Anda di halaman 1dari 40

USMS

020156 A Review of Heterogeneity Measures Used in Reservoir


Characterization
L.W. Lake, U. of Texas; J.L. Jensen, Heriot-Watt
University

Copyright 1989 Society of Petroleum Engineers


This manuscript was provided to the Society of Petroleum Engineers for distribution
and possible publication in an SPE journal. The material is subject to correction
by the author(s). Permission to copy is restricted to an abstract of not more than
300 words. Write SPE Book Order Dept., Library Technician, P.O. Box 833836,
Richardson, TX 75083-3836 U.S.A. Telex 730989 SPEDAL.
1

A REVIEW OF HETEROGENEITY MEASURES USED


IN RESERVOIR CHARACTERIZATION

Larry W. Lake1 and Jerry L. Jensen2


1The University of Texas at Austin
2Heriott-Watt University
ABSTRACT

Heterogeneity, the spatial variation in properties, is a ubiquitous feature


in all naturally-oc~urring permeable media and one of the most important
factors governing fluid flow. This importance being recognized from the earliest
days of the petroleum industry, a number of measures have been proposed and
used to represent this heterogeneity.
The most pronounced form of heterogeneity involves permeability. This
paper reviews the means by which this heterogneity is assessed. The
measures fall into two categories: static and dynamic. Static measures do not

account for fluid flow directly; they are exemplified by the Dykstra-Parsons and

Lorenz coefficients, whose statistical properties are reviewed here. Dynamic

measures include dispersivity and chanelling factors which directly relate to the
efficiency of a miscible displacement. We elucidate the distinction between
these two phenomena and review the means by which they are quanitified.

INTRODUCTION

Reservoir heterogeneity has long been recognized as an important factor

in governing reservoir performance. The petroleum literature abounds with

both theoretical and field studies examining the impact of heterogeneity and

describing techniques to assess and mitigate its effects. Usually, the theoretical
SPE 2 01 56
2

stud~~~had two objectives. The first is to raise the reader's awareness of


the (usually adverse) effects of heterogeneity. The second is to provide some
technique for applying the results obtained to situations of immediate interest to
the reader, despite the fact that each reservoir is uniquely heterogeneous.
The uniqueness of each reservoir, however, does not necessarily
preclude the utility of either theoretical or field heterogeneity studies.
Depending upon the recovery process chosen, two reservoirs may still behave
in very similar fashions. The essence of any heterogeneity study of these
reservoirs, then, will be to identify the features which impact the performance
and define (preferably quantitatively) their levels. Such a procedure requires
taking samples (hypothetical or real), making measurements of the appropriate
properties, and distilling from the data the necessary information.
This paper reviews the variety of methods investigators have used to
distill and convey information regarding the degree of heterogeneity. We begin
by defining heterogeneity. A general structure for heterogeneity measurements
is then presented. Finally, we review the various measures used and relate
them to the general structure.

DEFINITION OF HETEROGENEITY

Consider a large-rate displacement through any porous, permeable


medium using matched mobility and density, chemically inert, miscible fluids.
Heterogeneity is the quality of the medium which causes the flood front--the
boundary between the displacing and displaced fluids--to spread as the
displacement proceeds. For a homogeneous medium, the rate of spreading is
zero. As the degree of heterogeneity increases, the amount of spreading
increases. As we shall see below, spreading can take place both globally and

locally
SPE 2 015 6
3

This definition is flow-based. It depends not only upon variations in


porosity and permeability, but also upon the spatial relationships of these
variations. It does not, however, include the influences of gravity, capillarity, or
viscosity, whose effects may be altered by heterogeneity but which are not
caused solely by the properties of the medium.

Structure of Heterogeneity Measures


Most' heterogeneity measures concentrate on the level of permeability
variation in a reservoir, variations in other properties, such as porosity, cation
exchange capacity, and amounts of clay minerals, being often ignored. This
practice is usually justified by two arguments. First, permeability is a property
which, by its very nature, has a direct impact upon flow. Consequently, it must
be included in any measure of heterogeneity. Second, permeability variations
are typically much larger than variations of other properties; hence, changes in
permeability can easily dominate the influence of variations in other properties.
Whatever the reservoir properties involved, heterogeneity measures can
be classified into one of three groups.
1. Static measures ignoring correlation. These measures are based on
measured samples (usually core plugs) taken from the reservoir. The
spatial relationships of the samples are ignored. The samples are
treated as independent and representative data all coming from one
population. Permeability and (exceptionally) porosity data are used in
the calculation.
2. Static measures using correlation. These measures are based on
measured samples and qualitative assessments of correlation. The
correlations are made on a reservoir scale using fence diagrams or sand
isopachs.
SPE 2 01 56
4

3. Dynamic measures. A flow experiment is an essential part of these

measures. Some feature of the displacement effluent is measured to

give an indication of the heterogeneity. Such measures implicitly

account for all variations in reservoir properties which impact the process

under test.

Each measure type has advantages and disadvantages relative to the

other types. For example, an advantage to dynamic measures is that, if the

process used during the flow experiment closely parallels the process which is

expected to be applied to the reservoir, the results are most directly applicable

with a minimum of interpretation. Disadvantages include the cost, complexity,

and the selection of "representative" elements of reservoir for conducting flow

experiments.

In the following, we examine several heterogeneity measures

categorized according to the above structure. Wherever possible, the variability

or precision of the measures will also be discussed. To emphasize that many

measures are estimates based on sample measurements, we will often denote

estimated quantities by placing a "hat" (A) over the symbol.

STATIC MEASURES IGNORING CORRELATION

We consider three heterogeneity measures: the coefficient of variation,

the Dykstra-Parsons coefficient, and the Lorenz coefficient. These measures

are frequently used because of the relatively low cost to obtain estimates and

their effectiveness in conveying heterogeneity effects in models. All three use

permeability data but porosity data may be included. Permeability is assumed

to be a scalar quantity for all the measures.


