Sie sind auf Seite 1von 36

MATRIX PERMEABILITY MEASUREMENTS OF GAS SHALES: GAS SLIPPAGE

AND ADSORPTION AS SOURCES OF SYSTEMATIC ERROR

by

ERIC AIDAN LETHAM

A THESIS SUBMITTED IN PARTIAL FULFILLMENT OF


THE REQUIREMENTS FOR THE DEGREE OF

BACHELOR OF SCIENCE (HONOURS)

in

THE FACULTY OF SCIENCE

(Geological Science)

This thesis conforms to the required standard

………………………………………
Supervisor

THE UNIVERSITY OF BRITISH COLUMBIA

(Vancouver)

MARCH 2011

© Eric Aidan Letham, 2011


ABSTRACT

At pressures less than 4 MPa, matrix permeability of gas shales show a strong
dependence upon the pore pressure at which they are measured. This is the result of gas
slippage, a process that causes Darcy’s Law to break down during the low pressure flow of
gas through porous media. In gas shales, the range of pore pressures where gas slippage is
recognizable is larger than in conventional reservoir rocks. This is attributed to the smaller
pore systems found in tighter rocks. The type of gas flowing through porous media also
influences the matrix permeability, as the size of molecules is one of the factors that control
the mean free path of a gas. For this reason, matrix permeability determined using helium as
a probing gas will be different than when methane is used. Comparing this difference to the
underestimation of matrix permeability resulting from neglecting to correct for adsorption,
the latter is found to be negligible. Therefore, when trying to determine the matrix
permeability of a rock to natural gas, no condition was identified where using helium as a
probing gas would lead to a more accurate result than using methane.

ii
TABLE OF CONTENTS

TITLE PAGE………………………………………………….….…………….………..…....i
ABSTRACT…………………………………………………….……………………...……..ii
TABLE OF CONTENTS……………………………….…...…………….……….………...iii
LIST OF FIGURES………………………………….………..…….……….…….………....iv
LIST OF TABLES…………………………………………….………………...……….……v
ACKNOWLEDGEMENTS……………………………………..…………………….…...…vi
INTRODUCTION………………………...……………………….………………...…….…1
BACKGROUND AND THEORY………………………………………….………......……2
Permeability………………………………………………….…………………..……2
Gas Slippage …...……………..……………………………………………….……...3
Adsorption …...……………..……………………………………………….……….10
METHODS……………………………………………………………………………..……11
Measuring Matrix Permeability…………………………….……………..…………11
Measuring Porosity.…...……………..…………………………………..….…...…..16
RESULTS……………………………………………………………………...….……...….17
DISCUSSION………………….……………………………………………..……...…....…21
Implications for Reservoir Modeling.……………………….………..…………...…21
Implications for Permeability Measurement Techniques…………..………………..24
Implications for Probing Gas Selection……………………….……………………..25
CONCLUSION……………………………………………….…………………….….……26
REFERENCES CITED……………………………………….………………………...…...28

iii
LIST OF FIGURES

Figure 1. Velocity profile of fluid flow in a cylindrical pipe…………………….....................4


Figure 2. Schematic diagram of the influence of gas slippage on velocity profiles…………..5
Figure 3. Histogram of pore throat diameters for siliceous mudstones from the Barnett Shale,
Texas…………………………………………….……………………………….……………7
Figure 4. Pore throat size distribution of the Fontainebleau Sandstone……..….…............….8
Figure 5. Relationship of mean free path to gas pressure……………………………………..9
Figure 6. Relationship of mean free path to inverse of gas pressure………………….……..10
Figure 7. Schematic diagram of the experimental setup used to measure matrix
permeability…………………………………………………………………………….........13
Figure 8. Processing example of pulse decay data………………………………………….15
Figure 9. Schematic diagram of a Langmuir isotherm………………………………………16
Figure 10. Sample 1 matrix permeability as a function of pore pressure……………………19
Figure 11. Sample 1 matrix permeability as a function of inverse pore pressure……………20
Figure 12. Permeability as a function of pore pressure for Klinkenberg’s sample “A”……..22
Figure 13. Permeability as a function of inverse pore pressure for Klinkenberg’s sample
“A”…………………………………………………………………………………………...22
Figure 14. Permeability as a function of pore pressure for Klinkenberg’s sample “L”……..23
Figure 15. Permeability as a function of inverse pore pressure for Klinkenberg’s sample
“L”…………………………………………………………………………………………...23

iv
LIST OF TABLES

Table 1. Percent underestimation of matrix permeability at different pore pressures….........20

v
ACKNOWLEDGEMENTS

I would like to thank Dr. R. Marc Bustin for supervising the completion of this study.
I appreciate the time you were able to give me and your willingness to answer all of my
questions. It made for an enjoyable experience.
I would also like to thank Oyeleye Adeboye, Dr. Gareth Chalmers, Dr. Amanda
Bustin, and Kristal Li for their help in the lab, as well as Bryn Letham and Mary Lou Bevier
for their constructive criticism while I was writing this thesis.

