 2000 1936 1930 1928 1923 1918 1911 1878 1808 1709 1600 1467 1313 1161 1035 782 600 437 Rel Vol 0.9371 0.9422 0.9475 0.9532 0.9592 0.9657 0.9728 0.9805 0.989 0.9909 0.9927 0.9947 0.9966 0.9987 1 1.0014 1.0018 1.003 1.0042 1.0058 1.0139 1.0324 1.0625 1.1018 1.0611 1.2504 1.3694 1.502 1.9283 2.496 3.4464 360 CHAPTER 4 PVT Properties of Crude Oils It should be pointed out that in the case of light or volatile oils, the two slopes are not clearly defined. The pressure-volume plot shows a continuous shape with no clear discontinuity or break in volumetric behavior and consequently might produce an inaccurate bubble point pressure and corresponding volume “Vsat” measurements. Essentially, bubble point determination relies instead on the visual observation of forming the first bubble of gas that defines the bubble point. As outlined later in this chapter, Vsat is an important property that will impact the accuracy of all other PVT properties and, therefore, must be estimated accurately. n Oil density from the CCE test As shown in Table 4.5, the density of the oil at the saturation pressure is 0.6484 g/cm3. The density is determined directly from mass “m” of the fluid sample and fluid volume “V” measurements. Designated ρsat for the fluid density at the saturation pressure, it can be calculated from ρsat ¼ m Vsat Above the bubble point pressure, the density of the oil can be calculated by using the recorded relative volume at any pressure, to give ðm=Vsat Þ ρ ¼ sat ρ¼ Vrel Vp, T =Vsat (4.106) where ρ ¼ density at any pressure above the saturation pressure ρsat ¼ density at the saturation pressure Vrel ¼ relative volume at the pressure of interest Vp,T ¼ volume at the pressure “p” and reservoir (cell) temperature EXAMPLE 4.28 Given the experimental data in Table 4.5, verify the oil density values at 4000 and 6500 psi. Solution Applying Eq. (4.106) gives the following results: At 4000 psi: ρo ¼ ρsat 0:6484 ¼ ¼ 1:6714g=cm3 Vrel 0:9657 At 6500 psi: ρo ¼ ρsat 0:6484 ¼ ¼ 0:6919 Vrel 0:9371 PVT Measurements 361 n Isothermal compressibility coefficient from CCE test The instantaneous isothermal compressibility coefficient of any substance is defined as the rate of change in volume with respect to pressure per unit volume at a constant temperature. Mathematically, the substance isothermal compressibility “c” is defined by the following expression: c¼ 1 @V V @p T Applying the above expression for the crude oil instantaneous isothermal compressibility gives co ¼ 1 @V V @p T And equivalently can be written in terms of the relative volume: co ¼ 1 @ ðV=Vsat Þ ðV=Vsat Þ @p or co ¼ 1 @Vrel Vrel @p (4.107) Commonly, the relative volume data above the bubble point pressure is plotted as a function of pressure, as shown in Fig. 4.27. To evaluate instantaneous co at any pressure p, it is necessary approximate the 1 0.99 0.98 (Vrel)1 (co)average = –1 (Vrel)1 + (Vrel)2 (Vrel)1 – (Vrel)2 p1 – p 2 Relative volume 2 0.97 (Vrel)2 0.96 0.95 0.94 0.93 1500 2500 p1 3500 p2 4500 Pressure n FIGURE 4.27 Relative volume data above the bubble point pressure. 5500 6500 7500 362 CHAPTER 4 PVT Properties of Crude Oils derivative (@Vrel/@p) at the desired pressure. The pressure-volume curve shown in Fig. 4.27 can be represented by the following third degree polynomial: Vrel ¼ 1:4701 1013 p3 + 2:95895 109 p2 30:124 106 p + 1:048212 with @Vrel =@p ¼ 1:4701 1013 ð3Þp2 + 2:95895 109 ð2Þp 30:124 106 EXAMPLE 4.29 Using Fig. 4.27, evaluate co at 4500 psi. Solution n Relative volume at 4500 psi ¼ 0.9592 n @Vrel ¼ 1:4701 1013 ð3Þð4500Þ2 + 2:95895 109 ð2Þ4500 30:124 106 @p ¼ 12:424 106 Apply Eq. (4.107) to give 1 @Vrel 1 ¼ co ¼ 12:424 106 ¼ 12:953 106 psi1 Vrel @p 0:9592 Traditionally, the laboratory isothermal compressibility coefficients are reported as average compressibility coefficients at several ranges of pressure as shown Table 4.5. These values are determined by calculating the changes in the relative volume at the indicated pressure interval and applying the following expression: ðco Þaverage ¼ 1 ðVrel Þ1 ðVrel Þ2 ðVrel Þ1 + ðVrel Þ2 p1 p2 2 (4.108) The subscripts 1 and 2 represent the corresponding values at the lower and higher pressure ranges, respectively. EXAMPLE 4.30 Using the measured relative volume data in Table 4.5 for the Big Butte crude oil system, calculate the average oil compressibility in the pressure range of 2500– 2000 psi. Solution Apply Eq. (4.108) to give 1 ðVrel Þ1 ðVrel Þ2 ðVrel Þ1 + ðVrel Þ2 p1 p2 2 1 0:9657 0:9592 ðco Þaverage ¼ ¼ 13:51 106 psi1 0:9657 + 0:9592 4000 4500 2 ðco Þaverage ¼ PVT Measurements 363 n The traditional Y-function The relative volume data frequently require smoothing to correct for laboratory inaccuracies in measuring the total hydrocarbon volume, particularly below the saturation pressure and also at lower pressures. A dimensionless compressibility function, commonly called the Y-function, is used to smooth the values of the relative volume “Vrel” below the saturation pressure. The Y-function in its mathematical expression is ONLY defined below the saturation pressure; that is, p < pb, and given by the following equation: Y¼ psat p pðVrel 1Þ (4.109) where psat ¼ saturation pressure, psia p ¼ pressure, psia Vrel ¼ relative volume at pressure p Table 4.5 lists the computed values of the Y-function as calculated by using Eq. (4.109). To smooth the relative volume data below the saturation pressure, the Y-function is plotted as a function of pressure on a Cartesian scale. Depending on the classification of fluid type, the Y-function could form a straight line or a line with only a small curvature. Fig. 4.28 shows the Y-function versus pressure for the Big Butte crude oil system. The figure illustrates the erratic behavior of the laboratory-measured data near the bubble point pressure. The following steps summarize the simple procedure of smoothing and correcting the relative volume data. Step 1. Calculate the Y-function for all pressures below the saturation pressure using Eq. (4.109). Step 2. Plot the Y-function versus pressure on a Cartesian scale. Step 3. Determine the coefficients of the best straight-line fit of the Y-function data: Y ¼ a + bp (4.110) where a and b are the intercept and slope of the lines, respectively. Step 4. Recalculate the relative volume at all pressures below the saturation pressure from the following expression: Vrel ¼ 1 + psat p pða + bpÞ (4.111) 364 CHAPTER 4 PVT Properties of Crude Oils erratic behavior of the laboratory data n FIGURE 4.28 Y-Function vs. pressure. EXAMPLE 4.31 Smooth the laboratory recorded relative volume data of Table 4.5 with the best straight-line fit of the Y-function as given by Y ¼ a + bp where a ¼ 1:0981 b ¼ 0:000591 PVT Measurements 365 Notice that the laboratory report suggests that the bubble point pressure could be greater than the reported value. Solution Applying Eq. (4.111) gives Vrel ¼ 1 + psat p 1950:7 p ¼ 1+ pða + bpÞ pð1:0981 + 0:000591pÞ Notice that the pressure must be expressed in psia. Results are shown in the table below. Pressure (psia) Measured Vrel Smoothed Vrel 1950.7 1944.7 1942.7 1937.7 1932.7 1925.7 1892.7 1822.7 1723.7 1614.7 1481.7 1327.7 1175.7 1049.7 796.7 614.7 451.7 – – – – – – – – 1.0625 1.1018 1.1611 1.2504 1.3696 1.502 1.9283 2.496 3.4464 – 1.0014 1.0018 1.003 1.0042 1.0058 1.0139 1.0324 1.063 1.1028 1.1626 1.2532 1.3741 1.5091 1.9458 2.5328 3.529 As illustrated in the above example, erratic data from the CCE test were obtained around the saturation pressure and also at lower pressures. To visually detect the appearance of the gas or first droplet of liquid that marks the saturation pressure “psat” can be a tedious and difficult task, particularly when conducting a CCE test on volatile oil and retrograde gas systems. Erroneous measurement of the saturation pressure will lead to an incorrect corresponding value of the reference volume Vsat, which consequently, impacts the accuracy of the relative volume values. As detailed later in this chapter, correct relative volume data are required to generate the PVT properties of the crude oil system, including n n Oil FVF “Bo” above the pb The two-phase FVF “Bt” 366 CHAPTER 4 PVT Properties of Crude Oils n n Isothermal compressibility coefficient of the oil “co” The Extended Y-Function Hosein et al. (2014) introduced the concept of the extended Y-function, which is based on replacing the Vsat in the traditional Y-function with that of the laboratory-measured value of initial volume “Vi” at the highest pressure in the CCE test. Using Vi as a reference volume will eliminate the tedious process of estimating Vsat, particularly when performing laboratory PVT study on volatile oil and condensate systems. Hosein and coauthors’ proposed approach offers a major advantage of using Vi as a reference volume in that the Y-function calculations can be extended above the saturation pressure; that is, removing the limitation of Y-function applicability to only below the saturation pressure. Applying the approach would produce two straight lines (above and below the saturation pressure) that clearly identify the saturation pressure by the intersection of the two lines, eliminating the need to measure the saturation pressure by visually observing the appearance of the first bubble of gas or droplet of liquid. This extended relative volume “(Vrel)EXT” can be expressed by the following relationship: ðVrel ÞEXT ¼ Vt Vi To distinguish between the traditional Y-function and the newly introduced EXTENDED Y-function, Hosein and coauthors termed the new Y-function as the YEXT function to signify its applicability to applications above the saturation pressure. The YEXT function is given by " # pi p pi p ¼ log YEXT ¼ log pi ½ ðVt =Vi Þ 1 pi ðVrel ÞEXT 1 (4.111a) Ahmed (2015) suggests that in addition to applying the above expression, the “pi” can be also replaced with “p” and referring to the function as “YEXTp”: " pi p YEXTp ¼ log p ðVrel ÞEXT 1 # (4.111b) The two expressions above should be plotted jointly for the purpose of clearly identifying the saturation pressure “psat” and the corresponding saturation volume “Vsat.” Ahmed (2015) proposed the following modified procedure for analyzing the CCE laboratory data, which is summarized in the following steps: PVT Measurements 367 1. To convert laboratory relative volume “Vrel” data into “(Vrel)EXT,” divide each Vrel value by the relative volume “ðVi =Vsat Þpmax ” that corresponds to the highest pressure “pmax” in the CCE data set; that is, ðVrel ÞEXT ¼ ðVt =Vsat Þ Vt ¼ ðVi =Vsat Þpmax Vi 2. Calculate the extended Y-function values at each pressure using Eqs. (4.111a), (4.111b); that is, " # pi p YEXT ¼ log pi ðVrel ÞEXT 1 " # pi p YEXTp ¼ log p ðVrel ÞEXT 1 3. On a Cartesian (regular) scale, plot results from steps 1 and 2 as a function of pressure on the same graph; that is, YEXT, YEXTp, and (Vrel)EXT versus pressure. 4. The two extended Y-function plots will clearly confirm the laboratory-reported saturation pressure or correctly identify a correct saturation pressure “psat.” 5. From the “(Vrel)EXT” plot, identify relative volume value at the saturation pressure and designate the value as Vpsat EXT , which represents the correct ratio of “Vsat/Vi.” 6. Draw the best straight-line fit of the extended Y-function “YEXT” data below the saturation pressure and determine the coefficients “a&b” of the best fit to give YEXT ¼ a + bp 7. Smooth the extended relative volume “(Vrel)EXT” data by recalculating their values at all pressures below the saturation pressure from the following expression: ðVrel ÞEXT ¼ 1 + p p i pi 10 a + b p 8. The traditional “smoothed” relative volume “Vrel” data below the saturation pressure can then be generated as a function of pressure from the following expression: ðVrel ÞEXT Vrel ¼ Vpsat EXT The approach can be clarified further by reworking Example 4.31. 368 CHAPTER 4 PVT Properties of Crude Oils EXAMPLE 4.32 Smooth the laboratory recorded relative volume data of Table 4.5 for the extended Y-function approach and verify the reported saturation pressure of 1936 psi. Notice that the laboratory report suggests that the bubble point pressure could be greater than the reported value. Solution The application of the modified Y-function procedure are summarized in the following steps: 1. Convert laboratory relative volume “Vrel” data into “(Vrel)EXT” by dividing each Vrel by relative volume that corresponds to the highest pressure “pmax” in the CCE data set; that is, ðVrel ÞEXT ¼ ðVt =Vsat Þ Vt ¼ ðVi =Vsat Þpmax Vi p Vrel (Vrel)EXT 6500 6000 5500 5000 4500 4000 3500 3000 2500 2400 2300 2200 2100 2000 1936 1930 1928 1923 1918 1911 1878 1808 1709 1600 1467 1313 1161 1035 782 600 437 0.9371 0.9422 0.9475 0.9532 0.9592 0.9657 0.9728 0.9805 0.989 0.9909 0.9927 0.9947 0.9966 0.9987 1 1.0014 1.0018 1.003 1.0042 1.0058 1.0139 1.0324 1.0625 1.1018 1.0611 1.2504 1.3694 1.502 1.9283 2.496 3.4464 1 1.005442 1.0111098 1.017181 1.023583 1.03052 1.038096 1.046313 1.055384 1.057411 1.059332 1.061466 1.063494 1.065735 1.067122 1.068616 1.069043 1.070323 1.071604 1.073311 1.081955 1.101697 1.133817 1.175755 1.132323 1.334329 1.461317 1.602817 2.057731 2.663536 3.677729 PVT Measurements 369 2. Calculate the extended Y-function at pressure using Eqs. (4.111a), (4.111b). " # pi p pi p ¼ log pi ½ðVt =Vi Þ 1 pi ðVrel ÞEXT 1 YEXT ¼ log " # pi p YEXTp ¼ log p ðVrel ÞEXT 1 p (Vrel)EXT YEXT YEXTp 6500 6000 5500 5000 4500 4000 3500 3000 2500 2400 2300 2200 2100 2000 1936 1930 1928 1923 1918 1911 1878 1808 1709 1600 1467 1313 1161 1035 782 600 437 1 1.005442 1.011098 1.017181 1.023583 1.03052 1.038096 1.046313 1.055384 1.057411 1.059332 1.061466 1.063494 1.065735 1.067122 1.068616 1.069043 1.070323 1.071604 1.073311 1.081955 1.101697 1.133817 1.175755 1.132323 1.334329 1.461317 1.602817 2.057731 2.663536 3.677729 1.150272 1.141839 1.128138 1.11551 1.100447 1.083326 1.065451 1.045765 1.040874 1.037047 1.031919 1.027808 1.022504 1.019568 1.010578 1.008075 1.000568 0.993205 0.983634 0.938341 0.851138 0.741001 0.632375 0.767278 0.377828 0.250547 0.144491 0.08004 0.26309 0.45799 1.185035 1.21439 1.242081 1.275211 1.3113 1.352171 1.401243 1.460739 1.473576 1.488233 1.502409 1.518502 1.534388 1.545576 1.537934 1.535881 1.529502 1.52327 1.515287 1.477559 1.406853 1.321173 1.241168 1.413761 1.072477 0.998628 0.942464 0.839662 0.771668 0.714439 3. On a Cartesian (regular) scale, plot results from steps 1 and 2 as a function of pressure on the same graph; that is, YEXT, YEXTp, and (Vrel)EXT versus pressure. The plot is shown in Fig. 4.29. 370 CHAPTER 4 PVT Properties of Crude Oils 1.4 2 YEXTp = log pi – p p (Vrel)EXT –1 1.5 1.3 YEXT = log 8P 0.5 pi – p pi (Vrel)EXT –1 1.2 0.0 00 95 YEXT 1 42 6+ Vext 1.1 EX T = Y –0 .84 0 –0.5 (Vrel)EXT = Vt Vi –1 0 1000 2000 3000 4000 5000 6000 1 7000 n FIGURE 4.29 Extended Y-function vs. pressure. 4. The two extended Y-function plots show saturation pressure 1950 psia as compared with the reported saturation pressure of 1936 psi. 5. From the “(Vrel)EXT” plot, identify relative volume value at the “corrected” saturation pressure to give: Vpsat EXT 1:066261. It should be noted that the value of the (Vrel)EXT at laboratory-reported saturation pressure of 1936 is 1.067122. 6. Draw the best straight-line fit of the extended Y-function “YEXT” data below the saturation pressure and determine the coefficients “a&b” of the best fit to give YEXT ¼ a + bp ¼ 0:84426 + 0:000958p 7. Smooth the extended relative volume “(Vrel)EXT” data by recalculating their values at all pressures below the saturation pressure from the following expression: ðVrel ÞEXT ¼ 1 + p p 6500 p i ¼ 1+ pi 10a + bp 6500 100:84426 + 0:000958 p PVT Measurements 371 p Smoothed (Vrel)EXT 1950 1936 1930 1928 1923 1918 1911 1878 1808 1709 1600 1467 1313 1161 1035 782 600 437 1.066261 1.068549 1.069554 1.069892 1.070745 1.071607 1.072833 1.078896 1.093463 1.118727 1.154434 1.21271 1.307899 1.443168 1.598972 2.095052 2.688102 3.485354 8. Generate the traditional “smoothed” relative volume “Vrel” data below the saturation pressure from the following expression: ðVrel ÞEXT ðVrel ÞEXT Vrel ¼ ¼ Vpsat EXT 1:066261 p (Vrel)EXT Smoothed Vrel 6500 6000 5500 5000 4500 4000 3500 3000 2500 2400 2300 2200 2100 2000 1950 1936 1930 1.000000 1.005442 1.011098 1.017181 1.023583 1.03052 1.038096 1.046313 1.055384 1.057411 1.059332 1.061466 1.063494 1.065735 1.066261 1.068549 1.069554 0.937857 0.942961 0.948265 0.95397 0.959975 0.96648 0.973586 0.981292 0.989799 0.9917 0.993502 0.995503 0.997405 0.999506 1 1.002146 1.003088 372 CHAPTER 4 PVT Properties of Crude Oils 1928 1923 1918 1911 1878 1808 1709 1600 1467 1313 1161 1035 782 600 437 1.069892 1.070745 1.071607 1.072833 1.078896 1.093463 1.118727 1.154434 1.21271 1.307899 1.443168 1.598972 2.095052 2.688102 3.485354 1.003406 1.004205 1.005014 1.006163 1.011849 1.025512 1.049206 1.082694 1.137348 1.226622 1.353485 1.499607 1.964859 2.521054 3.268762 The laboratory density of 0.6484 g/cm3 measured at 1936 psi should be corrected to the new saturation pressure of 1950 psi to give ðρsat Þ1950psi ¼ ðVrel ÞEXT @ 1936psi Vrel ÞEXT @ 1950psi ρsat @1936psi ¼ 1:07122 0:6468 1:066261 ¼ 0:65142g=cm3 with all densities at higher pressure than 1950 psi given by ρ¼ ρsat 0:65142 ¼ Vrel Vrel p (Vrel)EXT Vrel Density 6500 6000 5500 5000 4500 4000 3500 3000 2500 2400 2300 2200 2100 2000 1950 1 1.005442 1.011098 1.017181 1.023583 1.03052 1.038096 1.046313 1.055384 1.057411 1.059332 1.061466 1.063494 1.065735 1.066261 0.937857 0.942961 0.948265 0.95397 0.959975 0.96648 0.973586 0.981292 0.989799 0.9917 0.993502 0.995503 0.997405 0.999506 1 0.694579 0.690819 0.686955 0.682847 0.678576 0.674009 0.669089 0.663835 0.658129 0.656868 0.655676 0.654358 0.653111 0.651737 0.651416 PVT Measurements 373 Differential Liberation Test The differential liberation (DL) test is perhaps the most common laboratory test that is conducted on crude oil samples. The DL process is summarized in the following steps and is illustrated schematically in Fig. 4.30: n n n n n n A crude oil sample is placed in a visual PVT cell at its bubble point pressure and reservoir temperature. The cell pressure is reduced in stages, usually 10–15 pressure increments, from the saturation pressure to atmospheric pressure. The liberated gas at each pressure stage is allowed to reach equilibrium with the cell remaining oil. The volume of the two phases (ie, remaining oil volume “VL” and liberated gas volume “Vgas”) are measured and recorded at each pressure level. The liberated gas is then displaced at constant cell pressure to a metering device (eg, gasometer) and the measured volume is corrected to SC and designated to as “(Vgas)sc.” The above depletion process is repeated at constant reservoir temperature until a pressure close to atmospheric pressure is reached. Having arrived at the last stage, the volume of the residual oil is first measured at last cell conditions, then corrected to SC (ie, 14.7 psia and 60°F), and designated as “VLsc.” pb VL p1<pb p2<p1 p1 Gas off (V gas)sc Vgas Vsat VL1 Vt Hg p sc Gas off (V gas) sc Gas off (V gas) sc Vgas VL2 Vt VL1 Hg p2 VLsc VL2 Hg = total volume at p&T Vt VL = remaining oil volume at p&T V Lsc = remaining oil volume at p sc&T sc V gas = liberated gas volume at p&T (Vgas)sc = liberated gas volume at psc&Tsc n FIGURE 4.30 Differential vaporization test at constant temperature “T.” Hg Hg Hg 14.7 psia 60°F 374 CHAPTER 4 PVT Properties of Crude Oils The following nomenclatures are used to describe the various hydrocarbon volume measurements associated with each pressure-depletion stage during the DL experiment: p ¼ cell pressure T ¼ cell temperature Vt ¼ total combined volume of the liberated gas and remaining oil at p&T VL ¼ remaining oil volume at p&T VLsc ¼ remaining oil volume at psc&Tsc Vgas ¼ liberated gas volume at p&T (Vgas)sc ¼ liberated gas volume at psc&Tsc It should be noted that the remaining oil at each depletion stage is subjected to continual compositional changes as it becomes progressively richer in the heavier components. In addition, the above described type of DL is characterized by a varying composition of the total hydrocarbon system; whereas the total system composition is kept unchanged during the CCE test. The combined data from the DL and CCE tests are considered sufficient to simulate the separation process taking place in the reservoir. Consider the illustration in Fig. 4.31 which depicts a possible depletion scenario of an oil reservoir and pressure profile around the production well. In region “A”, the reservoir pressure “pr” is greater than the bubble point pressure “pb” indicating an undersaturated oil system with the oil as the only moving phase, which is migrating (possibly by expansion) to region “B”. In region “B,” the reservoir pressure is slightly below the bubble point pressure, which signifies the start of the liberation of solution gas. However, the gas saturation “Sg” is presumably below the critical gas saturation “Sgc” indicating that the gas is immobile and will remain in contact with the oil, which resembles the CCE separation process. Therefore, region “B” is appropriately simulated by the data obtained from the flash liberation test (CCE test). Region “C” represents the pressure profile near the wellbore where the pressure is approaching the bottom-hole flowing pressure “pwf.” A large pressure sink occurs in region “B” causing the liberated gas to exceed Sgc allowing the gas to flow. Because the gas has higher mobility than oil, the liberated gas begins to flow leaving behind the oil that originally contained it. Consequently, this behavior follows the DL sequence. However, it should be pointed out that in the tubing and surface separation facilities, the gas and oil flow simultaneously in equilibrium resembling the flash expansion process. In general, the reservoir fluid flow dynamic is modeled by the combined flash and DL tests. PVT Measurements 375 Flash separation Oil Gas Well pr <<pb pr ≤ pb pr > pb Sg > Sgc Sg ≤ Sgc Sg = 0 Only oil is flowing Only oil is flowing Oil+gas are flowing Pressure C A B pb Oil+gas are flowing Only oil is flowing Only oil is flowing pwf Distance from wellbore Pressure profile around wellbore n FIGURE 4.31 Reservoir phase distribution in an oil reservoir. The experimental data obtained from the DL test include n n n n n n n n The differential oil FVF “Bod” The differential gas solubility; that is, solution GOR, “Rsd” Gas FVF “Bg” in ft3/scf Gas compressibility factor “Z” Total FVF “Btd” Composition of the liberated gas Gas specific gravity Density of the remaining oil as a function of pressure Differential Oil FVFs The differential oil FVFs, Bod (commonly called the relative oil volume factors), are calculated at all pressure levels by dividing the recorded oil volumes, VL, by the volume of residual oil, VLsc: Bod ¼ VL VLsc (4.112) 376 CHAPTER 4 PVT Properties of Crude Oils EXAMPLE 4.33 A DL test was conducted on a crude oil sample at 180°F. The sample exhibited a bubble point pressure of 3565 psig. The residual (remaining) oil volume was corrected to SC of 14.7 psia and 60°F to give 60 cm3. The following measurements were obtained from a differential expansion (DE) test: Pressure (psig) Total Volume (cm3) Liquid Volume (cm3) Gas Removed (scf) 3565 3000 2400 1800 1200 600 200 0 99.29514779 105.1180611 100.8688861 98.43652432 100.8219751 118.801706 168.9280437 63.85313132 99.29514779 92.66229773 87.02573049 82.03645568 77.70667102 73.45739134 70.05558226 63.85313132 0 0.086792948 0.07660421 0.067170194 0.061887145 0.059245621 0.043396474 0.061509785 Bold number represents boiling point pressure. Using the above reported DE measurements data, calculate the differential oil FVF “Bod”; that is, relative oil volume factor. Solution Calculate Bod by applying Eq. (4.112) to give Bod ¼ VL VL ¼ VLsc 60 Pressure (psig) Total Volume (cm3) Liquid Volume (cm3) 3565 3000 2400 1800 1200 600 200 0 0 psig, 60°F 99.29514779 105.1180611 100.8688861 98.43652432 100.8219751 118.801706 168.9280437 63.85313132 99.29514779 92.66229773 87.02573049 82.03645568 77.70667102 73.45739134 70.05558226 63.85313132 60 Gas Removed (scf) 0 0.086792948 0.07660421 0.067170194 0.061887145 0.059245621 0.043396474 0.061509785 Bod (bbl/ STB) 1.6549 1.7520 1.6811 1.6406 1.6804 1.9800 2.8155 1.0642 1.0000 Bold number represents boiling point pressure. Differential Gas Solubility “Rsd” The gas solubility (ie, differential solution GOR, “Rsd”) at any pressure is defined as the volume of gas that is dissolved (remaining) in solution at the pressure per stock-tank barrel of the corrected residual oil. PVT Measurements 377 Mathematically, the differential solution GOR, Rsd, is calculated by dividing the volume of gas remaining (dissolved) in solution at the specified pressure by the residual oil volume. EXAMPLE 4.34 Using the DE measurements data in Example 4.33, calculate the differential solution GOR “Rsd.” Solution Step 1. Convert the volume of the corrected residual oil volume as expressed in cubic centimeter to stock-tank barrel: Stock tank volume ¼ 60 ð30:48Þ3 5:615 ¼ 0:000377STB Step 2. Calculate TOTAL volume of the dissolved gas: Total volume of the dissolved gas ¼ 0.456606 scf Step 3. Calculate Rsd by dividing the volume of the dissolved gas at each pressure by 0.000377, to give Pressure (psig) Total Volume (cm3) Liquid Volume (cm3) Gas Removed (scf) 3565 3000 2400 1800 1200 600 200 0 0 psig, 60°F 99.29514779 105.1180611 100.8688861 98.43652432 100.8219751 118.801706 168.9280437 63.85313132 99.29514779 92.66229773 87.02573049 82.03645568 77.70667102 73.45739134 70.05558226 63.85313132 60 0 0.086792948 0.07660421 0.067170194 0.061887145 0.059245621 0.043396474 0.061509785 Total ¼ 0.456606 scf Bold number represents boiling point pressure. Gas FVF “Bg” The gas FVF is the defined ratio of the volume “(Vgas)p,T” occupied by n moles of gas at a specified pressure and temperature to the volume “(Vgas)sc” occupied by the same number of moles (ie, n moles) at SC. Mathematically, this is defined by Vgas p, T 3 ; ft =scf Bg ¼ Vgas sc (4.113) Dissolved (Gas scf) Rsd (scf/STB) 0.4566 0.3698 0.2932 0.2260 0.1642 0.1049 0.0615 0.0000 1210 980 777 599 435 278 163 0 378 CHAPTER 4 PVT Properties of Crude Oils EXAMPLE 4.35 Using the DE measurements data in Example 4.33, calculate the gas FVF. Solution As an Example of calculating Bg at 3000 psig: n Volume of the free gas at 3000 psig ¼ total volume-volume of the liquid Vgas 3000, 180 ¼ ð105:1180611 92:66229773Þ=ð30:48Þ3 ¼ 0:00044ft3 n Calculate Bg at 3000 psig and 180°F by applying the definition of this gas property: Bg ¼ 0:00044=0:086792948 ¼ 0:005068ft3 =scf n Pressure (psig) Total Volume (cm3) 3565 3000 2400 1800 1200 600 200 0 0 psig, 60°F 99.29514779 105.1180611 100.8688861 98.43652432 100.8219751 118.801706 168.9280437 63.85313132 The tabulated values of Bg are given below: Liquid Volume (cm3) Gas Removed (scf) Free Gas (ft3) Bg (ft3/scf) 99.29514779 92.66229773 87.02573049 82.03645568 77.70667102 73.45739134 70.05558226 63.85313132 60 0 0.086792948 0.07660421 0.067170194 0.061887145 0.059245621 0.043396474 0.061509785 Total ¼ 0.456606 scf No free gas 0.00044 0.00049 0.00058 0.00082 0.00160 0.00349 0.005068 0.006382 0.008622 0.01319 0.027028 0.080459 Bold number represents boiling point pressure. Gas Deviation Factor “Z” Using the volume of the free gas at any pressure and the corresponding volume at SC, the gas deviation factor can be calculated by applying the gas ESO at both conditions; that is, p Vgas p, T ¼ ZnRT to give : n ¼ p Vgas p, T =ðZRT Þ psc Vgas sc ¼ Zsc nRTsc to give : n ¼ p Vgas sc =ðZsc RTsc Þ where Zsc ¼ 1 Tsc ¼520°R T ¼ 640°R psc ¼ 14.7 psia PVT Measurements 379 Equating the above two expressions and solving for Z, gives Z¼ p Vgas p, T T T sc psc Vgas sc EXAMPLE 4.36 Using the DE measurements data in Example 4.33, calculate the gas FVF. Solution As a sample of the Z-factor calculations at 3000 psig: p Vgas p, T T sc Z¼ T psc Vgas sc Z¼ ð3000 + 14:7Þð0:00044Þ 520 ¼ 0:8445 180 + 460 14:7ð0:08679Þ The tabulated values of Z as a function of pressure are given below: Pressure (psig) Total Volume (cm3) Liquid Volume (cm3) Gas Removed (scf) Free Gas (ft3) Z 3565 3000 2400 1800 1200 600 200 0 0 psig, 60°F 99.29514779 105.1180611 100.8688861 98.43652432 100.8219751 118.801706 168.9280437 63.85313132 99.29514779 92.66229773 87.02573049 82.03645568 77.70667102 73.45739134 70.05558226 63.85313132 60 0 0.086792948 0.07660421 0.067170194 0.061887145 0.059245621 0.043396474 0.061509785 Total ¼ 0.456606 scf No free gas 0.00044 0.00049 0.00058 0.00082 0.00160 0.00349 0.844483 0.85174 0.864839 0.885583 0.918313 0.954804 Bold number represents boiling point pressure. Total FVF “Btd” From DE Test The two-phase (total) FVF from DE test is determined by applying the definition of this property as expressed mathematically by Eq. (4.69). In terms of differential data, Btd is given by where Btd ¼ Bod + ðRsbd Rsd ÞBg Rsbd ¼ gas solubility at the bubble point pressure, scf/STB Rsd ¼ gas solubility at any pressure, scf/STB Bod ¼ oil FVF at any pressure, bbl/STB Bg ¼ gas FVF, bbl/scf The generated PVT data in Examples 4.33–4.35 are used to calculate B. 380 CHAPTER 4 PVT Properties of Crude Oils EXAMPLE 4.37 Using solution data from Examples 4.33–4.35, calculate the total FVF. Solution Applying Eq. (4.115), to calculated Btd gives Btd ¼ Bod + ðRsbd Rsd ÞBg Btd ¼ Bod + ð1210 Rsd ÞBg Notice, Bg must be expressed in bbl/scf where 1 bbl ¼ 5.615. For example, at 3000 psig Btd ¼ 1:544 + ð1210 980Þ ð0:00507=5:615Þ ¼ 1:752bbl=STB Pressure (psig) Rsd (scf/STB) Bod (bbl/STB) Bg (ft3/scf) Btd (bbl/STB) 3565 3000 2400 1800 1200 600 200 0 1210 980 777 599 435 278 163 0 1.655 1.544 1.450 1.367 1.295 1.224 1.168 1.064 0.00507 0.00638 0.00862 0.01319 0.02703 0.08046 1.655 1.752 1.943 2.306 3.116 5.711 16.170 Bold number represents boiling point pressure. Table 4.6 shows results of the DL test conducted on a crude oil bottom-hole sample taken from the Big Butte Field in Montana. The graphical presentation of the data in terms of the relative oil volume and differential GOR are presented in Figs. 4.32 and 4.33, respectively. The test indicates that the differential GOR and differential relative oil volume at the bubble point pressure are 933 scf/STB and 1.730 bbl/STB, respectively. The symbols Rsdb and Bodb are used to represent these two values: Rsdb ¼ 933 scf=STB Bodb ¼ 1:730 bbl=STB Column 4 of Table 4.6 shows the relative total volume, Btd, from DL as calculated from the following expression: Btd ¼ Bod + ðRsdb Rsd ÞBg PVT Measurements 381 Table 4.6 Differential Liberation Data (Differential Vaporization at 247°F) Pressure (psig) 1 Rsda 2 Bodb 3 Btdc 4 ρ (g/cm3) 5 Z 6 Bgd 7 Incremental γ g 8 pb > 1936 1700 1500 1300 1100 900 700 500 300 185 120 0 933 841 766 693 622 551 479 400 309 242 195 0 1.73 1.679 1.639 1.6 1.563 1.525 1.486 1.44 1.382 1.335 1.298 1.099 1.73 1.846 1.982 2.171 2.444 2.862 3.557 4.881 8.138 13.302 20.439 0.7745 0.6484 0.6577 0.665 0.672 0.679 0.6863 0.6944 0.7039 0.7161 0.7256 0.7328 2.375 0.864 0.869 0.876 0.885 0.898 0.913 0.932 0.955 0.97 0.979 0.01009 0.01149 0.01334 0.01591 0.01965 0.02559 0.03626 0.06075 0.09727 0.14562 0.885 0.894 0.901 0.909 0.927 0.966 1.051 1.23 1.423 1.593 At 60°F ¼ 1.000. Gravity of residual oil ¼ 34.6°API at 60°F. Density of residual oil ¼ 0.8511 g/cm3 at 60°F. a Cubic feet of gas at 14.73 psia and 60°F per barrel of residual oil at 60°F. b Barrels of oil at indicated pressure and temperature per barrel of residual oil at 60°F. c Barrels of oil and liberated gas at indicated pressure and temperature per barrel of residual oil at 60°F. d Cubic feet of gas at indicated pressure and temperature per cubic feet at 14.73 psia and at 60°F. n FIGURE 4.32 Relative volume vs. pressure. 382 CHAPTER 4 PVT Properties of Crude Oils Solution GOR “Rsd” at 247°F 1000 Rsbd Solution GOR (scf/STB) “Rsd” 800 600 R sd 400 200 0 0 500 1000 1500 2000 Pressure n FIGURE 4.33 Solution gas-oil ratio vs. pressure. The gas deviation factor, Z, listed in column 6 represents the Z-factor of the liberated (removed) solution gas at the specific pressure. These values are calculated from the recorded gas volume measurements as follows: Z¼ ! Vgas p, T p T Tsc ! Vgas sc psc Column 7 of Table 4.6 that contains the gas FVF, Bg, which can be also determined in terms of the Z-factor by applying Eq. (3.52); that is, psc ZT Bg ¼ Tsc p Quality Checks of DE Data n Oil density from DE and CCE tests Plotting the saturated oil density data (ie, (p pb) from the DL test and those of the undersaturated density from the CCE should exhibit a smooth transition from the saturated to the undersaturated density values, as shown schematically in Fig. 4.34. A steep increase or erratic density values around the saturation pressure might indicate flow assurance problems; for example, asphaltene precipitation. PVT Measurements 383 Pressure-Volume Relations (at 247 °F) Differential Vaporization (at 247 °F) Pressure (psig) b>>1936 1700 1500 1300 1100 900 700 500 300 185 120 0 Relative Oil Volume Bod (B) Solution Gas/Oil Ratio Rsd (A) 933 1.730 841 1.679 766 1.639 693 1.600 622 1.563 551 1.525 479 1.486 400 1.440 309 1.382 242 1.335 195 1.298 0 1.099 @ 60⬚F = 1.000 Relative Total Volume Btd (C) 1.730 1.846 1.982 2.171 2.444 2.862 3.557 4.881 8.138 13.302 20.439 Gas Formation Incremental Gas Oil Deviation Volume Gravity Factor Density Factor 3 (Air = 1.000) (D) (g/cm ) Z 0.6484 0.6577 0.6650 0.6720 0.6790 0.6863 0.6944 0.7039 0.7161 0.7256 0.7328 0.7745 0.01009 0.01149 0.01334 0.01591 0.01965 0.02559 0.03626 0.06075 0.09727 0.14562 0.864 0.869 0.876 0.885 0.898 0.913 0.932 0.955 0.970 0.979 0.885 0.894 0.901 0.909 0.927 0.966 1.051 1.230 1.423 1.593 2.375 De Density ns ity e? en fro m Pressure (psig) 6500 6000 5500 5000 4500 4000 3500 3000 2500 2400 2300 2200 2100 2000 b>>1936 1930 1928 1923 1918 1911 1878 1808 1709 1600 1467 1313 1161 1035 782 600 437 DE lt ha p As rob y from Densit pb CCE Pressure n FIGURE 4.34 Quality checks of oil density from DE and CCE. n Quality checks of saturated density The saturated density from any particular separation test should be calculated and compared with the reported laboratory data. All the reported and calculated saturated density values should fall within a maximum of 1% of each other; otherwise, the validity of the test should be reconsidered. The following mathematical expression resulting from the mass material balance on the crude oil is used to calculate the saturated oil density: " # n 5 X 1 141:5 ρodb ¼ + 1:483 10 psc γ gi + 1 ðRsd Þi ðRsd Þi + 1 Bodb 131:5 + °APIRO i¼1 (4.114) where ρodb ¼ oil density in g/cm3 at pb from DE test, g/cm3 APIRO ¼ residual oil API gravity from DE test Relative Volume (A) 0.9371 0.9422 0.9475 0.9532 0.9592 0.9657 0.9728 0.9805 0.9890 0.9909 0.9927 0.9947 0.9966 0.9987 1.0000 1.0014 1.0018 1.0030 1.0042 1.0058 1.0139 1.0324 1.0625 1.1018 1.1611 1.2504 1.3694 1.5020 1.9283 2.4960 3.4464 Y-Function (B) Density (g/cm3) 0.6919 0.6882 0.6843 0.6803 0.6760 0.6714 0.6665 0.6613 0.6556 0.6544 0.6531 0.6519 0.6506 0.6493 0.6484 2.108 2.044 1.965 1.874 1.784 1.710 1.560 1.453 1.356 384 CHAPTER 4 PVT Properties of Crude Oils Bodb ¼ oil FVF at pb from DE test, bbl/STB Rsd ¼ gas solubility from DE test, scf/STB psc ¼ standard pressure (eg, 14.7), psia EXAMPLE 4.38 Using the following DE test data listed in Table 4.7, validate the saturated oil density. Solution Calculate the saturated density by applying Eq. (4.114), to give " # n 5 X 1 141:5 + 1:483 10 psc γ gi + 1 ðRsd Þi ðRsd Þi + 1 ρodb ¼ Bodb 131:5 + °APIRO i¼ 1 ρodb ¼ 1 141:5 + 1:483 105 ð14:7Þ½0:7559ð1050 924Þ 1:602 131:5 + 34:9 + 0:7376ð924 737Þ + 0:7272ð737 547Þ + 0:7448ð547 359Þ + 0:8649ð359 195Þ + 1:9116ð195 0Þ ¼ 0:6704g=cm3 The absolute error is 0.02% indicating a reliable data set. n Quality checks of the total FVF from DE and CCE tests The two-phase FVF “Bt” as a function of pressure can be calculated and compared graphically with those generated from the DE test to validate the quality of these tests. The two-phase FVF from the CCE test can be calculated below the saturation pressure from the following expression: ðBt ÞCCE ¼ Bodb Vrel (4.115) Table 4.7 DE Test Data for Example 4.38 Differential Liberation at 182°F Physical Properties i psia Bo Rs Z Bg Sp.gr. Density (g/cm3) 1 2 3 4 5 6 7 3009 2610 2012 1408 807 310 15 1.602 1.55 1.475 1.4 1.324 1.242 1.063 1050 924 737 547 359 195 0 0.777 0.779 0.802 0.855 0.929 1 0.00538 0.00699 0.01026 0.01891 0.05206 0.7559 0.7376 0.7274 0.7448 0.8649 1.9116 0.6699 0.6789 0.6929 0.7083 0.7265 0.7487 0.7994 Residual oil gravity at STP: 0.8494 [34.9°API]. Bold numbers represent boiling point pressure. PVT Measurements 385 Plotting (Bt)CCE and Btd as a function of pressure on the same plot is used to evaluate the quality of both tests. By visual inspection, if the two curves are overlapped it will indicate acceptable CCE and DE tests. EXAMPLE 4.39 Perform QCs on the total FVF using the DE and CCE data listed in Table 4.8. Table 4.8 DE and CCE Test Data for Example 4.39 DE Test Data p Rsd Bod Btd 3565 3000 2400 1800 1200 600 200 0 0 1210 980 777 599 435 278 163 0 0 1.655 1.655 1.544 1.752 1.450 1.943 1.367 2.306 1.295 3.116 1.224 5.713 1.168 16.180 1.064 1.064219 Bold number represents boiling point pressure. CCE Test Data Density 0.6397 0.6614 0.6823 0.7035 0.7231 0.7437 0.7610 0.7794 0.7794 Z Bg Sp.gr 0.843 0.850 0.863 0.884 0.916 0.953 0.00507 0.00638 0.00862 0.01319 0.02703 0.08046 0.741 0.720 0.710 0.712 0.757 0.876 1.662 1.661785 Pressure (psig) Relative Volume Density (g/cm3) 6500 6475 6023 5524 5019 4528 4027 3722 3669 3626 3604 3585 3565 3559 3540 3519 3480 3445 3356 3197 2938 2556 2078 1554 1151 876 709 565 0.9677 0.9679 0.9714 0.9759 0.9810 0.9866 0.9931 0.9976 0.9984 0.9990 0.9994 0.9997 1.0000 1.0005 1.0019 1.0036 1.0068 1.0098 1.0179 1.0341 1.0664 1.1328 1.2656 1.5353 1.9425 2.4526 2.9633 3.6454 0.6611 0.6609 0.6585 0.6555 0.6521 0.6484 0.6441 0.6413 0.6407 0.6403 0.6401 0.6399 0.6397 386 CHAPTER 4 PVT Properties of Crude Oils 18.000 16.000 14.000 12.000 10.000 Bt 8.000 DE Bt • CCE Bt 6.000 4.000 2.000 0.000 0 1000 2000 3000 Pressure 4000 Pressure Relative Density CCE Bt (psig) 6500 6475 6023 5524 5019 4528 4027 3722 3669 3626 3604 3585 3565 3559 3540 3519 3480 3445 3356 3197 2938 2556 2078 1554 1151 876 709 565 Volume 0.9677 0.9679 0.9714 0.9759 0.9810 0.9866 0.9931 0.9976 0.9984 0.9990 0.9994 0.9997 1.0000 1.0005 1.0019 1.0036 1.0068 1.0098 1.0179 1.0341 1.0664 1.1328 1.2656 1.5353 1.9425 2.4526 2.9633 3.6454 (g/cm3) 0.6611 0.6609 0.6585 0.6555 0.6521 0.6484 0.6441 0.6413 0.6407 0.6403 0.6401 0.6399 0.6397 (bbl/STB) 1.6014 1.60172 1.60767 1.61505 1.62349 1.6328 1.64358 1.65091 1.65223 1.65333 1.6539 1.6544 1.65492 1.6557 1.65811 1.6609 1.66623 1.67116 1.68453 1.71132 1.76487 1.87462 2.09449 2.54087 3.21465 4.05887 4.90396 6.0328 n FIGURE 4.35 Total formation volume factor from DE and CCE test data for Example 4.39. Solution n Calculate the Bt using the CCE test data by applying Eq. (4.115). ðBt ÞCCE ¼ Bodb Vrel ðBt ÞCCE ¼ 1:655Vrel Validate both tests by plotting (Bt)CCE and (Bt)DE as a function of pressure as shown in Fig. 4.35. n Fig. 4.35 shows an excellent match indicating quality laboratory data. n Separator Tests It should be noted that the plots presented in Figs. 4.32 and 4.33 for Bod and Rsd data as a function of pressure resemble the traditional appearance of the oil FVF “Bo” and the solution gas solubility “Rs” curves, leading to their misuse in reservoir calculations. The DL test simulates the behavior of the crude oil and the liberation of the solution gas in the reservoir; however, in the applications of the material balance equation and other reservoir engineering mathematical tools require properties of the produced fluid as related to the surface separation facilities, not those of the DE test. It should be noted that the values of Rsd and Bod from the DE test are determined based on using the PVT Measurements 387 n FIGURE 4.36 Impact of surface facilities on DE data. reservoir residual oil volume as a reference volume, whereas the oil FVF “Bo” and gas solubility “Rs” are based on using the volume of the residual oil at separator stock-tank conditions as the reference volume. As shown in Fig. 4.36, the process of producing the hydrocarbon system through the surface facilities must be included and accounted for to generate the required PVT properties of all aspects of reservoir engineering calculations. Separator tests are conducted to determine the changes in the volumetric behavior of the reservoir fluid as the fluid passes through the separator (or separators) and then into the stock-tank. The resulting volumetric behavior is influenced to a large extent by the operating conditions of the surface separation facilities in terms of their pressures and temperatures. The primary objective of conducting separator tests, therefore, is to provide the essential laboratory information necessary for determining the optimum surface separation conditions, which in turn maximize the stock-tank oil production. The laboratory flash separator unit is designed to simulate a full-scale field separation equipment. The laboratory unit usually operates from atmospheric pressure up to 2000 psig and temperatures from 40 to 250°F. The lab separation unit is used to determine the 388 CHAPTER 4 PVT Properties of Crude Oils optimum conditions for field separation by calculating the changes in FVF and GORs with relation to changes in separator pressure and temperature. In addition, it provides n n n volume of gas liberated at each stage, composition and specific gravity of the liberated gas from each stage, and composition of the stock-tank liquid and specific gravity. Results of the flash separation test, when appropriately combined with the DL test data, provide a means of obtaining the PVT parameters (Bo, Rs, and Bt) required for petroleum engineering calculations. The test involves placing a hydrocarbon sample at its saturation pressure and reservoir temperature in a PVT cell. The volume of the sample is measured as Vsat. The hydrocarbon sample then is displaced and flashed through a laboratory multistage separator system, commonly one to three stages. The pressure and temperature of these stages are set to represent the desired or actual surface separation facilities. The gas liberated from each stage is removed to measure its specific gravity, composition, and volume at SC. The volume of the remaining oil in the last stage (representing the stock-tank condition) is measured and recorded as (Vo)st. These experimental measured data can then be used to determine the oil FVF and gas solubility at the bubble point pressure as follows: Vsat ðVo Þst Vg sc Rsfb ¼ ðVo Þst Bofb ¼ (4.116) (4.117) where Bofb ¼ bubble point oil FVF, as measured by flash liberation, bbl of the bubble point oil/STB Rsfb ¼ bubble point solution GOR as measured by flash liberation, scf/STB (Vg)sc ¼ total volume of gas removed from separators, scf This laboratory procedure is repeated at a series of different separator pressures and a fixed temperature. It usually is recommended that four of these tests be used to determine the optimum separator pressure, which usually is considered the separator pressure that results in minimum oil FVF. At the same pressure, the stock-tank oil gravity is at a maximum and the total evolved gas (ie, the separator gas and the stock-tank gas) is at a minimum. A typical example of a set of separator tests for a two-stage separation system is shown in Table 4.9. The tabulated values in Table 4.9 indicate that PVT Measurements 389 Table 4.9 Separator Tests Separator Pressure (psig) Temperature (°F) GOR,a Rsfb °API at 60°F FVF,b Bofb 50 0 Total 100 0 Total 200 0 Total 300 0 Total 75 75 737 41 778 676 92 768 602 178 780 549 246 795 40.5 1.481 40.7 1.474 40.4 1.483 40.1 1.495 75 75 75 75 75 75 a GOR in cubic feet of gas at 17.65 psia and 60°F per barrel of stock-tank oil at 60°F. FVF is barrels of saturated oil at 2.620 psig and 220°F per barrel of stock-tank oil at 60°F. b optimum separator pressure is around 100 psia and is considered to be the separator pressure that results in the minimum oil FVF. It is important to note that the oil FVF varies from 1.474 to 1.495 bbl/STB, while the gas solubility ranges from 768 to 795 scf/STB. Table 4.9 indicates that the values of the crude oil PVT data depend on the method of surface separation. Table 4.10 presents the results of performing a separator test on the Big Butte crude oil. The DL data, as expressed in Table 4.6, shows that the solution GOR at the bubble point is 933 scf/STB, as compared with the measured value of 646 scf/STB from the separator test. This significant difference, as also shown in Figs. 4.37 and 4.38, is attributed to the fact that the processes of obtaining residual oil and stock-tank oil from bubble point oil are different. The DL is considered a multiple series of flashes at the elevated reservoir temperatures; the separator test generally is a one- or two-stage flash at low pressure and low temperature. The quantity of gas released is different, and the quantity of final liquid is different. Again, it should be pointed out that oil FVF, as expressed by Eq. (4.116), is defined as “the volume of oil at reservoir pressure and temperature divided by the resulting stock-tank oil volume after it passes through the surface separators.” Adjustment of DE Data to Separator Conditions The material balance calculations and black oil reservoir simulation require the PVT data of the crude oil including the oil FVF “Bo” and gas solubility “Rs” as a function of the pressure. Amyx et al. (1960) Table 4.10 Separator Test Data, Separator Flash Analysis Separator Flash Analysis Gas/Oil Ratio (scf/bbl) Flash Conditions (psig) (°F) (A) Gas/Oil Ratio (scf/STB) (B) StockTank Oil Gravity at 60°F (°API) Formation Volume Factor Bofb (C) Separator Volume Factor Specific Gravity of Flashed Gas (D) (Air 5 1.000) 1936 247 28 0 130 80 Oil Phase Density 0.6484 593 13 Rsfb 632 13 38.8 1.527 Rsfb = 646 GOR1 = 632 1.066 1.010 1.132* ** Bofb GOR2 = 13 Rsfb = GOR1 + GOR2 = 646 scf/STB Pb Bofb = (V)pb,T VST T, pb Tsep = 130°F psep = 28 psig n FIGURE 4.37 A comparison between Bodb and Bofb. Tsc = 80 psc = 0 psig (V)pb,T (V)ST = 1.527 bbl / STB 0.7823 0.8220 PVT Measurements 391 n FIGURE 4.38 A comparison between Rsdb and Rsfb. proposed a procedure for constructing the oil FVF and gas solubility curves using the DL data in conjunction with the experimental separator flash data for a given set of separator conditions. The method is summarized in the following steps. Step 1. Calculate the differential shrinkage factors at various pressures by dividing each relative oil volume factors, Bod, by the relative oil volume factor at the bubble point, Bodb: Sod ¼ Bod Bodb (4.118) where Bod ¼ differential relative oil volume factor at pressure p, bbl/STB Bodb ¼ differential relative oil volume factor at the bubble point pressure pb, psia, bbl/STB Sod ¼ differential oil shrinkage factor, bbl/bbl of bubble point oil The differential oil shrinkage factor has a value of 1 at the bubble point and a value less than 1 at subsequent pressures below pb. Step 2. Adjust the relative volume data by multiplying the separator (flash) FVF at the bubble point, Bofb, by the differential oil shrinkage factor, Sod, at each pressures. Mathematically, this relationship is expressed as follows: 392 CHAPTER 4 PVT Properties of Crude Oils Bo ¼ Bofb Bod ¼ Bofb Sod Bodb (4.119) where Bo ¼ oil FVF, bbl/STB Bofb ¼ oil FVF at pb from separator test Sod ¼ differential oil shrinkage factor, bbl/bbl of bubble point oil Step 3. Calculate the oil FVF at pressures above the bubble point pressure by multiplying the relative oil volume data, Vrel, as generated from the CCE test, by Bofb: Bo ¼ ðVrel ÞðBofb Þ (4.120) where Bo ¼ oil FVF above the bubble point pressure, bbl/STB Vrel ¼ relative oil volume, bbl/bbl. Step 4. Adjust the differential gas solubility data, Rsd, to give the required gas solubility factor, Rs: Rs ¼ Rsfb ðRsdb Rsd Þ Bofb Bodb (4.121) where Rs ¼ gas solubility, scf/STB Rsfb ¼ bubble point solution GOR from the separator test, scf/STB Rsdb ¼ solution GOR at the bubble point pressure as measured by the DL test, scf/STB Rsd ¼ solution GOR at various pressure levels as measured by the DL test, scf/STB It should be pointed out the Eqs. (4.119), (4.121) usually produce values less than 1 for Bo and negative values for Rs at low pressures. The calculated curves of Bo and Rs versus pressures must be manually drawn to Bo ¼ 1.0 and Rs ¼ 0 at atmospheric pressure. McCain (2002) suggested the following simple expression for adjusting the