cond-mat0606247/TPM.tex
1: 
2: \documentclass[oneside,english]{amsart}
3: \usepackage[T1]{fontenc}
4: \usepackage[latin1]{inputenc}
5: \usepackage{float}
6: \usepackage{graphicx}
7: \usepackage{amssymb}
8: 
9: 
10: 
11: 
12: 
13: 
14:  \numberwithin{equation}{section} %% Comment out for sequentially-numbered
15:  \numberwithin{figure}{section} %% Comment out for sequentially-numbered
16: 
17: \usepackage{lscape}
18: \usepackage[square]{natbib}
19: \usepackage{babel}
20: \makeatother
21: \begin{document}
22: 
23: \title{Permeability up-scaling using Haar Wavelets}
24: 
25: 
26: \author{V. Pancaldi,$^2$ K. Christensen,$^2$ P.R.
27: King$^1$ }
28: 
29: \maketitle \noindent $^{1}$Department of Earth Science and
30: Engineering, $^{2}$Department of Physics, Imperial College London,
31: United Kingdom
32: 
33: \begin{abstract}
34: In the context of flow in porous media, up-scaling is the coarsening
35: of a geological model and it is at the core of water resources research
36: and reservoir simulation. An ideal up-scaling procedure preserves
37: heterogeneities at different length-scales but reduces the computational
38: costs required by dynamic simulations. A number of up-scaling procedures
39: have been proposed. We present a block renormalization algorithm using
40: Haar wavelets which provide a representation of data based on averages
41: and fluctuations.
42: 
43: In this work, absolute permeability will be discussed for single-phase
44: incompressible creeping flow in the Darcy regime, leading to a finite
45: difference diffusion type equation for pressure. By transforming the
46: terms in the flow equation, given by Darcy's law, and assuming that
47: the change in scale does not imply a change in governing physical
48: principles, a new equation is obtained, identical in form to the original.
49: Haar wavelets allow us to relate the pressures to their averages and
50: apply the transformation to the entire equation, exploiting their
51: orthonormal property, thus providing values for the coarse permeabilities.
52: 
53: Focusing on the mean-field approximation leads to an up-scaling where
54: the solution to the coarse scale problem well approximates the averaged
55: fine scale pressure profile.
56: \end{abstract}
57: 
58: 
59: \section{Introduction}
60: 
61: The term up-scaling is used in reservoir engineering to refer to the
62: procedure by which a geological model is coarse grained into a flow
63: model. This is essential in modelling mass transport correctly to
64: gain an understanding of subsurface systems such as oil fields,
65: ground water flow and waste deposits. Correct estimation of the
66: transport properties of these reservoirs, including permeability, is
67: vital for their management. For example, in the contexts mentioned
68: above, good control of the fluid dynamics is necessary to ensure
69: optimization of recovery and the safety of the environment
70: \citep{Noet:futurestoch}. The procedure presented in this paper,
71: based on renormalization and wavelets, is a general coarse graining
72: technique, inspired by the wavelet treatment of the Ising model and
73: in line with the new developments that have been suggested in the
74: field of materials modelling \cite{Ism:W2005,Ism:W}.
75: 
76: In Section \ref{sub:Upscaling-techniques}, the main up-scaling
77: methods and related issues will be briefly reviewed. In Section
78: \ref{sec:Real-space-renormalization-and} a short account of
79: real-space renormalization and Haar wavelets will be given leading
80: to the description of the proposed method in Section
81: \ref{sec:A-renormalization-rule}. Numerical simulations and results
82: will be presented in Section \ref{sec:Numerical-simulations-and}.
83: 
84: The present paper is intended as a proof of concept of how wavelets
85: can be used in the field of upscaling by establishing a specific
86: formalism and applying it to the simplest cases. This allows us to
87: explore the underlying workings of the method, an essential step
88: towards the treatment of less trivial problems.
89: 
90: 
91: \subsection{\label{sub:Upscaling-techniques}Up-scaling techniques}
92: 
93: Numerous methods have been suggested for the coarsening of
94: permeability in geological systems, the simplest being averaging
95: techniques. As highlighted in a classic review \citep{Farmer:krev},
96: we can subdivide up-scaling methods into three categories:
97: deterministic, stochastic and heuristic. Further distinctions can be
98: made between analytical and numerical methods. The main issue with
99: up-scaling is the heterogeneity which characterizes natural porous
100: media on many different length scales. Heterogeneities range from
101: millimeters to kilometers, due to the great variety of types of
102: rocks and depositional processes that can be present in the same
103: system. Often, there is no clear division between the system size
104: and the length scales of the features or the size of the cells in
105: the model.
106: 
107: An analogy can be made between flow in porous media and currents
108: through resistors. This is possible because of the nature of the
109: equation for flow, Darcy's law, which is an elliptic equation
110: relating flow to the gradient of the pressure just like Ohm's law
111: relates current to voltage drop in conductors.
112: 
113: The problem of up-scaling is thus translated into solving the
114: Laplace-like differential equation, encouraging the application of
115: the wide range of methods which have been devised for this purpose
116: in other fields, for example field theoretical techniques,
117: perturbation expansions, effective medium theory, percolation
118: approaches or more simply finite differences and finite elements
119: methods, see Ref. \citep{Farmer:krev} for a recent review. A serious
120: drawback of these techniques, especially perturbation and effective
121: medium theory, is the underlying assumption that fluctuations in
122: permeability are small.
123: 
124: Renormalization offers an alternative, allowing for large
125: fluctuations in the system to be taken into account. Renormalization
126: techniques are a step-by-step approach where the system is coarsened
127: progressively, integrating out features on small length scales,
128: leading to the large scale effective permeability. Moreover,
129: renormalization can be applied to stochastic data sets by acting on
130: the probability distribution of the considered property rather than
131: on the single data points \cite{Hast:permup}.
132: 
133: With the exception of geological modelling techniques involving
134: object based methods and irregular grids, typically permeability
135: data is interpolated stochastically from the information gained at
136: precise locations in the reservoir. Hence, the emphasis is on
137: preserving the features of its statistical distribution rather than
138: the precise values. Furthermore, uncertainty pervades all stages of
139: reservoir modelling, from the measurement of permeability to the
140: estimation of the size of different rock type elements, rendering
141: statistical analysis the only viable tool to account for a range of
142: equiprobable scenarios which could represent the physical system
143: \cite{Chris:errors}.
144: 
145: Although there are various solutions to calculating effective
146: permeability for specific conditions, most of them have not been
147: implemented in the standard reservoir engineering packages for
148: industry. In practice, the methods of choice are often simple
149: averages, due to the ease and speed with which they can be
150: implemented and to the fact that precision in the estimation of
151: permeability in a specific location does not affect the uncertainty
152: implicit in the modelling process.