vi
INTRODUCTION

Natural gas resource estimates have grown significantly in the past decade (Energy
Information Administration, 2010) due to the recognition of gas shale reservoirs and growing
confidence in our ability to economically exploit them (Bustin, 2005; Natural Resources
Canada, 2008). Successful exploitation of gas shales in the United States has led to an
increase in North American supply. Coupled with the global economic recession beginning
in 2008, this has led to a decrease in natural gas prices (NRC, 2010). Lower profit margins
induced by this high supply-low demand (low price) situation create an environment wherein
producers must keep their costs low in order to remain profitable.
Aside from non-geological factors that determine the economics of a natural gas
deposit (e.g. proximity to transportation networks), the characteristics of a deposit that
govern its profitability are how much gas is in place and how easily it can be brought to the
surface. Specific to gas shales, which North America is currently shifting towards as its
dominant source of supply (Natural Resources Canada, 2009), the key parameters
determining these characteristics are total organic carbon content (TOC), maturity, mineral
matter, thickness, porosity, and permeability (Ross and Bustin, 2007a).
Given that permeability is such an important parameter for predicting reservoir
production, strides have been made to improve the accuracy of its measurement (Cui et al.,
2009). Building upon this, the present study identifies systematic errors in gas shale
permeability measurements, and provides suggestions for how they can be minimized.
Measuring gas shale permeability is problematic. Gas shales typically have
permeabilities much less than one millidarcy (Bustin, 2005), which are orders of magnitude
lower than conventional reservoir rocks. Therefore, traditional steady-flow measurements
require too much time to complete (Dicker and Smits, 1988; Cui et al., 2009). For this
reason, specialized methods for measuring permeability of tight (low permeability)
unconventional reservoir rocks - such as gas shales - have been developed that can be
completed in a shorter time frame. These methods include pressure pulse decay, crushed
sample dynamic pycnometry, desorption tests, and the analysis of Hg intrusion curves (Luffel
et al., 1993; Cui et al., 2009).

1
The methods currently employed to measure the permeability of gas shales are by no
means perfect, and refinement of these techniques is necessary before accurate results can be
obtained. One problem is that desorption, crushed sample dynamic pycnometry, and Hg
intrusion techniques do not take place under confining pressure (Cui et al., 2009), even
though permeability varies markedly with effective stress (Katsube, 2000; Murthy, 2008;
Dong et al., 2010). A second problem is that permeability is often assumed to be a property
of the rock itself, and therefore that it does not vary with the pressure and type of gas flowing
within it. However, this assumption breaks down under circumstances where the process of
gas slippage is significant. A third problem is that many techniques are used without
accounting for adsorption of gas onto the internal surfaces of the shale (Cui et al., 2009).
This thesis addresses the latter two problems by exploring how permeability measurements
change when different probing gases and pore pressures are used. By using different gases,
the size and adsorption characteristics can be varied. Changing the pore pressure causes
variation in the mean free path of the gas molecules. A comparison of permeability
measurements calculated with and without a mathematical correction for adsorption will also
be made. These experiments illuminate systematic errors that can be eliminated by refining
measurement techniques, which will result in more accurate permeability measurements.
This should lead to improved reservoir analysis, and therefore permit better economic
decisions.

BACKGROUND AND THEORY

Permeability
In 1856, Henry Darcy’s observations relating the discharge flux ( ) through porous
media to the pressure gradient ( ) and a constant of proportionality dependent upon the

permeability of the medium ( ) (see Equation 1) were published (Darcy, 1856).

Shortly thereafter, the relationship became known as Darcy’s Law. However, by the 1950’s,
it had been decided that permeability should be a property of the porous media only, meaning

2
it would be independent of the type of fluid flowing within it (Hubbert, 1957). Therefore,
viscosity ( was separated from the proportionality constant resulting in a variation of
Darcy’s Law

where is the permeability.


The permeability of shale gas reservoirs can be broken down into two distinct
categories: fracture permeability and matrix permeability. Fracture permeability characterizes
the ease with which fluid can flow through the natural fractures in the shale, as well as those
fractures created through hydrofracturing. On the other hand, matrix permeability is the ease
with which a fluid can flow within the intact portion of the shale. Fracture permeability
usually falls within the millidarcy range (Bustin et al., 2008), whereas matrix permeability is
in the microdarcy to nanodarcy range (Bustin, 2005). Although it is much lower than fracture
permeability, under some circumstances matrix permeability controls the production of a
shale gas reservoir (Luffel et al., 1993; Bustin et al., 2008).
The present study is concerned with matrix permeability and the methods used to
evaluate it. Compared to conventional reservoir rocks, gas shales generally have much
smaller pores and pore throats. Therefore, in a gas shale, collisions between molecules and
pore walls are more frequent in comparison to collisions between the molecules themselves.
Under these circumstances, the process of gas slippage (discussed below) becomes
important, resulting in matrix permeability that is dependent upon the type of gas flowing
through the shale as well as the gas pressure. Thus, it has been suggested that the ease with
which a gas can flow (permeability) within gas shales is better classed as diffusivity (Cui et
al., 2009). To be consistent with the literature, the ease with which gas can flow through the
matrix of gas shales shall be termed matrix permeability herein, even though the flow
incorporates both advective (Darcy) flow and diffusion. Therefore, matrix permeability is not
restricted to being a property of the rock, and may vary with the flow conditions.

Gas Slippage
Consider the idealized example where a pore throat is represented as a cylindrical
pipe. The two dimensional velocity profile of a fluid flowing through the pore throat would

3
be a hyperbolic curve, as depicted in Figure 1 (John and Haberman, 1971). The velocity is
zero at the pore wall, has a maximum in the middle of the pore throat, and an average
velocity of half the maximum. The volume flow rate ( ) of fluid flowing through the pore
throat can be quantified using Poiseuille’s equation for compressible fluids:

where is the radius of the pore throat, is the viscosity, is the length of the pore throat,
is the upstream pressure, is the downstream pressure, and is the average pressure
(Klinkenberg, 1941).

Figure 1. Two dimensional velocity profile for fluid flowing through a cylindrical pipe.
Length of arrows from zero baseline indicate magnitude of velocity.