differential gas solubility data, Rsd, to give the required gas solubility factor, Rs: Rs ¼ Rsd Rsfb Rsdb Step 5. Obtain the two-phase (total) FVF, Bt, by multiplying values of the relative oil volume, Vrel, below the bubble point pressure by Bofb: Bt ¼ ðBofb ÞðVrel Þ (4.122) PVT Measurements 393 where Bt ¼ two-phase FVF, bbl/STB Vrel ¼ relative oil volume below the pb, bbl/bbl. Similar values for Bt can be obtained from the DL test by multiplying the relative total volume, Btd by Bofb: Bt ¼ ðBtd ÞðBofb Þ=Bodb (4.123) EXAMPLE 4.40 The CCE test, DL test, and separator test for the Big Butte crude oil system are given in Tables 4.5, 4.6, and 4.10, respectively. Calculate the oil FVF at 4000 and 1100 psi, the gas solubility at 1100 psi, and the two-phase FVF at 1300 psi. Solution Step 1. Determine Bodb, Rsdb, Bofb, and Rsfb from Tables 4.9 and 4.10: Bodb ¼ 1:730 bbl=STB Rsdb ¼ 933 scf=STB Bofb ¼ 1:527 bbl=STB Rsfb ¼ 646 scf=STB Step 2. Calculate Bo at 4000 by applying Eq. (4.120): Bo ¼ ðVrel ÞðBofb Þ Bo ¼ ð0:9657Þð1:57Þ ¼ 1:4746 bbl=STB Step 3. Calculate Bo at 1100 psi by applying Eqs. (4.118), (4.119): Sod ¼ Bod Bodb Sod ¼ 1:563 0:9035 1:730 and Bo ¼ Bofb Sod Bo ¼ ð0:9035Þð1:5277Þ ¼ 1:379 bbl=STB Step 4. Calculate the gas solubility at 1100 psi using Eq. (4.121): Bofb Bodb 1:527 Rs ¼ 646 ð933 622Þ ¼ 371 scf=STB 1:730 Rs ¼ Rsfb ðRsdb Rsd Þ Note: Using the correction proposed by McCain gives Rsfb 646 ¼ 622 ¼ 431 scf=STB Rs ¼ Rsd Rsdb 933 394 CHAPTER 4 PVT Properties of Crude Oils Step 5. From the pressure-volume relations (ie, constant composition data) of Table 4.5, the relative volume at 1300 psi is 1.2579 bbl/bbl. Using Eq. (4.122), calculate Bt to give Bt ¼ ðBofb ÞðVrel Þ Bt ¼ ð1:527Þð1:2579Þ ¼ 1:921 bbl=STB Equivalently applying Eq. (4.123) gives Bt ¼ ðBtd ÞðBofb Þ=Bodb Bt ¼ ð2:171Þð1:527Þ=1:73 ¼ 1:916 bbl=STB Table 4.11 presents a complete documentation of the adjusted differential vaporization data for the Big Butte crude oil system. Figs. 4.39 and 4.40 compare graphically the adjusted values of Rs and Bo with those of unadjusted values. Note that no adjustments are needed for the gas FVF, oil density, or viscosity data. Dodson et al. (1953) proposed perhaps the most ideal laboratory procedure for obtaining crude oil PVT data. The methodology is commonly called the Dodson method or, alternatively, the composite liberation test. The proposed procedure is summarized below and is schematically illustrated in Fig. 4.41: n n n n n A large crude oil sample is placed in a PVT cell at reservoir temperature and bubble point pressure. A small sample of the crude oil is withdrawn to conduct a separator test and generate Bofb and Rsfb. The pressure is reduced below the saturation pressure, the evolved gas is removed at constant pressure, and its composition and properties are determined. A small oil sample is again withdrawn from the remaining oil and pumped through the separators to obtain Bo and Rs and other properties at this particular reduced pressure. The above process should be repeated at several progressively lower reservoir pressures until complete curves of Bo and Rs versus reservoir pressure have been obtained. It should be noted that the Dodson procedure is an expensive laboratory procedure and, therefore, is rarely conducted. Quality Checks of Separator Test Data. The saturated density as obtained from the separator test should be calculated and compared with the reported PVT Measurements 395 Table 4.11 Differentiation Liberation Data Adjusted to Separator Conditions Pressure (psig) Rsa Bob 6500 6000 5500 5000 4500 4000 3500 3000 2500 2100 2300 2200 2100 2000 b > 1936 1700 1500 1300 1100 900 700 500 300 185 120 0 646 646 646 646 646 646 646 646 646 646 646 646 646 646 646 564 498 434 371 309 244 175 95 36 1.431 1.439 1.447 1.456 1.465 1.475 1.486 1.497 1.51 1.513 1.516 1.519 1.522 1.525 1.527 1.482 1.446 1.412 1.379 1.346 1.311 1.271 1.22 1.178 1.146 Gas FVFc 0.01009 0.01149 0.01334 0.01591 0.01965 0.02559 0.03626 0.06075 0.09727 0.14562 Oil Density (g/cm3) 0.6919 0.6882 0.6843 0.6803 0.676 0.6714 0.6665 0.6613 0.6556 0.6544 0.6531 0.6519 0.6506 0.6493 0.6484 0.6577 0.665 0.672 0.679 0.6863 0.6944 0.7039 0.7161 0.7256 0.7328 0.7745 Separator conditions: First stage ¼ 28 psig at 130°F. Stock-tank ¼ 0 psig at 80°F. a Cubic feet of gas at 14.73 psia and 60°F per barrel of stock-tank oil at 60°F. b Barrel of oil at indicated pressure and temperature per barrel of stock-tank oil at 60°F. c Cubic feet of gas at indicated pressure and temperature per cubic feet at 14.73 psia and 60°F. laboratory data. The reported and calculated saturated density value should fall within a maximum of 1% of each other; otherwise, the validity of the test should be reconsidered. The following mathematical expression resulting from the mass material balance on the crude oil is used to calculate the saturated oil density: ρob ¼ " # n X 1 141:5 + 1:483 105 psc γ g GOR sep Bofb 131:5 + °APIST sep¼1 Oil/Gas Viscosity Ratio 19 21.3 23.8 26.6 29.8 33.7 38.6 46 52.8 58.4 396 CHAPTER 4 PVT Properties of Crude Oils n FIGURE 4.39 Adjusted gas solubility vs. pressure. n FIGURE 4.40 Adjusted oil formation volume factor vs. pressure. where ρob ¼ oil density in g/cm3 APIST ¼ stock-tank API gravity Bofb ¼ oil FVF at pb from the separator test bbl/STB (GOR)sep ¼ separator GOR, scf/STB γ g ¼ liberated gas gravity from each separator stage PVT Measurements 397 n FIGURE 4.41 Composite DE test “Dodson Procedure.” EXAMPLE 4.41 Table 4.12 results from a separator test performed on a bottom-hole crude oil sample. The test was conducted using a three-stage plus stock-tank laboratory separation arrangement. The test shows a saturated oil density of 0.669 g/cm3 at the bubble point pressure of 3009 psi and 182°F. The following additional measured data were obtained: n n Bofb ¼ 1.465 bbl/STB Rsfb ¼ 839 scf/STB The residual oil specific gravity and API gravity at stock-tank conditions are 0.8313 and 38.72, respectively. Validate the saturated oil density. Solution Calculate the saturated density from the following density expression: " # n 5 X 1 141:5 ρob ¼ + 1:483 10 psc γ g GOR sep Bofb 131:5 + °APIST sep¼1 to give ρob ¼ 1 141:5 + 1:483 105 ð14:7Þ½587 0:757 1:465 131:5 + 38:72 + 114 0:76 + 104 0:997 + 34 1:395 ¼ 0:669 398 CHAPTER 4 PVT Properties of Crude Oils Table 4.12 Separator Test Data, Separator Flash Analysis Stage Pressure Temperature Rs (scf/STB) Bo Gas Gravity ρo (g/cm3) 1 2 3 Stock-tank 3009 796 317 65 15 182 134 80 80 60 839 587 114 104 34 1.465 1.226 1.139 1.082 1 0.757 0.76 0.997 1.395 0.669 0.7206 0.7587 0.7781 0.8313 The calculated density is an exact match with the reported laboratory value, confirming the validity of the experimental saturated oil density. Extrapolation of Reservoir Fluid Data In partially depleted reservoirs or fields that originally existed at the bubble point pressure, it is difficult to obtain a fluid sample that represents the oil in the reservoir at the time of discovery. Also, in collecting fluid samples from oil wells, the possibility exists of obtaining samples with a saturation pressure that might be lower than or higher than the actual saturation pressure of the reservoir. In these cases, it is necessary to correct or adjust the laboratory PVT measured data to reflect the actual or desired saturation pressure, including n n n n constant composition expansion DE test oil viscosity data separator test data Correcting CCE Data. The correction procedure summarized in the following steps is based on calculating the Y-function value for each point below the “old” saturation pressure. Step 1. n Calculate the Y-function, as expressed by Eq. (4.107), for each point by using the old saturation pressure: Y¼ n n pold sat p pðVrel 1Þ Plot the values of the Y-function versus pressure on a Cartesian scale and draw the best straight line. Notice that points in the neighborhood of the saturation pressure may be erratic and need not be used. Calculate the coefficients a and b of the straight-line equation: Y ¼ a + bp PVT Measurements 399 n Recalculate the relative volume, Vrel, values by applying Eq. (4.109) and using the “new” saturation pressure and the coefficients “a&b” from step 3: Vrel ¼ 1 + pnew sat p pða + bpÞ (4.124) Step 2. To determine points above the “new” saturation pressure, proceed by the following steps: n Plot the “old” relative volume values above the “old” saturation pressure versus pressure on a regular scale and draw a smooth curve connecting points. IF a straight line is obtained, draw the best straight line through these points and determine the slope of line S. Note that the slope is negative; that is, S < 0. n Draw a smooth curve or straight line that passes through the point with coordinate Vrel ¼ 1, pnew and parallel to the line/curve sat of previous step. n Relative volume data above the new saturation pressure are read from the curve. If a straight line, generate the new relative volume data from the following expression at any pressure p: Vrel ¼ 1 S pnew sat p (4.125) where S ¼ slope of the line and p ¼ pressure. EXAMPLE 4.42 The pressure-volume relations of the Big Butte crude oil system are given in Table 4.5. The test indicates that the oil has a bubble point pressure of 1930 psig at 247°F. The Y-function for the oil system is expressed by the following linear equation: Y ¼ 1:0981 + 0:000591p Above the bubble point pressure, the relative volume data versus pressure exhibits a straight-line relationship with a slope, S, of 0.0000138. The field surface production and well testing data indicate that the actual bubble point pressure is approximately 2500 psig. Please reconstruct the pressure-volume data using the new reported saturation pressure. Solution Use Eqs. (4.124), (4.125): pnew 2500 p sat p ¼ 1+ pða + bpÞ pð1:0981 + 0:000591pÞ new Vrel ¼ 1 S psat p ¼ 1 ð0:0000138Þð2500 pÞ Vrel ¼ 1 + to give the results in the table below. 400 CHAPTER 4 PVT Properties of Crude Oils Pressure (psig) Old Vrel New Vrel Comments 6500 6000 5000 4000 3000 (pb)new 5 2500 2000 (pb)old 5 1936 1911 1808 1600 600 437 0.9371 0.9422 0.9532 0.9657 0.9805 0.9890 0.9987 1.0000 1.0058 1.0324 1.1018 2.4960 3.4404 0.9448 0.9517 0.9655 0.9793 0.9931 1.0000 1.1096 1.1299 1.1384 1.1767 1.1018 2.4960 3.4404 Eq. (4.125) Eq. (4.125) Eq. (4.125) Eq. (4.125) Eq. (4.125) Eq. (4.124) Eq. (4.124) Eq. (4.124) Eq. (4.124) Eq. (4.124) Eq. (4.124) Eq. (4.124) Bold number represents boiling point pressure. Correcting DL Data. The laboratory-measured relative oil volume Bod data must be corrected to account for the new bubble point pressure pnew b . The proposed procedure is summarized in the following steps: Step 1. Plot the Bod data versus gauge pressure on a regular scale. Step 2. Draw the best straight line through the middle pressure range of 30–90% pb. Step 3. Extend the straight line to the new bubble point pressure, as shown schematically in Fig. 4.42. Step 4. Transfer any curvature at the end of the original curve, that is, ΔBo1 at pold b , to the new bubble point pressure by placing ΔBo1 above or below the straight line at pnew b . Step 5. Select any differential pressure Δp below the pold b and transfer the corresponding curvature ΔBo1 to the pressure (pnew b –Δp). Step 6. Repeat this process and draw a curve that connects the generated Bod points with the original curve at the point of intersection with the straight line. Below this point no change is needed. The correction procedure for the solution to the GOR, Rsd, data is identical to that of the relative oil volume data. Correcting Oil Viscosity Data. The oil viscosity data can be extrapolated to a new, higher bubble point pressure by applying the following steps. Step 1. Defining the “fluidity” as the reciprocal of the oil viscosity (ie, 1/ μo), calculate the fluidity for each point below the original saturation pressure. Step 2. Plot fluidity versus pressure on a Cartesian scale (see Fig. 4.43). PVT Measurements 401 New (Bod)@ new pb Original (Bod)@ old pb + ΔB1 ΔB1 Original (Bod)@ old pb ΔB2 Original Bo curve ΔB1 ΔB2 Bod Δp Δp 0.9 pb 0.3 pb (pb)old n FIGURE 4.42 Adjusting Bod curve to reflect new bubble point pressure. . 1/m o . . . mo . (m ob)old (m ob)new 0 n FIGURE 4.43 Extrapolation of oil viscosity to new pb. Pressure (pb)old (pb)new (pb)new 402 CHAPTER 4 PVT Properties of Crude Oils mo A (m ob)old B (m ob)new 0 Pressure (pb)old (pb)new n FIGURE 4.44 Extrapolation of oil viscosity above the new pb. Step 3. Draw the best straight line through the points and extend it to the new saturation pressure, pold b . Step 4. New oil viscosity values above pold b are read from the straight line. To obtain the oil viscosity for pressures above the new bubble point pressure follow the following steps: pnew b Step 1. Plot the viscosity values for all points above the old saturation pressure on a Cartesian coordinate, as shown schematically in Fig. 4.44, and draw the best straight line through them, line A. draw Step 2. Through the point on the extended viscosity curve at pnew b a straight line (line B) parallel to A. Step 3. Viscosities above the new saturation pressure then are read from line A. Correcting the Separator Tests Data n Stock-tank GOR and gravity No corrections are needed for the stock-tank GOR and the stock-tank API gravity. n Separator GOR The gas solubility at the new bubble point pressure (Rsfb)new is changed in the same proportion as the differential ratio was changed; that is, old Rnew sfb ¼ Rsfb new Rsfb Rold sdb (4.126) PVT Measurements 403 n The separator GOR then is the difference between the new (corrected) gas solubility Rnew sfb and the unchanged stock-tank GOR. Formation volume factor The oil FVF at the new bubble point pressure “Bofb” is adjusted in the same proportion as the DL value: old Bnew ofb ¼ Bofb new Bodb Bold odb (4.127) EXAMPLE 4.43 Results of the DL and the separator tests on the Big Butte crude oil system are given in Tables 4.6 and 4.10, respectively. The field data indicates that the actual bubble point value is 2000 psi. The correction procedure for Bod and Rsd as described previously was applied, to give the following values at the new bubble point of 2000 psi: Bnew odb ¼ 2:013bbl=STB Rnew sbd ¼ 1134scf=STB Using the separator test data given in Table 4.10, calculate the gas solubility and the oil FVF at the new bubble point pressure. Solution Gas solubility is found from Eq. (4.126): new 1134 old Rsdb ¼ 646 ¼ R ¼ 785scf=STB Rnew sfb sfb 933 Rold sdb Separator GOR ¼ 785 13 ¼ 772scf=STB The oil FVF is found by applying Eq. (4.127): new old Bodb ¼ B Bnew ofb ofb Bold odb 2:013 Bob ¼ 1:527 ¼ 1:777bbl=STB 1:730 Laboratory Analysis of Gas-Condensate Systems A standard laboratory analysis on a gas-condensate sample consists of n n n recombination and analysis of the separator samples, measuring the pressure-volume relationship; that is, CCE test, and constant volume depletion (CVD) test. The above listed routine laboratory tests are detailed next. 404 CHAPTER 4 PVT Properties of Crude Oils Recombination of Separator Samples Obtaining a representative sample of the reservoir fluid is considerably more difficult for a gas-condensate fluid than for a conventional black oil reservoir. The principal reason for this difficulty is that the liquid may condense from the reservoir fluid during the sampling process, and if representative proportions of both liquid and gas are not recovered, then an erroneous composition is calculated. Because of the possibility of erroneous compositions and the limited volumes obtainable, subsurface sampling is seldom used in gas-condensate reservoirs. Instead, surface sampling techniques are used, and samples are obtained only after long-stabilized flow periods. During this stabilized flow period, volumes of liquid and gas produced in the surface separation facilities are accurately measured, and the fluid samples are then recombined in these proportions. The hydrocarbon composition of separator samples also are determined by chromatography, low-temperature fractional distillation, or a combination of both. Table 4.13 shows the Table 4.13 Reservoir Fluid Composition Basis of Recombination: 7407 scf Separator Gas/Barrel Separator Liquid Component Symbol Name N2 CO2 H2S Nitrogen Carbon dioxide Hydrogen sulfide Methane Ethane Propane i-Butane n-Butane i-Pentane n-Pentane Hexanes Heptanes Octanes Nonanes Decanes Undecanes Dodecanes Tridecanes C1 C2 C3 i-C4 n-C4 i-C5 n-C5 C6 C7 C8 C9 C10 C11 C12 C13 Reservoir Fluid (mol%) Separator Gas (mol%) Separator Liquid (mol%) Separator Liquid (wt.%) Molecular Weight Specific Gravity (Water 5 1.0) 0.063 1.154 0.000 0.002 0.470 0.000 0.001 0.242 0.000 28.01 44.01 34.08 0.809 0.818 0.801 0.055 1.069 0.000 80.655 11.141 3.935 0.941 1.087 0.385 0.234 0.191 0.090 0.074 0.040 0.010 16.006 8.871 7.900 3.501 5.491 4.330 4.112 7.754 7.754 9.629 5.953 4.110 2.824 2.053 1.819 3.002 3.118 4.073 2.379 3.731 3.652 3.468 7.812 9.272 12.057 8.325 6.445 4.854 3.864 3.722 16.04 30.07 44.10 58.12 58.12 72.15 72.15 86.18 95.90 107.07 119.51 134.17 147.00 161.00 175.00 0.300 0.356 0.507 0.563 0.584 0.624 0.631 0.664 0.706 0.738 0.766 0.779 0.790 0.801 0.812 72.624 10.859 4.428 1.259 1.634 0.875 0.716 1.130 1.105 1.261 0.775 0.519 0.351 0.255 0.226 (Continued) PVT Measurements 405 Table 4.13 Reservoir Fluid Composition Continued Basis of Recombination: 7407 scf Separator Gas/Barrel Separator Liquid Component Separator Gas (mol%) Symbol Name C14 C15 C16 C17 C18 C19 C20 C21 C22 C23 C24 C25 C26 C27 C28 C29 C30+ Tetradecanes Pentadecanes Hexadecanes Heptadecanes Octadecanes Nonadecanes Eicosanes Heneicosanes Docosanes Tricosanes Tetracosanes Pentacosanes Hexacosanes Heptacosanes Octacosanes Nonacosanes Triacontanes plus Total Molecular weight 100.000 20.56 Separator Liquid (mol%) Separator Liquid (wt.%) 1.417 1.120 0.833 0.713 0.576 0.484 0.339 0.280 0.239 0.186 0.147 0.128 0.096 0.075 0.060 0.047 0.171 3.148 2.698 2.161 1.976 1.691 1.489 1.091 0.951 0.854 0.691 0.569 0.518 0.401 0.329 0.271 0.219 0.929 100.000 85.54 100.00 Molecular Weight Specific Gravity (Water 5 1.0) 190.00 206.00 222.00 237.00 251.00 263.00 275.00 291.00 305.00 318.00 331.00 345.00 359.00 374.00 388.00 402.00 465.22 0.823 0.833 0.840 0.848 0.853 0.858 0.863 0.868 0.873 0.878 0.882 0.886 0.890 0.894 0.897 0.900 0.914 Reservoir Fluid (mol%) 0.176 0.139 0.103 0.089 0.072 0.060 0.042 0.035 0.030 0.023 0.018 0.016 0.012 0.009 0.007 0.006 0.021 100.000 28.64 Compositional Groupings of Reservoir Fluid Group mol% wt.% MW SG Tb Total fluid C7+ C10+ C20+ C30+ 100.000 5.350 2.210 0.220 0.021 100.000 26.113 14.464 2.531 0.345 28.64 139.76 187.45 330.17 465.22 0.781 0.822 0.882 0.914 N/A N/A N/A 1232 1410 Bold number represents boiling point pressure. hydrocarbon analyses of the separator liquid and gas samples taken from the Nameless field. The samples were combined with a GOR of 7407 scf per barrel of separator liquid at separator conditions of 640 psia and 100°F. The separator liquid density is 0.683 g/cm3 or 42.62 lb/ft3. The laboratory recombined data indicates that the overall well stream system contains 72.624 mol% methane and 5.35 mol% heptanes-plus. 406 CHAPTER 4 PVT Properties of Crude Oils The Consistency Test on the Recombined GOR As detailed earlier, a concise QC test on the laboratory-reported GOR can be performed by solving the component material balance relationship for the ratio “nL/nv” from yi nt nL xi ¼ zi nv nv zi which suggests plotting “yi/zi” versus “xi/zi” would produce a straight line with a negative slope value of “nL/nv.” The GOR can then be calculated from the following expression: GOR ¼ 2130:331ðρo Þpsep , Tsep ðMWÞLiquid ðnL =nv Þ ; scf=bbl Fig. 4.45 shows the straight line resulting from plotting “yi/zi” versus “xi/zi” using the laboratory gas and liquid samples listed in Table 4.13. The slope of the line in the figure is 0.1421, which is converted into GOR, to give GOR ¼ GOR ¼ 2130:331ðρo Þpsep , Tsep ðMWÞLiquid ðnL =nv Þ 2130:331ð43:62Þ ¼ 7467 scf=bbl ð85:54Þð0:1421Þ which shows a very acceptable match with the reported GOR of 7407 scf/bbl with an absolute error of 0.8%. yi / zi Symbol N2 CO2 C1 C2 C3 i-C4 n-C4 i-C5 n-C5 C6 C7 C8 C9 C10 C11 C12 C13 xi / zi n FIGURE 4.45 Recombined GOR consistency check. Molecular weight Density, g/cm3 Separator Gas yi Separator Liquid xi Reservoir Fluid zi (mol %) (mol %) (mol %) 0.063 1.154 80.655 11.141 3.935 0.941 1.087 0.385 0.234 0.191 0.090 0.074 0.040 0.010 0.002 0.470 16.006 8.871 7.900 3.501 5.491 4.330 4.112 7.754 8.264 9.629 5.953 4.110 2.824 2.053 1.819 0.055 1.069 72.624 10.859 4.428 1.259 1.634 0.875 0.716 1.130 1.105 1.261 0.775 0.519 0.351 0.255 0.226 34.13 85.54 0.683 xi /zi yi /zi 0.03636 0.43966 0.22040 0.81693 1.78410 2.78078 3.36047 4.94857 5.74302 6.86195 7.47873 7.63600 7.68129 7.91908 1.14545 1.07951 1.11058 1.02597 0.88866 0.74742 0.66524 0.44000 0.32682 0.16903 0.08145 0.05868 0.05161 0.01927 PVT Measurements 407 Separator Pressure: Separator Temperature: Separator Components Liquid (mol%) Methane Ethane Propane i-Butane n-Butane i-Pentane n-Pentane Hexane Heptane Octane Nonane 640 100 Separator Gas (mol%) 16.006 8.871 7.900 3.501 5.491 4.330 4.112 7.754 8.264 9.629 5.953 80.655 11.141 3.935 0.941 1.087 0.385 0.234 0.191 0.090 0.074 0.040 psia °F K = yi /xi log(K*p) Tb (°R) 5.039 1.256 0.498 0.269 0.198 0.089 0.057 0.025 0.011 0.008 0.007 200.95 343.01 332.21 549.74 415.94 665.59 470.45 734.08 490.75 765.18 541.76 828.63 556.56 845.37 615.37 911.47 668.74 970.57 717.84 1023.17 763.07 1070.47 3.509 2.905 2.503 2.236 2.103 1.755 1.561 1.198 0.843 0.692 0.634 Tc (°R) pc (psia) b b(1/Tb-1/T) 667.0 707.8 615.0 527.9 548.8 490.4 488.1 439.5 397.4 361.1 330.7 803.9 1412.6 1798.2 2037.3 2151.2 2383.7 2478.2 2795.3 3079.2 3344.4 3592.9 3.2694 2.9676 2.6880 2.4779 2.4272 2.2323 2.1991 2.0005 1.8044 1.6177 1.4412 4.0 C1 3.0 log(k*p) C2 C3 i-C4 n-C4 2.0 i-C5 n-C5 1.0 C6 C7 0.0 –2.5 C8 C9 –2.0 –1.5 –1.0 –0.5 0.0 0.5 b(1/Tb - 1/T) 1.0 n FIGURE 4.46 Hoffman plot consistency check on gas-condensate system. The Hoffman plot for the separator samples should be made to evaluate the consistency of the recombined fluid samples. Using the data in Table 4.13, the Hoffman plot shown in Fig. 4.46 indicates good consistency in the laboratory measurement data. Constant Composition Test The experimental test procedure, as shown schematically in Fig. 4.46, is similar to that conducted on a crude oil system that involves measuring the pressure-volume relations of the reservoir fluid at reservoir temperature with a PVT visual cell. The PVT cell allows the visual observation of the 1.5 2.0 2.5 3.0 408 CHAPTER 4 PVT Properties of Crude Oils condensation process and the measurement of liquid dropout (LDO) volume as the cell pressure is reduced below the saturation pressure. In general, the CCE test on a gas-condensate sample is performed to provide the engineer with the fundamental pressure-volume relationship of a retrograde gas system and other properties, including n n The dew point pressure “pd.” The relative volume “Vrel,” which is defined as the ratio of the total measured volume of the hydrocarbon system “Vt” to that of the volume at the dew point pressure “Vsat.” This property is defined by Eq. (4.105) as Vrel ¼ n Vt Vsat The liquid volume “VL” at each pressure below the dew point pressure is reported into two different dimensionless forms. The first form is based on the reference volume Vsat and other is based on Vt; that is, ðLDOÞCCE Vsat ¼ VL Vsat (4.128) ðLDOÞCCE Vt ¼ VL Vt (4.129) It should be pointed out the above two expressions are interrelated through the relative volume by the following relationship: CCE ðLDOÞCCE Vsat ¼ Vrel ðLDOÞVt n (4.130) The calculation of the traditional Y-function is not valid or applicable using the relative volume from the CCE test on retrograde gas. However, the two forms of extended Y-function along with the relative volume can be used to confirm the reported dew point pressure. The two extended Y-functions were given earlier by Eqs. (4.111a), (4.111b) as " # pi p YEXT ¼ log pi ðVrel ÞEXT 1 " # pi p YEXTp ¼ log p ðVrel ÞEXT 1 where ðVrel ÞEXT ¼ n Vrel ðVrel Þpmax The gas compressibility factor and gas density at the dew point pressure are calculated from the volumetric, mass, and number of moles using the following basic relationships: Zd ¼ pd Vsat nRT PVT Measurements 409 ρd ¼ m pd Ma ¼ Vsat Zd RT The Z-factor and gas density at pressures above the saturation pressure are calculated from Z ¼ Zd p ðVrel Þ pd ρ¼ ρd Vsat where Zd ¼ gas compressibility factor at the dew point pressure, pd pd ¼ dew point pressure, psia m ¼ weight of the sample Ma ¼ apparent molecular weight of the gas sample p ¼ pressure, psia It should be noted that the above relationships are only valid for p pd. Table 4.14 shows the dew point determination and the pressure-volume relations of the Nameless field. The dew point pressure of the system is reported as 4100 psi at 320°F. Table 4.14 Pressure-Volume Relations “CCE” of a Gas-Condensate System at 320°F Pressure (psia) Relative Volume (V/Vsat) Gas Density (g/cm3) 11,150 10,000 9000 8000 7000 6000 5000 4250 4100 4000 3800 3600 3500 3000 2500 2000 0.616 0.641 0.666 0.698 0.739 0.795 0.880 0.978 1.000 1.023 1.064 1.125 1.153 1.302 1.538 1.919 0.389 0.374 0.360 0.344 0.325 0.302 0.273 0.245 0.240 Relative Liquid Volume (%) 0.00 0.08 0.28 0.52 0.65 1.42 2.23 2.98 Deviation Factor (z) 1.572 1.466 1.371 1.276 1.183 1.090 1.006 0.950 0.938 (Continued) 410 CHAPTER 4 PVT Properties of Crude Oils Table 4.14 Pressure-Volume Relations “CCE” of a Gas-Condensate System at 320°F Continued Pressure (psia) Relative Volume (V/Vsat) 1500 1100 2.587 3.614 Gas Density (g/cm3) Relative Liquid Volume (%) Deviation Factor (z) 3.74 4.33 Notes: Relative volume ( V/ Vsat) is the fluid volume at the indicated pressure and temperature relative to the saturated fluid volume. Specific volume (ft3/lb) ¼ 1/density (g/cm3) 62.428. Density (lb/ft3) ¼ Density (g/cm3) 62.428. Relative liquid volume % is the volume of liquid at indicated p and T/total volume at saturation pressure. Bold numbers represent boiling point pressure. EXAMPLE 4.44 Using the data in Table 4.14, confirm the reported dew point pressure of 4100 psi. Solution Step 1. Calculate (Vrel)EXT, (Yrel)EXT, and (Yrel)EXTp with results as tabulated below. Pressure (psia) Relative Volume (Vt/Vsat) Gas Density (g/cm3) 11,150 10,000 9000 8000 7000 6000 5000 4250 4100 4000 3800 3600 3500 3000 2500 2000 1500 1100 0.616 0.641 0.666 0.698 0.739 0.795 0.880 0.978 1.000 1.023 1.064 1.125 1.153 1.302 1.538 1.919 2.587 3.614 0.389 0.374 0.360 0.344 0.325 0.302 0.273 0.245 0.240 h ðLDOÞVsat 0.00 0.08 0.28 0.52 0.65 1.42 2.23 2.98 3.74 4.33 iCCE (%) Z-Factor (Vrel)EXT YEXT YEXTp 1.572 1.466 1.371 1.276 1.183 1.090 1.006 0.950 0.938 1.00000 1.04058 1.08117 1.13312 1.19968 1.29058 1.42857 1.58766 1.62338 1.66071 1.72727 1.82630 1.87175 2.11364 2.49675 3.11526 4.19968 5.86688 1.39164 1.09061 0.87577 0.69968 0.53673 0.36798 0.23087 0.20525 0.17999 0.13830 0.08286 0.05961 0.04674 0.17515 0.32536 0.50511 0.68725 0.45234 0.46881 0.47099 0.47263 0.47038 0.45788 0.44133 0.44065 0.43223 0.42481 0.40451 0.39920 0.38729 0.36393 0.33503 0.30333 0.27352 Bold number represents boiling point pressure. Step 2. Plot Vrel, (Vrel)EXT, (Yrel)EXT, and (Yrel)EXTp as a function of pressure, as shown in Figs. 4.47 and 4.48, to give a dew point pressure of 4100 psi, confirming the reported value. PVT Measurements 411 100% gas p < pd pd pi > pd Vt Vsat Vt VL Hg Vrel = Vt Vsat (LDO)Vsat = VL Vsat (LDO)VTot = VL Vt n FIGURE 4.47 A schematic illustration of constant composition expansion (CCE) test. n FIGURE 4.48 Dew point pressure from CCE test data for Example 4.44. p << pd p << pd Vt VL Vt VL 412 CHAPTER 4 PVT Properties of Crude Oils CVD Test CVD experiments are performed on gas condensates and volatile oils to simulate reservoir depletion performance and the compositional variations in the producing gas. The test is designed to simulate LDO behavior in the pore as pressure drops below the saturation pressure. The basic assumption in the CVD test is that retrograde LDO will not have a sufficient liquid saturation that is greater than critical saturation to become mobile. In addition, it is also assumed that the reservoir pore volume will not change or impact the results of the test; hence, the name “constant volume depletion test.” The test provides a variety of useful and important information that is used in reservoir engineering and field processing calculations. The laboratory procedure of the test is shown schematically in Fig. 4.49 and summarized in the following steps: Step 1. A measured amount of a representative sample of the original reservoir fluid with a known overall composition of zi is charged to a visual PVT cell at the dew point pressure, as shown in Fig. 4.49A. The temperature of the PVT cell is maintained at the reservoir temperature, T, throughout the experiment. The initial volume at the saturation pressure “Vsat” is measured and used as a reference volume. It should be pointed out that the CVD test procedure n FIGURE 4.49 A schematic illustration of the constant volume depletion “CVD” test. PVT Measurements 413 can also be conducted on a volatile oil sample. In this case, the PVT cell initially contains liquid, instead of gas, at its bubble point pressure. Step 2. The initial gas compressibility factor is calculated from the real gas equation: Zd ¼ pd Vsat ni RT where pd ¼ dew point pressure, psia Vsat ¼ initial gas volume at dew point pressure, ft3 ni ¼ initial number of moles of the gas ¼ m/Ma R ¼ gas constant, 10.73 T ¼ temperature, °R Zd ¼ compressibility factor at dew point pressure m ¼ weight of the gas sample, lb Ma ¼ apparent molecular weight of the gas sample Step 3. The cell pressure is reduced from the saturation pressure to a predetermined level “p1” below the dew point pressure. This reduction in the pressure is achieved by withdrawing mercury from the cell, as illustrated in Fig. 4.49B. During this process and as the pressure drops below the dew point pressure, a second phase (retrograde liquid) is formed. The fluid in the cell is brought to equilibrium at p1 and the total volume (gas + liquid) is measured and recorded as “Vt1.” Step 4. Mercury is reinjected into the PVT cell at constant pressure “p1” while an equivalent volume of remaining equilibrium gas is simultaneously removed from the cell. When the initial volume “Vsat” is reached, mercury injection ceases. The volume of the retrograde liquid “VL” is visually measured as illustrated in Fig. 4.49C while the volume of the removed gas is measured and recorded as Vg1 ¼ Vt1 Vsat This step is designed to resemble the characteristics of the pressuredepletion process in retrograde gas reservoirs, whereas the retrograde liquid remaining immobile in the reservoir and the equilibrium gas is the only producing phase. Step 5. The removed gas is charged at constant pressure of p1 into a gasometer, where its volume is measured at SC and 414 CHAPTER 4 PVT Properties of Crude Oils recorded as (Vgp)sc. A gas chromatographic analysis is also performed to determine the composition of the removed gas, yi. The number of moles of gas produced is calculated from the expression np ¼ psc Vgp sc RTsc where np ¼ moles of gas produced (Vgp)sc ¼ volume of gas produced measured at SC, scf Tsc ¼ standard temperature, °R psc ¼ standard pressure, psia R ¼ 10.73 Step 6. The gas compressibility factor at cell pressure of p1 and temperature is calculated from the real gas EOS as follows: Z¼ p1 ðVt1 Vsat Þ np RT Step 7. Volume of the retrograde liquid “VL” is reported as a fraction of Vsat and referred to as liquid dropout and given by LDO ¼ VL Vsat where VL ¼ liquid volume in the PVT cell at any depletion pressure Vsat ¼ volume at dew point pressure Step 8. Another important property that is routinely reported is termed the two-phase compressibility factor “Ztwo-phase.” The two-phase compressibility factor represents the total compressibility of all the remaining fluids (gas and retrograde liquid) in the cell and is computed at pressure “p” from the real gas law as Ztwo-phase ¼ pVsat ni np RT (4.131) where (ni np) ¼ represents the remaining moles of fluid in the cell at pressure p ni ¼ initial moles in the cell np ¼ cumulative moles of gas removed Step 9. The volume of gas produced is usually expressed as a percentage of gas initially in place, and is calculated by dividing PVT Measurements 415 the cumulative volume of the produced gas by the gas initially in place, both at SC. It can be treated as the recovery factor “RF.” %Gp ¼ " X # np ðni Þoriginal 100 Because 1 mol occupies 379.4 scf at SC, percentage of cumulative gas produced can be expressed as %Gp ¼ " X # 379:4np 379:4ðni Þoriginal 100 ¼ "X # Vgp sc GIIP 100 or commonly as RF ¼ Gp 100 GIIP where Gp ¼ cumulative gas removed (produced) from the PVT cell pressure p, scf GIIP ¼ volume of gas initially placed in the PVT at saturation pressure, scf RF ¼ recovery factor, % It should be pointed out that Eq. (4.131) can be expressed in a more convenient form by replacing moles of gas (ie, ni and np) with their equivalent gas volumes: Zd p Ztwo-phase ¼ pd 1 ðRFÞ (4.132) where Zd ¼ gas deviation factor at the dew point pressure pd ¼ dew point pressure, psia p ¼ reservoir pressure, psia RF ¼ gas recovery factor Step 10. This experimental procedure is repeated over several pressure increments until a minimum test pressure is reached, after which the quantity and composition of the gas and retrograde liquid remaining in the cell are determined. 416 CHAPTER 4 PVT Properties of Crude Oils DISCUSSION OF RETROGRADE GAS LABORATORY DATA Results from CCE and CVD laboratory tests on a surface sample taken from the Nameless field are presented in Tables 4.15 and 4.16, and Fig. 4.50. A detailed description of these tests are given below. n The CCE test data presented in Table 4.15 indicates a dew point pressure of 5022 psig at 218°F. Notice that column 4 lists the measured liquid volumes from the CCE test as expressed with reference to Vsat as given by Eq. (4.128); that is, ðLDOÞCCE Vsat ¼ VL Vsat Table 4.15 Constant Composition Expansion at 218°F 1 2 Pressure Vrel ¼ VVsatt 8000 7500 7000 6500 6000 5659 pres 5337 5022 psat 4713 4454 4034 3712 3328 3032 2660 2383 1987 1714 1511 1353 1226 839 0.87823 0.89166 0.90677 0.92442 0.94518 0.96149 0.97930 1.00000 1.02746 1.05552 1.11360 1.17160 1.26275 1.35464 1.50981 1.66647 1.98235 2.30023 2.61926 2.93902 3.25925 4.86448 3 Density (g/cm3) 0.44768 0.44094 0.43359 0.42531 0.41597 0.40891 0.40148 0.39317 5 6 4 ðLDOÞVsat ¼ VL =Vsat Z-Factor Eg (Mscf/bbl) 0.00% 14.25% 20.65% 25.53% 27.33% 28.35% 28.70% 28.32% 28.32% 27.33% 26.44% 25.71% 25.20% 24.70% 22.86% LDO is reported with Vsat as the reference volume. 1.42422 1.35579 1.28702 1.21855 1.15029 1.10381 1.06045 1.01912 1.63789 1.61302 1.58592 1.55540 1.52095 1.49492 1.46750 1.43688 Discussion of Retrograde Gas Laboratory Data 417 Table 4.16 Depletion Study at 218°F (Hydrocarbon Analyses of Produced Well Stream, mol%) Reservoir Pressure (psig) Well Stream Components (D.P.) 5022 mol% 4100 mol% 3200 mol% 2300 mol% 1400 mol% 500 mol% 0 mol% Hydrogen sulfide Nitrogen Carbon dioxide Methane Ethane Propane iso-Butane n-Butane iso-Pentane n-Pentane Hexanes Heptanes-plus Totals 0.002 0.084 1.163 65.410 12.107 5.329 1.280 2.039 0.911 0.902 1.247 9.526 100.000 0.000 0.091 1.201 69.857 12.331 5.256 1.196 1.862 0.760 0.760 1.023 5.662 100.000 0.000 0.095 1.231 72.736 12.474 5.005 1.121 1.721 0.690 0.690 0.887 3.350 100.000 0.000 0.097 1.263 74.286 12.942 4.895 1.046 1.567 0.635 0.598 0.760 1.910 100.000 0.000 0.096 1.295 74.907 13.243 5.033 1.049 1.557 0.570 0.537 0.675 1.038 100.000 0.000 0.087 1.307 72.742 13.552 6.053 1.318 1.983 0.706 0650 0.805 0.796 100.000 0.000 0.061 1.047 56.939 12.190 8.511 2.698 4.519 2.254 2.148 1.633 8.000 100.000 Heptanes-plus (C7+) fraction characteristics 1 Molecular weight 2 Specific gravity 154.383 0.8003 145.652 0.7917 143.256 0.7886 142.671 0.7874 141.163 0.7859 137.776 0.7830 133.233 1.2591 0.000 0.000 24.284 167.771 27.681 191.238 26.030 179.833 22.856 157.903 18.951 130.928 16.507 114.041 1.0191 1.0191 0.8922 0.9181 0.8411 0.8426 0.8354 0.7858 0.8586 0.7295 0.9207 0.5610 N/A N/A 9.316 22.803 40.399 60.762 81.436 92.789 6.662 5.224 4.251 93.16 Mscf 5.054 3.684 2.780 228.03 Mscf 4.019 2.679 1.848 403.99 Mscf 3.471 2.093 1.264 607.62 Mscf 3.969 2.312 1.262 814.36 Mscf 11.189 8.860 6.567 927.89 Mscf Condensed retrograde liquid volume 3 HC pore volume % 4 Bbls/MMscf of DP gas Gas deviation factor 5 Equilibrium gas 6 Two-phase Cumulative produced well stream volume 7 Vol% of initial DP Gas 0.000 GPM from CVD well stream compositions 8 Propane plus (C3+) Butanes plus (C4+) Pentanes-plus (C5+) 9 Cumulative recovery Gp based on GIIP ¼ 1000 Mscf n 9.471 8.013 6.957 0 Mscf Table 4.16 contains the gas chromatographic analysis of the produced gas stream from the CVD test, which clearly shows the expected reduction trend in the concentration of C7+ and the associated increase in the methane fraction concentration with pressure depletion. Rows 1 and 418 CHAPTER 4 PVT Properties of Crude Oils 30.0% 25.0% LDO 20.0% Pressure (psig) Liquid Dropout % 5022 4100 3200 2300 1400 500 0 0.000 24.284 27.681 26.030 22.856 18.951 16.507 15.0% 10.0% 5.0% 0.0% 0 1000 2000 3000 4000 5000 6000 Pressure (psig) n FIGURE 4.50 Retrograde condensation during gas depletion at 218°F. n n n 2 of Table 4.16 show the molecular weights and specific gravities of the C7+ in the produced well stream. With pressure depletion, the concentrations of intermediate components (ie, C2 through C6, in the produced gas stream) are seen to decrease as they condense forming part of the LDO in the PVT cell. At lower pressures, however, their concentrations in the produced gas increase as they revaporize back to the gas phase, resulting in a reduction in the volume of the LDO. Row 3 in Table 4.16 lists the retrograde liquid dropout “LDO” as a function of pressure. The LDO is expressed as a percentage of initial volume “Vsat” at the dew point pressure. Fig. 4.50 shows the traditional graphical presentation of the LDO as a function of pressure. The cell volume at the dew point pressure is used as a reference volume that essentially represents the reservoir pore volume. The figure shows the dropped-out liquid is vaporized back into the gas phase at lower pressure conditions. This behavior occurs as a result of the transfer of components between the LDO phase and gas phase. Some gas-condensate systems exhibit what is referred to as a tail, where the liquid drops out very slowly before increasing toward a maximum. The initial gradual and small buildup of the condensate phase has been attributed to the contamination of collected samples with hydraulic fluids from various sources during drilling, production, and sampling. There is no firm evidence, Discussion of Retrograde Gas Laboratory Data 419 however, that the trailing cannot be the true characteristic of a real reservoir fluid. The CVD measurements indicate that the maximum LDO of 27.7% occurs around 3200 psi. For most retrograde reservoirs, the maximum LDO occurs near 2000–2500 psi. Cho et al. (1985) correlated the maximum LDO (expressed in %) with reservoir temperature “T” (in °R) and the overall mole fraction of the heptanes-plus fraction “zC7 + ” in the dew point mixture by the following expression: ðLDOÞmax ¼ 93:404 + 479:9zC7 + 19:73 ln ðT 460Þ The above expression can only be used to approximate the maximum LDO. Using the key data from Table 4.16 applicable to the above relationship, gives ðLDOÞmax ¼ 93:404 + 479:9ð0:09526Þ 19:73 ln ð218Þ ¼ 32:9% n It should be noted that the CVD test is performed to simulate the LDO occurring in the reservoir porous media as the pressure drops below the dew point pressure. This retrograde condensation process is routinely performed on a hydrocarbon sample placed in a PVT cell representing the “pore volume”; however, without including initial reservoir water. The LDO can be converted into oil saturation by accounting for the presence of the initial water saturation, to give So ¼ ðLDOÞð1 Swi Þ where So ¼ retrograde liquid (oil) saturation, % LDO ¼ liquid dropout, % Swi ¼ initial water saturation, fraction Using the LDO data in Table 4.16 and assuming an initial water saturation of 30%, the calculated values of the oil saturation as a function of pressure can be tabulated from the data below. n p LDO So 5022 4100 3200 2300 1400 500 0 0 24.284 27.681 26.03 22.856 18.951 16.507 0 16.9988 19.3767 18.221 15.9992 13.2657 11.5549 Rows 5 and 6 in Table 4.16 display the equilibrium gas Z-factor and the two-phase Z-factor, respectively, resulting from the laboratory CVD 420 CHAPTER 4 PVT Properties of Crude Oils n test. It should pointed out that when applying the traditional gas material balance in terms of plotting (p/Z) versus Gp, the two-phase compressibility factor must be used instead of single-phase Z-factor when applying this analysis. Row 7 in Table 4.16 represents cumulative “wellstream produced as a percentage of initial gas-in-place at the dew point pressure.” This is the total recovery factor of the well stream; that is, RF ¼ n Gp 100 GIIP The recovery factor data from the CVD test can be used directly with field production data to calculate future field performance and reserves if the hydrocarbon pore volume does not change with declining reservoir pressure. Frequently, the surface gas is processed to remove and liquify all hydrocarbon components that are heavier than methane, such as ethane, propane, and heavier gases. These liquefied components are called plant products or the liquid yield. The quantities of liquid products are expressed in gallons of liquid per thousand standard cubic feet of gas processed “gal/Mscf” and usually denoted as “GPM.” The GPM is commonly reported for C3+ through C5+ groups in the produced well streams at each pressure-depletion stage, as shown in Row 8 of Table 4.16. The following expression can be used for calculating the anticipated GPM for each component in the gas phase: psc yi Mi GPMi ¼ 11:173 Tsc γ oi Assuming psc ¼ 14.7 psia and Tsc ¼ 520°R, the above relationship is reduced to yi Mi GPMi ¼ 0:31585 γ oi with a total plant product GPM as given by the summation of GPM of individual components: GPM ¼ X GPMi i For example, the liquid yield of the C3+ group at CVD depletion pressure stage k is given by ðGPMÞC3 + ¼ C7 + X i¼C3 ðGPMi Þk ¼ 0:31585 C7 + X yi Mi i¼ C3 γ oi k Discussion of Retrograde Gas Laboratory Data 421 where yi ¼ mole fraction of component i in the gas phase Mi ¼ molecular weight of component i k ¼ pressure-depletion stage γ oi ¼ specific gravity of component i as a liquid at SC (Chapter 1, Table 1.1, column E) The plant products, or the Yield “Y,” can also be expressed in terms of STB/MMscf, to give yi Mi YSTB=MMscf i ¼ 7:5224 γ oi It should be pointed out that liquefying and recovery of these products depend on the plant separation efficiency “EP.” Because a complete recovery of these products is not feasible, a rule of thumb indicates 5– 25% of ethane, 80–90% of the propane, 95% or more of the butanes, and 100% of the heavier components can be recovered from a simple surface facility. Including the plant separation efficiency “EP” in calculating the plant GPM gives ðGPMÞC3 + ¼ C7 + X i¼C3 n ðGPMi Þk ¼ 0:31585 C7 + X yi Mi i¼C3 γ oi EP (4.133) k Row 9 in Table 4.16 displays the cumulative recovery during depletion, assuming 1000 Mscf of gas initially in place at the dew point pressure (ie, G ¼ 1000 Mscf). The listed cumulative gas “Gp” is essentially the product of GIIP times RF at each depletion pressure. For example, cumulative gas at a depletion pressure of 3200 psi is Gp ¼ GðRFÞ Gp ¼ 1000ð0:22803Þ ¼ 228:03Mscf It should be pointed out that all cumulative gas production listed in Row 9 in Table 4.16 assumes that the reservoir pressure is initially at the dew point. Whitson and Brule (2000) point out that, if the initial reservoir pressure “pi” does not equal the dew point pressure, the assumed gas-inplace “G” of 1000 Mscf and tabulated cumulative gas values (ie, Row 9) can be adjusted by using the following methodology. Step 1. Calculate the adjusted gas-in-place (GIIP)adj at the initial reservoir pressure pi as given by ðGIIPÞadj ¼ ðGÞpsat + ΔG with the added correction to gas-in-place as given by 422 CHAPTER 4 PVT Properties of Crude Oils ΔG ¼ G ðpi =Zi Þi 1 ðpd =Zd Þ where ΔG ¼ additional GIIP G ¼ gas initially in place at the dewpoint; that is, 100 Mscf (p/Z)i ¼ value of (p/Z) at initial reservoir pressure “pi” (p/Z)d ¼ value of (p/Z) at dew point pressure “pd” For example, Table 4.15 indicates that the initial reservoir pressure is 5659 psig with a Z-factor of 1.10381 with a dew point pressure of 5022 psig. Applying the additional ΔG gives ½ð5659 + 14:7Þ=1:10381 ΔG ¼ 106 1 ¼ 40Mscf ½ð5022 + 14:7Þ=1:0191 Step 2. Calculate wet gas initially in place at initial reservoir pressure from GIIP ¼ G + ΔG to give GIIP ¼ 1000 + 40 ¼ 1040Mscf Step 3. Based on initial reservoir pressure, pi, adjust the tabulated cumulative gas production data by applying the following expressions: For pressures p > pd Gp ¼ ðGIIPÞ pi pd ðpd =Zd Þ Zi Zd For pressures p pd Gp adj ¼ Gp + ΔG When p < pd, the above expression suggests that the entire tabulated cumulative gas production (Row 9) for all depletion pressures should be increased by ΔG. For example, using the CCE data in Tables 4.15 and 4.16 to calculate the cumulative well stream gas production at 5337, 5022, and 3200 psig: n ΔG ¼ 40 Mscf and GIIP ¼ 1040 Mscf n At 5337 psig, the Z-factor from Table 4.15 is 1.06045, then Gp ¼ ðGIIPÞ pi pd ðpd =Zd Þ Zi Zd Discussion of Retrograde Gas Laboratory Data 423 Gp ¼ n n 1040 5673:7 5351:7 ¼ 20 Mscf ð5036:7=1:01912Þ 1:10381 1:06045 At 5022 psi, the cumulative gas production Gp ¼ 0 + 40 ¼ 40 Mscf At 3200 psig, the cumulative gas production from Table 4.16 is 228.03 Mscf, the adjusted cumulative gas: Gp adj ¼ Gp + ΔG Gp adj ¼ 228:03 + 40 ¼ 268:03 Mscf EXAMPLE 4.45 Table 4.16 shows the well stream compositional analysis of the Nameless field. Using Eq. (4.133), calculate the maximum plant liquefied components recovery ðGPMÞC3 + using the following plant recovery efficiency “EP”: n n n n 5% for ethane 80% for propane 95% for butanes 100% for heavier components Solution Using the composition of well stream at the dew point pressure, calculate the GPM for each component by applying Eq. (4.133), to give psc yi Mi GPMi ¼ 11:173 EP Tsc γ oi 15:025 yi Mi yi Mi EP ¼ 0:3228 EP GPMi ¼ 11:173 γ oi γ oi 520 Construct the following working table: Component yi Mi γ oi C1 C2 C3 i-C4 n-C4 i-C5 n-C5 C6 C7+ 65.41 12.107 5.329 1.28 2.039 0.911 0.902 1.247 9.526 30.07 44.097 58.123 58.123 72.15 72.15 86.177 154.383 0.35619 0.50699 0.56287 0.58401 0.6247 0.63112 0.66383 0.8003 EP EP (GPM)i 0.05 0.165 0.8 1.1973 0.95 0.4055 0.95 0.6225 1 0.3397 1 0.3330 1 0.5227 1 5.9337 ðGPMÞC3 + ¼ 9:5194 The calculated value of ðGPMÞC3 + is within 2.3% of the reported laboratory value of 9.471 GPM. 424 CHAPTER 4 PVT Properties of Crude Oils CONDENSATE GOR AND CONDENSATE YIELD As the reservoir pressure drops below the dew point pressure of a retrograde gas system, a large amount of the heavy hydrocarbon components constituting most of the LDO is left immobile in the reservoir rather than produced at the surface. The liquid phase accumulating in the reservoir is generally richer in these heavy ends while the produced gas is depleted of heavy ends resulting in less condensate yield at the surface. The increase in the reservoir LDO and the corresponding decrease in the condensate yield continues with decreasing pressure until the maximum LDO is reached. Reaching the pressure at which the maximum LDO occurs signifies the start of the vaporization of the LDO components to the gas phase. This reverse process is illustrated in Fig. 4.51. It is important to differentiate between the condensate yield “Y” and condensate-gas-ratio “CGR” and also to develop an expression to describe the relationship between these two field measurements. Based on the illustration presented in Fig. 4.52, the following relationships apply: 280 30.0% O LD 200 20.0% LDO (%) CGR = Q /Q o gas 160 15.0% Y=Q o /Qs tream 10.0% 120 5.0% 80 0.0% 0 1000 2000 3000 Pressure (psig) 4000 n FIGURE 4.51 LDO, condensate yield, and condensate-gas-oil ratio during gas depletion at 218°F. 5000 40 6000 Y & CGR (STB/MMscf) 240 25.0% Condensate GOR and Condensate Yield 425 Qg The stream from a “Well” or “PVT cell” CGR = Qstream Qo Qg Qo Y < CGR n FIGURE 4.52 LDO, condensate yield, and condensate-gas-oil ratio relationship. Condensate