Sunteți pe pagina 1din 18

SPE 107954

Improved Permeability Prediction Relations for Low Permeability Sands


F.A. Florence, SPE, Texas A&M U., J.A. Rushing, SPE, Anadarko Petroleum Corp., K.E. Newsham, SPE, Apache Corp.,
and T.A. Blasingame, SPE, Texas A&M U.
Copyright 2007, Society of Petroleum Engineers
This paper was prepared for presentation at the 2007 SPE Rocky Mountain Oil & Gas
Technology Symposium held in Denver, Colorado, U.S.A., 1618 April 2007.
This paper was selected for presentation by an SPE Program Committee following review of
information contained in an abstract submitted by the author(s). Contents of the paper, as
presented, have not been reviewed by the Society of Petroleum Engineers and are subject to
correction by the author(s). The material, as presented, does not necessarily reflect any
position of the Society of Petroleum Engineers, its officers, or members. Papers presented at
SPE meetings are subject to publication review by Editorial Committees of the Society of
Petroleum Engineers. Electronic reproduction, distribution, or storage of any part of this paper
for commercial purposes without the written consent of the Society of Petroleum Engineers is
prohibited. Permission to reproduce in print is restricted to an abstract of not more than
300 words; illustrations may not be copied. The abstract must contain conspicuous
acknowledgment of where and by whom the paper was presented. Write Librarian, SPE, P.O.
Box 833836, Richardson, TX 75083-3836, U.S.A., fax 01-972-952-9435.

Abstract
This work addresses the problem of estimating Klinkenbergcorrected permeability from single-point, steady-state
measurements on samples from low permeability sands. The
"original" problem of predicting the corrected or "liquid
equivalent" permeability (i.e., referred to as the Klinkenbergcorrected permeability) has been under investigation since the
early 1940s in particular, using the application of "gas
slippage" theory to petrophysics by Klinkenberg.1
In the first part of our work, the applicability of the JonesOwens4 and Sampath-Keighin5 correlations for estimating the
Klinkenberg-corrected (absolute) permeability from singlepoint, steady-state measurements is investigated. We also
provide an update to these correlations using modern petrophysical data.
In the second part of our work, we propose and validate a new
"microflow" model for the evaluation of an equivalent liquid
permeability from gas flow measurements. This work is based
on a more detailed application of similar concepts employed
by Klinkenberg. In fact, we can obtain the Klinkenberg result
as an approximate form of our result. Our theoretical "microflow" result is given as a rational polynomial in terms of the
Knudsen number (the ratio of the mean free path of the gas
molecules to the characteristic flow length (typically the
radius of the capillary)).
The following contributions are derived from this work:
Validation and extension of the correlations proposed by JonesOwens and Sampath-Keighin for low permeability samples.
Development and validation of a new "microflow" model which
correctly represents gas flow in low permeability core samples.
This model is also applied as a correlation for prediction of the

equivalent liquid permeability in much the same fashion as the


Klinkenberg model, although our new model is substantially
more theoretical (and robust) as compared to the Klinkenberg
correction model.

Introduction
The gas slippage phenomenon typically occurs in the laboratory when gas flow experiments are conducted at low
pressures. Gas slippage is defined as the condition where the
mean free path of the gas molecules is no longer negligible
compared to the average effective rock pore throat radius
i.e., the gas molecules tend to "slip" on the surfaces of the
porous media. This effect yields an overestimation of the
measured gas permeability compared to the true absolute
permeability if it were measured using a liquid.
For flow in tubes, the gas slippage phenomenon has been
investigated since the end of the nineteenth century. The first
study of gas slippage in porous media was conducted by
Klinkenberg.1 The Klinkenberg model approximates a linear
relationship between the measured gas permeability and the
reciprocal absolute mean core pressure.2 This model has been
a consistent basis for the development of methods computing
the absolute liquid permeability of a core sample based on a
single data point i.e., single-point steady-state permeability
measurement methods.
Subsequent work focused on correlating the parameters of the
Klinkenberg model (i.e, the Klinkenberg-corrected permeability or equivalent liquid permeability (k) and the Klinkenberg
gas slippage factor (bK). Heid et al3 and Jones and Owens4
proposed two correlations similar in form between bK and k,
while Sampath and Keighin5 proposed a different form of
correlation using effective porosity () as a third parameter.
At a more theoretical level, the gas slippage phenomenon is
part of a larger area of study the theory of rarefied gases.
This theory has developed substantially in the last fifty years,
leading to a better understanding and more accurate modeling
of the gas slippage effect. The primary objectives of this work
are:
To compare and evaluate existing correlations for single-point,
steady-state measurement of permeability, and
To develop the concept of a new model for gas slippage in low
permeability sands.

F.A. Florence, J.A. Rushing, K.E. Newsham, and T.A. Blasingame

Overview of Existing Correlations


The "Klinkenberg" model was developed to account for the
discrepancies between permeabilities measured with gases and
liquids as flowing fluids. This model is based on the argument
developed by Kundt and Warburg6 that, at low pressures, or
more specifically, when the mean free path of the gas
molecules is within two orders of magnitude of the capillary
radius, the velocity of the gas molecules at the surface of the
porous medium is non-zero.
Klinkenberg proved that there exists an approximately linear
relationship between the measured gas permeability (ka) and
the reciprocal mean pressure ( 1 / p ), which is given as:
bK
k a k
1 ................................................................ (1)
p

Where bK is the "gas slippage factor" which is a constant that


relates the mean free path ( ) of the molecules at the mean
pressure ( p ) and the effective pore radius (r).
The bK-term is defined by:
bK 4c

, where c 1 .................................................... (2)


p
r

In Eq. 1, k is the "equivalent liquid permeability" (which is


also called Klinkenberg-corrected permeability). Klinkenberg
has shown that k is the true, absolute permeability of the
sample. Derivations of the Klinkenberg relations are given in
Appendix A.
Taking the Klinkenberg equation as fact, later work in this
area focused on the determination of the gas slippage factor.
In the late 1940s, Heid et al3 conducted a study including the
effects of pressure, pore size, and the type of porous medium
on permeability. Heid et al determined the gas slippage factor
(bK) and the (extrapolated) permeability (k) of 11 synthetic
cores and 164 natural core samples of representative sands
from different producing areas of the United States.

SPE 107594

Additional research in the 1970s focused on the gas slippage


effect because of the emergence of "unconventional gas
resources" as an important energy source. In particular, much
of the laboratory research in the 1970s/early 1980s addressed
permeability measurements in tight gas sands which typically exhibit permeabilities much lower than that of the core
samples Heid et al. studied.
As a case in point, Jones and Owens4 studied the effects of gas
slippage for more than one hundred samples from tight gas
sands. Although the database used by Jones and Owens is not
available, they did provide a correlation similar in form to the
one presented by Heid et al. for the relationship between the
gas slippage factor (bK) and the equivalent liquid permeability
(k). The Jones-Owens formula also confirms that the gas
slippage effect is more significant for lower permeability
rocks. The Jones-Owen result is given as:
bK 12.639(k ) 0.33 ......................................................... (4)

In 1981 (two years after the Jones and Owens work), Sampath
and Keighin5 studied 10 core samples from a tight gas sand
field in Uinta County, Utah, and published a formula relating
the gas slippage factor (bK) to the ratio of Klinkenbergcorrected permeability to effective porosity (i.e., k/). The
Sampath-Keighin correlation is given by:
0.53

k
bK 13.851

...................................................... (5)

The data and the correlation relation developed by SampathKeighin are shown in Fig. 2.

Based on the results of their core sample experiments shown


in Fig. 1, Heid et al3 developed a relationship between the gas
slippage factor (bK) and the corresponding equivalent liquid
permeability (k) which is written as:

bK 11.419(k ) 0.39 ........................................................ (3)

Fig. 2

Fig. 1

Heid et al data and correlation (ref. 3).

Sampath-Keighin data and correlation (ref. 5).