SPE 2 015 6
5

The Coefficient of Variation

The coefficient of variation, Cv, is defined to be

~Var(k)
C v = E(k)

where E(k) is the expectation of the permeability k and Var(k) is the

variance. Permeability is treated as a random variable in this equation. For the


case where permeability is log-normally distributed, ln(k)-N(J.L,cr2); that is, the

random variable ln(k) is normally distributed with mean Jl and variance cr2.

Because E(k) = exp(J.1+0.5cr2) and Var(k) = exp(2J.L+cr2) [exp(cr2)-1 ], we have

Cv = ~ exp(cr2)-1. ( 1)

Hence, Cv is independent of J.1 for log-normal distributions.

Cv is a measure of the variability relative to the mean value of the

permeability. In a homogeneous medium, Cv=O. In general, Cv is unbounded


but, for a set of n (2~n<oo) permeability data, Cv~1n-1 .1

Cv has not been used explicitly in reservoir modeling to measure

heterogeneity. It has been a model, however, in the development of

heterogeneity measures such as the Dykstra-Parsons coefficient and Polasek

and Hutchinson's heterogeneity factor.

The Dykstra-Parsons Coefficient

The most popular heterogeneity measure, the Dykstra-Parsons

coefficient, Vop, 4 is a measure based strictly on permeability variations. In their

original study, Dykstra and Parsons used minipermeameter measurements.4,


SPE 2 015 6
6

More commonly, however, core plug permeabilities are used. Exceptionally,


transient test data may be used. Vop has also been called the "coefficient of
permeability variation", "the variance", or "the variation".
The method used to estimate Vop, as described by Dykstra and Parsons,
involves ranking the permeability data in order of decreasing magnitude,
assigning probabilities to the data, and plotting the permeability and probability
values on log-normal probability paper (Fig. 1). A "best-fit" line to the points is
drawn and, if the points do not fall approximately on a straight line, more weight
is to be given to the central points than the points at the extremities. The line
represents an "equivalent" reservoir having a log-normal permeability
distribution. The line is used to determine the 84th percentile permeability, k0 .84
and the median permeability, k0.50 • The coefficient estimate then follows from

" ko.so - ko.S4 (2)


Vop = ko.so

This procedure, while calling for a log-normal probability plot to be made,


does not require that the permeabilities be log-normally distributed. Willhite
points out that pessimistic results from calculations based on the equivalent
reservoir may occur because the larger permeability layers would dominate the
predicted behavior. The misfit in behavior is also influenced by the subjective
element in drawing the "best-fit" line. Jensen and Lake demonstrate this
nonuniqueness, showing how several "equivalent" layered reservoirs
1\
(equivalent in the sense of having the same Vop) may have quite different
behavior. The non-quantitative nature of determ~ning the line also precludes
. 1\
any assessment of the statistical variability of Vop.
SPE 2 01 56
7

The above procedure assumes that each permeability datum represents


the same quantity (e.g., thickness or volume) of formation. This assumption is
required during the probability assignments. This assumption can be partly
verified in the case of core plug permeabilities by examining the choice of
coring points. In the case of transient test and minipermeameter data, the
volume of investigation can vary substantially between the measurements and
this should be accounted for in the probability weights assigned to the data.
A
Since 0 ~ ko.a4 ~ ko.so, then 0 ~ Vop ~ 1. The lower values (0 to 0.5)
represent cases of low heterogeneity, while the higher values (0. 7 to 1.0) reflect
reservoirs with large to extremely large levels of heterogeneity. Most reservoirs
have 0.5 ~ Vop ~ 0.9.8 Depending upon the nature of the permeability data, Vop
may be a measure of vertical heterogeneity or lateral and vertical heterogeneity.
A
For example, if Vop is obtained using horizontal core plug permeability data, it
will primarily reflect the variability in the horizontal permeability with depth.

Since it is common for depthwise permeability variations to be much larger than


A
areal variations in reservoirs, Vop may be a pessimistic measure of the overall

permeability heterogeneity. In a study of 22 fields, Lambert found that only


three had Dykstra-Parsons coefficients based on areal variation exceeding the
Vop based on depth variation and none of these situations were statistically
significant at the 2°/o level. This behavior reflects on the probability weighting
given to the data; weights obtained from cross-sections of strata do not
necessarily correspond to weights obtained from the strata volumes.
Craig2 points out that Eq. (2) should have the logarithms of the quantities

in the definition

V _ log(ko.so) - log(ko.a4)
- log (ko.so) ·
SPE 2 015 6
8

This indeed could be the case if the permeability distribution is log-normal and,

for some reason, we wanted to· have a measure of the variability in the log

transformed domain. In terms of the variation in the untransformed domain--

which is what fluid flow responds to--the quantity V is probably misleading for

two reasons. First, as Eq. (1) shows, the coefficient of variation strictly

depends upon the difference log(ko.so) - log(ko.a4) i.e., the ratio ko.so/ko.a4 with no

explicit dependence upon the median. V can change without Vop changing and

vice versa. Second, the form of Eq. (2) where, in the numerator, percentiles are

subtracted to give a measure of the variability and, in the denominator,

the median is used as a measure of central location is consistent with the

notion of "robust" statistics. From this point of view, averages and standard

deviations are less helpful statistics than percentiles and differences of

percentiles because the latter are less subject to influence by erroneous data.

After describing the estimation procedure, Dykstra and Parsons went

on to correlate values of Vop with oil recovery from core-scale water floods.

