Skip to main content

Part of the book series: SpringerBriefs in Earth Sciences ((BRIEFSEARTH))

  • 1522 Accesses

Abstract

This chapter develops representations in state-space notation of the porous-media flow equations derived in Chap. 1. For single-phase flow, the states are grid-block pressures, and for two-phase flow they are grid-block pressures and saturations. The inputs are typically bottom-hole pressures or total well flow rates, the outputs are typically bottom-hole pressures in those wells were the flow rates were prescribed, and phase rates in those wells were the bottom-hole pressures were prescribed. The use of matrix partitioning to describe the different types of inputs leads to a description in terms of nonlinear ordinary-differential and algebraic equations with (state-dependent) system, input, output and direct-throughput matrices. Other topics include generalized state-space representations, linearization, elimination of prescribed pressures, the tracing of stream lines, lift tables, computational aspects, and the derivation of an energy balance for porous-media flow.

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 39.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 54.99
Price excludes VAT (USA)
  • Compact, lightweight 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

Notes

  1. 1.

    The dependent variables follow from the physics of the problem. In case of multi-phase flow through porous media they are typically pressures, component masses or phase saturations; see Chap. 1. Here we use only a single dependent variable, but in general multiple dependent variables will occur, in which case multiple differential equations are required to describe the problem.

  2. 2.

    Typically, most of the values \( \bullet_{i} \) are equal to zero in a single equation. For example in the case of one-dimensional single-phase flow modeled with first-order finite differences, the only three non-zero values \( \bullet_{i} \) in the ith equation in a system of n equations with 1 < i < n are given by: \( \hat{e}_{i} \left( { \bullet_{i} ,{{d\left( { \bullet_{i} } \right)} \mathord{\left/ {\vphantom {{d\left( { \bullet_{i} } \right)} {dt}}} \right. \kern-0pt} {dt}}} \right) = \hat{f}_{i} \left( { \bullet_{i - 1} , \bullet_{i} , \bullet_{i + 1} ,\psi_{i} } \right). \)

  3. 3.

    An alternative name for the input matrix is distribution matrix because it distributes the inputs u over the states x.

  4. 4.

    Sometimes this form of generalized state equations is referred to as a descriptor system .

  5. 5.

    System equations expressed as \( {\mathbf{g}}\left( \ldots \right) = \bf{0} \) are sometimes referred to as equations in residual  form.

  6. 6.

    The random model errors are also referred to as random input, or as a stochastic forcing term.

  7. 7.

    Usually simply referred to as Jacobians.

  8. 8.

    In the systems-and-control literature this is known as a control-affine nonlinear equation. An affine function is a linear function plus a translation. Control affine functions are an important topic of study in nonlinear control theory.

  9. 9.

    In a numerical implementation, the inverse V −1 of the diagonal  matrix V can be computed very efficiently by just taking the reciprocals of the diagonal elements.

  10. 10.

    In grid blocks that are not penetrated by a well the prescribed flow rates are of course equal to zero.

  11. 11.

    I.e. we do not consider kinetic energy at a macroscopic level. We do take into account energy dissipation, which is the change of mechanical energy into thermal energy, or heat, and which at an atomic level can be interpreted as kinetic energy again.

  12. 12.

    Elevation-related potential energy plays an important role in well-bore flow. Most reservoirs have enough potential energy, at least initially, to lift the oil to surface naturally in the production wells . This lift effect is in most cases caused by the difference in density between oil and water, such that in an oil-filled well that drains a hydrostatically-pressured reservoir the oil will be lifted to surface because of elevation-related potential energy.

  13. 13.

    For a detailed derivation of pressure-related potential energy for the case of pressure-dependent rock and fluid compressibilities, see Hubbert (1940).

  14. 14.

    Energy delivered to a system through mechanical or hydraulic action is often referred to as work . Work (or energy) per unit time is then called power . In strict SI units time is expressed in s (seconds), energy and work in J (Joule) and power in W (Watt), such that 1 W is equal to 1 J/s.

  15. 15.

    In comparison, Eq. (2.127) was also a function of the well-bore pressures \( {\mathbf{\overset{\lower0.5em\hbox{$\smash{\scriptscriptstyle\smile}$}}{p} }}_{well} \) and \( {\bar{\mathbf{p}}}_{well} \).

  16. 16.