Equation 3 holds true for the flow of liquids in reservoir rocks (Klinkenberg, 1941).
However, for the low pressure flow of gas through microporous materials (i.e. low pressure
gas shales) equation 3 breaks down. This is because, for a gas, the velocity directly adjacent
to the pore wall is not zero, and varies with the mean free path of the gas.
The mean free path of a gas is the average distance a molecule travels before it
collides with another molecule (Knight, 2004). Therefore, at the pore wall there is a layer of
gas the thickness of the mean free path where, on average, molecules do not experience any
collisions with one another. Let us call this the MFP layer. Within the MFP layer, the average
velocity will be dictated by the average velocity of the gas molecules immediately adjacent to
the MFP layer (Klinkenberg, 1941). This is due to the fact that, on average, this is the last
place a collision took place between the molecules in the MFP layer and other gas molecules.
As can be seen in Figure 1, this means that the larger the mean free path, the faster the

4
average velocity in the immediate vicinity of the pore wall will be. This is depicted
schematically in Figure 2.

Pore Throat Wall


Velocity Profile
Zero Velocity

Increasing
Mean Free
Path and
Average
Velocity

Velocity

Figure 2. Schematic diagram depicting the impact of changing mean free path of a gas on its
velocity profile when flowing through a cylindrical pipe.

The equation for the mean free path of an ideal gas is as follows:


where λ is the mean free path, is the number of gas molecules, is the volume occupied
by the gas, and is the radius of the gas molecules (Knight, 2004). Equation 5 shows how

5
the molecular density term ( ) in equation 4 can be expressed in terms of pressure ,

temperature and Boltzmann’s constant .

For an isothermal situation, the mean free path of a gas is therefore proportional to the
inverse of pressure. Similarly, the mean free path is proportional to the inverse of the
molecular radius squared (equation 4). So, the lower the pressure and smaller the molecules,
the faster will be the average velocity in the pore throats. A higher average velocity leads to
an increased volume flow rate, which translates to higher matrix permeability.
The impact of mean free path on the volume flow rate of a compressible fluid was
quantified by Klinkenberg (1941) as follows:

in which is a proportionality factor and is the average mean free path over the pressure
range to . Although not quantified exactly, the value of is slightly less than 1
(Klinkenberg, 1941). The rightmost term in equation 6 quantifies the amount of gas slippage
taking place. If the mean free path ( ) approaches zero, the rightmost term becomes 1,
and equation 6 is reduce to equation 3. This means that no gas slippage is taking place.
Likewise, if the radius of the pore throat ( is much larger than the mean free path, the
rightmost term also becomes 1, indicating no gas slippage. For this reason, the effects of gas
slippage will be greater in rocks with smaller pores and pore throats.
It is therefore important to compare the pore size distribution of gas shales to those of
conventional reservoir rocks. Loucks et al. (2009) used capillary-pressure analysis to
determine the distribution of pore throat sizes in siliceous mudstones from the Mississippian
Barnett shale, a major gas shale play in the Fort Worth basin, Texas (Montgomery et al.,
2005). Their results are shown in Figure 3. It can be seen that most of the pore throats lie in
the nanometer size range. Lindquist et al. (2000) used synchroton X-ray tomographic images
to characterize the pore throat size distribution of the Fontainebleau sandstone (see figure 4).
These sandstones are fine grained, well sorted quartz arenites of Oligocene age located just
south of Paris, France (Haddad et al., 2006). In contrast to the mudstones from the Barnett
shale, most of the pore throats in the Fontainebleau sandstone are in the micrometer size

6
range. Assuming that this size relationship holds, it can be expected that gas shales, having
smaller pore throat radii, will experience more gas slippage than conventional reservoir rocks
(see equation 6). Consequently, matrix permeability of gas shales is more strongly dependent
on the mean free path of gas molecules flowing within them than conventional reservoir
rocks.

Figure 3. Histogram of pore-throat diameters calculated from capillary-pressure sample


analysis of four siliceous mudstone samples from the Barnett shale, Texas (Loucks et al.,
2009).

Two probing gases (gases used to experimentally evaluate the physical properties of a
rock) commonly used when measuring permeability are methane and helium. Methane is the
dominant component of natural gas, and therefore would naturally seem like a good selection
for a probing gas. However, complications (discussed below) are introduced due to the fact
that methane is not an inert gas (Cui et al., 2009). On the other hand, helium, a noble gas, is
often chosen as a probing gas because it behaves more like an inert gas. However, the
diameter of methane (0.38 nm) is 46% larger than that of helium (0.26 nm) (Cui et al., 2009).

7
probability

Figure 4. Pore throat size distribution for four rocks from the Fontainebleau Sandstone
(Lindquist et al., 2000).

As can be seen in equation 4, all else held constant, this size discrepancy will result in a
longer mean free path for helium than for methane (see Figure 5). A longer mean free path
translates into a larger volume flow rate (equation 6), which results in higher matrix
permeability.

8
Mean Free Path vs Pressure
25

20
Mean Free Path

15
Helium
(nm)

Methane
10
Power (Helium)
Power (Methane)
5

0
0 2 4 6 8
Pressure (MPa)

Figure 5. Relationship between the mean free path of a gas (calculated using equation 4) and
pressure for both helium and methane.

In Figure 5 it can be seen that the difference between the mean free path of helium
and methane decreases as pressure increases. In fact, the mean free path converges on a
common value of zero at infinite mean pressure (see Figure 6). Because the mean free path
directly affects the volume flow rate of gas through pores and pore throats (equation 6),
variations in matrix permeability with changing pore pressure follow the same trends found
in Figure 5 and Figure 6.

9
Mean Free Path vs Inverse Pore Pressure
25

20

15
Mean Free Path
(nm) Helium
10
Methane

0
0 0.5 1 1.5
Inverse Pressure (MPa-1)

Figure 6. Linear relationship between the mean free path of a gas and the inverse of pressure.
Note the trends for helium and methane converge on a common mean free path of zero at
infinite mean pressure.