Since their study, several reports11,12 have appeared examining and discussing

the merits of the correlation. Other studies11-16 have used the coefficient in

non-water flood applications. A significant feature of many studies2.4,12,14 is the

low sensitivity of models to variations in Vop when Vop ~ 0.5 while, for the large

heterogeneity cases, the models exhibit a large sensitivity. This behavior

reflects, in part, the nature of Vop which is to force all situations into a finite

range, 0 to 1, with the large heterogeneity situations being particularly poorly

discerned.
SPE 2 01 56
9

Modifications to the Dykstra-Parsons Procedure

Since Dykstra and Parson's study, several authors have introduced

subtle but significant changes to the procedure for Vop estimation while still
calling the result "the Dykstra-Parsons coefficient" or referencing Dykstra and
Parson's work. For instance, two authors17,18 do not mention the relative

weighting of the central and extreme portions of the plot to determine the "best-

fit" line. This omission increases the number of possible "equivalent" reservoirs.
Another, more significant change is that several descriptions12,15,19,2o do not

mention the "best-fit" line at all. The quantities ko.so and k0 . 84 are estimated

directly from the data. This redefinition eliminates the subjectivity of the line

fitting and permits quantitative analysis of Vop. Nonetheless, the estimates


obtained using the "best-fit" line procedure may be different from those obtained
using the percentiles based directly on the data.9
For the case ln(k)-N(J.L,cr2), ko.so=exp(J.L) and k0 . 84=exp(J.L-cr) so that
Vop=1-exp(-cr). Warren and Cosgrove12 point out that

Vop = 1 - exp[-~ln(kA/kH)] (3)

where kA and kH are the arithmetic and harmonic averages of the permeability

data. Eq. (3) follows from two features of the log-normal distribution:
E (k)=exp(J.1+0.5cr2) and E(k-1 )=exp( -J.L+0.5cr2). 3

The log-normal assumption also permits straight-forward computation of


A
the sampling bias and variability of V DP which arises because a limited set of n

data is being used in the estimation procedure. Jensen and Lake9 show that

mv = -0.7 49•cr2•exp( -cr)/n

and
SPE 20156
10

sv = 1.49•a•exp(-a)tfr1,

A
where mv is the bias and sv is the standard error of V DP· The bias is usually
quite small but, nonetheless, causes the estimates, on average, to
underestimate the true Vop. The situation is more significant in the case of the
standard error. For example, if Vop=0.60, then to have a two-in-three chance of
A
0.55::::;Vop::::;0.65 requires that 120 samples be used in the estimation.
A
When the permeability is not log-normally distributed, V DP values
obtained from estimates of ko.so and ko.s4 not longer bear a relationship to Vop
values obtained using either the "best-fit" line or assuming the permeabilities
are log-normally distributed .. Jensen and Currie21 establish some relationships
assuming a family of forms for the permeability distribution.
Lambert1 o also investigated the effects of including porosity, <f>, in the
A A
computation of Vop. She compared Vop obtained using permeability only with
A
V oP based on the ratio kl<f>. The ratio kl<f> reflects the interstitial velocity of a

single-phase fluid in the medium.22 Lambert found no significant difference

between the two Dykstra-Parsons estimates, emphasizing the relative


constancy of porosity compared to permeability.

Heterogeneity Measures Related to Vop


An extensive body of work exists relating Vop to reservoir performance. It
is, however, an imperfect measure. One way in which it could be improved is to
extend the scaling such that the large heterogeneity cases are more easily
discerned. Claridge23 suggested a measure VC' where

V ko.1s - ko.so
c- ko.so · (4)
SPE 2 015 6
11

He assumes ln(k)""N(J.L, a2) so that Eq. (4) may be expressed as

V c. ko.so - ko.s4
ko.s4

Vop and Vc are related by Vop = Vc/(1 +VcY. While Q$;Vop$;1, Q$;Vc$;oo so that

definition of the large heterogeneity cases is no longer constrained by the finite

limit. Claridge23 states that V c "gives a somewhat more linear relationship with
actual permeability variation." He could also have considered using a instead

of Vc since Vc=exp(a)-1, so that, considering the Taylor series expansion of

exp(•), they are equal to the first order. From the same type of analysis used by
Jensen and Lake,9 the bias and standard error properties of 'fJ c, when estimated
A A A
using the percentile estimates ko.so and ko.s4, are identical to those for Vop.
Jensen and Lake9 proposed a pair of heterogeneity measures, U and p,

where

U _ ko.1s- ko.s4
- ko.s4

and p represents a measure of the permeability distribution asymmetry. Like


Claridge's measure, O~U$;oo, however the median permeability has been

replaced by the 16th percentile, k 0 . 16 . By using a more extreme percentile


A A
estimate, U has moderately better error properties than Vop. To define a given
A
level of heterogeneity with any specified error, n points are required for Vop and
A
0.43n points are needed for U. Having two measure statistics complicates

modeling and comparisons of reservoir performance, but it begins to address

the nonuniqueness problem of Vop.


SPE 2 015 6
12

Jensen and Currie21 approach the problem of heterogeneity assessment

in a different way. Rather than propose a new measure, they suggest a more

efficient method of estimating V0 p. They use the asymmetry factor p along with
" which have 50°/o less
a maximum likelihood approach to obtain estimates Vop

statistical error than the traditional method. This means that only one-quarter of

the data required for a Vop estimate based on percentile estimates is needed to

" DP with the proposed method.


optai n V

The Lorenz Coefficient

Another measure of heterogeneity, suggested by Schmalz and Rahme,24

is the Lorenz coefficient. The Lorenz curve (Fig. 2) is a plot of cumulative flow

capacity, Fm, versus cumulative thickness, Hm, where

(5a)

and

(5b)

for a reservoir of n layers. The layers are arranged in order of decreasing

permeability so that k(1) is the layer with thickness h(1) and the largest

permeability while k(n) is the layer with thickness h(n) and the smallest

permeability. By definition, 0;S;Fm;S;1 and Q;S;Hm;S;1 for 1;S;m;S;n. Because of the


SPE 2 015 6
13

layer ordering, the Lorenz curve monotonically increases from m=1 to m=n with

a monotonically decreasing slope.