153: 
154: 
155: \section{\label{sec:Real-space-renormalization-and}Renormalization and Haar
156: Wavelet Transforms}
157: 
158: 
159: \subsection{Renormalization in up-scaling }
160: 
161: The concept of real-space renormalization has proved to be extremely
162: useful in estimating effective permeability efficiently
163: \citep{King:renkeff}. The basic idea behind this method is to start
164: with a lattice on which a property, in this case permeability, is
165: defined at each lattice cell. Successively the original cells are
166: grouped in a number of blocks, assigning new values for the
167: coarsened property. To avoid confusion, it is necessary to clarify
168: what is referred to by the words ``block'' and ``cell''. A cell is
169: the basic unit of the fine grid which typically characterizes the
170: geological model. Cell permeability is therefore what is commonly
171: referred to as fine permeability. A block is the basic unit of the
172: coarse grid used in flow simulations. The term block permeability
173: refers to the coarse equivalent permeability of the block,
174: calculated from the cell permeabilities through up-scaling
175: \cite{Wen:condrev}. This is clearly dependent on the boundary
176: conditions and is different from effective permeability, defined as
177: the permeability needed to relate the mathematical expectations of
178: the flow and of the pressure gradient. Due to the finite size of the
179: blocks it is only possible to consider equivalent permeability,
180: which ensures a match of flow patterns between the block and the
181: constitutive cells. After rescaling all the length scales, blocks
182: become cells and the result is a coarse-grained lattice with fewer
183: cells, but which still possesses the essential features of the
184: original system.
185: 
186: This procedure was first suggested by \cite{Kadanoff:blockspins} as
187: an efficient method to extrapolate the large scale behaviour of an
188: infinite system once fluctuations on smaller scales are averaged
189: out. The main advantage is that the procedure can be repeated until
190: the lattice has achieved the required coarseness with a low
191: computational cost, the algorithm being linear in the system size.
192: 
193: The renormalization transformation is by no means unique and many
194: different renormalization schemes have been proposed, some inspired
195: by an analogy between flow in porous media, percolation processes
196: and the flow of currents through resistors \cite{Will:Mathsoilren}.
197: 
198: Real-space transformations are a particular case of the more general
199: concept of the renormalization group. While the real-space version
200: already provides a versatile and fast technique for up-scaling, a
201: ``full'' real- and momentum-space renormalization method for
202: coarse-graining of subsurface reservoirs was presented by
203: Hristopoulos et al. \cite{hrist:RGoverview,hrist:RG1999}. This
204: general treatment has confirmed the applicability of the
205: renormalization concept to up-scaling, providing a solution of the
206: problem in all orders of perturbation, even for heterogeneous
207: systems where large fluctuations render other methods unsound.
208: 
209: \subsection{The Haar wavelet transform}
210: 
211: The mathematical concept of wavelets was first suggested in 1909 by
212: Haar \cite{Haar:wav}. It found its first application in the field of
213: seismology in 1989 in the work of Morlet \cite{Morelt:wav}
214: \emph{}and has since then been at the origin of a substantial number
215: of new approaches to various subjects, for example, biology
216: \cite{Wav:bio} and statistical mechanics \cite{Ism:W}. The basic
217: idea underlying wavelets is to decompose a function or a set of
218: data, in the continuous and discrete case respectively, into basic
219: components and their relative coefficients \cite{Daubechies:10lec}.
220: \emph{}In this sense it is very similar to a Fourier transform,
221: where the basic components are sines and cosines and the
222: coefficients are given by their amplitude. Wavelet transforms,
223: however, offer both spatial and frequency resolution. For this
224: reason they have been particularly successfully applied to the
225: analysis of signals where it is necessary to capture both underlying
226: periodic functions and specific localized features, which are almost
227: impossible to represent with periodic components.
228: 
229: At this point, however, a distinction between two different uses of
230: wavelets must be made. On one side, wavelets can be used to compress
231: information in terms of reducing the number of data points with a
232: filtering procedure. This has been applied extensively in the
233: context of up-scaling by Sahimi \cite{Sahimi:swa2004}, where a
234: filtering process reduces the number of permeability values in the
235: system without compromising its general statistics. On the other
236: side, a more {}``pervasive'' application of wavelets can lead to the
237: coarsening of permeability by acting on the flow equations
238: themselves. This approach has been suggested in statistical physics
239: to compress the information relative to spins and coupling constants
240: in the Ising model \cite{Ism:W} and then extended to include various
241: aspects of materials modelling \cite{Ism:W2005}.
242: 
243: The main point, already noted by Best \cite{Best:LandauGins}, is
244: that there is a striking similarity between the perspective of
245: renormalization and of wavelet transforms: both highlight the
246: features of a system in terms of large scale behaviour and
247: fluctuations away from it and both provide a connection between the
248: different relevant scales.
249: 
250: In this paper, the simplest type of discrete wavelet transform, the
251: Haar transform, is implemented in a renormalization method. Its
252: effect is to separate the average of the original data from the
253: fluctuations, expressed in terms of differences. Wavelets are
254: constructed through the scaling and shifting of the so called mother
255: wavelet.The Haar wavelet is defined as follows,
256: \cite{Daubechies:10lec}:
257: $\psi_{jk}\left(x\right)\equiv\psi\left(2^{j/2}x-k\right)$, where
258: $\psi\left(x\right)$ is the mother wavelet, $j\in z$ is the scale
259: parameter and $k$ is the shift. This leads to a Haar wavelet matrix
260: of the form: \[\mathbf{H}=\left[\begin{array}{cc}1 & \phantom{-}1\\1
261: & -1\end{array}\right].\]
262: 
263: 
264: For example, if we apply this transform to a $2\times1$ vector we
265: can obtain a new vector in terms of sums and differences of the
266: original values. As will be seen in Section
267: \ref{sec:A-renormalization-rule} this is a useful up-scaling scheme
268: valid in any dimension. This is a very simple transform, however,
269: the formalism described can be easily applied with any matrix
270: transform.
271: 
272: \subsection{The system: single-phase laminar flow}
273: 
274: The simple problem analysed in this paper is single-phase creeping
275: flow of a viscosity dominated incompressible fluid through a porous
276: medium. We will assume unit viscosity and ignore the effect of
277: gravity. The basic equation is Darcy's equation for flow,
278: $\mathbf{q=-K\nabla P}$, where $\mathbf{K}$ is permeability and
279: $\mathbf{\nabla P}$ is the gradient of pressure, combined with the
280: continuity equation, $\mathbf{\nabla\cdot q}=0$, which give rise to
281: a Laplace-like differential equation:
282: $\mathbf{\nabla}\cdot(\mathbf{K\nabla P})=0$.
283: 
284: The discretization was performed by specifying the permeability
285: values at the cell centres and assuming pressure to be piece-wise
286: linear across the cell. Transmissibility is equal to permeability in
287: the case of unit volume of the discretization grid cell:
288: $t_{i}=k_{i}/\Delta x$, where $\Delta x=1$ is the size of the grid
289: cell. Assuming transmissibility $t_{i}$ to be piecewise constant
290: with an interface between $t_{i}$ and $t_{i+1}$ at the cell boundary
291: and imposing flow conservation, the inter cell transmissibility,
292: $t_{ij}$ is found to be the harmonic mean of $t_{i}$ and $t_{j}$
293: \cite{Aziz:1979}.
294: 
295: \begin{equation}
296: t_{ij}=\frac{t_{i}t_{j}}{t_{i}+t_{j}}=\frac{1}{1/t_{i}+1/t_{j}};\,\,\,\,\,\,
297: t_{ij}\left(t_{j}=0\right)=0;\,\,\,\,\,\,
298: t_{ij}\left(t_{j}=\infty\right)=t_{i}=k_{i}\label{eq:trans}\end{equation}
299:  As described in \cite{Aziz:1979}, this constitutes a satisfactory
300: approximation if the properties do not change excessively between
301: adjacent cells. \footnote{In the literature, the term ``block'' is
302: used to refer to what we call cells. Our choice is motivated to
303: avoid confusion given our precise definition of a block. }
304: 
305: Assuming permeability to be a diagonal tensor, as in an isotropic
306: medium, mass balance equations for the system give rise to a
307: five-point scheme finite-difference equation expressed in matrix
308: form:
309: 
310: \begin{equation}
311: \mathbf{TP=R}.\label{Eq:main}
312: \end{equation}
313:  Here, for a one-dimensional system of linear size $N$, $\mathbf{T}$
314: is an $N\times N$ matrix of transmissibilities, $\mathbf{P}$ is an
315: $N\times1$ pressure vector and $\mathbf{R}$ is an $N\times1$
316: boundary condition vector \cite{King:Upk,Aziz:1979}.
