Skip to main content

Basics of Oil and Gas Flow in Reservoirs

  • Chapter
  • First Online:
Fundamentals and Practical Aspects of Gas Injection

Part of the book series: Petroleum Engineering ((PEEN))

  • 538 Accesses

Abstract

This chapter deals with basic concepts of oil and gas flow in porous media. In the first section (oil flow), the basics of oil flow in porous media and different flow equations, including viscous, Darcy, and Brinkman flow are introduced. Then, based on a new work the boundary between these kinds of flow are reported. Then, the well-known diffusivity equation, combination of continuity equation and Darcy velocity, is introduced. The relative permeability as an important concept in two-phase flow through porous media is described as well. After that, the forces affecting oil flow in porous media are introduced as well as different dimensionless numbers which reflect the relative importance of forces during oil flow. The gas flow through porous reservoir can occur as single or multiphase. A second phase releases as a result of pressure decline in reservoir and starts flowing when its saturation exceeds a minimum value, known as critical saturation. At these conditions, the relative permeability concept applies to take into account the multiphase flow. Also, the classic Darcy’s law equation for flow in porous media may fail in certain flowing conditions, and the non-Darcy flow equations need to be studied and applied for proper modelling of gas flow in porous medium. In addition, the role of wettability alteration on production enhancement of gas condensate reservoirs is a novel approach in gas reservoir performance. Also, the active forces during CO2 storage and the challenges of gas flow in unconventional gas reservoirs are two important subjects. These subjects will be covered in the second section (gas flow) of this chapter.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

eBook
USD 16.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 129.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 199.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Abbreviations

ANN:

Artificial neural network

BHP:

Bottom-hole pressure, psi

IPR:

Inflow performance curve

LSSVM:

Least square support vector machine

MGGP:

Multi-gene genetic programming

TPR:

Tubing performance curve

C:

Isothermal compressibility, pa1

Ct:

Total compressibility, 1/psi

E:

Non-Darcy effect

Fo:

Forchheimer number

f:

Fractional flow

G:

Cumulative gas (HCPV)

Grw:

Diffusional Grashof number

h:

Formation thickness, m

I0:

Modified Bessel function, first kind, zero-order

I1:

Modified Bessel function, first kind, the first order

k:

Permeability, md

K0:

Modified Bessel function, the second kind, zero-order

K1:

Modified Bessel function, the second kind, the first order

M:

\(\left( {{\text{M}} = \sqrt {{\text{zf}}\left( {\text{z}} \right)} } \right)\)

p:

Pressure, psi

p̄:

Average reservoir pressure, pa

p0:

Initial Pressure, psi

pp:

Pseudo-pressure

Pi:

Initial reservoir pressure, pa

Pe:

Peclet number

PD:

Dimensionless pressure, \(({\text{P}}_{{\text{D}}} = \frac{{2 {\uppi }{\text{K}}_{{\text{f}}} {\text{h}}\left( {{\text{P}}_{{\text{i}}} - {\text{P}}\left( {{\text{r}},{\text{t}}} \right)} \right)}}{{{\text{q}} {\upmu }}})\)

Pi:

Initial reservoir pressure, pa

Pwf:

Wellbore pressure during the drawdown period, pa

PDwf:

Dimensionless wellbore pressure, (\({\text{P}}_{{{\text{Dwf}}}} = \frac{{2 {\uppi }{\text{K}}_{{\text{f}}} {\text{h}}\left( {{\text{P}}_{{\text{i}}} - {\text{Pwf }}} \right)}}{{{\text{q}} {\upmu }}}\))

q:

Production rate, m3 s1

qD:

Dimensionless production, (\({\text{q}}_{{\text{D}}} = \frac{{{\text{q}} {\upmu }{\text{B}}}}{{{\text{k}}_{{\text{f}}} {\text{h}}\left( {{\text{P}}_{{\text{i}}} - {\text{Pwf}}} \right)}}\))

R:

Cylinder Radius, m

Ra:

Rayleigh number

Re:

Reynolds number

r:

Radius, ft

rc:

Pore radius, ft

re:

Reservoir outer boundary radius, m

rw:

Wellbore radius, m

rD:

Dimensionless radius, (\({\text{r}}_{{\text{D}}} = {\text{r}}/{\text{r}}_{{\text{w}}}\))

rDe:

Dimensionless outer boundary radius, (\({\text{r}}_{{{\text{eD}}}} = {\text{r}}_{{\text{e}}} /{\text{r}}_{{\text{w}}}\))

S:

A characteristic coefficient of the fractured rock proportional to the specific surface of the block

Sc:

Schmidt number

t:

Time, s

tD:

Dimensionless time, (\({\text{t}}_{{\text{D}}} = \frac{{{\text{K}}_{{\text{f}}} {\text{t}}}}{{\left[ {\left( {\emptyset {\text{C}}} \right)_{{\text{m}}} + \left( {\emptyset {\text{C}}} \right)_{{\text{f}}} } \right] {\upmu }{\text{r}}_{{\text{w}}}^{2} }}\))

U:

Velocity, ms1

U*:

Rate of mass flow per unit volume, represents the transfer of fluid between blocks and fractures, kg/s m3, (\({\text{U}}^{*} = \frac{{ {\uprho }{\text{SK}}_{{\text{m}}} \left( {{\text{P}}_{{\text{m}}} - {\text{P}}_{{\text{f}}} } \right)}}{\upmu }\))

U̅:

Flux, S1

v:

Velocity, ft/s

\(\overline{v}\) :

Average velocity (m/s)

v g :

General solution

v p :

Particular solution

z:

Compressibility factor/Laplace space variable

μ:

Viscosity, cp

β:

Coefficient of quadratic velocity term

γ:

Coefficient of cubic velocity term

σ:

Interfacial tension, dyne/ft

ρ:

Density, lb/ft3

v:

Velocity (ft/d)

Ø:

Porosity

λ:

Dimensionless matrix/fracture permeability ratio, (\({\uplambda } = \frac{{ {\upalpha }{\text{K}}_{1} {\text{r}}_{{\text{w}}}^{2} }}{{{\text{K}}_{2} }}\))

ω:

Fracture storage, (\({\upomega } = \frac{{\emptyset_{2} {\text{C}}_{2} }}{{\emptyset_{1} {\text{C}}_{1} + \emptyset_{2} {\text{C}}_{2} }}\))

θ:

Rock wettability

α:

Angle of deviation/interporosity flow shape factor, m2

c:

Capillary

D:

Dimensionless

f:

Fracture

gde:

Dry gas

gw:

Wet gas

id:

Dry gas injected

m:

Matrix

pD:

Recovery of wet gas

r:

Relative

W:

Wellbore

wc:

Critical water

References

  1. Bird RB, Stewart WE, Lightfoot EN. Transport phenomena, 2nd ed. (2002) n.d.

    Google Scholar 

  2. Brown GO. Henry Darcy and the making of a law. Water Resour Res. 2002;38:11.

    Article  Google Scholar 

  3. Saidi AM. Realism of flow mechanism in reservoirs. 2013.

    Google Scholar 

  4. Løvoll G, Méheust Y, Måløy KJ, Aker E, Schmittbuhl J. Competition of gravity, capillary and viscous forces during drainage in a two-dimensional porous medium, a pore scale study. Energy. 2005;30:861–72.

    Article  Google Scholar 

  5. Dake LP. Fundamentals of reservoir engineering, volume 8 of. Dev Pet Sci. 1978.

    Google Scholar 

  6. Neuman SP. Theoretical derivation of Darcy’s law. Acta Mech. 1977;25:153–70.

    Article  MATH  Google Scholar 

  7. Saidi AM. Reservoir engineering of fractured reservoirs (fundamental and Practical Aspects). Total; 1987.

    Google Scholar 

  8. Ahmed T. Reservoir engineering handbook. Gulf professional publishing; 2018.

    Google Scholar 

  9. Gavin L. Pre-Darcy flow: a missing piece of the improved oil recovery puzzle? In: SPE/DOE symposium on improved oil recovery. Society of Petroleum Engineers; 2004.

    Google Scholar 

  10. Amyx J, Bass D, Whiting RL. Petroleum reservoir engineering physical properties. 1960.

    Google Scholar 

  11. Ahmed T, McKinney P. Advanced reservoir engineering. Elsevier; 2011.

    Google Scholar 

  12. Klinkenberg LJ. The permeability of porous media to liquids and gases. In: Drilling and production practice. American Petroleum Institute; 1941.

    Google Scholar 

  13. Azin R, Mohamadi-Baghmolaei M, Sakhaei Z. Parametric analysis of diffusivity equation in oil reservoirs. J Pet Explor Prod Technol. 2017;7:169–79.

    Article  Google Scholar 

  14. Basak P. Non-Darcy flow and its implications to seepage problems. J Irrig Drain Div. 1977;103:459–73.

    Article  Google Scholar 

  15. Soni JP, Islam N, Basak P. An experimental evaluation of non-Darcian flow in porous media. J Hydrol. 1978;38:231–41.

    Article  Google Scholar 

  16. Xue-wu W, Zheng-ming Y, Yu-ping S, Xue-wei L. Experimental and theoretical investigation of nonlinear flow in low permeability reservoir. Procedia Environ Sci. 2011;11:1392–9.

    Article  Google Scholar 

  17. Siddiqui F, Soliman MY, House W, Ibragimov A. Pre-Darcy flow revisited under experimental investigation. J Anal Sci Technol. 2016;7:1–9.

    Article  Google Scholar 

  18. Prada A, Civan F. Modification of Darcy’s law for the threshold pressure gradient. J Pet Sci Eng. 1999;22:237–40.

    Article  Google Scholar 

  19. Hekecioglu I, Jiang Y. Flow through porous media of packed spheres saturated within water. J Fluids Eng. 1994;116:164–70.

    Article  Google Scholar 

  20. Civan F. Porous media transport phenomena. Wiley; 2011.

    Google Scholar 

  21. Evans RD, Civan F. Characterization of non-Darcy multiphase flow in petroleum bearing formation. Final report. Oklahoma Univ., Norman, OK (United States). School of Petroleum and …; 1994.

    Google Scholar 

  22. Dye AL, McClure JE, Miller CT, Gray WG. Description of non-Darcy flows in porous medium systems. Phys Rev E. 2013;87:33012.

    Article  Google Scholar 

  23. Mitchell J, Younger J. Abnormalities in hydraulic flow through fine-grained soils. Permeability and capillarity of soils. ASTM International; 1967.

    Google Scholar 

  24. Yu RZ, Bian YN, Lei Q, Yang ZM, Wang KJ. Low-velocity non-Darcy flow numerical simulation of the Da45 reservoir, Honggang Oilfield, Jilin, China. Pet Sci Technol. 2013;31:1617–24.

    Article  Google Scholar 

  25. Farmani Z, Azin R, Fatehi R, Escrochi M. Analysis of pre-Darcy flow for different liquids and gases. J Pet Sci Eng. 2018;168:17–31.

    Article  Google Scholar 

  26. Farmani Z, Farokhian D, Izadpanahi A, seifi F, Zahedizadeh P, Safari Z, et al. Pre-Darcy flow and Klinkenberg effect in dense, consolidated carbonate formations. Geotech Geol Eng. 2019;37. https://doi.org/10.1007/s10706-019-00841-0.

  27. Brinkman HC. A calculation of the viscous force exerted by a flowing fluid on a dense swarm of particles. Flow, Turbul Combust. 1949;1:27.

    Article  MATH  Google Scholar 

  28. De Lemos MJS. Turbulence in porous media: modeling and applications. Elsevier; 2012.

    Google Scholar 

  29. Dybbs A, Edwards RV. A new look at porous media fluid mechanics—Darcy to turbulent. Fundamentals of transport phenomena in porous media. Springer; 1984. p. 199–256

    Google Scholar 

  30. Hubbert MK. Darcy’ law: its physical theory and application to entrapment of oil and gas. History of geophysics: volume 3: the history of hydrology, vol. 3. 1987. p. 1–26.

    Google Scholar 

  31. Zeng Z, Grigg R. A criterion for non-Darcy flow in porous media. Transp Porous Media. 2006;63:57–69.

    Article  Google Scholar 

  32. Zimmerman RW, Al-Yaarubi A, Pain CC, Grattoni CA. Non-linear regimes of fluid flow in rock fractures. Int J Rock Mech Min Sci. 2004;41:163–9.

    Article  Google Scholar 

  33. Hassanizadeh SM, Gray WG. High velocity flow in porous media. Transp Porous Media. 1987;2:521–31.

    Article  Google Scholar 

  34. Zahedizadeh P, Safari Z, Behnamnia M, Sadeqi-Murche-khorti R, Azadian H, Mirghayed BR, et al. Boundary determination between darcy, brinkman, and viscous flows using effective parameters. Appl Math Comput 2021;Under Revi.

    Google Scholar 

  35. Terry RE, Rogers JB, Craft BC. Applied petroleum reservoir engineering. Pearson Education; 2015.

    Google Scholar 

  36. Mosavat N, Torabi F, Zarivnyy O. Developing new Corey-based water/oil relative permeability correlations for heavy oil systems. In: SPE heavy oil conference. Society of Petroleum Engineers; 2013.

    Google Scholar 

  37. Torabi F, Mosavat N, Zarivnyy O. Predicting heavy oil/water relative permeability using modified Corey-based correlations. Fuel. 2016;163:196–204.

    Article  Google Scholar 

  38. Sufi AH, Ramey Jr HJ, Brigham WE. Temperature effects on relative permeabilities of oil-water systems. In: SPE annual technical conference and exhibition. Society of Petroleum Engineers; 1982.