A
The Lorenz coefficient, Lc, is given by twice the area (the shaded region

in Fig. 2) between the Lorenz curve ABC and the diagonal AC. If the medium is

homogeneous, all the permeability values are identical and the Lorenz curve is
A
the straight line AC. Hence, Lc=O. Increasing levels of heterogeneity are

indicated by movement of the Lorenz curve, ABC, away from the diagonal AC
A A
with Lc increasing but always less than unity. Typical values for Lc for

reservoirs are in the range of 0.3 to 0.6.1 o

Equations (5) are discrete versions of a slightly more elaborate notion.25

If p(k) is the permeability probability density function,

00

F(x) = E~k) J kp{k)dk (6a)


X

and
00

H(x) = J p(k)dk. (6b)


X

x is a parameter which varies from 0 to oo along the Lorenz curve. For the

Lorenz curve to exist, E(k)<oo, which may not always be the case.9

The Lorenz coefficient is defined to be

00 k1
Lc = 1 - E~k) J p(k1) J k2p(k2)dk2dk1
0 0

and O~Lc~1. While the Lorenz curve is uniquely defined (to within a scale

factor) for any given distribution,2s Lc is not uniquely associated with any
SPE 2 015 6
14

particular permeability distribution.9 For the case of ln(k)-N(J.l,cr2),

Lc=erf(0.5cr)=erf[-0.5•1n(1-Vop)], where erf(•) is the error function.26

As was the case for the Dykstra-Parsons plot, there is an inherent issue

of probability associated with the Lorenz plot. Comparing Eqs. (5) and (6), we

observe that the probability of occurrence of a permeability value in the

reservoir is associated with a height of an interval. Depending on the volume

represented by each of the n intervals, the height may not accurately reflect the

frequency of occurrence of its respective interval.

Schmalz and Rahme24 related Lc to waterflood performance using both

the Stiles27 and the Suder and Calhoun2B models. Schauer,29 in a field study of

the performance of nine waterfloods, presented data relating Lc to the injection

required for the first production increase. Lake et al.3o used Lc when they

considered the impact of heterogeneity upon chemical flood recovery efficiency,

but they did not explicitly include it in their model.

Modifications to the Lorenz Procedure

The Lorenz procedure can be modified by including porosity in the


calculation.2,10,22,3o In place of the cumulative thickness, Hm, the cumulative

storage capacity, Cm, is used:

i=m
L<t>(i).h(i)
i=1
Cm=-.- - -
l=n
L<t>(i).h(i)
i=1
SPE 2 01 56
15

If porosity is constant, the Lorenz curve remains unaltered. The data must be
ordered according to the ratio kl<j>. The inclusion of porosity variations will not
A
substantially increase Lc.

Estimation of the Lorenz Coefficient


Estimation of Lc requires evaluating the area between the Lorenz curve
and the diagonal. This is easily done by computer using a simple algorithm
such as the. trapezoidal rule or Simpson's rule. Another approach is to use the
relationship between Lc and Gini's coefficient of concentration, G.1 In the
Appendix, we show that

(7)

is equivalent to using the trapezoidal rule for the area. Note that Eq. (7)
requires no ordering of the data and assumes all k's have equal probability.
An analytic assessment of the sampling properties of Lc, such as that
done for \top when permeability is log-normally distributed, is not simple; the
results can only be expressed in terms of integrals which must be numerically
evaluated. Jensen and Lake9 show the results of a Monte Carlo analysis,
assuming permeability to be log-normally distributed, for bias and standard
error. Lc is substantially negatively biased (greater than 5°/o) for small sample
sizes (n<40) and heterogeneous distributions (Lc>0.6). Thus, uncompensated
heterogeneity assessments using Lc will understate the heterogeneity in the
reservoir. It is, however, more precise than the Dykstra-Parsons estimator.
SPE 2 015 6
16

STATIC MEASURES USING CORRELATION


These measures, which all use reservoir correlation on the interwell
scale, are based on static data at the wellbore. Their main use is to make
inferences about the reservoir properties in the interwell region. The measures
aim to better convey the impact of the reservoir structure upon reservoir
performance. While moderately successful in accomplishing this aim, their
evaluation requires much more effort than the preceding measures.

Polasek and Hutchinson's Heterogeneity Factor


Polasek and Hutchinson 5 reported on a study covering 31 sedimentary
formation outcrops. 31 The first portion concentrated on the permeability
variations within the clean sands. Similar to the approach of Dykstra and
Parsons, 4 they generated log-normal probability plots and measured the slope
of a best-fit line to give an indication of the permeability variation. Although they
do not state how they calculated the slope, it appears from their data that the
slope is given by the difference between log1o(ko.1o) and log1o(ko.9s). Hence,
1\
their slope is approximately -1.21n(1-Vop). They observed that, while it was
useful to study the within-unit variations, those variations could be the least
significant with regard to large-scale fluid movement. The sand analysis
conveyed little information concerning the relative positions of the sand and
shale which could dominate the performance.
To give a quantitative measure of the amount and location of shaly
material, Polasek and Hutchinson proposed the heterogeneity factor, HF. It is
based on a layered model and comprises the following steps.
IPE 2 015 6
17

1. Establish a field-wide correlation from the logs and other available data.

Treat distinct intervals as separate units; otherwise, break a section down

into layers, each of about 5 feet (1.5m) in thickness.

2. At each well and for each unit or layer, estimate the net-to-gross ratio, nij,

where i is the well number (i=1, 2, ... , q wells) and j is the layer or unit

number U=1, 2, ... , nunits and/or layers).


3. Estimate the average net-to-gross ratio, Nj, and the standard deviation of

the net-to-gross ratio, Sj, for the jth unit or layer:

s.2J --q-1.
-
1 t
1=1
(n··IJ -N·)2
J

4. H F i• the heterogeneity factor for the jth unit or layer, is given by the

coefficient of variation of the net-to-gross ratio,


HFJ·-~
-N·· J

5. HF for each continuous deposit is obtained by averaging, on a volume

basis if possible, the HF's for all layers and/or units within the deposit.

A perfectly layered system will have HF=O. A heterogeneous, unlayered system


will have HF~-{(t1, since, as previously discussed, this is an upper limit on the

coefficient of variation.

Polasek and Hutchinson also listed the HF values for nine reservoirs and

their qualitative recovery behavior. While the correlation was generally good,
SPE 2 0 l!) 6
18

the authors point out that process parameters (e.g., mobility ratio) also may
have influenced the reservoir performances listed. They also report, for five
watered-out reservoirs, the mobility ratio, HF, and bypass factor. Again, the
correlation between HF and bypass factor was moderately good.