317: 
318: No-flow boundary conditions were imposed at the top and bottom of
319: the entire system by setting the cell permeabilities to zero, such
320: that also the transmissibilities in this region would be zero. A
321: pressure gradient in the horizontal direction was established by
322: setting permeability at the left and right boundaries equal to
323: infinity so as to generate transmissibilities at these interfaces
324: which are identical to the local permeabilities, see Equation
325: (\ref{eq:trans}). These global boundary conditions correspond to
326: imposing no flow at the top and bottom of the block and to a
327: constant pressure profile along the left and right boundaries.
328: Clearly, these boundaries can be rotated to calculate vertical
329: permeability. As outlined in \citep{Durlofsky:krev2005}, a different
330: choice of boundary conditions, for instance periodic, would not
331: alter the result significantly, given the local nature of the
332: up-scaling process.
333: 
334: In a system of dimension $d=1$, the matrix $\mathbf{T}$ has a
335: tridiagonal shape, arising from the coupling of each cell with its
336: two nearest neighbours and with itself, while in $d>1$ dimensions
337: further couplings are introduced leading to a diagonally dominant
338: sparse matrix with $2d$ non-zero off-diagonals, see Figure
339: \ref{cap:Transmissibility-matrix-structure}.
340: 
341: \begin{figure}
342:  $\begin{array}{ccc}
343: \resizebox{0.30\textwidth}{!}{\includegraphics{K41d.eps}}&
344: \resizebox{0.30\textwidth}{!}{\includegraphics{K42d.eps}}&
345: \resizebox{0.30\textwidth}{!}{\includegraphics{K43d.eps}}\\
346: [-0.3cm] \mbox{\small (a)} & \mbox{\small (b)} & \mbox{\small (c)}
347: \end{array}$
348: \caption{ \label{cap:Transmissibility-matrix-structure}Structure of
349: the transmissibility matrix $\mathbf{T}$ for $N=4$ in (a) $d=1$, (b)
350: $d=2$, and (c) $d=3$.}
351: \end{figure}
352: 
353: 
354: 
355: \section{\label{sec:A-renormalization-rule} Renormalization based on Haar
356: Wavelet Transforms }
357: 
358: 
359: \subsection{One-dimensional system}
360: 
361: As mentioned in Section \ref{sec:Real-space-renormalization-and},
362: wavelets can be used to decompose the behaviour of a system into
363: averages and fluctuations. For example, if we consider a
364: one-dimensional system consisting of two grid cells where pressure
365: is defined, the pressure can be expressed with the two cell values
366: or in terms of the average and semi-difference:
367: 
368: \begin{equation}
369: \mathbf{P'}=\mathbf{WP}=\left[\begin{array}{c}
370: \mathbf{\Sigma}\\
371: \Delta\end{array}\right],\end{equation} where the matrix
372: $\mathbf{W}$, the pressure vector $\mathbf{P}$, the average
373: $\Sigma$, and the semi-difference $\Delta$ are given by:
374: 
375: \begin{equation}
376: \mathbf{W}=\frac{1}{2}\left[\begin{array}{cc}
377: 1 & \phantom{-}1\\
378: 1 & -1\end{array}\right];\,\mathbf{P}=\left[\begin{array}{c}
379: p_{1}\\
380: p_{2}\end{array}\right];\,\Sigma=\frac{\begin{array}{c}
381: p_{1}+p_{2}\end{array}}{2};\,\Delta=\frac{\begin{array}{c}
382: p_{1}-p_{2}\end{array}}{2}.\label{Eq:1d}\end{equation}
383: 
384: 
385: The matrix $\mathbf{W}$ relates the original pressure variable
386: $\mathbf{P}$ to the new pressure variable $\mathbf{P'}$. Thus if we
387: operate on the pressure vector of Equation (\ref{Eq:main}) with
388: $\mathbf{W}$, a new pressure vector $\mathbf{P'}$ can be obtained,
389: where the first element is the average of the original pressures,
390: see Equation (\ref{Eq:1d}). This matrix is simply $1/2\,\mathbf{H}$,
391: where $\mathbf{H}$ is the Haar transform matrix for a $1\times2$
392: system.
393: 
394: Let us consider a $1\times N$ system, with $N=4$, that we want to
395: coarsen by a factor $n=2$ by transforming a $1\times4$ group of
396: cells into a $1\times2$ group of blocks. We will have:
397: 
398: 
399: \begin{equation}
400: \mathbf{W}=\frac{1}{2}\left[\begin{array}{cccc}
401: 1 & \phantom{-}1 & 0 & \phantom{-}0\\
402: 0 & \phantom{-}0 & 1 & \phantom{-}1\\
403: 1 & -1 & 0 & \phantom{-}0\\
404: 0 & \phantom{-}0 & 1 &
405: -1\end{array}\right];\,\mathbf{P}=\left[\begin{array}{c}
406: p_{1}\\
407: p_{2}\\
408: p_{3}\\
409: p_{4}\end{array}\right];\end{equation}
410: 
411: 
412: \begin{equation}
413: \mathbf{P'=\left[\mathbf{\begin{array}{c}
414: \mathbf{\Sigma}\\
415: \mathbf{\Delta}\end{array}}\right]};\mathbf{\,\Sigma}=\left[\begin{array}{c}
416: \frac{\begin{array}{c}
417: p_{1}+p_{2}\end{array}}{2}\\
418: \frac{\begin{array}{c}
419: p_{3}+p_{4}\end{array}}{2}\end{array}\right];\,\mathbf{\Delta}=\left[\begin{array}{c}
420: \frac{\begin{array}{c}
421: p_{1}-p_{2}\end{array}}{2}\\
422: \frac{\begin{array}{c}
423: p_{3}-p_{4}\end{array}}{2}\end{array}\right].\end{equation}
424: 
425: 
426: An important property of $\mathbf{W}$ is that the product
427: $\mathbf{WW^{T}}$ is the identity matrix multiplied by a factor of
428: $1/n$. $\mathbf{WW^{T}}$can be therefore inserted altering Equation
429: \ref{Eq:main} only by a factor of $n$:
430: 
431: \begin{equation}
432: \mathbf{TW^{T}WP}=\frac{1}{n}\mathbf{R}.\end{equation} To complete
433: the equation transformation we multiply by $\mathbf{W}$ on both
434: sides to obtain a new transmissibility matrix and a new boundary
435: condition vector applied to the transformed pressure:
436: 
437: \begin{equation}
438: \mathbf{\left(WTW^{T}\right)WP}=\frac{1}{n}\mathbf{WR}.\end{equation}
439: Defining the transformed variables,
440: 
441: \begin{equation}
442: \mathbf{T'=WTW^{T}};\mathbf{\,\,\,\,\, P'=WP};\mathbf{\,\,\,\,\,
443: R'=WR};\end{equation}
444: 
445: 
446: \noindent we have
447: 
448: \begin{equation}
449: \mathbf{T'P'=}\frac{1}{n}\mathbf{R'}.\label{eq:app}\end{equation} Up
450: to this point, the transformation has been completely reversible; in
451: fact, we have simply changed the variables with which we represent
452: the system. Now we approximate Equation (\ref{eq:app}) by ignoring
453: the fluctuations of the systems to preserve the large scale
454: behaviour. To do this, we define new variables
455: $\mathbf{\mathcal{P}}$ and $\mathbf{\mathcal{R}}$ composed of the
456: first $(N/2)$ elements of $\mathbf{P'}$ and $\mathbf{R'}$
457: respectively, and $\mathbf{\mathcal{T}}$ as the $(N/2)\times(N/2)$
458: upper left corner of $\mathbf{T'}$.