    Google Scholar 

  39. Honarpour MM. Relative permeability of petroleum reservoirs. CRC press; 2018.

    Google Scholar 

  40. Xu J, Guo C, Jiang R, Wei M. Study on relative permeability characteristics affected by displacement pressure gradient: experimental study and numerical simulation. Fuel. 2016;163:314–23.

    Article  Google Scholar 

  41. Dake LP. The practice of reservoir engineering. vol. 36. Elsevier; 2013.

    Google Scholar 

  42. Mohamadi-Baghmolaei M, Azin R, Sakhaei Z, Mohamadi-Baghmolaei R, Osfouri S. Novel method for estimation of gas/oil relative permeabilities. J Mol Liq. 2016;224:1109–16.

    Article  Google Scholar 

  43. Purcell WR. Capillary pressures-their measurement using mercury and the calculation of permeability therefrom. J Pet Technol. 1949;1:39–48.

    Article  Google Scholar 

  44. Fatt I, Dykstra H. Relative permeability studies. J Pet Technol 1951;3:249–56.

    Google Scholar 

  45. Burdine N. Relative permeability calculations from pore size distribution data. J Pet Technol. 1953;5:71–8.

    Article  Google Scholar 

  46. Corey AT. The interrelation between gas and oil relative permeabilities. Prod Mon. 1954;19:38–41.

    Google Scholar 

  47. Wyllie MRJ, Gardner GHF. The generalized Kozeny-Carman equation. Part 2. A novel approach to problems of fluid flow. World Oil 1958;146:210–28.

    Google Scholar 

  48. Wahl WL, Mullins LD, Elfrink EB. Estimation of ultimate recovery from solution gas-drive reservoirs. Trans AIME. 1958;213:132–8.

    Article  Google Scholar 

  49. Torcaso MA, Wyllie MRJ. A comparison of calculated krg/kro ratios with a correlation of field data. J Pet Technol. 1958;10:57–8.

    Article  Google Scholar 

  50. Brooks RH, Corey AT. Properties of porous media affecting fluid flow. J Irrig Drain Div. 1966;92:61–90.

    Article  Google Scholar 

  51. Honarpour M, Koederitz LF, Harvey AH. Empirical equations for estimating two-phase relative permeability in consolidated rock. J Pet Technol. 1982;34:2–905.

    Article  Google Scholar 

  52. Ibrahim MN, Koederitz LF. Two-phase relative permeability prediction using a linear regression model. In: SPE eastern regional meeting. Society of Petroleum Engineers; 2000.

    Google Scholar 

  53. Mulyadi H, Amin R, Kennaird AF. Practical approach to determine residual gas saturation and gas-water relative permeability. In: SPE annual technical conference and exhibition. Society of Petroleum Engineers; 2001.

    Google Scholar 

  54. Lomeland F, Ebeltoft E, Thomas WH. A new versatile relative permeability correlation. Paper SCA 2005-32 presented at international symposium of the Society of Core Analysts, Toronto, Canada; 2005.

    Google Scholar 

  55. Sigmund PM, McCaffery FG. An improved unsteady-state procedure for determining the relative-permeability characteristics of heterogeneous porous media (includes associated papers 8028 and 8777). Soc Pet Eng J. 1979;19:15–28.

    Article  Google Scholar 

  56. Chierici GL, Ciucci GM, Sclocchi G. Two-phase vertical flow in oil wells-prediction of pressure drop. J Pet Technol. 1974;26:927–38.

    Article  Google Scholar 

  57. Chierici GL. Novel relations for drainage and imbibition relative permeabilities. Soc Pet Eng J. 1984;24:275–6.

    Article  Google Scholar 

  58. Bastian P. Numerical computation of multiphase flow in porous media, Habilitation. Tech Fak Der Christ Kiel. 1999.

    Google Scholar 

  59. Li K, Horne RN. Comparison of methods to calculate relative permeability from capillary pressure in consolidated water‐wet porous media. Water Resour Res 2006;42.

    Google Scholar 

  60. Ahmadi MA, Zendehboudi S, Dusseault MB, Chatzis I. Evolving simple-to-use method to determine water–oil relative permeability in petroleum reservoirs. Petroleum. 2016;2:67–78.

    Article  Google Scholar 

  61. Ahmadi MA. Connectionist approach estimates gas–oil relative permeability in petroleum reservoirs: application to reservoir simulation. Fuel. 2015;140:429–39.

    Article  Google Scholar 

  62. Farmani Z, Azin R, Mohamadi-Baghmolaei M, Fatehi R, Escrochi M. Experimental and theoretical study of gas/oil relative permeability. Comput Geosci. 2019;23:567–81.

    Article  Google Scholar 

  63. van Golf-Racht TD. Fundamentals of fractured reservoir engineering, vol. 12. Elsevier; 1982.

    Google Scholar 

  64. Barenblatt GI, Zheltov IP, Kochina IN. Basic concepts in the theory of seepage of homogeneous liquids in fissured rocks [strata]. J Appl Math Mech. 1960;24:1286–303.

    Article  MATH  Google Scholar 

  65. Warren JE, Root PJ. The behavior of naturally fractured reservoirs. Soc Pet Eng J. 1963;3:245–55. https://doi.org/10.2118/426-PA.

    Article  Google Scholar 

  66. Zahedizadeh P, Safari Z, Ghaderi A, Azin R. Transition zone diagnosis in finite fractured reservoirs. Comput Geosci. 2020;24:1483–96.

    Article  MathSciNet  MATH  Google Scholar 

  67. Dong Z, Li W, Lei G, Wang H, Wang C. Embedded discrete fracture modeling as a method to upscale permeability for fractured reservoirs. Energies. 2019;12:812.

    Article  Google Scholar 

  68. Choi ES, Cheema T, Islam MR. A new dual-porosity/dual-permeability model with non-Darcian flow through fractures. J Pet Sci Eng. 1997;17:331–44.

    Article  Google Scholar 

  69. Hoteit H, Firoozabadi A. Multicomponent fluid flow by discontinuous Galerkin and mixed methods in unfractured and fractured media. Water Resour Res 2005;41.

    Google Scholar 

  70. Mousavi Nezhad M, Javadi AA, Abbasi F. Stochastic finite element modelling of water flow in variably saturated heterogeneous soils. Int J Numer Anal Methods Geomech. 2011;35:1389–408.

    Article  Google Scholar 

  71. Karimi-Fard M, Durlofsky LJ, Aziz K. An efficient discrete fracture model applicable for general purpose reservoir simulators. In: SPE reservoir simulation symposium proceedings. Society of Petroleum Engineers; 2003.

    Google Scholar 

  72. Fumagalli A, Pasquale L, Zonca S, Micheletti S. An upscaling procedure for fractured reservoirs with embedded grids. Water Resour Res. 2016;52:6506–25.

    Article  Google Scholar 

  73. Ding M, Kantzas A. Capillary number correlations for gas-liquid systems. J Can Pet Technol 2007;46.

    Google Scholar 

  74. Kiani M, Osfouri S, Azin R, Dehghani SAM. Impact of fluid characterization on compositional gradient in a volatile oil reservoir. J Pet Explor Prod Technol. 2016;6:835–44.

    Article  Google Scholar 

  75. Kiani M, Osfouri S, Azin R, Dehghani SAM. A comprehensive model for the prediction of fluid compositional gradient in two-dimensional porous media. J Pet Explor Prod Technol. 2019;9:2221–34.

    Article  Google Scholar 

  76. Azin R, Raad SMJ, Osfouri S, Fatehi R. Onset of instability in CO 2 sequestration into saline aquifer: scaling relationship and the effect of perturbed boundary. Heat Mass Transf. 2013;49:1603–12.

    Article  Google Scholar 

  77. Raad SMJ, Fatehi R, Azin R, Osfouri S, Bahadori A. Linear perturbation analysis of density change caused by dissolution of carbon dioxide in saline aqueous phase. J Mol Liq. 2015;209:539–48.

    Article  Google Scholar 

  78. Louriou C, Ouerfelli H, Prat M, Najjari M, Nasrallah S Ben. Gas injection in a liquid saturated porous medium. Influence of gas pressurization and liquid films. Transp Porous Media 2012;91:153–71.

    Google Scholar 

  79. Saidi AM. Twenty years of gas injection history into well-fractured Haft Kel field (Iran). In: International petroleum conference and exhibition of Mexico. Society of Petroleum Engineers; 1996. https://doi.org/10.2118/35309-MS.

  80. Alkindi AS, Muggeridge A, Al-Wahaibi YM. The influence of diffusion and dispersion on heavy oil recovery by VAPEX. In: International thermal operations and heavy oil symposium. Society of Petroleum Engineers; 2008.

    Google Scholar 

  81. Azin R, Kharrat R, Ghotbi C, Rostami B, Vossoughi S. Simulation study of the VAPEX process in fractured heavy oil system at reservoir conditions. J Pet Sci Eng. 2008;60:51–66.

    Article  Google Scholar 

  82. Etminan SR, Haghighat P, Maini BB, Chen ZJ. Molecular diffusion and dispersion coefficient in a propane-bitumen system: case of vapour extraction (VAPEX) process. In: SPE EUROPEC/EAGE annual conference and exhibition. Society of Petroleum Engineers; 2011.

    Google Scholar 

  83. Adewale A, Olalekan O, Kelani B, Abass I. Effects of Peclet number on miscible displacement process through the reservoir. In: Nigeria annual international conference and exhibition. Society of Petroleum Engineers; 2004.

    Google Scholar 

  84. Patel RD, Greenkorn RA. Prediction of recovery in miscible displacement in porous media using the dispersion equation. In: Fall meeting of the Society of Petroleum Engineers of AIME. Society of Petroleum Engineers; 1969.

    Google Scholar 

  85. Boustani A, Maini BB. The role of diffusion and convective dispersion in vapour extraction process. J Can Pet Technol 2001;40.

    Google Scholar 

  86. Sureshjani MH, Azin R, Gerami S. A roadmap for production data analysis in an Iranian offshore gas-condensate field n.d.

    Google Scholar 

  87. Heidari Sureshjani MH, Azin R, Lak A, Osfouri S, Chahshoori R, Sadeghi F, et al. Production data analysis in a gas-condensate field: methodology, challenges and uncertainties. Iran J Chem Chem Eng. 2016;35:113–27.

    Google Scholar 

  88. Sarvestani MT, Sureshjani MH, Gerami S, Azin R. Some remarks on production data analysis in layered gas reservoirs. n.d.

    Google Scholar 

  89. Wattenbarger RA. Well performance equations (1987 PEH Chapter 35). Pet Eng Handb. 1987.

    Google Scholar 

  90. Ertekin T, Abou-Kassem JH, King GR. Basic applied reservoir simulation. 2001.

    Google Scholar 

  91. Firoozabadi A, Katz DL. An analysis of high-velocity gas flow through porous media. J Pet Technol. 1979;31:211–6.

    Article  Google Scholar 

  92. Tek MR, Coats KH, Katz DL. The effect of turbulence on flow of natural gas through porous reservoirs. J Pet Technol. 1962;14:799–806.

    Article  Google Scholar 

  93. Ezeudembah AS, Dranchuk PM. Flow mechanism of Forchheimer’s cubic equation in high-velocity radial gas flow through porous media. In: SPE annual technical conference and exhibition. Society of Petroleum Engineers; 1982.

    Google Scholar 

  94. Holditch SA, Morse RA. The effects of non-Darcy flow on the behavior of hydraulically fractured gas wells (includes associated paper 6417). J Pet Technol. 1976;28:1–169.

    Article  Google Scholar 

  95. Skjetne E, Auriault J-L. High-velocity laminar and turbulent flow in porous media. Transp Porous Media. 1999;36:131–47.

    Article  MathSciNet  MATH  Google Scholar 

  96. Eburi Losoha S. Mathematical modeling, analysis and simulation of the productivity index for non-linear flow in porous media, with applications in reservoir engineering. 2007.

    Google Scholar 

  97. Saboorian-Jooybari H, Pourafshary P. Significance of non-Darcy flow effect in fractured tight reservoirs. J Nat Gas Sci Eng. 2015;24:132–43.

    Article  Google Scholar 

  98. Wu Y-S, Pruess K. Gas flow in porous media with Klinkenberg effects. Transp Porous Media. 1998;32:117–37.

    Article  Google Scholar 

  99. Whitson CH, Fevang O. Modeling gas condensate well deliverability, SPE 30714. 1995.

    Google Scholar 

  100. Fan L, Harris BW, Jamaluddin AJ. Understanding gas-condensate reservoirs. Oilf Rev. 2005:14–27. https://doi.org/10.4043/25710-MS.

  101. Jamiolahmady M, Danesh A, Tehrani DH, Duncan DB. A mechanistic model of gas-condensate flow in pores. Transp Porous Media. 2000;41:17–46.

    Article  Google Scholar 

  102. Henderson GD, Danesh A, Tehrani DH, Al-Kharusi B. The relative significance of positive coupling and inertial effects on gas condensate relative permeabilities at high velocity. In: SPE annual technical conference and exhibition. Society of Petroleum Engineers; 2000.

    Google Scholar 

  103. Mott R, Cable A, Spearing M. Measurements and simulation of inertial and high capillary number flow phenomena in gas-condensate relative permeability. In: SPE annual technical conference and exhibition. Society of Petroleum Engineers; 2000.

    Google Scholar 

  104. Danesh A, et al. Gas condensate recovery studies. 1994.

    Google Scholar 

  105. Longeron DG. Influence of very low interfacial tensions on relative permeability. Soc Pet Eng J. 1980;20:391–401.

    Article  Google Scholar 

  106. Coats KH. SPE 8284. An equation of state compositional model. 1980.

    Google Scholar 

  107. Amaefule JO, Handy LL. The effect of interfacial tensions on relative oil/water permeabilities of consolidated porous media. Soc Pet Eng J. 1982;22:371–81.