Alpay's Sand Index


Alpay32 proposed dividing a reservoir into "bands of genetic similarity". A
interval or zone may consist of several bands. For the reservoir under
consideration (the Cardium sands in the Pembina field), Alpay found that the
gamma ray log reflected the shale/sand proportions well. Thus, he developed a
sand index, Sl, based on the gamma ray measurement for each interval. 81=1-
lsh. where Ish is the shaliness index33 found in well log interpretation. Sl varies

from 0 to 1, with 0 representing little permeable sand, while 1 represents


virtually all "flow conductive" material.
Alpay32 made contour maps of Sl and the number of bands for each
interval. These represented, respectively, the variations in amount of
permeable sar1d and the stratigraphic complexity of the interval. He ·also
considered measures of Sl variability (the standard deviation and kurtosis) on a
well-by-well basis.
No results are given by Alpay32 to relate reservoir performance to either
stratigraphic complexity or Sl. He does state, however, that such an approach
has given insight into reservoir performance and variations between different
portions of fields.

Other Measures Using Correlation


Pirson17 briefly discusses three measures: a coefficient of stratification
(<;): a degree of lensing or lenticularity (A.); and a coefficient of thinning (t). <;=1
SPE 2 015 6
19

reflects prefect correlation of a reservoir property between two wells while <;=0

represents no continuity. A. is a direction dependent property; if A.=O for two

wells, the variability of the property is about the same at the two wells, whereas
A.;eO represents a change in the variability. t is the ratio of gross pay between

two wells. Pirson makes some general remarks about the relationship of these

three measures to reservoir performance, but does not give any specific

examples.

DYNAMIC MEASURES OF HETEROGENEITY

Dynamic measures are those which relate directly to or are derived

directly from fluid flow. These measures are more complicated than static

measures because they also depend on the properties of the fluids comprising

the flow and because they depend on the spatial arrangement or

autocorrelation of the heterogeneity. To lessen the fluid property dependence,

we consider, as before, only completely miscible fluids here, an assumption

which obviates dealing with the troublesome topics of wettability and local

surface characterization.

Channeling and Mixing

Even restricted to miscible fluids, heterogeneity manifests itself in flow

patterns in two basic ways. Figure 3 shows the two-dimensional spatial

arrangement of a large number of fluid elements (particles) which were

simultaneously released into a flowing stream at the inlet end of the medium

when t=O. The particle positions are summarized through isoconcentration

lines, contours of equal number of fluid particles at a particular instant t. (Of

course, no two particles can occupy the same location and isoconcentration

representations therefore imply a degree of averaging corresponding to the


SP.E 2 015 6
20

representative elementary volume 34 in a permeable medium.) In general, the


isoconcentration lines exhibit gross distortions about an average (x-fixed)
vertical position (channeling) and also deviations in position about a (y-fixed)
lateral position (mixing).
If we average the concentrations in Fig. 3 laterally the concentration
profile in the lower portion results. The variance of the particle positions in this
frame of reference represents the combined effects of channeling and mixing.
On the other hand, the lateral average of the variance in particle positions at a
fixed y, as in the upper portion of Fig. 3, gives an estimate of the local mixing.
The channeling+mixing or large-scale variance is aE and the mixing or small-

scale variance is a~ The difference between these two (aE-a~), the channeling

variance, is a measure of the tendency of the fluid flow to channel through the
medium. Subtracting variances like this implies that channeling and mixing
occur independently. Whether such independence, in fact, exists is unknown;
however, indirect evidence regarding the growth rate of unstable miscible
displacements 35 suggest that there is some degree of dependency. This
difference, which is always non-negative, approaches zero as the
isoconcentration lines become vertical and parallel. On the other hand, a~

approaches zero in the limit of all the particles remaining on the same
isoconcentration line during the flow.
The physical quantity describing the mixing is the dispersion coefficient
which, above some relatively low velocity, is proportional to the local fluid
velocity. The constant of proportionality is the dispersivity, the basic dynamic
measure of heterogeneity which describes mixing. A displacement dominated
by this type of mixing exhibits good lateral sweep efficiency and a variance
which grows asymptotically in proportion to (travel distance)112 or t1/2.
SPE 2 015 6
21

The channeling variance is generally represented in flow equations as a


non-linear relation between fluid flux and concentration. Historically, the
relation has no universal form, but many of the most useful models incorporate
heterogeneity factors which describe the local channeling. Unfortunately, this
description of channeling is nearly always combined with a description of
hydrodynamic channeling--viscous or gravity fingering. Such instabilities
certainly do enhance the propensity for channeling, but they add additional
complexity to the heterogeneity measures and they are not necessary for
channeling to occur. A displacement dominated by channeling exhibits little
mixing, poor lateral sweep efficiency and a variance which grows asymptotically
in proportion to travel distance or time35.

Autocorrelation
Autocorrelation is a statistical measure of the likelihood of two quantities
spatially separated by a distance h having the same value. There are a large
number of empirical functions describing autocorrelation. The impact of
autocorrelation functions on dynamic heterogeneity measures is discussed at
some length in Gelhar and Axness36 and Arya37. Correlation models range
from simple exponentials having only a single scale, through models having
multiple scales38 up to quite complex representations which have a continuum
of correlation scales39 Unfortunately, much of this complexity is actually
observed in the distribution of permeability in realistic media. 4 0
Autocorrelation is important because it determines whether the fluid flow
is mixing or channeling dominated. Flow in media where autocorrelation is
nonexistent or sm·all (compared to the medium length) tends to exhibit mixing 14.
Such media are called random or uncorrelated. The absence of channeling in
these cases is evident even at large heterogeneity levels and modest degrees
SPE 2 015 6
22

of instability 41 ,42 except when there is a large amount of mixing. When the

autocorrelation length is even a modest fraction of the medium length, the fluid

flow will tend to channel. Media with very long autocorrelation lengths are

strictly layered or uniformly stratified.


Of course, it is impossible to separate the effects of mixing and

channeling without at least two-dimensional observations of the fluid flow, a

feature which has been impossible until the advent of imaging. Thus,

researchers have tended to explain aE either solely as a mixing or a channeling


phenomenon.