459: 
460: \begin{center}$\mathbf{T}=\left[\begin{array}{cccc}
461: 2k_{1}+t_{12} & -t_{12} & 0 & 0\\
462: -t_{12} & t_{12}+t_{23} & -t_{23} & 0\\
463: 0 & -t_{23} & t_{23}+t_{34} & -t_{34}\\
464: 0 & 0 & -t_{34} & t_{34}+2k_{4}\end{array}\right];$\end{center}
465: 
466: \begin{center}$\mathbf{T'}=\left[\begin{array}{cccc}
467: 2k_{1}+t_{23} & -t_{23} & 2k_{1}-t_{23} & -t_{23}\\
468: -t_{23} & t_{23}+2k_{4} & t_{23} & t_{23}-2k_{4}\\
469: 2k_{1}-t_{23} & t_{23} & t_{23}+t_{34} & t_{23}\\
470: -t_{34} & t_{23}-2k_{4} & t_{23} &
471: 4t_{34}+2k_{4}+t_{23}\end{array}\right];$\end{center}
472: 
473: \begin{equation}
474: \mathbf{T'}=\left[\begin{array}{cc}
475: A & B\\
476: B^{T} &
477: C\end{array}\right];\,\,\,\,\,\,\,\,\,\,\,\mathbf{\mathcal{T}}=A=\left[\begin{array}{cc}
478: 2k_{1}+t_{23} & -t_{23}\\
479: -t_{23} & t_{23}+2k_{4}\end{array}\right].\end{equation}
480: 
481: 
482: To determine the coarse pressure, we invert the renormalised
483: transmissibility matrix $\mathcal{T}$ and multiply the resulting
484: pressure by $2$. This rescale is necessary to compensate for the
485: change from cell values to block values, which has doubled the size
486: of $\Delta x$.
487: 
488: \begin{equation}
489: \mathcal{TP}=\frac{1}{2}\mathcal{R};\,\,\,\,\mathcal{P}=\frac{1}{2}\mathcal{T}^{-1}\mathcal{R};\,\,\,\,\mathbf{P}_{coarse}=2\mathcal{P}.\label{eq:inversion}\end{equation}
490: 
491: 
492: \noindent Using $\mathbf{\mathcal{T}}$, $\mathbf{\mathcal{P}}$, and
493: $\mathbf{\mathcal{R}}$ corresponds to assuming that fluctuations of
494: pressures $\mathbf{\Delta}$, are negligible. In other words, we
495: represent the system in what is commonly called a mean-field
496: approximation where only the average behaviour of the pressure field
497: is considered. Hence, exploiting the orthonormal property of
498: $\mathbf{W}$, an expression for the coarse transmissibility can be
499: derived, by operating on Darcy's equation on the fine scale, leading
500: to a mean-field pressure solution. The general principle underlying
501: this method, can be applied in any dimension and to all problems
502: which require coarsening.
503: 
504: \subsection{Two- and three-dimensional systems}
505: 
506: In $d$-dimensions a similar treatment can be performed, where the
507: equivalent of a linear arrangement of $N$ cells is a $d$-hypercube
508: of linear size $N$ which we want to coarsen by a factor of $2$ in
509: each direction. In this case a convention for the ordering of the
510: pressures in the vector is needed. The coefficient in the
511: $\mathbf{W}$ matrix and the pressure rescale factor is now
512: $1/2^{d}$. Moreover, while it is easy to write down expressions for
513: the average and difference for two cell values, a complication
514: arises when cells are averaged in a dimension equal or higher than
515: two. In this case, the pressures are averaged 4 at a time and there
516: is no unique way to define their difference. For example, the
517: $\mathbf{W}$ matrix and $\mathbf{P'}$ for a $2\times2$ system can be
518: given by:
519: 
520: \begin{equation}
521: \mathbf{W}=\frac{1}{4}\left[\begin{array}{cccc}
522: 1 & \phantom{-}1 & \phantom{-}1 & \phantom{-}1\\
523: 1 & -1 & \phantom{-}1 & -1\\
524: 1 & \phantom{-}1 & -1 & -1\\
525: 1 & -1 & -1 &
526: \phantom{-}1\end{array}\right];\,\,\,\,\mathbf{P'}=\frac{1}{4}\left[\begin{array}{c}
527: p_{1}+p_{2}+p_{3}+p_{4}\\
528: p_{1}-p_{2}+p_{3}-p_{4}\\
529: p_{1}+p_{2}-p_{3}-p_{4}\\
530: p_{1}-p_{2}-p_{3}+p_{4}\end{array}\right],\end{equation} but this is
531: by no means the only valid choice. The constraints on $\mathbf{W}$
532: are that the top row should produce the pressures average, that
533: $\mathbf{WW^{T}}$ is proportional to the identity and that all rows
534: are orthonormal to the top one.
535: 
536: While in one dimension the flow follows a forced path, already in
537: two dimensions we can recover many of the characteristics of
538: transport phenomena. Moreover, when looking at the elements of the
539: matrix $\mathbf{\mathcal{T}}$ for the two-dimensional system, it was
540: noted that the block permeability can be obtained by performing a
541: specific average of the cell permeabilities, see Figure
542: \ref{cap:A-schematic-representation} and Appendix.
543: 
544: For a $4\times4$ system, the transmissibility matrix is
545: $16\times16$. When transformed with $\mathbf{W}$ and
546: $\mathbf{W^{T}}$ the matrix obtained is still $16\times16$, but
547: taking the first four rows and columns only, we get a $4\times4$
548: matrix. This can be compared to the transmissibility matrix of a
549: $2\times2$ system to deduce the relation between the permeabilities
550: at cell and block level, see Appendix.
551: 
552: 
553: \begin{figure}[H]
554: \begin{center}\includegraphics[%
555:   scale=0.25]{renbw.eps}\end{center}
556: 
557: \begin{center}\includegraphics[%
558:   scale=0.4]{upscheme13.eps}\end{center}
559: 
560: 
561: \caption{\label{cap:A-schematic-representation}A schematic
562: representation of the relation between cell and block permeabilities
563: and transmissibilities. One step in the renormalization algorithm.
564: (\textbf{a}) $8\times8$ permeability map. (\textbf{b}) The
565: $4\times4$ coarsened permeability map. Notice how a $4\times4$ group
566: of cells is substituted by a $2\times2$ group of blocks.