    Article  Google Scholar 

  108. Harbert LW. Low interfacial tension relative permeability. In: SPE annual technical conference and exhibition. Society of Petroleum Engineers; 1983.

    Google Scholar 

  109. Jamiolahmady M, Danesh A, Tehrani DH, Duncan DB. Positive effect of flow velocity on gas–condensate relative permeability: network modelling and comparison with experimental results. Transp Porous Media. 2003;52:159–83.

    Article  Google Scholar 

  110. Jamiolahmady M, Sohrabi M, Ireland S, Ghahri P. A generalized correlation for predicting gas–condensate relative permeability at near wellbore conditions. J Pet Sci Eng. 2009;66:98–110.

    Article  Google Scholar 

  111. Fulcher RA Jr, Ertekin T, Stahl CD. Effect of capillary number and its constituents on two-phase relative permeability curves. J Pet Technol. 1985;37:249–60.

    Article  Google Scholar 

  112. Asar H, Handy LL. Influence of interfacial tension on gas/oil relative permeability in a gas-condensate system. SPE Reserv Eng. 1988;3:257–64.

    Article  Google Scholar 

  113. Munkerud PK, Torsaeter O. The effects of interfacial tension and spreading on relative permeability in gas condensate systems. In: IOR 1995—8th European symposium on improved oil recovery. European Association of Geoscientists & Engineers; 1995. p. cp-107.

    Google Scholar 

  114. Henderson GD, Danesh A, Tehrani DH, Al-Shaidi S, Peden JM. Measurement and correlation of gas condensate relative permeability by the steady-state method. SPE J. 1996;1:191–202.

    Article  Google Scholar 

  115. Henderson GD, Danesh A, Tehrani DH, Al-Shaidi S, Peden JM. Measurement and correlation of gas condensate relative permeability by the steady-state method. SPE Reserv Eval Eng. 1998;1:134–40.

    Article  Google Scholar 

  116. Mott R, Cable A, Spearing M. A new method of measuring relative permeabilities for calculating gas-condensate well deliverability. In: SPE annual technical conference and exhibition. Society of Petroleum Engineers; 1999.

    Google Scholar 

  117. Chen HL, Wilson SD, Monger-McClure TG. Determination of relative permeability and recovery for North Sea gas condensate reservoirs. In: SPE annual technical conference and exhibition. Society of Petroleum Engineers; 1995.

    Google Scholar 

  118. Blom SMP, Hagoort J, Soetekouw DPN. Relative permeability at near-critical conditions. In: SPE annual technical conference and exhibition. Society of Petroleum Engineers; 1997.

    Google Scholar 

  119. Henderson GD, Danesh A, Tehrani DH, Peden JM. The effect of velocity and interfacial tension on relative permeability of gas condensate fluids in the wellbore region. J Pet Sci Eng. 1997;17:265–73.

    Article  Google Scholar 

  120. Delclaud J, Rochon J, Nectoux A. Investigation of gas/oil relative permeabilities: high-permeability oil reservoir application. In: SPE annual technical conference and exhibition. Society of Petroleum Engineers; 1987.

    Google Scholar 

  121. Kalaydjian F-M, Bourbiaux BJ, Lombard JM. Predicting gas-condensate reservoir performance: how flow parameters are altered when approaching production wells. In: SPE annual technical conference and exhibition. Society of Petroleum Engineers; 1996.

    Google Scholar 

  122. Jamiolahmady M, Sohrabi M, Ghahri P, Ireland S. Gas/condensate relative permeability of a low permeability core: coupling vs. inertia. SPE Reserv Eval Eng 2010;13:214–27.

    Google Scholar 

  123. Thomas O. Reservoir analysis based on compositional gradients. 2007.

    Google Scholar 

  124. Patzek TW, Silin DB, Benson SM, Barenblatt GI. On vertical diffusion of gases in a horizontal reservoir. Transp Porous Media. 2003;51:141–56.

    Article  Google Scholar 

  125. Kuchuk FJ, Saeedi J. inflow performance of horizontal wells in multilayer reservoirs. In: SPE annual technical conference and exhibition. Society of Petroleum Engineers; 1992.

    Google Scholar 

  126. Jatmiko W, Daltaban TS, Archer JS. Multi-phase flow well test analysis in multi-layer reservoirs. In: SPE annual technical conference and exhibition. Society of Petroleum Engineers; 1996.

    Google Scholar 

  127. Nikjoo E, Hashemi A. Pressure transient analysis in multiphase multi layer reservoirs with inter layer communication. In: SPE Europec/EAGE annual conference. Society of Petroleum Engineers; 2012.

    Google Scholar 

  128. Fetkovich MJ, Bradley MD, Works AM, Thrasher TS. Depletion performance of layered reservoirs without crossflow. SPE Form Eval. 1990;5:310–8.

    Article  Google Scholar 

  129. Park H, Horne RN. Well test analysis of a multilayered reservoir with formation crossflow. In: SPE annual technical conference and exhibition. Society of Petroleum Engineers; 1989.

    Google Scholar 

  130. Azin R, Sedaghati H, Fatehi R, Osfouri S, Sakhaei Z. Production assessment of low production rate of well in a supergiant gas condensate reservoir: application of an integrated strategy. J Pet Explor Prod Technol. 2019;9:543–60.

    Article  Google Scholar 

  131. Ahmed T, Evans J, Kwan R, Vivian T. Wellbore liquid blockage in gas-condensate reservoirs. In: SPE eastern regional meeting. Society of Petroleum Engineers; 1998.

    Google Scholar 

  132. Jamaluddin AKM, Ye S, Thomas J, D’Cruz D, Nighswander J. Experimental and theoretical assessment of using propane to remediate liquid buildup in condensate reservoirs. In: SPE annual technical conference and exhibition. Society of Petroleum Engineers; 2001.

    Google Scholar 

  133. Kossack CA, Opdal ST. Recovery of condensate from a heterogeneous reservoir by the injection of a slug of methane followed by nitrogen. In: SPE annual technical conference and exhibition. Society of Petroleum Engineers; 1988.

    Google Scholar 

  134. Marokane D, Logmo-Ngog AB, Sarkar R. Applicability of timely gas injection in gas condensate fields to improve well productivity. In: SPE/DOE improved oil recovery symposium. Society of Petroleum Engineers; 2002.

    Google Scholar 

  135. Narinesingh J, Alexander D. CO2 enhanced gas recovery and geologic sequestration in condensate reservoir: a simulation study of the effects of injection pressure on condensate recovery from reservoir and CO2 storage efficiency. Energy Procedia. 2014;63:3107–15.

    Article  Google Scholar 

  136. Sanger PJ, Hagoort J. Recovery of gas-condensate by nitrogen injection compared with methane injection. SPE J. 1998;3:26–33.

    Article  Google Scholar 

  137. Wang X, Indriati S, Valko PP, Economides MJ. Production impairment and purpose-built design of hydraulic fractures in gas-condensate reservoirs. In: International oil and gas conference and exhibition in China. Society of Petroleum Engineers; 2000.

    Google Scholar 

  138. Dehane A, Tiab D, Osisanya SO. Comparison of the performance of vertical and horizontal wells in gas-condensate reservoirs. In: SPE annual technical conference and exhibition. Society of Petroleum Engineers; 2000.

    Google Scholar 

  139. Asgari A, Dianatirad M, Ranjbaran M, Sadeghi AR, Rahimpour MR. Methanol treatment in gas condensate reservoirs: a modeling and experimental study. Chem Eng Res Des. 2014;92:876–90.

    Article  Google Scholar 

  140. Wang H, Chen E, Jia X, Liang L, Wang Q. Superhydrophobic coatings fabricated with polytetrafluoroethylene and SiO2 nanoparticles by spraying process on carbon steel surfaces. Appl Surf Sci. 2015;349:724–32.

    Article  Google Scholar 

  141. Mousavi MA, Hassanajili S, Rahimpour MR. Synthesis of fluorinated nano-silica and its application in wettability alteration near-wellbore region in gas condensate reservoirs. Appl Surf Sci. 2013;273:205–14.

    Article  Google Scholar 

  142. Kewen L, Abbas F. Experimental study of wettability alteration to preferential gas-wetting in porous media and its effects. SPE Reserv Eval Eng. 2000;3:139–49.

    Article  Google Scholar 

  143. Zhang S, Jiang G-C, Wang L, Qing W, Guo H-T, Tang X, et al. Wettability alteration to intermediate gas-wetting in low-permeability gas-condensate reservoirs. J Pet Explor Prod Technol. 2014;4:301–8.

    Article  Google Scholar 

  144. Jin J, Wang Y, Wang K, Ren J, Bai B, Dai C. The effect of fluorosurfactant-modified nano-silica on the gas-wetting alteration of sandstone in a CH4-liquid-core system. Fuel. 2016;178:163–71.

    Article  Google Scholar 

  145. Li K, Liu Y, Zheng H, Huang G, Li G. Enhanced gas-condensate production by wettability alteration to gas wetness. J Pet Sci Eng. 2011;78:505–9.

    Article  Google Scholar 

  146. Zoghbi B, Fahes MM, Nasrabadi H. Identifying the optimum Wettability conditions for the near-wellbore region in gas-condensate reservoirs. In: Tight gas completions conference. Society of Petroleum Engineers; 2010.

    Google Scholar 

  147. Sheydaeemehr M, Sedaeesola B, Vatani A. Gas-condensate production improvement using wettability alteration: a giant gas condensate field case study. J Nat Gas Sci Eng. 2014;21:201–8.

    Article  Google Scholar 

  148. Butler M, Trueblood JB, Pope GA, Sharma MM, Baran Jr JR, Johnson D. A field demonstration of a new chemical stimulation treatment for fluid-blocked gas wells. In: SPE annual technical conference and exhibition. Society of Petroleum Engineers; 2009.

    Google Scholar 

  149. Ehtesabi H, Ahadian MM, Taghikhani V, Ghazanfari MH. Enhanced heavy oil recovery in sandstone cores using TiO2 nanofluids. Energy Fuels. 2014;28:423–30.

    Article  Google Scholar 

  150. Giraldo J, Benjumea P, Lopera S, Cortés FB, Ruiz MA. Wettability alteration of sandstone cores by alumina-based nanofluids. Energy Fuels. 2013;27:3659–65.

    Article  Google Scholar 

  151. Hendraningrat L, Torsæter O. Effects of the initial rock wettability on silica-based nanofluid-enhanced oil recovery processes at reservoir temperatures. Energy Fuels. 2014;28:6228–41.

    Article  Google Scholar 

  152. Saboori R, Azin R, Osfouri S, Sabbaghi S, Bahramian A. Wettability alteration of carbonate cores by alumina-nanofluid in different base fluids and temperature. J Sustain Energy Eng. 2018;6:84–98.

    Article  Google Scholar 

  153. Basu BJ, Kumar VD, Anandan C. Surface studies on superhydrophobic and oleophobic polydimethylsiloxane–silica nanocomposite coating system. Appl Surf Sci. 2012;261:807–14.

    Article  Google Scholar 

  154. Zheng Y, He Y, Qing Y, Zhuo Z, Mo Q. Formation of SiO2/polytetrafluoroethylene hybrid superhydrophobic coating. Appl Surf Sci. 2012;258:9859–63.

    Article  Google Scholar 

  155. Nimittrakoolchai O-U, Supothina S. Preparation of stable ultrahydrophobic and superoleophobic silica-based coating. J Nanosci Nanotechnol. 2012;12:4962–8.

    Article  Google Scholar 

  156. Gao Q, Zhu Q, Guo Y, Yang CQ. Formation of highly hydrophobic surfaces on cotton and polyester fabrics using silica sol nanoparticles and nonfluorinated alkylsilane. Ind Eng Chem Res. 2009;48:9797–803.

    Article  Google Scholar 

  157. Lakshmi RV, Bera P, Anandan C, Basu BJ. Effect of the size of silica nanoparticles on wettability and surface chemistry of sol–gel superhydrophobic and oleophobic nanocomposite coatings. Appl Surf Sci. 2014;320:780–6.

    Article  Google Scholar 

  158. Li K, Zeng X, Li H, Lai X. Fabrication and characterization of stable superhydrophobic fluorinated-polyacrylate/silica hybrid coating. Appl Surf Sci. 2014;298:214–20.

    Article  Google Scholar 

  159. Hsieh C-T, Lai M-H, Cheng Y-S. Fabrication and superhydrophobicity of fluorinated titanium dioxide nanocoatings. J Colloid Interface Sci. 2009;340:237–42.

    Article  Google Scholar 

  160. Pazokifard S, Mirabedini SM, Esfandeh M, Farrokhpay S. Fluoroalkylsilane treatment of TiO2 nanoparticles in difference pH values: Characterization and mechanism. Adv Powder Technol. 2012;23:428–36.

    Article  Google Scholar 

  161. Yildirim A, Budunoglu H, Daglar B, Deniz H, Bayindir M. One-pot preparation of fluorinated mesoporous silica nanoparticles for liquid marble formation and superhydrophobic surfaces. ACS Appl Mater Interfaces. 2011;3:1804–8.

    Article  Google Scholar 

  162. Brassard J-D, Sarkar DK, Perron J. Synthesis of monodisperse fluorinated silica nanoparticles and their superhydrophobic thin films. ACS Appl Mater Interfaces. 2011;3:3583–8.

    Article  Google Scholar 

  163. Ramezani M, Vaezi MR, Kazemzadeh A. Preparation of silane-functionalized silica films via two-step dip coating sol–gel and evaluation of their superhydrophobic properties. Appl Surf Sci. 2014;317:147–53.

    Article  Google Scholar 

  164. Schaeffer DA, Polizos G, Smith DB, Lee DF, Hunter SR, Datskos PG. Optically transparent and environmentally durable superhydrophobic coating based on functionalized SiO2 nanoparticles. Nanotechnology. 2015;26:55602.