Dispersivity

Hydrologists generally use dispersion models to explain the large-scale

variance. The literature relating dispersivity to variances and correlation in fluid


properties is massive 14,36,3?, 43-45 and a detailed review is beyond the scope of

this paper. The most recent contribution by Dagan 45 shows how the tensorial

form of dispersion coefficient can be related to the tensorial form of the

autocovariance function. In one-dimensional, isotropic random fields, the


dispersivity a is related to the coefficient of variation of the velocity field Cv and
its autocorrelation length Q by.14

(3)

In this equation and all other versions of such statistical transport properties,

the property is related to the statistics of the velocity field, not to those of the

underlying permeability field. A general relation between permeability statistics


SPE 2 015 6
23

and velocity statistics does not exist, being influenced by flow geometry,
boundary conditions and even the type of calculation method employed.45.46

Channeling Factors
Reservoir engineers, on the other hand, have tended to represent a~
through channeling factors. Such factors have been intimately associated with
the propagation of instabilities47-53 and most workers have not attempted to
separately account for heterogeneity. One of the few which does, however, is
the heterogeneity factor HK in the work of Koval. 47 HK is defined as the pore
volumes of fluid required to bring the effluent concentratio~ of a stable, miscible
displacement up to 98°/o of the injected concentration. This is about the most
direct measure of fluid flow performance that a heterogeneity measure can have
because 1/HK now represents the pore volumes of fluid injected before
breakthrough. HK ranges between 1.5 and 5 in laboratory cores 47 but can be
much larger for field-scale displacements. For uniformly layered media HK is
empirically related to V DP by 13

(8)

See also reference 23. Such a relationship has not been established for other
types of heterogeneity. HK, as expressed through Eq. (8), does an excellent job
of "straightening out" the nonlinearity in V 0 p described previously. For
example, changing V 0 p from 0.2 to 0.3 changes HK (and therefore the
breakthrough and sweepout times) relatively little; however, changing V 0 p from
0.7 to 0.8, the normal range from routine core data, changes HK substantially.
SPE 2 015 6
24

Scaling
Unfortunately, the dynamic measures do not scale well. Scaling is the
process of calibrating a measure against laboratory data and then using the so-
calibrated model in predictions. The reasons behind the lack of scaling are
many-fold: sensitivity to fluid properties, scale-dependence of media
correlation, and lack of a sound theoretical base for these measures.
We can, however, tie all of these factors together through the notion of
flow regimes, illustrated schematically in Fig. 4. The large-scale variance of a
displacement in a realistic medium grows neither strictly linear with distance nor
with the square root of distance. Instead, there may be portions which are of
one type of the other separated by transitions regions. The change in the rate
of growth may be caused by a number of factors: progression to larger scales of
heterogeneity, stabilization of instabilities or the initiation of instabilities from
large-scale perturbations. The actual at in Fig. 4 grows initially with
(distance) 112 and then changes to linear growth.
Scaling entails matching at at some small scale (laboratory in Fig. 4)
and then using this match to prediction of at a large scale (field). Using a
mixing model, where the variance grows as (distance)1/2 throughout, yields a
prediction of aE which is too small. To match aE at the field scales requires an
increase in the apparent dispersivity and, indeed, experience has shown that a
does appear to increase with length of measurement scale.1 4 On the other
hand, scaling with a channeling model will overpredict the field-scale variance.
Experience has also shown that such models must be attenuated in order to
match field results. 49

Neither the increase in dispersivity nor the decrease in channeling factor


is desirable for prediction, particularly since both are supposed to be intrinsic
properties of the medium. The solution to the dilemma is (a) defining time-
SPE 2 015 6
25

dependent material properties which will cause the variance to grow in

whatever manner is desired or (b) to used both the channeling and mixing

models simultaneously. The latter solution is by far the better solution since

· time-dependent properties are tricky to handle in multiwell simulations.

SUMMARY

Flow through permeable media, even restricted to stable miscible

displacements, is too complex to be represented completely with simple

heterogeneity indicators. Nevertheless, such indicators do indicate many

features of such flows and may provide a basis for classification. Past

experience on the last point has not been encouraging; however, such

experience has generally been devoid of geological insight. Future work on

heterogeneity indicators will provide this insight by closer association with

geology and by relying more heavily on ideas related to spatial correlation.

ACKNOWLEDGEMENTS

Portions of this work were supported by U.S. Department of Energy

Contract No. 19-89BC14251.000, Annex II; the State of Texas Advanced

Technology Program; and the Center for Enhanced Oil and Gas Recovery

Research. Larry W. Lake holds the Shell Distinguished Chair in Petroleum

Engineering.

NOMENCLATURE

l
Cm = Cumulative storage capacity

Cv = coefficient of variation

E(•) = Expectation operator

erf(•) = Error function


SPE 20156
26

F(•) = Incomplete first moment

Fm = Cumulative flow capacity

G = Gini's coefficient of concentration

H(•) = Incomplete zeroth moment

HF = Polasek and Hutchinson's


heterogeneity factor

HK = Koval's heterogeneity factor

Hm = Cumulative thickness

k = permeability

K = sum of permeabilities

Lc = Lorenz heterogeneity measure

n = number of data

N = average net-to-gross ratio


N(J.t,a2) = Normally distributed variable with mean Jl and variance a2

nij = net-to-g ross ratio for ith well and jth layer

p = measure of permeability distribution asymmetry

q = number of wells

n = number of units or layers at the ith well

s = standard deviation of the net-to-g ross ratio

Sl = Alpay's sand index

t = Pirson's coefficient of thinning or time

u = Jensen and Lake's heterogeneity measure

v = Craig's coefficient of variation

Vc = Claridge's heterogeneity measure

Vop = Dykstra and Parson's heterogeneity measure

Var(•) = Variance operator

X = dummy variable
SEE 2 015 6
27

a. = dispersivity

<t> = porosity

Ish = shaliness index


"A = Pirson's degree of lensing
c; = Pirson's coefficient of stratification
Q = correlation length

<J'L = large-scale variance

as = small scale variance

Super- or sub-scripts

A = arithmetic average
H = harmonic average
(i) = ith element in an ordered set
0.16 = 16th percentile
0.50 = 50th percentile (median)
0.84 = 84th percentile
{\
= estimated quantity

APPENDIX A

We derive here the relationship between Gini's coefficient of

concentration, G, and the trapezoidal rule for estimating Lc. We assume that
each datum, ki, has probability~ and is ordered such that k 1 ~k 2 ~····~kn. From

Fig. 5, Lc=2x(area A) or 2x(area A + area B) - 1. Hence, from the trapezoidal


rule
SPE 2 015 6
28

n
where ko = 0 and K = L',ki. The following steps then yield the desired result.
i=1

k =2~Ki r±(ki-kj) + ±(ki+kj)+ i(krkil- i(kj+ki)J;


j=1 i=1 i=1 i=j i=j

1 n
2nKI, [
j=1
n

i=j
:»i -
.
Jkj];

and, since the last two of the three right-side terms are each zero. we obtain

The right-side of this expression is G. 1

REFERENCES

1. Kendall. M .• and Stuart. A.• The Advanced Theory of Statistics. MacMillan

Pub. Co .• New York (1977),1. 48-50.