Adsorption
It has been recognised that a significant portion of the natural gas contained within
organic rich shales is held in the adsorbed state (Chalmers and Bustin, 2007). Adsorbed gas
exists as a semi-liquid on the pore surfaces of the shale, and has a much higher density than
free gas in the compressed state (Cui et al., 2009). Because the computation of matrix
permeability for tight rocks is often based on mass balance equations, the presence of
adsorbed gas is a complication; the amount of gas contained within the sample ( ) cannot be
calculated using a simple formula involving the pore pressure ( ), temperature ( ), pore
volume ( and the gas constant ( :

The amount of gas contained within the sample when adsorption occurs will be greater than
that calculated using equation 7. For this reason, a matrix permeability calculated using a
mass balance equation without a correction for adsorption would be lower than the actual

10
matrix permeability of the sample (Cui et al., 2009). The degree of error is dependent upon
the pressure at which measurements are taken, as the ratio of gas held in the adsorbed state to
gas held in the compressed state is not linearly proportional to pressure (Cui et al., 2009).
Instead, at low pressures this ratio is high, leading to a larger underestimation of permeability
than at higher pressures where the ratio is lower. The magnitude of permeability
underestimation is also dependent upon the quantity and rank of organic matter in the
sample, as adsorbed gas is mostly found within the microporosity of organic matter, and
microporosity of organic matter is positively correlated with rank (Chalmers and Bustin,
2007).

METHODS

Measuring Matrix Permeability


Measuring the matrix permeability of gas shales using traditional steady flow
techniques is impractical due to the tight nature of these rocks; the flow is so slow that
expensive, very precise equipment would be required (Dicker and Smits, 1988; Cui et al.,
2009). Therefore, for this study matrix permeability measurements were made using the
pulse decay technique on core samples. This process involves monitoring the decay with time
of a pressure pulse imparted on the sample. Core samples were chosen over crushed samples
so that a confining stress could be applied. This is important, as matrix permeability is known
to vary with effective stress by more than an order of magnitude (Katsube, 2000; Murthy,
2008; Dong et al., 2010).
Figure 7 is a schematic representation of the experimental setup. The sample is held
in a Hoek cell between two platens, which also function as a pathway for gas to reach the
sample. The platens and sample are surrounded on their curved cylindrical sides by a plastic
membrane inside the cell. Confining pressure is applied using a hydraulic pump. The plastic
membrane separates the sample from the hydraulic oil, and also functions as a seal to prevent
the probing gas from escaping the system or traveling between the two reservoirs in a path
other than through the sample. A 1.4 cm3 upstream gas reservoir and a 1.2 cm3 downstream
gas reservoir are separated by the sample. The two reservoirs can also be directly connected

11
by opening valve 2. The upstream reservoir is adjoined to a pressurised tank containing the
probing gas, which can be shut off from the system using valve 1. The downstream reservoir
(and upstream reservoir if valve 2 is open) can be connected to or shut off from the
atmosphere using valve 3.
Samples were prepared by cutting 2.9 cm diameter cores from larger diameter core
samples on a drill press. The sample core ends were then cut off using a low speed circular
saw, leaving a roughly cylindrical sample of approximately 3 cm length and 2.9 cm diameter.
The ends of the sample were then machined into flat parallel surfaces. The machining
ensured that the samples were truly cylindrical, which was desirable to assure that the sample
ends met flush with the ends of the platens.
Pulse decay experiments were carried out using the following procedure. First, the
sample was placed in the Hoek cell and the platens fastened tightly against the flat faces of
the core. Second, a confining pressure equal to the desired effective stress of the
measurement was applied. Next, valves two and three were opened, leaving the whole system
open to the atmosphere. Once equilibrium was attained (i.e. the pressure everywhere in the
system was atmospheric), both the differential and absolute pressure transducers were
zeroed. Following this, valve three was shut. Both the gas pressure (which is also the pore
pressure) and confining pressure were then incrementally increased to the test values, taking
care not to exceed the desired effective stress. This was done to prevent fracturing of the
sample, and to ensure that the pore structure didn’t collapse beyond what resulted from the
desired effective stress. Next the sample was left to soak, which allowed the gas to migrate
into the shale. Because gas shales have such low matrix permeability, the soaking period
lasted up to ten hours. Once all pressure transducers stabilized at a constant value, meaning
equilibrium had been attained and the soak period had finished, valve two was shut.
Following this, valve three was opened long enough to decrease the pressure in the
downstream reservoir by 0.28 MPa.
The above steps generated a pressure gradient across the core that drove flow of the
probing gas from the upstream to the downstream reservoir. The differential pressure
between the two reservoirs was then measured with time and recorded by a computer
program at intervals of 0.0007 MPa. The rate of differential pressure decay is related to the
matrix permeability of the rock; differential pressure will decay more rapidly for higher

12
Figure 7. Schematic diagram of the experimental setup used to measure matrix permeability.

13
matrix permeability samples than lower matrix permeability samples. An exact value of
matrix permeability was obtained using equation 8 (Cui et al., 2009):

where is the matrix permeability, is the viscosity of the probing gas, is the
compressibility of the probing gas, is the length of the sample, is the cross sectional area
of the sample, and and are the volumes of the upstream and downstream reservoirs
respectively. is a quantity that reflects the rate at which differential pressure decays. To
obtain , the log of dimensionless differential pressure is plotted against time; is the slope
of this graph (see Figure 8). is defined as:

where is the first solution of the transcendental equation:

The values and are the ratio of the amount gas that can be held in the sample compared to
the upstream and downstream reservoirs respectively:

In equations 10 and 11, is the pore volume and is a factor quantifying the amount of gas
held in the adsorbed state in comparison to the amount of gas in the free state, defined as:

where is the effective adsorption porosity and is the porosity (Cui et al., 2009).
To ascertain , Langmuir isotherms (Langmuir, 1918) were determined. Figure 9
shows the typical shape and important parameters of a Langmuir isotherm, which is a curve
representing adsorbed gas content as a function of pressure with temperature held

14
constant. The Langmuir volume ( ) is the maximum amount of adsorbed gas the rock can
hold. The Langmuir pressure ( ) is the pressure at which the amount of adsorbed gas is one
half of the Langmuir volume. Assuming the adsorbed gas exists as a monolayer of molecules,
the two parameters and can be used to generate the Langmuir isotherm using the
equation:

where is the quantity of gas adsorbed and is the pore pressure. This Langmuir isotherm
can then be used to determine the effective adsorption porosity of the sample at the test
pressure. Dividing this by the porosity of the sample yields .

𝑥 𝑦
𝑠
𝑥

Figure 8. An example of how was calculated from the experimental data. The log of
differential pressure is plotted as a function of time. is the slope of this plot.

Langmuir isotherms for the sample used in this study were not measured directly.
Instead, the TOC content of the sample was measured, and by comparison with samples from
the same area and with similar TOC content and known Langmuir parameters, the

15
Langmuir volume and pressure of the sample used in this study was estimated. These
parameters were then used to correct the matrix permeability measurements for adsorption.
Matrix permeabilities were also calculated without correcting for adsorption (i.e. calculated
with equal to zero, see equations 8 through 13) and with a very high Langmuir volume (8
cm3/g) for means of comparison.

Langmuir
Volume (𝑉𝐿

Amount of 𝑉𝐿
Adsorbed
Gas

Langmuir
Pressure
Pressure

Figure 9. Schematic diagram of a Langmuir isotherm. The Langmuir volume is the maximum
amount of adsorbed gas the rock can contain and the Langmuir pressure is the pressure at
which half of the Langmuir volume is adsorbed.

As discussed above, both the process of gas slippage and the underestimation of
matrix permeabiltiy due to adsorption show a dependence on pore pressure as well as the
type of probing gas used. For that reason, matrix permeability measurements were made at a
variety of pore pressures and with different gases while holding all other variables known to
effect matrix permeability (e.g. effective stress) as constant as possible. Through doing so,
the effects of gas slippage and adsorption on matrix permeability were isolated.

Measuring Porosity
To determine porosity, both the skeletal density and bulk density were first measured.
The skeletal density is the mass per unit volume of rock, excluding the porosity. It is the

16
density of the rock if it were to have zero porosity. The bulk density is the mass per unit
volume of the rock as a whole. It therefore lies in between the skeletal density and the void
volume (pore) density. The difference between skeletal density and bulk density was then
used to determine porosity as follows:

To measure skeletal density, pycnometry was employed. This process involved the
expansion of helium between two previously separated cells with known volume, one of
which contained the sample. Samples were crushed and sieved to a size range of 0.84 to 0.59
mm so that measurement could be completed in a reasonable time frame (approximately 0.5
hours per sample). By comparison of the initial pressures of the two cells and the final
pressure of the two connected cells once equilibrium was established, Boyle’s Law

was used to determine the total volume occupied by the gas. Subtracting this volume from
the total volume of the two cells leaves the volume of the shale sample. Dividing the mass of
the sample by this volume gave the skeletal density.
The bulk density of the shale was determined using mercury immersion. Whole shale
samples were first weighed, and then submerged in a beaker of mercury. The very high
surface tension of mercury prevents it from entering the pores of the shale. The displacement
of mercury was therefore taken to be the volume of the shale sample. Bulk density was then
calculated by dividing the mass of the shale sample by its measured volume.

RESULTS

The sample used in this study was a green-grey parallel laminated siltstone from the
Horseshoe Canyon Formation in south central Alberta. The sample was found to have 13%
porosity and a TOC content of 0.77%. Comparison of the TOC content of this sample to
other rocks from the Horseshoe Canyon Formation with known Langmuir parameters led to
an estimated Langmuir pressure of 1.82 MPa and Langmuir volume of 0.11 cm3/g.

17
Figure 10 shows the results of matrix permeability measurements made on Sample 1
using both helium and methane. Measurements were made over a range of pore pressures
from 1 to 8 MPa. Correction for adsorption of methane using the estimated Langmuir
pressure and volume resulted in matrix permeability that was at most underestimated by
0.2%. This difference is so small that it was impossible to graph on this scale; the two values
overlapped. They are therefore both represented as red squares. Correction for adsorption
using a hypothetical Langmuir volume and pressure of 8 cm3/g and 2 MPa respectively is
represented as green triangles. A Langmuir volume of 8 cm3/g is high for a gas shale, and
would be characteristic of more organic rich materials such as coal. That being true, in this
hypothetical case, error induced by not correcting for adsorption represents the maximum
that would be expected for a gas shale. The percent of underestimation is tabulated in Table
1, and ranges from 10.0% at a pore pressure of 0.93 MPa to 1.2% at a pore pressure of 7.9
MPa.
The trends for both helium and methane are a decrease in matrix permeability with
increasing pore pressure (Figure 10). The matrix permeability to helium drops by 54% and
the matrix permeability to methane 33% over the measured interval. These trends become
more gradual at higher pressure, and closely resemble those for the mean free path of helium
and methane over the same interval of pore pressures (see Figure 5). Figure 11 shows the
relationship of inverse pore pressure to permeability, which is characterized by linear trends
for both helium and methane that converge upon a common matrix permeability at infinite
mean pressure of approximately 3.5 10-4 md. Based on this evidence, it is concluded that
the trends seen in matrix permeability as a function of pore pressure are the result of gas
slippage.
At a common pore pressure, the matrix permeability to helium is always greater than
the matrix permeability to methane. Again, this observation can be attributed to gas slippage;
the smaller diameter of helium compared to methane results in a larger mean free path, and
therefore more slippage.