yield: Y¼ Qo ; STB=MMscf Qstream Condensate GOR: CGR ¼ Qo ; STB=MMscf Qg Gas equivalent volume: Veq ¼ 0:133000 γo ; MMscf=STB Mo Total volume of gas stream: Qstream ¼ Veq Qo + Qg ; MMscf Combining the above three relationships gives Y¼ Qo Veq Qo + Qg 426 CHAPTER 4 PVT Properties of Crude Oils or equivalently as Y¼ CGR Veq CGR + 1 The above expression indicates that Y < CGR. THE MODIFIED MBE FOR RETROGRADE GAS RESERVOIRS It should be pointed that the volumetric material balance for conventional gas reservoirs, commonly called the tank model, is based on the single-phase Z-factor. This simple tank model is formulated by applying the following gas mole balance: np ¼ ni nr where np ¼ moles of gas produced ni ¼ moles of gas initially in the reservoir nr ¼ moles of gas remaining in the reservoir the gas moles can be replaced by their equivalents using the real gas law, to give the following conventional material balance equation “MBE”: psc Gp pi V pV ¼ RTsc Zi RT Zg RT However, for condensate reservoirs, the remaining number of models must expressed in terms of the two-phase compressibility factor instead of the single-phase Z-factor “Zg”; that is, psc Gp pi V pV ¼ RTsc Zi RT Ztwo-phase RT Arranging this equation gives p Ztwo-phase ¼ pi psc T Gp Zi Tsc V The initial reservoir gas volume “V” can be expressed in terms of the volume of gas at SC by V ¼ Bg G ¼ psc Zi T G Tsc pi The Modified MBE for Retrograde Gas Reservoirs 427 Combining the above relationships gives p pi ¼ Ztwo-phase Zi pi 1 Gp Zi G The above relationship suggests that a plot of (p/Z) versus Gp would result in a straight line with a slope of m ¼ [(pi/Zi)(1/G)]. Expressing the above material balance in terms of recovery factor “RF” gives p Ztwo-phase ¼ pi pi RF Zi Zi (4.134) where pi ¼ initial reservoir pressure Gp ¼ cumulative gas production, scf p ¼ current reservoir pressure V ¼ original gas volume, ft3 Zi ¼ gas deviation factor at pi Zg ¼ single-phase gas deviation factor at p T ¼ temperature, °R Ahmed (2015) developed the following expression to calculated the liquid Z-factor “ZL” using the CVD data at any pressure: ZL ¼ Ztwo-phase Zg ðLDOÞ Zg + Ztwo-phase ðLDO 1Þ To account for the initial water saturation, the above relationship can be modified to give ZL ¼ Ztwo-phase Zg ðLDOÞð1 Swi Þ Zg + Ztwo-phase ½LDOð1 Swi Þ 1 The material balance can be expressed in terms of remaining gas moles “nrg” and remaining oil moles “nrL” as np ¼ ni nrg nrL to give 1 ðLDOÞ ðLDOÞ pi pi ¼ Gp p Zi GZi Zg ZL or in terms of RF, 1 ðLDOÞ ðLDOÞ pi pi p ¼ RF Zi Zi Zg ZL (4.135) 428 CHAPTER 4 PVT Properties of Crude Oils The above modified material balance equation can be used to accurately calculate the remaining fluid in the reservoir at abandonment and also to investigate the impact of initial water saturation on recovery. Consistency Tests on the CCE and CVD Data Tables 4.17 and 4.18 shows detailed CCE and CVD laboratory data performed on a recombined separator samples from a retrograde gas field. The CCE test shows a dew point pressure of 3763 psia with a saturated gas density and Z-factor of 0.3986 g/cm3 and 0.79416, respectively. There are several simple measures that can be used to check the quality of the CCE and CVD laboratory data. Some of these measures are applied next using the reported values in Tables 4.17 and 4.18. Table 4.17 CCE and CVD Depletion Test Data CCE Test Data p (psia) Rel. Vol., Vrel Z 5022 4660 4498 4348 4211 4085 3942 3763 3618 3475 3348 3236 3135 3044 2960 2884 2744 2623 2515 2415 2208 2039 1773 0.89862 0.92153 0.93309 0.94462 0.95594 0.96724 0.98105 1 1.02283 1.04834 1.07387 1.09894 1.12388 1.14846 1.17312 1.19724 1.24672 1.29555 1.34469 1.3956 1.52101 1.64873 1.91461 0.95243 0.9063 0.88577 0.86681 0.84956 0.83388 0.81617 0.79416 LDO 5 VL/Vsat (%) 0 11.4454677 20.5684308 25.0748645 26.9899664 28.3105372 29.2283070 29.9966784 30.5535648 31.1680000 31.5725535 31.7212371 31.7778120 31.7738989 31.5566922 30.9209515 Gas Density (g/cm3) 0.424 0.4135 0.4083 0.4034 0.3986 0.3939 0.3884 0.381 LDO 5 VL/Vt (%) 0 11.19 19.62 23.35 24.56 25.19 25.45 25.57 25.52 25.00 24.37 23.59 22.77 20.89 19.14 16.15 (Continued) The Modified MBE for Retrograde Gas Reservoirs 429 Table 4.17 CCE and CVD Depletion Test Data Continued CVD Test Data Pressure LDO Z-Factor Ztwo-phase % 3763 3413 3113 2813 2513 2213 1913 1613 1313 1013 513 0 22.35 27.26 29.22 29.82 29.43 28.68 27.67 26.62 25.48 23.49 0.7942 0.7887 0.7884 0.793 0.7977 0.8071 0.8192 0.835 0.8552 0.8884 0.9374 0.7942 0.7672 0.7531 0.7351 0.7191 0.704 0.6875 0.6638 0.6353 0.5991 0.4604 0 6.1707 12.7807 19.2219 26.1969 33.568 41.145 48.55 56.1684 64.0568 76.1755 Table 4.18 Density of CO2 as a Function of Pressure and Temperature CO2 Density (g/cm3), Pressure (psi) Temperature (°F) 362.594 725.189 1087.78 1450.38 2175.57 2900.75 3625.94 4351.13 68 86 104 122 140 158 176 194 212 230 248 266 284 302 320 0.0527 0.0499 0.0476 0.0456 0.0437 0.0421 0.0406 0.0391 0.0378 0.0366 0.0354 0.0344 0.0334 0.0325 0.0316 0.1423 0.1251 0.1135 0.1052 0.0984 0.093 0.0883 0.0845 0.081 0.0778 0.0749 0.0722 0.0697 0.0674 0.0653 0.81 0.655 0.2305 0.1932 0.1726 0.1584 0.1469 0.1381 0.1305 0.1239 0.1187 0.1141 0.1094 0.1054 0.1018 0.855 0.782 0.638 0.3901 0.2868 0.2478 0.2215 0.2019 0.1877 0.1765 0.1673 0.1595 0.1525 0.1461 0.1403 0.901 0.85 0.785 0.705 0.604 0.504 0.43 0.373 0.333 0.304 0.28 0.262 0.2465 0.2337 0.2229 0.9335 0.8887 0.8415 0.7855 0.724 0.6605 0.5935 0.5325 0.4815 0.4378 0.4015 0.3718 0.347 0.3267 0.3089 0.96 0.919 0.8771 0.8347 0.7889 0.7379 0.6872 0.6359 0.588 0.5443 0.5053 0.4718 0.4419 0.4151 0.3918 0.9832 0.946 0.9077 0.8687 0.8292 0.7882 0.7466 0.704 0.663 0.623 0.5855 0.5517 0.52 0.4925 0.468 430 CHAPTER 4 PVT Properties of Crude Oils n The listed Z-factor at any pressure in the CCE test data must satisfy the following relationship: Z¼ pVrel Zdew pdew For example, at 4211 psia Z-factor ¼ 0.8495 Relative volume “Vrel” ¼ 0.95593 Gas density ¼ 03986 g/cm3 Applying the above relation gives Z¼ n ð4211Þð0:95594Þ 0:79415 ¼ 0:8495 3763 an exact match with the reported value. Similarly, the gas density at any pressure in the CCE test data must satisfy the following relationship: ρ¼ ρsat Vrel Applying the above relation gives the exact reported value: ρ¼ n 0:381 ¼ 0:3986g=cm3 0:95594 Below the dew point pressure, the following expression provides the link that exists between relative volume and both reported LDO as given by ðLDOÞCCE Vt ¼ ðLDOÞCCE Vsat Vrel At 2884 psia, the CCE test data shows Vrel ¼ 1.19724 ðLDOÞVsat ¼ 30.55356 ðLDOÞVt ¼ 25.52 Applying the above expression to give the exact match: ðLDOÞCCE Vt ¼ n ðLDOÞCCE 30:55356 Vsat ¼ ¼ 25:52 Vrel 1:19724 The values of the two-phase Z-factor in the CVD test data can be checked for accuracy by comparing laboratory data against values as calculated from the following relationship: Ztwo-phase ¼ Zd p pd 1 ðRFÞ Liquid Blockage in Gas-Condensate Reservoirs 431 n FIGURE 4.53 Liquid dropout from CCE and CVD tests. For example, at 2513 psia: Reported Ztwo-phase ¼ 0.7191 Recovery factor ¼ 26.1969% Applying the above equation gives Ztwo-phase ¼ n Zd p 0:7942 2513 ¼ ¼ 0:7186 pd 1 ðRFÞ 3763 1 0:261969 A match with less than 1%. A plot of the LDO obtained from the CVD data and those of two LDOs from the CCE test, as shown in Figs. 4.53 and 4.54 should reveal the following consistency characteristics as observed by Lawrence and Gupta (2009): n The CVD LDO should all be between the two CCE LDOs. n The CVD maximum LDO should occur at lower pressure than the CCE maximum LDO. LIQUID BLOCKAGE IN GAS-CONDENSATE RESERVOIRS Well productivity is a key parameter in the development of gas-condensate reservoirs. However, accurate prediction of well productivity can be difficult because of the need to understand and account for the complex flow behavior occurring in both. 432 CHAPTER 4 PVT Properties of Crude Oils p* CVD Region 1 Region 2 Gas Gas oil oil Gas Gas Region 3 Gas Gas Gas oil oil 0.6 Pressure Oil saturation 0.8 CCE oil Flowing oil + gas Oil 0.4 sat Flowing gas ura tion pro Flowing gas pe Pressure profile pd file 0.2 SO>Soc & kro> 0 SO ≤ Soc & kro= 0 So= 0 & kro= 0 pwf 0 rw rdew Radius (ft) re n FIGURE 4.54 Condensate liquid blockage. n n The extended well flow lines from wellbore to surface facilities. The behavior of the retrograde gas system with declining reservoir pressure. A brief discussion of the above issues are presented next. Behavior of Gas-Condensate Wells In conventional dry or wet gas reservoirs, the calculation of the bottom-hole flowing pressure “pwf” as a function of the wellhead pressure and the pressure loss in the tubing is a relatively simple task — resulting in an accurate calculation of the gas rate as given by Darcy’s equation: Qg ¼ ðp kk h rg re 1422T ln + s 0:75 pwf rw or in a traditional form as ! 2p dp μg Z (4.136) Liquid Blockage in Gas-Condensate Reservoirs 433 Qg ¼ kkrg h½mðpe Þ mðpwf Þ re + s 0:75 1422T ln rw with mðpe Þ ¼ ð pe o ! 2p dp μg Z where Qg ¼ gas flow rate, Mscf/day T ¼ reservoir temperature, °R s ¼ skin factor k ¼ absolute permeability, md krg ¼ relative permeability h ¼ thickness, ft re ¼ drainage radius, ft rw ¼ wellbore radius, ft m( pe) ¼ real gas potential as evaluated from “0 to pe,” psi2/cp m(pwf) ¼ real gas potential as evaluated from “0 to pwf,” psi2/cp In retrograde gas reservoirs, however, wells could experience a significant decrease in the well productivity when the pressure drops below the dew point pressure in any segment of flow line extended from wellbore to surface facilities. The gas rate calculation requires an accurate estimation of the wellbore flowing pressure, which is impacted by the pressure loss in the tubing and the additional wellhead back pressure. Estimating the bottom-hole flowing pressure is considered a difficult task to perform for several reasons, including: n n n n n n n Multiphase flow could occur in the tubing based on the pressure gradient. Well temperature distribution in the tubing will impact the dew point pressure of the flowing fluid, which, in turn, will impact pwf. In general, the dew point pressure is a strong function of temperature; for example, for a fixed overall composition “zi,” the dew point will decrease with decreasing temperature. Effect of liquid loading on bottom-hole flowing pressure. Different flow regimes as the pressure drops below the dew point pressure. Flow regime transitions in the tubing. Physical phenomena, such as turbulence flow. Thermodynamic phase changes with pressure and temperature occurring as the condensate gas leaves the wellbore to surface separation facilities. 434 CHAPTER 4 PVT Properties of Crude Oils Many studies have investigated wellbore pressure drop in gas-condensate wells based on well pressure gradient condensate wells. The pressure gradient may be divided into three classes: 1. Wellhead pressure and bottom-hole flowing pressure are both above dew point, indicating a single gas phase flow in the tubing. The condensate yield and condensate-gas ratio will remain constant as long as the separation conditions of surface facilities remain unchanged. 2. Wellhead pressure is below the fluid dew point pressure at wellhead temperature while the bottom-hole flowing pressure is above dew point. In this case, the condensate yield and CGR will remain relatively constant. 3. Wellhead pressure as well as the bottom-hole flowing pressure are both below the fluid dew point pressure, indicating two-phase flow. Behavior of Retrograde Gas Reservoirs When the bottom-hole flowing pressure drops below the dewpoint, a region of high liquid saturation might build up around the wellbore, causing a “liquid blockage” that reduces the gas flow and well productivity. Ahmed (2000) studied the effectiveness of lean gas, N2, and CO2 injection in reducing the liquid blockage. The author suggested the use of the Huff-n-Puff injection to vaporize the LDO, providing that the reservoir pressure is above the dew point pressure. The gas cycling, on the other hand, is a common and viable approach when the bottom-hole flowing pressure and reservoir pressure are both below the dew point pressure. Results of Huff-n-Puff study indicated the importance of selecting the volume of the injected gas with the conclusion that the injected gas volume and success of the injection is impacted to a large degree with stage of reservoir depletion at the start of the injection. Fevang and Whitson (1994) presented a comprehensive study on the performance of retrograde gas reservoirs through the identification of the following three regions: Region 1 As most of the pressure drop is located within a short distance from the wellbore, it will create a large pressure sink around the wellbore. A substantial accumulation of the LDO could occur near the wellbore when the bottomhole flowing pressure drops below the dew point pressure. As shown schematically in Figs. 4.53 and 4.54, this inner near-wellbore zone is identified as Region 1 and characterized by the simultaneous flow of oil and gas. The richness of the flowing well stream and the PVT properties of the reservoir gas will determine the amount of liquid saturation in Region 1. It should be pointed out that the assumption of the two-phase flow could only occur if Liquid Blockage in Gas-Condensate Reservoirs 435 n n n The retrograde gas is classified as rich retrograde system. The pressure is below the dew point pressure. The LDO saturation is above the critical saturation; that is, So > Soc to allow for the liquid to flow with the gas. As documented by several reservoir simulation studies, most of the retrograde gas well deliverability loss is caused by the reduction of the gas relative permeability “krg” due to the condensate buildup in Region 1. The simultaneous flow of gas and oil in Region 1, presumably flowing in equilibrium, resembles the prevailing equilibrium state of the CCE test condition. Therefore, the data obtained from the CCE test can be manipulated and combined to generate the relative permeability ratio “krg/kro” as a function of pressure. As detailed in the earlier discussion of the CCE test, the following three measured pressure-volume relationships are reported mathematically by Eqs. (4.105), (4.128), (4.129) as Vrel ¼ Vg + VL Vt ¼ Vsat Vsat ðLDOÞCCE Vsat ¼ ðLDOÞCCE Vt ¼ VL Vsat VL VL ¼ Vt Vg + VL Combining the above expressions gives Vg 1 ¼ 1 VL ðLDOÞCCE V t Assuming the simultaneous flow if Region 1 is governed by the Darcy’s steady state flow, to give krg khΔp =½ug ð ln ðre =rw Þ krg uo Vg ¼ ¼ VL ðkro khΔpÞ=½uo ð ln ðre =rw Þ kro ug Combining the above two expressions, gives " # μg krg 1 ¼ 1 CCE kro μo ðLDOÞVt (4.137) It should be pointed out the relative permeability ratio “krg/kro” as given by the above relationship can only be generated as a function of pressure by using PVT software as the viscosity ratio “μg/μo” cannot be obtained from the CCE process. Region 2 As shown in Figs. 4.53 and 4.54, Region 2 is the intermediate zone where condensate dropout originates. Based on the pressure profile around the wellbore that is bounded between the flowing bottom-hole 436 CHAPTER 4 PVT Properties of Crude Oils pressure and reservoir pressure, the dew point exits at a distance “rdew” from the wellbore. The dew point pressure at the radius rdew marks the boundary between Regions 3 and 2. As the single-phase original reservoir gas from Region 3 enters Region 2, the first droplets of liquid condense at the boundary. In Region 2, contrary to Region 1, the saturation “So” of the net condensate accumulation is considered below the critical saturation and, therefore, immobile. Region 2 is then characterized by a single-phase flow with the gas as the only flowing phase while the condensate liquid remains immobile. Because the gas is considered the only moving “removing phase” from Region 2 to Region 1, the pressure-saturation relationship in Region 2 is similar to that of the CVD process. The saturation pressure can then be obtained directly from the CVD experiment data after correcting for the initial water saturation “Swi” as So ¼ ð1 Swi ÞLDO It should be noted that the overall composition “zi” of the fluid in Region 1 is equivalent to that of the flowing gas from Region 2; in addition, the dew point pressure of the total (combined) fluid in Region 1 is identical to that of flowing gas from Region 2 as represented by p*. Region 3 This zone consists of single-phase gas, as the pressure in this zone is above the dewpoint pressure, with no hydrocarbon liquid in the region. The single-phase gas rate can be calculated accurately by from Eq. (4.136); that is, Qg ¼ ð pe kk h rg re + s 0:75 pwf 1422T ln rw ! 2p dp μg Z However, the above expression must be modified to account for the impact of the above three regions on the gas rate. Eq. (4.136) can be modified to give the gas rate as # ð pe " krg kh kro dp Rs + Qg ¼ re μo Bo μg Bgd 0:75 + s pwf 1422 ln rw (4.138) or # ð pe " krg kro dp Rs + Qg ¼ C μo Bo μg BE pwf (4.139) Liquid Blockage in Gas-Condensate Reservoirs 437 where the performance coefficient C is defined by C¼ kh re 0:75 + s 1422 ln rw where Qg ¼ gas flow rate, Mscf/day k ¼ absolute permeability, md h ¼ thickness, ft re ¼ drainage radius rw ¼ well base radius pe ¼ reservoir pressure at the external well drainage radius, psi p* ¼ dew point pressure of the gas entering Region 1 from Region 2 pwf ¼ bottom-hole flow pressure BgD ¼ dry gas FVF, bbl/scf Bo ¼ oil FVF of the condensate liquid, bbl/STB s ¼ skin factor The integral part of Eq. (4.139) that represents the pressure drop from pe to pwf is divided into three portions to describe the individual pressure drop in each region as ð pe pwf ! krg kro + Rs dp ¼ ðΔpÞRegion1 + ðΔpÞRegion2 + ðΔpÞRegion3 Bgd μg Bo μo where ðΔpÞRegion1 ¼ ðp pwf ðΔpÞRegion2 ¼ ðΔpÞRegion3 ¼ ! krg kro + Rs dp Bgd μg Bgd μg ð pd p ð pe pd ! krg dp Bgd μg ! 1 dp Bgd μg Fevang and Whitson (1994) presented a comprehensive treatment and detailed discussion of the methodology of implementing the above treatment of the three regions in reservoir simulation. Special Laboratory PVT Tests In addition to the previously described routine laboratory tests, a number of other projects may be performed for very specific applications. If a 438 CHAPTER 4 PVT Properties of Crude Oils reservoir is to be developed under gas injection, miscible gas injection, or a dry gas cycling scheme, a number of specialized tests must be conducted, including n n n n n the swelling test slim-tube test rising bubble test core flood other tests. The above tests are designed to provide the essential “optimized” key data that can be used when planning reservoir developing scenarios with miscible gas displacement. The displacement efficiency of reservoir oil by gas injection is highly pressure dependent and miscible displacement is only achieved at pressures greater than a certain minimum pressure, termed the minimum miscibility pressure “MMP.” The minimum miscibility pressure is defined as the “lowest pressure for which a given injected gas composition can develop miscibility through a multicontact process with a given reservoir oil at reservoir temperature.” The reservoir to which the process is applied must be operated at or above the MMP to develop miscibility. Reservoir pressures below the MMP result in immiscible displacements and consequently lower oil recoveries. It should be pointed out that the minimum miscibility pressure is completely independent of the reservoir heterogeneity; however, the MMP is a strong function of n n n Oil composition Composition of the injected gas Reservoir temperature Swelling Test As shown schematically in Fig. 4.55, a gas with a known composition (usually similar to proposed injection gas) is added to the original reservoir oil at varying proportions in a series of steps. After each addition of a certain measured volume of gas, the overall mixture is quantified in terms of the molar percentage of the injection gas, such as 10 mol% of N2. The PVT cell then is pressured to the saturation pressure of the new mixture (only one phase is present). The gas addition starts at the saturation pressure of the reservoir fluid (bubble point pressure if an oil sample or dew point pressure if a gas sample) and the addition of the gas continues to perhaps 80 mol% injected gas in the fluid sample. It should be pointed out that the saturation pressure may change from bubble point to dewpoint after significant Liquid Blockage in Gas-Condensate Reservoirs 439 n FIGURE 4.55 Schematic diagram of the swelling test. volumes of gas injection. As shown in Fig. 4.56, results from the test are expressed in terms of n n saturation pressure as a function of volume of gas injected per barrel of the original fluid (ie, scf/bbl). The swollen volume as a function of injected gas expressed in scf/bbl. The swollen volume is expressed as the ratio of volume of the saturated fluid mixture to the volume of the original saturated reservoir oil. The data collected from the swelling test is essential as it reveals the ability of the injected gas to: n n n n dissolve in the reservoir fluid swell the hydrocarbon system vaporize the hydrocarbon form multiphases in case of CO2 injection. Results from the swelling test must be used to tune EOS parameters for compositional modeling. 440 CHAPTER 4 PVT Properties of Crude Oils 2.00 swelling ratio (Vsat)new/(Vsat)orig psat 1.50 Original psat 1.00 0 500 1000 1500 2000 scf of injected gas / bbl of original fluid 2500 3000 n FIGURE 4.56 Saturation pressure and swelling ration as a function of volume of gas injected. Slim Tube Gas injection is considered to be one of the most effective EOR methods, particularly in developing light oil reservoirs. Depending on the injection pressure, there are three classifications of gas injection: n n n Miscible gas injection Partial miscible gas injection Immiscible gas injection The key parameter in the above classifications is based on the minimum miscibility pressure “MMP,” which is considered the most important parameter for assessing the applicability of any miscible gas flood for an oil reservoir. The classical thermodynamics definition of miscibility is the condition of pressure and temperature at which two fluids, when mixed in any proportion, form a single phase. For example, kerosene and crude oil are miscible at room conditions, whereas stock-tank and water are clearly immiscible. At miscible condition, the displacement efficiency exceeds 90%; however, if the flow is one-dimensional with no dispersive mixing, the displacement efficiency could reach 100%. A slim-tube miscibility test is performed to determine the MMP by using the equipment setup as shown schematically in Fig. 4.57. Typically the test is performed in a 40-ft coiled stainless steel tubing prepacked with glass beads or sand and is filled with oil at constant Liquid Blockage in Gas-Condensate Reservoirs 441 Coiled stainless steel tubing prepacked with glass beads Oil Inj. gas Sight glass Back pressure regulator Pump Gas meter Gas chromatograph n FIGURE 4.57 Schematic diagram of slim-tube apparatus. temperature. The coiled tube is characterized by a very high porosity and permeability. Unlike the previously described laboratory tests that concentrate specifically on fluid property variations with pressure and composition, the slim-tube test is specially designed to examine the flushing efficiency and fluid mixing during a miscible displacement process for a given gas composition. The slim-tube equipment and test procedure are designed to eliminate the effect of the following parameters: n n n Reservoir heterogeneities Impact of reservoir water Gravity effect The displacement of the oil from the slim tube is normally repeated four to six times, each at sequentially increasing pressure. The test is terminated after the injection of 1.2 pore volume of gas injected at each pressure. The fraction of the initial oil-in-place (ie, in the stainless steel coil) recovered after each injection is recorded and plotted as a function of pressure as shown in Fig. 4.58. Typically, pressure-recovery increases rapidly with increasing pressure and then starts to flatten and eventually forms a straight line starting at certain pressure that is defined as minimum miscibility pressure. It should be pointed out that different recovery levels, such as 80% at the gas breakthrough, or 90–95% ultimate recovery have also been suggested as the criteria for miscible displacement. The oil recovery, however, depends on the tube design and operating conditions. A slim tube may deliver only 80% oil recovery at miscible conditions. The evaluation of recovery changes Separator 442 CHAPTER 4 PVT Properties of Crude Oils 3 2 Lean Gas Injection 1 Recovery factor RF after 1.2 PV injected Test # 1 2 3 4 5 6 p 4000 4400 4600 4700 4900 5200 RF, % 76 85 93 96 99 99 Cumulative gas injected, PV Recovery factor MMP Pressure n FIGURE 4.58 Determination of MMP from the slim-tube oil recovery data. with displacement pressure, or gas enrichment, is more appropriate to determine miscibility conditions than searching for a high recovery. The most acceptable definition is the pressure, or enrichment level, at the break-over point of the ultimate oil recovery, as shown in Fig. 4.58. The recovery is expected to increase by increasing the displacement pressure, but the additional recovery above MMP is generally minimal. The effluent liquid is usually monitored and flashed to atmospheric conditions and produced gases are passed through GC for compositional analysis. It should be pointed out that static tests, where the injection gas and the reservoir oil are equilibrated in a cell, also are conducted to determine the mixture phase behavior. Although these tests do not closely simulate the dynamic reservoir conditions, they provide accurate, well-controlled data that are quite valuable, particularly for tuning the EOS used in modeling the process. Liquid Blockage in Gas-Condensate Reservoirs 443 It should be noted that the pressure-recovery factor curve shows a pressure that is lower than the MMP and designated as the vaporization pressure “pvap.” This is the pressure at which the recovery factor curve starts to rise with increasing pressure until it reaches the MMP. For all pressures below the vaporization pressure, the displacement is classified as immiscible. The following expression, called the miscibility function “α,” can be used to define and classify the type of gas injection: α¼ p pvap MMP pvap where “p” is the injected gas pressure in the reservoir. The above function suggests if (1) p pvap; indicates immiscible displacement and miscibility function is set to zero, α ¼ 0. (2) pvap p MMP; indicates partial miscible displacement with 0 < α < 1. (3) p MMP; indicates miscible displacement with α ¼ 1. The miscibility function is routinely used in black oil simulator, as an alternative to compositional simulator, to simulate miscible displacement by gas injection. The miscibility function is basically used to adjust the viscosity of the displacing gas and the displaced oil to account for gas solubility in oil or vaporization of oil components to the gas phase. In addition, α is used to adjust relative permeability curves based on type of displacement. Viscosity Adjustment The viscosity adjustment procedure is based on the following fluid mixing and property characteristics: n Mobility ratio: The mobility ratio “M,” which is defined as the ratio of the oil viscosity “μo” to that of injected gas or the solvent “μs”; that is, M¼ n μo μs Viscosity mixing parameter: This parameter is represented by ω and given by the following expression: ω¼ 1 h i 4 log 0:78 + 0:22ðMÞ1=4 log ðMÞ It should be noted that for complete miscible displacement, M ¼ 1 and ω ¼ 1. 444 CHAPTER 4 PVT Properties of Crude Oils n Mixture viscosity: The mixture is correlated with saturation of all phased and viscosity of oil and solvent (gas) by the following expression: μm ¼ 1 Sw So s + μS0:25 0:25 μ o n !4 s Mixed oil viscosity: The mixed oil viscosity is given by μom ¼ μ1ω μωm o n Mixed solvent (gas) viscosity: The mixed oil viscosity is given by ω μSm ¼ μ1ω S μm The above defined mixing parameters are used to the calculate the effective oil and solvent viscosity by applying the following relationships: μse ¼ ð1 αÞμs + αμsm μoe ¼ ð1 αÞμo + αμom Relative Permeability Adjustment The relative permeability data is obtained from a laboratory test conducted on a reservoir core saturated with crude oil and initial connate water. The crude is displaced from the core by immiscible gas. In reservoir simulation using a black oil simulator, the laboratory-generated relative permeability must be adjusted to account for the interaction and mixing that are taking place between the injected gas and oil. The adjustment is based calculating the mixture relative permeability parameter “krm” as given by Sg So Sorm krow + krg krm ¼ 1 Sw Sorm 1 Sw Sorm where Sorm is the residual oil saturation to miscible displacement, usually set between 5% and 10%. The effective (adjusted) relative permeability of the gas and oil are then given by So Sorm krm 1 Sw Sorm Sg krm krg eff ¼ ð1 αÞkrg + 1 Sw Sorm ðkro Þeff ¼ ð1 αÞkro + α Notice that no adjustment is needed if α ¼ 0. Liquid Blockage in Gas-Condensate Reservoirs 445 Rising Bubble Experiment Rising bubble apparatus is employed in another experimental technique, which is commonly used for quick and reasonable estimates of gas-oil miscibility. In this method, the miscibility is determined from the visual observations of changes in the shape and appearance of bubbles of injected gas as they rise through a visual high-pressure cell filled with the reservoir crude oil. A series of tests are conducted at different pressures or enrichment levels of the injected gas, and the bubble shape is continuously monitored to determine miscibility. This test is qualitative in nature, as miscibility is inferred from visual observations. Hence, some subjectivity is associated with the miscibility interpretation of this technique. Therefore, the results obtained from this test are somewhat arbitrary; however, this test is quite rapid and requires less than 2 h to determine miscibility. This method also is cheaper and requires smaller quantities of fluids compared to the slim tube. The subjective interpretations of miscibility from visual observations, lack of quantitative information to support the results, and some arbitrariness associated with miscibility interpretation are some of the disadvantages of this technique. Also, no strong theoretical background appears to be associated with this technique and it provides only reasonable estimates of gas-oil miscibility conditions. Minimum Miscibility Pressure Correlations The displacement efficiency of oil by gas depends highly on pressure, and miscible displacement is achieved only at pressures greater than a certain minimum, the minimum miscibility pressure. Slim-tube displacement tests are commonly used to determine the MMP for a given crude oil. The minimum miscibility pressure is defined as the pressure at which the oil recovery versus pressure curve (as generated from the slim-tube test) shows a sharp change in slope, the inflection point. There are several factors that affect MMP, including n n n n n n Reservoir temperature. Oil characteristics and properties, including the API gravity. Injected gas composition. Concentration of C1 and N2 in the crude oil. Oil molecular weight. Concentration of intermediate components (C2–C5) in the oil phase. There exists a distinct possibility between reservoir temperature and MMP because MMP increases as the reservoir temperature increases. The MMP also increases with high-molecular-weight oil and oils containing higher 446 CHAPTER 4 PVT Properties of Crude Oils concentrations of C1 and N2. However, the MMP decreases with increasing the percentage of the intermediates in the oil phase. To facilitate screening procedures and gain insight into the miscible displacement process, many correlations relating the MMP to the physical properties of the oil and the displacing gas have been proposed. It should be noted that, ideally, any MMP correlation should n n n Account for each parameter known to affect the MMP. Be based on thermodynamic or physical principles that affect the miscibility of fluids. Be directly related to the multiple-contact miscibility process. A variety of correlations or estimations of the MMP developed from regressions of slim-tube data. Although less accurate, these correlations are quick and easy to use and generally require only a few input parameters. Hence, they are very useful for a fast screening of a reservoir for various types of gas injection. They also are useful when detailed fluid characterizations are not available. One significant disadvantage of current MMP correlations is that the regressions use MMP from slim-tube data, which themselves are uncertain. Some MMP correlations require only the input of reservoir temperature and the API gravity of the reservoir fluid. Other, more accurate, correlations require reservoir temperature and the total C2–C6 content of the reservoir fluid. In nearly all the correlations, the methane content of the oil is assumed to not affect the MMP significantly. In general, the MMP correlations fall into the following two categories: n n Those dedicated to pure and impure CO2. Those that examine the MMP of other gases. The following section briefly reviews the correlations associated with each of these categories. Pure and Impure CO2 MMP It is well documented that the development of miscibility in a CO2 crude oil displacement is the result of extraction of some hydrocarbons from the oil by dense CO2. Many authors found considerable evidence that the extraction of hydrocarbons from a crude oil is strongly influenced by the density of CO2. Improvement of extraction with the increase in CO2 density that accompanies increasing pressure accounts for the development of miscibility. The presence of impurities can affect the pressure required to achieve Liquid Blockage in Gas-Condensate Reservoirs 447 miscible displacement. Some of the widely used correlations are outlined next, including the following: n n n n n n n Orr and Silva correlation Extrapolated vapor pressure (EVP) method Yellig and Metcalfe correlation Alston’s correlation National Petroleum Council (NPC) method Cronquist’s correlation Yuan’s correlation Orr and Silva Correlation. Orr and Silva (1987) developed a methodology for determining the MMP for pure and contaminated CO2/crude oil systems. Orr and Silva pointed out that the distribution of molecular sizes present in a crude oil has a significantly larger impact on the MMP than variations in hydrocarbon structure. The carbon-number distributions of the crude oil system are the only data needed to use the correlation. The authors introduced a weighted-composition parameter, F, based on a normalized partition coefficient, Ki, for the components C2 through C37 fractions. The specific steps of the method are summarized next. (1) From the chromatographic or the simulated compositional distribution of the crude oil, omit C1 and all the nonhydrocarbon components (ie, CO2, N2, and H2S) from the oil composition. Normalize the weight fractions (wi) of the remaining components (C2–C37). (2) Calculate the partition coefficient Ki for each component in the remaining oil fractions; that is, C2–C37, using the following equation: log ðKi Þ ¼ 0:761 0:04175Ci where Ci is the number of carbon atoms of component i. (3) Evaluate the weighted-composition parameter, F, from F¼ 37 X Ki wi 2 (4) Calculate the density of CO2 required to achieve miscibility from the following expressions: ρmmp ¼ 1:189 0:542F, for F < 1:467 ρmmp ¼ 0:42, for F > 1:467 (5) Given the reservoir temperature, find the pressure, which is assigned to the MMP, at which the CO2 density is equal to the required ρmmp. Kennedy (1954) presented the relationship of the density of CO2 as a 448 CHAPTER 4 PVT Properties of Crude Oils function of pressure and temperature in a tabulated as shown in Table 4.18. The tabulated values of CO2 density as a function of pressure and temperature can be used to estimate the MMP by entering the table with the calculated value of ρmmp and reservoir temperature to find the pressure that corresponds to the MMP. EVP Method. Orr and Jensen (1986) suggested that the vapor pressure curve of CO2 can be extrapolated and equated with the minimum miscibility pressure to estimate the MMP for low-temperature reservoirs (T < 120°F). A convenient equation for the vapor pressure has been given by Newitt et al. (1996): EVP ¼ 14:7exp 10:91 2015 255:372 + 0:5556ðT 460Þ with the EVP in psia and the temperature, T, in °R. EXAMPLE 4.46 Estimate the MMP for pure CO2 at 610°R using the EVP method. Solution Estimate the MMP for pure CO2 at 610°R using the EVP method: 2015 EVP ¼ 14:7exp 10:91 2098 psi 255:372 + 0:5556ð610 460Þ Researchers in the Petroleum Recovery Institute suggest equating the MMP with the vapor pressure of CO2 when the system temperature is below the critical temperature, Tc, of CO2. For temperatures greater than Tc, they proposed the following expression: MMP ¼ 1071:82893 10b with the coefficient b as defined by b ¼ ½2:772 ð1519=T Þ where MMP is in psia and T in °R. EXAMPLE 4.47 Estimate the MMP for pure CO2 at 610°R using the method proposed by the Petroleum Recovery Institute. Liquid Blockage in Gas-Condensate Reservoirs 449 Solution Step 1. Calculate the coefficient b: b ¼ ½2:772 ð1519=T Þ b ¼ ½2:772 ð1519=610Þ ¼ 0:28184 Step 2. Calculate MMP from MMP ¼ 1071:82893 10b MMP ¼ 1071:82893 100:28184 ¼ 2051 psi Yellig and Metcalfe Correlation. From their experimental study, Yellig and Metcalfe (1980) proposed a correlation for predicating the CO2 MMPs that use the temperature, T, in °R, as the only correlating parameter. The proposed expression follows: MMP ¼ 1833:7217 + 2:2518055ðT 460Þ + 0:01800674ðT 460Þ2 103949:93 T 460 Yellig and Metcalfe pointed out that, if the bubble point pressure of the oil is greater than the predicted MMP, then the CO2 MMP is set equal to the bubble point pressure. EXAMPLE 4.48 Estimate the MMP for pure CO2 at 610°R using the Yellig and Metcalfe method. Solution MMP ¼ 1833:7217 + 2:2518055ðT 460Þ + 0:01800674ðT 460Þ2 103949:93 T 460 MMP ¼ 1833:7217 + 2:2518055ð610 460Þ + 0:01800674ð610 460Þ2 103949:93 ¼ 1884psi T 460 Alston’s Correlation. Alston et al. (1985) developed an empirically derived correlation for estimating the MMPs for pure or impure CO2/oil systems. Alston and coworkers used the temperature, oil C5+ molecular weight, volatile oil fraction, intermediate oil fraction, and the composition of the CO2 stream as the correlating parameters. The MMP for pure CO2/oil systems is given by MMP ¼ 0:000878ðT 460Þ1:06 ðMC5 + Þ1:78 Xvol 0:136 Xint 450 CHAPTER 4 PVT Properties of Crude Oils where T ¼ system temperature, °R MC5 + ¼ molecular weight of pentane and heavier fractions in the oil phase Xint ¼ mole fraction of intermediate oil components (C2–C4, CO2, and H2S) Xvol ¼ mole fraction of the volatile (C1 and N2) oil components Contamination of CO2 with N2 or C1 has been shown to adversely affect the minimum miscibility pressure of carbon dioxide. Conversely, the addition of C2, C3, C4, or H2S to CO2 has been shown to lower the MMP. To account for the effects of the presence of contaminants in the injected CO2, the authors correlated impure CO2 MMP with the weighted average pseudocritical temperature, Tpc, of the injected gas and the pure CO2 MMP by the following expression: 168:893 87:8 Tpc 460 MMPimp ¼ MMP Tpc The pseudocritical temperature of the injected gas is given by Tpc ¼ X wi Tci 460 where MMP ¼ minimum miscibility pressure of pure CO2 MMPimp ¼ minimum miscibility pressure of the contaminated CO2 wi ¼ weight fraction of component i in the injected gas Tci ¼ critical temperatures of component i in the injected gas, in °R T ¼ system temperature, in °R It should be pointed out that the authors assigned a uniform critical temperature value of 585°R for H2S and C2 in the injected gas. Sebastian’s Method. Sebastian et al. (1985) proposed a similar corrective step that adjusts the pure CO2 MMP by an amount related to the mole average critical temperature, Tcm, of the injected gas by MMPimp ¼ ½CMMP where the correction parameter, C, is given by C ¼ 1:0 A 0:0213 0:000251A 2:35 107 A2 with A ¼ ½Tcm 87:89=1:8 X yi ðTci 460Þ Tcm ¼ Liquid Blockage in Gas-Condensate Reservoirs 451 where MMP ¼ minimum miscibility pressure of pure CO2 MMPimp ¼ minimum miscibility pressure of the contaminated CO2 yi ¼ mole fraction of component i in the injected gas Tci ¼ critical temperature of component i in the injected gas, °R To give a better fit to their data, the authors adjusted Tc of H2S from 212 to 125°F. NPC Method. The NPC proposed an empirical correlation that provides rough estimates of the pure CO2 MMPs. The correlation uses the API gravity and the temperature as the correlating parameters, as follows: Gravity (API) MMP (psi) <27° 27–30° >30° 4000 3000 1200 With reservoir temperature correction as given below: T (°F) Additional Pressure (psi) <120 120–150 150–200 200–250 0 200 350 500 Cronquist’s Correlation. Cronquist (1978) proposed an empirical equation that was generated from a regression fit on 58 data points. Cronquist characterizes the miscibility pressure as a function of T, molecular weight of the oil pentanes-plus fraction, and the mole percentage of methane and nitrogen. The correlation has the following form: MMP ¼ 15:988ðT 460ÞA with A ¼ 0:744206 + 0:0011038MC5 + + 0:0015279yc1 N2 where T ¼ reservoir temperature, °R, and yc1 N2 ¼ mole percentage of methane and nitrogen in the injected gas. Yuan’s Correlation. Yuan et al. (2005) developed a correlation for predicting minimum miscibility pressure for pure CO2 based on matching 41 experimental slim-tube MMP values. The authors correlated the MMP with the molecular weight of the heptanes-plus fraction, MC7 + ; temperature T, in °R; and mole percent of the intermediate components, CM; that is, C2–C6. 452 CHAPTER 4 PVT Properties of Crude Oils The correlation has the following form: " ðMMPÞpureCO2 ¼ a1 + a2 MC7 + + a3 CM + a4 + a5 MC7 + + a6 CM ðM C 7 + Þ2 # ðT 460Þ + a7 + a8 MC7 + + a9 ðMC7 + Þ2 + a10 CM ðT 460Þ2 with the coefficients as follows: a1 ¼ 1463.4 a2 ¼ 6.612 a3 ¼ 44.979 a4 ¼ 21.39 a5 ¼ 0.11667 a6 ¼ 8166.1 a7 ¼ 0.12258 a8 ¼ 0.0012283 a9 ¼ 4.052(106) a10 ¼ 9.2577(104) The authors proposed an additional correlation to account for the contamination of CO2 with methane. The correlation is valid for methane contents in the gas only up to 40%. The proposed expression takes the form ðMMPÞimpure ðMMPÞpureCO2 ¼ 1 + mðyCO2 100Þ with " a6 CM # ðT 460Þ ðMC7 + Þ2 h i + a7 + a8 MC7 + + a9 ðMC7 + Þ2 + a10 CM ðT 460Þ2 m ¼ a1 + a2 MC7 + + a3 CM + a4 + a5 MC7 + + with the coefficients as follows: a1 ¼ 0.065996 a2 ¼ 1.524(104) a3 ¼ 0.0013807 a4 ¼ 6.2384(104) a5 ¼ 6.7725(107) a6 ¼ 0.027344 a7 ¼ 2.6953(106) a8 ¼ 1.7279(108) a9 ¼ 43.1436 (1011) a10 ¼ 1.9566 (108) Lean Gas and Nitrogen Miscibility Correlations High-pressure lean gases and N2 injection have been successfully used as displacing fluids for EOR projects and also are widely used in gas cycling and pressure maintenance. In general, miscibility pressure with lean gas Liquid Blockage in Gas-Condensate Reservoirs 453 (eg, methane) increases with increasing temperature because the solubility of methane in hydrocarbons decreases with increasing temperature, which in turn causes the size of the two-phase region to increase. However, miscibility pressure with nitrogen decreases with increasing temperatures because the solubility of nitrogen in hydrocarbons increases. Firoozabadi and Aziz documented the successful use of lean gas/N2 as a high-pressure miscible gas injection in several oil fields. The are several correlations that can be used to estimate the MMP with lean gas, including the following: n n n n n Firoozabadi and Aziz’s correlation Hudgins’s correlation Glaso’s correlation Sebastian and Lawrence’s correlation Lange’s correlation Firoozabadi and Aziz’s Correlation Firoozabadi and Aziz (1986) proposed a generalized correlation that can predict MMP for N2 and lean gases. They used the concentration of oil intermediate components, temperature, and the molecular weight of C7+ as the correlating parameters. The authors defined the intermediate components as the sum of the mole % of C2–C5, CO2, and H2S in the crude oil. The correlation has the following form: MMP ¼ 9433 188 103 F + 1430 103 F2 with F¼ I MC7 + ðT 460Þ0:25 I xC2 C5 + xCO2 + xH2 S where I ¼ concentration of intermediates in the oil phase, mol% T ¼ temperature, °R MC7 + ¼ molecular weight of C7+ Conrard (1987) points out that most nitrogen and lean gas miscibility correlations overestimate the MMP as compared with experimental MMP. The author attributed this departure to the difference between MMP and saturation pressure and suggested the following expression to improve the predictive capability of the Firoozabadi and Aziz correlation: MMP ¼ 0:6909ðMMPÞFA + 0:3091pb 454 CHAPTER 4 PVT Properties of Crude Oils where MMP ¼ corrected minimum miscibility pressure, psi MMPF–A ¼ Firoozabadi and Aziz MMP, psi pb ¼ bubble point pressure, psi A very simplified expression that uses the Firoozabadi and Aziz form is given below for estimating MMP for a nitrogen-crude oil system: 1:8I MMP ¼ 75:652 TMC7 + 0:5236 Hudgins’s Correlation Hudgins et al. (1990) performed a comprehensive laboratory study of N2 miscible flooding for enhanced recovery of light crude oil. They stated that the reservoir fluid composition, especially the amounts of C1 through C5 fractions in the oil phase, is the major determining factor for miscibility. For pure N2 the authors proposed the following expression: MMP ¼ 5568 eR1 + 3641 eR2 with R1 ¼ 792:06xC2 C5 MC7 + T 0:25 2:158 106 ðC1 Þ5:632 R2 ¼ MC7 + T 0:25 where T ¼ temperature, °F C1 ¼ mole fraction of methane xC2 C5 ¼ sum of the mole fraction of C2–C5 in the oil phase Glaso’s Correlation Glaso (1990) investigated the effect of reservoir fluid composition, displacement velocity, column length of the slim tube, and temperature on slim-tube oil recovery with N2. Glaso used the following correlating parameters in developing his MMP relationship: n n n n Molecular weight of C7+; that is, MC7 + . Temperature T, in °R. Sum of the mole % intermediates (C2–C6) in the oil phase; that is, xC2 C5 . Mole % of methane in the oil phase; that is, xC1 . Liquid Blockage in Gas-Condensate Reservoirs 455 Glaso proposed the following relationships: For API < 40: MMP ¼ 80:14 + 35:35H + 0:76H 2 where H¼ MC0:88 ðT 460Þ0:11 7+ ðxC2 C6 Þ0:64 ðxC1 Þ0:33 For API > 40: MMP ¼ 648:5 + 2619:5H + 1347:6H2 where H¼ MC0:88 ðT 460Þ0:25 7+ ðxC2 C6 Þ0:12 ðxC1 Þ0:42 Sebastian and Lawrence’s Correlation Sebastian and Lawrence (1992) developed a correlation for predicting the MMP for nitrogen, N2. The authors used the following variables to develop their correlation: n n n n Molecular weight of C7+ in the oil phase, MC7 + . Mole fraction of methane in the oil phase, xC1 . Mole fraction of intermediate components (C2–C6 and CO2) in the oil phase, xm. Reservoir temperature, T, in °R. The proposed correlation has the following form: $ % 3283xC1 T 4:776ðxC1 Þ2 T 2 + 4:008xm T 2 MMPN2 ¼ 4603 MC7 + + 2:05MC7 + + 7:541T Lange’s Correlation Lange (1998) developed a generalized MMP correlation that can be used for a wide range of injected gases, crude oils, temperatures, and pressures. The correlation is based on representation of the physical and chemical properties of crude oil and injected gas through the well-known Hildebrand solubility parameters. The solubility parameter concept is widely used to judge the “goodness” of a solvent for a solute, such that a small difference in solubility 456 CHAPTER 4 PVT Properties of Crude Oils parameters between a solute and solvent suggests that the solute may dissolve in the solvent. The solubility parameter of a high-pressure gas depends primarily on its reduced density, ρr, which can be estimated with an equationof-state calculation and is expressed by the following relationship: pffiffiffiffiffiffi δgi ¼ 0:122ρri pci where ρri ¼ reduced density of component i in the injected gas, lb/ft3 pci ¼ critical pressure of component i in the injected gas, psia δgi ¼ solubility parameter of component i in the injected gas, (cal/cm3)0.5 The reduced density as defined previously is given by ρr ¼ ρi =ρci or δgi ¼ 0:122 ρi pffiffiffiffiffiffi pci ρci The solubility parameter of the injected gas, δgas, can be calculated by using the following mixing rules, based on an individual component’s solubility, δgi, and its volume fraction in the injected gas: δgas ¼ N X υi δgi (4.140) i¼1 where δgas ¼ solubility parameter of the injected gas (cal/cm3)0.5 δgi ¼ solubility parameter of component i in the injected gas (cal/cm3)0.5 vi ¼ volume fraction of component i in the injected gas, psia pci ¼ critical pressure of component i in the injected gas, psia ρi ¼ density of component i in the injected gas at injection pressure and temperature, lb/ft3 ρci ¼ critical density of component i in the injected gas, lb/ft3 The solubility of a crude oil depends primarily on its molecular weight and can be approximated by the following expression: δoil ¼ ð6:97Þ0:01M 0:00556ðT 460Þ (4.141) where M ¼ molecular weight of the reservoir oil and T ¼ reservoir temperature, °R. Lange points out that miscibility occurs when the solubility difference between the oil injected gas is around 3(cal/cm3)0.5, or Miscibilitycriterion : jδoil δgas j 3 0:4 Problems 457 The process of determining the MMP using Lange’s correlation is essentially a trial-and-error approach. A pressure is assumed and δgas is calculated; if the miscibility criterion, as just defined, is met, then the assumed pressure is the MMP; otherwise the process is repeated. Lange also developed an expression for estimating the residual oil saturation to miscible displacement, Sorm, which is very useful as a screening parameter. The correlation is based on data from a large pool of core floods with the following form: Sorm ¼ Sorw 0:12 δoil δgas 0:11 where Sorw is residual oil saturation to water flood. PROBLEMS 1. Tables 4.19–4.21 show the experimental results performed on a crude oil sample taken from the MTech field. The results include the CCE, DE, and separator tests. a. Select the optimum separator conditions and generate Bo, Rs, and Bt values for the crude oil system. Plot your results and compare to the unadjusted values. b. Assume that new field indicates that the bubble point pressure is better described by a value of 2500 psi. Adjust the PVT to reflect for the new bubble point pressure. 