SPE 2 015 6
29

2. Craig, F. F., The Reservoir Aspects of Waterflooding. SPE Monograph


Series, SPE, Richardson, Texas (1971 ).
3. Johnson, N. L., and Katz, S., Continuous Univariate Distributions -1,
John Wiley and Sons, New York (1970), 115-118.
4. Dykstra, H. and Parsons, R. L., .. The Prediction of Oil Recovery by Water
Flood," in Secondary Recovery of Oil jn the United States, American
Petroleum Institute, New York (1950), 160-174.
5. Polasek, T. L., and Hutchinson, C. A., .. Characterization of Non-
Uniformities with a Sandstone Reservoir from a Fluid Mechanics
Standpoint," Proc. of Seventh World Petroleum Congress, Elsevier Pub.
Co. Ltd. (1967), 2, 397-407.
6. Goggin, D. J., Thrasher, R., and Lake, L. W., "A Theoretical and
Experimental Analysis of Minipermeameter Response Including Gas
Slippage and High-Velocity Flow Effects," In Situ, 12 (1988), nos. 1 and
2, 79-116.
7. Warren, J. E., Skiba, F. F., and Price, H. S.,. "An Evaluation of the
Significance of Permeability Measurements," J. Pet. Tech. (Aug., 1961 ),
739-744.
8 Willhite, G. Paul, Waterflooding, SPE Textbook, Dallas (1986).
9. Jensen, J. L., and Lake, L. W., "The Influence of Sample Size and
Permeability Distribution on Heterogeneity Measures," SPE Reservoir
Engineering (May, 1988), 629-638.
10. Lambert, M. E., "A Statistical Study of Reservoir Heterogeneity," MS
Thesis, The University of Texas at Austin, 1981.
11. Mobarak, S., "Waterflooding Performance Using Dykstra-Parsons as
Compared With Numerical Model Performance," J. Pet. Tech. (Jan.,
SP.E 2 015 6
30

1975) 113-115 and Discussions Journal of Petroleum Technology (Aug.,


1975; Dec., 1975; Jan., 1976; and June, 1976).
12. Warren, J. E., and Cosgrove, J. J., "Prediction of Waterflood Behavior in a
Stratified System," Soc. Pet. Eng. J. (June, 1964), 149-157.
13. Paul, G. W., Lake, L. W., Pope, G. A., and Young, G. B.:, "A Simplified
Predictive Model for Micellar-Polymer Flooding," SPE 10733 presented
at the 1982 California Regional Meeting, San Francisco, March 24-26.
14. Arya, A., Hewett, T. A., Larson, R., and Lake, Larry W., "Dispersion and
Reservoir Heterogeneity," SPE Reservoir Engineering (Feb., 1988), 139-
148.
15. Craig, F. F., "Effect of Reservoir Description on Performance Predictions,"
J. Pet. Tech., (Oct., 1970), 1239-1245.
16. Orr, F. M., and Sageev, A., "Reservoir Characterization for the C0 2

Enhanced Oil Recovery Process," annual report, Contract No. AC21-


85MC22042, U. S. DOE (March, 1987).
17. Pirson, S. J., Oil Reservoir Engjneerjng, McGraw-Hill Book Co., New York
(1958), 88-92.
18. Smith, C. R., Mechanics of Secondary Oil Recovery, Robert E. Krieger
Pub. Co., Malabar (1975), 227-228.
19. Arps, J. J., "Estimation of Primary Oil and Gas Reserves," in Petroleum
Production Handbook, T. C. Frick (ed.), SPE of AIME, Dallas (1962) 2,
37-29.
20. Mungan, N., "Enhanced Oil Recovery Using Water as a Driving Fluid,"
World Oil (Sept. 1981 ),160.
21. Jensen, J. L., and Currie, I. D., "Improving Performance Prediction by
More Accurate Heterogeneity Assessment," SPE/DOE 17364 presented
SP.E 2 015 6
31

at the SPE/DOE Enhanced Oil Recovery Symposium held in Tulsa,


Oklahoma, April 17-20, 1988.
22. Lake, Larry W., Enhanced Oil Recovery, Prentice-Hall, Inc., New York
(1988), 195-196.
23. Claridge, E. L., "Control of Viscous Fingering in Enhanced Oil Recovery
Processes: Effect of Heterogeneities," SPE 7662, May, 1978.
24. Schmalz, J. P., and Rahme, H. S., "The Variation in Water Flood
Performance with Variation in Permeability Profile," Producers Monthly
(July, 1950), 9-12.
25. Dagum, C., "Lorenz Curve," in Encyclopedia of Statistical Sciences by S.
Katz and N. L. Johnson (eds), 5, 156-161.
26. Aitchison, J., and Brown, J. A. C., The Lognormal Ojstrjbutjon,
Cambridge University Press, Cambridge (1957), 112.
27. Stiles, W. E., "Use of Permeability Distribution in Water Flood
Calculations," Trans. AIME (1949), 186,9-13.
28. Suder, F. E., and Calhoun, J. C., "Waterflood Calculations," Drilling and
Production Practice, American Petroleum Institute (1949), 260-270.
29. Schauer, P. E., "Application of Empirical Data in Forecasting Water Flood
Behavior," SPE 934G presented at the 32nd Annual Fall meeting, Dallas,
Texas, October, 6-9,1957.
30. Lake, L. W., Stock, L. G., and Lawson, J. B., "Screening Estimation of
Recovery Efficiency and Chemical Requirements for Chemical Flooding,"
SPE 7069 presented at the Fifth Symposium on Improved Methods for
Oil Recovery held in Tulsa, Oklahoma, April 16-19, 1978.
31. Stalkup, F. 1., "Permeability Variations Observed at the Faces of
Crossbedded Sandstone Outcrops," in Reservoir Characterization by
SPE 2 015 6
32

Larry W. Lake and H. B. Carroll (eds), Academic Press, Inc., New York
(1986), 141-175.