Although based on a reduced number of samples, the Sampath-Keighin correlation (Eq. 5) is interesting at a theoretical
level since it is noted in the literature7 that the square root of
the Klinkenberg-corrected permeability/porosity ratio is considered to be a "characteristic length" (recall that permeability
has the dimension of length-squared). Eq. 2 shows that the

SPE 107954

Improved Permeability Prediction Relations for Low Permeability Sands

Klinkenberg gas slippage factor, bK, is inversely proportional


to the capillary radius where the capillary radius is related
to the square root of k/by:
r 8.886 10 6 k / ...................................................... (6)

Eq. 6 is derived analytically in Appendix B we note that r


is the capillary radius (expressed in centimeters) and k is the
Klinkenberg-corrected permeability (in md). Assuming standard conditions, it is possible to derive a general but rigorous
"square-root" k/correlation of the form:
0.5

k
bK

....................................................................... (7)

The "-term" in Eq. 7 is a parameter that depends on the type


of gas used in the core flow experiment we present Eq. 7
computed for different gases on various data correlation plots
later in this paper.
Core Data Used in Current Research
The core data used in our work consists of data gathered in the
literature and from industry sources. The "literature database"
is comprised of data from Heid et al and Sampath and
Keighin. The "industry database" is composed of two datasets
obtained from the Lower Cotton Valley sandstones in North
Louisiana.8,9 The first "Cotton Valley" dataset includes 12
core samples; the second has 18 core samples. We note that
both datasets utilized multi-point, steady-state permeability
measurements. For each sample, we have effective porosity as
well as several gas permeabilities (ka) measured at various
mean core pressures ( p ). Klinkenberg-corrected permeability
(k) and gas slippage factor (bK) are estimated using the
classical Klinkenberg plotting method.
We also obtained "older" datasets from a Gas Research
Institute (GRI) project conducted in the 1980s and early 1990s

Fig. 3

in the Travis Peak (East Texas) and Frontier (Wyoming)


formations (refs. 10-12). We should note that all of the GRI
datasets were derived from unsteady-state permeability
measurements; however, the new correlations and models
derived in our work are based on the multi-point, steady-state
data. We have included the unsteady-state data for (visual)
reference only the data are not used in the correlations.
Comparison of Core Data and Correlations
Fig. 3 presents a comparison of available core data with the
Jones and Owens and Heid et al correlations in the bK versus
k format. In Fig. 4 we present a comparison of the available
data with the Sampath-Keighin correlation and our generalized
square-root model given by Eq. 7 (bK versus k/format plot).
Figs. 3 and 4 serve to highlight trends in the data as well as
to establish existence of a significant vertical spread in the
data sets which can lead to errors in permeability estimates
for low permeability reservoirs.
Next, we compare the Jones-Owens correlation, the SampathKeighin correlation, and our new proposed square-root
correlation with the data from the Lower Cotton Valley
formation (or modern "standard" database). The comparison
of the Heid et al and Jones-Owens correlations with the
(Cotton Valley) steady-state data is shown on Fig 5 (bK vs.
k). The Sampath-Keighin correlation and our proposed
"square-root" correlation are compared to the steady state data
in Fig 6 (bK vs. k/).
For both of the Cotton Valley datasets, the steady-state
permeability measurements were performed using nitrogen as
the flowing fluid and we note that the nitrogen curve on
Fig. 6 overestimates the gas slippage factor (bK), whereas the
carbon dioxide curve provides a good fit. Regardless of the
"intercept" issue, our model matches the data trends well.

Comparison of the "bK vs. k" correlations by Heid et al (ref. 3) and Jones-Owen (ref. 4) with various
field and literature data, acquired with unsteady-state (USS) or steady-state (SS) techniques.

F.A. Florence, J.A. Rushing, K.E. Newsham, and T.A. Blasingame

Fig. 4

Fig. 5

SPE 107594

Comparison of the "bK vs. k/" correlations by Sampath and Keighin (ref. 5) and this work with various field
and literature data, acquired with unsteady-state (USS) or steady-state (SS) techniques.

Comparison of the steady-state data with the JonesOwens and Heid et al correlations. The Heid et al
correlation seems well adapted for these datasets,
whereas the Jones-Owens correlation generally
underestimates the gas slippage factor, especially for
the first Cotton Valley sample set.

Fig. 6

Comparison of the Cotton Valley steady-state data with


the Sampath-Keighin and proposed square-root correlations although measurements were performed with
nitrogen as the flowing gas, we note the square-root
correlation for carbon dioxide gives the best match (the
best fit is actually obtained for a -value of 34).

SPE 107954

Improved Permeability Prediction Relations for Low Permeability Sands

In Figs 7 and 8 we present the computed equivalent liquid


permeability plotted against the extrapolated Klinkenbergcorrected permeability on log-log scales for both "Cotton
Valley" datasets. We should note that the "Klinkenbergcorrected permeability" refers only to the value of permeability actually extrapolated from a Klinkenberg plot (k measured),
while the permeability values computed from correlations (k
computed) are referred as "equivalent liquid permeabilities."
However, we will use the same symbol (k) for either case.

Fig. 7

Fig. 8

Computed equivalent liquid permeability versus


extrapolated Klinkenberg-corrected permeability, for the
Lower Cotton Valley Sample No. 1 the Jones-Owens
and Sampath-Keighin correlations overestimate k and
the theoretical square-root model for nitrogen underestimates k.

Computed equivalent liquid permeability versus


extrapolated Klinkenberg-corrected permeability, for the
Lower Cotton Valley Sample No. 2 the Jones-Owens
and Sampath-Keighin correlations provide fair results
for k > 1 d. The theoretical square-root model for
nitrogen underestimates k, but performs well for low
permeability samples.

In Figs. 9 and 10 we present the average absolute relative


errors for the computed equivalent liquid permeability plotted
against the extrapolated Klinkenberg-corrected permeability.
This format helps us to assess the accuracy of the correlation
on a "point-by-point" basis for a particular dataset.
For Cotton Valley Sample No. 1, the Jones-Owens and Sampath-Keighin models generally overestimate the permeability.
In both cases (Cotton Valley samples 1 and 2), the theoretical
square-root model (nitrogen curve) underestimates the
Klinkenberg-corrected permeability but the overall performance of the square-root relation is comparable to the other
methods.

Fig. 9

Average absolute relative errors for equivalent liquid


permeability versus measured Klinkenberg-corrected
permeability, for Lower Cotton Valley Sample No. 1.
Errors are generally less than 35 percent.

Fig. 10 Average absolute relative errors for equivalent liquid


permeability versus measured Klinkenberg-corrected
permeability, for Lower Cotton Valley Sample No. 2.
Errors are generally less than 30 percent, although there
are numerous outliers which exhibit greater than 50
percent error.

As shown in Figs. 9 and 10, the three models for k exhibit a


fairly high absolute relative error (>50 percent for some cases)
however; the "clustering" for both datasets suggests that

F.A. Florence, J.A. Rushing, K.E. Newsham, and T.A. Blasingame

SPE 107594

most samples exhibit less than 30 percent error. Given the


data and the relatively approximate nature of the correlations,
we consider the performance to be good, perhaps very good.
While it is difficult to discriminate the "best" of the three
correlations based on our results, we believe the SampathKeighin correlation and the generalized square-root
correlations are more "consistent" with the theory and as
such, should be favored for this application.
Similar to previous work, the objective of our research is to
estimate the bK-term accurately, and then utilize this estimate
in Eq. 1 to yield the Klinkenberg-corrected permeability. A
general procedure for "single-point data" is to use any given
pair of ka and values to estimate bK, and then to use the bK
and p values to estimate k. In our case, we will use a model
(i.e., correlation) for bK and substitute this result into Eq. 1 to
estimate k.
As an alternative, one could simply establish a correlation of
k = f(ka) this process was proposed by Jones and Owens
based on the gas permeability (ka) estimated at 100 psig. This
correlation is given as:

k 10( 0.398 log ka 1.067 log ka 0.0825) , for 0.1 d k a 1 md


........................................................................................... (8)
The purpose of our work is to establish a theoretically robust
correlation of k f ( p ,,..., k a ) so we will not pursue a
result of the form of Eq. 8.