567: (\textbf{c}) Blow-up of a $2\times2$ group of cells. (\textbf{d})
568: Blow-up of a $2\times2$ group of blocks. \textbf{\emph{}}Properties
569: of cells are subscripted with numbers, properties of blocks with
570: letters. Permeabilities are indicated by $k$ and transmissibilities
571: with $t$, $k{}_{A}=\left(k_{1}+k_{2}\right)/2,\,
572: t{}_{AB}=\left(t_{23}+t_{67}\right)/2,\,
573: t{}_{AC}=\left(t_{59}+t_{610}\right)/2$.}
574: \end{figure}
575: 
576: 
577: Accordingly, a renormalization algorithm was implemented whereby
578: groups of $4\times4$ cells are progressively substituted by groups
579: of $2\times2$ blocks, until the required degree of coarsening in
580: permeability is achieved. This procedure is fast and can be further
581: improved with the use of parallel computing. Finally,
582: $\mathbf{\mathcal{T}}$ is inverted to obtain the coarse pressure,
583: see Equation (\ref{eq:inversion}). The procedure in $d$-dimensions
584: is as follows:
585: 
586: \begin{enumerate}
587: \item Start with a permeability map, linear size $N$ multiple of 4. Calculate
588: the pressure by inverting the transmissibility matrix, see Equation
589: (\ref{Eq:main}).
590: \item Subdivide the system into groups of $4^{d}$ cells. Substitute each
591: group with a new group of $2^{d}$ blocks, calculating the new
592: permeability according to the averaging rule, see Appendix and
593: Figure \ref{cap:A-schematic-representation}.
594: \item The new system has a factor of $2^{d}$ less cells. Calculate the
595: upscaled pressure by inverting the new transmissibility matrix and
596: rescaling, Equation (\ref{eq:inversion}).
597: \end{enumerate}
598: Clearly the higher the dimension, the bigger the advantage in
599: avoiding a double matrix multiplication.
600: 
601: It should be stressed that this renormalization scheme derives
602: directly from the representation of the problem in the mean-field
603: approximation and from the choice of $\mathbf{W}$ matrix. This
604: result relates the elimination of permeability fluctuations to the
605: smoothening of fluctuations in pressure, revealing the basic
606: principle underlying renormalization methods for up-scaling.
607: Importantly, it also represents the starting point for devising a
608: controlled method to include the effects of fluctuations in the
609: coarsening process.
610: 
611: 
612: \section{\label{sec:Numerical-simulations-and}Numerical Simulations and Heterogeneities}
613: 
614: 
615: \subsection{Stochastically generated correlated permeability}
616: 
617: To emphasize the importance of maintaining the statistical
618: properties of the permeability distribution, various realizations
619: were generated with the same moments. Permeability was simulated as
620: a random, log-normally distributed correlated variable on two- and
621: three-dimensional Cartesian regular grids with a moving average
622: technique \cite{Wall:cor}.
623: 
624: The starting point is an uncorrelated field, that is normally or
625: uniformly distributed random numbers are assigned to each cell. Then
626: the correlation is introduced by averaging these values with a
627: moving circle technique \cite{Wall:cor}. By the central limit
628: theorem, the new distribution remains normal, at least for
629: sufficiently big circles, independent of the statistics of the
630: initial data. Moreover, the correlation length is related to the
631: radius of the circle used in the averaging process. Permeability is
632: then taken to be the exponential of this distribution. Anisotropic
633: systems can be generated by using ellipses to account for different
634: rock types in the simulated reservoir.
635: 
636: The upscaled pressure was compared with the simple averaging of the
637: fine pressure. Errors were calculated as differences between the two
638: pressure solutions at the same coarsening level and then averaged
639: over the entire system. While the average error is a useful measure
640: of accuracy, localizing the discrepancies allows us to look for
641: their justification in view of heterogeneities.
642: 
643: 
644: \subsection{\label{sub:Analysis-of-heterogeneity}Analysis of heterogeneity in
645: permeability distribution}
646: 
647: The simplest test cases to be analysed are two layered systems where
648: exact analytical solutions are known. More precisely, the equivalent
649: permeability for flow parallel to the strata is the arithmetic
650: average of the different permeabilities, and for perpendicular flow
651: it is the harmonic average. In general, these two averages can be
652: shown to be respectively the lower and upper limit on the effective
653: permeability of any system \cite{Farmer:krev}. As can be expected
654: the new renormalization technique is just as good in these cases as
655: others of its kind. It must be noted that while renormalization
656: according to the resistor analogy produces a final number
657: corresponding to  the equivalent permeability, the last step of the
658: wavelet method can only lead to a $2\times2$ cell. A number can be
659: obtained afterwards, but this is necessarily going to be some kind
660: of average. For example, in the case of vertical layering, while at
661: the third coarsening step the renormalization method already has a
662: homogeneous character, the wavelet method still has a layered
663: structure. The correct result, that is, the harmonic mean, can be
664: recovered by taking the harmonic mean of the two layers. In the case
665: of a chess-board configuration, where resistor analogy
666: renormalization underestimates permeability with an error increasing
667: with the difference between the two layer permeabilities
668: \cite{Zimm:ren_comp}, the wavelet method overestimates it by an even
669: larger amount. In this case it is possible to show analytically that
670: the exact result should be the geometric mean \cite{Farmer:krev}
671: while the wavelet method result is the arithmetic mean, as is
672: expected given the averaging which takes place in the algorithm.
673: This is not ideal but at least the error can be predicted and its
674: source clearly identified, see Table
675: \ref{cap:Comparison-of-wavelet}.
676: \begin{table}[H]
677: \begin{center}\begin{tabular}{llll}
678: \hline $k_{1}$= 2500, $k_{2}$= 5000& $k_{\textrm{resistor}}$&
679: $k_{\textrm{wavelet}}$& $k_{\textrm{exact}}$\tabularnewline \hline
680: Perpendicular layering& 3333& 3333& 3333\tabularnewline Parallel
681: layering& 3750& 3750& 3750\tabularnewline Chess-board& 3429& 3750&
682: 3535.5\tabularnewline \hline
683: \end{tabular}\end{center}
684: 
685: 
686: \caption{\label{cap:Comparison-of-wavelet}Comparison of effective
687: permeability obtained by resistor and wavelet based renormalization
688: for layered and chess-board systems with cells of low ($k_{1}$) and
689: high ($k_{2}$) homogeneous permeability reduced to a single cell.
690: Both methods predict the exact results correctly for the layered
691: cases while both fail in the chess-board case. }
692: \end{table}
693: 
694: 
695: Initially two-dimensional systems were analysed so the method
696: described will refer to this case. Results are also presented in
697: three dimensions, where the procedure is identical in concept.
698: 
699: First, an analysis was made on the permeability distribution at each
700: up-scaling step, see Table \ref{cap:Distribution-of-permeabilities}.
701: 
702: %
703: \begin{table}[H]
704: \begin{center}\begin{tabular}{ccc}
705: \hline Cell size& Mean& Std\tabularnewline \hline $\phantom{1}$1&
706: 4902.9& 11.8597\tabularnewline $\phantom{1}$2& 4902.8&
707: 11.54\tabularnewline $\phantom{1}$4& 4902.9& 11.05\tabularnewline
708: $\phantom{1}$8& 4903.1& 10.24\tabularnewline 16& 4903.9&
709: $\phantom{1}$8.36\tabularnewline \hline
710: \end{tabular}\end{center}
711: 
712: 
713: \caption{\label{cap:Distribution-of-permeabilities}Statistics of
714: permeability distribution at each coarsening step, for a
715: $64\times64$ system, with correlation length $r=10$ averaged over 10
716: realizations. At each up-scaling step the cell size doubles. Notice
717: how the renormalization preserves the mean and how the standard
718: deviation starts to decrease considerably only when the cell size is
719: comparable to the correlation length.}
720: \end{table}
721: 
722: 
723: Once the permeability maps had been generated, pressure boundary
724: conditions were set on the left and right boundaries of the system.