2. A crude oil system exists at its bubble point pressure of 1708.7 psia and a temperature of 131°F. Given the following data: API ¼ 40° Average specific gravity of separator gas ¼ 0.85 Separator pressure ¼ 100 psig a. Calculate Rsb using 1. Standing’s correlation. 2. Vasquez and Beggs’s method. 3. Glaso’s correlation. 4. Marhoun’s equation. 5. Petrosky and Farshad’s correlation. b. Calculate Bob by applying methods listed in part a. 3. Estimate the bubble point pressure of a crude oil system with the following limited PVT data: API ¼ 35° T ¼ 160°F Rsb ¼ 700 scf/STB γ g ¼ 0.75 Use the five different methods listed in problem 2, part a. 458 CHAPTER 4 PVT Properties of Crude Oils Table 4.19 Pressure-Volume Relations of a Reservoir Fluid at 260°F (Constant Composition Expansion) Pressure (psig) Relative Volume 5000 4500 4000 3500 3000 2500 2300 2200 2100 2051 2047 2041 2024 2002 1933 1843 1742 1612 1467 1297 1102 863 653 482 0.946 0.953 0.9607 0.9691 0.9785 0.989 0.9938 0.9962 0.9987 1 1.001 1.0025 1.0069 1.0127 1.032 1.0602 1.0966 1.1524 1.2299 1.3431 1.5325 1.8992 2.4711 3.405 Table 4.20 Differential Vaporization at 260° p Rsd Btd Bod 2051 1900 1700 1500 1300 1100 900 700 500 300 170 0 1004 930 838 757 678 601 529 456 379 291 223 0 1.808 1.764 1.708 1.66 1.612 1.566 1.521 1.476 1.424 1.362 1.309 1.11 1.808 1.887 2.017 2.185 2.413 2.743 3.229 4.029 5.537 9.214 16.246 ρ (g/cm3) 0.6063 0.6165 0.6253 0.6348 0.644 0.6536 0.6635 0.6755 0.6896 0.702 0.7298 Z Bg (bbl/scf) γg 0.5989 0.88 0.884 0.887 0.892 0.899 0.906 0.917 0.933 0.955 0.974 0.00937 0.01052 0.01194 0.01384 0.01644 0.02019 0.02616 0.03695 0.06183 0.10738 0.843 0.84 0.844 0.857 0.876 0.901 0.948 0.018 1.373 2.23 Problems 459 Table 4.21 Separator Tests of Reservoir Fluid Sample Separator Pressure (psig) Separator Temperature Rs (scf/bbl) Rs (scf/STB) 200–0 71 71 72 72 71 71 70 70 431a 222 522 126 607 54 669 25 490b 223 566 127 632 54 682 25 100–0 50–0 25–0 API Bo 48.2 1.549 48.6 1.529 48.6 1.532 48.4 1.558 a c Sep Vol Factor Gas Sp.gr. 1.138d 1.006 1.083 1.006 1.041 1.006 1.02 1.006 0.739e 1.367 0.801e 1.402 0.869e 1.398 0.923e 1.34 Gas-oil ratio in cubic feet of gas at 60°F and 14.75 psi absolute per barrel of oil at indicated pressure and temperature. Gas-oil ratio in cubic feet of gas at 60°F and 14.75 psi absolute per barrel of stock-tank oil at 60°F. Formation volume factor in barrels of saturated oil at 2051 psi gauge and 260°F per barrel of stock-tank oil at 60°F. d Separator volume factor in barrels of oil at indicated pressure and temperature per barrel of stock-tank oil at 60°F. e Collected and analyzed in the laboratory. b c 4. A crude oil system exists at an initial reservoir pressure of 3000 psi and 185°F. The bubble point pressure is estimated at 2109 psi. The oil properties at the bubble point pressure are as follows: Bob ¼ 1.406 bbl/STB Rsb ¼ 692 scf/STB γ g ¼ 0.876 API ¼ 41.9° Calculate a. Oil density at the bubble point pressure. b. Oil density at 3000 psi. c. Bo at 3000 psi. 5. The PVT data as shown in Table 4.22 were obtained on a crude oil sample taken from the Nameless field. The initial reservoir pressure was 3600 psia at 160°F. The average specific gravity of the solution gas is 0.65. The reservoir contains 250 MMbbl of oil initially in place. The oil has a bubble point pressure of 2500 psi. a. Calculate the two-phase oil FVF at 1. 3200 psia. 2. 2800 psia. 3. 1800 psia. b. Calculate the initial oil in place as expressed in MMSTB. c. What is the initial volume of dissolved gas in the reservoir? d. Calculate the oil compressibility coefficient at 3200 psia. 460 CHAPTER 4 PVT Properties of Crude Oils Table 4.22 PVT Data for Problem 5 Pressure (psia) 3600 3200 2800 2500 2400 1800 1200 600 200 Rs (scf/STB) Bo (bbl STB) 567 554 436 337 223 143 1.31 1.317 1.325 1.333 1.31 1.263 1.21 1.14 1.07 6. The PVT data below was obtained from the analysis of a bottom-hole sample. p (psia) Relative Volume, V/Vsat 3000 2927 2703 2199 1610 1206 999 1.0000 1.0063 1.0286 1.1043 1.2786 1.5243 1.7399 a. Plot the Y-function versus pressure on rectangular coordinate paper. b. Determine the constants in the equation Y ¼ mp + b using the method of least squares. c. Recalculate relative oil volume from the equation. 7. A 295 cm3 of a crude oil sample was placed in a PVT at an initial pressure of 3500 psi. The cell temperature was held at a constant temperature of 220°F. A DL test was then performed on the crude oil sample with the recorded measurements as given in Table 4.23. Using the recorded measurements and assuming an oil gravity of 40°API, calculate the following PVT properties: a. Oil FVF at 3500 psi. b. Gas solubility at 3500 psi. c. Oil viscosity at 3500 psi. d. Isothermal compressibility coefficient at 3300 psi. e. Oil density at 1000 psi. 8. Experiments were made on a bottom-hole crude oil sample taken from the North Grieve field to determine the gas solubility and oil FVF as a Problems 461 Table 4.23 Crude Oil Sample for Problem 7 p (psi) T (°F) Total Volume (cm3) 3500 3300 3000a 2000 1000 14.7 220 220 220 220 220 60 290 294 300 323.2 375.2 – Volume of Liquids (cm3) Vol. Liberated Gas (cm3) Gas Gravity 290 294 300 286.4 271.5 179.53 0 0 0 0.1627 0.184 0.5488 – – – 0.823 0.823 0.823 a Bubble point pressure. function of pressure. The initial reservoir pressure was recorded as 3600 psia and reservoir temperature was 130°F. The data in Table 4.24 were obtained from the measurements. At the end of the experiments, the API gravity of the oil was measured as 40°. If the average specific gravity of the solution gas is 0.7, calculate a. Total FVF at 3200 psia. b. Oil viscosity at 3200 psia. c. Isothermal compressibility coefficient at 1800 psia. 9. You are producing a 35°API crude oil from a reservoir at 5000 psia and 140°F. The bubble point pressure of the reservoir liquids is 4000 psia at 140°F. Gas with a gravity of 0.7 is produced with the oil at a rate of 900 scf/STB. Calculate a. Density of the oil at 5000 psia and 140°F. b. Total FVF at 5000 psia and 140°F. 10. An undersaturated oil reservoir exists at an initial reservoir pressure 3112 psia and a reservoir temperature of 125°F. The bubble point of Table 4.24 Data for Problem 8 Pressure (psia) Rs (scf/STB) Bo (bbl/STB) 3600 3200 2800 2500 2400 1800 1200 600 200 567 567 567 567 554 436 337 223 143 1.31 1.317 1.325 1.333 1.31 1.263 1.21 1.14 1.07 462 CHAPTER 4 PVT Properties of Crude Oils Table 4.25 Pressure vs. Oil Relationship for Problem 10 Pressure (psia) Bo (bbl/STB) 3112 2800 2400 2000 1725 1700 1600 1500 1400 1.4235 1.429 1.437 1.4446 1.4509 1.4468 1.4303 1.4139 1.3978 the oil is 1725 psia. The crude oil has the pressure versus oil FVF relationship found in Table 4.25. The API gravity of the crude oil and the specific gravity of the solution gas are 40° and 0.65, respectively. Calculate the density of the crude oil at 3112 psia and 125°F. 11. A PVT cell contains 320 cm3 of oil and its bubble point pressure of 2500 psia and 200°F. When the pressure was reduced to 2000 psia, the volume increased to 335.2 cm3. The gas was bled off and found to occupy a volume of 0.145 scf. The volume of the oil was recorded as 303 cm3. The pressure was reduced to 14.7 psia and the temperature to 60°F while 0.58 scf of gas was evolved, leaving 230 cm3 of oil with a gravity of 42°API. Calculate a. Gas compressibility factor at 2000 psia. b. Gas solubility at 2000 psia. 12. The composition of a crude oil and the associated equilibrium gas is given below. The reservoir pressure and temperature are 3000 psia and 140°F, respectively. Component xi yi C1 C2 C3 n-C4 n-C5 C6 C7+ 0.40 0.08 0.07 0.03 0.01 0.01 0.40 0.79 0.06 0.05 0.04 0.02 0.02 0.02 References 463 The following additional PVT data are available: Molecular weight of C7+ ¼ 215 Specific gravity of C7+ ¼ 0.77 Calculate the surface tension 13. Estimate the MMP for pure CO2 at 630°R using the following methods: a. EVP method. b. Petroleum Recovery Institute method. c. Yellig and Metcalfe’s method. REFERENCES Abu-Khamsin, A., Al-Marhoun, M., 1991. Development of a new correlation for bubblepoint viscosity. Arab. J. Sci. Eng. 16 (2A), 99. Ahmed, T., 1988. Compositional modeling of Tyler and Mission Canyon formation oils with CO2 and lean gases. Final report submitted to Montana’s On a New Track for Science (MONTS), Montana National Science Foundation Grant Program. Ahmed, T., 2000. Removing well bore liquid blockage by gas injection. Paper 002–00, In: Presented at the Rio Oil and Gas Expo and Conference, Rio de Janeiro, Brazil, 16–19 October 2000. Ahmed, T., 2014. Equations of State and PVT Analysis. Course Material. Ahmed, T., 2015. Equations of State and PVT Analysis. Course Material. Alani, G.H., Kennedy, H.T., 1960. Volume of liquid hydrocarbons at high temperatures and pressures. Trans. AIME 219, 288–292. Al-Shammasi, A., 1999. Bubble-point pressure and formation volume factor correlations. SPE paper 53185, In: Presented at the SPE Middle East Conference, Bahrain, 20–23 February 1999. Alston, R.B., Kokolis, G.P., James, C.F., 1985. CO2 minimum miscibility pressure: a correlation for impure CO2 streams and live oil streams. Soc. Pet. Eng. J. 25 (2), 268–274. Amyx, J.M., Bass, D.M., Whiting, R., 1960. Petroleum Reservoir Engineering — Physical Properties. McGraw-Hill, New York, NY. Beal, C., 1946. The viscosity of air, water, natural gas, crude oils and its associated gases at oil field temperatures and pressures. Trans. AIME 165, 94–112. Beggs, H.D., Robinson, J.R., 1975. Estimating the viscosity of crude oil systems. J. Pet. Technol. 27 (9), 1140–1141. Brill, J., Beggs, D., 1973. A study of two-phase flow in inclined pipes. J. Pet. Technol. 25 (5), 607–617. Chew, J., Connally Jr., C.A., 1959. A viscosity correlation for gas-saturated crude oils. Trans. AIME 216, 23–25. Cho, S., Civan, F., Starting, K., 1985. A correlation to predict maximum condensation for retrograde fluids. Paper SPE 14268, In: Presented at the SPE Annual Meeting, Las Vegas, 22–25 September 1985. Conrard, P., 1987. Discussion of analysis and correlation of nitrogen and lean gas miscibility pressure. SPE Reserv. Eng. Craft, B., Hawkins, M., 1959. Applied Petroleum Reservoir Engineering. Prentice-Hall, Englewood Cliffs, NJ. 464 CHAPTER 4 PVT Properties of Crude Oils Cragoe, C., 1997. Thermodynamic Properties of Petroleum Products. U.S. Department of Commerce, Washington, DC. p. 97. Cronquist, C., 1978. Carbon dioxide dynamic displacement with light reservoir oils. In: Paper Presented at the U.S. Department of Energy Annual Symposium, Tulsa, 28–30 August 1978. Dindoruk, B., Christman, P., 2004. PVT properties and viscosity correlations for Gulf of Mexico oils. SPE Reserv. Eval. Eng. 7, 427–437. Dodson, C., Goodwill, D., Mayer, E., 1953. Application of laboratory PVT data tir engineering problems. Journal of Petroleum Technology. Fanchi, J.R., 1985. Calculation of parachors for composition simulation. J. Pet. Technol. 2049–2050. Fevang, D., Whitson, C., 1994. Accurate in-situ compositions in petroleum reservoirs. Paper SPE 28829, In: Presented at the European Petroleum Conference, London, 25–27 October 1994. Firoozabadi, A., Aziz, K., 1986. Analysis and correlation of nitrogen and lean gas miscibility pressure. SPE Reserv. Eng. 1 (6), 575–582. Firoozabadi, A., et al., 1988. Surface tension of reserve crude oil/gas systems. SPE Reserv. Eng. Glaso, O., 1980. Generalized pressure-volume-temperature correlations. J. Pet. Technol. 32 (5), 785–795. Glaso, O., 1990. Miscible displacement: recovery tests with nitrogen. SPE Reserv. Eng. 5 (1), 61–68. Haas, J., 1976. Physical properties of the coexisting phases and thermochemical properties of H2O. Geological survey bulletin. Hosein, R., Mayrhoo, R., McCain Jr., D.W., 2014. Determination of Bubble-Point and Dew-Point Pressure Without a Visual Cell. In: SPE 169947. Presented at the SPE Biennial Energy Resources conference held in Port of Spain, Trinidad, 9–11 June 2014. Hudgins, D.A., Liave, F., Chung, F., 1990. Nitrogen miscible displacement of light crude oil. SPE Reserv. Eng. 5 (1), 100–106. Katz, D., 1942. Drilling and Production Practice. American Petroleum Institute, Dallas, TX. Katz, D., Saltman, W., 1939. Surface tension of hydrocarbons. Ind. Eng. Chem. 31, 91. Kennedy, G., 1954. Pressure-volume-temperature relations in CO2. Am. J. Sci. 252, 225. Khan, S., et al., 1987. Viscosity correlations for Saudi Arabia oils. SPE paper 15720, Society of Petroleum Engineers, Dallas, TX. Lange, E., 1998. Correlation and prediction of residual oil saturation for gas-injectionenhanced oil-recovery processes. SPE Reserv. Eval. Eng. 1, 127–133. Lasater, J.A., 1958. Bubble-point pressure correlation. Trans. AIME 213, 379–381. Lawrence, J., Gupta, D., 2009. Quality assessment and consistency evaluation of hydrocarbon PVT data compositions. IPTC 13784, In: Presented at the IPTC Held in Doha, Qatar, 7–9 December 2009. Little, J.E., Kennedy, H.T., 1968. A correlation of the viscosity of hydrocarbon systems with pressure, temperature, and composition. Soc. Pet. Eng. J. 8, 157–162. Lohrenz, J., Bra, B.G., Clark, C.R., 1964. Calculating viscosities of reservoir fluids from their compositions. J. Pet. Technol. 16 (October), 1171–1176. Marhoun, M.A., 1988. PVT correlation for middle east crude oils. J. Pet. Technol. 40 (May), 650–666. References 465 McCain, W., 1991. The Properties of Petroleum Fluids. PennWell Publishing Company, Tulsa, OK. McCain, W., 2002. Analysis of black oil PVT reports revisited. Paper SPE 77386, In: Presented at the SPE Annual Meeting, San Antonio, TX, September 29–October 2, 2002. McCain, W., et al., 1988. The coefficient of isothermal compressibility of black oils at pressures below bubble point. SPE Form. Eval. 3, 659–662. McKetta, J., Wehe, A., 1962. Hydrocarbon/water and formation water correlations. In: Frick, T.C., Taylor, R.W. (Eds.), Petroleum Production Handbook, second ed. Society of Petroleum Engineers, Richardson, TX. Newitt, D.M., et al., 1996. Carbon dioxide. In: Din, F. (Ed.), Thermodynamic Functions of Gases, vol. 1. Butterworths, London, pp. 102–134. Orr, F.M., Jensen, C.M., 1986. Interpretation of pressure-composition phase diagram for CO2/crude-oil systems. Soc. Pet. Eng. J. 24, 485–497. Orr, F.M., Silva, M.K., 1987. Effect of oil composition on MMP—part 2. Correlation. SPE Reserv. Eng. 2, 479–491. Ostermann, R., et al., 1987. A correlation for increased gas gravity during pressure depletion. Paper SPE 16962, In: Presented at the SPE Annual Conference, Dallas, 27–30 September 1987. Petrosky, G.E., Farshad, F., 1993. Pressure-volume-temperature correlations for Gulf of Mexico crude oils. SPE paper 26644, In: Presented at the 68th Annual Technical Conference of the Society of Petroleum Engineers, Houston, 3–6 October 1993. Ramey, H., 1973. Correlation of surface and interfacial tensions of reserve fluids. SPE paper no. 4429, Society of Petroleum Engineers, Richardson, TX. Rowe, A., Chou, J., 1970. Pressure-volume-temperature correlation relation of aqueous NaCl solutions. J. Chem. Eng. Data 15, 61–66. Sebastian, H., Lawrence, D., 1992. Nitrogen minimum miscibility pressures. Paper SPE 24134, In: Presented at the SPE/DOE Meeting, Tulsa, OK, 22–24 April 1992. Sebastian, H.M., Wenger, R.S., Renner, T.A., 1985. Correlation of minimum miscibility pressure for impure CO2 streams. J. Pet. Technol. 37, 2076–2082. Standing, M.B., 1947. A pressure-volume-temperature correlation for mixtures of California oils and gases. Drilling and Production Practice. API, Washington, D.C, pp. 275–287. Standing, M.B., 1974. Petroleum Engineering Data Book. Norwegian Institute of Technology, Trondheim. Standing, M.B., 1977. Volumetric and Phase Behavior of Oil Field Hydrocarbon Systems. Society of Petroleum Engineers, Dallas, TX. pp. 125–126. Standing, M.B., 1981. Volumetric and Phase Behavior of Oil Field Hydrocarbon Systems, ninth ed. Society of Petroleum Engineers, Dallas, TX. Standing, M.B., Katz, D.L., 1942. Density of natural gases. Trans. AIME 146, 140–149. Sugden, S., 1924. The variation of surface tension. VI. The variation of surface tension with temperature and some related functions. J. Chem. Soc. 125, 32–39. Sutton, R.P., Farshad, F.F., 1984. Evaluation of empirically derived PVT properties for Gulf of Mexico crude oils. Paper SPE 13172, In: Presented at the 59th Annual Technical Conference, Houston, 1984. Trube, A.S., 1957. Compressibility of undersaturated hydrocarbon reservoir fluids. Trans. AIME 210, 341–344. 466 CHAPTER 4 PVT Properties of Crude Oils Vasquez, M., Beggs, H.D., 1980. Correlations for fluid physical property prediction. J. Pet. Technol. 32, 968–970. Weinaug, C., Katz, D.L., 1943. Surface tension of methane-propane mixtures. Ind. Eng. Chem. 25, 35–43. Whitson, C.H., Brule, M.R., 2000. Phase Behavior. SPE, Richardson, TX. Yellig, W.F., Metcalfe, R.S., 1980. Determination and prediction of CO2 minimum miscibility pressures. J. Pet. Technol. 32, 160–168. Yuan, H., et al., 2005. Simplified method for calculation of MMP. SPE paper 77381, Society of Petroleum Engineers, Dallas, TX. Chapter 5 Equations of State and Phase Equilibria A phase is the part of a system that is uniform in physical and chemical properties, homogeneous in composition, and separated from other, coexisting phases by definite boundary surfaces. The most important phases occurring in petroleum production are the hydrocarbon liquid phase and gas phase. Water is also commonly present as an additional liquid phase. These can coexist in equilibrium when the variables describing change in the entire system remain constant in time and position. The chief variables that determine the state of equilibrium are system temperature, system pressure, and composition. The conditions under which these different phases can exist are a matter of considerable practical importance in designing surface separation facilities and developing compositional models. These types of calculations are based on the following two concepts, equilibrium ratios and flash calculations, which are discussed next. EQUILIBRIUM RATIOS As indicated in Chapter 1, a system that contains only one component is considered the simplest type of hydrocarbon system. The qualitative understanding of the relationship that exists between temperature, T, pressure, p, and volume, V, of pure components can provide an excellent basis for understanding the phase behavior of complex hydrocarbon mixtures. In a multicomponent system, the partitioning and the tendency of components to escape between the liquid and gas phase are described by the equilibrium ratio “Ki” of a given component. The equilibrium ratio is defined as the ratio of the mole fraction of the component in the gas phase “yi” to the mole fraction of the component in the liquid phase “xi”. Mathematically, the relationship is expressed as Ki ¼ yi xi Equations of State and PVT Analysis. http://dx.doi.org/10.1016/B978-0-12-801570-4.00005-2 Copyright # 2016 Elsevier Inc. All rights reserved. (5.1) 467 468 CHAPTER 5 Equations of State and Phase Equilibria where Ki ¼ equilibrium ratio of component i yi ¼ mole fraction of component i in the gas phase xi ¼ mole fraction of component i in the liquid phase The above expression indicates that when the Ki-value for a given component is greater than 1, it signifies that the component tends to concentrate in the gas phase. Phase equilibria and other phase behavior calculations require accurate estimation equilibrium ratio values. At pressures below 100 psia, Raoult’s and Dalton’s laws for ideal solutions provide a simplified means of predicting equilibrium ratios. Raoult’s law states that the partial pressure, pi, of a component in a multicomponent system is the product of its mole fraction in the liquid phase, xi, and the vapor pressure of the component, pvi: pi ¼ xi pvi (5.2) where pi ¼ partial pressure of a component i, psia pvi ¼ vapor pressure of component i, psia xi ¼ mole fraction of component i in the liquid phase Dalton’s law states that the partial pressure of a component is the product of its mole fraction in the gas phase “yi” and system pressure “p”; that is, p i ¼ yi p (5.3) where p ¼ system pressure, psia. At equilibrium and in accordance with the previously cited laws, the partial pressure exerted by a component in the gas phase must be equal to the partial pressure exerted by the same component in the liquid phase. Therefore, equating the equations describing the two laws, gives xi pvi ¼ yi p By rearranging the preceding relationship and introducing the concept of the equilibrium ratio gives yi pvi ¼ ¼ Ki xi p (5.4) Eq. (5.4) suggests that for an “ideal solution” with an overall composition of “zi” that exist at low pressure and temperature exhibits the following two properties: n n The value of equilibrium ratio “Ki” for any component is independent of the overall composition. As shown in Fig. 5.1, the vapor pressure “pvi” is only a function of temperature, which implies that the K-value is only a function of system pressure “p” and temperature “T”. Equilibrium Ratios 469 n FIGURE 5.1 Vapor pressure chart for hydrocarbon components. Courtesy of the Gas Processors Suppliers Association. Published in the GPSA Engineering Data Book, Tenth Edition, 1987. 470 CHAPTER 5 Equations of State and Phase Equilibria Taking the logarithm of both sides of Eq. (5.4) and rearranging gives log ðKi Þ ¼ log ðpvi Þ log ðpÞ This relationship indicates that for a constant temperature, T (implying a constant pvi), the component equilibrium ratio “Ki” is a linear function of pressure with a slope of 1 when plotted on a log-log scale. It is appropriate at this stage to introduce and define the following nomenclatures: zi ¼ mole fraction of component i in the entire hydrocarbon mixture nt ¼ total number of moles of the hydrocarbon mixture, lb-mol nL ¼ total number of moles in the liquid phase nυ ¼ total number of moles in the vapor (gas) phase By definition, nt ¼ nL + nυ (5.5) Eq. (5.5) indicates that the total number of moles in the system is equal to the total number of moles in the liquid phase plus the total number of moles in the vapor phase. A material balance on the ith component results in zi nt ¼ xi nL + yi nv (5.6) where zint ¼ total number of moles of component i in the system xinL ¼ total number of moles of component i in the liquid phase yinv ¼ total number of moles of component i in the vapor phase Also, by the definition of the total mole fraction in a hydrocarbon system, we may write X zi ¼ 1 (5.7) i X xi ¼ 1 (5.8) i X yi ¼ 1 (5.9) i It is convenient to perform all the phase equilibria calculations on the basis of 1 mol of the hydrocarbon mixture; that is, nt ¼ 1. That assumption reduces Eqs. (5.5), (5.6) to nL + nυ ¼ 1 (5.10) xi nL + yi nυ ¼ zi (5.11) Combining Eqs. (5.4), (5.11) to eliminate yi from Eq. (5.11) gives xi nL + ðxi Ki Þnυ ¼ z Flash Calculations 471 Solving for xi yields zi nL + nυ Ki xi ¼ (5.12) Eq. (5.11) can also be solved for yi by combining it with Eq. (5.4) to eliminate xi, which gives yi ¼ zi Ki ¼ xi Ki nL + nυ Ki (5.13) Combining Eq. (5.12) with Eq. (5.8) and Eq. (5.13) with Eq. (5.9) results in X X xi ¼ i X i yi ¼ i X i zi ¼1 nL + nυ Ki (5.14) zi Ki ¼1 nL + nυ Ki (5.15) As X i yi X xi ¼ 0 i then X i X zi Ki zi ¼0 nL + nυ Ki n + L nυ Ki i Rearranging the above expression to give X zi ðKi 1Þ i nL + nυ Ki ¼0 Replacing nL in the above relationship with (1 nυ) yields f ðnυ Þ ¼ X zi ðKi 1Þ ¼0 nυ ðKi 1Þ + 1 i (5.16) Eqs. (5.12)–(5.16) provide the necessary relationships to perform volumetric and compositional calculations on a hydrocarbon system. These types of calculation are referred to as flash calculations. Details of performing flash calculations are described next. FLASH CALCULATIONS Flash calculations are an integral part of all reservoir and process engineering calculations. They are required whenever it is desirable to know the amounts (in moles) of hydrocarbon liquid and gas coexisting in a reservoir 472 CHAPTER 5 Equations of State and Phase Equilibria or a vessel at a given pressure and temperature. These calculations also are performed to determine the composition of the existing hydrocarbon phases. Given the overall composition of a hydrocarbon system at a specified pressure and temperature, flash calculations are performed to determine the moles of the gas phase, nυ, moles of the liquid phase, nL, composition of the liquid phase, xi, and composition of the gas phase, yi. The computational steps for determining nL, nv, yi, and xi of a hydrocarbon mixture with a known overall composition of zi and characterized by a set of equilibrium ratios, Ki, are summarized in the following steps: Step 1. Calculate number of moles of gas phase “nυ”: Eq. (5.16) can be solved for the number of moles of the vapor phase nυ by using the Newton-Raphson iteration technique. In applying this iterative technique, assume any arbitrary value of nυ between 0 and 1, such as nυ ¼ 0.5. A good assumed value may be calculated from the following relationship: nυ ¼ A=ðA + BÞ where A¼ X ½zi ðKi 1Þ i X 1 zi 1 B¼ Ki i Using the assumed value of nv, evaluate the function f(nυ) as given by Eq. (5.16); that is, f ðnυ Þ ¼ X zi ðKi 1Þ nv ðKi 1Þ + 1 i If the absolute value of the function f(nυ) is smaller than a preset tolerance, such as 106, then the assumed value of nυ is the desired solution. If the absolute value of f(nυ) is greater than the preset tolerance, then a new value of (nυ)new is calculated from the following expression: f ðnυ Þ ðnυ Þnew ¼ nυ 0 f ðnυ Þ with the derivative f 0 (nυ) as given by f 0 ðnυ Þ ¼ ( X zi ðKi 1Þ2 i ½nυ ðKi 1Þ + 12 ) Flash Calculations 473 where (nυ)new is the new value of nυ to be used for the next iteration. This procedure is repeated with the new value of nυ until convergence is achieved; that is, when jf ðnυ Þj eps or alternatively jðnυ Þnew ðnυ Þj eps in which eps is a selected tolerance, such as eps ¼ 106. Step 2. Calculate number of moles of liquid phase “nL”: The number of moles of the liquid phase “nL” can be determined from Eq. (5.10) to give nL ¼ 1 nυ Step 3. Calculate the composition of the liquid phase “xi”: Given nv and nL, calculate the composition of the liquid phase by applying Eq. (5.12): xi ¼ zi nL + nυ Ki Step 4. Calculate the composition of the gas phase “yi”: Determine the composition of the gas phase from Eq. (5.13): yi ¼ zi Ki ¼ xi Ki nL + nυ Ki EXAMPLE 5.1 A hydrocarbon mixture with the following overall composition is flashed in a separator at 50 psia and 100°F: Component zi C3 i-C4 n-C4 i-C5 n-C5 C6 0.20 0.10 0.10 0.20 0.20 0.20 Assuming an ideal solution behavior, perform flash calculations. Solution Step 1. Determine the vapor pressure pvi from the Cox chart (Fig. 5.1) and calculate the equilibrium ratios using Eq. (5.4). The results are shown below. 474 CHAPTER 5 Equations of State and Phase Equilibria Component zi pvi at 100°F Ki 5 pvi/50 C3 i-C4 n-C4 i-C5 n-C5 C6 0.20 0.10 0.10 0.20 0.20 0.20 190 72.2 51.6 20.44 15.57 4.956 3.80 1.444 1.032 0.4088 0.3114 0.09912 Step 2. Solve Eq. (5.16) for nυ using the Newton-Raphson method: f ðnυ Þ ðnυ Þn ¼ nυ 0 f ðnυ Þ Iteration nυ f(nv) 0 1 2 3 4 0.08196579 0.1079687 0.1086363 0.1086368 0.1086368 3.073(102) 8.894(104) 7.60(107) 1.49(108) 0.0 to give nυ ¼ 0.1086368. Step 3. Solve for nL: nL ¼ 1 nυ nL ¼ 1 0:1086368 ¼ 0:8913631 Step 4. Solve for xi and yi to yield xi ¼ zi nL + nυ Ki yi ¼ xi Ki The component results are shown below. Component zi Ki xi 5 zi/(0.8914 + 0.1086 Ki) yi 5 xi Ki C3 i-C4 n-C4 i-C5 n-C5 C6 0.20 0.10 0.10 0.20 0.20 0.20 3.80 1.444 1.032 0.4088 0.3114 0.09912 0.1534 0.0954 0.0997 0.2137 0.2162 0.2216 0.5829 0.1378 0.1029 0.0874 0.0673 0.0220 Flash Calculations 475 Binary System Note that, for a binary system (ie, a two-component system), flash calculations can be performed without restoring to the preceding iterative technique. Flash calculations can be performed by applying the following steps. Step 1. Solve for the composition of the liquid phase, xi. For a twocomponent system, Eqs. (5.8), (5.9) can be expanded as X xi ¼ x1 + x2 ¼ 1 i X yi ¼ y1 + y2 ¼ K1 x1 + K2 x2 ¼ 1 i Solving these expressions for the liquid composition, x1 and x2, gives x1 ¼ 1 K2 K1 K2 and x2 ¼ 1 x1 where x1 ¼ mole fraction of the first component in the liquid phase x2 ¼ mole fraction of the second component in the liquid phase K1 ¼ equilibrium ratio of the first component K2 ¼ equilibrium ratio of the second component Step 2. Solve for the composition of the gas phase, yi. From the definition of the equilibrium ratio, calculate the composition of the liquid as follows: y1 ¼ x1 K1 y2 ¼ x2 K2 ¼ 1 y1 Step 3. Solve for the number of moles of the vapor phase, nυ, and liquid phase, nL. Arrange Eq. (5.12) to solve for nυ by using the mole fraction and K-value of one of the two components to give nυ ¼ z1 x 1 x1 ðK1 1Þ and nL ¼ 1 nυ Exact results will be obtained if selecting the second component; that is, nυ ¼ z2 x 2 x2 ðK2 1Þ 476 CHAPTER 5 Equations of State and Phase Equilibria and nL ¼ 1 nυ where z1 ¼ mole fraction of the first component in the binary system x1 ¼ mole fraction of the first component in the liquid phase z2 ¼ mole fraction of the second component in the binary system x2 ¼ mole fraction of the second component in the liquid phase K1 ¼ equilibrium ratio of the first component K2 ¼ equilibrium ratio of the second component The equilibrium ratio, as calculated by Eq. (5.4) in terms of vapor pressure and system pressure, proved inadequate. Eq. (5.4) was developed based on the following assumptions: n n The vapor phase is an ideal gas as described by Dalton’s law. The liquid phase is an ideal solution as described by Raoult’s law. The combination of the above two assumptions is unrealistic and results in inaccurate predictions of equilibrium ratios at high pressures. EQUILIBRIUM RATIOS FOR REAL SOLUTIONS For a real solution, the equilibrium ratios are no longer a function of the pressure and temperature alone but also the composition of the hydrocarbon mixture. This observation can be stated mathematically as Ki ¼ K ðp, T, zi Þ Numerous methods have been proposed for predicting the equilibrium ratios of hydrocarbon mixtures. These correlations range from a simple mathematical expression to a complicated expression containing several compositional dependent variables. The following methods are presented: Wilson’s correlation, Standing’s correlation, the convergence pressure method, and Whitson and Torp’s correlation. Wilson’s Correlation Wilson (1968) proposed a simplified thermodynamic expression for estimating K-values. The proposed expression has the following form: Ki ¼ pci Tci exp 5:37ð1 + ωi Þ 1 p T (5.17) Equilibrium Ratios for Real Solutions 477 where pci ¼ critical pressure of component i, psia p ¼ system pressure, psia Tci ¼ critical temperature of component i, °R T ¼ system temperature, °R This relationship generates reasonable values for the equilibrium ratio when applied at low pressures. The main advantage of using Wilson’s expression to estimate the equilibrium ratios is in its ability to provide a consistent set of K-values at a specified pressure and temperature; that is, KC1 > KC2 > KC3 >, and so on. As detailed later in this chapter, equations of state rely on iterative technique to calculate the K-value with stating values based on Wilson’s equation. Standing’s Correlation Several authors, including Hoffmann et al. (1953), Brinkman and Sicking (1960), Kehn (1964), and Dykstra and Mueller (1965), suggested that any hydrocarbon or nonhydrocarbon component could be uniquely characterized by combining its boiling point temperature, critical temperature, and critical pressure into a characterization parameter, which is defined by the following expression: Fi ¼ bi ½1=Tbi 1=T (5.18) log ðpci =14:7Þ ½1=Tbi 1=Tci (5.19) with bi ¼ where Fi ¼ component characterization factor Tbi ¼ normal boiling point of component i, °R. Based on the characterization parameter “Fi,” Standing (1979) derived a set of equations that fit the graphical presentations of the experimentally determined equilibrium ratio of Katz and Hachmuth (1937). The proposed form of the correlation is based on an observation that plots of log(Ki p) versus Fi at a given pressure often form straight lines with a slope of “c” and intercept of “a”. The basic equation of the straight-line relationship is given by log ðKi pÞ ¼ a + cFi Solving for the equilibrium ratio, Ki, gives 478 CHAPTER 5 Equations of State and Phase Equilibria 1 Ki ¼ 10ða + cFi Þ p (5.20) where the coefficients “a” and “c” in the relationship are the intercept and the slope of the line, respectively. From six isobar plots of log(Ki p) versus Fi for 18 sets of equilibrium ratio values, Standing correlated the coefficients “a” and “c” with the pressure, to give a ¼ 1:2 + 0:00045p + 15 108 p2 c ¼ 0:89 0:00017p 3:5 108 p2 (5.21) (5.22) Standing pointed out that the predicted values of the equilibrium ratios of N2, CO2, H2S, and C1–C6 can be improved considerably by changing n n the correlating parameter “bi” of Eq. (5.19) the boiling point “Tbi” of these components. The author proposed the following optimized modified values: Component bi Tbi (°R) N2 CO2 H2S C1 C2 C3 i-C4 n-C4 i-C5 n-C5 C6 n-C6 n-C7 n-C8 n-C9 n-C10 470 652 1136 300 1145 1799 2037 2153 2368 2480 2738 2780 3068 3335 3590 3828 109 194 331 94 303 416 471 491 542 557 610 616 616 718 763 805 K-Value of the Plus Fraction When making flash calculations, the equilibrium ratio for each component including the plus fraction (eg, C7+) must be calculated at the prevailing pressure and temperature. Katz and Hachmuth (1937) proposed a Equilibrium Ratios for Real Solutions 479 “rule-of-thumb” that might be applied for light hydrocarbon mixtures that suggest the K-value of C7+ can be taken as 15% of the K of C7; that is, KC7 + ¼ 0:15KC7 Standing offered an alternative approach for determining the K-value of the heptanes and heavier fractions. By imposing experimental equilibrium ratio values for C7+ on Eq. (5.20), Standing calculated the corresponding characterization factors, Fi, for the plus fraction; that is, log FC7 + ¼ KC7 + p c a The calculated FC7 + values were used to specify or FIND the pure normal paraffin hydrocarbon having the K-value equivalent to that of the C7+ fraction. Standing suggested the following computational steps for determining the parameters b and Tb of the heptanes-plus fraction. Step 1. Determine from the following relationship the number of carbon atoms “n” of the normal paraffin hydrocarbon that have the K-value of the C7+ fraction: n ¼ 7:30 + 0:0075ðT 460Þ + 0:0016p (5.23) Step 2. Calculate the correlating parameter “b” and the boiling point “Tb” of that normal paraffin component from the following expression: b ¼ 1013 + 324n 4:256n2 (5.24) Tb ¼ 301 + 59:85n 0:971n2 (5.25) Values of “b” and “Tb” can then be used to calculate the characterization parameter “Fi” of that particular normal paraffin hydrocarbon and be used to calculate its K-value, which is assigned to the C7+. It is interesting to note that numerous experimental phase equilibria data suggest that the equilibrium ratio for carbon dioxide can be closely approximated by the following relationship: KCO2 ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi KC1 KC2 where KCO2 ¼ equilibrium ratio of CO2 at system pressure p and temperature T KC1 ¼ equilibrium ratio of methane at p and T KC2 ¼ equilibrium ratio of ethane at p and T 480 CHAPTER 5 Equations of State and Phase Equilibria It should be pointed out that Standing’s proposed correlations are valid for pressures and temperatures that are less than 1000 psia and 200°F, respectively. This limitation on the use of Standing method lends itself appropriately for performing separation calculations, which usually operate within the range of Standing’s limit of applications. A Simplified K-Value Method For light hydrocarbon mixtures that exist at pressure below 1000 psia, the K-value for the methane fraction can be estimated from the following relationship: B exp A T KC1 ¼ p with A ¼ 2:0 107 p2 0:0005p + 9:5073 B ¼ 0:0001p2 0:456p + 856 where p ¼ pressure, psia T ¼ temperature, °R. It is possible to correlate the equilibrium ratios of C2–C7 with that of the Kvalue of C1 by the following relationship: Ki ¼ KC1 Ri lnðpKC1 Þ with the component characterization parameter Ri for component “i” as given by Ri ¼ ai T bi The temperature, T, is in °R. Values of the coefficients ai and bi for C2–C7 are tabulated below: Component ai bi C2 C3 i-C4 n-C4 i-C5 n-C5 C6 C7 0.00665 0.00431 0.00327 0.00245 0.00203 0.00147 0.00094 0.00091 1.29114 1.269876 1.206566 1.1002 0.910736 0.737 0.528 0.516263 Equilibrium Ratios for Real Solutions 481 This correlation provides a very rough estimate of the K-values for the above listed components at pressures below 1000 psia. The K-value for the C7+ can be estimated from the “rule-of-thumb” equation (ie, KC7 + ¼ 0:15KC7 ). EXAMPLE 5.2 A hydrocarbon mixture with the following composition is flashed at 1000 psia and 150°F. Component zi CO2 N2 C1 C2 C3 i-C4 n-C4 i-C5 n-C5 C6 C7+ 0.009 0.003 0.535 0.115 0.088 0.023 0.023 0.015 0.015 0.015 0.159 If the molecular weight and specific gravity of C7+ are 150.0 and 0.78, respectively, calculate the equilibrium ratios using: n n n Wilson’s correlation Standing’s correlation Simplified method Solution Using Wilson’s method Step 1. Calculate the critical pressure, critical temperature, and acentric factor of C7+ by using the characterization method of Riazi and Daubert discussed in Chapter 2 and Example 2.1, to give Tc ¼ 1139:4°R pc ¼ 320:3 psia ω ¼ 0:5067 Step 2. Apply Eq. (5.17) to get the results shown below: Ki 5 pci Tci exp 5:37ð1 + ωi Þ 1 1000 610 482 CHAPTER 5 Equations of State and Phase Equilibria Component P0 c (psia) T0 c (°R) ω Ki CO2 N2 C1 C2 C3 i-C4 n-C4 i-C5 n-C5 C6 C7+ 1071 493 667.8 707.8 616.3 529.1 550.7 490.4 488.6 436.9 320.3 547.9 227.6 343.37 550.09 666.01 734.98 765.65 829.1 845.7 913.7 1139.4 0.225 0.040 0.0104 0.0986 0.1524 0.1848 0.2010 0.2223 0.2539 0.3007 0.5069 2.0923 16.343 7.155 1.263 0.349 0.144 0.106 0.046 0.036 0.013 0.00029 Using Standing’s method Step 1. Calculate the coefficients a and c from Eqs. (5.21), (5.22) to give a ¼ 1:2 + 0:00045p + 15 108 p2 a ¼ 1:2 + 0:00045 ð1000Þ + 15 108 ð1000Þ2 ¼ 1:80 c ¼ 0:89 0:00017p 3:5 108 p2 c ¼ 0:89 0:00017 ð1000Þ 3:5 108 ð1000Þ2 ¼ 0:685 Step 2. Calculate the number of carbon atoms, n, from Eq. (5.23) to give n ¼ 7:30 + 0:0075 ðT 460Þ + 0:0016p n ¼ 7:3 + 0:0075 ð150Þ + 0:0016 ð1000Þ ¼ 10:025 Step 3. Determine the parameter b and the boiling point, Tb, for the hydrocarbon component with n carbon atoms by using Eqs. (5.24), (5.25): b ¼ 1013 + 324n 4:256n2 b ¼ 1013 + 324ð10:025Þ 4:256ð10:025Þ2 ¼ 3833:369 Tb ¼ 301 + 59:85n 0:971n2 Tb ¼ 301 + 59:85ð10:025Þ 0:971ð10:025Þ2 ¼ 803:41°R Step 4. Apply Eqs. (5.18), (5.20) to give the results shown below: Fi ¼ bi ½1=Tbi 1=T 1 Ki ¼ 10ða + cFi Þ p Equilibrium Ratios for Real Solutions 483 Component bi Tbi Fi Ki CO2 N2 C1 C2 C3 i-C4 n-C4 i-C5 n-C5 C6 C7+ 652 470 300 1145 1799 2037 2153 2368 2480 2738 3833.369 194 109 94 303 416 471 491 542 557 610 803.41 2.292 3.541 2.700 1.902 1.375 0.985 0.855 0.487 0.387 0 1.513 2.344 16.811 4.462 1.267 0.552 0.298 0.243 0.136 0.116 0.063 0.0058 The Convergence Pressure Method Early high-pressure phase equilibria studies revealed that, when a hydrocarbon mixture of a fixed overall composition is held at a constant temperature as the pressure increases, the equilibrium values of all components converge toward a common value of unity at certain pressure. This pressure is termed the convergence pressure, pk, of the hydrocarbon mixture. The convergence pressure essentially is used to correlate the effect of the composition on equilibrium ratios. The concept of the convergence pressure can be better explained by examining Fig. 5.2. The figure is a schematic diagram of a typical set of equilibrium ratios plotted versus pressure on log-log paper for a hydrocarbon mixture held at a constant temperature. The illustration shows a tendency of the equilibrium ratios to converge isothermally to a value of Ki ¼ 1 for all components at a specific pressure; that is, convergence pressure. A different hydrocarbon mixture may exhibit a different convergence pressure. Standing (1977) suggested that the convergence pressure could be roughly correlated linearly with the molecular weight of the heptanes-plus fraction. Whitson and Torp (1981) expressed this relationship by the following equation: pk ¼ 60MC7 + 4200 where MC7 + is the molecular weight of the heptanes-plus fraction. (5.26) 484 CHAPTER 5 Equations of State and Phase Equilibria Constant temperature Me tha ne 10 Eth an Pro e Equilibrium ratio “K-value” pa ne Bu tan 1.0 es Pen tan es He xan es 0.1 Hep tan es 1000 Pressure 100 PK 10,000 n FIGURE 5.2 Equilibrium ratios for a hydrocarbon system. Whitson and Torp reformulated Wilson’s equation [Eq. (5.17)] to yield more accurate results at higher pressures by incorporating the convergence pressure into the correlation, to give Ki ¼ A1 pci pci Tci exp 5:37Að1 + ωi Þ 1 pk p T (5.27) with the exponent “A” roughly approximated by the following expression: A¼ 1 0:7 p pk where p ¼ system pressure, psig pk ¼ convergence pressure, psig T ¼ system temperature, °R ωi ¼ acentric factor of component i Equilibrium Ratios for the Plus Fractions 485 EXAMPLE 5.3 Rework Example 5.2 and calculate the equilibrium ratios by using Whitson and Torp’s method. Solution Step 1. Determine the convergence pressure from Eq. (5.26) to give pk ¼ 60MC7 + 4200 ¼ 9474 Step 2. Calculate the coefficient A: 0:7 p A¼1 pk A¼1 1000 0:7 ¼ 0:793 9474 Step 3. Calculate the equilibrium ratios from Eq. (5.27) to give the results shown below: pci 0:7931 pci Tci Ki ¼ exp 5:37Að1 + ωi Þ 1 9474 1000 610 Component pc T0 c (°R) ω Ki CO2 N2 C1 C2 C3 i-C4 n-C4 i-C5 n-C5 C6 C7+ 1071 493 667.8 707.8 616.3 529.1 550.7 490.4 488.6 436.9 320.3 547.9 227.6 343.37 550.09 666.01 734.98 765.65 829.1 845.7 913.7 1139.4 0.225 0.040 0.0104 0.0986 0.1524 0.1848 0.2010 0.2223 0.2539 0.3007 0.5069 2.9 14.6 7.6 2.1 0.7 0.42 0.332 0.1794 0.150 0.0719 0.683(103) EQUILIBRIUM RATIOS FOR THE PLUS FRACTIONS The equilibrium ratios of the plus fractions often behave in a manner different from the other components of a system. This is because the plus fraction, in itself, is a mixture of components. Several techniques have been proposed for estimating the K-value of the plus fractions. Three methods are presented next. 486 CHAPTER 5 Equations of State and Phase Equilibria n n n Campbell’s method Winn’s method Katz’s method Campbell’s Method Campbell (1976) proposed that the plot of log(Ki) versus (Tci)2 for each component is a linear relationship for any hydrocarbon system. Campbell suggested that, by drawing the best straight line through the points for propane through hexane components, the resulting line can be extrapolated to obtain the K-value of the plus fraction. Alternatively, a plot of log (Ki) versus (1=Tbi ) of each fraction in the mixture could also produce a straight line that can be extrapolated to obtain the equilibrium ratio of the plus fraction from the reciprocal of its average boiling point. Winn’s Method Winn (1954) proposed the following expression for determining the equilibrium ratio of heavy fractions with a boiling point above 210°F: KC + ¼ KC7 ðKC2 =KC7 Þb (5.28) where KC + ¼value of the plus fraction KC7 ¼ K-value of n-heptane at system pressure, temperature, and convergence pressure KC2 ¼ K-value of ethane b ¼ volatility exponent Winn correlated, graphically, the volatility component, b, of the heavy fraction, with the atmosphere boiling point, as shown in Fig. 5.3. This graphical correlation can be expressed mathematically by the following equation: b ¼ a1 + a2 ðTb 460Þ + a3 ðTb 460Þ2 + a4 ðTb 460Þ3 + a5 =ðT 460Þ (5.29) where Tb ¼ boiling point, °R a1–a5 ¼ coefficients with the following values: a1 ¼ 1:6744337 a2 ¼ 3:4563079 103 a3 ¼ 6:1764103 106 a4 ¼ 2:4406839 109 a5 ¼ 2:9289623 102 Equilibrium Ratios for the Plus Fractions 487 n FIGURE 5.3 Volatility exponent “b”. Katz’s Method Katz et al. (1959) suggested that a factor of 0.15 times the equilibrium ratio for the heptane component gives a reasonably close approximation to the equilibrium ratio for heptanes and heavier. This suggestion is expressed mathematically by the following equation: KC7 + ¼ 0:15KC7 (5.30) K-values for Nonhydrocarbon Components Lohrenze et al. (1963) developed the following set of correlations that express the K-values of H2S, N2, and CO2 as a function of pressure, temperature, and the convergence pressure, pk. For H2 S : " p 0:8 1399:2204 6:3992127 + lnðKH2 S Þ ¼ 1 pk T # 18:215052 1112446:2 ln ðpÞ 0:76885112 + T T2 488 CHAPTER 5 Equations of State and Phase Equilibria For N2 : p 0:4 1184:2409 11:294748 0:90459907 ln ðpÞ ln KN 2 ¼ 1 pk T For CO2 : p 0:6 152:7291 7:0201913 ln ðpÞ ln KCO 2 ¼ 1 pk T # 1719:2856 644740:69 1:8896974 + T T2 where T ¼ temperature, °R p ¼ pressure, psia pk ¼ convergence pressure, psia VAPOR-LIQUID EQUILIBRIUM CALCULATIONS The vast amount of experimental and theoretical work performed on equilibrium ratio studies indicates their importance in solving phase equilibrium problems in reservoir and process engineering. The basic phase equilibrium calculations include: n n n the dew point pressure “Pd” bubble point pressure “Pb” separator calculations. The use of the equilibrium ratio in performing these applications is discussed next. Dew Point Pressure The dew point pressure, pd, of a hydrocarbon system is defined as the pressure at which an infinitesimal quantity of liquid is in equilibrium with a large quantity of gas. For a total of 1 lb-mol of a hydrocarbon mixture (ie, nt ¼ 1), the following conditions are applied at the dew point pressure: nL 0 nv 1 yi ¼ zi Imposing the above constraints on Eq. (5.14), gives X i X X zi zi zi ¼ ¼ ¼1 nL + nυ Ki 0 + ð1:0ÞKi Ki i i where zi is the overall composition of the hydrocarbon system. (5.31) Vapor-Liquid Equilibrium Calculations 489 The solution of Eq. (5.31) for the dew point pressure, pd, involves a trialand-error process, as summarized in the following steps: Step 1. Assume a trial value of pd. A good starting value can be obtained by combining Wilson’s equation (ie, Eq. (5.17)) with Eq. (5.31) to give X zi i Ki ¼ 8 > < X> 9 > > = zi ¼ 1 > > p T i > : ci exp 5:37ð1 + ωi Þ 1 ci > ; pd T (5.32) Solving the above expression for pd to give an initial estimate of pd as pd ¼ 1 8 > < X> 9 > > = zi > > T i > :pci exp 5:37ð1 + ωi Þ 1 ci > ; T (5.33) Step 2. Using the assumed dew point pressure, calculate the equilibrium ratio, Ki, for each component at the system temperature. Step 3. Calculate the summation of Eq. (5.31) using the K-values of step 2 X zi =Ki i Step 4. The correct value of the dew point pressure is obtained when the sum 1; however, if n the sum ≪ 1, repeat steps 2 and 3 at a higher initial value of pressure n the sum ≫ 1, repeat the calculations with a lower initial value of pd. The above outlined procedure using Wilson’s correlation is intended to only provide approximations of the K-values and Pd. These approximations are used as initial values in equations of state application. However, using Eq. (5.33) will generate a very low value for the dew point pressure; perhaps close to the lower dew point value. The main reason is that when applying Eq. (5.33), the term “zi/[Pc exp(5.37(1 + ω)(1 Tc/T))]” for the plus fraction is much greater than the nearest component value by over 40 times, resulting in a large value of the denominator of Eq. (5.33). Because it’s only an approximation, the plus fraction should be excluded. 