Fig. 11 Klinkenberg-corrected (extrapolated) permeability (k)


versus measured gas permeability (ka).

Klinkenberg Model Revisited


We begin our discussion of the Klinkenberg concept by first
returning to the fundamentals we will utilize the same
measurements as with the correlations presented in the
previous section (i.e., ka, p , , and backpressure (if used))
but we now consider the rigorous mechanisms of flow at the
microscopic (and smaller) scales. As a first step, we accept
the Klinkenberg approximation as a "first order" type of
estimate Eq. 1, repeated below for clarity:
bK
k a k
1 ................................................................ (1)
p

We "define" k as the true, equivalent liquid permeability of


the system, but we will also use estimates of k estimated
using the Klinkenberg correction as a standard to correlate
against. That is, we will use the Klinkenberg-corrected permeability (k) as reported (or calculated) from our data
sources as the reference permeability. Obviously, we would
prefer to use a more rigorous estimate, but for the present
study, we must prove the concept model (i.e., the "microflow"
model presented in this section) against some standard and
we believe that the Klinkenberg-corrected estimates are
appropriate for that purpose.
In Fig. 11, we present the Klinkenberg-corrected (equivalent
liquid) permeability (k) plotted against the measured gas
permeability (ka) on log-log coordinates as a "correlation" to
ensure that, at least directionally, Eq. 1 is valid.

As a primer for the development of our microflow model, we


consider the phenomenon of gas slippage at a fundamental
level as a phenomenon which occurs as a subset of a much
larger area of study known as the theory of rarefied gases.
With the development of the common interest for aeronautics
and aerospace, as well as more recently, the development of
MEMS (Micro Electro-Mechanical Systems), the flow of
gases at very low pressure or, more aptly, the flow of gas
molecules with a high mean free path (the mean free path
being inversely proportional to the pressure) has been
thoroughly studied in the last half century.
In the pursuit of a new microflow model, it then seems natural
to revisit the Klinkenberg model, if for no other reason than to
confirm that our new microflow model reverts to the
Klinkenberg model under certain conditions (which it does).
The flow regime for a gas flowing in a micro-channel is
typically determined by the value of the Knudsen number (Kn)
which is defined as:13

Kn
......................................................................... (9)
lchar

In Eq. 9, is the mean free path of the gas molecules (i.e., the
average distance (length) between two consecutive molecular
interactions) and lchar is the characteristic length of the flow
geometry (e.g., channel height, pipe radius).
For our purposes, the Knudsen number is difficult to define
rigorously for a porous medium but for the sake of
argument, we presume that some "characteristic length" can be
estimated for a porous medium. More generally, we assume
that we can define the Knudsen number based on the properties of the porous medium.

SPE 107954

Improved Permeability Prediction Relations for Low Permeability Sands

The classical definition of the mean free path from thermodynamics is:
( p, T ) / 2

1
RT

p
M

[ ( p, T )] ......................... (10)

where:
p = mean pressure (absolute)
T = absolute temperature
= viscosity of the gas at T and p
R = universal gas constant
M = molecular weight of the gas
This particular definition (Eq. 10) is based on the kinetic
theory for a perfect gas.14
In addition to the definitions above, we also must consider
various other flow regimes as follows (see Fig. 12):
Continuum Flow Regime: For Kn < 0.01, the mean free
path of the gas molecules is negligible compared to the
characteristic dimension of the flow geometry (i.e., the lchar
-parameter). In this case the continuum hypothesis of fluid
mechanics is applicable (i.e., the system is described by the
Navier-Stokes equations).
"Slip-Flow" Regime: For 0.01 < Kn < 0.1, the mean free
path is no longer negligible, and the slippage phenomenon
appears in the "Knudsen" layer (layer of gas molecules
immediately adjacent to the wall).
"Transition" Regime: For 0.1 < Kn < 10.
Free Molecular Flow Regime: For Kn > 10, the flow is
dominated by diffusive effects.

The volumetric gas flowrate (q) flowing through a capillary of


radius r and length L is given by:13
r 4
4 Kn p
q
(1 ( Kn) Kn)
1
........................... (11)
8
1 Kn L

The (Kn)-term is defined by Karniadakis and Beskok13 as:

128
( Kn)
tan -1 4 Kn 0.4 ............................................. (12)
152

Using the same capillary model as Klinkenberg,1 we derived


the following "microflow" model which relates the measured
permeability to gas (ka), the equivalent liquid permeability
(k), and the Knudsen number (Kn):
4 Kn
k a k [1 ( Kn) Kn]
1
..................................... (13)
1 Kn

Eq. 13 is rigorously derived and should be valid for all low


pressure/low velocity flow regimes which exist for gas flow in
porous media. We have not investigated the application of Eq.
13 for high pressure and/or high velocity flow.
Correlation of the Knudsen Number with Porosity,
Permeability and Mean Pressure
The Knudsen number (Kn), unlike the mean pressure ( p ),
cannot be measured by direct laboratory measurements. Our
objective is to design a model for estimating the equivalent
liquid permeability (k) using single-point, steady-state measurements (i.e., using a single pair of ka and p values). To
remedy this issue, we propose a "pseudo" Knudsen number
(Knp) which is defined as a function of the "typically"
measured parameters (i.e., p , , ka).
Assuming that the core flow experiments are conducted at
isothermal conditions (i.e. there is no temperature gradient in
the core) and that variations in viscosity are negligible over
the range of pressures considered, then the Knudsen number
(defined by Eq. 9) is inversely proportional to both the mean
pressure ( p ) and the characteristic flow geometry length
(lchar). Typically the pore throat radius (defined by Eq. 6 as a
function of the equivalent liquid permeability (k) and the
porosity ()) will be used to estimate lchar.
Substituting Eq. 12 into Eq. 13 yields

128
4 Kn p
k a k
1
tan -1 4 Kn p 0.4 Kn p
1

15

1 Kn p

........................................................................................ (14)
Fig. 12

Limits of the different flow regimes, as a function of


characteristic length lchar, and reciprocal mean free
path normalized (1 atm, 300K). Lines defining various
flow regimes are based on flow of air at isothermal
conditions (Modified from ref. 13).

Karniadakis and Beskok13 developed a unified model for gas


flow in micro-tubes (see Appendix C) where this model is
valid over the entire range of flow regimes.

where Knp is the "pseudo" Knudsen number. In order to


utilize Eq. 14, we must assume that the mean pressures,
porosities, and the gas and equivalent liquid permeabilities are
available. To solve for the "pseudo" Knudsen number (Knp)
for a particular case, we then rearrange Eq. 14 into the
following "root solution" form:

128
4 Kn p
k a k
1
tan -1 4 Kn p 0.4 Kn p
1
0
2
15

1 Kn p

........................................................................................ (15)

F.A. Florence, J.A. Rushing, K.E. Newsham, and T.A. Blasingame

Applying the Pseudo-Knudsen Number Concept


To demonstrate an application of the "pseudo" Knudsen
number (Knp), we only use the Lower Cotton Valley data (No.
2) (ref. 9), from which we derive two separate correlations for
Knp (see Appendix D). Our first correlation relates Knp to the
reciprocal mean pressure, the porosity, and the gas permeability as follows:
Kn p 0.4

1 0.598 0.352
ka

............................................ (16)
p

As implied, Eq. 16 was derived for the specific case of the


Cotton Valley No. 2 data, in a process where Knp (Eq. 15) was
estimated using known data for ka, p , and k. After Knp was
obtained as the root of Eq. 15, we then correlated these Knp
values with the ka, p , and data to yield Eq. 16.