18
Relationship Between Permeability and Pore Pressure -
Sample 1
1.00E-03

8.00E-04
Matrix Permeability (md)

6.00E-04

4.00E-04

2.00E-04

0.00E+00
0 1 2 3 4 5 6 7 8
Pore Pressure (MPa)

Helium Methane Methane - Langmuir Volume 8cc/g

Figure 10. Results from matrix permeability measurements on Sample 1 over a range of pore
pressures and using different probing gases.

19
Pore Pressure (MPa) Percent underestimation of permeability – no
correction compared to Langmuir Volume of 8
cm3/g and Langmuir Pressure of 2 MPa
0.93 10.0
1.78 6.7
2.71 4.6
3.60 3.4
4.44 2.7
6.23 1.7
7.90 1.2

Table 1. Percent underestimation of matrix permeability when adsorption is ignored for


Sample 1 using a hypothetical Langmuir volume of 8 cm3/g.

Permeability as a Function of Inverse Pore Pressure -


Sample 1
1.20E-03

1.00E-03
Matrix Permeability (md)

8.00E-04

Helium
6.00E-04
Methane

4.00E-04 Linear (Helium)


Linear (Methane)
2.00E-04

0.00E+00
0 0.2 0.4 0.6 0.8 1 1.2
Inverse Pore Pressure (MPa-1)

Figure 11. Matrix permeability as a function of inverse pore pressure for Sample 1. Note that
the trends for both helium and methane converge on a common matrix permeability at
infinite pore pressure.

20
DISCUSSION

Implications for Reservoir Modeling


By analyzing the influence of pore pressure on permeability, Klinkenberg (1941) was
able to recognise the process of gas slippage in petroleum reservoir rocks. However, he
concluded that in formations of moderately high permeability, gas slippage was “not of first
importance for practical purposes” (Klinkenberg, 1941). He went on to state that for
formations of lower permeability (2-10 millidarcies in Klinkenberg’s experiments), the
percent difference of permeability measured at atmospheric pressure compared to that
indicated by extrapolation to infinite mean pressure was considerable (up to 100%).
However, this discrepancy was only significant at atmospheric pressure, and therefore is of
little applicability to modeling a natural gas reservoir, where pressures are much higher than
atmospheric.
Contrastingly, the results of the present study indicate that for rocks of much lower
permeability (e.g. gas shales with matrix permeability in the nanodarcy range), gas slippage
results in significant variation of matrix permeability over a larger range of pressures,
including those experienced in natural gas reservoirs. Under these circumstances, gas
slippage is important for practical purposes, as is discussed below. For means of comparison,
the results from two of Klinkenberg’s 1941 permeability experiments on core samples are
shown in Figure 12 and Figure 14. When compared with Figure 10, the expanded pressure
range where gas slippage is important becomes obvious (note Figure 10 has the same
pressure scale as figures 12 and 14). In comparing the results of Klinkenberg’s study, of
notable significance is that the decay of matrix permeability with increasing pore pressure is
more gradual for the lesser permeable rock (Figure 14) than the more permeable rock (Figure
12).
The results of this study are particularly important for low pressure gas shale
reservoirs, such as the Antrim and New Albany, where typical initial reservoir pressures are
2.8 MPa and 2.1 - 4.1 MPa respectively (Curtis, 2002). During production of these reservoirs,
the reservoir pressures will drop from their initial values, which will result in increased gas
slippage and therefore higher matrix permeabilities. For instance, typical drawdown curves

21
Relationship of Permeability to Pore Pressure
- Klinkenberg's Sample "A"
350

300

250
Permeability (md)

200

150

100

50

0
0 1 2 3 4 5 6 7 8
Pressure (MPa)

Figure 12. Permeability to air as a function of pore pressure for sample “A” in Klinkenberg’s
experiments. Data from Klinkenberg, 1941.

Relationship of Permeability to Inverse Pore


Pressure - Klinkenberg's Sample "A"
350

300

250
Permeability (md)

200

150

100

50

0
0 200 400 600 800 1000 1200
Inverse Pore Pressure (MPa-1)

Figure 13. Permeability to air as a function of inverse pore pressure for sample “A” in
Klinkenberg’s experiments. Data from Klinkenberg, 1941.

22
Relationship of Permeability to Pore Pressure
- Klinkenberg's Sample "L"
5

4.5
Permeability (md)

3.5

2.5

2
0 1 2 3 4 5 6 7 8
Pressure (MPa)

Figure 14. Permeability to hydrogen as a function of pore pressure for sample “L” in
Klinkenberg’s experiments. Data from Klinkenberg, 1941.

Relationship of Permeability to Inverse Pore


Pressure - Klinkenberg's Sample "L"
5
4.5
4
3.5
Permeability (md)

3
2.5
2
1.5
1
0.5
0
0 1 2 3 4 5 6 7
Inverse Pore Pressure (MPa-1)

Figure 15. Permeability to hydrogen as a function of inverse pore pressure for sample “L” in
Klinkenberg’s experiments. Data from Klinkenberg, 1941.