    Article  Google Scholar 

  165. Wang H, Fang J, Cheng T, Ding J, Qu L, Dai L, et al. One-step coating of fluoro-containing silica nanoparticles for universal generation of surface superhydrophobicity. Chem Commun. 2008:877–9.

    Google Scholar 

  166. Motlagh NV, Birjandi FC, Sargolzaei J, Shahtahmassebi N. Durable, superhydrophobic, superoleophobic and corrosion resistant coating on the stainless steel surface using a scalable method. Appl Surf Sci. 2013;283:636–47.

    Article  Google Scholar 

  167. Saboori R, Azin R, Osfouri S, Sabbaghi S, Bahramian A. Synthesis of fluorine-doped silica-coating by fluorosilane nanofluid to ultrahydrophobic and ultraoleophobic surface. Mater Res Express. 2017;4:105010.

    Google Scholar 

  168. Saboori R, Azin R, Osfouri S, Sabbaghi S, Bahramian A. Wettability alteration of carbonate rocks from strongly liquid-wetting to strongly gas-wetting by fluorine-doped silica coated by fluorosilane. J Dispers Sci Technol. 2018;39:767–76.

    Article  Google Scholar 

  169. Sakhaei Z, Azin R, Naghizadeh A, Osfouri S, Saboori R, Vahdani H. Application of fluorinated nanofluid for production enhancement of a carbonate gas-condensate reservoir through wettability alteration. Mater Res Express. 2018;5:35008.

    Article  Google Scholar 

  170. Sakhaei Z, Mohamadi-Baghmolaei M, Azin R, Osfouri S. Study of production enhancement through wettability alteration in a super-giant gas-condensate reservoir. J Mol Liq. 2017;233:64–74.

    Article  Google Scholar 

  171. Wu Y-S, Li J, Ding D, Wang C, Di Y. A generalized framework model for the simulation of gas production in unconventional gas reservoirs. SPE J. 2014;19:845–57.

    Article  Google Scholar 

  172. Guo C, Xu J, Wu K, Wei M, Liu S. Study on gas flow through nano pores of shale gas reservoirs. Fuel. 2015;143:107–17.

    Article  Google Scholar 

  173. Sun Z, Li X, Shi J, Zhang T, Sun F. Apparent permeability model for real gas transport through shale gas reservoirs considering water distribution characteristic. Int J Heat Mass Transf. 2017;115:1008–19.

    Article  Google Scholar 

  174. Wu K, Chen Z, Li X, Guo C, Wei M. A model for multiple transport mechanisms through nanopores of shale gas reservoirs with real gas effect–adsorption-mechanic coupling. Int J Heat Mass Transf. 2016;93:408–26.

    Article  Google Scholar 

  175. Sun Z, Shi J, Wang K, Miao Y, Zhang T, Feng D, et al. The gas-water two phase flow behavior in low-permeability CBM reservoirs with multiple mechanisms coupling. J Nat Gas Sci Eng. 2018;52:82–93.

    Article  Google Scholar 

  176. Clarkson CR, Qanbari F. A semi-analytical method for forecasting wells completed in low permeability, undersaturated CBM reservoirs. J Nat Gas Sci Eng. 2016;30:19–27.

    Article  Google Scholar 

  177. Wang G, Ren T, Wang K, Zhou A. Improved apparent permeability models of gas flow in coal with Klinkenberg effect. Fuel. 2014;128:53–61.

    Article  Google Scholar 

  178. Lee J. Well testing. Society of Petroleum Engineers; 1982.

    Google Scholar 

  179. Odeh AS, Babu DK. Comparison of solutions of the nonlinear and linearized diffusion equations. SPE Reserv Eng. 1988;3:1–202.

    Article  Google Scholar 

  180. Van Everdingen AF, Hurst W. The application of the Laplace transformation to flow problems in reservoirs. J Pet Technol. 1949;1:305–24. https://doi.org/10.2118/949305-G.

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Reza Azin .

Editor information

Editors and Affiliations

Appendices

Appendix 3.1: The Viscous Flow Equation in Cylindrical Coordinates

The viscous equation is obtained in cylindrical coordinates according to Eq. (3.60) where \(B = \frac{1}{\mu }\frac{dP}{{dz}}\).

$$\frac{{d^{2} v}}{{dr^{2} }} + { }\frac{1}{r}{ }\frac{dv}{{dr^{ } }} = B$$
(3.60)

This equation is an inhomogeneous ordinary differential equation because of its non-zero right-hand side. To solve this equation, first, the homogeneous part (B = 0) is solved and the general solution (vg) is obtained, then the inhomogeneous part (B ≠ 0) is calculated, and the particular solution (vp) is achieved, and finally, the total answer is resulted as in Eq. (3.61):

$$v_{\left( r \right)} = v_{g} + { }v_{p}$$
(3.61)

To solve the homogeneous part, the right-hand side of Eq. (3.60) is considered equal to zero and the general solution (vg) is calculated. Therefore, Eq. (3.62) is obtained as follows:

$$\frac{{d^{2} v_{g} }}{{dr^{2} }} + { }\frac{1}{r}{ }\frac{{dv_{g} }}{{dr^{ } }} = 0{ }$$
(3.62)

Using the variation of variables in Eq. (3.63), Eq. (3.62) rewritten as Eq. (3.64):

$$u = { }\frac{{dv_{g} }}{dr}{ }.{ }\frac{du}{{dr}} = { }\frac{{d^{2} v_{g} }}{{dr^{2} }}{ }$$
(3.63)
$$\frac{du}{{dr}} + { }\frac{1}{r}{ }u = 0$$
(3.64)

By arranging and integrating from Eq. (3.64), Eq. (3.65) is obtained as follows:

$$u = { }\frac{{C_{1} }}{r}$$
(3.65)

where C1 is a constant. By combining Eqs. (3.63) and (3.65) and reintegration, Eq. (3.66) is achieved:

$$v_{g} = { }C_{1} \ln \left( r \right) + { }C_{2}$$
(3.66)

where C2 is constant. Based on the point that the left-hand side of Eq. (3.60) is a second-order differential equation, the particular solution of equation (vp) is a relationship in the order of two. Therefore, the particular solution of the equation and its derivatives are obtained as Eq. (3.67):

$$v_{p} = ar^{2} + br + c \cdot \frac{{dv_{p} }}{{dr^{ } }} = 2ar + b \cdot \frac{{d^{2} v_{p} }}{{dr^{2} }} = 2a$$
(3.67)

where a, b, and c are constant. Using Eq. (3.67), Eq. (3.60) can be rewritten as Eq. (3.68):

$$\left( {2a} \right) + { }\frac{1}{r}\left( {2ar + b} \right) - B = 2a + 2a + { }\frac{b}{r} - B = 0$$
(3.68)

Assuming b = 0 in Eq. (3.68), the relationship of constants B and a results as Eq. (3.69):

$$a = { }\frac{B}{4}$$
(3.69)

Therefore, assuming c = 0 in Eq. (3.67), the particular solution of the differential equation is obtained as Eq. (3.70):

$$v_{p} = { }\frac{B}{4}r^{2}$$
(3.70)

The overall solution of the differential equation is expressed as Eq. (3.71) using Eqs. (3.61), (3.66) and (3.70):

$$v = { }\frac{B}{4}r^{2} + { }C_{1} \ln \left( r \right) + { }C_{2}$$
(3.71)

Boundary conditions are employed to calculate the constants C1 and C2. Applying the boundary conditions of Eqs. (3.72) and (3.73) in Eq. (3.71), these constants are calculated according to Eq. (3.74):

$$r = \pm R,\quad v = 0$$
(3.72)
$$r = 0,\quad \frac{dv}{{dr}} = 0$$
(3.73)
$$C_{1} = 0{ }.C_{2} = { } - \frac{B}{4}R^{2}$$
(3.74)

Therefore, by substitution of Eq. (3.74) and B value in Eq. (3.71), the viscous flow velocity profile equation in a cylinder can be obtained as Eq. (3.2).

$$v = \frac{1}{4\mu }{ }\frac{dp}{{dz}}\left( {R^{2} - { }r^{2} } \right)$$

To calculate the average value of a function in cylindrical coordinates, Eq. (3.75) is used.

$$\overline{v} = \frac{{\int_{0}^{R} {vrdr} }}{{\int_{0}^{R} {rdr} }}$$
(3.75)

Therefore, by substituting Eq. (3.2) in Eq. (3.75), the value of the average velocity can be calculated according to Eq. (3.76).

$$\overline{v} = \frac{1}{4\mu }{ }\frac{dp}{{dz}}\left( {\frac{{\int_{0}^{R} {r\left( {R^{2} - { }r^{2} } \right)} dr}}{{\int_{0}^{R} {rdr} }}} \right) = \frac{1}{4\mu }{ }\frac{dp}{{dz}}\left( {\frac{{\frac{{{\text{R}}^{4} }}{2} - \frac{{R^{4} }}{4}}}{{\frac{{R^{2} }}{2}}}} \right)$$
(3.76)

By simplifying Eq. (3.76), the viscous flow average velocity equation in a cylinder is obtained as Eq. (3.3).

$$\overline{v} = { }0.125\frac{{R^{2} }}{\mu }\frac{dp}{{dz}}$$

Appendix 3.2: The Brinkman Equation in Cylindrical Coordinates

Brinkman equation in cylindrical coordinates is known as Eq. (3.77):

$$r^{2} \frac{{d^{2} v}}{{dr^{2} }} + r\frac{dv}{{dr}} - { }Ar^{2} v = r^{2} B$$
(3.77)

In Eq. (3.77), parameter B is equal to the constant value of \(\frac{1}{\mu }\frac{dP}{{dz}}\), and parameter A is equal to K−1. This equation is an ordinary inhomogeneous differential equation. To solve this equation, both the general (vg) and the particular (vp) solutions which respectively indicate the homogeneous (r2B = 0) and inhomogeneous part (r2B ≠ 0) are calculated. According to Eq. (3.61), the sum of the particular and general solutions is equal to the total solution of the equation. To calculate vg, first Eq. (3.77) is homogenized as Eq. (3.78):

$$r^{2} \frac{{d^{2} v}}{{dr^{2} }} + r\frac{dv}{{dr}} - { }Ar^{2} v = 0$$
(3.78)

The answer of Eq. (3.78) with its similarity to the modified Basel differential equation (reference) is obtained as Eq. (3.79):

$$v_{g} = C_{1} I_{0} \left( {r\sqrt A } \right) + C_{2} K_{0} \left( {r\sqrt A } \right)$$
(3.79)

where I1 and K1 are the Modified Bessel functions of the first and second kind in the zero-order. According to the point that the left-hand side of Eq. (3.77) is a second-order differential equation, the particular solution of equation (vp) is in the same order. In this case, the particular solution and its derivatives can be assumed as Eq. (3.80):

$$v_{p} = ar^{2} + br + c \cdot \frac{{dv_{p} }}{dr} = 2ar + b \cdot \frac{{d^{2} v_{p} }}{{dr^{2} }} = 2a$$
(3.80)

where a, b, and c are constant. Therefore, Eq. (3.77) can be rewritten as Eq. (3.81) using Eq. (3.80).

$$r^{2} \left( {2a} \right) + r\left( {2ar + b} \right) - { }Ar^{2} \left( {ar^{2} + br + c} \right) = Br^{2}$$
(3.81)

According to Eq. (3.79), it can be seen that for the validity of this equation, the values a and b must be equal to zero (a = b = 0) and the value of c must be equal to –B/A (c = −B/A). Therefore, by placing these constants in Eq. (3.80), the particular solution can be calculated as Eq. (3.82).

$$v_{g} = - \frac{B}{A}$$
(3.82)

Finally, providing particular and general solutions and also considering Eq. (3.61), the velocity function is calculated as Eq. (3.83).

$$v_{\left( r \right)} = C_{1} I_{0} \left( {r\sqrt A } \right) + C_{2} K_{0} \left( {r\sqrt A } \right) - \frac{B}{A}$$
(3.83)

Boundary conditions are employed to calculate the constants C1 and C2 in Eq. (3.81). By applying the boundary conditions of Eq. (3.73) in Eq. (3.83), Eq. (3.84) can be written as follows:

$$C_{1} \sqrt A I_{1} \left( {0*\sqrt A } \right) - C_{2} \sqrt A K_{1} \left( {0*\sqrt A } \right) = 0$$
(3.84)

According to the Bessel function properties, the values I1(0) and K1(0) are equal to zero and infinity, respectively. Therefore, for the validity of Eq. (3.82), the value of C2 must be equal to zero (C2 = 0). Also, by applying the boundary conditions of Eq. (3.72) and the value of C2 in Eq. (3.83), Eq. (3.85) is achieved. By solving Eq. (3.85), the C1 constant results in Eq. (3.86).

$$C_{1} I_{0} \left( {R\sqrt A } \right) - \frac{B}{A} = 0$$
(3.85)
$$C_{1} = \frac{B}{{AI_{0} \left( {R\sqrt A } \right)}}$$
(3.86)

Finally, by placing the constants C1 and C2 as well as the values A and B in Eq. (3.83), the velocity function is calculated as Eq. (3.87).

$$v_{\left( r \right)} = \frac{K}{\mu }\frac{dP}{{dz}}\left( {1 - \frac{{I_{0} \left( {\frac{r}{\sqrt K }} \right)}}{{I_{0} \left( {\frac{R}{\sqrt K }} \right)}}{ }} \right)$$
(3.87)