SPE 107594

Substituting Eq. 18 into Eq. 15 and rearranging yields:


0.4

1 a1 a2
ka
128
a1 a2
-1 1

1 2 tan 4
a0 k
a0 k

k 15
p
p

1 a1 a2
a0 k
1 1/

............................................................................................ (19)

At this point, Eq. 16 can be substituted into Eq. 15 and this


result can be solved directly for the equivalent liquid
permeability (k). Obviously, k data must be available to
"calibrate" Eq. 15 (i.e., to estimate Knp). However, we believe
that the calibration of Eq. 16 may (in future work) be reduced
to the determination of the intercept term (the ka exponent is
approximately -0.5 and the exponent may be imposed as 0.5
similar to the "square-root" correlations).
As an alternative, we can follow the exact same calibration
procedure, but this time substitute k for ka in Eq. 16 (i.e., the
Knp = f(ka, p ,) correlation). For this case, we obtain:
Kn p 2.62

1 0.553 0.3897
k

.......................................... (17)
p

The equivalent liquid permeability values computed using Eq.


15 (based on the values of Knp from Eq. 16 or 17) are
correlated with the "given" Klinkenberg-corrected permeability data for this case in Fig. 13.

Fig. 13

Computed (Klinkenberg) equivalent liquid permeability


(k) versus Klinkenberg-corrected (extrapolated)
permeability (k). (pseudo-Knudsen number approach,
implicit and explicit relations for Knp)

We recognize that Klinkenberg-corrected permeability data


may not always be available in practice and the process of
solving simultaneously for the pseudo-Knudsen number (Knp)
and the equivalent liquid permeability (k) will require an
implicit formulation (an analog in the field of phase behavior
would be an equation-of-state (EOS) which is implicit in fluid
density).

The "trick" is to obtain the "calibration" coefficients (a0, a1,


and a2) in Eq. 19. As noted earlier, Eq. 19 can be formulated
as an "equation-of-state" and the coefficients (a0, a1, and a2)
can be tuned using non-linear regression provided that
there are sufficient data measurements, and that such measurements are taken for samples from the same depositional
sequence.

In such a case, (i.e., an implicit relation) we would have to


determine coefficients for the Knp = f(k, p , ) simultaneously with the estimation of k using Eq. 15. Writing the
correlation for the pseudo-Knudsen number in general form,
we have:
1
Kn p a0 k a1a2 ........................................................ (18)
p

We must note that we do not in any manner propose that Eq.


18 is "universal" (i.e., one set of coefficients (a0, a1, and a2)
does not apply to all possible data cases, Eq. 18 must be
calibrated for each dataset). For this case, we did formulate
Eq. 19 as a "fully implicit" solution using non-linear regression and we achieved following results:
a0 = 0.93
a1 = -0.49
a2 = 0.13
The results of our "fully implicit" correlation are shown in
Fig. 14.

where Eq. 18 provides a "correlation" for the variables in this


problem and this relation allows us, in theory, to solve Eq.
15 for the equivalent liquid permeability (k) in an implicit
fashion.

SPE 107954

Improved Permeability Prediction Relations for Low Permeability Sands

will provide significant improvement for the estimation of


equivalent liquid permeability from typical steady-state (gas)
permeability measurements.
Nomenclature
Variables
b
bK
c
Kn
Knp
ka
k

lchar

=
=
=
=
=
=
=
=
=

p
p
q
r
Fig. 14

Computed (Klinkenberg) equivalent liquid permeability


(k) versus Klinkenberg-corrected (extrapolated)
permeability (k). (fully implicit formulation for the
pseudo-Knudsen number).

Summary and Conclusions


Summary
The development of the microflow model for permeability
prediction in low permeability gas sands offers an alternative
to the classical Klinkenberg model, which is the long-time
reference for measuring permeability. Although the Klinkenberg model is the universally accepted standard, we believe it
may yield poor results for very or ultra low permeability core
samples.
Our new model has been validated using field data from a low
permeability reservoir case, but this "proof" is by no means
exhaustive additional validation is warranted.
Conclusions
1. The Jones-Owens, Sampath-Keighin, and our "Square-Root"
correlations may be used satisfactorily for single point, steadystate measurements as mechanisms to estimate the equivalent
liquid permeability. The Sampath-Keighin and Square-Root
correlations should be preferred based on the theoretical
formulations for these models.
2. The microflow model presented in this work is promising
as it provides a second-order correction for gas slippage
(beyond the "first-order" Klinkenberg formulation. The use
of such a correction should be especially relevant for low to
ultra low permeability core samples.

Recommendations/Comment
We note that the theoretical square-root correlation (Appendix
B) has a limited accuracy in this work. However, the squareroot model validates use of the permeability/porosity ratio in
modeling the gas slippage factor and serves to connect the
classical Klinkenberg model with our new "microflow model."
Upon further validation, we believe that our microflow model

=
=
=
=
=
=

General slip coefficient


Klinkenberg gas slippage factor, psi
Proportionality constant
Knudsen number, dimensionless
Pseudo-Knudsen number, dimensionless
Apparent gas permeability, md
Equivalent liquid permeability or Klinkenbergcorrected permeability, md
Characteristic length of the flow geometry, cm
Mean core pressure, psia
Differential pressure, psi
Gas flowrate, cc/sec
Effective pore radius, cm
Rarefaction coefficient parameter, dimensionless
Gas viscosity, cp.
Mean free path of the gas molecules, cm
Porosity, fraction

References
1. Klinkenberg, L.J.: "The Permeability of Porous Media to Liquid
and Gases," paper presented at the API 11th Mid Year Meeting,
Tulsa, Oklahoma (May 1941); in API Drilling and Production
Practice (1941), 200-213
2. RP40, Recommended Practices for Core Analysis, 2nd edition,
API, Washington, DC (1998).
3. Heid, J.G., et al: "Study of the Permeability of Rocks to
Homogeneous Fluids," in API Drilling and Production Practice
(1950), 230-246
4. Jones, F.O. and Owens, W.W.: "A laboratory Study of Low
Permeability Gas Sands," paper SPE 7551 presented at the 1979
SPE Symposium on Low-Permeability Gas Reservoirs, May 2022, 1979, Denver, Colorado.
5. Sampath, K. and Keighin, C.W.: "Factors Affecting Gas Slippage
in Tight Sandstones," paper SPE 9872 presented at the 1981
SPE/DOE Low Permeability Symposium, May 27-29, 1981,
Denver, Colorado.
6. Kundt, A. and Warburg, E.: "ber Reibung und Wrmelei-tung
verdnnter Gase," Poggendorfs Annalen der Physik und Chemie
(1875), 155, 337.
7. Jones, S.C.: "Using the Inertial Coefficient, , to Characterize
Heterogeneity in Reservoir Rock," paper SPE 16949 presented at
the 1987 SPE Annual Technical Conference and Exhibition,
September 27-30, 1987, Dallas, Texas
8. Lower Cotton Valley Formation Core Report Anadarko
Petroleum Corp., (2005)
9. Lower Cotton Valley Formation Core Report Anadarko
Petroleum Corp., (2006)
10. Travis Peak Formation Core Report Well Howell No. 5, S.A.
Holditch, (1986)
11. Travis Peak Formation Core Report Well S.F.E. No. 2, S.A.
Holditch, (1987).
12. Frontier Formation Core Report Well S.F.E. 4-24, S.A.
Holditch, (1991).

10

F.A. Florence, J.A. Rushing, K.E. Newsham, and T.A. Blasingame

13. Karniadakis, G.E. and Beskok, A.: Micro-flows, Fundamentals


and Simulation, Springer-Verlag, New-York (2002).
14. Loeb, L.B.: The Kinetic Theory of Gases, second edition,
McGraw-Hill Co. Inc., New York City (1934).
15. NIST Chemistry Webbook, NIST, http:\\webbook.nist.gov.

Appendix A: Derivation of Klinkenberg Equation


This appendix presents our derivation of the Klinkenberg
equation (ref. 1). Note that our derivation uses components
that Klinkenberg did not employ originally (i.e., elements of
the literature for the kinetic theory of gases (ref. 14)).
Gas Flow Through a Straight Capillary. We consider a
capillary tube of radius (R0) and length (L) with its axis
coincident with the x-axis. Gas is flowing through the
capillary as a result of a pressure drop (p = (p1 - p2)) across
the length of the capillary tube. Assume that the velocity of
the flowing gas (v) is a function only of the distance r from the
x-axis. Consider now a cylindrical shell of length (L) between
the cylinders of radii r and r+
r (see Fig. A.1). The force (F1)
acting on the cross-section of this shell is the normal pressure
due to the flowing gas. In derivative form, this force is:
dF1 2pr
r ................................................................ (A-1)
Assuming steady-state fluid motion, this force is balanced by
the viscous drag exerted by the flowing gas on the outer and
the inner surfaces (along the x-axis) of the cylindrical shell.
The viscous drag is defined by:
F S

dv
.................................................................... (A-2)
dr

where:

Coefficient of viscosity of the gas.