725: These were taken to be fixed at 100 on the left and 50 on the right.
726: The drop in pressure across the system is a fundamental factor in
727: determining the errors in the estimates. However, the use of
728: relative errors mitigates this effect and the same boundary
729: conditions were used in all the simulations. A pressure profile was
730: obtained at each renormalization step inverting the corresponding
731: transmissibility matrix with the correct renormalized boundary
732: conditions and compared to an equally coarsened pressure obtained by
733: successively averaging fine pressure on $2\times2$ cells, see Figure
734: \ref{cap:Wavelet-transform-based}. The process was repeated $10$
735: times to generate a distribution of results.
736: 
737: %
738: \begin{figure}[H]
739: \begin{center}\includegraphics[%
740:   scale=0.25]{good32k.eps}\includegraphics[%
741:   scale=0.25]{good32p.eps}\includegraphics[%
742:   scale=0.25]{good16av.eps}\end{center}
743: 
744: \begin{center}\includegraphics[%
745:   scale=0.25]{good16k.eps}\includegraphics[%
746:   scale=0.25]{good16p.eps}\includegraphics[%
747:   scale=0.25]{gooddiff.eps}\end{center}
748: 
749: 
750: \caption{\label{cap:Wavelet-transform-based}Wavelet renormalization
751: of a permeability map from $32\times32$ to $16\times16$.
752: (\textbf{a1}) Fine scale permeability. (\textbf{a2}) Fine scale
753: pressure solution obtained from fine scale permeability.
754: (\textbf{a3}) Average of the fine scale pressure solution
755: ($2\times2$ cells averaged). (\textbf{b1}) Wavelet renormalized
756: coarse permeability. (\textbf{b2}) Coarse pressure solution obtained
757: from coarse permeability. (\textbf{b3}) Modulus of relative error,
758: $\left|\textrm{a3-b2}\right|$/a3. In this case the relative
759: difference between the averaged fine scale pressure (\textbf{a3})
760: and the coarse pressure (\textbf{b2}) is within 2\%. This procedure
761: was repeated for systems with varying permeability ranges and with
762: different heterogeneities, simulating different rock types.}
763: \end{figure}
764: 
765: 
766: When averaged over many realizations, the absolute error between the
767: averaged fine scale pressure and the coarse pressure obtained from
768: the wavelet upscale was consistently found to be of order $10^{-3}$.
769: For this kind of systems, errors in a single realizations did not
770: exceed 5\%.
771: 
772: As expected, the error was found to be higher with higher standard
773: deviation of the permeability, see Table
774: \ref{cap:Comparison-of-error r=3D1}, but only for very heterogeneous
775: systems, where the standard deviation is an order of magnitude
776: larger than the mean.
777: 
778: %
779: \begin{table}[H]
780: \begin{center}\begin{tabular}{ccc}
781: \hline $\sigma/\mu$& Mean relative error (10$^{-3}$)& Std of error
782: (10$^{-3}$)\tabularnewline \hline $\phantom{1}$0.1& $-$5.23&
783: 3.41\tabularnewline $\phantom{1}$0.2& $-$0.74& 3.47\tabularnewline
784: $\phantom{1}$0.4& $-$0.84& 3.34\tabularnewline $\phantom{1}$0.8&
785: $-$1.46& 3.15\tabularnewline $\phantom{1}$1$\phantom{.1}$&
786: $\phantom{-}$0.58& 3.64\tabularnewline $\phantom{1}$2$\phantom{.1}$&
787: $\phantom{-}$1.82& 3.52\tabularnewline 10$\phantom{.1}$&
788: $\phantom{-}$0.79& 4.71\tabularnewline \hline
789: \end{tabular}\end{center}
790: 
791: 
792: \caption{\label{cap:Comparison-of-error r=3D1}Comparison of mean and
793: standard deviation of the relative error at different standard
794: deviation of permeability and same correlation length $r$=3,
795: $\mu$=10000, averaged over entire system. All data averaged over 27
796: realizations of $32\times32$ systems being upscaled to $16\times16$.
797: }
798: \end{table}
799: 
800: 
801: %
802: \begin{table}[H]
803: \begin{center}\begin{tabular}{ccc}
804: \hline Correlation length& Mean relative error (10$^{-3}$)& Std of
805: error (10$^{-3}$)\tabularnewline \hline 1& $-$0.52&
806: 3.85\tabularnewline 2& $-$0.49& 3.51\tabularnewline 3&
807: $\phantom{-}$0.67& 3.46\tabularnewline 4& $\phantom{-}$0.64&
808: 3.45\tabularnewline 5& $\phantom{-}$0.66& 3.43\tabularnewline \hline
809: \end{tabular}\end{center}
810: 
811: 
812: \caption{\label{cap:Comparison-of-error std=3D}Comparison of error
813: for different correlation lengths but same standard deviation,
814: $\sigma$=1000, $\mu=1000$ (average of multiple realizations of
815: $32\times32$ systems being upscaled to $16\times16$, see text for
816: details about the number of realizations). Notice a very weak
817: dependence of the standard deviation of the error on the correlation
818: length that seems to suggest that a more correlated system can be
819: upscaled more accurately.}
820: \end{table}
821: 
822: 
823: 
824: Next, a comparison between realizations with varying correlation
825: length $r$, expressed in terms of grid cells, and equal standard
826: deviation in permeability was made. A different number of
827: realizations were averaged depending on the correlation length of
828: the system, considering that each subsystem of linear size equal to
829: the correlation length constitutes a sample in statistical terms
830: (number of realizations = $3r^{2}$). As can be seen in Table
831: \ref{cap:Comparison-of-error std=3D}, the more the field is
832: correlated, that is, the larger the value of $r$, the better the
833: wavelet renormalization method approximates the fine scale pressure
834: average. However, even at a radius of correlation equal to one grid
835: cell, the average standard deviation of the error is within 0.4\%.
836: 
837: While the error averaged over the entire system can be misleadingly
838: small, due to cancellations which occur between positively and
839: negatively biased results at specific locations, the standard
840: deviation of the error over the system can be taken as a faithful
841: indicator of the performance of the method.
842: 
843: %
844: \begin{figure}
845: \begin{center}\includegraphics[%
846:   scale=0.25]{k32easy.eps}\includegraphics[%
847:   scale=0.25]{p32easy.eps}\end{center}
848: 
849: \begin{center}\includegraphics[%
850:   scale=0.25]{p16easy.eps}\includegraphics[%
851:   scale=0.25]{p16reneasy.eps}\end{center}
852: 
853: \begin{center}\includegraphics[%
854:   scale=0.25]{erreasy.eps}\includegraphics[%
855:   scale=0.25]{errren.eps}\end{center}
856: 
857: 
858: \caption{\label{cap:Wavres comp}Comparison of resistor and wavelet
859: renormalization. (\textbf{a}1) Fine scale permeability. (a2) Fine
860: scale pressure solution obtained from fine scale permeability. (b)
861: Coarse pressure solution obtained with wavelet method. (c) Coarse
862: pressure solution obtained with resistor method. (d) Modulus of
863: relative error of pressure obtained with the wavelet method with
864: respect to the pressure average. (e) Modulus of relative error of
865: pressure obtained with the resistor method with respect to the
866: pressure average. The error in the wavelet renormalization is of
867: order $10^{-3}$ because the permeability field is fairly
868: homogeneous. However, the resistor renormalization is less effective
869: even in this ideal case. }
870: \end{figure}
871: 
872: 
873: A comparison with the resistor renormalization performed according
874: to \cite{King:immis} (equation 2.1), can be seen in Figure
875: (\ref{cap:Wavres comp}). It is possible to develop a more accurate
876: resistor renormalization algorithm by considering the anisotropy
877: generated by the upscaling process. However, this algorithm is not
878: as immediate as the wavelet renormalization algorithm to implement,
879: requiring the definition of two transmissibilities per cell. It must
880: be noted that, while the wavelet method is geared towards
881: reproducing the average pressure, the resistor method is based on
882: flux conservation, thus it is not surprising that the results of the
883: two methods differ.