To calculate the average value of a function in a cylindrical system Eq. (3.75) is used. Therefore, by placing Eq. (3.140) in Eq. (3.75), Eq. (3.88) is obtained as follows:

$$\overline{v} = \frac{K}{\mu }\frac{dP}{{dz}}\left( {\frac{{\int_{0}^{R} {rdr} }}{{\int_{0}^{R} {rdr} }} - \frac{{\int_{0}^{R} {rI_{0} \left( {\frac{r}{\sqrt K }} \right)} }}{{\int_{0}^{R} {rdr} I_{0} \left( {\frac{R}{\sqrt K }} \right)}}} \right) = \frac{K}{\mu }\frac{dP}{{dz}}\left( {1 - \frac{{\left[ {r\sqrt K I_{1} \left( {\frac{r}{\sqrt K }} \right)} \right]_{0}^{R} }}{{\left[ {\frac{{r^{2} }}{2}} \right]_{0}^{R} I_{0} \left( {\frac{R}{\sqrt K }} \right)}}} \right)$$
(3.88)

where I1 is the modified Bessel function of the first kind in the order of one. By solving and simplification of Eq. (3.88), the average velocity function for the Brinkman flow in a cylindrical system is calculated as Eq. (3.89).

$$\overline{v} = { }\frac{dP}{{dz}}\frac{K}{\mu }{ }\left( {1 - \frac{{2\sqrt K I_{1} \left( {\frac{R}{\sqrt K }} \right)}}{{RI_{0} \left( {\frac{R}{\sqrt K }} \right)}}} \right)$$
(3.89)

Appendix 3.3: Linear Diffusivity Equation Solution

This solution is provided by John Lee for radial flow solution in an infinite-acting homogenous reservoir [178]. The basic partial differential equation is given in dimensionless format

$$\frac{{\partial^{2} p_{D} }}{{\partial r_{D}^{2} }} + \frac{1}{{r_{D} }}\frac{{\partial p_{D} }}{{\partial r_{D} }} = \frac{{\partial p_{D} }}{{\partial t_{D} }}$$
(3.90)

where

$$r_{D} = \frac{r}{{r_{w} }}$$
(3.91)
$$p_{D} = p_{DC} \frac{kh}{{qB_{\mu } }}\left( {p_{i} - p_{r} } \right)$$
(3.92)
$$t_{D} = t_{DC} \frac{k}{{Q\mu C_{t} r_{w}^{2} }}t$$
(3.93)

where \(t_{DC}\) and \(p_{DC}\) are given by

The “initial” condition is given as

$$p_{D} \left( {r_{D} \cdot t_{D} = 0} \right) = 0\quad ({\text{uniform}}\,{\text{pressure}}\,{\text{distribution}})$$
(3.94)

The constant rate inner boundary condition is

$$\left[ {r_{D} \frac{{\partial p_{D} }}{{\partial r_{D} }}} \right]_{{r_{D} = 1}} = - 1\quad ({\text{constant}}\,{\text{flow}}\,{\text{rate}}\,{\text{at}}\,{\text{the}}\,{\text{well}})$$
(3.95)

The “infinite-acting” outer boundary condition is given by

$$p_{D} \left( {r_{D} \to \infty \cdot t_{D} } \right) = 0$$
(3.96)

Rewriting Eq. (3.90):

$$\frac{1}{{r_{D} }}\frac{\partial }{{\partial r_{D} }}\left[ {r_{D} \frac{{\partial p_{D} }}{{\partial r_{D} }}} \right] = \frac{{\partial p_{D} }}{{\partial t_{D} }}$$
(3.97)

The Boltzmann transform variable, \(\varepsilon_{D}\), is defined as

$$t_{D} = ar_{D}^{b} t_{D}^{c}$$
(3.98)

where in this problem

a = \(1/4\), b = 2, c = −1.

Which yields

$$\varepsilon_{D} = \frac{{r_{D}^{2} }}{{4t_{D} }}$$
(3.99)

Expanding \(\frac{{\partial^{2} p_{D} }}{{\partial r_{D}^{2} }}\)

$$\frac{\partial }{{\partial r_{D} }}\left[ {\frac{{\partial p_{D} }}{{\partial r_{D} }}} \right] + \frac{1}{{r_{D} }}\frac{{\partial p_{D} }}{{\partial r_{D} }} = \frac{{\partial p_{D} }}{{\partial t_{D} }}$$
(3.100)

Applying the chain rule

$$\frac{{\partial p_{D} }}{\partial x} = \frac{{\partial t_{D} }}{\partial x}\frac{{\partial p_{D} }}{{\partial t_{D} }}$$
(3.101)

Which combined with Eq. (3.100) gives

$$\frac{{\partial \varepsilon_{D} }}{{\partial r_{D} }}\frac{\partial }{{\partial \varepsilon_{D} }}\left[ {\frac{{\partial \varepsilon_{D} }}{{\partial r_{D} }}\frac{{\partial p_{D} }}{{\partial \varepsilon_{D} }}} \right] + \frac{1}{{r_{D} }}\frac{{\partial \varepsilon_{D} }}{{\partial r_{D} }}\frac{{\partial p_{D} }}{{\partial \varepsilon_{D} }} = \frac{{\partial \varepsilon_{D} }}{{\partial t_{D} }}\frac{{\partial p_{D} }}{{\partial \varepsilon_{D} }}$$
(3.102)

Expanding

$$\frac{{\partial \varepsilon_{D} }}{{\partial r_{D} }}\left[ {\frac{\partial }{{\partial \varepsilon_{D} }}\left( {\frac{{\partial \varepsilon_{D} }}{{\partial r_{D} }}} \right)\frac{{\partial p_{D} }}{{\partial \varepsilon_{D} }} + \frac{{\partial \varepsilon_{D} }}{{\partial r_{D} }}\frac{{\partial^{2} p_{D} }}{{\partial \varepsilon_{D}^{2} }}} \right] + \left[ {\frac{1}{{r_{D} }}\frac{{\partial \varepsilon_{D} }}{{\partial r_{D} }} - \frac{{\partial \varepsilon_{D} }}{{\partial t_{D} }}} \right]\frac{{\partial p_{D} }}{{\partial t_{D} }} = 0$$
(3.103)

Isolating terms

$$\left[ {\frac{{\partial \varepsilon_{D} }}{{\partial r_{D} }}} \right]^{2} \frac{{\partial^{2} p_{D} }}{{\partial \varepsilon_{D}^{2} }} + \left[ {\frac{{\partial \varepsilon_{D} }}{{\partial r_{D} }}\frac{\partial }{{\partial \varepsilon_{D} }}\left( {\frac{{\partial \varepsilon_{D} }}{{\partial r_{D} }}} \right) + \frac{1}{{r_{D} }}\frac{{\partial \varepsilon_{D} }}{{\partial r_{D} }} - \frac{{\partial \varepsilon_{D} }}{{\partial t_{D} }}} \right]\frac{{\partial p_{D} }}{{\partial t_{D} }} = 0$$
(3.104)

Dividing through by \(\left( {\partial \varepsilon_{D} /\partial r_{D} } \right)^{2}\) gives

$$\frac{{\partial^{2} p_{D} }}{{\partial \varepsilon_{D}^{2} }} + \frac{1}{{\left( {\partial \varepsilon_{D} /\partial r_{D} } \right)^{2} }}\left[ {\frac{{\partial \varepsilon_{D} }}{{\partial r_{D} }}\frac{\partial }{{\partial \varepsilon_{D} }}\left( {\frac{{\partial \varepsilon_{D} }}{{\partial r_{D} }}} \right) + \frac{1}{{r_{D} }}\frac{{\partial \varepsilon_{D} }}{{\partial r_{D} }} - \frac{{\partial \varepsilon_{D} }}{{\partial t_{D} }}} \right]\frac{{\partial p_{D} }}{{\partial t_{D} }} = 0$$
(3.105)

Reducing the \(\partial ()/\partial \varepsilon_{D}\) term we have

$$\frac{{\partial^{2} p_{D} }}{{\partial \varepsilon_{D}^{2} }} + \frac{1}{{\left( {\partial \varepsilon_{D} /\partial r_{D} } \right)^{2} }}\left[ {\frac{\partial }{{\partial r_{D} }}\left( {\frac{{\partial \varepsilon_{D} }}{{\partial r_{D} }}} \right) + \frac{1}{{r_{D} }}\frac{{\partial \varepsilon_{D} }}{{\partial r_{D} }} - \frac{{\partial \varepsilon_{D} }}{{\partial t_{D} }}} \right]\frac{{\partial p_{D} }}{{\partial t_{D} }} = 0$$
(3.106)

Which can be further reduced to

$$\frac{{\partial^{2} p_{D} }}{{\partial \varepsilon_{D}^{2} }} + \frac{1}{{\left( {\partial \varepsilon_{D} /\partial r_{D} } \right)^{2} }}\left[ {\frac{{\partial^{2} \varepsilon_{D} }}{{\partial r_{D}^{2} }} + \frac{1}{{r_{D} }}\frac{{\partial \varepsilon_{D} }}{{\partial r_{D} }} - \frac{{\partial \varepsilon_{D} }}{{\partial t_{D} }}} \right]\frac{{\partial p_{D} }}{{\partial t_{D} }} = 0$$
(3.107)

Completing the factorization of the \(\left( {\partial \varepsilon_{D} /\partial r_{D} } \right)^{2}\) gives

$$\frac{{\partial^{2} p_{D} }}{{\partial \varepsilon_{D}^{2} }} + \left[ {\frac{1}{{\left( {\partial \varepsilon_{D} /\partial r_{D} } \right)^{2} }}\frac{{\partial^{2} \varepsilon_{D} }}{{\partial r_{D}^{2} }} + \frac{1}{{r_{D} }}\frac{1}{{\partial \varepsilon_{D} /\partial r_{D} }} - \frac{1}{{\left( {\partial \varepsilon_{D} /\partial r_{D} } \right)^{2} }}\frac{{\partial \varepsilon_{D} }}{{\partial t_{D} }}} \right]\frac{{\partial p_{D} }}{{\partial t_{D} }} = 0$$
(3.108)

Using Eq. (3.99) the following derivation is taken

$$\frac{{\partial \varepsilon_{D} }}{{\partial t_{D} }} = \frac{\partial }{{\partial t_{D} }}\left( {\frac{{r_{D}^{2} }}{{4t_{D} }}} \right) = \frac{{r_{D}^{2} }}{4}\frac{\partial }{{\partial t_{D} }}\left( {\frac{1}{{t_{D} }}} \right) = \frac{ - 1}{{t_{D} }}\frac{{r_{D}^{2} }}{{4t_{D} }} = \frac{ - 1}{{t_{D} }}\varepsilon_{D}$$
(3.109)
$$\frac{{\partial \varepsilon_{D} }}{{\partial r_{D} }} = \frac{\partial }{{\partial r_{D} }}\left( {\frac{{r_{D}^{2} }}{{4t_{D} }}} \right) = \frac{2}{{4t_{D} }}r_{D} = \frac{2}{{r_{D} }}\frac{{r_{D}^{2} }}{{4t_{D} }} = \frac{2}{{r_{D} }}t_{D}$$
(3.110)
$$\frac{{\partial^{2} \varepsilon_{D} }}{{\partial r_{D}^{2} }} = \frac{\partial }{{\partial r_{D} }}\left[ {\frac{\partial }{{\partial r_{D} }}\left( {\frac{{r_{D}^{2} }}{{4t_{D} }}} \right)} \right] = \frac{\partial }{{\partial r_{D} }}\left( {\frac{2}{{4t_{D} }}r_{D} } \right) = \frac{2}{{4t_{D} }} = \frac{2}{{r_{D}^{2} }}\frac{{r_{D}^{2} }}{{4t_{D} }} = \frac{2}{{r_{D}^{2} }}t_{D}$$
(3.111)

Substituting Eqs. (3.109)–(3.111) into Eq. (3.108) gives

$$\frac{{\partial^{2} p_{D} }}{{\partial \varepsilon_{D}^{2} }} + \left[ {\frac{1}{{\left( {2\varepsilon_{D} /r_{D} } \right)^{2} }}\frac{2}{{r_{D}^{2} }}\varepsilon_{D} + \frac{1}{{r_{D} }}\frac{1}{{(2\varepsilon_{D} /r_{D} )}} - \frac{1}{{\left( {\partial \varepsilon_{D} /\partial r_{D} } \right)^{2} }}\left( {\frac{ - 1}{{t_{D} }}\varepsilon_{D} } \right)} \right]\frac{{\partial p_{D} }}{{\partial \varepsilon_{D} }} = 0$$
(3.112)

Reducing

$$\frac{{\partial^{2} p_{D} }}{{\partial \varepsilon_{D}^{2} }} + \left[ {\frac{1}{{2\varepsilon_{D} }} + \frac{1}{{2\varepsilon_{D} }} + 1} \right]\frac{{\partial p_{D} }}{{\partial \varepsilon_{D} }} = 0$$
(3.113)

Finally

$$\frac{{\partial^{2} p_{D} }}{{\partial \varepsilon_{D}^{2} }} + \left[ {1 + \frac{1}{{\varepsilon_{D} }}} \right]\frac{{\partial p_{D} }}{{\partial \varepsilon_{D} }} = 0$$
(3.114)

where Eq. (3.114) is the “Boltzmann” transformed differential equation.