32. AI pay, 0. A., "A Practical Approach to Defining Reservoir Heterogeneity,"


J. Pet. Tech. (July, 1972), 841-848.
33. Dewan, J. T., Essentials of Modern Open-Hole Log Interpretation,
Pennwell Pub. Co., Tulsa (1983), 248-249.
34. Bear, J., Dynamics of Fluids in Porous Media, American Elsevier
Publishing Co. Inc., New York City (1977).
35. Waggoner, John and Lake, Larry W., "A Detailed Look at Simple Viscous
Fingering," presented at the II Simposio lnternacional sabre
Recuperacion Mejorada de Crude Trabajos Tecnicos, 24 al 27 de
Febrero de 1987, Maracaibo, Venezuela.
36. Gelhar, L. W. and Axness, C. L., "Three-Dimensional Stochastic Analysis
of Macrodispersion in Aquifers," Water Resour. Res., 19, 161-180, 1983.
37. Arya, A., "Dispersion and Reservoir Heterogeneity," Ph.D. Dissertation,
The University of Texas at Austin, 1986.
38. Burrough, P. A., "Multiscale Sources of Spatial Variation in Soil. I. The
Application of Fractal Concepts to Nested Levels of Soil Variation," J.
Soil Science, 34, (1983), 577-597.
39. Hewett, T.A., "Fractal Distributions of Reservoir Heterogeneity and Their
Influence on Fluid Transport," SPE 15386 presented at the 61 st Annual ·
Technical Conference, New Orleans, October 5-8, 1986.
40. Goggin, David J., "Geologically-Sensible Modelling of the Spatial
Distribution of Permeability in Eolian Deposits: Page Sandstone
(Jurassic), Northern Arizona," Ph.D. Dissertation, The University of Texas
at Austin, 1988.
SPE 2 015 6
33

41. Waggoner, John R., Ph.D. Dissertation, The University of Texas at Austin,
in progress 1989.
42. Araktinga, U. G. and Orr, F. M. Jr., "Viscous Fingering in Heterogeneous
Porous Media," SPE 18095, presented at the 63rd Annual Technical
Conference and Exhibition in Houston, Texas, October 2-5, 1988.
43. Dagan, G., "Stochastic Modeling of Groundwater by Unconditional and
Conditional Probabilities, 2. The Solute Transport," Water Resour. Res.,
18, 835-848, 1982
44. Gelhar, L. W. and Axness, C. L., "Stochastic Analysis of Macrodispersion
in a Stratified Aquifer," Water Resour. Res., 15, 1387-1397.
45. Dagan, Gideon, "Time-Dependent Macrodispersion of Solute Transport
in Anisotropic Heterogeneous Aquifers, Water Resour. Res., 24, no. 9, pp.
1491-1500.
46. Yang, An-Ping, Ph.D. Dissertation, The University of Texas at Austin, in
progress 1989.
47. Koval, E. J., "A Method for Predicting the Performance of Unstable
Miscible Displacement in Heterogeneous Media," Soc. Pet. Eng. J., 3
(June 1963}, 145-154.
48. Dougherty, E. L., "Mathematical Model of an Unstable Miscible
Displacement in Heterogeneous Media," Soc. Pet. Eng. J., 3 (June
1963}, 155-163.
49. Todd, M. R. and Longstaff,W. J., "The Development, Testing, and
Application of a Numerical Simulator for Predicting Miscible Flood
Performance," J. Pet. Tech., (July 1972}, 874-882.
50. Fayers, F. J., "An Approximate Model with Physically Interpretable
Parameters for Representing Viscous Fingering," SPE Reservoir
Engineering (May 1988), 551-558.
SPE 2 015 6
34

51. Fishlock, T. P. and Rodwell, W. R., "Improvements in the Numerical


Simulation of Carbon Dioxide Displacement," presented at the Second
European Symposium on Enhanced Oil Recovery, ARTEP, Paris,
France, Nov. 8-10, 1982,317.
52. Fayers, F. J. and Newley, T. M. J., "Detailed Validation of an Empirical
Model for Fingering with Gravity Effects," SPE Reservoir Engineering
(May 1988), 542-550.
53. Young, L. C., "The Use of Dispersion Relationship to Model Adverse
Mobility Ratio Miscible Displacements," SPE 14899, presented at the 5th
joint SPE/DOE symposium on Enhanced Oil Recovery, Tulsa, Oklahoma
(1986).
SEE 2 015 6

"Best-fit" line

0.02 0.16 0.50 0.84 0.98

Probability Scale

Figure 1. The Dykstra-Parsons coefficient calculation procedure.


SP.E 2 015 6

1.0

0.0
0.0 1.0

Figure 2. Flow capacity-storage capacity curves and the Lorenz coefficient.


SP.E 2 015 6

Mixing:
Average cr8

C(x,y,t)

lsoconcentrat1on

Mixing+
Channeling

C(x,t)

Figure 3. Schematic illustration of the variances of channeling and mixing.


SPE 20156

Channeling
(-t')

\
Q)
()
c:
<U
·.:::::
<U
>Q) '-Mixing
<U
()
(-t 1/2)
CJ)
I
Q)
0>
L.
<U
_J

Lab Field
Scale Scale

Distance

Figure 4. Schematic illustration of scaling with dynamic heterogeneity measures.


SPE 2 015 6

1.0

Lorenz curve " " '

I Area B I

0.0
0.0 H1 1.0

Figure 5. Gini's approximation to the Lorenz coefficient.

Anda mungkin juga menyukai