S = Surface considered (2rL for the inner surface,
2
(r+
r)L for the outer surface).
The gas velocity reaches its maximum on the axis of the cylinder (i.e., r=0) and decreases radially towards the wall to a
velocity of zero therefore the velocity gradient dv/dr is
negative. The viscous drag exerted on the inner surfaces is:
dFinner Surface 2Lr

dv
............................................. (A-3)
dr

The component of the viscous drag on the outer surface of the


cylindrical shell, where the gas velocity is lower, is:
dv
d
v
r
dr
................... (A-4)
dFOuter Surface 2L(r
r)
dr

The resulting viscous drag exerted on the entire shell is given


by:
dF2 dFOuterSurface dFInnerSurface

The equilibrium yields dF1 = dF2, hence:


dv
d
v
r
dv
dr
............... (A-5)
2pr
r 2Lr
2L(r
r)
dr
dr

Expanding the second term on the right-hand side of Eq. A-5

SPE 107594

yields:
dv
d (v
r)
dr
2L(r
r)

dr
dv
d 2 v dv
d 2v

2L
r

r (r
)
r2
dr

dr 2 dr
dr 2

...................................................................................... (A-6)
In Eq. A-6, the term
r2(d2v/dr2) is negligible compared to the
other terms neglecting this term and substituting Eq. A-6
into Eq. A-5 yields:
dv

dr
dv
d 2 v dv
2L
r

r (r
)
dr

dr 2 dr

2pr
r 2Lr

...................................................................................... (A-7)
Expanding and simplifying Eq. A-7 gives:
d 2 v dv
rp L
r
.................................................. (A-8)
2
dr

dr

Dividing Eq. A-8 by (-rL) yields the following differential


equation:
d 2v
dr 2

1 dv
p

........................................................ (A-9)
r dr
L

A particular solution of Eq. A-9 (i.e., a non-linear differential


equation) is:
p 2
v A0
r ........................................................... (A-10)
4 L

where A0 is a constant fixed by the boundary conditions. Two


different boundary conditions can be used slip-flow at the
wall or no slip-flow at the wall.
If no slip-flow is assumed at the wall, the condition is v = 0
p
R0 2 . The velocity solution
4 L

for r = R0; this imposes A0


to Eq. A-10 is then:

p
v
( R0 2 r 2 ) ...................................................... (A-11)
4L

The gas flowrate (qg) is obtained by integrating the gas flowrate through the considered cylinder shell section (v(2rdr))
over the cross-section of the capillary tube, which yields the
Poiseuille equation:
p

2r
( R0 2 r 2 ) dr
R0 4 ................ (A-12)

4 L
8 L
0

qg

R0

If the gas slippage is considered at the wall, the condition can


be written v = v0 for r= R0 (where v0 is non-zero). Kundt and
Warburg (ref. 6) proved that the velocity at the wall (v0) is
proportional to the velocity gradient at the wall (
dv / dr
r R0 ),
as given by:

SPE 107954

Improved Permeability Prediction Relations for Low Permeability Sands

v0 c
dv / dr
r R ................................................... (A-13)
0

In Eq. A-13, c is a constant with a value slightly less than 1 (as


given by Kundt and Warburg) and is the average mean free
path of the gas molecules (i.e., the mean free path of the gas
molecules at the mean pressure p (see ref. 14)), defined by
p
p1 p2
/ 2 ). The velocity gradient at the wall is obtained
by differentiating Eq. A-10:

11

Rewriting Eq. A-20 in terms of superficial velocity (vl):


q
vl l
Atube
1 R 2 1 p
N 0
3
8 L

( R0 2 = Atube) .............................. (A-21)

dv
p

r ............................................................... (A-14)
dr
2 L

Applying the condition given by Eq. A-13 to Eq. A-14 yields:


v0 c

p
R0 ............................................................ (A-15)
2 L

Using Eq. A-12 in Eq. A-9, at r = R0, and rearranging yields:

p
A0
R0 2 2cR0 ............................................... (A-16)
4 L

Substituting Eq. A-16 for A0 in Eq. A-10 gives:

p
v
R0 2 r 2 2cR0 .......................................... (A-17)
4 L

As previously done for Eq. A-12, the gas flowrate (qg) is


obtained by integrating v(2rdr) over the whole radius of the
capillary:

0 4 L

qg

R0 p

R0 2 r 2 2cR0 dr

p 4 4c

R0
1

8 L
R0
R 2
1 p 4c
0 R0 2
1

8
L R0

..................................................................................... (A-18)
Re-writing Eq. A-18 in terms of superficial velocity (vl):
q
vg l
Atube
q
R 2 1 p 4c
l 0
1

2
8 L R0
R0

..................................................................................... (A-19)
Gas Flow Through an Idealized Porous Medium. The
"Klinkenberg" idealized porous medium is composed of solid
material through which the capillaries are oriented randomly
and have all the same radius, r. The direction of flow is
parallel to one of the planes of the cube, and let there be N
capillaries. The system for Klinkenberg's idealized porous
medium is shown in Fig. A-1.
The liquid flow rate (ql) through the capillary is given by
Poiseuille's law as: (3-dimensional flow, so Klinkenberg used
1/3 of the total flow for a particular direction).
1 R 2
1 p
ql N 0 R0 2
3
8
L

( R02 = Atube) ................... (A-20)

Fig. A.1

System schematic for a tube (Loeb (ref. 14)).

Darcy's law gives:


ql kl ( L2 )

1 p
L

(L2 = Acube (i.e., rock sample))..... (A-22)

Rewriting Eq. A-22 in terms of superficial velocity (vl):


q
vl l
Acube
1 p
kl
L

(L2 = Acube (i.e., rock sample))..... (A-23)

where kl is the permeability to liquid of the porous medium


which is also the absolute permeability of the porous medium
(provided that the saturation is 100 percent i.e., singlephase flow). Equating Eqs. A-19 and A-20 and simplifying
yields:
1 R 2
kl N 0 .............................................................. (A-24)
3
8

If we assume that that the flowing fluid is a gas, with a gas


slippage condition at the wall (v0 0), the gas flowrate (qg) is
given by Eq. A-18:
qg
R 2 1 p 4c
vg
0
1

................................ (A-19)
8 L R0
R0 2
2

1 R
Substituting kl N 0 into Eq. A-19 for yields:
3
8

v g kl

1 p 4c
1

................................................ (A-25)
L R0

For the same gas flow rate, Darcy's law gives:


v g k a

1 p
.............................................................. (A-26)
L

Where ka is the apparent permeability to gas. Setting Eq. A25 equal to Eq. A-26 and simplifying, yields:
4c
k a kl
1
......................................................... (A-27)
R0

As the mean free path ( ) is proportional to the reciprocal


mean pressure ( p ), we can define:

12

F.A. Florence, J.A. Rushing, K.E. Newsham, and T.A. Blasingame

4c bK

................................................................... (A-28)
r
p

where bK is a coefficient of proportionality, known as the


Klinkenberg gas slippage factor. Substituting Eq. A-28 for
4c/ r in Eq. A-24 gives the Klinkenberg equation:
bK
k a kl
1 (kl k for our definition) ................... (A-26)
p