Initial and boundary condition in terms of Boltzmann transform are

$$p_{D} \left( {r_{D} \cdot t_{D} = 0} \right) = 0$$
(3.115)

where for \(t_{D} \to 0; \varepsilon_{D} \to \infty\), which gives

$$p_{D} \left( { \varepsilon_{D} \to \infty } \right) = 0$$
(3.116)

The outer boundary condition, Eq. (3.94),

$$p_{D} \left( { r_{D} \to \infty .t_{D} } \right) = 0$$
(3.117)

Or as \(r_{D} \to \infty\); \(\varepsilon_{D} \to \infty\) which yields

$$p_{D} \left( { \varepsilon_{D} \to \infty } \right) = 0$$
(3.117)

where Eqs. (3.116) and (3.118) are the same which illustrates that the Boltzmann transform “collapses” 2 conditions into 1. Combining this observation with the inner boundary condition, we have 2 “boundary” conditions. Coupling this observation with the fact that Eq. (3.114) is only a function of the Boltzmann variable, \(\varepsilon_{D}\), we can solve Eq. (3.100) uniquely. Note that the “collapsing” of the initial and outer boundary conditions must occur or the Boltzmann transform is technically invalid. Recalling the constant rate inner boundary condition, Eq. (3.95)

$$\left[ { r_{D} + \frac{{\partial p_{D} }}{{\partial r_{D} }}} \right]_{{r_{D} = 1}} = - 1\,or\,\left[ { r_{D} + \frac{{\partial p_{D} }}{{\partial r_{D} }}} \right]_{{r_{D} \to 0}} = - 1\quad ({\text{line}}\,{\text{source}}\,{\text{condition}})$$
(3.119)

Or

$$\left[ { r_{D} \frac{{\partial \varepsilon_{D} }}{{\partial r_{D} }}\frac{{\partial p_{D} }}{{\partial \varepsilon_{D} }}} \right]_{{r_{D} \to 0}} = \left[ { r_{D} \left( {\frac{2}{{r_{D} }}\varepsilon_{D} } \right)\frac{{\partial p_{D} }}{{\partial \varepsilon_{D} }}} \right]_{{r_{D} \to 0,\varepsilon_{D} \to 0}} = 2\left[ { \varepsilon_{D} + \frac{{\partial p_{D} }}{{\partial \varepsilon_{D} }}} \right]_{{\varepsilon_{D} \to 0}} = - 1$$
(3.120)

Which can be rearranged to yield

$$\left[ {\varepsilon_{D} \frac{{\partial p_{D} }}{{\partial \varepsilon_{D} }}} \right]_{{\varepsilon_{D} \to 0}} = \frac{ - 1}{2}$$
(3.121)

Making the following variable of substitution

$$v = \frac{{dp_{D} }}{{d\varepsilon_{D} }}$$
(3.122)

Substituting Eq. (3.122) into Eq. (3.124) and noting that use of ordinary derivatives

$$\frac{dv}{{d\varepsilon_{D} }} + \left[ {1 + \frac{1}{{\varepsilon_{D} }}} \right]v = 0$$
(3.123)
$$\frac{1}{v}dv = - \left[ {1 + \frac{1}{{\varepsilon_{D} }}} \right]d\varepsilon_{D} = - d\varepsilon_{D} - \frac{1}{{\varepsilon_{D} }}d\varepsilon_{D}$$
(3.124)

Integrating

$$\ln \left( v \right) = - \varepsilon_{D} - \ln \left( {\varepsilon_{D} } \right) + \beta ;\,\beta = {\text{constant}}\,{\text{of}}\,{\text{integration}}$$
(3.125)

Exponentiation

$$v = {\text{exp}}\left[ { - \varepsilon_{D} - \ln \left( {\varepsilon_{D} } \right) + \beta } \right]$$
(3.126)

Or

$$v = {\text{exp}}\left[ { - \varepsilon_{D} \left] {{\text{exp}}} \right[ - \ln \left( {\varepsilon_{D} } \right)} \right]{\text{exp}}\left[ \beta \right]$$
(3.127)

Which reduces to

$$v = \frac{\alpha }{{\varepsilon_{D} }}{\text{exp}}\left[ { - \varepsilon_{D} } \right]$$
(3.128)

where \(\alpha = {\text{exp}}\left[ \beta \right]\), i.e., the constant of integration. Recalling Eq. (3.120) and combining gives

$$\frac{{dp_{D} }}{{d\varepsilon_{D} }} = \frac{\alpha }{{\varepsilon_{D} }}{\text{exp}}\left[ { - \varepsilon_{D} } \right]$$
(3.129)

Multiplying through by \(\varepsilon_{D}\) gives

$$\varepsilon_{D} \frac{{dp_{D} }}{{d\varepsilon_{D} }} = \alpha {\text{exp}}\left[ { - \varepsilon_{D} } \right]$$
(3.130)

Substitution of Eq. (3.128) into Eq. (3.119) gives

$$\alpha \mathop {\lim }\limits_{{\varepsilon_{D} \to 0}} \left[ {{\text{exp}}\left[ { - \varepsilon_{D} } \right]} \right] = \frac{ - 1}{2}$$
(3.131)

Or

$$\alpha = - 1/2$$
(3.132)

Substitution of Eq. (3.132) into Eq. (3.129) gives

$$\frac{{dp_{D} }}{{d\varepsilon_{D} }} = \frac{ - 1}{{2\varepsilon_{D} }}{\text{exp}}\left[ { - \varepsilon_{D} } \right]$$
(3.133)

Separating and integrating Eq. (3.133) gives

$$\int\limits_{{p_{D} = 0}}^{{p_{D} }} {dp_{D} } = \frac{ - 1}{2}\int\limits_{{\varepsilon_{D} = 0}}^{{\varepsilon_{D} }} {\frac{1}{{\varepsilon_{D} }}e^{{ - \varepsilon_{D} }} } d\varepsilon_{D}$$
(3.134)

where we note that \(p_{D} = 0\) at \(\varepsilon_{D} = \infty\) is the initial outer boundary condition. Completing the integration and reversing the limits we have

$$p_{D} = \frac{1}{2}\int\limits_{{\varepsilon_{D} = \frac{{r_{D}^{2} }}{{4t_{D} }}}}^{\infty } {\frac{1}{y}e^{{ - {\text{y}}}} } dy$$
(3.135)

We note that the integral in Eq. (3.135) is the exponential integral, E1(x), which is given by

$${\text{E}}_{1} \left( {\text{x}} \right) = \int\limits_{X}^{\infty } {\frac{1}{y}e^{{ - {\text{y}}}} } dy$$
(3.136)

Combining Eqs. (3.135) and (3.136) gives our final result

$$p_{D} \left( {r_{D} ,t_{D} } \right) = \frac{1}{2}{\text{E}}_{1} \left( {\frac{{r_{D}^{2} }}{{4t_{D} }}} \right)$$
(3.137)

Appendix 3.4: Solution of Non-linear Diffusivity Equation

The analytical solution of non-linear diffusivity equation needs changing variables in order to perform linear equation to be solved analytically.

$$\frac{{\partial^{2} p}}{{\partial r^{2} }} + \frac{1}{r}\frac{\partial p}{{\partial r}} + c\left( {\frac{\partial p}{{\partial r}}} \right)^{2} = \frac{\emptyset \mu c}{{kt}}\frac{\partial p}{{\partial t}}$$
(3.138)
$$\Delta p = p - p_{i} \cdot T = \frac{kt}{{\emptyset \mu c}}\,and\,\Delta p = \frac{1}{c}\ln p^{*}$$
(3.139)

The linearized form is presented as [179]:

$$\frac{{\partial^{2} p^{*} }}{{\partial r^{2} }} + \frac{1}{r}\frac{{\partial p^{*} }}{\partial r} = \frac{{\partial p^{*} }}{\partial T}$$
(3.140)

The linear equation can be solved using Laplace transform and concludes:

$$\Delta p = \left( {\frac{ - 1}{c}} \right)\left( {{\text{ln}}\left( {1 + \lambda \left( {\ln \left( {\frac{T}{{r_{w}^{2} }}} \right) + 0.2319} \right)} \right) + \frac{0.5772\lambda }{{1 + \lambda \left( {\frac{{{\text{ln}}T}}{{r_{w}^{2} }} + 0.2319} \right)}}} \right)$$
(3.141)
$$\lambda = \frac{q\mu c}{{4\pi kh}}$$
(3.142)

where, h is formation thickness.

Appendix 3.5: The Solution of the Warren-Root Equation in the Different Boundary Condition

3.1.1 Warren-Root Equations

Reservoir pressure distribution is achieved according to the mathematical model. Different reservoir boundary conditions can be applied to this end. The Warren-Root model is used to calculate pressure distribution in fractured reservoirs. Equations (3.28) and (3.29) are used for fluid velocity through matrix and fractures media, respectively:

$${\vec{\mathbf{U}}}_{{\mathbf{m}}} = - \frac{{{\mathbf{K}}_{{\mathbf{m}}} }}{{{\varvec{\upmu}}}}{\mathbf{gra}}{\mathbf{d}}\left( {{\mathbf{P}}_{{\mathbf{m}}} } \right)$$
$${\vec{\mathbf{U}}}_{{\mathbf{f}}} = - \frac{{{\mathbf{K}}_{2} }}{{{\varvec{\upmu}}}}{\mathbf{grad}}\left( {{\mathbf{P}}_{{\mathbf{f}}} } \right)$$

Subscripts m and f denote to matrix and fracture media, respectively. K is absolute permeability in (m2), U is velocity in (ms−1), P is pressure in (Pa), and μ is dynamic viscosity in (Pa s). To calculate matrix and fracture pressure distribution, the mass conservation equation is formulated as follows:

$$\frac{{\partial \left( {\emptyset_{{\text{m}}}\uprho } \right)}}{{\partial {\text{t}}}} + {\text{div}}\left( {\uprho {\overline{\text{U}}}_{{\text{m}}} } \right) + {\text{U}}^{*} = 0$$
(3.143)
$$\frac{{\partial \left( {\emptyset_{{\text{f}}}\uprho } \right)}}{{\partial {\text{t}}}} + {\text{div}}\left( {\uprho {\overline{\text{U}}}_{{\text{f}}} } \right) + {\text{U}}^{*} = 0$$
(3.144)
$${\text{U}}^{*} = \frac{{\uprho \,{\text{SK}}_{{\text{m}}} }}{\upmu }\left( {{\text{P}}_{{\text{m}}} - {\text{P}}_{{\text{f}}} } \right)$$
(3.145)

The rate of mass flow per unit volume is defined by U* in (kg s−1 m−3), which indicates fluid transfer between matrix and fracture in quasi-steady-state condition. is fluid velocity in (ms−1), and φ is porosity, which represents fluid storage capacity in each media. S is a characteristic coefficient of fractured rock proportional to the specific surface of a block, and ρ is fluid density in (kg m−3). For slightly compressible fluids, density is calculated by Eq. (3.146).

$$\uprho =\uprho _{0} \left( {1 + {\text{CP}}} \right)$$
(3.146)

where C is fluid compressibility in (Pa−1), and indicates dependency of fluid volume on the pressure. Equations (3.147) and (3.148) are derived by combining Eqs. (3.143) to (3.146).

$$\emptyset_{{\text{m}}} {\text{C}}_{{\text{m}}} \frac{{\partial {\text{P}}_{{\text{m}}} }}{{\partial {\text{t}}}} + \frac{{{\text{SK}}_{{\text{m}}} }}{\upmu }\left( {{\text{P}}_{{\text{m}}} - {\text{P}}_{{\text{f}}} } \right) = 0$$
(3.147)
$$\emptyset_{{\text{f}}} {\text{C}}_{{\text{f}}} \frac{{\partial {\text{P}}_{{\text{f}}} }}{{\partial {\text{t}}}} - \frac{{{\text{K}}_{{\text{f}}} }}{\upmu }\left( {\frac{1}{{\text{r}}}\frac{\partial }{{\partial {\text{r}}}}\left( {{\text{r}}\frac{{\partial {\text{P}}_{{\text{f}}} }}{{\partial {\text{r}}}}} \right)} \right) + \frac{{{\text{SK}}_{{\text{m}}} }}{\upmu }\left( {{\text{P}}_{{\text{m}}} - {\text{P}}_{{\text{f}}} } \right) = 0$$
(3.148)

Dimensionless variables in Eqs. (3.149)–(3.153) are used to reduce the number of variables.

$${\text{P}}_{{\text{D}}} = \frac{{2\uppi {\text{K}}_{{\text{f}}} {\text{h}}\left( {{\text{P}}_{{\text{i}}} - {\text{P}}\left( {{\text{r}},{\text{t}}} \right)} \right)}}{{{\text{q}}\upmu }}$$
(3.149)
$${\text{r}}_{{\text{D}}} = \frac{{\text{r}}}{{{\text{r}}_{{\text{w}}} }}$$
(3.150)
$${\text{t}}_{{\text{D}}} = \frac{{{\text{K}}_{{\text{f}}} {\text{t}}}}{{\left( {{\text{C}}_{{\text{m}}} \emptyset_{{\text{m}}} + {\text{C}}_{{\text{f}}} \emptyset_{{\text{f}}} } \right)\upmu {\text{r}}_{{\text{w}}}^{2} }}$$
(3.151)
$$\uplambda = \frac{{\upalpha {\text{K}}_{{\text{m}}} {\text{r}}_{{\text{w}}}^{2} }}{{{\text{K}}_{{\text{f}}} }}$$
(3.152)
$$\upomega = \frac{{\emptyset_{{\text{f}}} {\text{C}}_{{\text{f}}} }}{{\emptyset_{{\text{m}}} {\text{C}}_{{\text{m}}} + \emptyset_{{\text{f}}} {\text{C}}_{{\text{f}}} }}$$
(3.153)

Dimensionless pressure is defined by PD, rD is dimensionless radius, and tD is dimensionless time. Equations (3.152) and (3.153) represent fracture reservoir parameters, which indicate interporosity (λ) and fracture storage capacity (ω).