    The general expression for the inverse of a 2 × 2 block matrix is given by \( \left[ {\begin{array}{*{20}c} {{\mathbf{V}}_{11} } & {{\mathbf{V}}_{12} } \\ {{\mathbf{V}}_{21} } & {{\mathbf{V}}_{22} } \\ \end{array} } \right]^{ - 1} = \left[ {\begin{array}{*{20}c} {{\tilde{\mathbf{V}}}_{1}^{ - 1} } & { - {\mathbf{V}}_{11}^{ - 1} {\mathbf{V}}_{12} {\tilde{\mathbf{V}}}_{2}^{ - 1} } \\ { - {\mathbf{V}}_{22}^{ - 1} {\mathbf{V}}_{21} {\tilde{\mathbf{V}}}_{1}^{ - 1} } & {{\tilde{\mathbf{V}}}_{2}^{ - 1} } \\ \end{array} } \right] \), where \( {\tilde{\mathbf{V}}}_{ 1} = {\mathbf{V}}_{ 1 1} { - }{\mathbf{V}}_{ 1 2} {\mathbf{V}}_{ 2 2}^{ - 1} {\mathbf{V}}_{ 2 1} \) and \( {\tilde{\mathbf{V}}}_{ 2} \, = \,{\mathbf{V}}_{ 2 2} { - }{\mathbf{V}}_{ 1 1}^{ - 1} {\mathbf{V}}_{ 1 2} \) are the Schur complements of V 11 and V 22 respectively; see e.g. Friedland (1986), pp. 479–481. Using the property that equally sized diagonal matrices commute, we can derive Eq. (2.151) from this more general expression.

  17. 17.

    This distinction is not very clear cut. For example to compute the oil and water ‘input’ rates, we make use of the fractional flows around the wells which are a direct function of the saturations, i.e. of state variables. In this sense the rates also contain indirect output information on the saturations around the wells.

  18. 18.

    Here, b is an arbitrary right-hand side.

  19. 19.

    Even if the reservoir is above bubble point such that it contains only oil and water and no free gas, the flow in the well bore will be three-phase because associated gas  will be released from the oil as the well-bore pressure decreases at increasing elevations above the reservoir. Alternatively, the lift table entries can be chosen as tubing-head pressure, oil rate, gas-oil ratio, and water-cut. Whatever the choice of the table entries, it is assumed that the fluid properties at standard conditions and the well-bore geometry do not change during the reservoir simulation.

  20. 20.

    See the footnote on page 46.

References

  • Batycky RP, Blunt MJ, Thiele MR (1997) A 3D field-scale streamline-based reservoir simulator. SPE Reserv Eng 12(4):246–254. doi:10.2118/36726-PA

    Google Scholar 

  • Bratvedt F, Gimse T, Tegnander C (1996) Streamline computations for porous media flow including gravity. Transp Porous Media 25(1):63–78. doi:10.1007/BF00141262

    Article  Google Scholar 

  • Brill JP, Mukherjee H (1999) Multi-phase flow in wells. SPE Monograph Series 17, SPE, Richardson

    Google Scholar 

  • Datta-Gupta A, King MJ (2007) Streamline simulation: Theory and Practice, SPE Textbook Series, 11. SPE, Richardson

    Google Scholar 

  • Friedland B (1986) Control system design—An introduction to state-space methods. McGraw-Hill, Reprinted in 2005 by Dover, New York

    Google Scholar 

  • Hubbert MK (1940) The theory of ground-water motion. J Geol 48(8):785–944

    Article  Google Scholar 

  • Karney BW, Seneviratne A (1991) Application of energy concepts to groundwater flow: time step control and integrated sensitivity analysis. Water Resour Res 27(12):3225–3235

    Article  Google Scholar 

  • King MJ, Datta-Gupta A (1998) Streamline simulation: a current perspective. In Situ 22(1):91–140

    Google Scholar 

  • Lanczos C (1970) The variational principles of mechanics, 4th edn. University of Toronto Press, Toronto. Reprinted in 1986 by Dover, New York

    Google Scholar 

  • Langhaar HL (1962) Energy methods in applied mechanics. Wiley, New York

    Google Scholar 

  • Oliver DS, Reynolds AC, Liu N (2008) Inverse theory for petroleum reservoir characterization and history matching. Cambridge University Press, Cambridge

    Book  Google Scholar 

  • Pollock DW (1988) Semi analytical computation of path lines for finite-difference models. Ground Water 26(6):743–750

    Article  Google Scholar 

  • Zienckiewicz OC, Taylor RL (1989) The finite element method, 4th ed., vol 1–2. McGraw-Hill, London

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jan Dirk Jansen .

Rights and permissions

Reprints and permissions

Copyright information

© 2013 The Author(s)

About this chapter

Cite this chapter

Jansen, J.D. (2013). System Models. In: A Systems Description of Flow Through Porous Media. SpringerBriefs in Earth Sciences. Springer, Heidelberg. https://doi.org/10.1007/978-3-319-00260-6_2

Download citation

Publish with us

Policies and ethics