Appendix B: Derivation of a Theoretical SquareRoot Correlation


Theoretical Capillary Radius. Considering the flow of a
liquid through a capillary tube of inner radius r and length L,
the fluid flow rate (q) is given by Poiseuille
s law as:
r 2 1 p 1 p 4
q r 2

r ....................................... (B-1)
8 L
8 L

where is the viscosity of the fluid and p is the pressure


drop across the length of the tube. Considering a cylinder of
idealized porous medium with a radius R0 and a length L,
composed of n identical capillary tubes (such as described
above), with the same orientation parallel to the axis of the
cylinder. We note that for this discussion (as opposed to our
previous work in this section), that R0 is now defined as the
outer radius of the bulk core sample. By definition, the
porosity of a porous medium is given by:
Void Volume
r 2 L
r2

n
n
................................. (B-2)
2
Bulk Volume
R0 L
R02

SPE 107594

2
r 2

n
r 8k r 2 8k ........................................ (B-7)
R2
0

Or, solving Eq. B-7 for the "equivalent capillary" radius, r, we


obtain:
r 2 2 k / ............................................................. (B-8)

In Eq. B-8, the "units" of permeability depend on the units


system. For consistency, we must apply a conversion factor,
C0, as a multiplier (e.g., when k is in md and r is in cm C0 =
3.141510-6 cm / md , recall that 1 Darcy = 9.8692310-9
cm2). The general form of Eq. B-8 with the units conversion
factor (C0) is:

r 2 2 C0 k / .................................................. (B-9)
Using C0 = 3.141510-6, Eq. B-9 becomes:
r 8.886 10 6 k / .................................................. (B-10)

Theoretical Square-Root Correlation. To correct for the


effects of gas slippage in permeability measurements, Klinkenberg (ref. 1) derived an approximate linear relationship
between the measured gas permeability (ka) and the reciprocal
mean pressure ( p ). This result is given as:
bK
k a k
1 ......................................................... (B-11)
p

The total volumetric flow rate (qtot) of a fluid flowing through


this cylinder is defined by multiplying Eq. B-1 by the number
of tubes, n. This gives:

In Eq. B-11, k is the Klinkenberg-corrected permeability and


bK is the Klinkenberg gas slippage factor. The gas slippage
factor (bK) is a "constant" which relates the mean free path of
the gas molecules ( ) at the mean (absolute) pressure ( p )
and the effective pore-throat radius (r), as given by:

r2 1
p
1 p 4
(r 2 )
n
r ............................ (B-3)
8
L
8 L

4c bK

................................................................... (B-12)
r
p

qtot n

For flow in a porous medium, Darcy's law is defined as:


k p
qtot A
................................................................ (B-4)
L

In Eq. B-4, k is the absolute permeability of the porous


medium, and A is the cross-sectional area of the cylinder.
Since A = R02, Eq. B-4 yields:
k
p
qtot (R02 )
.......................................................... (B-5)

Equating Eqs. B-3 and B-5, and rearranging yields:


4

r
k R02 ................................................................... (B-6)
8

Rearranging Eq. B-6 gives:


2
r 2

n
r 8k
R2
0

Substituting Eq. B-2 into the above result, we have:

In Eq. B-12, c is a constant very close to unity (ref. 6). We


will consider c = 1 for the remainder of this derivation.
Rearranging Eq. B-12 yields:
4
bK p ................................................................... (B-13)
r

Substituting r in Eq. B-13 with Eq. B-10 gives:


0.5

k
4
bK
p

8.886 10

.................................... (B-14)

In Eq. B-14, the mean free path of the gas molecules is


defined by: (ref. 14)
( p, T ) / 2

1
RT

p
M

[ ( p, T )] ...................... (B-15)

In this definition, R is the universal gas constant, M is the


molecular weight of the gas and is the viscosity of the gas at
the corresponding thermodynamic state (i.e., is a function of
the absolute pressure, p , absolute temperature, T, and
composition of the gas). For clarity, the subsequent formulae
are expressed in S.I. units, unless otherwise specified.

SPE 107954

Improved Permeability Prediction Relations for Low Permeability Sands

The square-root correlation developed in this work conforms


to the following assumptions:
The temperature in the core sample is uniform during
steady-state flow (assumed: T = 298 K).
The gas used for the experiments is nitrogen (M =
28.01348 kg/kg-mole).
Assuming that, in the range of pressure applied in the
laboratory, nitrogen behaves as an ideal gas, the variation
of the nitrogen viscosity with pressure is negligible; hence
N 2 (T , p) N2 ,1 atm (T ) .

Table B.1

13

Intercept for Eq. B-20, various gases..

Flowing Gas
hydrogen
helium
air
nitrogen
carbon dioxide

Molecular
Weight
(kg/kg-mole)
2.0159
4.0026
28.9586
28.01348
44.0095

Gas Viscosity at
1atm and 298 K
(Pa.s)
8.845x10-6
1.985x10-5
1.842x10-5
1.781x10-5
1.503x10-5

Eq. B-20
Intercept
(psi)
80.236
127.802
44.106
43.345
29.181

These estimates are based on the fluid properties of ref.2 and


the NIST correlations (ref.15) as appropriate.

Nitrogen viscosity as a function of temperature ( N2 ,1 atm (T ) ,


expressed in Pa.s) is computed using Sutherland's equation:
(ref. 2)

13.85T 1.5
N 2 ,1 atm (T )
10 7 ................................... (B-16)
T 102
The value of the universal gas constant, R = 8,314 J/K/kgmole. The product p in Eq. B-15 can be calculated using
Eq. B-15 and B-16:
p / 2

1
p

8314 298 13.85 2981.5

10 7 p 0.00664
28.01348
298 102

........................................................................................ (B-17)
To obtain the radius, r, in meters (recall that the base unit of
the mean free path is the meter), Eq. B-10 (expressed in
traditional units) becomes;
r 8.886 10 8 k /

Fig. B.1

Using the result from Eq. B-17 and Eq. B-18 in Eq. B-14
yields:
0.5

4 0.00664 k
bK

8.886 10 8

0.5

k
2.9885 105

........ (B-19)

Where bK is expressed in Pa. Converting bK to psi yields:


0.5

k
bK 43.345

.................................................... (B-20)

The significance of Eq. B-20 is that this result establishes the


basis for the bK versus k / correlation given by Sampath
and Keighin (ref. 5). The exponent (-0.5) is established
rigorously from theory as shown above.
The intercept (43.345) is based on the following assumptions:
ambient temperature (T = 298 K) and
ambient pressure (1 atm).
However, these assumptions are probably not significant
compared to the assumption of nitrogen as the reference gas.
In Table B.1 we summarize our estimates of the intercept
coefficient for hydrogen, helium, air, nitrogen, and carbon
dioxide. We have also shown the intercept of the square-root
model plotted against molecular weight for hydrogen, helium,
air, nitrogen, and carbon dioxide in Fig. B.1.

(B-18)

Intercept of the square root model versus molecular


weight of the gas used in the measurements the
viscosity of the different gases except H 2 is similar
(see Table B.1), hence the p-term (Eq. B-17) varies
approximately with the reciprocal square-root of the
molecular weight.

Appendix C: A Rigorous Microflow Model Applied to


the Problem of Gas Flow through Porous Media
Unified Flow Model Gas Flow in Pipes. Karniadakis and
Beskok (ref. 13) developed a unified model that predicts volumetric and mass flow rates for gas flow in channels and pipes
over the entire Knudsen regime (i.e., all flow regimes). The
Karniadakis-Beskok "microflow" model is given (without
derivation) as:

1 p
4 Kn
q lchar 4
[1 ( Kn) Kn]
1
.................(C-1)
8
L
1 bKn

where:
q
Volumetric flow rate in the conduit, cc/sec
lchar
Characteristic length of the flow geometry (e.g.,
channel height, pipe radius), cm
L Length of conduit, cm
p
Pressure drop across the length of the conduit, atm
= Gas viscosity at temperature and pressure, cp
b
= Dimensionless slip coefficient, (b is defined as -1)

Kn) = Dimensionless term in the rarefaction coefficient


In Eq. C-1, the Knudsen number (Kn) is defined by:

14

F.A. Florence, J.A. Rushing, K.E. Newsham, and T.A. Blasingame

Kn
..................................................................... (C-2)
lchar

terms in S.I. units):