23
for the Antrim Shale indicate that reservoir pressures as low as 1.0 MPa are reached during
the final stages of production (Martini et al., 2003).
Even though most gas shales have initial pressures higher than those of the Antrim
and New Albany (Curtis, 2002), the effects of gas slippage are still important for these
reservoirs. To produce natural gas from the subsurface, a pressure gradient is established
between the wellbore and the surrounding reservoir. Pressure in the well bore is kept low in
order to maintain flow. Therefore, near the well bore, it is expected that the gas pressure is
within the pressure range where gas slippage significantly influences the matrix permeability
of the shale.
In a reservoir model, the accuracy of the input parameters will dictate how accurate
the results are. For this reason alone, the implications of this study are important; care should
be taken to accurately determine how matrix permeability will evolve over the pressure range
of production. However, it is important to recognise that not all gas shale reservoir
production is controlled by the matrix permeability (Bustin et al., 2008). Only under certain
circumstances will matrix permeability have strong control on the outcomes of a model.
Nonetheless, for these cases, the trends observed in this study should be considered.

Implications for Permeability Measurement Techniques


Aside from pressure pulse decay experiments on core samples, two alternate methods
commonly used to determine the matrix permeability of gas shales are crushed sample
dynamic pycnometry and canister desorption tests. Crushed sample dynamic pycnometry
involves a similar setup to that used when determining the skeleton density of samples in this
study. However, of interest is the rate of change of pressure upon allowing gas to expand
between the two cells of known volume, one containing the sample. In a theoretical
discussion of this method, Cui et al. (2009) suggested that matrix permeability would show
variation depending on the pressures used. The analytical results of the present study support
this conclusion. This being the case, when trying to determine the matrix permeability of a
high pressure gas shale reservoir, crushed sample dynamic pycnometry measurements
performed at low pressures will yield an overestimation of matrix permeability. Adding the
fact that this method does not take place under confining pressure, the overestimation could
be orders of magnitude in size.

24
Canister desorption tests involve retrieving core from the subsurface and immediately
placing it in a sealed canister that is maintained at reservoir temperature. The cumulative
volume of gas released from the sample is then measured as a function of time. From this
data, matrix permeability can be calculated (Cui et al., 2009). However, during the transition
from the subsurface to the canister, it is inevitable that the pore pressure in the sample will
drop by some amount. Additionally, once the sample is sealed in the canister, pore pressure
will continue to drop throughout the duration of the experiment. If the pressure drops down
into the range where matrix permeability is sensitive to pore pressure, the result would be a
measured matrix permeability higher than that of interest.

Implications for Probing Gas Selection


When measuring matrix permeability to natural gas, intuitively it would be a good
idea to use helium as a probing gas so that the systematic error induced by using methane
without a correction for adsorption can be avoided. This would save the time and money
required to determine the Langmuir isotherm of the sample. However, using helium, a gas
significantly smaller than the average molecules in natural gas, also leads to systematic error;
the smaller diameter of helium leads to a larger mean free path, which translates to higher
volume flow rates through pore throats, and therefore higher matrix permeability. As is seen
in Figure 10, even for a hypothetical unrealistically high Langmuir volume for a shale, the
systematic error caused by not correcting for adsorption is dwarfed by the systematic error
resulting from the increased gas slippage taking place when using helium as a probing gas.
Both the systematic errors induced by not correcting for adsorption when using
methane as a probing gas and by increased levels of gas slippage when using helium as a
probing gas decay as pressure is increased; at high pressures, the mean free path of either gas
is so small that the effects of gas slippage become negligible, and the ratio of gas held in the
free state to gas held in the adsorbed state becomes so high that the effects of adsorption on
matrix permeability can be ignored. Under this circumstance, it might be concluded that
matrix permeabilities determined using either helium or methane as a probing gas will be
similar. In this study, not enough data was taken at high pressures to assess whether or not
this is true. However, some studies have indicated that, because of its smaller size, helium
will be able to access pores and percolation pathways that methane cannot (Ross and Bustin,

25
2007b). This could conceivably influence the matrix permeability. It is therefore concluded
that when trying to determine the matrix permeability of a rock to natural gas, no situation
has been identified in this study where using helium as a probing gas will give a more
accurate result than methane.

CONCLUSION

Increased scarcity of easily retrievable natural gas has resulted in the exploitation of
tight unconventional resources. These include gas shales, which in comparison to
conventional reservoir rocks are characterized by much smaller pores and pore throats, and
therefore considerably lower matrix permeability. Due to gas slippage, the smaller pore
structure also results in dependence of matrix permeability on pore pressure over a wider
range of pore pressures than in conventional reservoir rocks. This became evident when
comparing permeability trends from Klinkenberg’s 1941 study of conventional core samples
to those measured for gas shales in this study.
Even with an unrealistically large Langmuir volume, underestimation of matrix
permeability due to not correcting for adsorption was at most 10%. However, using helium as
a probing gas (as is common practice to avoid correcting for adsorption) resulted in
overestimation of methane matrix permeability by as much as 65%. This is attributed to
helium, a smaller molecule, experiencing more gas slippage. Therefore, when determining
the matrix permeability to natural gas (of which methane is the dominant component) no
circumstance was identified in this study where using helium as a probing gas gave more
accurate results than using methane.
By measuring matrix permeability using methane as a probing gas and taking care to
complete measurements at the correct pore pressure, systematic error can be avoided. This
will result in more accurate model parameters. In turn, this should permit better economic
decisions in a time when profit margins on natural gas are thin.
The sample used in this study had a measured matrix permeability in the 10-4 md
range. Gas shales can have matrix permeabilities orders of magnitude smaller than this.

26
Further study should therefore be completed using tighter rocks than that used in this study. It
would be expected that tighter rocks would have even smaller pore structures, and therefore
would show an even stronger dependence of matrix permeability on pore pressure.