At the initial time, there is a pressure equilibrium state in the reservoir. Overall reservoir pressure is equal to reservoir initial pressure in this condition. Therefore, at the initial times, Eq. (3.154) is used as follows:

$${\text{P}} = {\text{P}}_{{\text{i}}} \quad {\text{t}} = 0$$
(3.154)

Dimensionless pressure at initial time is resulted by a combination of Eqs. (3.149) and (3.154).

$${\text{P}}_{{{\text{Dm}}}} = 0\quad {\text{t}}_{{\text{D}}} = 0$$
(3.155)

Equations (3.147) and (3.148) are rewritten in dimensionless forms by combining Eqs. (3.149)–(3.153) as follows:

$$\left( {1 -\upomega } \right)\frac{{\partial {\text{P}}_{{{\text{Dm}}}} }}{{\partial {\text{t}}_{{\text{D}}} }} -\uplambda \left( {{\text{P}}_{{{\text{Df}}}} - {\text{P}}_{{{\text{Dm}}}} } \right) = 0$$
(3.156)
$$-\upomega \frac{{\partial {\text{P}}_{{{\text{Df}}}} }}{{\partial {\text{t}}_{{\text{D}}} }} + \frac{1}{{{\text{r}}_{{\text{D}}} }}\frac{\partial }{{\partial {\text{r}}_{{\text{D}}} }}\left( {{\text{r}}_{{\text{D}}} \frac{{\partial {\text{P}}_{{{\text{Df}}}} }}{{\partial {\text{r}}_{{\text{D}}} }}} \right) + \left( {1 -\upomega } \right)\frac{{\partial {\text{P}}_{{{\text{Dm}}}} }}{{\partial {\text{t}}_{{\text{D}}} }} = 0$$
(3.157)

Equation (3.157) is rewritten as Eq. (3.158):

$$\frac{{\partial {\text{P}}_{{{\text{Df}}}}^{2} }}{{\partial^{2} {\text{r}}_{{\text{D}}} }} + { }\frac{1}{{{\text{r}}_{{\text{D}}} }}\frac{{\partial {\text{P}}_{{{\text{Df}}}} }}{{\partial {\text{r}}_{{\text{D}}} }} = \left( {1 -\upomega } \right)\frac{{\partial {\text{P}}_{{{\text{Dm}}}} }}{{\partial {\text{t}}_{{\text{D}}} }} +\upomega \frac{{\partial {\text{P}}_{{{\text{Df}}}} }}{{\partial {\text{t}}_{{\text{D}}} }}$$
(3.158)

There are several methods for solving partial differential equations. Laplace transform is one of these methods. Laplace transform is applied to solve differential Eqs. (3.156) and (3.158). Consequently Eq. (3.159) is resulted by Eq. (3.156) Laplace transform:

$$\left( {1 -\upomega } \right){\text{zP}}_{{{\text{Dm}}}} \left( {{\text{z}},{\text{r}}_{{\text{D}}} } \right) - {\text{P}}_{{{\text{Dm}}}} \left( {{\text{t}} = 0} \right) =\uplambda \left( {{\text{P}}_{{{\text{Df}}}} \left( {{\text{z}},{\text{r}}_{{\text{D}}} } \right) - {\text{P}}_{{{\text{Dm}}}} \left( {{\text{z}},{\text{r}}_{{\text{D}}} } \right)} \right)$$
(3.159)

The Laplace transform variable is known as z. By substituting Eq. (3.159) in the initial condition (Eq. (3.155)), matrix dimensionless pressure is resulted as the following equation:

$${\text{P}}_{{{\text{D}}m}} \left( {\text{z}} \right) = \frac{{\uplambda {\text{P}}_{{{\text{Df}}}} \left( {{\text{z}} \cdot {\text{r}}_{{\text{D}}} } \right)}}{{\left( {1 -\upomega } \right){\text{z}} +\uplambda }}$$
(3.160)

To obtain the fracture pressure equation, the Laplace transform of Eq. (3.158) is rewritten as Eq. (3.161).

$$\frac{{\partial^{2} {\text{P}}_{{{\text{Df}}}} \left( {{\text{z}} \cdot {\text{r}}_{{\text{D}}} } \right)}}{{\partial {\text{r}}_{{\text{D}}}^{2} }} + \frac{1}{{{\text{r}}_{{\text{D}}} }}\frac{{\partial {\text{P}}_{{{\text{Df}}}} \left( {{\text{z}}.{\text{r}}_{{\text{D}}} } \right)}}{{\partial {\text{r}}_{{\text{D}}} }} = { }\left( {1 -\upomega } \right){\text{zP}}_{{{\text{Dm}}}} \left( {{\text{z}}.{\text{r}}_{{\text{D}}} } \right) +\upomega {\text{zP}}_{{{\text{Df}}}} \left( {{\text{z}}.{\text{r}}_{{\text{D}}} } \right)$$
(3.161)

Equation (3.160) is replaced in Eq. (3.161) and Eq. (3.162) is resulted as follows:

$$\frac{{{\text{d}}^{2} {\text{P}}_{{{\text{Df}}}} \left( {{\text{z}}.{\text{r}}_{{\text{D}}} } \right)}}{{{\text{dr}}_{{\text{D}}}^{2} }} + \frac{1}{{{\text{r}}_{{\text{D}}} }}\frac{{{\text{dP}}_{{{\text{Df}}}} \left( {{\text{z}}.{\text{r}}_{{\text{D}}} } \right)}}{{{\text{dr}}_{{\text{D}}} }} = [{\text{zf}}\left( {\text{z}} \right)]{\text{ P}}_{{{\text{Df}}}} \left( {{\text{z}},{\text{r}}_{{\text{D}}} } \right)$$
(3.162)

In Eq. (3.162), f(z) is a function of ω and λ and is described as Eq. (3.31):

$${\mathbf{f}}\left( {\mathbf{z}} \right) = \frac{{{{\varvec{\upomega}}}\left( {1 - {{\varvec{\upomega}}}} \right){\mathbf{z}} + {{\varvec{\uplambda}}}}}{{\left( {1 - {{\varvec{\upomega}}}} \right){\mathbf{z}} + {{\varvec{\uplambda}}}}}$$

The solution of the partial differential Eq. (3.162) is necessary to calculate fracture dimensionless pressure. Therefore, rD2 is multiplied by Eq. (3.162) and Eq. (3.163) is obtained as a standard form of Bessel equation.

$${\text{r}}_{{\text{D}}}^{2} \frac{{{\text{d}}^{2} {\text{P}}_{{{\text{Df}}}} \left( {z.{\text{r}}_{{\text{D}}} } \right)}}{{{\text{dr}}_{{\text{D}}}^{2} }} + {\text{r}}_{{\text{D}}} \frac{{{\text{dP}}_{{{\text{Df}}}} \left( {{\text{z}}.{\text{r}}_{{\text{D}}} } \right)}}{{{\text{dr}}_{{\text{D}}} }} - \left( {{\text{r}}_{{\text{D}}}^{2} {\text{zf}}\left( {\text{z}} \right)} \right){\text{P}}_{{{\text{D}}2}} \left( {{\text{z}}.{\text{r}}_{{\text{D}}} } \right) = 0$$
(3.163)

Therefore, the solution of Eq. (3.163) is written as Eq. (3.30) by I0 and K0 parameters, which are the first and second type of modified Bessel functions in zero-order, respectively.

$${\mathbf{P}}_{{{\mathbf{Df}}}} \left( {{\mathbf{z}}.{\mathbf{r}}_{{\mathbf{D}}} } \right) = {\mathbf{C}}_{1} {\mathbf{I}}_{0} \left( {{\mathbf{M}} {\mathbf{r}}_{{\mathbf{D}}} } \right) + {\mathbf{C}}_{2} {\mathbf{K}}_{0} \left( {{\mathbf{M}} {\mathbf{r}}_{{\mathbf{D}}} } \right)$$

Bessel equation constants are introduced by C1 and C2, and boundary conditions are applied to obtain these values. To simplify calculations, \(M = \sqrt {{\text{zf}}\left( {\text{z}} \right)}\) is considered. Constants of Eq. (3.30) are calculated by different boundary conditions substitution, and pressure distribution equations are obtained as a result.

3.1.2 Constant Production Rate in Closed Outer Boundary

Well production rate is kept constant in constant production rate conditions. A choke valve is usually used to stabilize the production rate in operating conditions. This case is more common in hydrocarbon reservoirs production. The wellbore boundary condition is written as Eq. (3.164) in this case:

$$\left( {\frac{{\partial {\text{P}}}}{{\partial {\text{r}}}}} \right)_{{{\text{r}}_{{\text{w}}} }} = {\text{constant}}\quad {\text{r }} = {\text{r}}_{{\text{w}}}$$
(3.164)

Some hydrocarbon reservoirs are confined with faults or impermeable layers. No fluid flows through the outer boundary in these closed reservoirs. Also, for a reservoir with several production wells, the closed boundary can be assumed among the wells. The closed boundary conditions are expressed in Eq. (3.165):

$$\left( {\frac{{\partial {\text{P}}_{ } }}{{\partial {\text{r}}}}} \right)_{{{\text{r}}_{{\text{e}}} }} = 0\quad {\text{r}} = {\text{r}}_{{\text{e}}}$$
(3.165)

Darcy equation is applied to make a relationship between flow and pressure for an inner boundary condition (the well condition) as Eq. (3.166). In this case, C1 and C2 are calculated by replacing the boundary conditions (Eqs. (3.164) and (3.165)) in Eq. (3.30).

$${\text{q}} = - 2\uppi {\text{r}}_{{\text{w}}} {\text{h }}\frac{{{\text{K}}_{{\text{f}}} }}{\upmu }{ }\frac{{\partial {\text{P}}_{{\text{f}}} }}{{\partial {\text{r}}}}\quad {\text{r}} = {\text{r}}_{{\text{w}}}$$
(3.166)

Equation (3.165) is written for closed boundary condition and dimensionless form of Eqs. (3.165) and (3.166) is resulted as Eqs. (3.168) and (3.167), respectively.

$$\frac{{\partial {\text{P}}_{{{\text{Df}}}} }}{{\partial {\text{r}}_{{\text{D}}} }} = - 1\quad {\text{r}}_{{\text{D}}} = 1$$
(3.167)
$$\frac{{\partial {\text{P}}_{{{\text{Df}}}} }}{{\partial r_{D} }} = 0\quad {\text{r}}_{{\text{D}}} = {\text{r}}_{{{\text{De}}}}$$
(3.168)

By differentiating Eq. (3.30) and integrating with Eq. (3.167), results can be represented as the following equations:

$${\text{C}}_{1} {\text{M}}\,{\text{I}}_{1} \left( {{\text{M}}\,{\text{r}}_{{\text{D}}} } \right) - {\text{C}}_{2} {\text{M}}\,{\text{K}}_{1} \left( {{\text{M}}\,{\text{r}}_{{\text{D}}} } \right) = \frac{{\partial {\text{P}}_{{{\text{Df}}}} }}{{\partial {\text{r}}_{{\text{D}}} }}$$
(3.169)
$${\text{C}}_{1} {\text{MI}}_{1} \left( {\text{M }} \right) - {\text{C}}_{2} {\text{M}}\,{\text{K}}_{1} \left( {\text{M }} \right) = \frac{ - 1}{z}$$
(3.170)

The first and second types of modified Bessel functions in the first order are known as I1 and K1. Equation (3.171) is resulted by differentiating Eq. (3.30) and integrating with Eq. (3.168).

$${\text{C}}_{1} {\text{M}}\,{\text{I}}_{1} \left( {{\text{M}}\,{\text{r}}_{{{\text{De}}}} { }} \right) - {\text{C}}_{2} {\text{M}}\,{\text{K}}_{1} \left( {{\text{M}}\,{\text{r}}_{{{\text{De}}}} { }} \right) = 0$$
(3.171)

A system of two linear Eqs. (3.170) and (3.171) is formed to calculate C1 and C2:

$${\text{C}}_{1} = \frac{{ - {\text{K}}_{1} \left( {{\text{M}}\,{\text{r}}_{{{\text{De}}}} { }} \right)}}{{{\text{z}}\left[ {\left( {{\text{K}}_{1} \left( {{\text{M}}\,{\text{r}}_{{{\text{De}}}} } \right){ }} \right) \times {\text{M}}\,{\text{I}}_{1} \left( {\text{M }} \right) - {\text{I}}_{1} \left( {{\text{M}}\,{\text{r}}_{{{\text{De}}}} { }} \right) \times {\text{M}}\,{\text{K}}_{1} \left( {\text{M }} \right)} \right]}}$$
(3.172)
$${\text{C}}_{2} = \frac{{ - {\text{I}}_{1} \left( {{\text{Mr}}_{{{\text{De}}}} { }} \right)}}{{{\text{ z}}\left[ {{\text{K}}_{1} \left( {{\text{ M}}\,{\text{r}}_{{{\text{De}}}} } \right) \times {\text{M}}\,{\text{I}}_{1} \left( {\text{M }} \right) - {\text{I}}_{1} \left( {{\text{M}}\,{\text{r}}_{{{\text{De}}}} { }} \right) \times {\text{M}}\,{\text{K}}_{1} \left( {\text{M }} \right)} \right]}}$$
(3.173)

Equation (3.32) is resulted by substitution C1 and C2 in Eq. (3.30) as follows:

$${\mathbf{P}}_{{{\mathbf{Df}}}} \left( {{\mathbf{z}}.{\mathbf{r}}_{{\mathbf{D}}} } \right) = \frac{{{\mathbf{K}}_{1} \left( {{\mathbf{M}}\,{\mathbf{r}}_{{{\mathbf{De}}}} } \right) {\mathbf{I}}_{0} \left( {{\mathbf{M}}\,{\mathbf{r}}_{{\mathbf{D}}} } \right) + {\mathbf{I}}_{1} \left( {{\mathbf{M}}\,{\mathbf{r}}_{{{\mathbf{De}}}} } \right){\mathbf{K}}_{0} \left( {{\mathbf{M}}\,{\mathbf{r}}_{{\mathbf{D}}} } \right) }}{{{\mathbf{z}} {\mathbf{M}} \left[ {{\mathbf{K}}_{1} \left( {\mathbf{M}} \right) {\mathbf{I}}_{1} \left( {{\mathbf{M}}\,{\mathbf{r}}_{{{\mathbf{De}}}} } \right) - {\mathbf{K}}_{1} \left( {{\mathbf{M}}\,{\mathbf{r}}_{{{\mathbf{De}}}} } \right) {\mathbf{I}}_{1} \left( {\mathbf{M}} \right)} \right]}}$$

Therefore, the dimensionless pressure equation in a closed fractured reservoir and constant production rate is given by Eq. (3.32).