The core flow experiments were performed at a constant
room temperature (T = 298 K).
The gas used for the experiments was nitrogen (M =
28.01348 lb/lbmole).
The nitrogen viscosity was computed using Sutherland's
equation (ref. 2):

where is the mean free path of the gas molecules (i.e., the
average distance (length) between two consecutive molecular
interactions).
We use a value of -1 for the general slip coefficient (b) as
recommended by Karniadakis and Beskok. The role of the
rarefaction coefficient [1+
Kn) Kn] is to account for the
transition between the "slip-flow" regime (for which Klinkenberg model was developed) and the "free molecular flow"
regime. In the "slip-flow" regime (i.e., 0.01 < Kn < 0.1), the
rarefaction coefficient is equal to 1 (i.e., = 0); in the "free
molecular flow" regime (i.e., Kn > 10), the volumetric flowrate is independent of the Knudsen number and the parameter
tends toward a constant value (for Kn ).
For reference, the "Knudsen" flow regimes are defined as
follows below and are illustrated graphically in Fig. 12 in the
text. .
Continuum Flow Regime: For Kn < 0.01, the mean free
path of the gas molecules is negligible compared to the
characteristic dimension of the flow geometry (i.e., the lchar
-parameter). In this case the continuum hypothesis of fluid
mechanics is applicable (i.e., the system is described by the
Navier-Stokes equations).
"Slip-Flow" Regime: For 0.01 < Kn < 0.1, the mean free
path is no longer negligible, and the slippage phenomenon
appears in the "Knudsen" layer (layer of gas molecules
immediately adjacent to the wall)
"Transition" Regime: For 0.1 < Kn < 10.
Free Molecular Flow Regime: For Kn > 10, the flow is
dominated by diffusive effects.

The variation of the -parameter as a function of Kn is represented using: (see ref. 13)
( Kn) 0

2
tan -1
c Kn c 2 .......................................... (C-3)

where 0, c1 and c2 are constants. The values for the constants


c1 and c2 are respectively 4.0 and 0.4, and the parameter 0 is
given by:
64
1
64
0

................................................... (C-4)
3 4 15
1

Validation of the Microflow Model Concept. In order to


validate the applicability of the concept of the unified flow
model (Eq. C-1) for the problem of steady-state gas flow
through a porous medium, we computed the Knudsen number
for a given data set of 11 core samples tested at various mean
pressures.
The mean free path of gas molecules is defined by:
( p, T ) 2 /

1
RT

p
M

[ ( p , T )] ........................ (C-5)

The mean free path ( ) is computed using the following (all

SPE 107594

13.85T 1.5
N 2 ,1 atm (T )
T 102
N 2 (T , p) N 2 ,1 atm (T ) 0.12474 0.123688 p
1.05452 10 3 p 2 1.5052 10 6 p 3
.................................................................................................(C-6)

The characteristic length of the conduit (lchar) is required in


order to estimate the required Knudsen numbers. The
"typical" characteristic length is estimated from capillary pressure data or an equivalent single capillary concept and
these "lengths" are typically the capillary radii. We use this
exercise simply to evaluate one definition over another
conceptually, there is no perfect definition because we are
investigating flow in porous media, not uniform capillaries.
The definitions we consider for lchar are:
The computed average pore throat radius estimated from
capillary pressure data (these data are given in the various
reports from which the sample data were extracted).
The theoretical "capillary radius" given by (see Appendix
B):
lchar 8.886 10 6 k / ................................................... (C-7)

The minimum and maximum computed Knudsen numbers


(based on mean pressures) for each sample are presented in
Table C-1. Since most of the Knudsen numbers are greater
than 0.1, the "Transition" Regime (0.1 < Kn < 10) is the
dominant type of flow regime therefore, the rarefaction
coefficient defined in the previous section plays a major role
for modeling the volumetric flow rate. In concept, this observation validates the application of the microflow model for
(steady-state) gas flow in porous media.
Table C.1

Minimum and maximum Knudsen numbers corresponding to the models used to represent the
characteristic length lchar.

Sample ID
1
2
3
4
5
6
7
8
9
10
12

Kn Estimated from the


Average Pore Throat Radius
min Kn
max Kn
0.15
0.34
0.15
0.36
0.43
0.97
0.06
0.15
0.15
0.34
0.37
0.72
0.15
0.34
0.05
0.09
0.15
0.33
0.18
0.34
0.18
0.34

Kn Estimated from
Eq. C-7
min Kn
max Kn
0.64
1.51
0.32
0.77
0.40
0.91
0.72
1.67
0.12
0.27
0.41
0.80
0.22
0.50
0.16
0.26
0.28
0.63
0.28
0.52
0.31
0.57

SPE 107954

Improved Permeability Prediction Relations for Low Permeability Sands

A Rigorous Microflow Model for Gas Flow in an Idealized


Porous Medium. The "microflow" model is defined in the
previous section as:

1 p
4 Kn
q lchar 4
[1 ( Kn) Kn]
1
................ (C-1)
8
L
1 bKn

Poiseuille's law for fluid flow in a pipe (or tube) is given by:

1 p
q lchar 4
......................................................... (C-8)
8
L

while Darcy's law for fluid flow in a porous media is:


q k Acore

1 p
............................................................ (C-9)
L

In Eq. C-9, A is the cross-sectional area of the porous medium


(perpendicular to the direction of the flow). Following the
procedure given by Klinkenberg (ref. 1), we can equate
Poiseuille's and Darcy's laws to yield an expression for the
permeability (k). This procedure requires us to equate Eqs. C8 and C-9, which yields:
4
l
k char ................................................................ (C-10)
8 Acore

Substituting Eq. C-10 into Eq. C-1, we have:


k
p
q Acore
[1 ( Kn) Kn]

4 Kn
1

.................. (C-11)
1 bKn

where:
k
= Permeability, D
q
Volumetric flow rate in the conduit, cc/sec
lchar
The characteristic length of the flow geometry (e.g.,
channel height, pipe radius), cm
L
Length of conduit, cm
p
Pressure drop across the length of the conduit, atm

= Gas viscosity at temperature and pressure, cp


b
= Dimensionless slip coefficient (b is defined as -1)

Kn) = Dimensionless rarefaction coefficient


Kn
= Knudsen number, dimensionless
Multiplying through Eq. C-11 by

L
and using b=-1, we
A p

obtain:
q

L
4 Kn
k [1 ( Kn) Kn]
1
...................... (C-12)
A p
1 bKn

where we note that the left-hand-side of Eq. C-12 is simply the


"gas" permeability (ka) as defined by Darcy's law (i.e., the
"uncorrected" permeability). Making this reduction, we have
our base form:
4 Kn
k a k [1 ( Kn) Kn]
1
............................... (C-13)
1 Kn

Eq. C-13 provides an independent relation between the apparent gas permeability (ka), the slip-corrected permeability
(or Klinkenberg-corrected permeability) (k), and the
Knudsen number (Kn). We now need to finalize Eq. C-13 by
substitution of the relations for
Kn). We substitute Eq. C-3
into Eq. C-1 (and assume that c1 = 4.0 and c2 = 0.4 in Eq. C-

15

3), which yields a direct relation for


Kn) of the form of:

128
( Kn)
tan -1 4 Kn 0.4 .........................................(C-14)
2
15

We now substitute Eq. C-14 into Eq. C-13 to yield our formal
(or complete) result for this work:

128
4 Kn
k a k
1
tan -1 4 Kn 0.4 Kn
1

2

1 Kn
15

.....................................................................................(C-15)
Klinkenberg Model as a Simplified Microflow Model for
Slip-Flow Regime. We can prove "mathematically" (using
rigorous assumptions) that the model developed by
Klinkenberg is actually a simplification of the microflow
model for the case of the slip-flow regime (0.01 < Kn < 0.1).
Recalling the Klinkenberg equation:
bK
k a k
1 ..........................................................(C-16)
p

where:
k = Klinkenberg-corrected permeability, D
ka
Apparent gas permeability, D
bK

Klinkenberg gas slippage factor, atm


p

Mean core pressure, atm


The Klinkenberg gas slippage factor (bK) is defined by
Klinkenberg as a coefficient of proportionality:
4c bK

....................................................................(C-17)
r
p

where c is a constant close to 1 (see ref. 6) and r is the


effective pore throat radius (in our case r is the characteristic
length lchar).
Applying Eq. C-17 to Eq. C-16 yields:

k a k
1 4c .........................................................(C-18)
r

Recall the base form of our microflow model (Eq. C-13):


4 Kn
k a k [1 ( Kn) Kn]
1
................................(C-13)
1 Kn

We assume that we are in a slip-flow regime, then by definition, (Kn) = 0 (see ref. 13). Hence Eq. C-13 becomes:
4 Kn
k a k
1
.....................................................(C-19)
1 Kn

Also, for 0.01 < Kn < 0.1, we have:


1
1 Kn o( Kn) .................................................(C-20)
1 Kn

Hence:
4 Kn
4 Kn 4 Kn 2 o( Kn 2 )
......................................(C-21)
1 Kn
4 Kn

By definition of the Knudsen number (Eq. C-2):

16

F.A. Florence, J.A. Rushing, K.E. Newsham, and T.A. Blasingame

Kn
..................................................................... (C-2)
lchar

Eq. C-19 now yields:


k a k
1 4 Kn

.................................................... (C-22)
k
1 4

lchar

Eq. C-22 is very similar (and almost identical) to Eq. C-18,


provided that the constant c in Eq. C-18 is equal to one and the
mean free path of the gas molecules used in the microflow
model ( ) is defined as the average mean free path of the gas
molecules ( in the Klinkenberg model, evaluated at the
mean core pressure). We can consider that the Klinkenberg
model is an approximation of the microflow model.
Appendix D: Correlations for the "Pseudo-Knudsen"
Number Application to a Tight Gas Example
(North Louisiana, USA)
Determination of the Pseudo-Knudsen Numbers. This data
set (ref. 9) includes 18 core samples where the Klinkenberg
analysis (ref. 1) was performed in order to estimate the liquid
equivalent permeability (k). The measured mean pressure,
the permeability to gas, and the porosity are given for each
sample and using the mean pressure and permeability to
gas data, a Klinkenberg plot is constructed, and the gas
slippage factor (bK) and the Klinkenberg-corrected
permeability (k) are estimated.
As noted, the porosity of the sample is also given, but no
reference conditions were given, so we assumed 300 K (80
Deg F) and 1 atm. The data summary is given in Table D.1.
Table D.1

Summary of the core data (given data and computed results).

SPE 107594

128
4 Kn
k a k
1
tan -1 4 Kn 0.4 Kn
1
0
2

15

1 Kn

...................................................................................... (D-1)
It is easily proven that for any pair of values for ka and k, Eq.
D-1 has a unique positive solution: we refer to this solution as
the "pseudo Knudsen" number.
Correlation of the Pseudo-Knudsen Number with Pressure, Permeability and Porosity. We have developed two
correlations to relate the "pseudo-Knudsen" numbers (Knp)
and the available data (ka, k, , p). The first correlation given
below relates the "pseudo-Knudsen" number (Knp) to the
reciprocal mean pressure, the porosity (), and the gas
permeability (ka):
Kn p 0.4

1 0.598 0.352
ka

.......................................... (D-2)
p

Applying Eq. D-2 to Eq. D-1 yields:


128

1
1
k a k
1
tan 1
4[0.4 k a 0.5980.352 ] 0.4 0.4 k a 0.5980.352
p
152

1 0.598 0.352
0. 4 k a

1 0.598 0.352

1 0.4 k a

...................................................................................... (D-3)
The equivalent liquid permeability can then be easily
computed by:
128

1
1
k k a /
1
tan 1
4[0.4 k a 0.5980.352 ] 0.4 0.4 k a 0.5980.352
2
p
p

15

1 0.598 0.352
0.4 k a

1 0.598 0.352

1 0.4 k a

Sample
ID

k
(md)

bK
(psi)

1-8
2-10
2-22
3-8
3-48
1-1
1-5
1-7
2-2
2-7
2-8
2-12
2-28
3-4
3-6
3-10
3-36
3-55

0.0001
0.1322
0.0168
0.0439
0.0035
0.00007
0.00004
0.00003
0.0004
0.0124
0.0025
0.0025
0.0052
0.0002
0.0999
0.0067
0.0168
0.0013

565.3
14.2
32.5
32.8
83.6
753.5
801.0
771.0
165.1
26.8
77.1
215.1
58.6
303.1
26.4
84.8
39.2
183.7

(fraction)
0.023
0.063
0.056
0.078
0.07
0.029
0.02
0.017
0.036
0.06
0.051
0.055
0.037
0.04
0.066
0.071
0.057
0.066

Since we cannot compute the Knudsen number explicitly, the


goal of this step is to evaluate a "pseudo Knudsen" number
(Knp) using the relation below: (see Appendix C and ref. 13)

...................................................................................... (D-4)
For the second correlation we replaced the gas permeability
with the equivalent liquid permeability:
Kn p 2.62

1 0.553 0.3897
k

........................................ (D-5)
p

Substitution of Eq. D-5 into Eq. D-1 yields:


128

1
1
k a k
1
tan 1
4[ 2.62 k 0.5530.3897 ] 0.4 2.62 k 0.5530.3897
p
p
152

1
2.62 k 0.5530.3897
4
p

1
0.553 0.3897

1 2.62 k

...................................................................................... (D-6)
Rearranging Eq. D-6 yields:

SPE 107954

Improved Permeability Prediction Relations for Low Permeability Sands

17

128

1
1
k a k
1
tan 1
4[ 2.62 k 0.5530.3897 ] 0.4 2.62 k 0.5530.3897
2
p
p
15

1
2.62 k 0.5530.3897
4
p

1
0

1
0.5530.3897
1 2.62 k

....................................................................................... (D-7)
Eq. D-7 is an implicit relation where k is solved as a root of
this relation. Fig. D.1 presents the "pseudo-Knudsen" numbers obtained with Eqs. D-2 and D-5 plotted against the reference "pseudo-Knudsen" numbers (recall that we defined the
reference "pseudo-Knudsen" number as solution of Eq. D-1).

Fig. D.1

Fig. D.2

Computed equivalent liquid permeability versus


reference Klinkenberg-corrected permeability (extrapolated from a Klinkenberg plot (ka versus reciprocal
mean pressure)) both models present reasonably
good results.

Fig. D.3

Plot of residuals computed with Eq. D-8 against


equivalent liquid permeability (k) for an example
point from our data set Sample 1-8, ka = 0.0012699
md, p = 45.27 psig, = 0.023. The Klinkenbergcorrected permeability for this sample is k =
0.000125 md. The root solution for this case yields k
= 0.00018 md, which implies an absolute relative error
of 30.56 percent (compared to the "Klinkenberg"
permeability).

Correlated "pseudo-Knudsen" number (Eq. D-2 and


Eq. D-5) versus reference "pseudo-Knudsen" number
(solution of Eq. D-1) the implicit model (Eq. D-5)
achieves a better estimation of Knp than the explicit
model (Eq. D-2).

In Fig. D.2 we present the equivalent liquid permeabilities


computed using Eqs. D-4 and D-7 plotted against the
reference Klinkenberg-corrected permeability. Figs. D.3 and
D.4 show the behavior of the following function of k for a
given data point (e.g., Sample 1-8, ka = 0.0012699 md, p =
45.27 psig, = 0.023).

18

Fig. D.4

F.A. Florence, J.A. Rushing, K.E. Newsham, and T.A. Blasingame

Plot of absolute value of residuals computed with Eq.


D-8 against equivalent liquid permeability (k) for the
same point as in Fig. D.3 Sample 1-8, ka =
0.0012699 md, p = 45.27 psig, = 0.023. The Klinkenberg-corrected permeability for this sample is k =
0.000125 md. The root solution for this case yields k
= 0.00018 md, which implies an absolute relative error
of 30.56 percent (compared to the "Klinkenberg"
permeability).

SPE 107594

S-ar putea să vă placă și