884: 
885: 
886: \subsection{Shales}
887: 
888: One of the major drawbacks of the renormalization proposed by
889: \cite{King:renkeff} is its imprecise treatment of shales. When the
890: permeability contrast between adjacent cells is high, for example at
891: the interface between permeable rock such as sandstone, and
892: impermeable elements such as shales, the analogy with resistors
893: gives inaccurate predictions. This results in a deformation of
894: shales which can lead to misjudgment of the reservoir connectivity.
895: Typically, shales have a large aspect ratio and they are distributed
896: horizontally often constituting a barrier to flow in the vertical
897: direction. A successful alternative approach to shales is given in
898: Ref. \cite{Begg:vertsh}, where the permeability is related to the
899: length of the path going around the shale bodies.
900: 
901: Shales were implemented in the following way: some of the sites of
902: the system were chosen at random and shales of random size and
903: aspect ratio were created by setting the permeabilities in the area
904: to a very small value ($10^{-13}$). Another conventional way of
905: implementing shales into a model is to make correlation very
906: anisotropic. This causes areas of low permeability to naturally
907: emerge with the correct aspect ratio and orientation. However, the
908: chosen method provides a much greater difference between the low
909: permeability of the shales and the distribution of the permeability
910: in the sand, which is often characteristic of physical systems.
911: 
912: As can be seen in Figure \ref{cap:Shale error} and Table
913: \ref{cap:shalestable}, shales are correctly upscaled unless their
914: size becomes comparable to the size of the cells.
915: 
916: %
917: \begin{table}[H]
918: \begin{center}\begin{tabular}{ccccc}
919: \hline Max width& Max height& Shale fraction (\%)& Mean
920: error(10$^{-3}$)& Std of error(10$^{-3}$)\tabularnewline \hline
921: $\phantom{1}$2& $\phantom{1}$2& $\phantom{1}$3.2&
922: $\phantom{1}$18.29& $\phantom{1}$16.2\tabularnewline $\phantom{1}$2&
923: $\phantom{1}$2& 33.4& 113.85& 100.8\tabularnewline 16&
924: $\phantom{1}$5& 16.4& $\phantom{11}$9.1$\phantom{1}$&
925: $\phantom{1}$25.3\tabularnewline 16& $\phantom{1}$5& 33.4&
926: $\phantom{11}$6.73& $\phantom{1}$27.1\tabularnewline 16&
927: $\phantom{1}$5& 57.6& $\phantom{11}$3.1$\phantom{1}$&
928: $\phantom{1}$24.9\tabularnewline $\phantom{1}$5& 16& 18.8&
929: $\phantom{1}$48.7& $\phantom{1}$46$\phantom{1}$\tabularnewline
930: $\phantom{1}$5& 16& 36$\phantom{.1}$& $\phantom{1}$26.8&
931: $\phantom{1}$47.9\tabularnewline $\phantom{1}$5& 16& 52.2&
932: $\phantom{1}$20.3& $\phantom{1}$60$\phantom{1}$\tabularnewline
933: \hline
934: \end{tabular}\end{center}
935: 
936: 
937: \caption{\label{cap:shalestable}Error in up-scaling a system with
938: shales with different aspect ratio. Shale permeability set to
939: $10^{-13}$ . All values were averaged over 3 runs. Notice that
940: vertical and small shales are associated with a bigger error.}
941: \end{table}
942: 
943: 
944: When either the shale fraction or the sand fraction approaches the
945: percolation threshold, the error of the \emph{}wavelet method
946: calculated with respect to the average of the fine pressure solution
947: can be of order $10^{-1}$. In this case shales will either cover the
948: entire system or tend to disappear. Anisotropy also plays an
949: important role. Shales perpendicular to the flow seem to represent
950: more of a problem, since they oppose the pressure gradient, see
951: Table \ref{cap:shalestable}, bottom three entries. For example, in
952: Figure \ref{cap:Shale error}, the largest error occurs in the lower
953: central region where a vertical barrier disappears in the coarsening
954: process. However, in this situation, it is debatable that averaging
955: the pressure profile can be of any use. Visually, it is clear that
956: the upscaled pressure profile reproduces the fine scale pressure
957: profile with reasonable accuracy. The resistor renormalization, as
958: defined in Section \ref{sub:Analysis-of-heterogeneity}, produces
959: very unsatisfactory results. It must be noted that, even at the fine
960: scale, pressure in very nearly zero permeability areas is poorly
961: defined.
962: 
963: 
964: \begin{figure}[H]
965: \begin{center}\includegraphics[%
966:   scale=0.25]{sk32.eps}\includegraphics[%
967:   scale=0.25]{sp32.eps}\includegraphics[%
968:   scale=0.25]{spav.eps}\end{center}
969: 
970: \begin{center}\includegraphics[%
971:   scale=0.25]{sk16.eps}\includegraphics[%
972:   scale=0.25]{sp16.eps}\includegraphics[%
973:   scale=0.25]{serr.eps}\end{center}
974: 
975: \begin{center}\includegraphics[%
976:   scale=0.25]{sk16r.eps}\includegraphics[%
977:   scale=0.25]{sp16r.eps}\includegraphics[%
978:   scale=0.25]{serrr.eps}\end{center}
979: 
980: 
981: \caption{\label{cap:Shale error}Wavelet transform based real-space
982: renormalization of a permeability map with vertical shales from
983: $32\times32$ to $16\times16$. (\textbf{a1}) Fine scale permeability.
984: (\textbf{a2}) Fine scale pressure solution obtained from fine scale
985: permeability. (\textbf{a3}) Average of the fine scale pressure
986: solution ($2\times2$ cells averaged). (\textbf{b1})
987: Wavelet-renormalized coarse permeability. (\textbf{b2}) Coarse
988: pressure solution obtained from b1. (\textbf{b3}) Modulus of
989: relative error, $\left|\textrm{a3-b2}\right|$/a3. (\textbf{c1})
990: Resistor-renormalized coarse permeability. (\textbf{c2}) Coarse
991: pressure solution obtained from c1. (\textbf{c3}) Modulus of
992: relative error, $\left|\textrm{c3-c2}\right|$/c3. The relative
993: difference between the averaged fine scale pressure and the coarse
994: pressure from wavelet renormalization reaches 16\% with an average
995: of 6\%. Resistor renormalization clearly doesn't produce the
996: required result. The shale permeability is set to $10^{-13}$and also
997: the shales are distributed across the direction of flow, generating
998: a worst case scenario.}
999: \end{figure}
1000: 
1001: 
1002: \subsection{Three-dimensional systems}
1003: 
1004: As already mentioned, the wavelet renormalization method for
1005: up-scaling is easily extended to three-dimensional systems. In this
1006: case, no flow was assumed in two directions and pressures were
1007: specified on the boundaries of the third direction. The only
1008: difference in the procedure between two- and three-dimensional
1009: systems is the structure and size of the matrix $\mathbf{W}$. As for
1010: the two-dimensional case, by observing the structure of the
1011: transformed transmissibility matrix $\mathcal{T}$, a renormalization
1012: scheme can be devised to produce the coarse permeability avoiding
1013: the matrix multiplications. The algorithm substitutes cubes of
1014: linear size 4 cells by cubes of half the linear size. Preliminary
1015: runs confirm that the upscale procedure is approximately as accurate
1016: as it is in the two-dimensional case.