490 CHAPTER 5 Equations of State and Phase Equilibria EXAMPLE 5.4 A natural gas reservoir has the following composition: Component zi C1 C2 C3 i-C4 n-C4 i-C5 n-C5 C6 C7+ 0.7778 0.0858 0.0394 0.0083 0.016 0.0064 0.0068 0.0078 0.0517 The heptanes-plus fraction is characterized by a molecular weight and specific gravity of 150 and 0.78, respectively. The reservoir exists at 250°F; estimate the dew point pressure of the gas mixture. Solution Step 1. Calculate the critical properties of the C7+ using the Riazi-Daubert equation: θ ¼ a ðMÞb γ c exp½d M + eγ + f γ M Tc ¼ 544:2ð150Þ0:2998 ð0:78Þ1:0555 exp 1:3478 104 ð150Þ 0:61641ð0:78Þ + 0 ¼ 1139:4°R pc ¼ 4:5203 104 ð150Þ0:8063 ð0:78Þ1:6015 exp 1:8078 103 ð150Þ 0:3084ð0:78Þ + 0 ¼ 320:3 psia Tb ¼ 6:77857ð150Þ0:401673 ð0:78Þ1:58262 exp 3:77409 103 ð150Þ 2:984036ð0:78Þ 4:25288 103 ð150Þð0:78Þ ¼ 825:26°R Use Edmister’s Eq. (2.21) to estimate the acentric factor: 3½log ðpc =14:70Þ 1 7½Tc =Tb 1 3½log ð320:3=14:7Þ ω¼ 1 ¼ 0:5067 7½1139:4=825:26 1 ω¼ Step 2. Construct the following working table and approximate the dew point pressure by applying Eq. (5.33): pd ¼ 8 > < X> 1 9 > > = zi > T > :pci exp 5:37ð1 + ωi Þ 1 ci > ; T i> Vapor-Liquid Equilibrium Calculations 491 Component zi Pci Tci ωi Sum C1 C2 C3 i-C4 n-C4 i-C5 n-C5 C6 C7+ 0.7778 0.0858 0.0394 0.0083 0.016 0.0064 0.0068 0.0078 0.0517 667.8 707.8 616.3 529.1 550.7 490.4 488.6 436.9 320.3 343.37 550.09 666.01 734.98 765.65 829.1 845.7 913.7 1139.4 0.0104 0.0986 0.1524 0.1848 0.201 0.2223 0.2539 0.3007 0.5067 7.0699E-05 3.2101E-05 4.357E-05 1.9623E-05 4.8166E-05 3.9247E-05 5.0404E-05 0.00013244 0.02153184 n n Total sum of the last column ¼ 0.02197 to give a dew point pressure ¼ 1/0.02197 ¼ 45 psi Excluding the C7+, sum ¼ 0.000436 to give a dew point pressure ¼ 1/0.000436 ¼ 2292 psi Bubble Point Pressure At the bubble point pb, the hydrocarbon system is essentially liquid, except for an infinitesimal amount of vapor. For a total of 1 lb-mol of the hydrocarbon mixture, the following conditions are applied at the bubble point pressure: nL 1 nv 0 xi ¼ zi Obviously, under these conditions, xi ¼ zi. Imposing the above constraints to Eq. (5.15) yields X i X zi K i X zi K i ¼ ¼ ðzi K i Þ ¼ 1 nL + nv Ki 1 + ð0ÞKi i i (5.34) Following the procedure outlined for the dew point pressure determination, Eq. (5.34) is solved for the bubble point pressure, pb, by assuming various pressures and determining the pressure that produces K-values satisfying Eq. (5.34). During the iterative process, if X ðzi Ki Þ < 1, then the assumed pressure is high i X ðzi Ki Þ > 1, then the assumed pressure is low i Wilson’s equation can be used to give an estimate of the bubble point pressure for equations of state application: 492 CHAPTER 5 Equations of State and Phase Equilibria X zi i pci Tci ¼1 exp 5:37ð1 + ωÞ 1 pb T Solving for the bubble point pressure gives pb X i Tci zi pci exp 5:37ð1 + ωÞ 1 T (5.35) EXAMPLE 5.5 A crude oil reservoir exits at 200°F with a composition and associated component properties as tabulated below: Component zi Pci Tci ωi C1 C2 C3 i-C4 n-C4 i-C5 n-C5 C6 C7+ 0.42 0.05 0.05 0.03 0.02 0.01 0.01 0.01 0.40 667.8 707.8 616.3 529.1 550.7 490.4 488.6 436.9 230.4 343.37 550.09 666.01 734.98 765.65 829.1 845.7 913.7 1279.8 0.0104 0.0986 0.1524 0.1848 0.2010 0.2223 0.2539 0.3007 1.0653 The heptanes-plus fraction is characterized by the following measured properties: ðMÞC7 + ¼ 216:0 ðγ ÞC7 + ¼ 0:8605 ðTb ÞC7 + ¼ 977°R Estimate the bubble point pressure. Solution Step 1. Calculate the critical pressure and temperature of the C7+ by the RiaziDaubert equation, Eq. (2.4), to give pc ¼ 3:12281 109 ð977Þ2:3125 ð0:8605Þ2:3201 ¼ 230:4psi Tc ¼ 24:27870ð977Þ0:58848 ð0:8605Þ0:3596 ¼ 1279:8°R Step 2. Calculate the acentric factor by employing the Edmister correlation [Eq. (2.21)] to yield ω¼ 3½ log ðpc =14:70Þ ¼ 1:0653 7½Tc =Tb 1 Vapor-Liquid Equilibrium Calculations 493 Step 3. Construct the following table: Component zi Pci Tci ωi C1 C2 C3 i-C4 n-C4 i-C5 n-C5 C6 C7+ 0.42 0.05 0.05 0.03 0.02 0.01 0.01 0.01 0.40 667.8 707.8 616.3 529.1 550.7 490.4 488.6 436.9 230.4 343.37 550.09 666.01 734.98 765.65 829.1 845.7 913.7 1279.8 0.0104 0.0986 0.1524 0.1848 0.2010 0.2223 0.2539 0.3007 1.0653 Step 4. Estimate the bubble point pressure from Eq. (5.35) as follows: X Tci zi pci exp 5:37ð1 + ωÞ 1 pb T i Component zi Pci Tci ωi Eq. (5.35) C1 C2 C3 i-C4 n-C4 i-C5 n-C5 C6 C P7+ 0.42 0.05 0.05 0.03 0.02 0.01 0.01 0.01 0.4 667.8 707.8 616.3 529.1 550.7 490.4 488.6 436.9 320.3 343.37 550.09 666.01 734.98 765.65 829.1 845.7 913.7 1279.8 0.0104 0.0986 0.1524 0.1848 0.201 0.2223 0.2539 0.3007 1.0633 4620.64266 133.639009 45.2146052 12.6894565 6.6436649 1.630693 1.34909342 0.58895645 0.01761403 4822.41575 The estimated bubble point pressure is 4822 psia; notice the impact of the methane on the estimated saturation pressure. Separator Calculations Produced reservoir fluids are complex mixtures of different physical characteristics. As a well stream flows from the high-temperature, high-pressure petroleum reservoir, it experiences pressure and temperature reductions. Gases evolve from the liquids and the well stream changes in character. The physical separation of these phases is by far the most common of all field-processing operations and one of the most critical. The manner in which the hydrocarbon phases are separated at the surface influences the stock-tank oil recovery. The principal means of surface separation of gas and oil is the conventional stage separation. 494 CHAPTER 5 Equations of State and Phase Equilibria Stage separation is a process in which gaseous and liquid hydrocarbons are flashed (separated) into vapor and liquid phases by two or more separators. These separators usually are operated in a series at consecutively lower pressures. Each condition of pressure and temperature at which hydrocarbon phases are flashed is called a stage of separation. Traditionally, the stock tank is considered a separate stage of separation. Example of two-stage separation processes are shown in Fig. 5.4. To describe the liquid-gas separation process, it is convenient to group the composition of a hydrocarbon mixture into three groups of components: n n n Volatile (light) components: The group includes nitrogen, methane, and ethane. Intermediate components: This group comprises components with intermediate volatility such as propane through hexane and CO2. Heavy components: This group contains components of less volatility such as heptane and heavier components. Mechanically, there are two types of gas-oil separation a. differential separation b. flash separation. n FIGURE 5.4 Schematic drawing of two-stage separation processes. Vapor-Liquid Equilibrium Calculations 495 Differential Separation Process In this process, the pressure is reduced in stages and the liberated gas is removed from contact with the oil as soon as the pressure is reduced. Initially, the liberated gas is composed mainly of lighter components with intermediate and heavy components remaining in the liquid. When the gas is separated in this manner, the maximum amount of heavy and intermediate components remain in the liquid resulting in a minimum shrinkage of the oil. A greater stock-tank oil recovery will be obtained from the differential liberation process due to the fact that the lighter components liberated earlier at higher pressures and is not present at lower pressures to attract the intermediate and heavy components and pull them into the gas phase. Flash Separation In this flash liberation process, the liberated gas remains in contact with the oil until its instantaneous removal at the final separation pressure. A maximum proportion of intermediate and heavy components are attracted into the gas phase by this process resulting in a maximum oil shrinkage and, therefore, lower oil recovery. It should be pointed out that the separation process of the solution gas taking place in the reservoir as the pressure drops below the bubble point pressure can be described experimentally by the gas differential liberation process. The main reason for this classification is that due to the high mobility of the liberated gas, the gas moves faster than the oil toward the producing well, a process similarly simulated in the laboratory through a differential liberation test by removing the liberated gas as soon as it forms. However, there is a brief period when the liberated gas saturation is less that its critical saturation; that is, gas is immobile and remains in contact with the oil, which can be described by flash liberation. The flow from the wellbore through surface facilities can be described by a series of flash separations. However, it is believed that the produced hydrocarbon system undergoes a flash liberation in the primary separator followed by a differential process as the actual separation occurs where gas and liquid are removed from contact to a subsequent stage of separations. As the number of stages increases, the differential aspect of the overall separation becomes greater. The purpose of stage separation then is to reduce the pressure on the produced oil in steps to maximize the stock-tank oil recovery results. Separator calculations are basically performed to determine 496 CHAPTER 5 Equations of State and Phase Equilibria a. Optimum separation conditions in terms of separator pressure and temperature. b. Composition of the separated gas and oil phases. c. Oil formation volume factor. d. Producing gas-oil ratio. e. API gravity of the stock-tank oil. Stage separation of oil and gas is accomplished with a series of separator stages operating at sequentially reduced pressures when the liquid is discharged from a higher pressure separator into a lower pressure separator. The objective is to select the optimum set of working separators pressures that yield the maximum liquid recovery of liquid at stock-tank conditions. The optimization problem is constrained because the pressure of the first stage (primary separator) must be less and also impacted by the wellhead pressure and the pressure loss occurring between the two nodes. Most of optimization is performed on the subsequent separation stages, particularly the second stage. The optimization procedure is based on and supported by the following field and phase behavior observations: n n n When the separator pressure is high, large amounts of light components remain in the liquid phase; however, they are liberated at the stock-tank conditions along with other valuable components (intermediate and heavy components) that attracted to the to light-end molecules. This process, when it occurs, will result in a large amount of the stock-tank oil in the gas phase at the stock tank. If the pressure is too low, large amounts of light components are separated from the liquid, and they attract substantial quantities of intermediate and heavier components resulting in a substantial stocktank oil recovery. An intermediate pressure, called the optimum separator pressure, should be selected to maximize the oil volume accumulation in the stock tank. Selecting the optimum pressure will yield 1. A maximum in the stock-tank API gravity. 2. A minimum in the oil formation volume factor (ie, less oil shrinkage). 3. A minimum in the producing gas-oil ratio (gas solubility). This separator pressure optimization can be performed by conducting a separator test. Typical results of the test are tabulated below with the concept of determining the optimum separator pressure is shown in Fig. 5.5 by graphically plotting the API gravity, Bo, and Rs as a function of pressure, which indicates an optimum separator pressure of 100 psig. Vapor-Liquid Equilibrium Calculations 497 Rs API Rs Bo Optimum separator pressure API First stage separator pressure n FIGURE 5.5 Effect of separator pressure on API, Bo, and GOR. Psep (psig) Tsep (°F) Rsfb °API Bofb 50 to 0 Total 100 to 0 Total 200 to 0 Total 300 to 0 Total 75 75 737 41 778 676 92 768 602 178 780 549 246 795 40.5 1.481 40.7 1.474 40.4 1.483 40.1 1.495 75 75 75 75 75 75 The computational steps of the “separator calculations” are described by the following steps and in conjunction with Fig. 5.6, which schematically shows a bubble point reservoir flowing into a surface separation unit consisting of n-stages operating at successively lower pressures. Step 1. Calculate the volume of oil occupied by 1 lb-mol of the crude oil with a given overall composition “zi” at the reservoir pressure 498 CHAPTER 5 Equations of State and Phase Equilibria Gas removed 1 Gas removed 2 (nv)1 yi (nv)2 yi xi (nL)1 zi xi (nL)3 Gas removed Gas removed 3 (nv)3 yi xi (nL)2 Separator 1 Separator 2 Stage 1 Stage 2 n (nL)st = (nL)i i =1 Separator 3 Stage 3 Stock-tank Well stream “zi” Pb 1 mol Vo, ft3 n FIGURE 5.6 Schematic illustration of n-separation stages. (or bubble point pressure) and temperature. This volume, denoted as Vo, is calculated by recalling and applying the equation that defines the number of moles; that is, n¼ m ρo Vo ¼ ¼1 Ma Ma (5.36) Solving for the oil volume gives X Vo ¼ Ma ¼ ρo ðzi M i Þ i ρo (5.37) where m ¼ total weight of 1 lb-mol of the crude oil, lb/mol Vo ¼ volume of 1 lb-mol of the crude oil at reservoir conditions, ft3/mol Ma ¼ apparent molecular weight Mi ¼ molecular weight of component i ρo ¼ density of the reservoir oil at the bubble point pressure, lb/ft3 It should be pointed out that the density can be calculated from the crude oil composition using the Standing-Katz or Alani-Kennedy method. Vapor-Liquid Equilibrium Calculations 499 Step 2. Given the composition of the hydrocarbon stream “zi” to the first separator and the operating conditions of the separator, calculate a set of equilibrium ratios that describes the hydrocarbon mixture. Step 3. Assuming a total of 1 mol of the feed (ie, nt ¼ 1) entering the first separator, perform flash calculations to obtain n compositions of the separated gas “yi” and liquid “xi” n number of moles of the gas “nv” and the liquid “nL”. Designating these moles as (nL)1 and (nυ)1, the actual number of moles of the gas and the liquid leaving the first separation stage are ½nυ1 a ¼ ðnt Þðnυ Þ1 ¼ ð1Þðnυ Þ1 ½nL1 a ¼ ðnt ÞðnL Þ1 ¼ ð1ÞðnL Þ1 where [nv1]a ¼ actual number of moles of vapor leaving the first separator [nL1]a ¼ actual number of moles of liquid leaving the first separator. Step 4. Using the composition of the liquid “xi” leaving the first separator as the feed for the second separator “zi ¼ xi” and calculate the equilibrium ratios of the hydrocarbon mixture at the prevailing pressure and temperature of the second-stage separator. Step 5. Assume 1 mol of the feed from first-stage separator to the second stage, perform flash calculations to determine the compositions and quantities of the gas and liquid leaving the second separation stage. The actual number of moles of the two phases then is calculated from ½nυ2 a ¼ ½nL1 a ðnυ Þ2 ¼ ð1ÞðnL Þ1 ðnυ Þ2 ½nL2 a ¼ ½nL1 a ðnL Þ2 ¼ ð1ÞðnL Þ1 ðnL Þ2 where [nv2]a, [nL2]a ¼ actual moles of gas and liquid leaving separator 2 (nv)2, (nL)2 ¼ moles of gas and liquid as determined from flash calculations Step 6. The previously outlined procedure is repeated for each separation stage, including the stock-tank storage, and the calculated moles and compositions are recorded. The total number of moles of liberated solution gas from all stages are calculated as ðnν Þt ¼ n X ðnva Þi ¼ ðnν Þ1 + ðnL Þ1 ðnν Þ2 + ðnL Þ1 ðnL Þ2 ðnν Þ3 + ⋯ + ðnL Þ1 …ðnL Þn1 ðnν Þn i¼1 500 CHAPTER 5 Equations of State and Phase Equilibria In a more compact form, this expression can be written as ðnν Þt ¼ ðnν Þ1 + " # n i1 Y X ðnν Þi ðnL Þj i¼2 (5.38) j¼1 where (nv)t ¼ total moles of liberated gas from all stages, lb-mol/mol of feed n ¼ number of separation stages. The total moles of liquid remaining in the stock tank can also be calculated as ðnL Þst ¼ nL1 nL2 …nLn or ðnL Þst ¼ n Y ðnL Þi (5.39) i¼1 where (nL)st ¼ number of moles of liquid remaining in the stock tank. Step 7. Calculate the volume, in scf, of all the liberated solution gas from Vg ¼ 379:4ðnυ Þt (5.40) where Vg ¼ total volume of the liberated solution gas scf/mol of feed. Step 8. Determine the volume of stock-tank oil occupied by (nL)st moles of liquid from ðVo Þst ¼ ðnL Þst ðMa Þst ðρo Þst (5.41) where (Vo)st ¼ volume of stock-tank oil, ft3/mol of feed (Ma)st ¼ apparent molecular weight of the stock-tank oil (ρo)st ¼ density of the stock-tank oil, lb/ft3. This desnsity is calculated from the oil composition at stock-tank Step 9. Calculate the specific gravity and the API gravity of the stock-tank oil by applying the expression γo ¼ ðρo Þst 62:4 (5.42) Vapor-Liquid Equilibrium Calculations 501 Step 10. Calculate the total gas-oil ratio (gas solubility) Rs: GOR ¼ Vg ð5:615Þð379:4Þðnv Þt ¼ ðVo Þst =5:615 ðnL Þst ðMÞst =ðρo Þst GOR ¼ 2130:331ðnv Þt ðρo Þst ðnL Þst ðMÞst (5.43) where GOR ¼ gas-oil ratio, scf/STB. Step 11. Calculate the oil formation volume factor from the relationship Bo ¼ Vo ðVo Þst Substituting Eqs. (5.36), (5.41) with equivalent term in the above expression gives Bo ¼ Ma ðρo Þst ρo ðnL Þst ðMa Þst (5.44) where Bo ¼ oil formation volume factor, bbl/STB Ma ¼ apparent molecular weight of the feed (Ma)st ¼ apparent molecular weight of the stock-tank oil ρo ¼ density of crude oil at reservoir conditions, lb/ft3 The separator pressure can be optimized by calculating the API gravity, GOR, and Bo in the manner just outlined at different assumed pressures. The optimum pressure corresponds to a maximum in the API gravity and a minimum in gas-oil ratio and oil formation volume factor. EXAMPLE 5.6 A crude oil, with the composition as follows, exists at its bubble point pressure of 1708.7 psia and a temperature of 131°F. The surface separation facilities include two separators and stock tank with the following operating conditions: Stage # Pressure (psia) Temperature (°F) 1 2 Stock tank 400 350 14.7 72 72 60 502 CHAPTER 5 Equations of State and Phase Equilibria The composition of the crude oil “feed” to the primary separator is tabulated below: Mi Component zi CO2 N2 C1 C2 C3 i-C4 n-C4 i-C5 n-C5 C6 C7+ 0.0008 0.0164 0.2840 0.0716 0.1048 0.0420 0.0420 0.0191 0.1912 0.0405 0.3597 44.0 28.0 164.0 30.0 44.0 584.0 58.0 72.0 72.0 86.0 252 Given that the molecular weight and specific gravity of C7+ are 252 and 0.8429, calculate n n n Oil formation volume factor at the bubble point pressure Bofb Gas solubility at the bubble point pressure Rsfb. Stock-tank density and the API gravity of the hydrocarbon system. Solution Step 1. Calculate the apparent molecular weight of the crude oil to give X Ma ¼ zi Mi ¼ 113:5102 Step 2. Calculate the oil density at bubble point pressure by using the Standing and Katz correlations to yield ρob ¼ 44:794 lb=ft3 Step 3. Using the pressure and temperature of the primary separator (ie, 400 psia and 72°F) calculate the K-value for each component by applying the Standing’s K-value, to give 1 Ki ¼ 10ða + cFi Þ p Component zi Ki CO2 N2 C1 C2 C3 i-C4 n-C4 i-C5 n-C5 C6 C7+ 0.0008 0.0164 0.2840 0.0716 0.1048 0.0420 0.0420 0.0191 0.0191 0.0405 0.3597 3.509 39.90 8.850 1.349 0.373 0.161 0.120 0.054 0.043 0.018 0.0021 Vapor-Liquid Equilibrium Calculations 503 Step 4. Perform flash calculations on the hydrocarbon stream to the first separator to give nL ¼ 0:7209 nv ¼ 0:29791 Step 5. Solve for xi and yi to give xi ¼ zi nL + nv Ki yi ¼ xi Ki Component zi Ki xi yi Mi CO2 N2 C1 C2 C3 i-C4 n-C4 i-C5 n-C5 C6 C7+ 0.0008 0.0164 0.2840 0.0716 0.1048 0.0420 0.0420 0.0191 0.0191 0.0405 0.3597 3.509 39.90 8.850 1.349 0.373 0.161 0.120 0.054 0.043 0.018 0.0021 0.0005 0.0014 0.089 0.0652 0.1270 0.0548 0.0557 0.0259 0.0261 0.0558 0.4986 0.0018 0.0552 0.7877 0.0880 0.0474 0.0088 0.0067 0.0014 0.0011 0.0010 0.0009 44.0 28.0 16.0 30.0 44.0 58.0 58.0 72.0 72.0 86.0 252 Step 6. Use the calculated liquid composition as the feed for the second separator or flash the composition at the operating condition of the separator. The results are shown in the table below, with nL ¼ 0.9851 and nυ ¼ 0.0149. Component zi Ki xi yi Mi CO2 N2 C1 C2 C3 i-C4 n-C4 i-C5 n-C5 C6 C7+ 0.0005 0.0014 0.089 0.0652 0.1270 0.0548 0.0557 0.0259 0.0261 0.0558 0.4986 3.944 46.18 10.06 1.499 0.4082 0.1744 0.1291 0.0581 0.0456 0.0194 0.00228 0.0005 0.0008 0.0786 0.0648 0.1282 0.0555 0.0564 0.0263 0.0264 0.0566 0.5061 0.0018 0.0382 0.7877 0.0971 0.0523 0.0097 0.0072 0.0015 0.0012 0.0011 0.0012 44.0 28.0 16.0 30.0 44.0 58.0 58.0 72.0 72.0 86.0 252 Step 7. Repeat the calculation for the stock-tank stage, to give the results in the following table, with nL ¼ 0.6837 and nυ ¼ 0.3163. 504 CHAPTER 5 Equations of State and Phase Equilibria Component zi Ki xi yi CO2 N2 C1 C2 C3 i-C4 n-C4 i-C5 n-C5 C6 C7+ 0.0005 0.0008 0.0784 0.0648 0.1282 0.0555 0.0564 0.0263 0.0264 0.0566 0.5061 81.14 1159 229 27.47 6.411 2.518 1.805 0.7504 0.573 0.2238 0.03613 0000 0000 0.0011 0.0069 0.0473 0.0375 0.0450 0.0286 0.02306 0.0750 0.7281 0.0014 0.026 0.2455 0.1898 0.3030 0.0945 0.0812 0.0214 0.0175 0.0168 0.0263 Step 8. Calculate the actual number of moles of the liquid phase at the stocktank conditions from Eq. (5.39): ðnL Þst ¼ n Y ðnL Þi i¼1 ðnL Þst ¼ ð1Þð0:7209Þð0:9851Þð0:6837Þ ¼ 0:48554 Step 9. Calculate the total number of moles of the liberated gas from the entire surface separation system: nυ ¼ 1 ðnL Þst ¼ 1 0:48554 ¼ 0:51446 Step 10. Calculate apparent molecular weight of the stock-tank oil from its composition, to give ðMa Þst ¼ ΣMi xi ¼ 200:6 Step 11. Using the composition of the stock-tank oil, calculate the density of the stock-tank oil by using Standing’s correlation, to give ðρo Þst ¼ 50:920 Step 12. Calculate the API gravity of the stock-tank oil from the specific gravity: γ ¼ ρo =62:4 ¼ 50:920=62:4 ¼ 0:81660°=60° API ¼ ð141:5=γ Þ 131:5 ¼ ð141:5=0:816Þ 131:5 ¼ 41:9 Step 13. Calculate the gas solubility from Eq. (5.43) to give Rs ¼ 2130:331ðnv Þt ðρo Þst ðnL Þst ðMÞst Rs ¼ 2130:331ð0:51446Þð50:92Þ ¼ 573:0 scf=STB 0:48554ð200:6Þ Step 14. Calculate Bo from Eq. (5.44) to give Equations of State 505 Bo ¼ Ma ðρo Þst ρo ðnL Þst ðMa Þst Bo ¼ ð113:5102Þð50:92Þ ¼ 1:325bbl=STB ð44:794Þð0:48554Þð200:6Þ To optimize the operating pressure of the separator, these steps should be repeated several times under different assumed pressures and the results, in terms of API, Bo, and Rs, should be expressed graphically and used to determine the optimum pressure. EQUATIONS OF STATE An equation of state (EOS) is an analytical expression relating the pressure, p, to the temperature, T, and the volume, V. A proper description of this PVT relationship for real hydrocarbon fluids is essential in determining the volumetric and phase behavior of petroleum reservoir fluids and predicting the performance of surface separation facilities; these can be described accurately by equations of state. In general, most equations of state require only the critical properties and acentric factor of individual components. The main advantage of using an EOS is that the same equation can be used to model the behavior of all phases, thereby assuring consistency when performing phase equilibria calculations. The best known and the simplest example of an equation of state is the ideal gas equation, expressed mathematically by the expression p¼ RT V (5.45) where V ¼ gas volume in ft3 per 1 mol of gas, ft3/mol. The PVT relationship is used to describe the volumetric behavior of hydrocarbon gases at pressures close to the atmospheric pressure for which it was experimentally derived. However, the extreme limitations of the applicability of Eq. (5.45) prompted numerous attempts to develop an equation of state suitable for describing the behavior of real fluids at extended ranges of pressures and temperatures. A review of recent developments and advances in the field of empirical cubic equations of state are presented next, along with samples of their applications in petroleum engineering. Four equations of state are discussed here: n n n n van der Waals Redlich-Kwong Soave-Redlich-Kwong Peng-Robinson 506 CHAPTER 5 Equations of State and Phase Equilibria The Van der Waals Equation of State “VdW EOS” In developing the ideal gas EOS, two assumptions were made: n n The volume of the gas molecules is insignificant compared to the total volume and distance between the molecules. There are no attractive or repulsive forces between the molecules. Van der Waals (1873) attempted to eliminate these two assumptions in developing an empirical equation of state for real gases. In his attempt to eliminate the first assumption, van der Waals pointed out that the gas molecules occupy a significant fraction of the volume at higher pressures and proposed that the volume of the molecules, denoted by the parameter b, be subtracted from the actual molar volume, V, in Eq. (5.45), to give p¼ RT V b where the parameter b is known as the covolume and considered to reflect the volume of molecules. The variable V represents the actual volume in ft3 per 1 mol of gas. To eliminate the second assumption, van der Waals subtracted a corrective term, denoted by a/V2, from this equation to account for the attractive forces between molecules. In a mathematical form, van der Waals proposed the following expression: p¼ RT a V b V2 (5.46) where p ¼ system pressure, psia T ¼ system temperature, °R R ¼ gas constant, 10.73 psi-ft3/lb-mol, °R V ¼ volume, ft3/mol a ¼ “attraction” parameter b ¼ “repulsion” parameter The two parameters, a and b, are constants characterizing the molecular properties of the individual components. The symbol a is considered a measure of the intermolecular attractive forces between the molecules. Eq. (5.46) shows the following important characteristics: 1. At low pressures, the volume of the gas phase is large in comparison with the volume of the molecules. The parameter b becomes negligible in comparison with V and the attractive forces term, a/V2, becomes insignificant; therefore, the van der Waals equation reduces to the ideal gas equation [Eq. (5.45)]. Equations of State 507 2. At high pressure (ie, p ! 1) the volume, V, becomes very small and approaches the value b, which is the actual molecular volume, as mathematically given by lim V ðpÞ ¼ b p!1 The van der Waals or any other equation of state can be expressed in a more generalized form as follows: p ¼ prepulsion pattraction (5.47) where the repulsion pressure term, prepulsion, is represented by the term RT/(V b), and the attraction pressure term, pattraction, is described by a/V2. In determining the values of the two constants, a and b, for any pure substance, van der Waals observed that the critical isotherm has a horizontal slope and an inflection point at the critical point, as shown in Fig. 5.7. This observation can be expressed mathematically as follows: @p ¼0 @V Tc, pc 2 @ p ¼0 @V 2 Tc, pc (5.48) Differentiating Eq. (5.46) with respect to the volume at the critical point results in @p RTc 2a ¼ + ¼0 @V Tc , pc ðVc bÞ2 Vc3 Liquid Gas H VdW EOS critical point condition [ Pressure (5.49) ∂p ∂2p ] T , p ,V = 0 & [ 2 ] T , p ,V = 0 c c c ∂V ∂V c c c T1 < T2 < T3 < Tc < T4 C Pc Liquid – Vapor G F B Vc A Volume V n FIGURE 5.7 An idealized pressure-volume relationship for a pure component. T4 Tc T3 T E 2 T1 508 CHAPTER 5 Equations of State and Phase Equilibria 2 @ p 2RTc 6a ¼ + ¼0 @V 2 Tc , pc ðVc bÞ3 Vc4 (5.50) Solving Eqs. (5.49), (5.50) simultaneously for the parameters a and b gives 1 Vc 3 8 a¼ RTc Vc 9 b¼ (5.51) (5.52) Eq. (5.51) suggests that the covolume “b” is approximately 0.333 of the critical volume “Vc” of the substance. Numerous experimental studies revealed that the co-volume “b” is in the range 0.24–0.28 of the critical volume and pure component. Applying Eq. (5.48) to the critical point (i.e., by setting T ¼ Tc, p ¼ pc, and V ¼ Vc) and combining with Eqs. (5.51), (5.52) yields pc Vc ¼ ð0:375ÞRTc (5.53) Eq. (5.53) indicates that regardless of the type of the substance, the Van der Waals EOS gives a universal critical gas compressibility factor “Zc” of 0.375. Experimental studies show that Zc values for substances are ranging between 0.23 and 0.31. Eq. (5.53) can be combined with Eqs. (5.51), (5.52) to give more convenient and traditional expressions for calculating the parameters a and b, to yield a ¼ Ωa R2 Tc2 pc (5.54) RTc pc (5.55) b ¼ Ωb where R ¼ gas constant, 10.73 psia-ft3/lb-mol-°R pc ¼ critical pressure, psia Tc ¼ critical temperature, °R Ωa ¼ 0.421875 Ωb ¼ 0.125 Eq. (5.46) can also be expressed in a cubic form in terms of the volume “V” as follows: p¼ RT a 2 V b V Equations of State 509 Rearranging gives RT 2 a ab V b+ V + V ¼0 p p p 3 (5.56) Eq. (5.56) is referred to as the van der Waals two-parameter cubic equation of state. The term two-parameter refers to parameters a and b. The term cubic equation of state implies that the equation have three possible values for the volume “V,” at least one of which is real. Perhaps the most significant features of Eq. (5.56) are that it describes the liquid-condensation phenomenon and the passage from the gas to the liquid phase as the gas is compressed. These important features of the van der Waals EOS are discussed next in conjunction with Fig. 5.8. Consider a pure substance with a p-V behavior as shown in Fig. 5.8. Assume that the substance is kept at a constant temperature, T, below its critical temperature. At this temperature, Eq. (5.56) has three real roots (volumes) for each specified pressure, p. A typical solution of Eq. (5.56) at constant temperature, T, is shown graphically by the dashed isotherm, the constant temperature curve DWEZB, in Fig. 5.8. The three values of V are the intersections B, E, and D on the horizontal line, corresponding to a fixed value of the pressure. This dashed calculated line (DWEZB) then appears to give a continuous transition from the gaseous phase to the liquid phase, but in reality, the transition is abrupt and discontinuous, with both liquid and vapor existing along the straight horizontal line DB. Examining the graphical solution of Eq. (5.56) shows that the largest root (volume), indicated by point D, corresponds to the volume of the saturated vapor, while the smallest positive volume, indicated by point B, corresponds to the volume of the C Pressure Pure component isotherm as calculated from VdW EOS W D P B E Z Volume n FIGURE 5.8 Pressure-volume diagram for a pure component. T 510 CHAPTER 5 Equations of State and Phase Equilibria saturated liquid. The third root, point E, has no physical meaning. Note that these values become identical as the temperature approaches the critical temperature, Tc, of the substance. Eq. (5.56) can be expressed in a more practical form in terms of the compressibility factor, Z. Replacing the molar volume, V, in Eq. (5.56) with ZRT/p gives RT 2 a ab V3 ¼ b + V + V ¼0 p p p 2 RT ZRT a ZRT ab + ¼0 V3 b + p p p p p or Z 3 ð1 + BÞZ2 + AZ AB ¼ 0 (5.57) where A¼ ap R2 T 2 (5.58) bp RT (5.59) B¼ Z ¼ compressibility factor p ¼ system pressure, psia T ¼ system temperature, °R Eq. (5.57) yields one real root1 in the one-phase region and three real roots in the two-phase region (where system pressure equals the vapor pressure of the substance). In the two-phase region, the largest positive root corresponds to the compressibility factor of the vapor phase, Zv, while the smallest positive root corresponds to that of the liquid phase, ZL. An important practical application of Eq. (5.57) is density calculations, as illustrated in the following example. EXAMPLE 5.7 A pure propane is held in a closed container at 100°F. Both gas and liquid are present. Calculate, using the van der Waals EOS, the density of the gas and liquid phases. 1 In some super critical regions, Eq. (5.57) can yield three real roots for Z. From the three real roots, the largest root is the value of the compressibility with physical meaning. Equations of State 511 Solution Step 1. Determine the vapor pressure, pv, of the propane from the Cox chart. This is the only pressure at which two phases can exist at the specified temperature: pv ¼ 185 psi Step 2. Calculate parameters a and b from Eqs. (5.54), (5.55), respectively: a ¼ Ωa R2 Tc2 ð10:73Þ2 ð666Þ2 ¼ 0:421875 ¼ 34,957:4 pc 616:3 and b ¼ Ωb RTc 10:73ð666Þ ¼ 0:125 ¼ 1:4494 616:3 pc Step 3. Compute the coefficients A and B by applying Eqs. (5.58), (5.59), respectively: A¼ ap ð34, 957:4Þð185Þ ¼ ¼ 0:179122 R2 T 2 ð10:73Þ2 ð560Þ2 B¼ bp ð1:4494Þð185Þ ¼ ¼ 0:044625 RT ð10:73Þð560Þ Step 4. Substitute the values of A and B into Eq. (5.57), to give Z 3 ð1 + BÞZ2 + AZ AB ¼ 0 Z 3 1:044625Z2 + 0:179122Z 0:007993 ¼ 0 Step 5. Solve the above third-degree polynomial by extracting the largest and smallest roots of the polynomial using the appropriate direct or iterative method, to give Zv ¼ 0:72365 ZL ¼ 0:07534 Step 6. Solve for the density of the gas and liquid phases using Eq. (2.17): ρg ¼ pM ð185Þð44:0Þ ¼ ¼ 1:87 lb=ft3 Z v RT ð0:72365Þð10:73Þð560Þ and ρL ¼ pM ð185Þð44Þ ¼ ¼ 17:98 lb=ft3 ZL RT ð0:07534Þð10:73Þð560Þ The van der Waals equation of state, despite its simplicity, provides a correct description, at least qualitatively, of the PVT behavior of substances in the liquid and gaseous states. Yet, it is not accurate enough to be suitable for design purposes. With the rapid development of computers, the equation-of-state approach for calculating physical properties and phase equilibria proved to be a powerful tool, and much energy was devoted to the development of new and accurate equations 512 CHAPTER 5 Equations of State and Phase Equilibria of state. Many of these equations are a modification of the van der Waals EOS and range in complexity from simple expressions containing 2 or 3 parameters to complicated forms containing more than 50 parameters. Although the complexity of any equation of state presents no computational problem, most authors prefer to retain the simplicity found in the van der Waals cubic equation while improving its accuracy through modifications. All equations of state generally are developed for pure fluids first and then extended to mixtures through the use of mixing rules. Redlich-Kwong’s Equation of State “RK EOS” Redlich and Kwong (1949) observed that the van der Waals a/V2 term does not contain the system temperature to account for its impact on the intermolecular attractive force between the molecules. Redlich and Kwong demonstrated that by a simple adjustment of the van der Waals’ a/V2 term to explicitly include the system temperature could considerably improve the prediction of the volumetric and physical properties of the vapor phase. Redlich and Kwong replaced the attraction pressure term with a generalized temperature dependence term, as given in the following form: p¼ RT a pffiffiffi V b V ðV + bÞ T (5.60) in which T is the system temperature, in °R. Redlich and Kwong noted that, as the system pressure becomes very large (ie, p ! 1) the molar volume, V, of the substance shrinks to about 26% of its critical volume, Vc, regardless of the system temperature. Accordingly, they constructed Eq. (5.60) to satisfy the following condition: b ¼ 0:26Vc (5.61) Imposing the critical point conditions (as expressed by Eq. (5.48)) on Eq. (5.60) and solving the resulting two equations simultaneously, gives a ¼ Ωa R2 Tc2:5 pc (5.62) RTc pc (5.63) b ¼ Ωb where Ωa ¼ 0.42747 Ωb ¼ 0.08664. Equations of State 513 It should be noted that by equating Eq. (5.63) with (5.61) gives 0:26Vc ¼ Ωb RTc pc Arranging the above expression: pc Vc ¼ 0:33RTc (5.64) Eq. (5.64) shows that the RK EOS produces a universal critical compressibility factor (Zc) of 0.333 for all substances. As indicated earlier, the critical gas compressibility ranges from 0.23 to 0.31 for most of the substances, with an average value of 0.27. Replacing the molar volume, V, in Eq. (5.60) with ZRT/p and rearranging gives Z3 Z2 + A B B2 Z AB ¼ 0 (5.65) where A¼ ap R2 T 2:5 (5.66) bp RT (5.67) B¼ As in the van der Waals EOS, Eq. (5.65) yields one real root in the one-phase region (gas-phase region or liquid-phase region) and three real roots in the two-phase region. In the latter case, the largest root corresponds to the compressibility factor of the gas phase, Zv, while the smallest positive root corresponds to that of the liquid, ZL. EXAMPLE 5.8 Rework Example 5.7 using the Redlich-Kwong equation of state. Solution Step 1. Calculate the parameters a, b, A, and B. a ¼ Ωa R2 Tc2:5 ð10:73Þ2 ð666Þ2:5 ¼ 0:4247 ¼ 914, 1101:1 pc 616:3 b ¼ Ωb RTc ð10:73Þð666Þ ¼ 0:08664 ¼ 1:0046 pc 616:3 A¼ ð914, 110:1Þð185Þ ap ¼ ¼ 0:197925 R2 T 2:5 ð10:73Þ2 ð560Þ2:5 B¼ bp ð1:0046Þð185Þ ¼ ¼ 0:03093 RT ð10:73Þð560Þ Step 2. Substitute the parameters A and B into Eq. (5.65) and extract the largest and the smallest roots, to give 514 CHAPTER 5 Equations of State and Phase Equilibria Z3 Z2 + A B B2 Z AB ¼ 0 Z 3 Z2 + 0:1660384Z 0:0061218 ¼ 0 Largest root : Z v ¼ 0:802641 Smallest root : Z L ¼ 0:0527377 Step 3. Solve for the density of the liquid phase and gas phase: ρL ¼ pM ð185Þð44Þ ¼ ¼ 25:7 lb=ft3 ZL RT ð0:0527377Þð10:73Þð560Þ ρv ¼ pM ð185Þð44Þ ¼ ¼ 1:688 lb=ft3 Zv RT ð0:0802641Þð10:73Þð560Þ Redlich and Kwong extended the application of their equation to hydrocarbon liquid and gas mixtures by employing the following mixing rules: " #2 n X pffiffiffiffi am ¼ xi ai (5.68) i¼1 bm ¼ n X ½xi bi (5.69) i¼1 where n ¼ number of components in the mixture ai ¼ Redlich-Kwong a parameter for the ith component as given by Eq. (5.62) bi ¼ Redlich-Kwong b parameter for the ith component as given by Eq. (5.63) am ¼ parameter a for mixture bm ¼ parameter b for mixture xi ¼ mole fraction of component i in the liquid phase To calculate am and bm for a hydrocarbon gas mixture with a composition of yi, use Eqs. (5.68), (5.69) and replace xi with yi: " #2 n X pffiffiffiffi am ¼ yi ai i¼1 n X bm ¼ ½yi bi i¼1 Eq. (5.65) gives the compressibility factor of the gas phase or the liquid with the coefficients A and B as defined by Eqs. (5.66), (5.67) as A¼ am p R2 T 2:5 B¼ bm p RT The application of the Redlich and Kwong equation of state for hydrocarbon mixtures can be best illustrated through the following two examples (Examples 5.9 and 5.10). Equations of State 515 EXAMPLE 5.9 Calculate the density of a crude oil with the composition at 4000 psia and 160°F given in the table below. Use the Redlich-Kwong EOS. Component xi M pc Tc C1 C2 C3 n-C4 n-C5 C6 C7+ 0.45 0.05 0.05 0.03 0.01 0.01 0.40 16.043 30.070 44.097 58.123 72.150 84.00 215 666.4 706.5 616.0 527.9 488.6 453 285 343.33 549.92 666.06 765.62 845.8 923 1287 Solution Step 1. Determine the parameters ai and bi for each component using Eqs. (5.62), (5.63): ai ¼ 0:47274 R2 Tc2:5 pc bi ¼ 0:08664 RTc pc The results are shown in the table below. Component xi C1 C2 C3 n-C4 n-C5 C6 C7+ M pc 0.45 16.043 0.05 30.070 0.05 44.097 0.03 58.123 0.01 72.150 0.01 84.00 0.40 215 Tc 666.4 343.33 706.5 549.92 616.0 666.06 527.9 765.62 488.6 845.8 453 923 285 1287 ai bi 161,044.3 0.4780514 493,582.7 0.7225732 914,314.8 1.004725 1,449,929 1.292629 2,095,431 1.609242 2,845,191 1.945712 1.022348 (107) 4.191958 Step 2. Calculate the mixture parameters, am and bm, from Eqs. (5.68), (5.69), to give " #2 n X pffiffiffiffi am ¼ xi ai ¼ 2,591, 967 i¼1 and bm ¼ n X ½xi bi ¼ 2:0526 i¼1 Step 3. Compute the coefficients A and B by using Eqs. (5.66), (5.67), to produce 516 CHAPTER 5 Equations of State and Phase Equilibria A¼ am p 2,591, 967ð4000Þ ¼ ¼ 9:406539 R2 T 2:5 10:732 ð620Þ2:5 B¼ bm p 2:0526ð4000Þ ¼ ¼ 1:234049 RT 10:73ð620Þ Step 4. Solve Eq. (5.65) for the largest positive root, to yield Z3 Z2 + 6:934845Z 11:60813 ¼ 0 ZL ¼ 1:548126 Step 5. Calculate the apparent molecular weight of the crude oil: X xi Mi ¼ 110:2547 Ma ¼ Step 6. Solve for the density of the crude oil: ρL ¼ pMa ð4000Þð100:2547Þ ¼ ¼ 38:93 lb=ft3 ZL RT ð10:73Þð620Þð1:548120Þ Note that calculating the liquid density as by Standing’s method gives a value of 46.23 lb/ft3. EXAMPLE 5.10 Using the the RK EOS, calculate the density of a gas phase with the composition at 4000 psia and 160°F shown in the following table. Component yi M pc Tc C1 C2 C3 C4 C5 C6 C7+ 0.86 0.05 0.05 0.02 0.01 0.005 0.005 16.043 30.070 44.097 58.123 72.150 84.00 215 666.4 706.5 616.0 527.9 488.6 453 285 343.33 549.92 666.06 765.62 845.8 923 1287 Solution Step 1. Determine the parameters ai and bi for each component using Eqs. (5.62), (5.63): ai ¼ 0:47274 R2 Tc2:5 pc bi ¼ 0:08664 RTc pc The results are shown in the table below. Equations of State 517 Component yi C1 C2 C3 C4 C5 C6 C7+ M pc 0.86 16.043 0.05 30.070 0.05 44.097 0.02 58.123 0.01 72.150 0.005 84.00 0.005 215 Tc 666.4 343.33 706.5 549.92 616.0 666.06 527.9 765.62 488.6 845.8 453 923 285 1287 ai bi 161,044.3 0.4780514 493,582.7 0.7225732 914,314.8 1.004725 1,449,929 1.292629 2,095,431 1.609242 2,845,191 1.945712 1.022348(107) 4.191958 Step 2. Calculate am and bm using Eqs. (5.68), (5.69), to give " #2 n X pffiffiffiffi am ¼ yi ai ¼ 241,118 i¼1 X bi xi ¼ 0:5701225 bm ¼ Step 3. Calculate the coefficients A and B by applying Eqs. (5.66), (5.67), to yield A¼ am p 241,118ð4000Þ ¼ ¼ 0:8750 R2 T 2:5 10:733 ð620Þ2:5 B¼ bm p 0:5701225ð4000Þ ¼ ¼ 0:3428 RT 10:73ð620Þ Step 4. Solve Eq. (5.65) for Zv, to give Z 3 Z 2 + 0:414688Z 0:29995 ¼ 0 Z v ¼ 0:907 Step 5. Calculate the apparent molecular weight and density of the gas mixture: P Ma ¼ yi Mi ¼ 20:89 ρv ¼ pMa ð4000Þð20:89Þ ¼ ¼ 13:85 lb=ft3 Zv RT ð10:73Þð620Þð0:907Þ Soave-Redlich-Kwong Equation of State “SRK EOS” One of the most significant milestones in the development of cubic equations of state is accredited to Soave (1972) with his proposed modification of the RK EOS attraction pressure term. Soave replaced the explicit temperature term (a/T0.5) in Eq. (5.60) with a more generalized temperaturedependent term, denoted by [a α (T)]; that is, p¼ RT aαðT Þ V b V ðV + bÞ (5.70) 518 CHAPTER 5 Equations of State and Phase Equilibria where α(T) is a dimensionless parameter that is defined by the following expression: h pffiffiffiffiffiffiffiffiffiffi i2 αðT Þ ¼ 1 + m 1 T=Tc Or equivalently in terms of reduced temperature “Tr”: pffiffiffiffiffi 2 α ðT Þ ¼ 1 + m 1 T r (5.71) Soave imposed the following conditions on the a dimensionless parameter α(T): n n α(T) ¼ 1 when the reduced temperature Tr, ¼ 1; that is, α(T ¼ Tc) ¼ 1 when T/Tc ¼ 1 At temperatures other than the critical temperature, Soave regressed on the vapor pressure of pure components to develop a temperature correction parameter “m” that correlated with the acentric factor “ω” by the following relationship: m ¼ 0:480 + 1:74ω 0:176ω2 (5.72) where Tr ¼ reduced temperature, T/Tc ω ¼ acentric factor of the substance T ¼ system temperature, °R For simplicity and convenience, the temperature-dependent term “α(T)” is replaced by the symbol “α” for the remainder of this chapter. For any pure component, the constants “a” and “b” in Eq. (5.70) are found by imposing the classical van der Waals’s critical point constraints on Eq. (5.70) and solving the resulting two equations, to give a ¼ Ωa R2 Tc2 pc (5.73) RTc pc (5.74) b ¼ Ωb where Ωa ¼ 0:42747 Ωb ¼ 0:08664 Edmister and Lee (1986) pointed out that the following constrain satisfies the critical isotherm as given by ðV Vc Þ3 ¼ V 3 ½3Vc V 2 + 3Vc2 V Vc3 ¼ 0 (5.75) Equations of State 519 Eq. (5.70) also can be expressed into a cubic form, to give V3 RT 2 aα bRT ðaαÞb V + b2 V ¼0 p p p p (5.76) At the critical point, Eqs. (5.75), (5.76) are identical with α ¼ 1. Equating the corresponding coefficients in both equations gives RTc pc (5.77) a bRTc b2 pc pc (5.78) ab pc (5.79) 3Vc ¼ 3Vc2 ¼ Vc3 ¼ Solving these equations simultaneously for parameters a and b will give relationships identical to those given by Eqs. (5.73), (5.74), in addition: n rearranging Eq. (5.77) gives 1 pc Vc ¼ RTc 3 which indicates that the SRK EOS gives a universal critical gas compressibility factor “Zc” of 0.333. n combining Eq. (5.74) with (5.77), gives the expected value of the covolume of 26% of the critical volume; that is, b ¼ 0:26Vc The compressibility factor “Z” can be introduced into Eq. (5.76) by replacing the molar volume “V” with (ZRT/p) and rearranging to give Z3 Z2 + A B B2 Z AB ¼ 0 (5.80) with A¼ ðaαÞp ðRT Þ2 B¼ bp RT (5.81) (5.82) 520 CHAPTER 5 Equations of State and Phase Equilibria where p ¼ system pressure, psia T ¼ system temperature, °R R ¼ 10.730 psia ft3/lb-mol °R EXAMPLE 5.11 Rework Example 5.7 and solve for the density of the two phases using the SRK EOS. Solution Step 1. Determine the critical pressure, critical temperature, and acentric factor from Table 1.1 of Chapter 1, to give Tc ¼ 666:01°R pc ¼ 616:3 psia ω ¼ 0:1524 Step 2. Calculate the reduced temperature: Tr ¼ T=Tc ¼ 560=666:01 ¼ 0:8408 Step 3. Calculate the parameter m by applying Eq. (5.72), to yield m ¼ 0:480 + 1:574ω 0:176ω2 m ¼ 0:480 + 1:574ð0:1524Þ 0:176ð1:524Þ2 Step 4. Solve for parameter α using Eq. (5.71), to give pffiffiffiffiffi 2 α ¼ 1 + m 1 Tr pffiffiffiffiffiffiffiffiffiffiffiffiffiffi 2 α ¼ 1 + 0:7052 1 0:8408 ¼ 1:120518 Step 5. Compute the coefficients a and b by applying Eqs. (5.73), (5.74), to yield α ¼ Ωa R2 Tc2 10:732 ð666:01Þ2 ¼ 35, 427:6 ¼ 0:42747 pc 616:3 α ¼ Ωb RTc 10:73ð666:01Þ ¼ 0:08664 ¼ 1:00471 616:3 pc Step 6. Calculate the coefficients A and B from Eqs. (5.81), (5.82) to produce A¼ ðaαÞp ð35, 427:6Þð1:120518Þ185 ¼ ¼ 0:203365 R2 T 2 10:732 ð560Þ2 B¼ bp ð1:00471Þð185Þ ¼ ¼ 0:034658 RT ð10:73Þð560Þ Step 7. Solve Eq. (5.80) for ZL and Zv: Z 3 Z2 + A B B2 Z + AB ¼ 0 Z 3 Z2 + 0:203365 0:034658 0:0346582 Z + ð0:203365Þð0:034658Þ ¼ 0 Solving this third-degree polynomial gives ZL ¼ 0:06729 Zv ¼ 0:80212 Equations of State 521 Step 8. Calculate the gas and liquid density, to give ρv ¼ pM ð185Þð44:0Þ ¼ 1:6887 lb=ft3 ¼ Z v RT ð0:802121Þð10:73Þð560Þ ρL ¼ pM ð185Þð44:0Þ ¼ ¼ 20:13 lb=ft3 L Z RT ð0:06729Þð10:73Þð560Þ The Mixture Mixing Rule To use Eq. (5.80) with mixtures, mixing rules are required to determine the terms (aα) and b for the mixtures. Soave adopted the following mixing rules: ðaαÞm ¼ XX i pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi xi xj ai aj αi αj 1 kij (5.83) X ½xi bi (5.84) j bm ¼ i with A¼ ðaαÞm p ðRT Þ2 (5.85) and B¼ bm p RT (5.86) The parameter kij is an empirically determined correction factor called the binary interaction coefficient (BIC), which is included to characterize any binary system formed by components i and j in the hydrocarbon mixture. These binary interaction coefficients are used to model the intermolecular interaction through empirical adjustment of the (aα)m term as represented mathematically by Eq. (5.83). They are dependent on the difference in molecular size of components in a binary system and they are characterized by the following properties: n The interaction between hydrocarbon components increases as the relative difference between their molecular weights increases: ki, j + 1 > ki, j n Hydrocarbon components with the same molecular weight have a binary interaction coefficient of zero: ki, j ¼ 0 522 CHAPTER 5 Equations of State and Phase Equilibria n The binary interaction coefficient matrix is symmetric: ki, j ¼ kj, i Slot-Petersen (1987) and Vidal and Daubert (1978) presented theoretical background to the meaning of the interaction coefficient and techniques for determining their values. Graboski and Daubert (1978) and Soave (1972) suggested that no binary interaction coefficients are required for hydrocarbon-hydrocarbon pairs (eg, kij ¼ 0), except between C1 and C7+. Whiston and Brule (2000) proposed the following expression for calculating the binary interaction coefficient between methane and the heavy fractions: kC1 C7 + ¼ 0:18 h 16:668 vci 1:1311 + ðvci Þ1=3 i6 where vci is the heavy fraction critical volume, which can be estimated from vci ¼ 0:4804 + 0:06011 Mi + 0:00001076 ðMi Þ2 where vci ¼ critical volume of the C7+, or its lumped component, ft3/lbm Mi ¼ molecular weight If nonhydrocarbons are present in the mixture, including binary interaction parameters can greatly improve the volumetric and phase behavior predictions of the mixture by the SRK EOS. Reid et al. (1987) proposed the following tabulated binary interaction coefficients for applications with the SRK EOS: Component N2 CO2 H2S N2 CO2 H2S C1 C2 C3 i-C4 n-C4 i-C5 n-C5 C6 C7 + 0 0 0.12 0.02 0.06 0.08 0.08 0.08 0.08 0.08 0.08 0.08 0 0 0.12 0.12 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0 0 0 0.08 0.07 0.07 0.06 0.06 0.06 0.06 0.05 0.03 In applying Eq. (5.75) for the compressibility factor of the liquid phase “ZL” the composition of the liquid “xi” is used to calculate the coefficients A and B Equations of State 523 of Eqs. (5.85), (5.86) through the use of the mixing rules as described by Eqs. (5.83), (5.84). For determining the compressibility factor of the gas phase, Zv, the previously outlined procedure is used with composition of the gas phase, yi, replacing xi. EXAMPLE 5.12 A two-phase hydrocarbon system exists in equilibrium at 4000 psia and 160°F. The system has the composition shown in the following table. Component xi yi pc Tc ωi C1 C2 C3 C4 C5 C6 C7+ 0.45 0.05 0.05 0.03 0.01 0.01 0.40 0.86 0.05 0.05 0.02 0.01 0.005 0.0005 666.4 706.5 616.0 527.9 488.6 453 285 343.33 549.92 666.06 765.62 845.8 923 1160 0.0104 0.0979 0.1522 0.1852 0.2280 0.2500 0.5200 The heptanes-plus fraction has the following properties: m ¼ 215 Pc ¼ 285 psia Tc ¼ 700°F ω ¼ 0:52 Assuming kij ¼ 0, calculate the density of each phase by using the SRK EOS. Solution Step 1. Calculate the parameters α, a, and b by applying Eqs. (5.71), (5.73), (5.74): pffiffiffiffiffi 2 α ¼ 1 + m 1 Tr R2 Tc2 pc RTc b ¼ 0:08664 pc a ¼ 0:42747 The results are shown in the table below. Component pc Tc ωi αi ai bi C1 C2 C3 C4 C5 C6 C7+ 666.4 706.5 616.0 527.9 488.6 453 285 343.33 549.92 666.06 765.62 845.8 923 1160 0.0104 0.0979 0.1522 0.1852 0.2280 0.2500 0.5200 0.6869 0.9248 1.0502 1.1616 1.2639 1.3547 1.7859 8689.3 21,040.8 35,422.1 52,390.3 72,041.7 94,108.4 232,367.9 0.4780 0.7225 1.0046 1.2925 1.6091 1.9455 3.7838 524 CHAPTER 5 Equations of State and Phase Equilibria Step 2. Calculate the mixture parameters (aα)m and bm for the gas phase and liquid phase by applying Eqs. (5.83), (5.84), to give the following. For the gas phase, using yi: XX pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi yi yj ai aj αi αj 1 kij ¼ 9219:3 ðaαÞm ¼ X i j bm ¼ ½yi bi ¼ 0:5680 i For the liquid phase, using xi: XX pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi xi xj ai aj αi αj 1 kij ¼ 104,362:9 ðaαÞm ¼ X i j bm ¼ ½xi bi ¼ 1:8893 i Step 3. Calculate the coefficients A and B for each phase by applying Eqs. (5.85), (5.86), to yield the following. For the gas phase, A¼ ðaαÞm p ð9219:3Þð4000Þ ¼ ¼ 0:8332 R2 T 2 ð10:73Þ2 ð620Þ2 B¼ bm p ð0:5680Þð4000Þ ¼ ¼ 0:3415 RT ð10:73Þð620Þ For the liquid phase, A¼ ðaαÞm p ð104, 362:9Þð4000Þ ¼ ¼ 9:4324 R2 T 2 ð10:73Þ2 ð620Þ2 B¼ bm p ð1:8893Þð4000Þ ¼ ¼ 1:136 RT ð10:73Þð620Þ Step 4. Solve Eq. (5.80) for the compressibility factor of the gas phase, to produce Z3 Z2 + A B B2 Z + AB ¼ 0 Z 3 Z2 + ð0:8332 0:3415 0:34152ÞZ + ð0:8332Þð0:3415Þ ¼ 0 Solving this polynomial for the largest root gives Zv ¼ 0:9267 Step 5. Solve Eq. (5.80) for the compressibility factor of the liquid phase, to produce Z3 Z 2 + A B B2 Z + AB ¼ 0 Z 3 Z 2 + ð9:4324 1:136 1:1362ÞZ + ð9:4324Þð1:136Þ ¼ 0 Solving the polynomial for the smallest root gives ZL ¼ 1:4121 Step 6. Calculate the apparent molecular weight of the gas phase and liquid phase from their composition, to yield the following: Equations of State 525 For the gas phase: Ma ¼ X For the liquid phase: Ma ¼ yi Mi ¼ 20:89 X xi Mi ¼ 100:25 Step 7. Calculate the density of each phase. ρ¼ pMa RTZ For the gas phase: ρv ¼ ð4000Þð20:89Þ ¼ 13:556 lb=ft3 ð10:73Þð620Þð0:9267Þ For the liquid phase: ρL ¼ ð4000Þð100:25Þ ¼ 42:68 lb=ft3 ð10:73Þð620Þð1:4121Þ It is appropriate at this time to introduce and define the concept of the fugacity and the fugacity coefficient of the component. The fugacity “f” is a measure of the molar Gibbs energy of a real gas. It is evident from the definition that the fugacity has the units of pressure; in fact, the fugacity may be looked on as a vapor pressure modified to correctly represent the tendency of the molecules from one phase to escape into the other. In a mathematical form, the fugacity of a pure component is defined by the following expression: f o ¼ p exp Z p Z1 dp p o (5.87) where fo ¼ fugacity of a pure component, psia p ¼ pressure, psia Z ¼ compressibility factor The ratio of the fugacity to the pressure, f/p, is called the fugacity coefficient “Φ”. This dimensionless property is calculated from Eq. (5.87) as fo ¼ Φ ¼ exp p Z p Z1 dp p o Soave applied this generalized thermodynamic relationship to Eq. (5.70) to determine the fugacity coefficient of a pure component, to give o f A Z+B ¼ ln ðΦÞ ¼ Z 1 ln ðZ BÞ ln ln p B Z (5.88) For a hydrocarbon multicomponent mixture that exits in the two-phase region, the component fugacity in each phase is introduced to develop a criterion for thermodynamic equilibrium. Physically, the fugacity of a 526 CHAPTER 5 Equations of State and Phase Equilibria component i in one phase with respect to the fugacity of the component in a second phase is a measure of the potential for transfer of the component between phases. The phase with the lower component fugacity accepts the component from the phase with a higher component fugacity. Equal fugacities of a component in the two phases results in a zero net transfer. A zero transfer for all components implies that the hydrocarbon system is in thermodynamic equilibrium. Therefore, the condition of the thermodynamic equilibrium can be expressed mathematically by fiv ¼ fiL , 1 i n (5.89) where f vi ¼ fugacity of component i in the gas phase, psi f Li ¼ fugacity of component i in the liquid phase, psi n ¼ number of components in the system The fugacity coefficient of component i in a hydrocarbon liquid mixture or hydrocarbon gas mixture is a function of the system pressure, mole fraction, and fugacity of the component. The fugacity coefficient is defined as fv For a component i in the gas phase: Φvi ¼ i yi p (5.90) fL For a component i in the liquid phase: ΦLi ¼ i xi p (5.91) where Φvi ¼ fugacity coefficient of component i in the vapor phase ΦLi ¼ fugacity coefficient of component i in the liquid phase. Equilibrium Ratio “Ki” The equilibrium ratio is defined mathematically by Eq. (5.1) as Ki ¼ yi =xi When two phases exist in equilibrium, it suggests that fiL ¼ fiv , which leads to redefining the K-value in terms of the fugacity of components as Ki ¼ fiL =ðxi pÞ ΦLi ¼ Φvi fiv ðyi pÞ (5.92) Reid et al. (1987) defined the fugacity coefficient of component i in a hydrocarbon mixture by the following generalized thermodynamic relationship: 1 lnðΦi Þ ¼ RT Z 1 V @p RT @ni V dV ln ðZ Þ (5.93) Equations of State 527 where V ¼ total volume of n moles of the mixture ni ¼ number of moles of component i Z ¼ compressibility factor of the hydrocarbon mixture By combining the above thermodynamic definition of the fugacity with the SRK EOS [Eq. (5.70)], Soave proposed the following expression for the fugacity coefficient of component i in the liquid phase: ln L bi ðZL 1Þ fi A 2Ψ i bi B ¼ ln ΦLi ln 1 + L ln ZL B xi p bm B ðaαÞm bm Z (5.94) where Ψi ¼ X pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi xj ai aj αi αj 1 kij (5.95) j ðaαÞm ¼ XX i pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi xi xj ai aj αi αj 1 kij (5.96) j Eq. (5.94) is also used to determine the fugacity coefficient of component in the gas phase, Φvi , by using the composition of the gas phase, yi, in calculating A, B, Zv, and other composition-dependent terms; or, bi ðZv 1Þ A 2Ψ i bi B ln 1 + v ln Φvi ¼ lnðZv BÞ bm B ðaαÞm bm Z where X pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi yj ai aj αi αj 1 kij j XX pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ðaαÞm ¼ yi yj ai aj αi αj 1 kij Ψi ¼ i j Gibbs Energy and Selecting the Correct Z-Factor In solving the cubic Z-factor equation for hydrocarbon mixtures, one or three real roots may exist. When three roots exist with different values designated as n n n Largest root: ZLargest Middle root: Zmiddle Smallest root: ZSmallest, The middle root ZMiddle, always is discarded as it has nonphysical value or meaning. However, selecting of the correct root between the remaining two 528 CHAPTER 5 Equations of State and Phase Equilibria roots is important to avoid phase equilibria convergence problems, which will lead the EOS to generate unrealistic results. The correct root is chosen as the one with the lowest normalized Gibbs energy function, g*. The normalized Gibbs energy function is defined in terms of the composition of the hydrocarbon phase and the individual component fugacity in that phase. Mathematically, the normalized Gibbs energy function is defined by the following expressions: Normalized Gibbs energy for gas phase: ggas ¼ n X yi ln fiv i¼1 Normalized Gibbs energy for liquid phase: gLiquid ¼ n X xi ln fiL i¼1 Therefore, assume that a liquid phase with a composition of xi has three Z-factor roots; the middle root is discarded automatically and the remaining two are designated ZL1 and ZL2 . To select the correct root, calculate the normalized Gibbs energy function using the two remaining roots: gZL ¼ 1 n X xi ln fiL i¼1 n X xi ln fiL gZL ¼ 2 i¼1 in which the fugacity fLi is determined from Eq. (5.94) using the two remaining roots, ZL1 and ZL2 , as L bi ðZL1 1Þ A 2Ψ i bi B ln 1 + fi zL1 ¼ xi p + exp lnðZL1 BÞ bm B ðaαÞm bm ZL1 L bi ðZL2 1Þ A 2Ψ i bi B ln 1 + lnðZL2 BÞ fi zL2 ¼ xi + exp bm B ðaαÞm bm ZL2 If gZL1 <gZL2 then ZL1 is chosen as the correct root; otherwise, ZL2 is chosen. It should be pointed out that the above procedure is applied for the gas phase if three Z-factor values are obtained with largest as referred to as Zg1 and smallest as Zg2; with the normalized Gibbs energy function as given by gZg1 ¼ n X yi ln fiv gZg2 ¼ n X yi ln fiv i¼1 i¼1 Equations of State 529 in which the fugacity f vi is determined by using the two remaining roots, Zg1 and Zg2, as v bi Zg1 1 A 2Ψ i bi B ln 1 + fi zg1 ¼ xi p + exp ln Zg1 B bm B ðaαÞm bm Zg1 v bi Zg2 1 A 2Ψ i bi B ln 1 + fi zg2 ¼ xi + exp ln Zg2 B bm B ðaαÞm bm Zg2 If gZg1 < gZg2 then Zg1 is chosen as the correct root; otherwise, Zg2 is chosen. Chemical Potential The chemical potential “μi” is a quantity that measures how much energy a component brings to a mixture as the component transfers from one phase to another. In a mixture, the chemical potential is defined as the slope of the free energy “G” of the system with respect to a change in the number of moles of that specific component. Thus, it is the partial derivative of the free energy with respect to the amount of that specific component, which is given the name chemical potential: @G μi ¼ @ni p, T , nj In a mixture, the normalized Gibbs free energy “g” is not the same in both phases because their compositions are different. Instead, the criterion for equilibrium is defined mathematically by i @G h gas ¼0 ¼ μi μliquid i @ni This criterion implies that, at equilibrium, the chemical potential of each component μi has the same value in each phase: liquid μgas i ¼ μi The chemical potential of component i is related to the activity a^i by μi ¼ μoi + RT lnða^i Þ with fi a^i ¼ o fi 530 CHAPTER 5 Equations of State and Phase Equilibria where a^i ¼ activity of component i in mixture at p and T μoi ¼ chemical potential of pure of component i at p and T f oi ¼ fugacity of pure component i at p and T fi ¼ fugacity of component i in mixture at p and T MODIFICATIONS OF THE SRK EOS To improve the pure component vapor pressure predictions by the SRK equation of state, Groboski and Daubert (1978) proposed a new expression for calculating parameter m of Eq. (5.72). The proposed relationship, which originated from analyzing extensive experimental data for pure hydrocarbons, has the following form: m ¼ 0:48508 + 1:55171ω 0:15613ω2 (5.97) Sim and Daubert (1980) pointed out that, because the coefficients of Eq. (5.97) were determined by analyzing vapor pressure data of lowmolecular-weight hydrocarbons, it is unlikely that Eq. (5.97) will suffice for high-molecular-weight petroleum fractions. Realizing that the acentric factors for the heavy petroleum fractions are calculated from an equation such as the Edmister correlation or the Lee and Kesler correlation, the authors proposed the following expressions for determining parameter m: n If the acentric factor is determined using the Edmister correlation, then m ¼ 0:431 + 1:57ωi 0:161ðωÞ2i n (5.98) If the acentric factor is determined using the Lee and Kesler correction, then m ¼ 0:315 + 1:60ωi 0:166ω2i (5.99) Elliot and Daubert (1985) stated that the optimal binary interaction coefficient, kij, would minimize the error in the representation of all the thermodynamic properties of a mixture. Properties of particular interest in phase equilibrium calculations include bubble point pressure, dew point pressure, and equilibrium ratios. The authors proposed a set of relationships for determining interaction coefficients for asymmetric mixtures2 that contain 2 An asymmetric mixture is defined as one in which two of the components are considerably different in their chemical behavior. Mixtures of methane with hydrocarbons of 10 or more carbon atoms can be considered asymmetric. Mixtures containing gases such as nitrogen or hydrogen are asymmetric. Modifications of the SRK EOS 531 methane, N2, CO2, and H2S. Referring to the principal component as i and other fractions as j, Elliot and Daubert proposed the following expressions: Binary Interaction between N2 – Components kij ¼ 0:107089 + 2:9776kij1 (5.100) Binary Interaction between CO2 – Components kij ¼ 0:08058 0:77215kij1 1:8404 kij1 (5.101) Binary Interaction between H2S – Components kij ¼ 0:07654 + 0:017921kij1 (5.102) Methane with components > C9 (Nonane) kij ¼ 0:17985 + 2:6958kij1 + 10:853 kij1 2 (5.103) where kij1 ¼ 2 εi εj 2εi εj (5.104) and εi ¼ pffiffiffiffi 0:480453 ai bi (5.105) The two parameters, ai and bi, of Eq. (5.105) were defined previously by Eqs. (5.73), (5.74); that is, a ¼ Ωa R2 Tc2 pc b ¼ Ωb RTc pc Volume Translation (Volume Shift) Parameter The major drawback in the SRK EOS is that the critical compressibility factor takes on the unrealistic universal critical compressibility of 0.333 for all substances. Consequently, the molar volumes typically are overestimated; that is, densities are underestimated. Peneloux et al. (1982) developed a procedure for improving the volumetric predictions of the SRK EOS by introducing a volume correction parameter, ci, into the equation. This third parameter does not change the vapor/liquid 532 CHAPTER 5 Equations of State and Phase Equilibria equilibrium conditions determined by the unmodified SRK equation (ie, the equilibrium ratio Ki) but it modifies the liquid and gas volumes. The proposed methodology, known as the volume translation method, uses the following expressions: L Vcorr ¼ VL X ðxi ci Þ (5.106) i v Vcorr ¼ Vv X ðy i c i Þ (5.107) i where VL ¼ uncorrected liquid molar volume (ie, VL ¼ ZLRT/p), ft3/mol Vv ¼ uncorrected gas molar volume Vv ¼ ZvRT/p, ft3/mol VLcorr ¼ corrected liquid molar volume, ft3/mol Vvcorr ¼ corrected gas molar volume, ft3/mol xi ¼ mole fraction of component i in the liquid phase yi ¼ mole fraction of component i in the gas phase The authors proposed six schemes for calculating the correction factor, ci, for each component. For petroleum fluids and heavy hydrocarbons, Peneloux and coworkers suggested that the best correlating parameter for the volume correction factor ci is the Rackett compressibility factor, ZRA. The correction factor then is defined mathematically by the following relationship: ci ¼ 4:437978ð0:29441 ZRA ÞTci =pci (5.108) where ci ¼ volume shift coefficient for component i, ft3/lb-mol Tci ¼ critical temperature of component i, °R pci ¼ critical pressure of component i, psia The parameter ZRA is a unique constant for each compound. The values of ZRA, in general, are not much different from those of the critical compressibility factors, Zc. If their values are not available, Peneloux et al. proposed the following correlation for calculating ci: ci ¼ ð0:0115831168 + 0:411844152ωi Þ where ωi ¼ acentric factor of component i. Tci pci (5.109) Modifications of the SRK EOS 533 EXAMPLE 5.13 Rework Example 5.12 by incorporating the Peneloux et al. volume correction approach in the solution. Key information from Example 5.12 includes: For gas: Zv ¼ 0:9267, Ma ¼ 20:89 For liquid: Z L ¼ 1:4212, Ma ¼ 100:25 T ¼ 160°F, and p ¼ 4000psi Solution Step 1. Calculate the correction factor ci using Eq. (5.109): Tci ci ¼ ð0:0115831168 + 0:411844152ωi Þ pci The results are shown in the table below. Component xi C1 C2 C3 C4 C5 C6 C P7+ pc Tc 0.45 666.4 343.33 0.05 706.5 549.92 0.05 616.0 666.06 0.03 527.9 765.62 0.01 488.6 845.8 0.01 453 923 0.40 285 1160 ωi ci cixi yi ciyi 0.0104 0.00839 0.003776 0.86 0.00722 0.0979 0.03807 0.001903 0.05 0.00190 0.1522 0.07729 0.003861 0.05 0.00386 0.1852 0.1265 0.00379 0.02 0.00253 0.2280 0.19897 0.001989 0.01 0.00198 0.2500 0.2791 0.00279 0.005 0.00139 0.5200 0.91881 0.36752 0.005 0.00459 0.38564 0.02349 Step 2. Calculate the uncorrected volume of the gas and liquid phase using the compressibility factors as calculated in Example 5.12. Vv ¼ RTZv ð10:73Þð620Þð0:9267Þ ¼ ¼ 1:54119ft3 =mol p 4000 VL ¼ RTZL ð10:73Þð620Þð1:4121Þ ¼ ¼ 2:3485ft3 =mol p 4000 Step 3. Calculate the corrected gas and liquid volumes by applying Eqs. (5.106), (5.107). X L ¼ VL ðxi ci Þ ¼ 2:3485 0:38564 ¼ 1:962927ft3 =mol Vcorr i v Vcorr ¼ Vv X ðyi ci Þ ¼ 1:54119 0:02394 ¼ 1:5177 ft3 =mol i Step 4. Calculate the corrected compressibility factors. v Zcorr ¼ ð4000Þð1:5177Þ ¼ 0:91254 ð10:73Þð620Þ L ¼ Zcorr ð4000Þð1:962927Þ ¼ 1:18025 ð10:73Þð620Þ 534 CHAPTER 5 Equations of State and Phase Equilibria Step 5. Determine the corrected densities of both phases. ρ¼ pMa RTZ ð4000Þð20:89Þ ¼ 13:767lb=ft3 ð10:73Þð620Þð0:91254Þ ð4000Þð100:25Þ ρL ¼ ¼ 51:07lb=ft3 ð10:73Þð620Þð1:18025Þ ρv ¼ As pointed out by Whitson and Brule (2000), when the volume shift is introduced to the EOS for mixtures, the resulting expressions for fugacity are h L pi fi modified ¼ fiL original exp ci RT and h v pi fi modified ¼ fiv original exp ci RT This implies that the fugacity ratios are unchanged by the volume shift: L L fi f iv modified ¼ v original fi modified fi modified The Peng-Robinson Equation of State Peng and Robinson (1976a) conducted a comprehensive study to evaluate the use of the SRK equation of state for predicting the behavior of naturally occurring hydrocarbon systems. They illustrated the need for an improvement in the ability of the equation of state to predict liquid densities and other fluid properties, particularly in the vicinity of the critical region. As a basis for creating an improved model, Peng and Robinson (PR) proposed the following expression: p¼ RT aα V b ðV + bÞ2 cb2 where a, b, and α have the same significance as they have in the SRK model, and parameter c is a whole number optimized by analyzing the values of the terms Zc and b/Vc as obtained from the equation. It is generally accepted that Zc should be close to 0.28 and b/Vc should be approximately 0.26. An optimized value of c ¼ 2 gives Zc ¼ 0.307 and (b/Vc) ¼ 0.253. Based on this value Modifications of the SRK EOS 535 of c, Peng and Robinson proposed the following equation of state (commonly referred to as PR EOS): p¼ RT aα V b V ðV + bÞ + bðV bÞ (5.110) Imposing the classical critical point conditions [Eq. (5.48)] on Eq. (5.110) and solving for parameters a and b yields a ¼ Ωa R2 Tc2 pc (5.111) RTc pc (5.112) b ¼ Ωb where Ωa ¼ 0:45724 Ωb ¼ 0:07780 This equation predicts a universal critical gas compressibility factor, Zc, of 0.307, compared to 0.333 for the SRK model. Peng and Robinson also adopted Soave’s approach for calculating parameter α: pffiffiffiffiffi 2 α ¼ 1 + m 1 Tr (5.113) where m ¼ 0.3796 + 1.54226ω 0.2699ω2. Peng and Robinson (1978) proposed the following modified expression for m that is recommended for heavier components with acentric values ω > 0.49: m ¼ 0:379642 + 1:48503ω 0:1644ω2 + 0:016667ω3 (5.114) Rearranging Eq. (5.110) into the compressibility factor form, gives Z3 + ðB 1ÞZ 2 + A 3B2 2B Z AB B2 B3 ¼ 0 (5.115) where A and B are given by Eqs. (5.81), (5.82) for pure component and by Eqs. (5.85), (5.86) for mixtures; that is, A¼ B¼ with ðaαÞm ¼ bm ¼ XX ðRT Þ2 bm p RT pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi xi xj ai aj αi αj 1 kij i j X ½xi bi i ðaαÞm p 536 CHAPTER 5 Equations of State and Phase Equilibria EXAMPLE 5.14 Using the composition given in Example 5.12, calculate the density of the gas phase and liquid phase by using the Peng-Robinson EOS. Assume kij ¼ 0. The components are described in the following table. Component xi yi pc Tc ωi C1 C2 C3 C4 C5 C6 C7+ 0.45 0.05 0.05 0.03 0.01 0.01 0.40 0.86 0.05 0.05 0.02 0.01 0.005 0.0005 666.4 706.5 616.0 527.9 488.6 453 285 343.33 549.92 666.06 765.62 845.8 923 1160 0.0104 0.0979 0.1522 0.1852 0.2280 0.2500 0.5200 The heptanes-plus fraction has the following properties: M ¼ 215 pc ¼ 285 psia Tc ¼ 700°F ω ¼ 0:582 Solution Step 1. Calculate the parameters a, b, and α for each component in the system by applying Eqs. (5.111)–(5.113): R2 Tc2 pc RTc b ¼ 0:07780 pc pffiffiffiffiffi 2 α ¼ 1 + m 1 Tr a ¼ 0:45724 The results are shown in the table below. Component pc C1 C2 C3 C4 C5 C6 C7+ Tc 666.4 343.33 706.5 549.92 616.0 666.06 527.9 765.62 488.6 845.8 453 923 285 1160 ωi m αi ai 0.0104 1.8058 0.7465 9294.4 0.0979 1.1274 0.9358 22,506.1 0.1522 0.9308 1.0433 37,889.0 0.1852 0.80980 1.1357 56,038.9 0.2280 0.73303 1.2170 77,058.9 0.2500 0.67172 1.2882 100,662.3 0.5200 0.53448 1.6851 248,550.5 bi 0.4347 0.6571 0.9136 1.1755 1.6091 1.4635 3.4414 Step 2. Calculate the mixture parameters (aα)m and bm for the gas and liquid phase, to give the following. Modifications of the SRK EOS 537 For the gas phase, XX pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi yi yj ai aj αi αj 1 kij ¼ 10,423:54 ðaαÞm ¼ i j X bm ¼ ðyi bi Þ ¼ 0:862528 i For the liquid phase, XX pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi xi xj ai aj αi αj 1 kij ¼ 107,325:4 ðaαÞm ¼ i j X bm ¼ ðxi bi Þ ¼ 1:696543 i Step 3. Calculate the coefficients A and B, to give the following. For the gas phase, A¼ ðaαÞm p ð10, 423:54Þð4000Þ ¼ ¼ 0:94209 R2 T 2 ð10:73Þ2 ð620Þ2 B¼ bm p ð0:862528Þð4000Þ ¼ ¼ 0:30669 RT ð10:73Þð620Þ For the liquid phase, A¼ ðaαÞm p ð107, 325:4Þð4000Þ ¼ ¼ 9:700183 R2 T 2 ð10:73Þ2 ð620Þ2 B¼ bm p ð1:696543Þð4000Þ ¼ ¼ 1:020078 RT ð10:73Þð620Þ Step 4. Solve Eq. (5.115) for the compressibility factor of the gas and liquid phases to give Z3 + ðB 1ÞZ 2 + A 3B2 2B Z AB B2 B3 ¼ 0 For the gas phase: Substituting for A ¼ 0.94209 and B ¼ 0.30669 in the above equation gives Z 3 + ðB 1ÞZ2 + A 3B2 2B2 2B Z AB B2 B3 ¼ 0 h i Z 3 ¼ ð0:30669 1ÞZ 2 + 0:94209 3ð0:30669Þ2 2ð0:30669Þ Z h i ð0:94209Þð0:30669Þ ð0:30669Þ2 ð0:30669Þ3 ¼ 0 Solving for Z gives Z v ¼ 0:8625 For the liquid phase: Substituting A ¼ 9.700183 and B ¼ 1.020078 in Eq. (5.115) gives Z + ðB 1ÞZ 2 + A 3B2 2B Z AB B2 B3 ¼ 0 h i Z3 ¼ ð1:020078 1ÞZ2 + ð9:700183Þ 3ð1:020078Þ2 2ð1:020078Þ Z h i ð9:700183Þð1:020078Þ ð1:020078Þ2 ð1:020078Þ3 ¼ 0 3 538 CHAPTER 5 Equations of State and Phase Equilibria Solving for Z gives ZL ¼ 1:2645 Step 5. Calculate the density of both phases. Density of the vapor phase: pMa Zv RT ð4000Þð20:89Þ ρv ¼ ¼ 14:566lb=ft3 ð10:73Þð620Þð0:8625Þ ρv ¼ Density of the liquid phase: pMa Zv RT ð4000Þð100:25Þ ¼ 47:67lb=ft3 ρL ¼ ð10:73Þð620Þð1:2645Þ ρL ¼ Applying the thermodynamic relationship, as given by Eq. (5.88), to Eq. (5.111) yields the following expression for the fugacity of a pure component: pffiffiffi # " Z+ 1+ 2 B f A pffiffiffi ln ¼ lnðΦÞ ¼ Z 1 lnðZ BÞ pffiffiffiffiffiffi ln p Z 1 2 B 2 2B (5.116) The fugacity coefficient of component i in a hydrocarbon liquid mixture is calculated from the following expression: L bi ðZL 1Þ f ln ln ZL B ¼ ln ΦLi ¼ xi p bm " L pffiffiffi # Z + 1 + 2B A 2Ψ i bi pffiffiffi ln pffiffiffi ð aα Þ b Z L + 1 2B 2 2B m m (5.117) where the mixture parameters bm, B, A, Ψ i, and (aα)m are defined previously. Eq. (5.117) also is used to determine the fugacity coefficient of any component in the gas phase by replacing the composition of the liquid phase, xi, with the composition of the gas phase, yi, in calculating the compositiondependent terms of the equation: ln v bi ðZ v 1Þ f A 2Ψ i bi lnðZ v BÞ pffiffiffi ¼ ln ΦLi ¼ yi p bm 2 2B ðaαÞm bm " pffiffiffi # v Z + 1 + 2B pffiffiffi ln Z v + 1 2B Modifications of the SRK EOS 539 The table below shows the set of binary interaction coefficient, kij, traditionally used when predicting the volumetric behavior of hydrocarbon mixture with the Peng-Robinson equation of state. Component CO2 N2 H2S C1 C2 C3 i-C4 n-C4 i-C5 n-C5 C6 C7 C8 C9 C10 CO2 N2 H2S C1 C2 C3 i-C4 n-C4 i-C5 n-C5 C6 C7 C8 C9 C10 0.105 0.025 0.070 0 0.130 0.010 0.085 0.005 0 0.125 0.090 0.080 0.010 0.005 0 0.120 0.095 0.075 0.035 0.005 0.000 0 0.115 0.095 0.075 0.025 0.010 0.000 0.005 0 0.115 0.100 0.070 0.050 0.020 0.015 0.005 0.005 0 0.115 0.100 0.070 0.030 0.020 0.015 0.005 0.005 0.000 0 0.115 0.110 0.070 0.030 0.020 0.010 0.005 0.005 0.000 0.000 0 0.115 0.115 0.060 0.035 0.020 0.005 0.005 0.005 0.000 0.000 0.000 0 0.115 0.120 0.060 0.040 0.020 0.005 0.005 0.005 0.000 0.000 0.000 0.000 0 0.115 0.120 0.060 0.040 0.020 0.005 0.005 0.005 0.000 0.000 0.000 0.000 0.000 0 0.115 0.125 0.055 0.045 0.020 0.005 0.005 0.005 0.000 0.000 0.000 0.000 0.000 0.00 0 0 0 0 0.135 0.130 0 Note that kij ¼ kij. To improve the predictive capability of the PR EOS when applying for mixtures containing nonhydrocarbon components (eg, N2, CO2, and CH4), Nikos et al. (1986) proposed a generalized correlation for generating the binary interaction coefficient, kij. The authors correlated these coefficients with system pressure, temperature, and the acentric factor. These generalized correlations originated with all the binary experimental data available in the literature. The authors proposed the following generalized form for kij: kij ¼ λ2 Trj2 + λ1 Trj + λ0 (5.118) where i refers to the principal component, N2, CO2, or CH4, and j refers to the other hydrocarbon component of the binary. The acentric-factordependent coefficients, λ0, λ1, and λ2 are determined for each set of binaries by applying the following expressions: For nitrogen-hydrocarbons, 2 λ0 ¼ 0:1751787 0:7043 log ωj 0:862066 log ωj (5.119) 2 λ1 ¼ 0:584474 1:328 log ωj + 2:035767 log ωj (5.120) 540 CHAPTER 5 Equations of State and Phase Equilibria 2 λ2 ¼ 2:257079 + 7:869765 log ωj + 13:50466 log ωj + 8:3864½ log ðωÞ3 (5.121) Nikos et al. also suggested the following pressure correction: kij0 ¼ kij 1:04 4:2 105 p (5.122) where p is the pressure in psi. For methane-hydrocarbons, 2 λ0 ¼ 0:01664 0:37283 log ωj + 1:31757 log ωj (5.123) λ1 ¼ 0:48147 + 3:35342 log ðωi Þ 1:0783½ log ðωi Þ2 (5.124) λ2 ¼ 0:4114 3:5072 log ðωi Þ 0:78798½ log ðωi Þ2 (5.125) For CO2-hydrocarbons, λ0 ¼ 0:4025636 + 0:1748927 log ωj λ1 ¼ 0:94812 0:6009864 log ωj (5.127) λ2 ¼ 0:741843368 + 0:441775 log ðωi Þ (5.128) (5.126) For the CO2 interaction parameters, the following pressure correction is suggested: kij0 ¼ kij 1:044269 4:375 105 p (5.129) Stryjek and Vera (1986) proposed an improvement in the reproduction of vapor pressures of a pure component by the PR EOS in the reduced temperature range from 0.7 to 1.0, by replacing the m term in Eq. (5.113) with the following expression: mo ¼ 0:378893 + 1:4897153 0:17131848ω2 + 0:0196554ω3 (5.130) To reproduce vapor pressures at reduced temperatures below 0.7, Stryjek and Vera further modified the m parameter in the PR equation by introducing an adjustable parameter, m1, characteristic of each compound to Eq. (5.113). They proposed the following generalized relationship for the parameter m: pffiffiffiffiffi m ¼ mo + m1 1 + Tr ð0:7 Tr Þ where Tr ¼ reduced temperature of the pure component m0 ¼ defined by Eq. (5.130) m1 ¼ adjustable parameter (5.131) Modifications of the SRK EOS 541 For all components with a reduced temperature above 0.7, Stryjek and Vera recommended setting m1 ¼ 0. For components with a reduced temperature greater than 0.7, the optimum values of m1 for compounds of industrial interest are tabulated below, Component m1 Nitrogen Carbon dioxide Water Methane Ethane Propane Butane Pentane Hexane Heptane Octane Nonane Decane Undecane Dodecane Tridecane Tetradecane Pentadecane Hexadecane 0.01996 0.04285 0.0664 0.0016 0.02669 0.03136 0.03443 0.03946 0.05104 0.04648 0.04464 0.04104 0.0451 0.02919 0.05426 0.04157 0.02686 0.01892 0.02665 Due to the totally empirical nature of the parameter m1, Stryjek and Vera could not find a generalized correlation for m1 in terms of pure component parameters. They pointed out that these values of m1 should be used without changes. Jhaveri and Youngren (1984) pointed out that, when applying the PengRobinson equation of state to reservoir fluids, the error associated with the equation in the prediction of gas-phase Z-factors ranged from 3% to 5% and the error in the liquid density predictions ranged from 6% to 12%. Following the procedure proposed by Peneloux and coworkers (see the SRK EOS), Jhaveri and Youngren introduced the volume correction parameter, ci, to the PR EOS. This third parameter has the same units as the second parameter, bi, of the unmodified PR equation and is defined by the following relationship: ci ¼ Si bi (5.132) where Si is a dimensionless parameter, called the shift parameter, and bi is the Peng-Robinson covolume, as given by Eq. (5.112). 542 CHAPTER 5 Equations of State and Phase Equilibria As in the SRK EOS, the volume correction parameter, ci, does not change the vapor/liquid equilibrium conditions, equilibrium ratio Ki. The corrected hydrocarbon phase volumes are given by the following expressions: L Vcorr ¼ VL X ðxi ci Þ v ¼ Vv Vcorr X ðy i c i Þ i¼1 i¼1 where VL, Vv ¼ volumes of the liquid phase and gas phase as calculated by the unmodified PR EOS, ft3/mol VLcorr, Vvcorr ¼ corrected volumes of the liquid and gas phase, ft3/mol Whitson and Brule (2000) point out that the volume translation (correction) concept can be applied to any two-constant cubic equation, thereby eliminating the volumetric deficiency associated with the application of EOS. Whitson and Brule extended the work of Jhaveri and Youngren and tabulated the shift parameters, Si, for a selected number of pure components. These tabulated values, which follow, are used in Eq. (5.132) to calculate the volume correction parameter, ci, for the Peng-Robinson and SRK equations of state. Component PR EOS N2 CO2 H2S C1 C2 C3 i-C4 n-C4 i-C5 n-C5 n-C6 n-C7 n-C8 n-C9 n-C10 0.1927 0.0817 0.1288 0.1595 0.1134 0.0863 0.0844 0.0675 0.0608 0.0390 0.0080 0.0033 0.0314 0.0408 0.0655 SRK EOS 0.0079 0.0833 0.0466 0.0234 0.0605 0.0825 0.0830 0.0975 0.1022 0.1209 0.1467 0.1554 0.1794 0.1868 0.2080 Jhaveri and Youngren proposed the following expression for calculating the shift parameter for the C7+: SC7 + ¼ 1 d ð M Þe Modifications of the SRK EOS 543 where M ¼ molecular weight of the heptanes-plus fraction d, e ¼ positive correlation coefficients The authors proposed that, in the absence of the experimental information needed for calculating e and d, the power coefficient e could be set equal to 0.2051 and the coefficient d adjusted to match the C7+ density with the values of d ranging from 2.2 to 3.2. In general, the following values may be used for C7+ fractions, by hydrocarbon family: Hydrocarbon family d e Paraffins Naphthenes Aromatics 2.258 3.044 2.516 0.1823 0.2324 0.2008 To use the Peng-Robinson equation of state to predict the phase and volumetric behavior of mixtures, one must be able to provide the critical pressure, the critical temperature, and the acentric factor for each component in the mixture. For pure compounds, the required properties are well defined and known. Nearly all naturally occurring petroleum fluids contain a quantity of heavy fractions that are not well defined. These heavy fractions often are lumped together as a heptanes-plus fraction. The problem of how to adequately characterize the C7+ fractions in terms of their critical properties and acentric factors has been long recognized in the petroleum industry. Changing the characterization of C7+ fractions present in even small amounts can have a profound effect on the PVT properties and the phase equilibria of a hydrocarbon system as predicted by the Peng-Robinson equation of state. The usual approach for such situations is to “tune” the parameters in the EOS in an attempt to improve the accuracy of prediction. During the tuning process, the critical properties of the heptanes-plus fraction and the binary interaction coefficients are adjusted to obtain a reasonable match with experimental data available on the hydrocarbon mixture. Recognizing that the inadequacy of the predictive capability of the PR EOS lies with the improper procedure for calculating the parameters a, b, and α of the equation for the C7+ fraction, Ahmed (1991) devised an approach for determining these parameters from the following two readily measured physical properties of C7+: the molecular weight, M7+, and the specific gravity, γ 7+. The approach is based on generating 49 density values for the C7+ by applying the Riazi and Daubert correlation. These values were subsequently subjected to 10 temperature and 10 pressure values in the range of 60–300°F and 14.7–7000 psia, respectively. The Peng-Robinson EOS was then 544 CHAPTER 5 Equations of State and Phase Equilibria applied to match the 4900 generated density values by optimizing the parameters a, b, and α using a nonlinear regression model. The optimized parameters for the heptanes-plus fraction are given by the following expressions. For the parameter α of C7+, " rffiffiffiffiffiffiffiffi!#2 520 α¼ 1+m 1 T (5.133) with m as defined by m¼ D A4 A7 + A2 M7 + + A3 M72 + + + A5 γ 7 + + A6 γ 27 + + M7 + γ7 + A0 + A1 D (5.134) with parameter D as defined by the ratio of the molecular weight to the specific gravity of the heptanes-plus fraction: D¼ M7 + γ7 + where M7+ ¼ molecular weight of C7+ γ 7+ ¼ specific gravity of C7+ A0-A7 ¼ coefficients as given in the table below Coefficient a A0 A1 A2 A3 A4 A5 A6 A7 b m 2.433525 10 6.8453198 36.91776 8.3201587 103 1.730243 102 5.2393763 102 2 6 0.18444102 10 6.2055064 10 1.7316235 102 2 9 3.6003101 10 9.0910383 10 1.3743308 105 7 3.4992796 10 13.378898 12.718844 2.838756 107 7.9492922 10.246122 1.1325365 107 3.1779077 7.6697942 6.418828 106 1.7190311 2.6078099 7 For parameters a and b of C7+, the following generalized correlation is proposed: " a or b ¼ 3 X i¼0 Ai D i # " # 6 X A4 A7 i4 + + Ai γ 7 + + D γ 7+ i¼5 (5.135) The coefficients A0–A7 are included in the table above. To further improve the predictive capability of the Peng-Robinson EOS, Ahmed (1991) Modifications of the SRK EOS 545 optimized the coefficients a, b, and m for nitrogen, CO2, and methane by matching 100 Z-factor values for each of these components. Using a nonlinear regression model, the optimized values in the table below are recommended. Component a CO2 N2 C1 1.499914 10 4.5693589 103 7.709708 103 4 b m (Eq. (5.133)) 0.41503575 0.4682582 0.46749727 0.73605717 0.97962859 0.549765 To provide the modified PR EOS with a consistent procedure for determining the binary interaction coefficient, kij, the following computational steps are proposed. Step 1. Calculate the binary interaction coefficient between methane and the heptanes-plus fraction from kC1 C7 + ¼ 0:00189T 1:167059 where the temperature, T, is in °R. Step 2. Set kCO2 -N2 ¼ 0:12 KCO2 -hydrocarbon ¼ 0:10 kN2 -hydrocarbon ¼ 0:10 Step 3. Adopting the procedure recommended by Pedersen, Thomassen, and Fredenslund (1989), calculate the binary interaction coefficients between components heavier than methane (ie, C2, C3, etc.) and the heptanes-plus fraction from kCn C7 + ¼ 0:8kCðn1Þ C7 + where n is the number of carbon atoms of component Cn; for example, Binary interaction coefficient between C2 and C7 + : kC2 C7 + ¼ 0:8kC1 C7 + Binary interaction coefficient between C3 and C7 + : kC3 C7 + ¼ 0:8kC2 C7 + Step 4. Determine the remaining kij from " kij kiC7 + Mj 5 ðM i Þ5 ðMC7 + Þ5 ðMi Þ5 # 546 CHAPTER 5 Equations of State and Phase Equilibria where M is the molecular weight of any specified component. For example, the binary interaction coefficient between propane, C3, and butane, C4, is " kC3 C4 ¼ kC3 C7 + ðM C 4 Þ5 ðM C 3 Þ5 # ðMC7 + Þ5 ðMC3 Þ5 Summary of Equations of State A summary of the different equations of state discussed in this chapter are listed in the table below in the form of p ¼ prepulsion pattraction EOS prepulsion pattraction a b Ideal RT V RT V b RT V b RT V b RT V b 0 0 0 a V2 R2 Tc2 pc R2 Tc2:5 Ωa pc R2 Tc2 Ωa pc R2 Tc2 Ωa pc Ωb vdW RK SRK PR a pffiffiffi V ð V + bÞ T aαðT Þ V ð V + bÞ aαðT Þ V ðV + bÞ + bðV bÞ Ωa RTc pc RTc Ωb pc RTc Ωb pc RTc Ωb pc EQUATION-OF-STATE APPLICATIONS Most petroleum engineering applications rely on the use of the equation of state due to its simplicity, consistency, and accuracy when properly tuned. Cubic equations of state have found widespread acceptance as tools that permit the convenient and flexible calculation of the complex phase behavior of reservoir fluids. Some of these applications are presented in this section, including: n n n n n n Equilibrium ratio “K-value” Dew point pressure “pd” Bubble point pressure “pb” Three-phase flash calculations Vapor pressure “pv” Simulating PVT laboratory experiments. Equilibrium Ratio A flow diagram is presented in Fig. 5.9 to illustrate the procedure of determining equilibrium ratios of a hydrocarbon mixture. For this type of calculation, the system temperature, T, the system pressure, p, and the overall Equation-of-State Applications 547 Given: zi, P, T Assume KAi KA i = Ki yi, nV ZV, FVi Xi, nL Perform flash calculations Z L, F Li Calculate Ki e > 0.0001 No, set Tci pci )] Ki = p exp[5.37(1 + wi)(1− T n KA i = Ki Test for convergence e ≤ 0.0001 Ki −1 ≤ e KA i where e ≈ 0.0001 i=1 Yes Solution gives: Ki, xi, yi, nL, nV, Z L, Z V n FIGURE 5.9 Flow diagram of equilibrium ratio determination by EOS. composition of the mixture, zi, must be known. The procedure is summarized in the following steps in conjunction with Fig. 5.9. Step 1. Assume an initial value of the equilibrium ratio for each component in the mixture at the specified system pressure and temperature. Wilson’s equation can provide the starting values of Ki: KiA ¼ pci exp½5:37ð1 + ωi Þð1 Tci =T Þ p where KA i is an assumed equilibrium ratio for component i. Step 2. Using the overall composition and the assumed K-values, perform flash calculations to determine xi, yi, nL, and nυ. Step 3. Using the calculated composition of the liquid phase, xi, determine the fugacity coefficient ΦLi for each component in the liquid phase by applying Eq. (5.117): ln ΦLi " L pffiffiffi # L Z + 1+ 2 B bi ðZL 1Þ A 2Ψ i bi pffiffiffi ln ln Z B pffiffiffi ¼ bm ZL 1 2 B 2 2B ðaαÞm bm 548 CHAPTER 5 Equations of State and Phase Equilibria Step 4. Repeat step 3 using the calculated composition of the gas phase, yi, to determine Φvi . " v pffiffiffi # L bi ðZv 1Þ Z + 1+ 2 B A 2Ψ i bi v pffiffiffi ln ln Φi ¼ lnðZ BÞ pffiffiffi bm 2 2B ðaαÞm bm Zv 1 2 B Step 5. Calculate the new set of equilibrium ratios from Ki ¼ ΦLi Φvi Step 6. Check for the solution by applying the following constraint: n X 2 Ki =KiA 1 ε i¼1 where ε ¼ preset error tolerance, such as 0.0001 n ¼ number of components in the system. If these conditions are satisfied, then the solution has been reached. If not, steps 1–6 are repeated using the calculated equilibrium ratios as initial values. Dew Point Pressure A saturated vapor exists for a given temperature at the pressure at which an infinitesimal amount of liquid first appears. This pressure is referred to as the dew point pressure, pd. The dew point pressure of a mixture is described mathematically by the following two conditions: yi ¼ zi , for 1 i n and nv ¼ 1 n X zi ¼1 K i i1 (5.136) (5.137) Applying the definition of Ki in terms of the fugacity coefficient to Eq. (5.137) gives n X zi i¼1 ki " n X ¼ i¼1 # X zi f v zi i ¼1 ¼ ΦLi zi pd ΦLi =Φvi or pd ¼ n v X f i i¼1 ΦLi Equation-of-State Applications 549 This above equation is arranged to give f ðpd Þ ¼ n v X f i i¼1 ΦLi pd ¼ 0 (5.138) where pd ¼ dew point pressure, psia f vi ¼ fugacity of component i in the vapor phase, psia ΦLi ¼ fugacity coefficient of component i in the liquid phase Eq. (5.35) can be solved for the dew point pressure by using the NewtonRaphson iterative method. To use the iterative method, the derivative of Eq. (5.138) with respect to the dew point pressure, pd, is required. This derivative is given by the following expression: " # n ΦLi @fiv =@pd fiv @ΦLi =@pd @f X 1 ¼ L 2 @pd i¼1 Φ (5.139) i The two derivatives in this equation can be approximated numerically as follows: v @f v f ðpd + Δpd Þ fiv ðpd Δpd Þ ¼ i @pd 2Δpd (5.140) L @ΦLi Φ ðpd + Δpd Þ ΦLi ðpd Δpd Þ ¼ i @pd 2Δpd (5.141) and where Δpd ¼ pressure increment (eg, 5 psia) fiv ðpd + Δpd Þ ¼ fugacity of component i at (pd + Δpd) f vi (pd–Δpd) ¼ fugacity of component i at (pd Δpd) ΦLi ðpd + Δpd Þ ¼ fugacity coefficient of component i at (pd + Δpd) ΦLi (pd–Δpd) ¼ fugacity coefficient of component i at (pd Δpd) ΦLi ¼ fugacity coefficient of component i at pd The computational procedure of determining pd is summarized in the following steps. Step 1. Assume an initial value for the dew point pressure, pA d. Step 2. Using the assumed value of pA d , calculate a set of equilibrium ratios for the mixture by using any of the previous correlations, such as Wilson’s correlation. 550 CHAPTER 5 Equations of State and Phase Equilibria Step 3. Calculate the composition of the liquid phase, that is, the composition of the droplets of liquid, by applying the mathematical definition of Ki, to give xi ¼ zi Ki It should be noted that yi ¼ zi. L Step 4. Calculate pA d using the composition of the gas phase, zi and Φi , using the composition of liquid phase, xi, at the following three pressures: n n n pA d pA d + Δpd pA d –Δpd where pA d is the assumed dew point pressure and Δpd is a selected pressure increment of 5–10 psi. Step 5. Evaluate the function f(pd) [ie, Eq. (5.138)] and its derivative using Eqs. (5.139)–(5.141). Step 6. Using the values of the function f(pd) and the derivative @f/@pd as determined in step 5, calculate a new dew point pressure by applying the Newton-Raphson formula: pd ¼ pA d f ðpd Þ=½@f =@pd (5.142) Step 7. The calculated value of pd is checked numerically against the assumed value by applying the following condition: jpd pA dj5 If this condition is met, then the correct dew point pressure, pd, has been found. Otherwise, steps 3–6 are repeated using the calculated pd as the new value for the next iteration. A set of equilibrium ratios must be calculated at the new assumed dew point pressure from ki ¼ ΦLi Φvi Bubble Point Pressure The bubble point pressure, pb, is defined as the pressure at which the first bubble of gas is formed. Accordingly, the bubble point pressure is defined mathematically by the following equations: yi ¼ zi and nL ¼ 1:0 (5.143) n X ½zi Ki ¼ 1 (5.144) i¼1 Equation-of-State Applications 551 Introducing the concept of the fugacity coefficient into Eq. (5.144) gives 2 L 3 fi n n 6 X X 7 ΦLi z 6zi i pb 7 ¼ 1 zi v ¼ v 5 4 Φi Φi i¼1 i¼1 Rearranging, pb ¼ n L X f i Φvi i¼1 or f ðpb Þ ¼ n L X f i i¼1 p b ¼0 v Φi (5.145) The iteration sequence for calculation of pb from the preceding function is similar to that of the de-point pressure, which requires differentiating that function with respect to the bubble point pressure: " # n Φvi @fiL =@pb fiL @Φvi =@pb @f X 1 ¼ v 2 @pb i¼1 Φ (5.146) i The phase envelope can be constructed by either calculating the saturation pressures, pb and pd, as a function of temperature or calculating the bubble point and dew point temperatures at selected pressures. The selection to calculate a pressure or temperature depends on how steep or flat the phase envelope is at a given point. In terms of the slope of the envelope at a point, the criteria are n Calculate the bubble point or dew point pressure if @ lnðpÞ lnðp2 Þ lnðp1 Þ @ lnðT Þ lnðT Þ lnðT Þ < 20 2 1 n Calculate the bubble point or dew point temperature if @ lnðpÞ lnðp2 Þ lnðp1 Þ @ lnðT Þ lnðT Þ lnðT Þ > 20 2 1 Three-Phase Equilibrium Calculations Two- and three-phase equilibria occur frequently during the processing of hydrocarbon and related systems. Peng and Robinson (1976b) proposed a three-phase equilibrium calculation scheme of systems that exhibit a water-rich liquid phase, a hydrocarbon-rich liquid phase, and a vapor phase. 552 CHAPTER 5 Equations of State and Phase Equilibria In applying the principle of mass conversation to 1 mol of a waterhydrocarbon in a three-phase state of thermodynamic equilibrium at a fixed temperature, T, and pressure, p, gives nL + nw + nv ¼ 1 (5.147) nL xi + nw xwi + nv yi ¼ zi (5.148) n n n n X X X X xi ¼ xwi ¼ yi ¼ zi ¼ 1 (5.149) i i¼1 i¼1 i¼1 where nL, nw, nυ ¼ number of moles of the hydrocarbon-rich liquid, the waterrich liquid, and the vapor, respectively xi, xwi, yi ¼ mole fraction of component i in the hydrocarbon-rich liquid, the water rich liquid, and the vapor, respectively The equilibrium relations between the compositions of each phase are defined by the following expressions: Ki ¼ yi ΦLi ¼ xi Φvi (5.150) Kwi ¼ yi Φw ¼ i xwi Φvi (5.151) and where Ki ¼ equilibrium ratio of component i between vapor and hydrocarbonrich liquid Kwi ¼ equilibrium ratio of component i between the vapor and waterrich liquid ΦLi ¼ fugacity coefficient of component i in the hydrocarbon-rich liquid Φvi ¼ fugacity coefficient of component i in the vapor phase Φw i ¼ fugacity coefficient of component i in the water-rich liquid Combining Eqs. (5.147)–(5.151) gives the following conventional nonlinear equations: 3 2 X X6 6 xi ¼ 4 i¼1 i¼1 7 7¼1 5 Ki Ki + Ki Kwi zi nL ð1 Ki Þ + nw (5.152) Equation-of-State Applications 553 3 2 X X6 7 zi ðKi =Kwi Þ 7¼1 6 xwi ¼ 5 4 K i i¼1 i¼1 n ð1 K Þ + n Ki + Ki L i w Kwi 3 2 X yi ¼ i¼1 X6 6 4 i¼1 7 7¼1 5 Ki Ki + Ki Kwi zi Ki nL ð1 Ki Þ + nw (5.153) (5.154) Assuming that the equilibrium ratios between phases can be calculated, these equations are combined to solve for the two unknowns, nL and nυ, and hence xi, xwi, and yi. The nature of the specific equilibrium calculation is what determines the appropriate combination of Eqs. (5.152)–(5.154). The combination of these three expressions then can be used to determine the phase and volumetric properties of the three-phase system. There are essentially three types of phase behavior calculations for the threephase system: bubble point prediction, dew point prediction, and flash calculation. Peng and Robinson (1980) proposed the following combination schemes of Eqs. (5.152)–(5.154). For the bubble point pressure determination, X X xi xwi ¼ 0 i i and " X # yi 1 ¼ 0 i Substituting Eqs. (5.152)–(5.154) in this relationships gives f ðnL , nw Þ ¼ X zi ð1 Ki =Kwi Þ ¼0 nL ð1 Ki Þ + nw ðKi =Kwi Ki Þ + Ki (5.155) zi K i 1¼0 nL ð1 Ki Þ + nw ðKi =Kwi Ki Þ + Ki (5.156) i and gðnL , nw Þ ¼ X i For the dew point pressure, X X xwi yi ¼ 0 "i # i X xi 1 ¼ 0 i 554 CHAPTER 5 Equations of State and Phase Equilibria Combining with Eqs. (5.152)–(5.154) yields f ðnL , nw Þ ¼ X zi Ki ð1=Kwi 1Þ ¼0 nL ð1 Ki Þ + nw ðKi =Kwi Ki Þ + Ki (5.157) zi 1¼0 nL ð1 Ki Þ + nw ðKi =Kwi Ki Þ + Ki (5.158) i and gðnL , nw Þ ¼ X i For flash calculations, X X xi yi ¼ 0 i i " # X xwi 1 ¼ 0 i or f ðnL , nw Þ ¼ X zi ð1 Ki Þ ¼0 nL ð1 Ki Þ + nw ðKi =Kwi Ki Þ + Ki (5.159) zi Ki =Kwi 1:0 ¼ 0 nL ð1 Ki Þ + nw ðKi =Kwi Ki Þ + Ki (5.160) i and gðnL , nw Þ ¼ X i Note that, in performing any of these property predictions, we always have two unknown variables, nL and nw, and between them, two equations. Providing that the equilibrium ratios and the overall composition are known, the equations can be solved simultaneously using the appropriate iterative technique, such as the Newton-Raphson method. The application of this iterative technique for solving Eqs. (5.159), (5.160) is summarized in the following steps. Step 1. Assume initial values for the unknown variables nL and nw. Step 2. Calculate new values of nL and nw by solving the following two linear equations: nL nw new 1 f ðnL , nw Þ @f =@nL @f =@nw nL ¼ gðnL , nw Þ nw @g=@nL @g=@nw where f(nL, nw) ¼ value of the function f(nL, nw) as expressed by Eq. (5.159) and g(nL, nw) ¼ value of the function g(nL, nw) as expressed by Eq. (5.160). The first derivative of these functions with respect to nL and nw is given by the following expressions: Equation-of-State Applications 555 ð@f =@nL Þ ¼ ( X zi ð1 Ki Þ2 ) ½nL ð1 Ki Þ + nw ðKi =Kwi Ki Þ + Ki 2 ( ) X zi ð1 Ki ÞðKi =Kwi Ki Þ i¼1 ð@f =@nw Þ ¼ i¼1 ð@g=@nL Þ ¼ ( X ½nL ð1 Ki Þ + nw ðKi =Kwi Ki Þ + Ki 2 zi ðKi =Kwi Þð1 Ki Þ ) ½nL ð1 Ki Þ + nw ðKi =Kwi Ki Þ + Ki 2 ( ) X zi ðKi Kwi ÞðKi =Kwi Ki Þ i¼1 ð@g=@nw Þ ¼ ½nL ð1 Ki Þ + nw ðKi =Kwi Ki Þ + Ki 2 i¼1 Step 3. The new calculated values of nL and nw then are compared with the initial values. If no changes in the values are observed, then the correct values of nL and nw have been obtained. Otherwise, the steps are repeated with the new calculated values used as initial values. Peng and Robinson (1980) proposed two modifications when using their equation of state for three-phase equilibrium calculations: n The first modification concerns the use of the parameter α as expressed by Eq. (5.113) for the water compound. Peng and Robinson suggested that, when the reduced temperature of this compound is less than 0.85, the following equation should be applied: 0:5 2 αw ¼ 1:0085677 + 0:82154 1 Trw (5.161) where n Trw is the reduced temperature of the water component; that is, Trw ¼ T/ Tcw. The second important modification of the PR EOS is the introduction of alternative mixing rules and new interaction coefficients for the parameter (aα). A temperature-dependent binary interaction coefficient was introduced into the equation, to give ðaαÞwater ¼ XXh i 0:5 i xwi xwj ai aj αi αj 1 τij (5.162) j where τij ¼ temperature-dependent binary interaction coefficient and xwi ¼ mole fraction of component i in the water phase. Peng and Robinson proposed graphical correlations for determining this parameter for each aqueous binary pair. Lim and Ahmed (1984) expressed these graphical correlations mathematically by the following generalized equation: 556 CHAPTER 5 Equations of State and Phase Equilibria τij ¼ a1 T 2 pci 2 T pci + a3 + a2 pcj Tci Tci pcj (5.163) where T ¼ system temperature, °R Tci ¼ critical temperature of the component of interest, °R pci ¼ critical pressure of the component of interest, psia pcj ¼ critical pressure of the water compound, psia Values of the coefficients a1, a2, and a3 of this polynomial are given in the table below for selected binaries. Component i a1 a2 a3 C1 C2 C3 n-C4 n-C6 0 0 18.032 0 49.472 1.659 2.109 9.441 2.800 5.783 0.761 0.607 1.208 0.488 0.152 For selected nonhydrocarbon components, values of interaction parameters are given by the following expressions: For N2-H2O binary, τij ¼ 0:402ðT=Tci Þ 1:586 (5.164) where τij ¼ binary parameter between nitrogen and the water compound Tci ¼ critical temperature of nitrogen, °R. For CO2-H2O binary, 2 T T 0:503 τij ¼ 0:074 + 0:478 Tci Tci (5.165) where Tci is the critical temperature of CO2. In the course of making phase equilibrium calculations, it is always desirable to provide initial values for the equilibrium ratios so the iterative procedure can proceed as reliably and rapidly as possible. Peng and Robinson adopted Wilson’s equilibrium ratio correlation to provide initial K-values for the hydrocarbon/vapor phase: Ki ¼ pci Tci exp 5:3727ð1 + ωi Þ 1 p T Equation-of-State Applications 557 While for the water-vapor phase, Peng and Robinson proposed the following expression for Kwi: Kwi ¼ 106 ½pci T=ðTci pÞ Vapor Pressure The calculation of the vapor pressure of a pure component through an equation of state usually is made by the same trial-and-error algorithms used to calculate vapor-liquid equilibria of mixtures. Soave (1972) suggests that the van der Waals (vdW), Soave-Redlich-Kwong (SRK), and the PengRobinson (PR) equations of state can be written in the following generalized form: p¼ RT aα V b V 2 + μVb + wb2 (5.166) with R2 Tc2 pc RTc b ¼ Ωb pc a ¼ Ωa where the values of u, w, Ωa, and Ωb for three different equations of state are given in the table below. EOS u w Ωa Ωb vdW SRK PR 0 1 2 0 0 1 0.421875 0.42748 0.45724 0.125 0.08664 0.07780 Soave introduced the reduced pressure, pr, and reduced temperature, Tr, to these equations, to give aαp αpr ¼ Ωa Tr R2 T 2 (5.167) bp pr ¼ Ωb Tr RT (5.168) A Ωa α ¼ B Ωb Tr (5.169) A¼ B¼ and where pr ¼ p=pc Tr ¼ T=Tc 558 CHAPTER 5 Equations of State and Phase Equilibria In the cubic form and in terms of the Z-factor, the three equations of state can be written as vdW : Z3 Z2 ð1 +BÞ + ZA AB ¼0 SRK : Z 3 Z2 + Z A B B2 AB ¼ 0 PR : Z 3 Z2 ð1 BÞ + Z A 3B2 2B AB B2 B2 ¼ 0 And the pure component fugacity coefficient is given by o f A vdW : ln ¼ Z 1 lnðZ BÞ p Z o f A B ¼ Z 1 lnðZ BÞ SRK : ln ln 1 + p B Z pffiffiffi # o " Z+ 1+ 2 B f A pffiffiffi ¼ Z 1 lnðZ BÞ pffiffiffi ln PR : ln p Z 1 2 B 2 2B A typical iterative procedure for the calculation of pure component vapor pressure at any temperature, T, through one of these EOS is summarized next. Step 1. Calculate the reduced temperature; that is, Tr ¼ T/Tc. Step 2. Calculate the ratio A/B from Eq. (5.169). Step 3. Assume a value for B. Step 4. Solve the selected equation of state, such as SRK EOS, to obtain ZL and Z (ie, smallest and largest roots) for both phases. Step 5. Substitute ZL and Zv into the pure component fugacity coefficient and obtain ln(fo/p) for both phases. Step 6. Compare the two values of f/p. If the isofugacity condition is not satisfied, assume a new value of B and repeat steps 3–6. Step 7. From the final value of B, obtain the vapor pressure from Eq. (5.168): B ¼ Ωb ðpv =pc Þ Tr pv ¼ BTr pc Ωb Solving for pv gives SIMULATING OF LABORATORY PVT DATA BY EQUATIONS OF STATE Before attempting to use an EOS-based compositional model in a full-field simulation study, it is essential the selected equation of state is capable of achieving a satisfactory match between EOS results and all the available Simulating of Laboratory PVT Data by Equations of State 559 PVT test data. It should be pointed out that an equation of state generally is not predictive without tuning its parameters to match the relevant experimental data. Several PVT laboratory tests were presented in Chapter 4 and are repeated here for convenience to illustrate the procedure of applying an EOS to simulate these experiments, including n n n n n n n n Constant volume depletion. Constant composition expansion. Differential liberation. Flash separation tests. Composite liberation. Swelling tests. Compositional gradients. Minimum miscibility pressure. Constant Volume Depletion Test A reliable prediction of the pressure depletion performance of a gascondensate reservoir is necessary in determining reserves and evaluating field-separation methods. The predicted performance is also used in planning future operations and studying the economics of projects for increasing liquid recovery by gas cycling. Such predictions can be performed with aid of the experimental data collected by conducting constant volume depletion (CVD) tests on gas condensates. These tests are performed on a reservoir fluid sample in such a manner as to simulate depletion of the actual reservoir, assuming that retrograde liquid appearing during production remains immobile in the reservoir. The CVD test provides five important laboratory measurements that can be used in a variety of reservoir engineering predictions: n n n n n Dew point pressure. Composition changes of the gas phase with pressure depletion. Compressibility factor at reservoir pressure and temperature. Recovery of original in-place hydrocarbons at any pressure. Retrograde condensate accumulation; that is, liquid saturation. The laboratory procedure of the CVD test (with immobile condensate), as shown schematically in Fig. 5.10, is summarized in the following steps. Step 1. A measured amount, m, of a representative sample of the original reservoir fluid with a known overall composition of zi is charged to a visual PVT cell at the dew point pressure, pd (section A). The temperature of the PVT cell is maintained at the reservoir 560 CHAPTER 5 Equations of State and Phase Equilibria Constant volume depletion (CVD) Gas Pd Vi Original gas @ Pd P P Gas Gas Vt Vi Condensate Condensate Hg Hg Hg A B C n FIGURE 5.10 Schematic illustration of the CVD test. temperature, T, throughout the experiment. The initial volume, Vi, of the saturated fluid is used as a reference volume. Step 2. The initial gas compressibility factor is calculated from the real gas equation: Zdew ¼ pd Vi ni RT ni ¼ m Ma (5.170) with where pd ¼ dew point pressure, psia Vi ¼ initial gas volume, ft3 ni ¼ initial number of moles of the gas Ma ¼ apparent molecular weight, lb/lb-mol m ¼ mass of the initial gas in place, lb R ¼ gas constant, 10.73 T ¼ temperature, °R zd ¼ compressibility factor at dew-point pressure with the gas initially in place as expressed in standard units, scf, given by G ¼ 379:4ni (5.171) Simulating of Laboratory PVT Data by Equations of State 561 Step 3. The cell pressure is reduced from the saturation pressure to a predetermined level, p. This can be achieved by withdrawing mercury from the cell, as illustrated in Fig. 5.10, section B. During the process, a second phase (retrograde liquid) is formed. The fluid in the cell is brought to equilibrium and the total volume, Vt, and volume of the retrograde liquid, VL, are visually measured. This retrograde volume is reported as a percent of the initial volume, Vi: SL ¼ VL 100 Vi This value can be translated into condensate saturation in the presence of water saturation by So ¼ ð1 Sw ÞSL =100 Step 4. Mercury is reinjected into the PVT cell at constant pressure p while an equivalent volume of gas is removed. When the initial volume, Vi, is reached, mercury injection ceases, as illustrated in Fig. 5.10, section C. The volume of the removed gas is measured at the cell conditions and recorded as (Vgp)p,T. This step simulates a reservoir producing only gas, with retrograde liquid remaining immobile in the reservoir. Step 5. The removed gas is charged to analytical equipment, where its composition, yi, is determined and its volume is measured at standard conditions and recorded as (Vgp)sc. The corresponding moles of gas produced can be calculated from the expression Vgp sc np ¼ 379:4 (5.172) where np ¼ moles of gas produced (Vgp)sc ¼ volume of gas produced measured at standard conditions, scf. Step 6. The compressibility factor of the gas phase at cell pressure and temperature is calculated by Vgp p, T Vactual Z¼ ¼ Videal Videal where Videal ¼ RTnp p (5.173) 562 CHAPTER 5 Equations of State and Phase Equilibria Combining Eqs. (5.172), (5.173) and solving for the compressibility factor Z gives 379:4p Vgp p, T Z¼ RT Vgp sc (5.174) Another property, the two-phase compressibility factor, also is calculated. The two-phase Z-factor represents the total compressibility of all the remaining fluid (gas and retrograde liquid) in the cell and is computed from the real gas law as Ztwo-phase ¼ pVi ni np ðRT Þ (5.175) where (ni np) ¼ represents the remaining moles of fluid in the cell ni ¼ initial moles in the cell np ¼ cumulative moles of gas removed Step 7. The volume of gas produced as a percentage of gas initially in place is calculated by dividing the cumulative volume of the produced gas by the gas initially in place, both at standard conditions: "X %Gp ¼ # Vgp sc G 100 (5.176) or equivalently as " X # np 100 %Gp ¼ ðni Þoriginal This experimental procedure is repeated several times, until a minimum test pressure is reached, after which the quantity and composition of the gas and retrograde liquid remaining in the cell are determined. This test procedure also can be conducted on a volatile oil sample. In this case, the PVT cell initially contains liquid, instead of gas, at its bubble point pressure. Simulating of the CVD Test by EOS In the absence of the CVD test data on a specific gas-condensate system, predictions of pressure-depletion behavior can be obtained by using any of the well-established equations of state to compute the phase behavior when the composition of the total gas-condensate system is known. The stepwise computational procedure using the Peng-Robinson EOS as a Simulating of Laboratory PVT Data by Equations of State 563 representative equation of state now is summarized in conjunction with the flow diagram shown in Fig. 5.11. Step 1. Assume that the original hydrocarbon system with a total composition zi occupies an initial volume of 1 ft3 at the dew point pressure pd and system temperature T: Vi ¼ 1 Step 2. Using the initial gas composition, calculate the gas compressibility factor Zdew at the dew point pressure from Eq. (5.115): Z3 + ðB 1ÞZ 2 + A 3B2 2B Z AB B2 B3 ¼ 0 Given: zi, Pd, T Calculate Zdew from: Z3 + (B − 1) Z2 + (A − 3 B2 − 2B)Z − (AB − B2 − B3) = 0 Assume Vi =1 ft3 Vp Calculate ni ni = ZRT = (1) (pd) Zdew TR Assume depletion pressure “P” Calculate Ki for zi @ P see Fig. 15-11 VL = (nL)actual ZL RT P Solution; gives: Ki, xi, yi, nL, nV, ZL, ZV Calculate volume of condensate VL Calculate volume of gas Vg Calculate TOTAL volume Vt = Vg + VL Determine moles of gas removed nP & remaining nr xi (nL)actual + yi nr zi = (nL)actual + nr n FIGURE 5.11 Flow diagram for EOS simulating CVD test. Combine the remaining phase and determine new Zi Vg = (nV)actual ZV RT (Vg)P = Vt − 1 np = P(Vg)P ZL RT nr = (nV)actual − np P 564 CHAPTER 5 Equations of State and Phase Equilibria Step 3. Calculate the initial moles in place by applying the real gas law: ni ¼ ð1Þðpd Þ Zdew RT (5.177) where pd ¼ dew point pressure, psi Zdew ¼ gas compressibility factor at the dew point pressure. Step 4. Reduce the pressure to a predetermined value p. At this pressure level, the equilibrium ratios (K-values) are calculated from EOS. Results of the calculation at this stage include n n n n n n n Equilibrium ratios (K-values). Composition of the liquid phase (ie, retrograde liquid), xi. Moles of the liquid phase, nL. Composition of the gas phase, yi. Moles of the gas phase, nυ. Compressibility factor of the liquid phase, ZL. Compressibility factor of the gas phase, Zv. The composition of the gas phase, yi, should reasonably match the experimental gas composition at pressure, p. Step 5. Because flash calculations, as part of the K-value results, are usually performed assuming the total moles are equal to 1, calculate the actual moles of the liquid and gas phases from ðnL Þactual ¼ ni nL (5.178) ðnυ Þactual ¼ ni nυ (5.179) where nL and nυ are moles of liquid and gas, respectively, as determined from flash calculations. Step 6. Calculate the volume of each hydrocarbon phase by applying the expression VL ¼ ðnL Þactual ZL RT p (5.180) Vg ¼ ðnv Þactual Zv RT p (5.181) where VL ¼ volume of the retrograde liquid, ft3/ft3 Vg ¼ volume of the gas phase, ft3/ft3 Simulating of Laboratory PVT Data by Equations of State 565 ZL, Zv ¼ compressibility factors of the liquid and gas phase T ¼ cell (reservoir) temperature, °R As Vi ¼ 1, then SL ¼ ðVL Þ100 This value should match the experimental value if the equation of state is properly tuned. Step 7. Calculate the total volume of fluid in the cell: Vt ¼ VL + Vg (5.182) where Vt ¼ total volume of the fluid, ft3. Step 8. As the volume of the cell is constant at 1 ft3, remove the following excess gas volume from the cell: Vgp p, T ¼ Vt 1 (5.183) Step 9. Calculate the number of moles of gas removed: np ¼ p Vgp p, T (5.184) Zv RT Step 10. Calculate cumulative gas produced as a percentage of gas initially P in place by dividing cumulative moles of gas removed, np, by the original moles in place: %Gp ¼ " X # np ðni Þoriginal 100 Step 11. Calculate the two-phase gas deviation factor from the relationship: Ztwo-phase ¼ ðpÞð1Þ ni np RT (5.185) Step 12. Calculate the remaining moles of gas (nυ)r by subtracting the moles produced (np) from the actual number of moles of the gas phase (nυ)actual: ðnυ Þr ¼ ðnυ Þactual np (5.186) 566 CHAPTER 5 Equations of State and Phase Equilibria Step 13. Calculate the new total moles and new composition remaining in the cell by applying the molal and component balances, respectively: ni ¼ ðnL Þactual + ðnυ Þr ; (5.187) xi ðnL Þactual + yi ðnυ Þr ni (5.188) zi ¼ Step 14. Consider a new lower pressure and repeat steps 4–13. Constant Composition Expansion Test Constant composition expansion experiments, commonly called pressure/ volume tests, are performed on gas condensates or crude oil to simulate the pressure/volume relations of these hydrocarbon systems. The test is conducted for the purposes of determining Saturation pressure (bubble point or dew-point pressure). Isothermal compressibility coefficients of the single-phase fluid in excess of saturation pressure. Compressibility factors of the gas phase. Total hydrocarbon volume as a function of pressure. n n n n The experimental procedure, shown schematically in Fig. 5.12, involves placing a hydrocarbon fluid sample (oil or gas) in a visual PVT cell at Voil = 100% P2 > Pb P1 >> Pb 100% Oil Voil ≈ 100% Voil = 100% Vt1 Oil Voil < 100% P3 = Pb Vt2 Oil Voil << 100% P4 << Pb P5 <<< Pb Gas Gas Vsat Oil Vt4 Oil Hg (a) (b) n FIGURE 5.12 Constant composition expansion test. (c) (d) (e) Vt5 Simulating of Laboratory PVT Data by Equations of State 567 reservoir temperature and a pressure in excess of the initial reservoir pressure (Fig. 5.12, section A). The pressure is reduced in steps at constant temperature by removing mercury from the cell, and the change in the hydrocarbon volume, V, is measured for each pressure increment. The saturation pressure (bubble point or dew point pressure) and the corresponding volume are observed and recorded (Fig. 5.12, section B). The volume at the saturation pressure is used as a reference volume. At pressure levels higher than the saturation pressure, the volume of the hydrocarbon system is recorded as a ratio of the reference volume. This volume, commonly termed the relative volume, is expressed mathematically by the following equation: Vrel ¼ V Vsat (5.189) where Vrel ¼ relative volume V ¼ volume of the hydrocarbon system Vsat ¼ volume at the saturation pressure Also, above the saturation pressure, the isothermal compressibility coefficient of the single-phase fluid usually is determined from the expression c¼ 1 @Vrel Vrel @p T (5.190) where c ¼ isothermal compressibility coefficient, psi1. For gas-condensate systems, the gas compressibility factor, Z, is determined in addition to the preceding experimentally determined properties. Below the saturation pressure, the two-phase volume, Vt, is measured relative to the volume at saturation pressure and expressed as Relativetotalvolume ¼ Vt Vsat (5.191) where Vt ¼ total hydrocarbon volume. Note that no hydrocarbon material is removed from the cell; therefore, the composition of the total hydrocarbon mixture in the cell remains fixed at the original composition. 568 CHAPTER 5 Equations of State and Phase Equilibria Simulating of the CCE Test by EOS The simulation procedure utilizing the Peng-Robinson equation of state is illustrated in the following steps. Step 1. Given the total composition of the hydrocarbon system, zi, and saturation pressure (pb for oil systems, pd for gas systems), calculate the total volume occupied by 1 mol of the system. This volume corresponds to the reference volume Vsat (volume at the saturation pressure). Mathematically, the volume is calculated from the relationship Vsat ¼ ð1ÞZRT psat (5.192) where Vsat ¼ volume of saturation pressure, ft3/mol psat ¼ saturation pressure (dew point or bubble point pressure), psia T ¼ system temperature, °R Z ¼ compressibility factor, ZL or Zv depending on the type of system Step 2. The pressure is increased in steps above the saturation pressure, where the single phase still exists. At each pressure, the compressibility factor, ZL or Zv, is calculated by solving Eq. (5.115) and used to determine the fluid volume. V¼ ð1ÞZRT p where V ¼ compressed liquid or gas volume at the pressure level p, ft3/mol Z ¼ compressibility factor of the compressed liquid or gas, ZL or Zv p ¼ system pressure, p > psat, psia R ¼ gas constant; 10.73 psi ft3/lb-mol °R The corresponding relative phase volume, Vrel, is calculated from the expression Vrel ¼ V Vsat Step 3. The pressure is then reduced in steps below the saturation pressure psat, where two phases are formed. The equilibrium ratios are calculated and flash calculations are performed at each pressure level. Results of the calculation at each pressure level include Ki, xi, yi, nL, nv, Zv, and ZL. Because no hydrocarbon material is removed during pressure depletion, the original moles (ni ¼ 1) and Simulating of Laboratory PVT Data by Equations of State 569 composition, zi, remain constant. The volumes of the liquid and gas phases can then be calculated from the expressions VL ¼ ð1ÞðnL ÞZL RT p (5.193) Vg ¼ ð1Þðnυ ÞZ v RT p (5.194) and Vt ¼ VL + Vg where nL, nυ ¼ moles of liquid and gas as calculated from flash calculations ZL, Zv ¼ liquid and gas compressibility factors Vt ¼ total volume of the hydrocarbon system Step 4. Calculate the relative total volume from the following expression: Relativetotalvolume ¼ Vt Vsat Differential Liberation Test The test is carried out on reservoir oil samples and involves charging a visual PVT cell with a liquid sample at the bubble point pressure and reservoir temperature, as described in detail in Chapter 4. The pressure is reduced in steps, usually 10–15 pressure levels, and all the liberated gas is removed and its volume, Gp, is measured under standard conditions. The volume of oil remaining, VL, also is measured at each pressure level. Note that the remaining oil is subjected to continual compositional changes as it becomes progressively richer in the heavier components. This procedure is continued to atmospheric pressure, where the volume of the residual (remaining) oil is measured and converted to a volume at 60°F, Vsc. The differential oil formation volume factors Bod (commonly called the relative oil volume factors) at all the various pressure levels are calculated by dividing the recorded oil volumes, VL, by the volume of residual oil, Vsc: Bod ¼ VL Vsc (5.195) The differential solution gas-oil ratio Rsd is calculated by dividing the volume of gas in the solution by the residual oil volume. Typical table and in Fig. 5.13. 570 CHAPTER 5 Equations of State and Phase Equilibria Bod Rsd B od R sd Pressure n FIGURE 5.13 Bod and Rsd versus pressure. Pressure (psig) Rsda Bodb Shrinkage (bbl/bbl) 2620 2350 2100 1850 1600 1350 1100 850 600 350 159 0 0 854 763 684 612 544 479 416 354 292 223 157 0 0 1.6 1.554 1.515 1.479 1.445 1.412 1.382 1.351 1.32 1.283 1.244 1.075 at 60°F ¼ 1.000 1.0000 0.9713 0.9469 0.9244 0.9031 0.8825 0.8638 0.8444 0.8250 0.8019 0.7775 0.6250 a b Cubic feet of gas at 14.65 psia and 60°F per barrel of residual oil at 60°F. Barrels of oil at indicated pressure and temperature per barrel of residual oil at 60°F. It should be pointed out that reporting the experimental data relative to the residual oil volume at 60°F gives the relative oil volume curve the appearance of the formation volume factor curve, leading to its misuse in reservoir calculations. A better method of reporting these data is in the form of a shrinkage curve, as shown in Fig. 5.14. The relative-oil volume data in Fig. 5.13 can be converted to a shrinkage curve by dividing each Sod Simulating of Laboratory PVT Data by Equations of State 571 Pressure n FIGURE 5.14 Oil shrinkage curve versus pressure. relative-oil volume factor Bod by the relative-oil volume factor at the bubble point, Bodb: Sod ¼ Bod Bodb (5.196) where Bod ¼ differential relative-oil volume factor at pressure p, bbl/STB Bodb ¼ differential relative-oil volume factor at the bubble point pressure, pb, psia, bbl/STB Sod ¼ differential oil shrinkage factor, bbl/bbl, of bubble point oil The shrinkage curve has a value of 1 at the bubble point and a value less than 1 at subsequent pressures below pb. In this suggested form, the shrinkage describes the changes in the volume of the bubble point oil as reservoir pressure declines. The results of the differential liberation test, when combined with the results of the flash separation test, provide a means of calculating the oil formation volume factors and gas solubility as a function of reservoir pressure. These calculations are outlined in the next section. Simulating the Differential Liberation Test by EOS The simulation procedure of the test by the Peng-Robinson EOS is summarized in the following steps and in conjunction with the flow diagram shown in Fig. 5.15. 572 CHAPTER 5 Equations of State and Phase Equilibria Given: zi, Pb, T Calculate ZL from: Z3 + (B − 1) Z2 + (A − 3B2 − 2B)Z − (AB − B2 − B3) = 0 Assume ni = 1 mole Calculate: Vsat (1) Z LRT nZRT Vsat = P = Pb Assume depletion pressure P < Pb Calculate Ki for zi @ P See Fig. 15-11 VL = (nL)actual ZLRT P Calculate volume of condensate VL Solution gives: Ki, xi, yi, nL, nV, ZL, ZV Calculate volume of gas Vg (nV)actual ZVRT P GP = 379.4 (nV)actual (GP)total = Set: zi = xi and ni = nL No Vg = ΣG p i=1 Last pressure ? Calculate Bod & Rsd n FIGURE 5.15 Flow diagram for EOS simulating DE test. Step 1. Starting with the saturation pressure, psat, and reservoir temperature, T, calculate the volume occupied by a total of 1 mol (ie, ni ¼ 1) of the hydrocarbon system with an overall composition of zi. This volume, Vsat, is calculated by applying Eq. (5.192): Vsat ¼ ð1ÞZRT psat Step 2. Reduce the pressure to a predetermined value of p at which the equilibrium ratios are calculated and used in performing flash calculations. The actual number of moles of the liquid phase, with a composition of xi, and the actual number of moles of the gas phase, with a composition of yi, then are calculated from the expressions Simulating of Laboratory PVT Data by Equations of State 573 ðnL Þactual ¼ ni nL ðnυ Þactual ¼ ni nυ where (nL)actual ¼ actual number of moles of the liquid phase (nυ)actual ¼ actual number of moles of the gas phase nL, nυ ¼ number of moles of the liquid and gas phases as computed by performing flash calculations Step 3. Determine the volume of the liquid and gas phase from VL ¼ ZL RT ðnL Þactual p (5.197) Vg ¼ Zv RT ðnυ Þactual p (5.198) where VL, Vg ¼ volumes of the liquid and gas phases, ft3 ZL, Zv ¼ compressibility factors of the liquid and gas phases. The volume of the produced (liberated) solution gas as measured at standard conditions is determined from the relationship Gp ¼ 379:4ðnυ Þactual (5.199) where Gp ¼ gas produced during depletion pressure p, scf. The total cumulative gas produced at any depletion pressure, p, is the cumulative gas liberated from the crude oil sample during the pressure reduction process (sum of all the liberated gas from previous pressures and current pressure) as calculated from the expression p X Gp p ¼ Gp psat where (Gp)p is the cumulative produced solution gas at pressure level p. Step 4. Assume that all the equilibrium gas is removed from contact with the oil. This can be achieved mathematically by setting the overall composition, zi, equal to the composition of the liquid phase, xi, and setting the total moles equal to the liquid phase: zi ¼ x i ni ¼ ðnL Þactual Step 5. Using the new overall composition and total moles, steps 2–4 are repeated. When the depletion pressure reaches the atmospheric pressure, the temperature is changed to 60°F and the residual-oil 574 CHAPTER 5 Equations of State and Phase Equilibria volume is calculated. This residual-oil volume, calculated by Eq. (5.197), is referred to as Vsc. The total volume of gas that evolved from the oil and produced (Gp)Total is the sum of the all-liberated gas including that at atmospheric pressure: 14:7 X Gp Total ¼ Gp psat Step 6. The calculated volumes of the oil and removed gas then are divided by the residual-oil volume to calculate the relative-oil volumes (Bod) and the solution GOR at all the selected pressure levels from Bod ¼ VL Vsc h i ð5:615Þðremaining gas in solutionÞ ð5:615Þ Gp Total Gp p Rsd ¼ ¼ Vsc Vsc (5.200) Flash Separation Tests Flash separation tests, commonly called separator tests, are conducted to determine the changes in the volumetric behavior of the reservoir fluid as the fluid passes through the separator (or separators) and then into the stock tank. The resulting volumetric behavior is influenced to a large extent by the operating conditions; that is, pressures and temperatures of the surface separation facilities. The primary objective of conducting separator tests, therefore, is to provide the essential laboratory information necessary for determining the optimum surface separation conditions, which in turn maximize the stock-tank oil production. In addition, the results of the test, when appropriately combined with the differential liberation test data, provide a means of obtaining the PVT parameters (Bo, Rs, and Bt) required for petroleum engineering calculations. The laboratory procedure of the flash separation test involves the following steps. Step 1. The oil sample is charged into a PVT cell at reservoir temperature and its bubble point pressure. Step 2. A small amount of the oil sample, with a volume of (Vo)pb, is removed from the PVT cell at constant pressure and flashed through a multistage separator system, with each separator at a fixed pressure and temperature. The gas liberated from each stage (separator) is removed and its volume and specific gravity measured. The volume of the remaining oil in the last stage (stock-tank condition) is recorded as (Vo)st. Simulating of Laboratory PVT Data by Equations of State 575 Step 3. The total solution gas-oil ratio and the oil formation volume factor at the bubble point pressure are then calculated: Bofb ¼ ðVo Þpb =ðVo Þst Rsfb ¼ Vg sc =ðVo Þst (5.201) (5.202) where Bofb ¼ bubble point oil formation volume factor, as measured by flash liberation, bbl of the bubble point oil/STB Rsfb ¼ bubble point solution gas-oil ratio, as measured by flash liberation, scf/STB (Vg)sc ¼ total volume of gas removed from separators, scf Step 4. Repeat steps 1–3. A typical example of a set of separator tests for a two-stage separation is shown in the table below. Separator pressure (psig) Temperature (°F) Rsfba °API Bofbb 50 to 0 75 75 40.5 1.481 100 to 0 75 75 40.7 1.474 200 to 0 75 75 40.4 1.483 300 to 0 75 75 737 41 P ¼ 778 676 92 P ¼ 768 602 178 P ¼ 780 549 246 P ¼ 795 40.1 1.495 a GOR in cubic feet of gas at 14.65 psia and 60°F per barrel of stock-tank oil at 60°F. FVF in barrels of saturated oil at 2620 psig and 220°F per barrel of stock-tank oil at 60°F. b The differential liberation data must be adjusted to account for the separation process taking place at the separator condition. The procedure, as discussed in Chapter 4, is based on multiplying the flash oil formation factor at the bubble point, Bofb, by the differential oil shrinkage factor Sod (as defined by Eq. (5.196)) at various reservoir pressures. Mathematically, this relationship is expressed as follows: Bo ¼ Bofb Sod (5.203) 576 CHAPTER 5 Equations of State and Phase Equilibria where Bo ¼ oil formation volume factor, bbl/STB Bofb ¼ bubble point oil formation volume factor, bbl of the bubble point oil/STB Sod ¼ differential oil shrinkage factor, bbl/bbl of bubble point oil The adjusted differential gas solubility data are given by Rs ¼ Rsfb ðRsdb Rsd Þ where Bofb Bodb (5.204) Rs ¼ gas solubility, scf/STB Rsfb ¼ bubble point solution gas-oil ratio as defined by Eq. (5.202), scf/ STB Rsdb ¼ solution gas-oil ratio at the bubble point pressure as measured by the differential liberation test, scf/STB Rsd ¼ solution gas-oil ratio at various pressure levels as measured by the differential liberation test, scf/STB These adjustments typically produce lower formation volume factors and gas solubilities than the differential liberation data. This adjustment procedure is illustrated graphically in Figs. 5.16 and 5.17. n FIGURE 5.16 Adjustment of oil-volume curve to separate conditions. Simulating of Laboratory PVT Data by Equations of State 577 n FIGURE 5.17 Adjustment of gas-in-solution curve to separator conditions. Composite Liberation Test The laboratory procedures for conducting the differential liberation and flash liberation tests for the purpose of generating Bo and Rs versus p relationships are considered approximations to the actual relationships. Another type of test, called a composite liberation test, has been suggested by Dodson et al. (1953) and represents a combination of differential and flash liberation processes. The laboratory test, commonly called the Dodson test, provides a better means of describing the PVT relationships. The experimental procedure for this composite liberation, as proposed by Dodson et al., is summarized in the following steps. Step 1. A large representative fluid sample is placed into a cell at a pressure higher than the bubble point pressure of the sample. The temperature of the cell then is raised to reservoir temperature. Step 2. The pressure is reduced in steps by removing mercury from the cell, and the change in the oil volume is recorded. The process is repeated until the bubble point of the hydrocarbon is reached. 578 CHAPTER 5 Equations of State and Phase Equilibria Step 3. A carefully measured small volume of oil is removed from the cell at constant pressure and flashed at temperatures and pressures equal to those in the surface separators and stock tank. The liberated gas volume and stock-tank oil volume are measured. The oil formation volume factor, Bo, and gas solubility, Rs, then are calculated from the measured volumes: Bo ¼ ðVo Þp, T =ðVo Þst Rs ¼ Vg sc =ðVo Þst (5.205) (5.206) where (Vo)p,T ¼ volume of oil removed from the cell at constant pressure p (Vg)sc ¼ total volume of the liberated gas as measured under standard conditions (Vo)st ¼ volume of the stock-tank oil Step 4. The volume of the oil remaining in the cell is allowed to expand through a pressure decrement and the gas evolved is removed, as in the differential liberation. Step 5. Following the gas removal, step 3 is repeated and Bo and Rs calculated. Step 6. Steps 3–5 are repeated at several progressively lower reservoir pressures to secure a complete PVT relationship. Note that this type of test, while more accurately representing the PVT behavior of complex hydrocarbon systems, is more difficult and costly to perform than other liberation tests. Consequently, these experiments usually are not included in a routine fluid property analysis. Simulating of the Composite Liberation Test by EOS The stepwise computational procedure for simulating the composite liberation test using the Peng-Robinson EOS is summarized next in conjunction with the flow chart shown in Fig. 5.18. Step 1. Assume a large reservoir fluid sample with an overall composition of zi is placed in a cell at the bubble point pressure, pb, and reservoir temperature, T. Step 2 Remove 1 lb-mol (ie, ni ¼ 1) of the liquid from the cell and calculate the corresponding volume at pb and T by applying the PengRobinson EOS to estimate the liquid compressibility factor, ZL: ðVo Þp, T ¼ ð1ÞZL RT pb (5.207) Simulating of Laboratory PVT Data by Equations of State 579 Given: zi, Pb, T Assume n = 1 mole Calculate: VP,T Flash zi at surface separation conditions. See Fig.15-11 Calculate Z L from: Z3 + (B − 1) Z2 + (A − 3B2 − 2B)Z − (AB − B2 − B3) = 0 (1) ZLRT nZRT VP,T = P = P Solution gives: Ki, xi, yi, nL, nV, ZL, ZV Calculate B0 and Rs Assume P < Pb and Perform flash calculations Set: zi = xi and ni = 1 P No Last pressure ? Yes Results Bo & Rs n FIGURE 5.18 Flow diagram for EOS simulating composite liberation test. where (Vo)p,T ¼ volume of 1 mol of the oil, ft3. Step 3. Flash the resulting volume (step 2) at temperatures and pressures equal to those in the surface separators and stock tank. Designating the resulting total liberated gas volume (Vg)sc and the volume of the stock-tank oil (Vo)st, calculate the Bo and Rs by applying Eqs. (5.205), (5.206), respectively: Bo ¼ ðVo Þp, T =ðVo Þst Rs ¼ Vg sc =ðVo Þst Step 4. Set the cell pressure to a lower pressure level, p, and using the original composition, zi, generate the Ki values through the application of the PR EOS. 580 CHAPTER 5 Equations of State and Phase Equilibria Step 5. Perform flash calculations based on zi and the calculated ki values. Step 6. To simulate the differential liberation step of removing the equilibrium liberated gas from the cell at constant pressure, simply set the overall composition, zi, equal to the mole fraction of the equilibrium liquid, x, and the bubble point pressure, pb, equal to the new pressure level, p. Mathematically, this step is summarized by the following relationships: zi ¼ x i pb ¼ p Step 7. Repeat steps 2–6. Swelling Tests Swelling tests should be performed if a reservoir is to be depleted under gas injection or a dry gas cycling scheme. The swelling test can be conducted on gas-condensate or crude oil samples. The purpose of this laboratory experiment is to determine the degree to which the proposed injection gas will dissolve in the hydrocarbon mixture. The data that can be obtained during a test of this type include n n The relationship of saturation pressure and volume of gas injected. The volume of the saturated fluid mixture compared to the volume of the original saturated fluid. The test procedure, as shown schematically in Fig. 5.19, is summarized in the following steps. Step 1. A representative sample of the hydrocarbon mixture with a known overall composition, zi, is charged to a visual PVT cell at saturation pressure and reservoir temperature. The volume at the saturation pressure is recorded and designated (Vsat)orig. Step 2. A predetermined volume of gas with a composition similar to the proposed injection gas is added to the hydrocarbon mixture. The cell is pressured up until only one phase is present. The new saturation pressure and volume are recorded and designated ps and Vsat, respectively. The original saturation volume is used as a reference value and the results are presented as relative total volumes: Vrel ¼ Vsat =ðVsat Þorig where Vrel ¼ relative total volume (Vsat)orig ¼ original saturation volume. Simulating of Laboratory PVT Data by Equations of State 581 Swelling test Injection gas Psat (Vsat)orig Original fluid (gas or oil) @ saturation pressure Psat (Psat)new ie Pb or Pd P Vt Hg Original fluid + injection gas B n FIGURE 5.19 Schematic illustration of the Swelling test. Step 3. Repeat step 2 until the mole percent of the injected gas in the fluid sample reaches a preset value (around 80%). Typical laboratory results of a gas-condensate swelling test with lean gas of a composition are shown Figs. 5.20 and 5.21. These two figures show the gradual increase in the saturation pressure and the swollen volume with each addition of the injection lean gas. As shown in Figs. 5.20 and 5.21, the dew point increased from 3428 to 4880 psig and relative total volume from 1 to 2.5043 after the final addition. Simulating of the Swelling Test by EOS The simulation procedure of the swelling test by the PR EOS is summarized in the following steps. Step 1. Assume the hydrocarbon system occupies a total volume of 1 bbl at the saturation pressure, ps, and reservoir temperature, T. Calculate the initial moles of the hydrocarbon system from the following expression: ni ¼ 5:615psat ZRT Original fluid + injection gas Hg Hg A (Vsat)new C 582 CHAPTER 5 Equations of State and Phase Equilibria 5000 Injected gas 4800 New dewpoint pressure 4600 Component yi CO2 0 N2 0 4400 C1 0.9468 4200 C2 0.0527 C3 0.0005 4000 3800 Cum. gas inj, scf/bbl 3600 0 1 3428 1.1224 3635 572 1.3542 4015 1523 1.9248 4610 2467 2.5043 4880 3200 0 500 1000 1500 2000 2500 3000 Dewpoint pressure, psig 190 3400 3000 Swollen volume, bbl/bbl scf/bbl of dewpoint gas n FIGURE 5.20 Dew point pressure during swelling of reservoir gas with lean gas at 200°F. Relative volume vs. volume of gas injected 2.6 Injected gas Swelling ratio, (Vsat)new/(Vsat)orig 2.4 2.2 2 1.8 Component yi CO2 0 N2 0 C1 0.9468 C2 0.0527 C3 0.0005 1.6 Cum. gas inj, scf/bbl 1.4 1.2 1 0 500 1000 1500 2000 scf/bbl of the dewpoint gas 2500 3000 Swollen volume, bbl/bbl Dewpoint pressure, psig 0 1 3428 190 1.1224 3635 572 1.3542 4015 1523 1.9248 4610 2467 2.5043 4880 n FIGURE 5.21 Reservoir gas swelling with lean gas at 200°F as expressed in terms. where ni ¼ initial moles of the hydrocarbon system psat ¼ saturation pressure “pb or pd” depending on the type of the hydrocarbon system, psia Z ¼ gas or liquid compressibility factor Step 2. Given the composition yinj of the proposed injection gas, add a predetermined volume (as measured in scf) of the injection gas to Simulating of Laboratory PVT Data by Equations of State 583 the original hydrocarbon system and calculate the new overall composition by applying the molal and component balances, respectively: nt ¼ ni + ninj , zi ¼ yinj ninj + ðzsat Þi ni nt with ninj ¼ Vinj 379:4 where ninj ¼ total moles of the injection gas Vinj ¼ volume of the injection gas, scf (zsat)i ¼ mole fraction of component i in the saturated hydrocarbon system Step 3. Using the new composition, zi, calculate the new saturation pressure, pb or pd, using the Peng-Robinson EOS as outlined in this chapter. Step 4. Calculate the relative total volume (swollen volume) by applying the relationship Vrel ¼ Znt RT 5:615psat where Vrel ¼ swollen volume psat ¼ saturation pressure. Step 5. Steps 2–4 are repeated until the last predetermined volume of the lean gas is combined with the original hydrocarbon system. Compositional Gradients Vertical compositional variation within a reservoir is expected during early reservoir life. One might expect the reservoir fluids to have attained equilibrium at maturity due to molecular diffusion and mixing over geological time. However, the diffusive mixing may require many tens of million years to eliminate compositional heterogeneities. When the reservoir is considered mature, it often is assumed that fluids are at equilibrium at uniform temperature and pressure and the thermal and gravitational equilibrium are established, which lead to the uniformity of each component fugacity throughout all the coexisting phases. For a single phase, the uniformity of fugacity is equivalent to the uniformity of concentration. 584 CHAPTER 5 Equations of State and Phase Equilibria The pressure and temperature, however, are not uniform throughout a reservoir. The temperature increases with depth with a gradient of about 1°F/100 ft. The pressure also changes according to the hydrostatic head of fluid in the formation. Therefore, compositional variations within a reservoir, particularly those with a tall column of fluid, should be expected. Table 5.1 shows the fluid compositional variation of a North Sea reservoir at different depths. Note that the methane concentration decreases from 72.30 to 54.92 mol% over a depth interval of only 81 m, indicating a major change in composition that must be considered when estimating reserves and production planning. In general, the hydrocarbon mixture is expected to get richer in heavier compounds and contains less of light components (such as methane) with increasing depth. The variations in the composition and temperature of fluid with depth will result in changes of saturation pressure as a function of depth. In the oil column, the bubble point pressure generally decreases with Table 5.1 Variations in Fluid Composition with Depth in a Reservoir Fluid D, Well 1 C, Well 2 B, Well 2 A, Well 2 Depth (meters subsea) Nitrogen Carbon dioxide Methane Ethane Propane i-Butane n-Butane n-Pentane Hexanes Heptanes Octanes Nonanes Decanes Undecanes-plus Molecular weight Undecanes-plus characteristics Molecular weight Specific gravity 3136 0.65 2.56 72.3 8.79 4.83 0.61 1.79 0.75 0.86 1.13 0.92 0.54 0.28 3.49 33.1 3156 0.59 2.48 64.18 8.85 5.6 0.68 2.07 0.94 1.24 2.14 2.18 1.51 0.91 6 43.6 3181 0.6 2.46 59.12 8.18 5.5 0.66 2.09 1.09 1.49 3.18 2.75 1.88 1.08 9.25 55.4 3217 0.53 2.44 54.92 9.02 6.04 0.74 2.47 1.33 1.71 3.15 2.96 2.03 1.22 10.62 61 260 0.848 267 0.8625 285 0.8722 290 0.8768 Simulating of Laboratory PVT Data by Equations of State 585 7500. De wp oin o Reserv tp 7600. re ss ur e ir press 7700. p g 144 g 7800. dient Depth/ft ; psi / ft ure gra D 7900. Sat P GOC Res P pd = pb = pi 8000. 8100. 3300 3400 p o D 144 3500 o ; psi / ft 3600 t oin -p e ble sur b s Bu pre 3700 3800 3900 Pressure/psia n FIGURE 5.22 Determination of GOC from pressure gradients. decreasing methane concentration, whereas the gas condensate dew point pressure increases with increasing heavy fractions, as shown in Fig. 5.22. As illustrated in Fig. 5.22, the GOC is defined as the depth at which the fluid system changes from a mixture with a dew point to a mixture with a bubble point. This may occur at a saturated condition, in which the GOC “gas” is in thermodynamic equilibrium with the GOC “oil” at the gas-oil contact reservoir pressure. By this definition, the following equality applies: pb ¼ pd ¼ pGOC where pb ¼ bubble point pressure pd ¼ dew point pressure pGOC ¼ reservoir pressure at GOC Fig. 5.23 shows an undersaturated GOC that describes the transition from dew point gas to bubble point oil through a “critical” or “new-critical” mixture with critical temperatures that equal the reservoir temperature. However, the critical pressure of this critical GOC mixture is lower than reservoir pressure. 586 CHAPTER 5 Equations of State and Phase Equilibria Pressure & temperature in Dewpo t press Retrograde gas-cap ure Depth re pres oint blep Bub on ure carb dro press peratu oir tem f hy TCo rvoir Reserv Oil rim Rese sure A compositional transition zone that contains a near critical hydrocarbon mixture that separate gas-cap gas from oil n FIGURE 5.23 Concept of undersaturated GOC; near critical mixture separates gas-cap gas from oil rim. Minimum Miscibility Pressure Several schemes for predicting MMP based on the equation-of-state approach have been proposed. Benmekki and Mansoori (1988) applied the PR EOS with different mixing rules that were joined with a newly formulated expression for the unlike three-body interactions between the injection gas and the reservoir fluid. Several authors used upward pressure increments to locate the MMP, taking advantage through extrapolation of the fact the K-values approach unity as the limiting coincident tie line is approached. A simplified scheme proposed by Ahmed (1997) is based on the fact that the MMP for a solvent/oil mixture generally is considered to correspond to the critical point at which the K-values for all components converge to unity. The following function, miscibility function, is found to provide accurate miscibility criteria as the solvent/hydrocarbon mixture approaches the critical point. This miscibility function, as defined next, approaches 0 or a very small value as the pressure reaches the MMP: FM ¼ " !# n X ½Φðyi Þvi 0 zi 1 ½Φðzi ÞLi i¼1 where FM ¼ miscibility function and Φ ¼ fugacity coefficient. This expression shows that, as the overall composition, zi, changes by gas injection and reaches the critical composition, the miscibility function is Tuning EOS Parameters 587 monotonically decreasing and approaching 0 or a negative value. The specific steps of applying the miscibility function follow. Step 1. Select a convenient reference volume (eg, one barrel) of the crude oil with a specified initial composition and temperature. Step 2. Combine an incremental volume of the injection gas with the crude oil system and determine the overall composition, zi. Step 3. Calculate the bubble point pressure, pb, of the new composition by applying the PR EOS (as given by Eq. (5.145)) Step 4. Evaluate the miscibility function, FM, at this pressure using the calculated values Φ(zi), Φ(yi), and the overall composition at the bubble point pressure. Step 5. If the value of the miscibility function is small, then the calculated pb is approximately equal to the MMP. Otherwise, repeat steps 2–7. TUNING EOS PARAMETERS Equations-of-state calculations are frequently burdened by the large number of fractions necessary to describe the hydrocarbon mixture for accurate phase behavior simulation. However, the cost and computer resources required for compositional reservoir simulation increase considerably with the number of components used in describing the reservoir fluid. For example, solving a two-phase compositional simulation problem of a reservoir system that is described by Nc components requires the simultaneous solution of 2Nc + 4 equations for each grid block. Therefore, reducing the number of components through lumping has been a common industry practice in compositional simulation to significantly speed up the simulation process. As discussed previously in this chapter, calculating EOS parameters (ie, ai, bi, and αi) requires component properties such as Tc, pc, and ω. These properties generally are well defined for pure components; however, determining these properties for the heavy fractions and lumped components rely on empirical correlations and the use of mixing rules. Inevitably, these empirical correlations and the selected mixing rules introduce uncertainties and errors in equation-of-state predictions. Tuning the selected equation of state is an important prerequisite to provide meaningful and reliable predictions. Tuning an equation of state refers to adjusting the parameters of the selected EOS to achieve a satisfactory match between the laboratory fluid PVT data and EOS results. The experimental data used should be closely relevant to the reservoir fluid and other recovery processes implemented in the field. The main objective of tuning EOS is to match all the available laboratory data, including 588 CHAPTER 5 Equations of State and Phase Equilibria n n n n n n n Saturation pressures; dew point and bubblepoint pressure DE data (Bod, Rsd, Bgd, ρ, etc.) CCE data (Y-function, Co, μ, ρ, etc.) CVD data (LDO, RF, Z-factor, etc.) Swelling tests data Separator tests data MMP Manual adjustments through trial and error or by an automatic nonlinear regression algorithm are used to adjust the EOS parameters to achieve a match between laboratory and EOS results. The regression variables are based on selecting a number of EOS parameters that may be adjusted or tuned to achieve a match between the experimental and EOS predictions. The following EOS parameters are commonly used as regression variables: n n n Properties of the undefined fractions that include Tc, pc, and ω or alternatively through the adjustments of Ωa and Ωb of these fractions. The Ω modifications should be interpreted as modifications of the critical properties, as they are related by the expressions a ¼ Ωa R2 Tc2 pc b ¼ Ωb RTc pc Binary interaction coefficient, kij, between the methane and C7+ fractions. When an injection gas contains significant amounts of nonhydrocarbon components, CO2 and N2, the binary interaction, kij, between these fractions and methane also should be modified. The regression is a nonlinear mathematical model that places global upper and lower limits on each regression variable, ΩaC1 , ΩbC1 ; and so forth. The nonlinear mathematical model determines the optimum values of these regression variables that minimize the objective function, defined as exp Nexp E Ecal X j j F¼ Wj Eexp j j where Nexp ¼ total number of experimental data points ¼ experimental (laboratory) value of observation j, such as pd, pb, Eexp j or ρob Tuning EOS Parameters 589 Ecal j ¼ EOC-calculated value of observation j, such as pd, pb, or ρob Wj ¼ weight factor on observation j The values of the weight factors in the objective function are internally set with default or user override values. The default values are 1.0 with the exceptions of values of 40 and 20 for saturation pressure and density, respectively. Note that the calculated value of any observation, such as Ecal j , is a function of the all-regression variables: Ecal j ¼ f ðpC1 ,TC1 ,…Þ The stepwise regression procedure in tuning EOS parameters is summarized next. Step 1. Split the plus fraction to C35 + or C45 + Step 2. Lump the extended compositional analysis into an optimum number of pseudocomponents; for example, F1, F2, F3: Comp Mol% CO2 N2 C1 C2 C3 i-C4 n-C4 i-C5 n-C5 C6 F1 F2 F3 MW 0.25 44.01 0.88 28.01 23.94 16.04 11.67 30.07 9.36 44.1 1.39 58.12 4.61 58.12 1.5 72.15 2.48 72.15 3.26 84 19.8198 111.46 13.5902 161.57 6.63966 241.13 Sp.gr Pc Tc 0.818 1071 547.9 0.8094 493.1 227.49 0.3 666.4 343.33 0.3562 706.5 549.92 0.507 616 666.06 0.5629 527.9 734.46 0.584 550.6 765.62 0.6247 490.4 829.1 0.6312 489.6 845.8 0.6781 457.1 910.1 0.7542 409.4 1047.55 0.8061 308 1199.96 0.8526 229.5 1363.62 Vc ω Tb 0.0342 0.0514 0.0991 0.0788 0.0737 0.0724 0.0702 0.0679 0.0675 0.064 0.06272 0.06299 0.06355 0.231 0.0372 0.0105 0.0992 0.1523 0.1852 0.1995 0.228 0.2514 0.2806 0.3653 0.4718 0.6199 351 140 201 333 416 471 491 542 557 607 714.42 864.2 1036.9 Step 3. Match PVT data by regressing on the following EOS parameters: • • • • • BIC: “kij” Critical pressure: Pc Critical temperature: Tc Shift volume: C Acentric factor: ω A summary of the first is described in the following tabulated form with the EOS regression parameters represented by “x”: 590 CHAPTER 5 Equations of State and Phase Equilibria Component N2 CO2 C1 C2 C3 i-C4 n-C4 i-C5 n-C5 C6 F1 F2 F3 kij pc Tc c ω x x x x x x x x x x x x x x x x x x x x Step 4. Having achieved a satisfactory match with the experimental data and obtained the “optimum” values of EOS parameters for C1, F1, F2, F3, and F4, further reduce the number of fractions representing the system by groping N2 with C1 and CO2 with C2. Match the available PVT data by only regression on the parameters of the newly combined groups, as shown below by “x”. Compare results with those obtained in step 2. If satisfactory, proceed to match with further grouping as shown in step 4. Component kij Pc Tc c ω N2 + C1 CO2 + C2 C3 i-C4 n-C4 i-C5 n-C5 C6 F1 F2 F3 x x x x x x x x x x x x Step 5. Continue the regression process on EOS parameters on each newly lumped groups without adjusting EOS parameters formed in the previous step. This step is summarized in the following two tabulated forms: Tuning EOS Parameters 591 Component kij N2 + C1 CO2 + C2 C3 i-C4 n-C4 i-C5 n-C5 C6 F1 F2 F3 x Component kij N2 + C1 CO2 + C2 C3+ i-C4 n-C4 i-C5 n-C5+ C6 F1 F2 F3 x Pc Tc c ω x x x x x x x x Pc Tc c ω x x x x x x x x x x x x x x As pointed out in Chapter 4, the Lohrenz, Bray, and Clark (LBC) viscosity correlation has become a standard in compositional reservoir simulation. When observed viscosity data are available, values of the coefficients a1– a5 in Eq. (4.86) and the critical volume of C7+ are adjusted and used as tuning parameters until a match with the experimental data is achieved. These adjustments are performed independent of the process of tuning equation-of-state parameters to match other PVT data. The key relationships of the LBC viscosity model follow: μob ¼ μo + 0 a1 + a2 ρr + a3 ρ2r + a4 ρ3r + a5 ρ4r ξm 1 n C B X C B ½ðxi Mi Vci Þ + xC7 + VC7 + Cρo B A @ i¼1 i62C7 + ρr ¼ Ma where a1 ¼ 0.1023 a2 ¼ 0.023364 a3 ¼ 0.058533 a4 ¼ 0.040758 a5 ¼ 0.0093324 4 0:0001 592 CHAPTER 5 Equations of State and Phase Equilibria PROBLEMS 1. A hydrocarbon system has the following overall composition: Component zi C1 C2 C3 i-C4 n-C4 i-C5 n-C5 C6 C7+ 0.30 0.10 0.05 0.03 0.03 0.02 0.02 0.05 0.40 given System pressure ¼ 2100 psia System temperature ¼ 150°F Specific gravity of C7+ ¼ 0.80 Molecular weight of C7+ ¼ 140 Calculate the equilibrium ratios of this system assuming both an ideal solution and real solution behavior. 2. A well is producing oil and gas with the following compositions at a gasoil ratio of 500 scf/STB: Component xi yi C1 C2 C3 n-C4 n-C5 C6 C7+ 0.35 0.08 0.07 0.06 0.05 0.05 0.34 0.60 0.10 0.10 0.07 0.05 0.05 0.03 given Current reservoir pressure ¼ 3000 psia Bubble point pressure ¼ 2800 psia Reservoir temperature ¼ 120°F Molecular weight of C7+ ¼ 160 Specific gravity of C7+ ¼ 0.823 Problems 593 Calculate the overall composition of the system; that is, zi. 3. The hydrocarbon mixture with the following composition exists in a reservoir at 234°F and 3500 psig: Component zi C1 C2 C3 C4 C5 C6 C7+ 0.3805 0.0933 0.0885 0.0600 0.0378 0.0356 0.3043 The C7+ has a molecular weight of 200 and a specific gravity of 0.8366. Calculate a. The bubble point pressure of the mixture. b. The compositions of the two phases if the mixture is flashed at 500 psia and 150°F. c. The density of the resulting liquid phase. d. The density of the resulting gas phase. e. The compositions of the two phases if the liquid from the first separator is further flashed at 14.7 psia and 60°F. f. The oil formation volume factor at the bubble point pressure. g. The original gas solubility. h. The oil viscosity at the bubble point pressure. 4. A crude oil exists in a reservoir at its bubble point pressure of 2520 psig and a temperature of 180°F. The oil has the following composition: Component xi CO2 N2 C1 C2 C3 i-C4 n-C4 i-C5 n-C5 C6 C7+ 0.0044 0.0045 0.3505 0.0464 0.0246 0.0683 0.0083 0.0080 0.0080 0.0546 0.4224 The molecular weight and specific gravity of C7+ are 225 and 0.8364. The reservoir initially contains 122 MMbbl of oil. The surface facilities 594 CHAPTER 5 Equations of State and Phase Equilibria consist of two separation stages connected in series. The first separation stage operates at 500 psig and 100°F. The second stage operates under standard conditions. a. Characterize C7+ in terms of its critical properties, boiling point, and acentric factor. b. Calculate the initial oil in place in STB. c. Calculate the standard cubic feet of gas initially in solution. d. Calculate the composition of the free gas and the composition of the remaining oil at 2495 psig, assuming the overall composition of the system remains constant. 5. A pure n-butane exists in the two-phase region at 120°F. Calculate the density of the coexisting phase using the following equations of state: a. van der Waals. b. Redlich-Kwong. c. Soave-Redlich-Kwong. d. Peng-Robinson. 6. A crude oil system with the following composition exists at its bubble point pressure of 3250 psia and 155°F. Component xi C1 C2 C3 C4 C5 C6 C7+ 0.42 0.08 0.06 0.02 0.01 0.04 0.37 If the molecular weight and specific gravity of the heptanes-plus fraction are 225 and 0.823, calculate the density of the crude using the a. Standing-Katz density correlation. b. Alani-Kennedy density correlation. c. van der Waals correlation. d. Redlich-Kwong EOS. e. Soave-Redlich-Kwong correlation. f. Soave-Redlich-Kwong correlation with the shift parameter. g. Peneloux volume correction. h. Peng-Robinson EOS. References 595 7. The following reservoir oil composition is available: Component Mol% Oil C1 C2 C3 i-C4 n-C4 i-C5 n-C5 C6 C7+ 40.0 8.0 5.0 1.0 3.0 1.0 1.5 6.0 34.5 The system exists at 150°F. Estimate the MMP if the oil is to be displaced by a. Methane. b. Nitrogen. c. CO2. d. 90% CO2 and 10% N2. e. 90% CO2 and 10% C1. 8. The compositional gradients of the Nameless field are as follows: C1 + N2 C7+ C2–C6 TVD, ft 69.57 67.44 65.61 63.04 62.11 61.23 58.87 54.08 6.89 8.58 10.16 12.54 13.44 14.32 16.75 22.02 23.54 23.98 24.23 24.42 24.45 24.45 24.38 23.9 4500 4640 4700 4740 4750 4760 4800 5000 Estimate the location of the GOC. REFERENCES Ahmed, T., 1991. A practical equation of state. (translation). SPE Reserv. Eng. 6 (1), 137–146. Ahmed, T., 1997. A generalized methodology for minimum miscibility pressure. In: Paper SPE 39034 presented at the Latin American Petroleum Conference, Rio de Janeiro, Brazil, August 30-September 3, 1997. 596 CHAPTER 5 Equations of State and Phase Equilibria Benmekki, E., Mansoori, G., 1988. Minimum miscibility pressure prediction with EOS. SPE Reserv. Eng. Brinkman, F.H., Sicking, J.N., 1960. Equilibrium ratios for reservoir studies. Trans. AIME 219, 313–319. Campbell, J.M., 1976. Gas conditioning and processing. In: Campbell Petroleum Series, vol. 1. Campbell, Norman, OK. Dodson, C., Goodwill, D., Mayer, E., 1953. Application of laboratory PVT data to engineering problems. J. Pet. Technol. 5 (12), 287–298. Dykstra, H., Mueller, T.D., 1965. Calculation of phase composition and properties for lean-or enriched-gas drive. Soc. Petrol. Eng. J. 5 (3), 239–246. Edmister, W., Lee, B., 1986. Applied Hydrocarbon Thermodynamics, In: second ed. vol. 1. Gulf, Houston, TX, p. pp. 52. Elliot, J., Daubert, T., 1985. Revised procedure for phase equilibrium calculations with soave equation of state. Ind. Eng. Chem. Process. Des. Dev. 23, 743–748. Graboski, M.S., Daubert, T.E., 1978. A modified soave equation of state for phase equilibrium calculations, 1. Hydrocarbon system. Ind. Eng. Chem. Process. Des. Dev. 17, 443–448. Hoffmann, A.E., Crump, J.S., Hocott, R.C., 1953. Equilibrium constants for a gascondensate system. Trans. AIME 198, 1–10. Jhaveri, B.S., Youngren, G.K., 1984. Three-parameter modification of the Peng-Robinson equation of sate to improve volumetric predictions. In: Paper SPE 13118, presented at the SPE Annual Technical Conference, Houston, September 16–19, 1984. Katz, D.L., Hachmuth, K.H., 1937. Vaporization equilibrium constants in a crude oilnatural gas system. Ind. Eng. Chem. 29, 1072. Katz, D., et al., 1959. Handbook of Natural Gas Engineering. McGraw-Hill, New York. Kehn, D.M., 1964. Rapid analysis of condensate systems by chromatography. J. Pet. Technol. 16 (4), 435–440. Lim, D., Ahmed, T., 1984. Calculation of liquid dropout for systems containing water. In: Paper SPE 13094, presented at the 59th Annual Technical Conference of the SPE, Houston, September 16–19, 1984. Lohrenze, J., Clark, G., Francis, R., 1963. A compositional material balance for combination drive reservoirs. J. Pet. Technol. Nikos, V., et al., 1986. Phase behavior of systems comprising north sea reservoir fluids and injection gases. J. Pet. Technol. 38 (11), 1221–1233. Pedersen, K., Thomassen, P., Fredenslund, A., 1989. Characterization of gas condensate mixtures. In: Advances in Thermodynamics. Taylor and Francis, New York. Peneloux, A., Rauzy, E., Freze, R., 1982. A consistent correlation for Redlich-KwongSoave volumes. Fluid Phase Equilib. 8, 7–23. Peng, D., Robinson, D., 1976a. A new two constant equation of state. Ind. Eng. Chem. Fundam. 15 (1), 59–64. Peng, D., Robinson, D., 1976b. Two and three phase equilibrium calculations for systems containing water. Can. J. Chem. Eng. 54, 595–598. Peng, D., Robinson, D., 1978. The Characterization of the Heptanes and Their Fractions: Research Report 28. Gas Producers Association, Tulsa, OK. Peng, D., Robinson, D., 1980. Two and three phase equilibrium calculations for coal gasification and related processes. In: Thermodynamics of Aqueous Systems with Industrial Applications. In: ACS Symposium Series no. 133. References 597 Redlich, O., Kwong, J., 1949. On the thermodynamics of solutions. An equation of state. Fugacities of gaseous solutions. Chem. Rev. 44, 233–247. Reid, R., Prausnitz, J.M., Sherwood, T., 1987. The Properties of Gases and Liquids, fourth ed. McGraw-Hill, New York. Sim, W.J., Daubert, T.E., 1980. Prediction of vapor-liquid equilibria of undefined mixtures. Ind. Eng. Chem. Process. Des. Dev. 19 (3), 380–393. Slot-Petersen, C., 1987. A systematic and consistent approach to determine binary interaction coefficients for the Peng-Robinson equation of state. In: Paper SPE 16941, presented at the 62nd Annual Technical Conference of the SPE, Dallas, September 27–30, 1987. Soave, G., 1972. Equilibrium constants from a modified Redlich-Kwong equation of state. Chem. Eng. Sci. 27, 1197–1203. Standing, M.B., 1977. Volumetric and Phase Behavior of Oil Field Hydrocarbon Systems. Society of Petroleum Engineers of AIME, Dallas, TX. Standing, M.B., 1979. A set of equations for computing equilibrium ratios of a crude oil/ natural gas system at pressures below 1,000 psia. J. Pet. Technol. 31 (9), 1193–1195. Stryjek, R., Vera, J.H., 1986. PRSV: an improvement Peng-Robinson equation of state for pure compounds and mixtures. Can. J. Chem. Eng. 64, 323–333. Van der Waals, J.D., 1873. On the Continuity of the Liquid and Gaseous State. Sijthoff, Leiden, The Netherlands (Ph.D. dissertation). Vidal, J., Daubert, T., 1978. Equations of state—reworking the old forms. Chem. Eng. Sci. 33, 787–791. Whiston, C., Brulé, M., 2000. “Phase Behavior.” Monograph volume 20, Society of Petroleum Engineers. Richardson, TX. Whitson, C.H., Brule, M.R., 2000. Phase Behavior. SPE, Richardson, TX. Whitson, C.H., Torp, S.B., 1981. Evaluating constant volume depletion data. In: Paper SPE 10067, presented at the SPE 56th Annual Fall Technical Conference, San Antonio, October 5–7, 1981. Wilson, G., 1968. A modified Redlich-Kwong EOS, application to general physical data calculations. In: Paper 15C, presented at the Annual AIChE National Meeting, Cleveland, May 4–7, 1968. Winn, F.W., 1954. Simplified nomographic presentation, hydrocarbon vapor-liquid equilibria. Chem. Eng. Prog. Symp. Ser. 33 (6), 131–135. Index Note: Page numbers followed by f indicate figures, t indicate tables, and b indicate boxes. A Abu-Khamsin and Al-Marhoun correlation, 308 Acentric factor, of propane, 13–14 Acid number, 73 Ahmed modified method, 159–164, 161b Ahmed’s splitting method, 150–155, 152b Alani-Kennedy method, 498 crude oil density calculation, 250–254, 251t, 253b n-Alkanes boiling point data, 89t GC data calibration, 88–89, 88f Alston’s correlation, MMP, 449–450 American Petroleum Institute (API) gravity, 71 crude oil, 239–241, 240b Aniline point, 73 Apparent molecular weight, 191–192 ASTM distillation curve, 134–135 B Beal’s correlation dead oil viscosity, 306 undersaturated oil viscosity, 309 Beggs-Robinson correlation dead oil viscosity, 306 saturated oil viscosity, 308 Behavior gas-condensate wells, 432–434 retrograde gas reservoirs, 434–437 Behrens and Sandler’s lumping scheme, 170–175, 173t Bergman’s method, 129–133, 132b Binary interaction coefficient (BIC), 521–522 Binary system, 475–476 characteristic of, 22 pressure-composition diagram for, 25–30, 26–27f pressure-volume diagram for, 23f properties, 521–522 two-component system, 22 Black oil approach, 342–344 simulator, 340–341 Boiling point curve, 4, 78 graphical correlation, 133–136 range, 78 temperature, 4 Bubble point curve, 3, 34 two-component system, 23 Bubble point pressure, 2, 268–275 equation of state, 550–551 Glaso’s correlation, 272–273, 273b Marhoun’s correlation, 273–274, 274b Petrosky-Farshad’s correlation, 274–275, 275b Standing’s correlation, 269–271, 270–271b two-component system, 22–23 vapor-liquid equilibrium, 491–493, 492b Vasquez-Beggs’s correlation, 271–272, 272b n-Butane, saturated liquid and gas density, 17b C Campbell’s method, 486 Carbon atoms critical properties and acentric factor, 119b measured mole, 183f methane content vs. depth, 182f molecular weight vs. depth, 183f saturation pressure vs. depth, 182f Carr-Kobayashi-Burrows’s correction method, 205–207, 224–228 Cavett’s correlations, 97 petroleum fraction, 112 CCE test. See Constant composition expansion (CCE) test Chemical potential, mixture mixing rule, 529–530 Chew-Connally correlation, saturated oil viscosity, 307–308 Cloud point, 72 Composite liberation test EOS, 578–580, 579f experimental procedure, 577–578 Compositional approach, 343 Compositional gradients, EOS, 583–585, 584t, 585–586f Compositional simulators, 343 Compressibility factors. See also Gas compressibility factor direct calculation of, 212 Standing-Katz, 207–209 Condensate GOR, and condensate yield, 424–426, 424–425f Conradson carbon residue (CCR), 73 Constant composition expansion (CCE) test, 268, 355–372, 356f, 357t, 359f, 566f EOS, 568–569 experimental procedure, 566–567 gas-condensate system, 567 laboratory analysis, 407–411, 409t isothermal compressibility coefficient from, 361, 361f oil density from, 360, 382, 383f pressure-volume relations from, 359f quality checks of total FVF from, 384, 385b Constant volume depletion (CVD) test, 560f EOS, 562–566, 563f gas-condensate systems laboratory analysis, 412–415, 412f laboratory procedure, 559–562 reservoir engineering predictions, 559 Continuous distribution function, 170 Convergence pressure method, 483–485 Cox chart, 11, 13b Cricondenbar, 33 Cricondentherm, 33 Critical compressibility factor correlations, 101 Critical density, 14–16 Critical point, 3, 33 Critical pressure, 3 graphical correlations, 137–139, 139f Critical temperature, 3 graphical correlation, 137, 138f Critical volume, 3 Cronquist’s correlation, MMP, 451 599 600 Index Crude oil API gravity, 239–241, 240b assay molecular and chemical characteristics, 71–73 plus fraction characterization methods, 93–139 stimulated distillation by GC, 87–93 TBP, 77–87 undefined components, 77 well-defined components, 73–76 characteristic properties of, 39–41 classification, 35–43 heptanes-plus fraction in, 174b liquid shrinkage for, 43f low-shrinkage oil, 39 near-critical crude oil, 41 ordinary black oil, 37 volatile crude oil, 39 Crude oil density, 239–240, 242–260 Alani-Kennedy’s method, 250–254, 251t, 253b Standing-Katz’s method, 243–249, 244f, 247b, 247–248f Crude oil viscosity, 304–315, 305f dead oil correlations, 306, 310b Beal’s correlation, 306 Beggs-Robinson’s correlation, 306 Glaso’s correlation, 307 saturated oil viscosity, 307, 310b Abu-Khamsin and Al-Marhoun, 308 Beggs-Robinson correlation, 308 Chew-Connally correlation, 307–308 undersaturated oil viscosity, 308–309, 310b Beal’s correlation, 309 Khan’s correlation, 309 Vasquez-Beggs’s correlation, 309–311 Cubic average boiling point (CABP), 123 CVD test. See Constant volume depletion (CVD) test D Dalton’s law, 468 Darcy’s equation, for gas and oil flow rates, 332 Dead oil viscosity, 304, 306, 310b Beal’s correlation, 306 Beggs-Robinson’s correlation, 306 Glaso’s correlation, 307 Density, 11 crude oil, 239–240, 242–260 Alani-Kennedy’s method, 250–254, 251t, 253b Standing-Katz’s method, 243–249, 244f, 247b, 247–248f liquid, estimation Katz’s method, 254–258, 255–256f, 257b, 259b Standing’s method, 258–260, 258b saturated oil DE test, 384b quality checks of, 383, 384b undersaturated oil, 296–297 Density-temperature diagram n-butane, 17 propane, 14–16, 15f DE test. See Differential expansion (DE) test Dew point curve, 3, 34 two-component system, 23 Dew point pressure, 2 equation of state, 548–550 swelling tests, 582f two-component system, 22–23 vapor-liquid equilibrium, 488–491, 490b Differential expansion (DE) test Dodson procedure, 394, 397f oil density from, 382, 383f saturated oil density, 384b total FVF from, 379–382, 380b, 385b, 386f quality check, 384, 385b Differential liberation (DL) test, 373–386, 373f, 376b differential gas solubility, 376–377, 377b differential oil FVF, 375–376, 376b differential solution gas-oil ratio, 569–570, 570f EOS, 571–574, 572f gas deviation factor “Z”, 378–379 gas FVF, 377–378, 378–379b quality checks of DE data, 382–386 relative-oil volume factor, 569–570, 570f Differential separation process, 495 Distillation process GC, 87–93 TBP analysis, 78–80, 79f DL test. See Differential liberation (DL) test Dodson test, 394, 397f, 577–578 Dranchuk and Abu-Kassem’s method, 214–215 Dranchuk-Purvis-Robinson method, 216–224 Dry gas reservoirs, 50–59, 51f E ECN. See Effective carbon numbers (ECN) Edmister correlation, 100–101, 530 Edmister’s equation, 110 Effective carbon numbers (ECN), 90–93 Enhanced oil recovery (EOR) projects, 315, 452–453 Equation of state (EOS), 505–530 applications, 546 bubble point pressure, 550–551 cubic, 341–344 definition, 190, 505 dew point pressure, 548–550 equilibrium ratio, 546–548, 547f model, thermodynamic-based, 344 parameters, 342 Peng-Robinson, 341–342, 534–546, 536b PVT test data, 558–587 composite liberation test, 577–580, 579f compositional gradients, 583–585, 584t, 585–586f constant composition expansion test, 566–569, 566f constant volume depletion test, 559–566, 560f, 563f differential liberation test, 569–574, 572f flash separation tests, 574–576 minimum miscibility pressure, 586–587 swelling tests, 580–583, 581–582f real gas, 414 Redlich-Kwong, 512–517, 513b, 515–516b simulator, compositional, 342 Soave-Redlich-Kwong, 517–521 example, 520b, 523b mixture mixing rule, 521–530 modifications, 530–546 volume translation method, 531–534, 533b Starling-Carnahan, 212–213 three-phase equilibrium calculations, 551–557 tuning, 342–344, 442 tuning parameters, 587–591 two-parameter cubic, 509 van der Waals, 506–512, 510b vapor pressure, 557–558 Equilibrium ratios, 467–471 development, 476 Index 601 equation of state, 546–548, 547f for hydrocarbon system, 484f mixture mixing rule, 526–527 for plus fractions, 485–488 Campbell’s method, 486 Katz’s method, 487 K-values, 487–488 Winn’s method, 486–487 for real solutions, 476–485 convergence pressure method, 483–485 example, 481b K-value of plus fraction, 478–480 simplified K-value method, 480–483 Standing’s correlation, 477–483 Wilson’s correlation, 476–477 Whitson and Torp’s method, 485b Extended Y-function, 366, 368b, 370f Extrapolated vapor pressure (EVP) method, MMP correlation, 448–449, 448b F Fanchi’s linear equation, 316 Final boiling point (FBP), 78 Firoozabadi and Aziz’s correlation, MMP, 453–454 Flash calculation, 471–476, 473b Flash liberation test, 374 Flash point, 72 Flash separation, 495–505 Flash separation test, 574–576 gas-in-solution curve, 577f laboratory procedure, 574–575 oil-volume curve, 576f Fluid density, 16–17, 16f Formation volume factor (FVF). See also Oil formation volume factor (FVF); Total formation volume factor (FVF) gas, 377–378, 378–379b water, 323–324 Freeze point, 73 Fugacity, 525–526 Fusion, 10 FVF. See Formation volume factor (FVF) G Gas composition, 198b definition, 189 deviation factor “Z”, 378–379 expansion factor, 222–223 formation volume factor, 221–222 FVF, 377–378, 378–379b high-molecular weight, 207–212 properties, 189, 191 sample conditioning, 335 separator, 335–336 specific gravity of, 241–242, 241b Gas-cap reservoir, 35 classification, 36f dry, 53f retrograde, 53f Gas chromatography (GC), 345–346, 346f stimulated distillation by, 87–93, 90f Gas compressibility factor, 196 calculation, 212 chart, 197–201, 198f for natural gases, 196–197 Gas-condensate reservoir, liquid blockage in, 431–457 Gas-condensate systems laboratory analysis, 254 constant composition test, 407–411, 409t CVD test, 412–415, 412f retrograde gas laboratory data, 416–423 recombination of separator samples, 404–407, 416–417t, 418f, 423b consistency test on recombined GOR, 404t, 406–407, 406–407f, 409t Gas-condensate wells, behavior of, 432–434 Gas density, 193, 204b of n-butane, 17b Gas-down-to (GDT) level, 60–61 Gas flow rate, 222b Darcy’s equation for, 332 Gas-oil contact (GOC), 59–64 compositional changes, 55f determination, 55f discovery well in gas column, 60–62f locating, 54t reservoir pressure, 585, 585–586f true vertical depth, 63b undersaturated, 65, 65f well-bore and stock-tank density, 56f Gas-oil ratio (GOR), 37–43, 241–242 consistency check, 355f consistency test on recombined, 404t, 406–407, 406–407f, 409t material balance consistency test on, 352–355 surface, 332 Gas-oil separation differential separation process, 495 flash separation, 495–505 types of, 494–495 Gas reservoir, 34 categories, 43 compositions, 52f dry gas reservoirs, 50–59 MBE for retrograde, 426–431 consistency tests on CCE and CVD Data, 428–431, 428–429t, 431–432f near-critical gas-condensate reservoirs, 46 retrograde gas reservoirs, 44–46 wet gas reservoirs, 47–50 Gas solubility, 260–268, 261f, 268b differential, 376–377, 377b Glaso’s correlation, 264–265, 265b Marhoun’s correlation, 265–266, 266b Petrosky-Farshad’s correlation, 266–268, 267b Standing’s correlation, 261–263, 262b Vasquez-Beggs’s correlation, 263–264, 264b water, 326 Gas viscosity, 223–224, 227b atmospheric correlation, 225f example, 227b Gas-water contact (GWC), 61 Generalized correlations, 94–122 Gibbs energy, mixture mixing rule, 527–529 Glaso’s correlation bubble point pressure, 272–273, 273b dead oil viscosity, 307 gas solubility, 264–265, 265b MMP correlation, 454–455 oil FVF, 278 total FVF, 301, 302b GOC. See Gas-oil contact (GOC) GOM crude oil system. See Gulf of Mexico (GOM) crude oil system GOR. See Gas-oil ratio (GOR) Graphical correlations boiling points, 133–136 critical pressure, 137–139, 139f critical temperature, 137, 138f molecular weight, 136–137 Gross heat of combustion (high heating value), 73 Gulf of Mexico (GOM) crude oil system, 266–268 hydrocarbon system, 279 PVT correlations for, 318–322 GWC. See Gas-water contact (GWC) 602 Index H I Hall and Yarborough correlations, 104 Hall-Yarborough’s method, 212–214 Heptanes-plus fraction, 78, 110b, 479–480 bubble point pressure, 492 in crude oil system, 174b dew point pressure, 490 molar distribution of, 180b molecular weight and specific gravity, 117b Hoffman plot, 349–350, 351f Hong’s mixing rules, 168–169 Hudgins’s correlation, MMP, 454 Hydrocarbon analysis of recombined reservoir fluid samples, 347t in petroleum reservoirs, 1 vapor pressure chart for, 12f weights and volumes calculation, 18b Hydrocarbon phase behavior crude oil system, 35–43 gas-oil contact, 59–64 gas reservoirs, 43–59 multicomponent systems, 32 oil reservoirs, 34–35 phase rule, 65–67 pressure-temperature diagram, 32–34 pure component saturated liquid density, 19–21 pure component volumetric properties, 11–19 p-x diagram, 25–30 reservoirs and reservoir fluids, 34–59 single-component systems, 1–21 three-component systems, 30–31 two-component systems, 22–30 undersaturated GOC, 65 Hydrocarbon-plus fractions crude oil assay, 71–77 lumping schemes, 165–180 PNA determination, 122–133 splitting schemes, 141–164 undefined components, 77 well-defined components, 73–76 Hydrocarbon system. See also specific types of hydrocarbon system components, vapor pressure, 469f equilibrium ratios for, 484f gas composition, 209b total mole fraction in, 470 IBP. See Initial boiling point (IBP) Ideal gas. See also Natural gas apparent molecular weight, 191–192 behavior of, 190–196, 200b definition, 189 density, 193 example, 191b, 195b specific gravity, 194–196, 194b specific volume, 193 standard volume, 193 Ideal gas law, 190 Ideal solution, 468–470 Initial boiling point (IBP), 78 Interfacial tension, 315–318 Inverse lever rule, 28–30 Isobutene, liquid and vapor calculation, 29b Isothermal compressibility coefficient from CCE test, 361, 361f of crude oil, 282–293 saturated, 283, 291–293, 292b undersaturated, 282–291 K Katz’s method, 487 liquid density estimation, 254–258, 255–256f, 257b, 259b Kesler and Lee correlations, 97–99 petroleum fraction, 113 Khan’s correlation, undersaturated oil viscosity, 309 K-values for nonhydrocarbon components, 487–488 of plus fraction, 478–480 simplified method, 480–483 L Lange’s correlation, MMP, 455–457 LBC viscosity correlation. See Lohrenz, Bray, and Clark (LBC) viscosity correlation Lean gas and nitrogen miscibility correlation, MMP, 452–457 Lee and Kesler correlation, 530 Lee-Gonzalez-Eakin’s method, 228–234, 229b Lee’s lumping scheme, 176–180 Lee’s mixing rules, 169–170 Lever rule, 31 LGF. See Logistic growth function (LGF) Liquid blockage, in gas-condensate reservoirs, 431–457 Liquid density estimation Katz’s method, 254–258, 255–256f, 257b, 259b Standing’s method, 258–260, 258b Liquid-gas separation process, 494 Logistic growth function (LGF), 163–164, 164f Log-log scale, 470 Lohrenz, Bray, and Clark (LBC) viscosity correlation, 591–592 Lohrenz’s splitting method, 147–148 Low-shrinkage oil, 39 oil-shrinkage curve, 40f phase diagram for, 39f Lumping schemes Behrens and Sandler’s lumping scheme, 170–175, 173t Hong’s mixing rules, 168–169 Lee’s lumping scheme, 176–180 Lee’s mixing rules, 169–170 Whitson’s lumping scheme, 165–168 M MABP. See Molar average boiling point (MABP) Magoulas and Tassios correlations, 104 carbon atoms, 120 heptanes-plus fraction, 118 Marhoun’s correlation bubble point pressure, 273–274, 274b gas solubility, 265–266, 266b oil FVF, 279 total FVF, 302–304, 302b Material balance (MB) consistency test, 350–352 on GOR, 352–355 quality test, 353t Material balance equation (MBE) oil FVF, 280–282 for retrograde gas reservoirs, 426–431 consistency tests on CCE and CVD data, 428–431, 428–429t, 431–432f MB. See Material balance (MB) MBE. See Material balance equation (MBE) MDT tool. See Modular dynamic testing (MDT) tool Mean average boiling point (MeABP), 123 Melting curve, 10 Melting-point temperature, 10 Index 603 Microsoft Solver, 90–92 Minimum miscibility pressure (MMP), 438, 441, 586–587 determination, 442f pure and impure CO2, 446–452, 448b Minimum miscibility pressure (MMP) correlations, 445–452 Alston’s correlation, 449–450 Cronquist’s correlation, 451 EVP method, 448–449, 448b Firoozabadi and Aziz’s correlation, 453–454 Glaso’s correlation, 454–455 Hudgins’s correlation, 454 Lange’s correlation, 455–457 lean gas and nitrogen miscibility correlations, 452–457 NPC method, 451 Orr and Silva correlation, 447–448 pure and impure CO2 MMP, 446–452, 448b Sebastian and Lawrence’s correlation, 455 Sebastian’s method, 450–451 Yellig and Metcalfe correlation, 449, 449b Yuan’s correlation, 451–452 Mixture mixing rule, SRK EOS, 521–530 chemical potential, 529–530 correct Z-factor selection, 527–529 equilibrium ratio, 526–527 Gibbs energy, 527–529 MMP. See Minimum miscibility pressure (MMP) Modified Katz’s splitting method, 143–147 Modular dynamic testing (MDT) tool, 329 Molar average boiling point (MABP), 123 Molar density, 12 Molar volume, 12 Molecular weight graphical correlation, 136–137 methods correlations, 108–121, 109t, 109f Mole fraction, 192–193 Multicomponent hydrocarbon systems, 32 pressure-temperature diagram, 32–34, 33f Multiple hydrocarbon samples, 181–184 N National Petroleum Council (NPC) method, MMP correlation, 451 Natural gas, 189. See also Ideal gas Carr-Kobayashi-Burrows’s correction method, 205–207, 224–228 compressibility of, 216–221 Dranchuk and Abu-Kassem’s method, 214–215 Dranchuk-Purvis-Robinson method, 216–224 gas compressibility factor, 196–197 Hall-Yarborough’s method, 212–214 hydrocarbon components, 202 Lee-Gonzalez-Eakin’s method, 228–234, 229b Newton-Raphson iteration technique, 215 physical properties, 189 Trube’s pseudoreduced compressibility, 219–220f viscosity, 223–224, 227b Wichert-Aziz’s correction method, 203–205 Near-critical crude oil, 41 liquid-shrinkage curve for, 42f phase diagram for, 42f Near-critical gas-condensate reservoirs, 46, 46–47f Near-critical reservoirs, 34 Net heat of combustion (lower heating value), 73 Newton-Raphson iteration technique, 215, 472 Newton-Raphson method, 554 Nonhydrocarbon adjustment methods Carr-Kobayashi-Burrows’s correction method, 205–207 Wichert-Aziz’s correction method, 203–205 Nonhydrocarbon components, Z-factor effect, 196–203 NPC method. See National Petroleum Council (NPC) method O Oil flow rates, Darcy’s equation, 332 properties, undersaturated, 294–297, 294f reservoir phase distribution in, 375f Oil density. See also Crude oil density from CCE test, 360 undersaturated, 296–297 Oil-down-to (ODT) level, 59–60, 59f Oil formation volume factor (FVF), 275–282, 281b differential, 375–376, 376b Glaso’s correlation, 278 Marhoun’s correlation, 279 material balance equation, 280–282 Petrosky-Farshad’s correlation, 279 pressure relationship, 276, 276f Standing’s correlation, 277 undersaturated, 295–296, 296b Vasquez-Beggs’s correlation, 277–278 Oil reservoir, 34 classification of, 34–35 phase distribution, 375f Optimum separator pressure, 496 Ordinary black oil, 37 liquid-shrinkage curve for, 38f p-T diagram for, 38f Orr and Silva correlation, MMP, 447–448 P Parachor, 315–316 Pedersen’s splitting method, 148–150, 149b Peng-Robinson equation of state (PR EOS), 534–546, 536b Peng-Robinson’s method, 124–129, 127b Perfect gas, 189 Perfect gas law, 196 Petroleum fraction acentric factor, 111b critical properties, 111b molecular weight, 111b Petrosky-Farshad’s correlation bubble point pressure, 274–275, 275b gas solubility, 266–268, 267b oil FVF, 279 undersaturated oil compressibility estimation, 290, 290b Phase behavior, 1 Phase diagram, 1, 32 of binary mixture, 25f dry gas, 51f low-shrinkage oil, 39f near-critical crude oil, 42f near-critical gas-condensate reservoir, 46f retrograde system, 44f ternary, 32f wet gas, 47f Phase envelope, 3, 33 impact of composition, 36f 604 Index Phase equilibria, 468 calculations, 470, 488–505 Phase rule, 65–67 Plus fraction equilibrium ratios for, 485–488 Campbell’s method, 486 Katz’s method, 487 Winn’s method, 486–487 K-value of, 478–480, 487–488 Plus fraction characterization methods Cavett’s correlations, 97 critical compressibility factor correlations, 101 Edmister’s correlations, 100–101 generalized correlations, 94–122 Hall and Yarborough correlations, 104 Kesler and Lee correlations, 97–99 Magoulas and Tassios correlations, 104 molecular weight methods correlations, 108–121 properties, 93–94 recommended characterizations, 121–122 Riazi and Daubert generalized correlations, 95–96 Rowe’s characterization method, 101–102 Sancet correlations, 108 Silva and Rodriguez correlations, 107 Standing’s correlations, 103 Twu’s correlations, 105–107 Watansiri-Owens-Starling correlations, 99–100 Willman-Teja correlations, 103–104 Winn and Sim-Daubert correlations, 99 PNA determination, 122–133 Pour point, 72 Power function form, 160, 160f PR EOS. See Peng-Robinson equation of state (PR EOS) Pressure-composition diagram, for binary systems, 25–30, 26–27f Pressure-temperature diagram multicomponent systems, 32–34, 33f ordinary black oil, 38f single-component system, 4, 10f, 11 two-component systems, 24f Pressure-volume diagram for binary system, 23f pure component, 2–3, 2f Pressure-volume relation, from CCE test, 359f Pressure-volume-temperature (PVT) correlations for GOM crude oil system, 318–322 test data, EOS, 558–587 composite liberation test, 577–580, 579f compositional gradients, 583–585, 584t, 585–586f constant composition expansion test, 566–569, 566f constant volume depletion test, 559–566, 560f, 563f differential liberation test, 569–574, 572f flash separation tests, 574–576 minimum miscibility pressure, 586–587 swelling tests, 580–583, 581–582f Pressure-volume-temperature (PVT) measurements, 342–415 differential liberation test, 373–386, 373f, 376b differential gas solubility, 376–377, 377b differential oil FVF, 375–376, 376b gas deviation factor “Z”, 378–379 gas FVF, 377–378, 378–379b quality checks of DE data, 382–386 routine laboratory PVT tests compositional analysis of reservoir fluid, 345–355, 346f, 347t quality assessment of laboratory compositional analysis, 346–355, 350f separator tests, 386–403, 387f, 389–390t, 390–391f adjustment of DE data to separator conditions, 389–398, 395t, 396–397f, 398t extrapolation of reservoir fluid data, 398–403 Pressure/volume tests. See Constant composition expansion (CCE) test Propane existence state, 13b saturated densities of, 19 saturated liquid density of, 21b in underground cavern, 18b vapor pressure of, 14b Pseudocomponent, from TBP test, 78–79 Pseudofractions, 74 Pseudoization, 165 Pure component physical properties for, 5t pressure-volume phase diagram, 2–3, 2f pressure-volume relationship, 507f, 509f Rackett’s compressibility factor, 20, 20t, 21b saturated liquid density, 19–21 vapor pressure, 557–558 volumetric properties, 11–19 PVT. See Pressure-volume-temperature (PVT) Q Quality lines, 24, 33 R Rackett compressibility factor, 20, 20t, 21b, 532 Raoult’s law, 468 Real gases, behavior of, 196–203 Real solutions, equilibrium ratios for, 476–485 convergence pressure method, 483–485 example, 481b K-value of plus fraction, 478–480 simplified K-value method, 480–483 Standing’s correlation, 477–483 Wilson’s correlation, 476–477 Recombined GOR, consistency test on, 404t, 406–407, 406–407f, 409t Redlich-Kwong equation of state (RK EOS), 512–517, 513b, 515–516b Reduced temperature, 13 Refractive index (RI), 72 Reid vapor pressure (RVP), 72 Repeat formation tester (RFT) tool, 51–52, 56–57 Reservoir classification, 34–59, 35f Reservoir fluids classification of, 34–59 laboratory analysis, 327 black oil simulator, 340–341 compositional modeling, 341–342 fluid sampling, 328–338 representative sample, 339–340, 341f separator samples, 331–337, 331f, 333f, 337t in situ representative sample, 339–340 subsurface sampling, 329–330, 330f, 334 unrepresentative sample, 339–340, 341f well conditioning, 327–328 wellhead sampling, 337–338 Index 605 Reservoir water properties, 323–327 gas solubility in water, 326 water FVF, 323–324 water isothermal compressibility, 326–327 water viscosity, 324–326 Retention time, 87 Retrograde gas reservoir, 44–46, 44f behavior of, 434–437 MBE for, 426–431 Riazi and Daubert generalized correlations, 95–96 Riazi-Daubert correlation, carbon atoms, 120 Riazi-Daubert equation, 110 petroleum fraction, 111 Rising bubble test, 445 RK EOS. See Redlich-Kwong equation of state (RK EOS) Routine laboratory PVT tests compositional analysis of reservoir fluid, 345–355, 346f, 347t quality assessment of laboratory compositional analysis, 346–355, 350f Rowe’s characterization method, 101–102 Rowe’s correlations carbon atoms, 119 heptanes-plus fraction, 117 RVP. See Reid vapor pressure (RVP) S Sancet correlations, 108 Saturated gas, 3 Saturated liquid, 3 of n-butane, 17b density curve, 16–17 Saturated oil density DE test, 384b quality checks of, 383, 384b isothermal compressibility coefficient, 283, 291–293, 292b reservoir, 35 viscosity, 304, 307, 310b Abu-Khamsin and Al-Marhoun correlation, 308 Beggs-Robinson correlation, 308 Chew-Connally correlation, 307–308 Saturated vapor density curve, 16–17 Schlumberger Repeat Formation Testing (RFT), 329 Sebastian and Lawrence’s correlation, MMP, 455 Sebastian’s method, MMP correlation, 450–451 Separation process differential, 495 flash, 495–505 gas-in-solution curve, 577f gas-oil separation, 494–495 liquid-gas, 494 n-separation stages, 498f oil-volume curve, 576f two-stage, 494, 494f vapor-liquid equilibrium calculations, 493–505 Separator gas, 335–336 Separator liquid samples, 336 Separator shrinkage factor, 334 Separator test. See also Flash separation test PVT measurement, 386–403, 387f, 389–390t, 390–391f adjustment of DE data to separator conditions, 389–398, 395t, 396–397f, 398t extrapolation of reservoir fluid data, 398–403 Silva and Rodriguez correlations, 107, 111 Simulated distillation (SD), GC, 87–93, 90f Simulator black oil, 340–341 compositional, 343 Single carbon number (SCN), 74 coefficients of equation, 77t concept of, 75t generalized physical properties, 76t Single-component hydrocarbon system, 1–21 degrees of freedom, 66b physical properties, 5t pressure-temperature diagram of, 4, 10f, 11 pressure-volume diagram, 2f saturated liquid density, 19–21 volumetric properties, 11–19 Slim tube test, 440–444, 441–442f relative permeability adjustment, 444 viscosity adjustment, 443–444 Smoke point, 73 Soave-Redlich-Kwong equation of state (SRK EOS), 517–521 example, 520b, 523b mixture mixing rule, 521–530 chemical potential, 529–530 correct Z-factor selection, 527–529 equilibrium ratio, 526–527 Gibbs energy, 527–529 modifications, 530–546 volume translation method, 531–534, 533b Solution gas, specific gravity of, 240b, 241–242 Special laboratory PVT tests, 437–457 minimum miscibility pressure correlations, 445–452 Alston’s correlation, 449–450 Cronquist’s correlation, 451 EVP method, 448–449, 448b Firoozabadi and Aziz’s correlation, 453–454 Glaso’s correlation, 454–455 Hudgins’s correlation, 454 Lange’s correlation, 455–457 lean gas and nitrogen miscibility correlations, 452–457 NPC method, 451 Orr and Silva correlation, 447–448 pure and impure CO2 MMP, 446–452, 448b Sebastian and Lawrence’s correlation, 455 Sebastian’s method, 450–451 Yellig and Metcalfe correlation, 449, 449b Yuan’s correlation, 451–452 rising bubble test, 445 slim tube test, 440–444, 441–442f relative permeability adjustment, 444 viscosity adjustment, 443–444 swelling test, 438–439, 439–440f Specific gravity, 194–196 example, 194b wet gases, 230–234, 232b Specific volume, 12, 193 Splitting schemes, 141–164 Ahmed modified method, 159–164, 161b Ahmed’s method, 150–155, 152b condensate-gas system, 144b discrete and continuous compositional distribution, 141f exponential and left-skewed distribution functions, 142f Lohrenz’s method, 147–148 modified Katz’s method, 143–147 Pedersen’s method, 148–150, 149b Whitson’s method, 155–159, 158b 606 Index SRK EOS. See Soave-Redlich-Kwong equation of state (SRK EOS) Standard volume, 193 Standing-Katz chart compressibility factor, 207–209 Z-factor, 212 Standing-Katz’s method, 498 crude oil density calculation, 243–249, 244f, 247b, 247–248f Standing’s correlation, 103, 477–483 bubble point pressure, 269–271, 270–271b carbon atoms, 120 gas solubility, 261–263, 262b heptanes-plus fraction, 118 oil FVF, 277 total FVF, 300–301, 302b undersaturated oil compressibility estimation, 290–291, 290b Standing’s method, liquid density estimation, 258–260, 258b Starling-Carnahan equation of state, 212–213 Straight-line relationship, 15, 15f Sublimation pressure curve, 10 Sulfur content, 72 Surface GOR, 332 Surface samples recombination procedure, 338–339, 338f Surface tension, 315–318 Swelling test, 438–439, 439–440f, 581f dew point pressure, 582f equation of state, 581–583 procedure, 580–581 reservoir gas, 582f T Three-component hydrocarbon systems, 30–31 at constant temperature and pressure, 32f degrees of freedom, 67b features of, 31 properties of, 31f Three-parameter gamma probability function, 155–156, 156f Three-phase equilibrium calculation, EOS, 551–557 Tie line, 28–31 Total formation volume factor (FVF), 298–304, 299f CCE test, quality check, 384, 385b Glaso’s correlation, 301, 302b Marhoun’s correlation, 302–304, 302b Standing’s correlation, 300–301, 302b Total mole fraction, in hydrocarbon, 470 Traditional Y-function, 363, 364f, 364b Trial-and-error process, 489 Trube’s correlation, undersaturated oil compressibility estimation, 284–289, 285–287f, 288b True boiling point (TBP), 77–87 condensate-gas specific gravity vs. molecular weight, 83f distillation process, 78–80, 79f mathematical representation of, 86f, 87 pseudocomponent from, 78–79 stock-tank oil sample, 80f volatile oil specific gravity vs. molecular weight, 82f Watson characterization factor, 80–82, 84 Tuning equation of state parameters, 543, 587–591 Two-component hydrocarbon system, 22–30 amounts of liquid and vapor, 28, 29f degrees of freedom, 67b isobutene, liquid and vapor calculation, 29b nomenclatures, 27–28 pressure-composition diagram, 25–30, 26–27f pressure-temperature diagram, 24f pressure-volume isotherm, 22f volatile component, 28 Two-parameter cubic equation of state, 509 Two-phase region, 3 Two-stage separation processes, 494, 494f Twu’s correlations, 105–107 U Undefined petroleum fractions, 77 Undersaturated GOC, 65, 65f Undersaturated oil compressibility estimation Petrosky-Farshad’s correlation, 290, 290b Standing’s correlation, 290–291, 290b Trube’s correlation, 284–289, 285–287f, 288b Vasquez-Beggs’s correlation, 289, 290b density, 296–297 FVF, 295–296, 296b isothermal compressibility coefficient, 282–291 properties, 294–297, 294f reservoir, 34 viscosity, 305, 308–309, 310b Beal’s correlation, 309 Khan’s correlation, 309 Vasquez-Beggs’s correlation, 309–311 Universal oil products, 80–81 V VABP. See Volume average boiling point (VABP) Van der Waals equation of state (VdW EOS), 506–512, 510b Vapor-liquid equilibrium bubble point pressure, 491–493, 492b dew point pressure, 488–491, 490b separator calculations, 493–505 Vapor phase, weights and volumes of, 18b Vapor pressure, 4 chart for hydrocarbon components, 12f curve, 4–10 equation of state, 557–558 hydrocarbon components, 469f Vasquez-Beggs’s correlation bubble point pressure, 271–272, 272b gas solubility, 263–264, 264b oil FVF, 277–278 undersaturated oil compressibility estimation, 289, 290b undersaturated oil viscosity, 309–311 VdW EOS. See Van der Waals equation of state (VdW EOS) Viscosity. See also specific types of viscosity correlations from oil composition, 311–315, 317b Little-Kennedy’s correlation, 314–315 Lohrenz-Bray-Clark’s (LBC) correlation, 312–314 of gas, 223–224, 227b slim tube test, 443–444 water, 324–326 Volatile crude oil, 39 liquid-shrinkage curve for, 41f p-T diagram for, 40f Volume average boiling point (VABP), 122–123, 134–135 Volume fraction, 192 Volume translation method, SRK EOS, 531–534, 533b Index 607 W Watansiri-Owens-Starling correlations, 99–100 petroleum fraction, 115 Water-oil contact (WOC), 56 Watson characterization factor, and true boiling point, 80–82, 84 Weight average boiling point (WABP), 122–123 Weight fraction, 192–193 Well-defined petroleum fractions, 73–76 Wet gas reservoirs, 47–50, 47f, 50b specific gravity, 230–234, 232b Whitson’s splitting method, 155–159, 158b Wichert-Aziz’s correction method, 203–205 Willman and Teja correlations, 103–104 petroleum fraction, 116 Wilson’s correlation, 476–477 Winn and Sim-Daubert correlation, 99 petroleum fraction, 114 Winn’s method, 486–487 WOC. See Water-oil contact (WOC) Y Yellig and Metcalfe correlation, MMP, 449, 449b Y-function extended, 366, 368b, 370f traditional, 363, 364f, 364b Yuan’s correlation, MMP, 451–452 Z Z-factor nonhydrocarbon components effect, 196–203 selection, 527–529 Standing-Katz chart, 212