27
REFERENCES

Bustin, R.M. 2005. Gas Shales Tapped for Big Play: AAPG Explorer, February.

Bustin, A.M.M., Bustin, R. M., and Cui, X., 2008, Importance of fabric on the production of
gas shales: SPE Paper 114167-PP presented at the 2008 Society of Petroleum Engineers
Unconventional Reservoirs Conference, 10-12 February, Keystone, Colorado, USA.

Chalmers, G. R. L., and Bustin, R. M., 2007, The organic matter distribution and methane
capacity of the Lower Cretaceous strata of Northeaster, British Columbia, Canada:
International Journal of Coal Geology, v. 70 p. 223-239.

Cui, X., Bustin, A. M. M., and Bustin, R. M., 2009, Measurements of gas permeability and
diffusivity of tight reservoir rocks: Different approaches and their applications:
Geofluids, v.9 p. 208-223.

Curtis, J. B., 2002, Fractured shale-gas systems: American Association of Petroleum


Geologists Bulletin, v. 86. P. 1921-1938.

Darcy, H., 1856, Les Fontaines Publiques de la Ville de Dijon, Victor Dalmont, Paris, 647 p.

Dicker, A. I. and Smits, R. M., 1988, A practical approach for determining permeability from
laboratory pressure-pulse decay measurement:, Paper SPE 17578 prepared for the 1988
International Meeting on Petroleum Engineering, November 1-4, Tianjin, China.

Dong, J., Hsu, J., Wu, W., Shimamoto, T., Hung, J., Yeh, E., Wu, Y., and Sone, H., 2010,
Stress-dependence of the permeability and porosity of sandstone and shale from TCDP
Hole-A: International Journal of Rock Mechanics & Mining Science, v. 47 p1141-1157.

Energy Information Administration, 2010, Proved reserves of natural gas,


http://www.eia.gov/cfapps/ipdbproject/iedindex3.cfm?tid=3&pid=3&aid=6&cid=&syid
=2001& eyid=2010&unit=TCF (20 November, 2010).

Haddad, S. C., Worden, R. H., Prior, D. J., and Smalley, P. C., 2006, Quartz cement in the
Fontainebleau sandstone, Paris Basin, France: Crystallography and implications for
mechanisms of cement growth: Journal of Sedimentary Research, v. 76, p. 244-256.

Hubbert, M. K., 1957, Darcy’s law and the field equations of the flow of underground fluids:
Hydrological Science Journal, v. 2 p. 23-59.

28
John, J. E. A., and Haberman, W., 1971, Introduction to fluid mechanics: New Jersey,
Pretence-Hall, 488 p.

Katsube, T. J., 2000, Shale permeability and pore-structure evolution characteristics: Natural
Resource Canada Current Research 2000-E15.

Klinkenberg, L. J., 1941, The permeability of porous media to liquids and gases: Drilling and
Production Practice, v. 41, p. 200-213.

Knight, R. D., 2004, Physics for Scientists and Engineers: A Strategic Approach: San
Francisco, Addison Wesley.

Lindquist, W. B., Venkatarangan, A., Dunsmuir, J., and Wong, T., 2000, Pore and throat size
distributions measured from synchrotron X-ray tomographic images of Fontainebleau
sandstones: Journal of Geophysical Research, v. 5, p. 509-521.

Loucks, R., Reed, R., Ruppel, S., and Jarvie, D., 2009, Morphology, genesis, and distribution
of nanometer-scale pores in siliceous mudstones of the mississippian Barnett shale:
Journal of Sedimentary Research, v. 79 p. 848-861.

Luffel, D. L., Hopkins, C. W., and Schettler, P. D., 1993, Matrix permeability measurements
of gas productive shales: Paper SPE 26633 presented at the 1993 SPE Annual Technical
Conference and Exhibition, 3-6 October, Huston, TX, USA.

Martini, A. M., Walter, L. M., Ku, T. C. W., Budai, J. M., McIntosh, J. C., and Schoell, M.,
2003, Microbial production and modification of gases in sedimentary basins: A
geochemical case study from a Devonian shale gas play, Michigan basin: American
Association of Petroleum Geologists Bulletin, v. 87, p. 1355-1375.

Montgomery, S. L., Jarvie, D. M., Bowker, K. A., and Pollastro, R. M., 2005, Mississippian
Barnett Shale, Fort Worth basin, north-central Texas: Gas-shale play with multi-trillion
cubic foot potential: AAPG Bulletin, v. 89 p. 155-175.

Natural Resources Canada, 2010, Canadian Natural Gas: Monthly Market Update,
http://nrcan.gc.ca/eneene/sources/natnat/2010/apravr-eng.php (20, November, 2010).

Natural Resources Canada, 2009, North America natural gas – heating season and winter
update.

29
Natural Resources Canada, 2008, Canadian natural gas: review of 2007/08 & outlook to
2020.

Pathi, V. S. M., 2008, Factors affecting the permeability of gas shales [MSc Dissertation],
Vancouver, University of British Columbia.

Ross, D. J. K. and Bustin, R. M., 2007a, Shale gas potential of the Lower Jurassic
Gordondale Member, northeastern British Columbia, Canada: Bulletin of Canadian
Petroleum Geology, v. 55 p. 51-75.

Ross, D. J. K. and Bustin, R. M., 2007b, Impact of mass balance calculations on adsorption
capacities in microporous shale gas reservoirs: Fuel, v. 86 p.

Vermesse, J., Vidal, D., and Malbrunot, P., 1996, Gas adsorption on zeolites at high pressure:
Langmuir, v. 12 p. 4190-4196.

30

Das könnte Ihnen auch gefallen