3.1.3 Constant Production Rate in Constant Pressure Outer Boundary

For infinite reservoirs with high-pressure supplier, there is no pressure drop at the reservoir outer boundary, and the outer boundary pressure equals to reservoir initial pressure, as shown by Eq. (3.174). The dimensionless form of this equation is presented as Eq. (3.175).

$${\text{P}}_{{\text{e}}} = {\text{P}}_{{\text{i}}} \quad {\text{r }} = {\text{r}}_{{\text{e}}}$$
(3.174)
$${\text{P}}_{{{\text{Dm}}}} = {\text{P}}_{{{\text{Df}}}} = 0\quad {\text{r}}_{{\text{D}}} = {\text{r}}_{{{\text{De}}}}$$
(3.175)

Equation (3.175) is replaced in Eq. (3.30) and Eq. (3.176) is resulted as follows:

$${\text{C}}_{1} {\text{I}}_{0} \left( {{\text{M}}\,{\text{r}}_{{{\text{De}}}} { }} \right) + {\text{C}}_{2} {\text{K}}_{0} \left( {{\text{M}}\,{\text{r}}_{{{\text{De}}}} { }} \right) = 0$$
(3.176)

A system of two linear Eqs. (3.170) and (3.176) is formed to calculate C1 and C2 in this case:

$${\text{C}}_{1} = - \frac{{{\text{K}}_{0} \left( {{\text{M}}\,{\text{r}}_{{{\text{De}}}} { }} \right)}}{{{\text{z M}}\left[ {{\text{K}}_{0} \left( {{\text{M}}\,{\text{r}}_{{{\text{De}}}} } \right){\text{I}}_{1} \left( {\text{M }} \right) + {\text{I}}_{0} \left( {{\text{M}}\,{\text{r}}_{{{\text{De}}}} { }} \right){\text{K}}_{1} \left( {\text{M }} \right)} \right]}}$$
(3.177)
$${\text{C}}_{2} = \frac{{{\text{I}}_{0} \left( {{\text{M}}\,{\text{r}}_{{{\text{De}}}} { }} \right)}}{{{\text{ z M}}\left[ {{\text{K}}_{0} \left( {{\text{M}}\,{\text{r}}_{{{\text{De}}}} } \right){\text{I}}_{1} \left( {\text{M }} \right) + {\text{I}}_{0} \left( {{\text{M}}\,{\text{r}}_{{{\text{De}}}} } \right){\text{K}}_{1} \left( {\text{M}} \right)} \right]}}$$
(3.178)

Equation (3.33) is resulted by replacing C1 and C2 in Eq. (3.30).

$${\mathbf{P}}_{{{\mathbf{Df}}}} \left( {{\mathbf{z}}.{\mathbf{r}}_{{\mathbf{D}}} } \right) = \frac{{{\mathbf{K}}_{0} \left( {{\mathbf{M}}\,{\mathbf{r}}_{{\mathbf{D}}} } \right){\mathbf{I}}_{0} \left( {{\mathbf{M}}\,{\mathbf{r}}_{{{\mathbf{De}}}} } \right) - {\mathbf{K}}_{0} \left( {{\mathbf{M}}\,{\mathbf{r}}_{{{\mathbf{De}}}} } \right){\mathbf{I}}_{0} \left( {{\mathbf{M}}\,{\mathbf{r}}_{{\mathbf{D}}} } \right)}}{{ {\mathbf{z}} {\mathbf{M}}\left[ {{\mathbf{K}}_{0} \left( {{\mathbf{M}}\,{\mathbf{r}}_{{{\mathbf{De}}}} } \right){\mathbf{I}}_{1} \left( {{\mathbf{M}} } \right) + {\mathbf{K}}_{1} \left( {{\mathbf{M}} } \right){\mathbf{I}}_{0} \left( {{\mathbf{M}}\,{\mathbf{r}}_{{{\mathbf{De}}}} } \right)} \right]}}$$

Equation (3.33) is applicable for dimensionless flow rate calculation at constant pressure boundary in case of constant pressure production condition.

3.1.4 Constant Pressure Production in Closed Outer Boundary

The wellbore pressure is kept constant, and flow rate changes i.e., it decreases with time in constant pressure production condition. In operating conditions, well pressure is maintained constant by using some gauges that alter the production rate. This production condition is mostly used in reservoirs with a high risk of gas or water coning. The wellbore boundary condition, in this case, is written as Eq. (3.179):

$${\text{P}}_{ } = {\text{constant}}\quad {\text{r }} = {\text{r}}_{{\text{w}}}$$
(3.179)

Production rate equation is obtained using the wellbore pressure equation suggested by Van Everdingen and Hurst [180]:

$${\text{q}}_{{\text{D}}} \left( {{\text{z}}.{\text{r}}_{{\text{D}}} } \right) = \frac{1}{{{\text{z}}^{2} {\text{ P}}_{{{\text{Dwf}}}} \left( {{\text{z}},{\text{r}}_{{\text{D}}} } \right)}}$$
(3.180)

The dimensionless production equation (Eq. (3.34)) is obtained by combining Eqs. (3.32) and (3.180).

$${\mathbf{q}}_{{\mathbf{D}}} \left( {{\mathbf{z}}.{\mathbf{r}}_{{\mathbf{D}}} } \right) = \frac{{ {\mathbf{M}}\left[ { {\mathbf{K}}_{1} \left( {\mathbf{M}} \right) {\mathbf{I}}_{1} \left( {{\mathbf{M}}\,{\mathbf{r}}_{{{\mathbf{De}}}} } \right) - {\mathbf{K}}_{1} \left( {{\mathbf{M}}\,{\mathbf{r}}_{{{\mathbf{De}}}} } \right) {\mathbf{I}}_{1} \left( {\mathbf{M}} \right)} \right] }}{{{\mathbf{z}}[{\mathbf{K}}_{1} \left( {{\mathbf{M}}\,{\mathbf{r}}_{{{\mathbf{De}}}} } \right) {\mathbf{I}}_{0} \left( {{\mathbf{M}}\,{\mathbf{r}}_{{\mathbf{D}}} } \right) + {\mathbf{I}}_{1} \left( {{\mathbf{M}}\,{\mathbf{r}}_{{{\mathbf{De}}}} } \right){\mathbf{K}}_{0} \left( {{\mathbf{M}}\,{\mathbf{r}}_{{\mathbf{D}}} } \right)]}}$$

The dimensionless flow rate at a closed boundary fractured reservoir with constant pressure production is calculated by Eq. (3.34).

3.1.5 Constant Pressure Production in Constant Pressure Outer Boundary

In this case, the inner and outer boundary conditions are described by Eqs. (3.179) and (3.174). The Van Everdingen and Hurst relations are applied as the same as the previous section. Equation (3.35) is concluded by combining Eqs. (3.33) and (3.180):

$${\mathbf{q}}_{{\mathbf{D}}} \left( {{\mathbf{z}}.{\mathbf{r}}_{{\mathbf{D}}} } \right) = \frac{{ {\mathbf{M}}\left[ {{\mathbf{K}}_{0} \left( {{\mathbf{M}}\,{\mathbf{r}}_{{{\mathbf{De}}}} } \right){\mathbf{I}}_{1} \left( {{\mathbf{M}} } \right) + {\mathbf{K}}_{1} \left( {{\mathbf{M}} } \right){\mathbf{I}}_{0} \left( {{\mathbf{M}}\,{\mathbf{r}}_{{{\mathbf{De}}}} } \right)} \right]}}{{{\mathbf{z}} \left[ {{\mathbf{K}}_{0} \left( {{\mathbf{M}}\,{\mathbf{r}}_{{\mathbf{D}}} } \right){\mathbf{I}}_{0} \left( {{\mathbf{M}}\,{\mathbf{r}}_{{{\mathbf{De}}}} } \right) - {\mathbf{K}}_{0} \left( {{\mathbf{M}}\,{\mathbf{r}}_{{{\mathbf{De}}}} } \right){\mathbf{I}}_{0} \left( {{\mathbf{M}}\,{\mathbf{r}}_{{\mathbf{D}}} } \right)} \right]}}$$

Equation (3.35) is applicable for dimensionless flow rate calculations at constant pressure reservoir boundary in the case of constant pressure production.

3.1.6 Warren-Root Analytical Solution

The original solution provided by Warren and Root considers infinite outer boundary conditions and also the constant production rate for an inner boundary condition. The inner boundary condition is presented as Eq. (3.164) and the outer boundary condition is defined as Eq. (3.181):

$${\text{P}} = {\text{P}}_{{\text{i}}} \quad {\text{r }} = \infty$$
(3.181)

This assumption is applied to constant production rate conditions [65]. This simplification is applied to Eq. (3.32). The infinite values of the first and second-order Bessel function tend to infinite and zero, respectively. Thus, Eq. (3.36) is a simplified form of Eq. (3.32) which is resulted by using the Warren-Root assumptions:

$${\mathbf{P}}_{{\mathbf{D}}} \left( {{\mathbf{z}}.{\mathbf{r}}_{{\mathbf{D}}} } \right) = \frac{{{\mathbf{K}}_{0} \left( {{\mathbf{M}}\,{\mathbf{r}}_{{\mathbf{D}}} } \right) }}{{{\mathbf{z}} {\mathbf{M}}\,{\mathbf{K}}_{1} \left( {\mathbf{M}} \right) }}$$

The Laplace inverse calculation of Eq. (3.36) is not possible by conventional analytical methods. An approximate method for calculating K0 and K1 is possible by considering primary terms of Bessel functions:

$${\mathbf{K}}_{0} \left( {\mathbf{M}} \right) = - {{\varvec{\upgamma}}} - {\mathbf{Ln}}\left( {\frac{{{\mathbf{r}}_{{\mathbf{D}}} }}{2}{\mathbf{M}}^{2} } \right)$$
(3.182)
$${\mathbf{K}}_{1} \left( {\mathbf{M}} \right) = \frac{1}{{{\mathbf{M}}^{2} }}$$
(3.183)

In Eq. (3.182), ɤ is Euler number and equals 0.5772. This approximation is applicable only for M < 0.01 values. Also, rD = 1 is assumed in the above equations. Equation (3.184) is obtained by substituting Eqs. (3.182) and (3.183) in Eq. (3.36) and dimensionless pressure is calculated by numerical inverse Laplace method, as follows:

$${\text{P}}_{{\text{D}}} \left( {{\text{t}}_{{\text{D}}} .1} \right) = \frac{1}{2}[0.80908 + {\text{Ln }}\left( {{\text{t}}_{{\text{D}}} } \right) + {\text{Ei}}\left( {\frac{{ -\uplambda {\text{t}}_{{\text{D}}} }}{{\upomega \left( {1 -\upomega } \right)}}} \right) - {\text{ Ei}}\left( {\frac{{ -\uplambda {\text{t}}_{{\text{D}}} }}{{\left( {1 -\upomega } \right)}}} \right)$$
(3.184)

Exponential integral, which is called Ei function, is defined as Eq. (3.185):

$${\text{Ei}}\left( { - {\text{X}}} \right) = - \int\limits_{{\text{X}}}^{\infty } {\frac{{{\text{e}}^{{ - {\text{u}}}} }}{{\text{u}}}} {\text{du}}$$
(3.185)

The Ei function values limit to zero at a very long time. Therefore, Eq. (3.184) is converted to Eq. (3.37):

$${\mathbf{P}}_{{\mathbf{D}}} \left( {{\mathbf{t}}_{{\mathbf{D}}} .1} \right) = \frac{1}{2}\left[ {0.80908 + {\mathbf{Ln}} \left( {{\mathbf{t}}_{{\mathbf{D}}} } \right)} \right] = \frac{1}{2}{\mathbf{Ln}} \left( {2.25{\mathbf{t}}_{{\mathbf{D}}} } \right) = 1.15\log \left( {2.25{\mathbf{t}}_{{\mathbf{D}}} } \right)$$

Therefore, Warren and Root solved the inverse Laplace of dimensionless pressure equation in a fractured reservoir with simple assumptions in specific conditions (M < 0.01). The solution provided by Warren and Root is a simplified form of the general solution shown in Eq. (3.32) for M < 0.01 at wellbore (rD = 1) [63].

Rights and permissions

Reprints and permissions

Copyright information

© 2022 The Author(s), under exclusive license to Springer Nature Switzerland AG

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Azin, R., Izadpanahi, A., Zahedizadeh, P. (2022). Basics of Oil and Gas Flow in Reservoirs. In: Azin, R., Izadpanahi, A. (eds) Fundamentals and Practical Aspects of Gas Injection. Petroleum Engineering. Springer, Cham. https://doi.org/10.1007/978-3-030-77200-0_3

Download citation

  • DOI: https://doi.org/10.1007/978-3-030-77200-0_3

  • Published:

  • Publisher Name: Springer, Cham

  • Print ISBN: 978-3-030-77199-7

  • Online ISBN: 978-3-030-77200-0

  • eBook Packages: EnergyEnergy (R0)

Publish with us

Policies and ethics