1017: 
1018: 
1019: \section{Conclusion}
1020: 
1021: An up-scaling method, based on the Haar wavelet transform and
1022: real-space renormalization was presented. Its advantages are speed,
1023: due to the underlying renormalization algorithm, and a rigorous
1024: mathematical derivation of the up-scaling rule.
1025: 
1026: This algorithm emerges from a mean-field picture of the solution to
1027: Darcy's equation, which is at the heart of the success of
1028: renormalization methods. The renormalization scheme is a consequence
1029: of the choice of $\mathbf{W}$ matrix, in this case the aim is to
1030: obtain the coarse permeability map that would generate the average
1031: pressure profile. A different matrix would lead to mathematically
1032: valid results, for example one where the block permeability is taken
1033: to be equal to the value at the top left cell in the group
1034: constituting the block. However, the present choice attempts to
1035: minimize the information loss inherent in the coarsening process
1036: while preserving the algorithm simplicity to ensure its efficiency.
1037: 
1038: Within this context, the lowest degree, mean-field approximation, in
1039: which all fluctuations are neglected, performs well in two and three
1040: dimensions. The main problems with this method are encountered when
1041: there is a high contrast in permeability, such as in the case of
1042: shales, which leads to sharp pressure changes that inevitably get
1043: smoothened out. The resistor renormalization fails even more
1044: drastically in this case. A different wavelet matrix choice would
1045: improve the performance of the method. It is nevertheless
1046: foreseeable that the emerging renormalization scheme would not be as
1047: easy to implement as the one presented. An exact solution could also
1048: be obtained, including all the fluctuation terms, however, the
1049: computational power required would be equivalent to performing the
1050: fine-scale solution.
1051: 
1052: At present, the method can be used as a fast upscaling technique
1053: able to cope with heterogeneities. The formalism introduced
1054: highlights how a very crude renormalization scheme is satisfactory
1055: in treating sufficiently homogeneous systems and how upscaling
1056: methods can be constructed to match the specifications of the
1057: problem and the required results. The resistor method is based on an
1058: analogy with current laws and is therefore a statement of
1059: conservation of flux. It is possible that by defining Darcy's
1060: equation in terms of fluxes rather than permeability and pressure,
1061: one might be able to find a matrix analogy to the resistor method
1062: that will reproduce the resistor upscaling rule in the same way as
1063: the current $\mathbf{W}$ matrix produces the renormalization scheme
1064: that was proposed. The present framework can be applied to other
1065: problems, such as advective transport, leading to insights into the
1066: general issue of how operators change as a consequence of
1067: coarsening.
1068: 
1069: It is hoped that further study will shed light on the effect of
1070: adding fluctuations to the mean-field approximation, allowing the
1071: choice between different degrees of accuracy depending on the
1072: available computational time.
1073: 
1074: 
1075: 
1076: \section{Acknowledgements}
1077: 
1078: V.P. gratefully acknowledges funding from the Department of Earth
1079: Science and Engineering, Imperial College London and the authors are
1080: thankful to the anonymous referees for very useful comments.
1081: 
1082: \begin{landscape}
1083: \section{Appendix}
1084: 
1085: In the following we reproduce the structure of the matrices
1086: discussed in the text. The structure of the transmissibility matrix
1087: for a $4\times4$ system:
1088: 
1089: \begin{flushleft}$\mathbf{T}=\left[\begin{array}{cccccccc}
1090: 2k_{1}+t_{1,2}+t_{1,5} & -t_{1,2} & 0 & 0 & -t_{1,5} & 0 & ... & 0\\
1091: -t_{2,1} & 2k_{2}+t_{2,3}+t_{2,5} & -t_{2,3} & 0 & 0 & -t_{2,5} & ... & 0\\
1092: ... & ... & ... & ... & ... & ... & ... & -t_{15,16}\\
1093: 0 & 0 & 0 & 0 & 0 & . & -t_{16,15} &
1094: 2k_{16}+t_{16,15}+t_{16,12}\end{array}\right]$\end{flushleft}
1095: 
1096: \noindent \begin{flushleft}The upper corner of the transformed
1097: matrix: $\mathbf{\mathcal{T}}=\mathbf{WTW^{T}}$\end{flushleft}
1098: 
1099: \noindent \begin{flushleft}\[
1100: \mathbf{\mathcal{T}}=\left[\begin{array}{cccc}
1101: k_{1}+k_{2}+\frac{t_{23}+t_{67}}{2}+\frac{t_{59}+t_{610}}{2} & -\frac{t_{23}+t_{67}}{2} & -\frac{t_{59}+t_{610}}{2} & 0\\
1102: -\frac{t_{23}+t_{67}}{2} & k_{3}+k_{4}+\frac{t_{23}+t_{67}}{2}+\frac{t_{711}+t_{812}}{2} & 0 & -\frac{t_{711}+t_{812}}{2}\\
1103: -\frac{t_{59}+t_{610}}{2} & 0 & k_{13}+k_{14}+\frac{t_{59}+t_{610}}{2}+\frac{t_{1011}+t_{1415}}{2} & -\frac{t_{1011}+t_{1415}}{2}\\
1104: 0 & -\frac{t_{711}+t_{812}}{2} & -\frac{t_{1011}+t_{1415}}{2} &
1105: k_{15}+k_{16}+\frac{t_{711}+t_{812}}{2}+\frac{t_{1011}+t_{1415}}{2}\end{array}\right]\]
1106: \end{flushleft}
1107: 
1108: The transmissibility matrix for a $2\times2$ system, the dash
1109: indicates that the properties refer to the $2\times2$ system :
1110: 
1111: \noindent \begin{flushleft}\[ \mathbf{T'}=\left[\begin{array}{cccc}
1112: 2k'_{1}+t'_{1,2}+t'_{1,3} & -t'_{1,2} & -t'_{1,3} & 0\\
1113: -t'_{2,1} & 2t'_{2}+t'_{2,1}+t'_{2,4} & 0 & -t'_{2,4}\\
1114: -t'_{1,3} & 0 & 2t'_{3}+t'_{3,1}+t'_{3,4} & -t'_{3,4}\\
1115: 0 & -t'_{2,4} & -t'_{3,4} &
1116: 2k'_{4}+t'_{2,4}+t'_{3,4}\end{array}\right]\]
1117: \end{flushleft}
1118: 
1119: Relationship between permeability and transmissibility in the
1120: upscaled system ($k'_{i}$, $t'_{ij}$) and in the fine scale system
1121: ($k{}_{i}$, $t{}_{ij}$):
1122: 
1123: \[
1124: k'_{1}=\frac{k_{1}+k_{2}}{2},\,\,\,\,
1125: t'_{12}=\frac{t_{23}+t_{67}}{2},\,\,\,\,
1126: t'_{13}=\frac{t_{59}+t_{610}}{2}\,\,\,\,\textrm{etc}.\]
1127: 
1128: 
1129: \end{landscape}
1130: 
1131: %\begin{thebibliography}
1132: 
1133: 
1134: \bibliographystyle{amsplain}
1135: \bibliography{biblio}
1136: 
1137: \end{document}
1138: