Sei sulla pagina 1di 42

SPE 131791

Coalbed Methane: Current Evaluation Methods, Future


Technical Challenges
C.R. Clarkson, University of Calgary, and R.M. Bustin, University of British Columbia

Copyright 2010, Society of Petroleum Engineers

This paper was prepared for presentation at the SPE Unconventional Gas Conference held in Pittsburgh, Pennsylvania, USA, 2325 February 2010.

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 have not been
reviewed by the Society of Petroleum Engineers and are subject to correction by the author(s). The material does not necessarily reflect any position of the Society of Petroleum Engineers, its
officers, or members. Electronic reproduction, distribution, or storage of any part of this paper 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 SPE copyright.

Abstract
Coalbed methane (CBM) produced from subsurface coal deposits, has been produced commercially now for over 30 years in
North America, and relatively recently in Australia, China and India. Historical challenges to predicting CBM well
performance and long-term production have included: accurate estimation of gas-in-place (including quantification of in-situ
adsorbed gas storage); estimation of initial fluid saturations (in saturated reservoirs) and mobile-water-in-place, estimation of
the degree of under-saturation (undersaturated coals produce mainly water above desorption pressure); estimation of initial
absolute permeability (system); selection of appropriate relative permeability curves; estimation of absolute permeability
changes as a function of depletion; prediction of produced gas composition changes as a function of depletion; accounting for
multi-layer behavior, and accurate prediction of cavity or hydraulic fracture properties. These challenges have primarily been
a result of the unique reservoir properties of CBM. Much progress has been made in the past decade to evaluate fundamental
properties of coal reservoirs, but there is still work to be done to obtain accurate estimates of some basic reservoir properties.
In recent years, horizontal wells and more complex well architectures and stimulation methodologies have been
implemented to improve recovery of CBM. These more complex development options bring with them a new set of
challenges for operators producing CBM. The exploitation of more geologically-complex coal with poorer reservoir quality
will necessitate new and inventive ways to develop the existing natural gas resources and possibly combine this with new
methods to extract energy from the coal in-situ. Development planning in these scenarios will become increasingly complex
as will evaluation methods.
The purpose of the current work is to review the state-of-the-art in CBM reservoir property and stimulation efficiency
evaluation and speculate on possible CBM development scenarios for the future and the technical challenges they will bring.
Current and future work required to meet these challenges will be discussed in the hope that industry, academia, and
government bodies alike will be proactive in the development of solutions that will make future CBM recovery efficient,
economic, and environmentally friendly.

Introduction
The presence of gas associated with coal has been long recognized due to explosions and outbursts in underground coal mines
(Flores 1998). Initial interest in production of coalbed methane was motivated by the need to pre-drain methane prior to
mining for safety reasons. Degasification of underground mines was, and continues to be, carried out by horizontal in-seam
boreholes in the mine and vertical boreholes in advance of mining. Coalbed methane (CBM) research in the United States was
initially undertaken by the Bureau of Mines (USBM) and Department of Energy (DOE) in the 1970s. In 1971 the USBM
began an experimental gas drainage program at 11 different sites across the United States using oil and gas technology
(Schraufnagel and Schafer 1996). The United States Bureau of Mines completed a 5 hole pilot project in the late 1970s in the
Black Warrior Basin of Alabama and later expanded their program to a 23-hole demonstration project that showed that some
73% of the in-situ methane could be produced. In the early 1980s the Gas Research Institute (GRI) worked to develop
technology for recovery of CBM which culminated in the mid 1990s with publication of their excellent volumes on coalbed
methane assessment and development, much of which remains relevant today.
Widespread interest in coalbed methane as a natural gas source in the U.S. was a result of the U.S. Congress enacting the
Crude Oil Windfall Profit Tax in 1980 and specifically the Section 29 tax credit. The purpose of the act was to promote
domestic production of fuels from unconventional deposits which were those deposits that were deemed difficult and /or
expensive to produce. These unconventional deposits specifically included coalbed methane, tight gas sands, gas produced
2 SPE 131791

from Devonian shale and oil produced from shale or tar sands. The Section 29 Tax Credit applied to any unconventional gas
wells placed in service after 1979 and before January 1st, 1993. The tax credit amounted to $3.00 per barrel oil equivalent and
was adjusted for inflation. In 1990 the tax credit was worth $0.87/mcf. The tax credit applied to gas produced and sold until
December 31, 2002. During the years where wells qualified for tax credits, exploration took place in all major coal basins in
the U.S. and provided a major stimulus to the exploration and development of what was at that time sub- or marginally-
economic deposits. Subsequently a flurry of coalbed methane exploration occurred worldwide mainly as result of the success
in the early 1990s in the San Juan Basin Fruitland coal fairway where production rates from coals commonly exceeded 1
mmcf/day and rates up to 13 mmcf/day had been recorded. Nowhere however was the success of San Juan Basin duplicated
and interest in coalbed methane declined until the late 1990s when renewed exploration was driven by higher gas prices. The
first commercial CBM success outside the U.S. occurred in the dry Horseshoe Canyon Coals in Western Canada. In recent
years the most notable coalbed methane development outside of North America has occurred in Queensland, Australia where
well production rates commonly exceed 1 mmcf/day from vertical wells and short lateral horizontal wells. Outside of North
America and Australia, coalbed methane exploration to date has led to little or no commercial development even though
substantial exploration has taken place and large volumes of gas have been proven to be present. Failure of coalbed methane
prospects has often been due to low permeability of the coals and resulting sub-economic production rates rather than low gas
contents. Currently as a result of the combination of comparatively low gas prices and the success of gas shales, coalbed
methane exploration in most parts of the world has waned, with a few notable exceptions.
Worldwide the magnitude and distribution of potential resources of coalbed methane mimics the distribution of coal
resources. The global resources of coal gas is poorly constrained due to most coal resource assessments being limited to coals
that are thick enough for commercial mining and to those coals that occur at mineable depths. Additionally, little exploration
or development for coal gas has occurred in many countries and methods of resource assessment are not standardized. The
largest coal resources and hence coal bed methane deposits occur in 14 countries and regions (Table 1) with the largest
resources occurring in Russia, China, United States, Australia and Canada. The resources in Russia are largely unexplored and
hence undeveloped and much of the current coalbed methane production in China is coal mine methane, where gas is produced
firstly to degasify coals as part of underground mining operations. The lack or paucity of coal gas production from Russia has
been mainly a result of ample supply of conventional natural gas. Development in China, India, and many European countries
have been largely unsuccessful due to the low permeability of deep and/or high rank coals that have yielded sub-economic
production rates even though large quantities of gas exist. In still other countries such as Indonesia, Vietnam and southern
Africa the lack of market and infrastructure has hampered development.

CBM Resource In CBM Recoverable


Country/Regio
Place (TCF) Recource TCF
n
Russia 450-2000+ 200
China 700-1270 100
United States 500-1500 140
Australia/New 500-1000 120
ZealandCanada 360-460 90
Indonesia 340-450 50
Southern Africa 90-220 30
Western Europe 200 20
Ukrain 170 25
eTurkey 50-110 10
India 70-90 20
Kazakhstan 40-60 10
South America + 50+ 10
Mexico Poland 20-50 5

Table 1 Estimated coalbed methane resources and recoverable resources of the major coal areas of the World (modified from
Kuuskraa, 2009).

In the US about 20 BCF/day of gas is produced from coal seams annually which is about 8% of U.S. domestic production
(EIA 2009). In Canada current gas production from coals seams is about 1 BCF/day, all of which is from the Western
Canadian Sedimentary Basin and of this greater than 90% is from the dry Horseshoe Canyon coals. In Australia coalbed
methane production has grown from effectively zero in 1996 to about 275 mmcf/day in 2009 and is anticipated to grow
exponentially as the gas market expands with construction of liquefied natural gas (LNG) terminals.
In this paper we review the state-of-theart in CBM reservoir characterization and stimulation efficiency evaluation and
speculate on possible future development. The paper is not meant to be exhaustive; we consider those aspects which we
believe have had and will have the greatest impact on future CBM exploitation. The paper is also tainted by personal bias and
interest of the authors. In the last five years our understanding of ultra-tight reservoirs that include many coals has benefited
SPE 131791 3

from the tremendous interest in gas shale exploitation. Additionally most of the novel theoretical and experimental research
on CBM during the same time period has been motivated by the potential for sequestration of carbon dioxide in coals.

State-of-the-Art in Reservoir Property and Stimulation Efficiency Analysis


Coal is a unique source and reservoir for gas. Because of the many unique properties of coal, some of which are summarized
below, quantifying reservoir properties, stimulation efficiencies and forecasting production rates of coal is challenging and
requires unique analytical tools and theoretical analyses. During early burial, biogenic gas, and during later burial,
thermogenic gas, is generated some of which is retained in the coal in the adsorbed state due to the high surface area of the
organic matter. Since the gas in the adsorbed state has a comparatively high density, significant volumes of gas can be stored
even at relatively low reservoir pressures. Compared to conventional reservoir rocks coal has a very low Youngs Modulus
and high Poissons ratio, and coal invariably has one or two sets of closely spaced extension fractures (cleat) formed during the
normal course of coalification and/or subsequent deformation. The solid matrix between coal fractures (cleat) is comparatively
impervious and the matrix swells significantly with gas adsorption and shrinks with gas desorption. These properties combine
to result in a reservoir in which fracture permeability may increase or decrease with production and in a reservoir that has a
propensity to have pre-existing fines and/or one in which fines are produced during drilling and completion.
Coal is considered (modeled) as a dual or three porosity system comprised of micropores (and transitional to mesopores)
and macropores. The macroporosity includes fractures, phyteral porosity, and intergranular porosity. Although the movement
of gas and fluids through coal is still imperfectly understood, it is generally assumed that the macroporosity is responsible for
the permeability of coal and the microporosity is the major contributor to the surface area of the coal and hence is responsible
for gas storage by adsorption.
Gas production from CBM wells occurs in response to reduction of pressure in the reservoir by pumping water from the
coal, in the case of wet coals and reduction in gas pressure by production for dry coals. As pressure is reduced, gas desorbs
from the coal matrix (as defined by the adsorption isotherm) and moves by diffusion through the matrix to the natural fractures
or larger pores. The fracture/cleat system, which has a much higher permeability than the matrix, is generally initially water
saturated and hence during production, relatively permeability changes occur.

Gas Capacity and Gas Content Determination. The reservoir capacity of coal beds includes porosity available to free gas
in pores and fractures, adsorbed gas principally on the organic fraction, and solution gas in bitumen and water. The gas
content is the amount of in-place gas which is the sum of sorbed, solution plus free gas. If the gas content is the same as the
gas capacity, the coal is referred to as saturated. If less gas is present than the capacity, the coal is referred to as undersaturated
and a decline in reservoir pressure is required to optimize production.
Reservoir Capacity. Adsorbed plus solution gas capacity is determined by isotherm analyses. An excellent detailed
summary of methods is provided by White et al. (2005). Adsorption capacity as measured by adsorption isotherms using
volumetric or gravimetric techniques has been applied for almost 100 years to coals and the study of adsorption phenomena in
other materials dates from much earlier (i.e. Langmuir 1916; Sinkinson and Turner 1926).
There has been little advancement in analytical analyses of coalbed methane capacity or determination of gas content from
coal since routine measurements began about 30 years ago. Adsorption capacity of coals continues to be routinely measured
using a volumetric Boyles Law type apparatus based on the following mass balance calculation for pure gases for sample
with mass, ms, (@STP conditions; T = 273.15 K, P = 0.101325 MPa):

T PrefI 1 PrefI I I 1

V ads = T STD Vref (Vvoid Vs Psc Psc .......................................................................(1)
PSTD ms z z z z
The sample void volume (Vvoid), which is the volume in the sample cell not occupied by solid sample (includes free space in
the sample chamber and accessible porosity in the sample), must be corrected for the volume occupied by the adsorbate (Vs) by
assuming a molar density of the adsorbate which is normally assumed to be at its boiling point at atmospheric pressure and Z
is the gas compressibility. ms = sample mass, Pref = reference cell pressure before and after expansion to the sample cell and Psc
= sample cell pressure before dosing with gas and after equilibrium is reached. If the adsorbate volume is neglected, the
Gibbs isotherm is obtained. Using the Gibbs excess approach (Gibbs 1961) the amount of gas adsorbed is determined
experimentally by (Sircar 1985):
nsorbed = ntotal - cgasVvoid........................................................................................................................... .............................(2)
Where n total is the total amount of gas in the system and cgasV void is the gas occupying the void volume calculated from the
molar concentration in the gas phase, cgas utilizing an equation-of-state for the gas at various pressure/temperature conditions.
The void volume in these calculations includes free space within the sample bomb and porosity within the sample not occupied
by sorbate. During the adsorption experiments the void volume progressively decreases as the sorbate occupies space which is
considered in the mass balance calculations.
4 SPE 131791

Currently there are no standard methods of adsorption analyses and inter-laboratory results tend to show dispersion even
though most laboratories tend to follow the same general procedure (Goodman et al. 2007) albeit most commercial
laboratories are seldom forthcoming about the details of their methods. Yu et al. (2007) suggest some of the inter-laboratory
dispersion maybe due to volume of reference and sample cells used but conclude the importance of experimental procedures
remain unclear. Other potential errors in adsorption analyses have been explored by Sakurovs et al. (2009) who suggest errors
in void volume measurement or helium density may account for discrepancy between laboratories.
In the last decade advances in adsorption analyses of coals has been mainly driven by researchers investigating the carbon
dioxide sequestration potential of coals. Recent studies (discussed later) find the void volume during adsorption analyses
decreases due to swelling of the organic matter during adsorption, the amount of which is organic matter- and gas-dependent.
Since carbon dioxide adsorption results in substantial swelling of coal, consideration of volume change due to swelling may be
an important consideration in established adsorption capacity of this gas.
To calculate the adsorbed gas component (or excess adsorption), precise measurement of the void volume (V void) is
required (Eq. 1). Helium expansions are used for void volume calculations as it is considered to give precise measurement of
the void space, hence sample volume by difference (Mavor et al. 1990). Helium is used as a choice of volumetric fluid for two
reasons: 1) it has a small kinetic diameter which enables penetration to the finest microporosity (Singh and Kakati 2000); and
2) it is commonly assumed that helium has a low adsorption coefficient at room temperature and moderate pressure (up to 9
MPa). However these assumptions have been questioned since solid atoms can attract helium (Starzewski and Grillet 1989;
Rouquerol et al. 1999) with helium even adsorbing on inert solids such as silicates (Gumma and Talu 2003). Recent studies
by Ross and Bustin (2007) have shown that using helium to measure the void volume may overestimate the void volume
available to other gases since shales and coal have pores systems so fine they act as molecular sieves to other gases- the
amount of which is sample dependent. If such is the case the adsorption capacity of coals (and shales) will be underestimated-
the amount by which will depend on the pore system and total sorption capacity.
Multicomponent gas adsorption analyses has received much recent interest as a result of the need to better understand and
model carbon dioxide and flue gas sequestration and the potential for enhanced coalbed methane production by gas sweeps.
The Extended Langmuir Model (Arii et al. 1992; Mavor 1996) is the most common model used to calculate multicomponent
gas adsorption based on pure gas isotherms. Although the Extended Langmuir Model (ELM) is still routinely used, recent
studies have found that thermodynamic approaches better match mixed gas adsorption isotherms (i.e. Clarkson and Bustin
2000). Clarkson and Bustin (2000) showed that the ideal adsorbed solution theory (IAS) using the Dubinin-Radushkevich (D-
R) and Dubinin-Astakhov (D-A) theory better predicts experimental binary methane-carbon dioxide adsorption analyses than
the ELM. The IAS/D-A theory provides better predictions where carbon dioxide selectivity varies with pressure. The IAS
theory predicts decreasing selectivity with increasing carbon dioxide concentration. Clarkson and Bustin (2000) and Bustin
(2009) investigated the role of moisture and coal composition and showed that moisture decreases selectivity of adsorption of
carbon dioxide but the effects of coal composition are less obvious. Clarkson and Bustin (2000) further showed IAS/D-A is
more accurate than the ELM and that the results of IAS are strongly depended on choice of pure gas isotherms. Manik et al.
(2002) have also shown the IAS model provides better predications of experimental data in binary and tertiary gas systems.
Clarkson (2003) provided a model in which the vacancy solution and Dubinin-Polanyi theories are combined in a
multicomponent model which further considers gas compositional shifts in produced gas. Many of the recent studies related to
sequestration and enhanced coalbed methane production have been summarized by Gasem et al (2009) and Pan and Connell
(2008). Pan and Connell (2008) compare the results of IAS, ELM and the 2 Dimensional Equation-of-State (2D EOS) for two
coals. They found the IAS and ELM models yielded similar results for both coals whereas the 2D EOS model yielded results
different from either the IAS or ELM models. With the large number of new researchers from various fields and varying
expertise pursuing the potential for sequestration of carbon dioxide in coal, we anticipate major ongoing research in the area of
multicomponent adsorption theory.
Free gas and solution gas capacity in coals has generally not been considered in quantifying reservoir capacity or gas
content of shales. Studies by Bustin and Clarkson (1999) and Bustin and Bustin (2009) show however that in low rank coals,
in which pores are comparatively large and surface areas comparatively small, free gas may comprise a significant component
of the reservoir capacity and gas content. This is particularly the case in dry coals where the pore and fracture systems are not
saturated with water such as the dry Horseshoe Canyon coals in Alberta. Quantifying the free gas capacity of coals is through
analytical methodologies akin to those used in gas shales- namely core preservation at the well site, measurement of pore
volume and water saturation. Solution gas has similarly been ignored in most coals. In low rank coals such of those of the
Powder River Basin of Wyoming or Horseshoe Canyon coals of Alberta, however, solution gas can be shown to be a
measureable and economically important part of the resources (Bustin and Bustin 2009).
Reservoir Gas Content. There has been comparatively little advancement in gas content determination of coals over the
last 35 years. Gas content of coal is still determined almost exclusively by desorption testing using the techniques described
by McLennan et al. (1995). The principals of desorption testing were initially laid out by Kissell et al. (1973) of the U.S.
Bureau of Mines (USBM) for measurements of gas content of coal. Subsequently minor modifications have been
recommended (Australian Standard 1991; McLennan et al. 1995) but overall the techniques utilized today are essentially those
of the USBM with some refinement. Desorption testing is based on the premise that gas stored in coal in the adsorbed state
takes considerable time to desorb and hence if a sample at the well site is sealed, subsequent gas that is desorbed can be
volumetrically determined and used as a measure of the gas-in-place in the reservoir. Additional samples can be collected for
SPE 131791 5

composition analyses. Samples for desorption analyses are collected at the well site and immediately placed in sealed
canisters and brought to the temperature of the drilling fluid, or at a calculated temperature depending on sample and well
conditions. Desorption tests can be performed on conventional core, cuttings, or sidewall core.
The largest error in desorption tests are in the estimate of gas lost prior to sealing the sample (the lost gas). The amount
of lost gas is dependent on the sample retrieval time, the particle size (diffusivity), temperature, total gas content, degree of
saturation of the sample, and nature of the drilling fluid (i.e. air vs. water). The most reliable data are assumed to be from
wire-line retrieved core because of rapid core recovery compared to conventional core and the finer particle size of cuttings
and sidewall core results in a greater proportion of gas lost prior to sealing the samples in the canister. The lost gas
component can vary from 5 to >50% of the total gas content and hence its measurement is critical for accurate gas-in-place
calculations and is the largest source of error in gas content determination. The theory and utility of the different methods of
lost gas calculation have been reviewed by Mavor et al. (1994), Mavor (1996), Diamond and Schatzel (1998) and Bustin
(2005). Four different methods are commonly applied to measuring lost gas volumes. They are:

1. USBM Direct Method


2. Smith and Williams
3. Amoco method
4. CBM Solutions method

The USBM Direct method is the most common method (Mavor 1996). The Smith and Williams method is particularly suited
for well cuttings and the CBM Solutions method is the most innovative. These methods are briefly outlined below.
The USBM Direct Method is the most widely used method of measuring lost gas. The method proposed by Kissell et al
(1973) uses two alternate calculations for lost time. If the well was drilled with water or other dense fluid, desorption is
considered to begin when the sample was half way to the surface. The total lost time is then:

Ttotal = (tsealed-tsurface)+(tsurface-tre)/2.................................................................................................................................................(3)

tsealed= the time core is sealed in canister


tsurface= time core reached the surface
tre= time sample retrieval begin

If the hole is drilled with air or mist, it is assumed that desorption begins at first penetration of the coal. The lost time is then
calculated as:

Ttotal = tsealed-tpenet...........................................................................................................................................................................(4)
tpenet= time coal first penetrated

The lost gas is estimated based on graphical extraction of the cumulative desorbed gas vs. the square root of desorption
time to the lost gas time. Since the extrapolation is based on the initial linear portion of the desorbed gas vs. square root of
time, it is critical that desorbed gas volumes be measured frequently and normally at the temperature of the drilling fluid or
reservoir.
The Smith and Williams method (Smith and Williams 1981) is a variation of the direct method developed for determining
gas content of drill cuttings. The Smith and Williams method is based on the surface time ratio (STR) and the lost time ratio
(LTR).

Surface time ratio = (ts-to)/ts..........................................................................................................................................................(5)


Lost time ratio = ts/t25......................................................... ..................................................(6)

based on time to reach surface (to), time sample to be sealed (ts) and time to desorb 25% of gas (t25).
The surface time ratio and lost time ratio provide a correction factor which when multiplied by the measured desorbed gas
is considered to yield the gas content. Because of the finer particle size of the cuttings, residual gas is not considered in the
calculations.
The Amoco method for estimating lost gas volume is based on:

6 2 D
Gc = ci 1 2 exp t GcL ............................................................................................................................. .......(7)
r2

Where Gc is the desorbed gas content, Gci, is the initial (or total) gas content (excluding residual gas), and GcL is the lost gas
content. The unknowns are determined using linear regression (McLennan et al. 1995).
6 SPE 131791

The CBM Solutions methodology (Bustin 2005) is based on the premise that it is impossible to determine the best
analytical or test a theoretical method for determining lost gas without knowing what the actual lost gas is. To develop a more
robust lost gas method, CBM Solutions experimented with cuttings, side wall core, conventionally retrieved core and wire-line
retrieved core for both coals and shale and developed specific sampling and lost gas protocols for each sampling technique and
lithology. In the laboratory experiments samples of different particle size are saturated in the lab and then desorption
simulations are carried out (Fig. 1). Thus it possible to design a lost gas procedure that fits the samples as well as access the
magnitude of the potential error in lost gas determination for varying lost gas times and sample types.
under estimation of lost gas
1 hour = -41%
.5 hour = -17%
150
cubic centimetres

100

50

0
-1.5 -1 -0.5
-2 0.5 1 1.5
0
2
-50
square root of time (hours)

Fig. 1 Lost gas simulation. Cuttings samples (.033-.132 inch equivalent diameter) are fully saturated with gas and then the
sample is desorbed with decreasing pressure and temperature to simulate sample recovery during drilling whilst
desorbed gas is quantified, This data shows that using the Direct Method of lost gas analyses with a 1 hour lost gas
time, the error in measurement is 41% and if lost gas time is 0.5 hours the error is 17%.

During the early time period the samples are typically maintained either at the reservoir temperature or mud temperature is
used during the collection of data during the lost gas period. The reasoning is that diffusion rates are temperature-dependent.
However these temperatures either over estimate or underestimate the actual average temperature history of the sample
following coring and sealing the samples in canisters. The most appropriate temperature values to use are well-specific and
must be calculated on site based on measurements of the mud and bottomhole temperature, mud circulation rate, size of
samples, recovery time and thermal diffusivity of the samples.
During the last ten years there has been growing acceptance of the use of cuttings samples from coal seams to determine
gas content. Early experiments such as reported Mavor et al. (1991; 1994) and summarized by McLennan et al. (1995) found
that cuttings samples yielded consistently lower gas contents than obtained from core samples from the same well bore. Such
comparisons are biased since cuttings samples obtained during coring operations are invariable much finer than those collected
during drilling. Studies by Bustin (2005) have shown that isolating and desorbing a representative size fraction of cuttings and
calibrating the lost curve with experiments with desorption experiments results in gas contents that are comparable to wire line
retrieved core on adjacent wells. Bell (2005) on behalf of the Energy Utility Board in Alberta carried out an independent basin
wide comparison of cuttings and core desorption tests using the methods described by Bustin (2005). Bell (2005) found that
cuttings samples compared favorably with wire-line retrieved core results. Cuttings are now used routinely for desorption in
the Powder River Basin and Western Canadian Sedimentary Basin for quantifying gas content.
Methods that have been attempted to directly measure coal gas content have met with either limited success or application.
One downhole technique that may hold promise for gas-in-place determination if certain ridged conditions are met is RAMAN
spectroscopy. RAMAN spectroscopy can be used to measure gas in solution (produced water) from which the partial pressure
of methane is obtained. If it assumed that the partial pressure of methane in the coal is equivalent to gas in solution, and if a
representative coal isotherm, is available the gas content of the coal can be determined (Lamarre and Pope 2007).
Other methods that have been proposed for many years (i.e. Hawkins et al. 1992) are inference of gas content in coal via
use of convnetional well logs. The density log for example can be used to infer ash content which in turn may be correlated
with gas content if it assumed (or measured) that gas content varies solely due to dilution by mineral matter (i.e. Bhanja and
Srivastava 2008). Such methods however require measured gas content data from the same coals for calibration for credible
results.
SPE 131791 7

Gas Deliverability Determination. Evaluation and understanding the deliverability of gas from coal requires quantitative
assessment of both matrix and coal seam flow properties. Measurements and modeling of coal seam deliverability requires a
mixture of well tests and laboratory analyses both of which have seen substantial refinement during the last decade.
Flow through coal is generally considered to occur by different mechanisms, depending on whether flow is through the
matrix or the fractures (Fig. 2).

Fig. 2 A conceptual model for the flow of gas and water through coal. Note single-phase flow of gas is thought to occur
through the coal matrix via the mechanism of diffusion, whereas water and gas flow through the fracture (cleat) by
Darcy flow.

We now discuss the transport mechanisms associated with the coal matrix and fractures, as well as the means for quantifying
flow at both scales.
Matrix Flow. Gas production in the past has been generally considered limited by fracture permeability (discussed below)
rather than matrix diffusion rates. Mavor (1996, p. 4.28) for example suggests that it is unlikely that reservoirs with
conventionally permeability suitable for economic production would be significantly hindered by low diffusivity. In practise
and theory however, whether or not economic production is rate limited by matrix diffusion or coal seam fracture permeability
will depend on effective fracture spacing which determines diffusion flow length, fracture permeability and diffusivity of the
matrix (Cui and Bustin 2006; Bustin et al. 2008). Although most coals have very regular cleat (Fig. 2) commonly spaced
much less than 1 cm, it does not necessarily follow that these fractures have effective permeability as supported by the
extremely low permeability of many coals with well developed cleat such as those of the Mannville Group in Alberta and the
Cameo coals of Colorado.
Movement of gas out of the coal micro- and transitional mesopores has long been considered controlled by diffusion (rather
than viscous- Darcy flow). Diffusion is generally modeled using Ficks Law, which assumes that mass flow rate is
proportional to the surface area and the diffusion coefficient of the material. Diffusion occurs in the direction of lower
concentration. Diffusion is generally considered (comprised of) in terms of three components (Cussler 1997):

1. Bulk diffusion - diffusion dominated by intermolecular interactions and encompasses the diffusion of one species
through different molecular species
2. Knudsen diffusion - diffusion where molecular and pore wall interactions are important
3. Surface diffusion - diffusion that occurs in the sorbed state fluid without mass transfer into the free gas state

Since the sorbed gas at the sorbed/free gas interface is assumed to be in equilibrium, the adsorption isotherm provides the
boundary relationship between sorbed and free gas. The earlist and simplest models (so called equilibrium models) of CBM
reservoirs ignore diffusion by assuming the gas concentration in the matrix (primary porosity) is equal to the sorption isotherm
storage capacity at the pressure of the free gas. The equilibrium model is effectively a single porosity model whereas more
recent and sophisticated models are dual or triple (bidisperse) porosity models.
8 SPE 131791

Dual porosity models take in to consideration the movement of gas from the primary (matrix) to secondary (fracture/cleat)
porosity systems. These models consider the movement of gas from the matrix to fractures as the source for the fractures.
The dual porosity formulation of gas production of Warren and Root (1963) modified to take into consideration matrix
diffusion (Mavor 1995) is as follows:

q gm = .2697V c cD G c Gs .............................................................................................................................................(8)

qgm = gas production rate from coal matrix (Mscf/D)


Vc = volume of coal matrix
= coal matrix shape factor, cm-2
D = diffusion coefficient, cm2/sec
Gc = average gas content of the matrix (scf/ton)
Gs = storage capacity of the coal matrix evaluated at the average pressure of the natural fracture system, scf/ton

The Warren and Root shape factor () is based on matrix geometry and requires a geometric length term, which
corresponds to the distance either between adjacent fractures or half the distance. Because the length term is difficult to
assess, it is more convenient to determine the effective diffusivity, which is the ratio of the diffusion constant to the square of
the average flow length. Diffusivity controls the rate of mass transfer from the matrix porosity to the secondary (fracture)
porosity. Recently Moro and Wattenbarger (2009) have reviewed shape factor values and formulations of different authors
that are commonly applied and test shape factors using numerical simulations and provide formulations for boundary
conditions of constant rate and constant pressure.
The dual porosity formulation of diffusion is still widely utilized however many recent studies have utilized a bidisperse
(Ruckenstein et al. 1971) diffusion model which is considered to more aptly represent the multiple pore sizes present in coal.
The bidisperse models currently in use have been recently summarized by Wei et al. (2007). The bidisperse models assume
two step gas diffusion with the micropore system dominated by surface diffusion and meso- and macropore system by pore
diffusion (Fig. 3). In the bidisperse representation, gas adsorption is in the micropores and mesopores and macropores provide
storage for free gas and the flow path between adsorbed gas stored in the micropores and fractures.
As a result of the interest in carbon dioxide sequestration in coal seams there has in the last several years been a large
amount of research as to the importance of diffusion, counter diffusion and adsorption kinetics on mass transport in coal. The
visualisation is that injected carbon dioxide into the fracture network results in counter diffusion and competitive adsorption in
the micro- and mesopores. Elegant mathematical formulations such as those of Yi et al. (2008) suggest that co-diffusion of gas
molecules facilitate gas transport in the presence of competitive adsorption whereas counter diffusion diminished mass
transport.

Equilibrium No diffusion
Single-porosity Simulator Cleat porosity
Approach

Non-equilibrium One-stage diffusion Micro-porosity


Dual-porosity Simulator
Approach Only micropore Cleat porosity
diffusion

Micro-porosity
Two=stage diffusion
Bidisperse pore
Triple-porosity Simulator
diffusion model Both micro- and Meso-/Macro porosity
macropore diffusion
Cleat porosity

Fig. 3 Comparison of major elements of diffusion models applied to coal (modified from Wei et al. 2007)

Measurement of Diffusion. Theoretical models of gas flux through the coal matrix described above have advanced rapidly
the last few years however there has been limited progress in our ability to measure matrix gas flux. Commercial reservoir
simulators for the most part utilize sorption time as a proxy for diffusivity of the coal matrix (Mavor and Nelson 1997).
Sorption time () is defined as the time required to desorb 63.2% of the gas from whole core desorption tests. Sorption time is
related to the diffusivity, particle shape and particle size.
SPE 131791 9

Most commonly diffusion is measured during canister desorption tests. From the Direct Method of lost gas calculation
(where desorption takes place at reservoir temperature):

D= bi .....................................................................................................................................................................(9)
{ }2
r2 2031 G ci

D= diffusion . coefficient cm2/sec


r2 = characteristic diffusion length cm
Gci= initial gas content (scf/ton)
bi = slope of desorbed curve vs. square root of time (scf/ton*hr.5)

The diffusivity can also be derived from the sorption time (Mavor 1996; King et al. 1986):

G
c = 1 e Dt
......................................................................................................................................................................(10)
G ci

where the characteristic sorption time is defined as:

1
= .....................................................................................................................................................................................(11)
D

Gc = cumulative gas desorbed (scf/ton)


t = time
Gci = initial gas content
= sorption time seconds
= coal matrix shape factor, cm-2

The coal shape factors (as summarized earlier) commonly used are 60 for cubes, 15 for spheres and 8 for cylinders.
The diffusion coefficient or diffusivity determined from sorption time from canister tests assumes the temperature,
diffusion length and diffusion rate of the canister samples is similar to that of the reservoir and desorption of gases occurs at
constant rate. Such assumptions however are probably rarely realistic. Coal generally has a well developed fracture fabric and
hence canister samples are rarely coherent samples with constant diffusion length. Additionally since coals are highly
compressible, assuming that diffusion at unconfined conditions in a canister is the same as coal confined in the subsurface is
probably rarely if ever realistic as suggested by permeability measurements of coal under varying confining pressures (ie.
Bustin 1997). Pulse-decay methods on confined cores at reservoir temperatures provides the most realistic conditions for
quantifying diffusion or permeability of tight rocks and coal, however methods in common practice use non reservoir gases
(generally helium or nitrogen) other than reservoir gas (Brace et al. 1968; Trimmer et al. 1980; Yamada and Jones 1980; Hsieh
et al. 1981; Bourbie and Walls 1982; Dicker and Smits 1988; Jones 1997).
In a recent paper Cui et al. (2009) discuss the failure of measurement of permeability and diffusivity in tight rocks, using
current method and provide new models for measurement of permeability and/or diffusion from pulse-decay type analyses of
crushed and confined samples, canister desorption and from adsorption isotherms. Cui et al. show that the use of helium as
routinely applied to measure porosity, permeability, and diffusivity result in non-systematic errors due to the molecular sieving
effect of the fine pore structure to larger molecules such as carbon dioxide and methane. Utilizing gases with larger adsorption
potentials than helium, such as nitrogen and all reservoir gases, to measure porosity or permeability of rocks with high surface
area is a viable alternative, but requires correcting for adsorption in the analyses. Effectively adsorption during pulse-decay
measurements of diffusivity or permeability comprises an additionally storativety term that most be incorporated in the
analyses (Fig. 4). Cui et al. (2009) provide new formulations of models that explicitly correct for adsorption during pulse-
decay measurements of core under reservoir conditions, as well as on crushed samples used to approximate permeability or
diffusivity. Cui et al. (2009) use gas adsorption isotherms to calculate an effective porosity due to adsorption which is equal
to:

s 1 ) qL pL
= . ...............................................................................................................................................(12)
a V std c g ( p L + p 2
(
)
10 SPE 131791

Where a is defined as the effective adsorption porosity, q L and p L are Langmuir volume and pressure respectively, s is the
skeleton density of porous samples, and Vstd is the molar volume of gas at standard pressure and temperature (i.e. 273.15 K
and 101325 Pa).

Fig. 4 Effective porosity, a (%), contributed by gas adsorption. The effective adsorption porosity is estimated from Eq. 12
without the factor of (1-) by assuming an ideal gas at T = 25C and s = 1.5 g cm3. (A) Langmuir isotherm volume is 1
cm3/ g and (B) Langmuir volume is 10 cm3/g. Modified from Cui et al. (2009).

Fracture Flow. Flow of water and gas through the coal natural fracture (cleat) system is often modeled using some form of
Darcys Law (pressure-driven flow). In some instances, when fracture spacing is dense, flow of gas and water to vertical wells
may be modeled using the radial inflow equation, assuming pseudo-steady state flow of gas and water through a single-
porosity system:

rg
[ R wf
]
qg = ............(13)
[
T ln(re / rw ) .75 + s + Dqg ]

( )
qw = rw R wf ..............(14)
1 4 12 w Bw [ln (re / rw]) .75 +
s

.
Eq. 13 and 14 serve to illustrate the key components of fracture flow:
1. Absolute permeability (k)
2. Relative permeability (krg and krw)
3. Skin
SPE 131791 11

4. Reservoir pressure p R [or m( p R ) ] and drawdown pR wf or [m( R


m( wf )]
5. Reservoir thickness
6. Fluid properties

Drainage radius, which is critical to reserves discussions below, is of lesser importance to fracture flow, given that it
appears in the natural log term. Although all 6 components are important for estimating flow through cleats, our discussion of
how to estimate these components will focus on 1, 2 and 3. The skin term in Eq. 13 and 14 is related the well completion and
stimulation method and is discussed in a later section.
Absolute permeability in coal is highly dependent upon the existence, frequency, orientation (relative to current in-situ
stresses), fracture height and degree of mineral in-filling (Laubach et al. 1998). A common model for describing cleat porosity
and permeability in coal is the matchstick model (Seidle 1992). The permeability is extremely sensitive to fracture aperture,
with which it has a cubic relationship. This model is simplistic because fracture apertures are assumed to be uniform, and
permeability is assumed to be isotropic. In reality, fracture permeability may be anisotropic due to the greater continuity of
face cleats relative to butt cleats, and the variation of cleat apertures. Typical permeability anisotropies observed in coal are
2:1 4:1, and this is expected to have some impact on development planning.
Any process acting to modify the cleat aperture will have a strong effect on absolute permeability. In coal reservoirs, there
are two physical processes that will act to change the physical dimension of the fracture apertures: 1) changes in effective
stress and 2) matrix shrinkage. With process 1), because the fracture pore volume is highly compressible, with pore volume
compressibilities typically on the order of 10-4 psi-1, increases in effective stress due to pore pressure depletion can cause the
fracture apertures to decrease in width, which in turn causes a reduction in absolute permeability. In some coal reservoirs,
process 2) will cause absolute permeability to increase with depletion, because the coal matrix will shrink during desorption,
causing an increase in fracture apertures. Analytical models that take into account both processes are reviewed in Palmer
(2009).
With gas adsorption, coal swells and with desorption coal shrinks and the degree of swelling or shrinking has been long
recognized to impact fracture permeability of coal. Studies such as those of Somerton et al. (1975), Durucan and Edwards
(1986), Gray (1987), Harpalani and Chen (1989), Palmer and Mansoori (1996), Shi and Durucan (2004), and Cui and Bustin
(2005, 2006) to name but a few, have investigated the degree of swelling and shrinkage of coal to various gases under various
reservoir conditions. These studies for the most part conclude that coal swelling is linearly proportional to volume of gas
sorbed and the impact on permeability of swelling (or shrinkage during production) is dependent on the mechanical moduli of
the coal. For linking coal volume change to change in reservoir permeability assumptions are required with respect to the
strain model. Cui et al. (2007) consider the impact of assuming uniaxial strain (purely vertical strain), plain strain (horizontal
strain predominant) and the general stress case and geomechanical properties on permeability. They suggest the uniaxial strain
and plain strain models are end member cases and actual strain conditions lie between them depending on local geology. In
the last few years coal swelling and shrinkage has been extensively measured and modeled from the perspective of carbon
dioxide sequestration and enhance coalbed methane production. Experimental studies universally show the degree of swelling
caused by various gases follows the sequence H2<N2<CH4<CO2<H2S<SO2 which compares with each gases relative
adsorption affinity (Fig. 5 and Fig. 6). Higher rank coals, which have higher adsorption capacities, show the greatest degree of
swelling (and shrinkage on desorption). The importance of coal composition on swelling independent of sorption capacity has
yet to be fully resolved. Cui et al. (2007), Pini et al. (2009) and Chikatamarla et al. (in press) describe experimental results in
which coal sample swelling and permeability and diffusion were monitored during flooding of coal core confined under
reservoir conditions. These studies quantify the reduction in permeability caused by coal swelling with the greatest
permeability loss being associated with the stronger adsorbing gases and mechanically weakest coals.
12 SPE 131791

40.0

35.0
Volume adsorbed (cc/g, daf) 30.0 H2
25.0 N2
CH4
20.0
CO2
15.0 H2S
10.0 SO2

5.0

0.0
0.0 1.0 2.0 3.0 4.0 5.0 6.0
Pressure (MPa)

Fig. 5 Comparison of adsorption of a semi-anthracite coal (Mist Mountain Formation, Alberta). From Chikatamarla et al. (in
press).

Fig. 6 Volumetric strain for 3 coal Canadian coals with N2, CH4, CO2 and H2S. The Ardley coal has rank equivalent to a vitrinite
reflectance of 0.46% and the Wolf Mountain and Quinsam coals have vitrinite reflectance values of 0.62%. From
Chikatamarla et al. (in press).

In addition to absolute permeability changes, relative permeability is known to have a strong impact on CBM production
profiles. As 2-phase CBM reservoirs are dewatered, the effective permeability to gas and water changes. The changes in
effective permeability that are fluid-saturation dependent are modeled using relative permeability curves, which are
notoriously difficult to obtain.
We now discuss methods for estimating absolute permeability, relative permeability and skin (or stimulation efficiency).
SPE 131791 13

Absolute Permeability Estimation. Absolute permeability in coal may be estimated from laboratory core tests, well test
analysis, advanced production data analysis and from reservoir simulation history-matching. At present, we believe the most
reliable and accurate method for establishing absolute permeability is from carefully designed well tests, preferably performed
during single-phase flow (water or gas) conditions, as can exist in undersaturated wells flowing above desorption pressure, or
saturated reservoirs with negligible mobile water production (dewatered coal wells, or dry CBM reservoirs). For completeness
we will briefly discuss each of these techniques, in approximate order of reliability.

1. Permeability Derived from Pressure Transient Analysis. A variety of conventional well test designs have been used
for CBM reservoirs including (pre-frac and post-frac) injection-falloff tests, (post-frac) flow/buildup, tank and slug tests
(Mavor and Saulsberry 1996). In addition to the examples provided by Mavor and Saulsberry, other examples of well test
applications that have appeared in the SPE literature in the past 20 years (only select references are given) are given in Table 2
(at end of paper).
In recent years, theoretical advancements have been made for analyzing before- and after-fracture closure data associated
with injection tests where injection rates are high enough to initiate and propagate a hydraulic fracture. The new test/analysis
procedures circumvent some of the limitations of conventional well test analysis techniques applied to data where frac
pressures have been exceeded during the test. Recently, pre-frac diagnostic injection tests (DFIT) have been used in tight gas
formations to obtain information for hydraulic fracturing operations (leakoff mechanism, closure time and pressure) and
reservoir pressure and permeability, the later coming from after-closure analysis, provided pseudolinear and pseudoradial flow
regimes are reached during the test (Barree et al. 2007). These tests have also been tried in coal reservoirs (Ramurthy et al.
2002). In the Ramurthy et al. study, the possible leakoff mechanisms for CBM reservoirs were reviewed, and both before-
and after-closure analysis was applied to Canadian and U.S. coal examples. The before-closure analysis requires rock property
and hydraulic fracture data that is not commonly available, and hence those results tend to be qualitative. Further, it is not
clear whether a hydraulic fracture is actually created in some coal DFIT tests, or the cleat system is simply inflated and fluid is
propagated along an existing cleat system with no new fracture created (R.D. Barree, personal communication). The hydraulic
fracture network may also be complex [see below and Palmer et al. (1993)]. Finally, after-closure analysis requires
pseudoradial flow to develop, which may not occur in a reasonable timeframe for low-permeability coals.
In addition to new theoretical developments, some advancement in well test implementation technology has also been
made. For example, wireline-conveyed formation evaluation tools have recently been used (Schlachter 2007) to isolate coal
intervals and perform drawdown-buildup tests that can be analyzed using CBM PTA techniques.
There are several reservoir complexities associated with coalbed methane that serve to complicate design and
interpretation of well tests:
Gas desorption (saturated CBM reservoirs or undersaturated reservoirs below desorption pressure)
Multi-phase flow (saturated CBM reservoirs or undersaturated reservoirs below desorption pressure, 2-phase flow
of gas and water)
Multi-layer behavior
Additional reservoir heterogeneities: dual porosity behavior, composite behavior
Non-static absolute permeability (dependence on effective stress and/or adsorption/desorption)

An excellent discussion of the theory associated with CBM well test analysis, and design and execution of single- and
multi-well tests was provided by Mavor and Saulsberry (1996). In that work, the authors address how to account for the
effects of desorption using a desorption compressibility term (Eq. 15), and multi-phase flow using a total fluid flow rate (gas +
water) and either a multiphase total mobility (which assumes negligible fluid saturation gradients) or multiphase flow potential
(Eq. 16).

Bg cV L p L
cd = ..............(15)
32.0368(p L + p R)
2

p k ( p) ( )
M (p ) = rg + k rw p dp ...............(16)
pb B
g g wBw

Kamal and Six (1989) also suggested the use of a modified pseudopressure function to account for multi-phase flow. Use
of multi-phase flow potential requires relating relative permeability to gas and water, which are functions of fluid saturation, to
pressure; Mavor and Saulsberry suggest different approaches for doing this one of the approaches mentioned was to
calculate relative permeability ratio as a function of time, using Eq. 17 below, then estimating estimating fluid saturation as a
function of time by interpolating the values from a table containing relative permeability ratios/relative permeability to gas and
water as a function of saturation. The saturation values are then assumed to be a linear function of pressure; multi-rate tests
14 SPE 131791

must be performed to develop this relationship. As recognized by Mavor and Saulsberry, the calculations are sensitive to the
shape of the relative permeability curve, which must be assumed or derived from other sources. Further, the relationship
between saturation and pressure may not be linear, particularly for longer tests.

,1000 qg
g Bg
krg .5 6146
= .....(17)
krw qw w Bw

Mavor and Saulsberry (1996) also discuss approaches for designing and analyzing tests where multi-layer behavior occurs
the use of production logging to estimate fluid flow rates from each zone, as well as gas-water ratio (based upon
gradiomanometer data) was mentioned. As with single-layer analysis, the individual zone gas-water ratio data may be used to
derive relative permeability ratios, which in turn can be used to derive fluid saturations and relative permeability to each phase
by interpolating a table with relative permeability data. Multi-layer testing in single-phase reservoirs was recently discussed
by Bastian et al. (2005) and Clarkson (2009), using the Horseshoe Canyon play in Western Canada as an example. In the
complex multi-seam Horseshoe Canyon coals, operators will often periodically run spinner surveys prior to isolating the coal
seams or zones with bridge-plug assemblies and packers with tandem gauges hung beneath them. The spinner surveys are run
to assist with allocation of production back to the individual seams or coal zones; the individual zone/seam buildups are
analyzed from reservoir pressure, kh and skin. A typical configuration for isolating zones/seams is given in Fig. 7.
In addition to multi-layer behavior, additional heterogeneities in the coal may complicate CBM welltest interpretation, so it
is recommended that a software package with the ability to model and interpret a variety of possible reservoir types is selected
for analysis. For example, again using the Horseshoe Canyon coal as an example, multi-layer and composite reservoir
characteristics are quite common, and rarely (in our experience), dual porosity behavior (Fig. 8). We selected the Horseshoe
Canyon coal to illustrate this because portions of the play exhibit primarily single-phase (gas) flow as noted by Mavor and
Saulsberry (1996), interpretation of reservoir heterogeneities in coal can be complicated if multi-phase flow effects occur.

perfs

Fig. 7 Schematic diagram illustrating a multi-zone well test in a Horseshoe Canyon CBM well. Modified from Bastian et al.
(2005)
SPE 131791 15

Selected model: Dual porosity (one possible interpretation)

Fig. 8 Modeling of CBM heterogeneities using commercial well test software. Data is from a pressure-buildup performed on
a single coal-seam in the Horseshoe Canyon play. Feketes F.A.S.T. WellTest T M software package was used for
analysis. Modified from Clarkson (2009).

The impact of non-static absolute permeability on the well testing strategy and analysis has also addressed in detail in
recent history. Absolute permeability in coal is known to change during primary depletion operations, either due to changes
in effective stress, which act to reduce permeability, or desorption (commonly referred to as matrix shrinkage), which acts to
increase permeability during depletion. Which effect has the dominant control on permeability change depends upon the level
of depletion of the coal, coal mechanical properties, coal fabric, in-situ stress conditions etc. An early study by Schwerer and
Pavone (1984) investigated the effects of pressure-dependent permeability in coal on conventional well test analysis of single-
phase, constant production or injection rate or constant pressure (simulated) data, and suggested that a series of constant-rate
tests be performed to ascertain the pressure-dependence of permeability. In the past decade, PTA methods have been used to
quantify absolute permeability changes in coal during depletion. Mavor and Vaughn (1998) observed that well test data could
be used to quantify permeability gains due to matrix shrinkage effects in the Fruitland coal of the San Juan Basin; absolute
permeability estimated from DSTs were compared to estimates from shut-in tests performed later in the life of the wells. More
recently, Gierhart et al. (2007) performed timed-lapse PTA on Fruitland coal wells within and outside the Fruitland coal
fairway in the San Juan Basin (Colorado portion); multiple shut-in/buildup tests were performed during the production life of a
well, resulting in multiple estimates of permeability which could then be plotted versus reservoir pressure, also estimated from
each test. These time-lapse pressure transient analyses were then analyzed for functional form, which have been used to
validate analytical models developed to predict permeability changes in coal (ex. Palmer et al. 2007, Clarkson et al. 2008b).
Example results from a time-lapse PTA are given in Fig. 9.
100.0

y = 5.3581E+02e-6.5340E-03x

10.0
Permeability, md

y = 4.8069E+02e-8.9604E-03x
1.0

0.1
800 750 700 650 600 550 500 450 400
Reservoir Pressure, PSIA

Effective Permeability to Gas Absolute Permeability Expon. (Absolute Permeability) Expon. (Effective Permeability to Gas)

Fig. 9 Example of estimated (effective and absolute) permeabilities as a function reservoir pressure obtained from time-
lapse PTA of a Fruitland coal well (used with permission from Roger Gierhart). Effective permeabilities (red dots)
obtained from the analysis were converted to absolute permeabilities (black dots) using estimated saturations and
relative permeability curves (Gierhart et al. 2007). Note that the trend in the data has been assumed to be exponential
with pressure.
16 SPE 131791

In addition to quantifying permeability changes during depletion, Mavor and Gunter (2006) used well testing to develop a
method for predicting porosity and permeability changes as a function of fracture pore pressure and gas content and
composition in support of enhanced coalbed methane recovery (ECBM) and CO2 sequestration projects. The approach used
was to 1) measure the gas composition and initial permeability of the reservoir, preferably using a production test followed by
shut-in test; 2) estimate the amount of pressure-induced strain using an injection test with water and/or a weekly adsorbing gas
(ex. N2); and 3) estimate the amount of sorption-induced strain using an injection test with a strongly-adsorbing gas. One of
the unique aspects of this study is the attempt to separate relative permeability effects from absolute permeability changes.

2. Permeability Derived from Production Data. A relative newcomer to the techniques available for permeability
estimation of coal, type-curve and straight-line methods (analogous to those methods used for pressure-transient analysis) have
recently been used for analyzing dry (single-phase gas) and 2-phase (gas+water) CBM reservoirs. Use of these techniques are
subject to the same (reservoir-related) complications as well test analysis; a summary of the historical work on CBM
production analysis was provided by Clarkson et al. (2007) and is not repeated here.
As discussed in Clarkson et al. (2009), advanced production analysis techniques for CBM reservoirs can be classified as: 1)
type-curve methods; 2) straight-line methods; 3) analytical and numerical simulation; and 4) empirical methods. In this
section, we only discuss the first 2 methods.
Type-curve methods involve matching production data (normalized for back-pressure variations) to dimensionless
analytical/empirical solutions reservoir and stimulation properties can be extracted from the match. For CBM, the
dimensionless variable definitions require pseudovariables that account for both pressure- and saturation-dependent reservoir
properties (+ desorption), and superposition time functions to account for variable rate/flowing pressure data. Use of these
modified variables, in combination with dimensionless variables developed for conventional reservoirs, has allowed CBM
production/flowing pressure data to be analyzed in an analogous fashion to conventional reservoirs. Both single-phase (gas)
and 2-phase (gas+water) CBM wells have been analyzed using this approach. Because 2-phase CBM reservoirs require an
estimate of relative permeability, Clarkson et al. (2009) recommended that type-curve analysis be performed in parallel with
analytical and numerical simulation such that the output from simulation history-matching may be used for pseudovariable
calculation during type-curve matching agreement between type-curve and simulation-derived permeability and stimulation
values should provide for more unique estimates. Single-phase (gas or water) analyses generally are more straight-forward
except in cases where non-static permeability effects occur, or reservoir heterogeneities (such as multi-layer) exist.
Straight-line techniques for production analysis are analogous to those used in pressure transient analysis; once the correct
flow regime is identified using derivatives or other diagnostic tools, data corresponding to that flow regime is plotted on a
specialized Cartesian plot designed to linearize the dataset the y-axis variable is typically rate-normalized pressure or
pseudopressure-drop, and the x-axis is superposition time or normalized cumulative production. For transient flow analysis,
permeability can only uniquely determined if a radial flow period is apparent, from which permeability can be extracted from
the radial flow Cartesian plot (Hager and Jones 2001). For boundary-dominated flow, permeability may be extracted from the
flowing material balance plot, if skin is known. To date, transient flow straight-line techniques have only been applied to dry
coal. The flowing material balance technique, which is applicable to boundary-dominated flow only, can be used for single-
phase and 2-phase flow CBM reservoirs (Clarkson et al. 2008a).
Two examples of the use of advanced production analysis methods for permeability estimation in coal are given below
these examples are considered by the authors to represent the state-of-the-art in production analysis for CBM reservoirs. In the
first dry coal case (Fig. 10), a rigorous, integrated methodology for estimating permeability from production data was applied.
Derivatives are used to identify the radial and boundary-dominated flow regimes (Fig. 10a); the radial flow plot is used to
estimate permeability (actually kh) as is the type-curve match. Once skin is estimated from radial flow analysis and type-curve
match, the skin may used to estimate permeability from the flowing material balance plot. Note that permeability values
estimated from the three analysis techniques are consistent (Clarkson 2009).
SPE 131791 17

a) Derivative Plot b) Radial Flow Plot

1.0E+5 3.0E+04

Radial flow 2.5E+04


(zero slope)
d(m(p)/q)/dlog(tca*)

[m(pi)-m(pwf)]/qg
1.0E+4 2.0E+04

Boundary- 1.5E+04
dominated
1.0E+3 flow (unit 1.0E+04
slope)
5.0E+03

1.0E+2 0.0E+00
1.0E+0 1.0E+1 1.0E+2 1.0E+3 1.0E+4 0 2 4 6
tca*, days Superposition Time

c) Fetkovich/Blasingame Type-Curves d) Flowing Material Balance

10.00
1.4E-1

1.2E-1

Normalized Rate, scf/D/psi2/cp


1.0E-1
1.00

8.0E-2
qDd

6.0E-2

0.10 4.0E-2

2.0E-2

0.0E+0
0.01 0 50 100 150
0.0001 0.001 0.01 0.1 1 10 Normalized Cumulative Production, MMscf
tDd

Fig. 10 Analysis of Horseshoe Canyon CBM well for permeability: (a) identification of flow regimes using derivative plot, (b)
radial flow analysis, (c) Fetkovich type-curve match, (d) flowing material balance analysis. Modified from Clarkson
(2009).

In the 2-phase analysis example (Fig. 11), an integrated analysis approach is also applied in which straight-line and type-
curve methods are used in parallel, but in this case, the fluid saturation and pressures obtained from simulation matching are
used to calculate the pseudovariables in the dimensionless terms for the type-curves. This approach was suggested by
Clarkson et al. (2009) and is believed to yield a more unique analysis. The permeability obtained from simulation and type-
curve analyses are in agreement; the skin values calculated from these methods were then used to calculate permeability from
flowing material balance analysis.
In these two examples, reservoir heterogeneities such as multi-layer effects and dual porosity behavior were not
encountered. Dual porosity effects have not been observed in production signatures for CBM reservoirs analyzed by the
authors to date; multi-layer effects are known to occur in some CBM reservoirs, and Clarkson (2009) suggests a procedure for
analyzing such reservoirs, including a layer flowing material balance method. If non-static permeability effects occur, the type-
curve and straight-line methods may be altered to account for these effects, as discussed in Clarkson et al. (2009).
18 SPE 131791

a) 2-P CBM Forecast b) Fetkovich/Blasingame Type-Curves

10.00
1000 1000

1.00
qg (Mscf/D)

qw (STB/D)
100 100

qDd
0.10

10 10
0 100 200 300 400 500 0.01
time (days) 0.0001 0.001 0.01 0.1 1 10
Modelqg Actualqg Actualqw Modelqw tDd

c) Pratikno-Blasingame Type-Curves d) 2-P CBM Flowing Material Balance

5.0E-02

Normalized Rate, scf/D/psi2/cp


4.0E-02
10.00

3.0E-02
qDd

1.00
2.0E-02

0.10 1.0E-02

0.0E+00
0.01 0 200 400 600 800 1000 1200
0.001 0.01 0.1 1 10 Normalized Cumulative Production, MMscf
tDd

Fig. 11 Match of Uinta Basin CBM well gas and water rate data using (a) analytical simulation model and gas rate data on the
Fetkovich type-curves (b), Pratikno-Blasingame type-curves (c), and flowing material balance plot (d). Modified from
Clarkson et al. (2009).

Production analysis methods have been used by the authors to estimate permeability at well locations that are then mapped
for use as a starting point for field reservoir simulation. In Fig. 12, effective permeability to gas (kg) at 600 psia was estimated
using the technique described in Clarkson et al. (2008a) and mapped. kg at 600 psia was selected as the base permeability in
relative permeability calculations because below 600 psia, the kg changes were assumed to be dominated by absolute
permeability changes and above 600 psia, kg was assumed to be a function of both fluid saturation and pressure. The resulting
permeability map generated from production methods proved to be very useful for simulation work. The map in Fig. 12 was
generated using Microsoft Excel.
Production data analysis has also been used to quantify permeability changes as a function of depletion in an analogous
fashion to the time-lapse pressure transient analysis methods discussed above. Clarkson and McGovern (2003) used
production data, flowing- and shut-in pressure and PVT data in combination with the radial flow equation to solve for effective
permeability to gas (kg), which was then plotted as a function of reservoir pressure. Gierhart et al. (2006) performed a similar
analysis, but solved for the backpressure coefficient (C) as a function of pressure using Eq. 18. In the Gierhart et al. (2006)
study, changes in C were assumed to be primarily due to absolute permeability changes in the Fruitland coal fairway.
Consistent with the time-lapse PTA studies of Zahner (1997) and later Gierhart et al. (2007), the trend in permeability growth
with pressure at low-pressures in the Fruitland coal fairway were shown to be exponential in nature. Permeability changes
from production analysis were later used to calibrate analytical models for permeability change, as discussed in Clarkson et al.
(2008b), and were incorporated into advanced production data analysis methods (Clarkson et al. 2008a, 2009).

qg .............(18)
C=
(p R2 wf )
n
SPE 131791 19

72-78
66-72
60-66
54-60
48-54
42-48
36-42
30-36
24-30
18-24
12-18
6-12
0-6

Fig. 12 Map of effective permeability (md) to gas (at 600 psia) generated using permeability estimates at well locations from
production analysis. Modeled area is approximately 72 square miles.

Reservoir simulation history-matching is a form of advanced production data analysis, and may also be used to estimate
absolute permeability in coal, provided robust estimates of other reservoir and stimulation properties are available. Our
preference is to obtain estimates of absolute permeability using PTA, or the methods described above, as a starting point in
field simulation; the permeability values, or trend in these values with depletion, are then tweaked to fine-tune matches to
well dynamic data.

3. Permeability Derived from Core Analysis. There are several complications leading to potential inaccuracies
associated with absolute permeability estimation from core (Mavor 1996): obtaining a representative sample of the reservoir;
obtaining a competent sample for analysis; recreating in-situ stress conditions; and recreating in-situ fluid saturations. A
primary concern is whether or not the absolute permeability obtained from core measurement is representative of the reservoir
contacted by producing wells; larger scale heterogeneities not present in the core, such as joints, may contribute greatly to flow
to wells, but often will not be represented in the core. Further, highly-fractured coals, which usually represent the more
permeable reservoir (unless the fractures are mineral-filled), are also very friable, meaning they may not be preserved for
analysis; more competent coal, which often has a higher mineral matter content with fewer fractures, generally will represent
lower permeability reservoir, hence core analysis may be biased towards lower permeability zones of the reservoir.
Hydrostatic testing at various confining pressures is often performed in an attempt to represent more realistic reservoir
conditions, but it is believed that the uniaxial stress condition is more representative. Mavor (1996) summarizes additional
measurement-related complications associated with core-derived permeabilities. Based upon core analysis performed up to the
mid-90s, Mavor recommended that core-derived permeabilities not be used to forecast production from CBM wells.
In the past several years, there have been numerous studies attempting to determine the impact of effective stress and
sorption-induced strain on permeability estimation. This work has been done in response to interest in the use of non-
hydrocarbon gases such as CO 2 and nitrogen for enhanced recovery of coalbed methane (ECBM) and CO 2 sequestration.
Table 3 (at end of paper) provides examples of core-tests for permeability that have appeared in the literature in the past 5
years (only select references are given). Most examples are from the SPE literature; the authors note that there have been
many more studies appearing in other technical journals. Many of these tests have been performed using hydrostatically-
stressed cores, and make little attempt to reproduce in-situ (stress, fluid saturation etc.) conditions. One exception is a recent
study by Mitra et al. (2008) who attempted to reproduce in-situ conditions using triaxially-tested stressed cores (at reservoir
temperature and moisture conditions) - permeability to various gases at varying pore-pressures were measured.
20 SPE 131791

In addition to quantifying the effects of stress and sorption-induced strain on coal permeability, core-scale tests have also
been used to quantify the effect of other coal properties (such as lithotype composition) on permeability. For example,
Clarkson and Bustin (1997) measured coal lithotype permeabilities using a non steady-state probe permeameter the finely
laminated nature of the coals tested necessitated that probe measurements be used to quantify lithotype controls on
permeability instead of full-diameter core tests. Differences in lithotype permeabilities were noted with bright coals having the
highest permeability (because of associated macro-fracturing), compared to other lithotypes. Additional testing would be
required to correct the profile measurements to in-situ conditions. Wold and Jeffrey (1999) noted cm-scale heterogeneity in
permeabilities derived from coal cores subjected to effective stress, and quantified differences in dull and bright coal
permeability.
There appear to be few studies where core-based estimates have been compared to field-derived estimates. One exception
is the work of Wold and Jeffrey (1999), who compared core- and well test derived results for Sydney Basin coals; in that
study, core-derived permeabilities compared favorably in magnitude to well test derived results, although max/min
permeabilities orientations did not agree. Also, Enever and Hennig (1997) compared laboratory- and welltest-derived
permeabilies for Australian coals; the relationship between permeability and effective stress for lab and field data was
qualitatively similar. Further work of this nature will be required to ascertain the usefulness of core-derived permeabilities for
forecasting wells in the field; it is clear however that core-based studies are useful for ascertaining the mechanistic controls on
permeability changes in coal reservoirs.

Relative Permeability Estimation. As with absolute permeability estimation, relative permeability in coal has been
quantified from core analysis, production analysis, and simulation history-matching. Unfortunately, far less work has been
done to measure relative permeability in coal in the past 20 years, despite the important control relative permeability has on
CBM production characteristics. Ham and Kantzas (2008) recently summarized the findings of relative permeability studies
in coals; 9 studies since 1973 were cited for comparison, we have summarized the results of 9 studies focused on core-based
absolute permeability measurements since 2005 (Table 3). The reason for the lack of core-based measurements of relative
permeability is related to the difficulty of the measurements, as well as the difficulty in obtaining and preserving samples.
Mavor (1996) summarized some core-based studies performed up to that point and discussed some of the nuances of
measurement and reporting. Mavor also summarized the procedures associated with steady-state and unsteady-state
measurements. The properties of coal that make relative permeability measurement difficult include (Ham and Kantzas):
friable and heterogeneous nature of coal, which makes it difficult to obtain a representative sample, low porosity of fracture
network, stress-dependence of permeability, and adsorption of gases onto coal.
Additional studies not summarized by Ham and Kantzas (2008) include Conway et al. (1995), Clarkson et al. (2007) and
Shedid and Rahman (2009), the latter of which post-dated the paper by Ham and Kantzas. Shedid and Rahman (2009)
provided relative permeability data for Australian coals (derived from Johnson, Bossier and Nauman method). The Conway et
al. (1995) study is noteworthy not only because relative permeability data for Blue Creek coal (Oak Grove mine, Black
Warrior Basin) was reported (measured at Marathon Oil Companys Research Center), but also because a quartz micro model
was used to visually investigate the effects of experimental factors on data gathered. Further, those authors suggested that
relative permeability can be derived in the field using well test analysis, analogous to what has been done for solution gas
drive reservoirs (ex. Al-Khalifah et al. 1987). They note that relative permeability curves derived from reservoir simulation
history-matching is problematic because all other reservoir properties input into the simulation (absolute horizontal and
vertical permeability, porosity) as well as reservoir description and near wellbore characterization, must be accurate. Finally,
they noted inconsistent results in absolute permeability estimation from well tests if relative permeability curves derived from
another formation/basin and simulation history-matching was used in the analysis. Conway et al. (1995) highlight many of the
issues with the status of relative permeability estimation in coal that do not appear to have been addressed in the literature
effectively in the last 15 years.
Clarkson et al. (2007) suggested that relative permeability in coal can be obtained from production analysis by solving the
radial flow equation to gas and water for effective permeability to those phases, then selecting a base permeability (in their
study, base permeability was selected as the effective permeability to gas at a point when relative permeability effects were
considered negligible i.e. when the coal was dewatered). Mobile water saturation was determined from water material
balance. The approach (with some procedural differences) is analogous to that suggested by Fetkovich (1986) for obtaining
gas/oil relative permeabilities. The approach has been used for prolific (saturated) CBM wells in the Fruitland coal fairway
where the transient flow period is short, and the coal reservoir exhibits tank-type behavior. The approach will be problematic
in low permeability CBM reservoirs exhibiting lengthy transient flow periods (Oswaldo Nunoz, personal communication).
This is illustrated in Fig. 13 where relative permeabilities were obtained using a modified version of the Clarkson et al. method
a tank-model was used to history-match production profiles obtained from another tank-model and from a numerical
simulator, and the resulting water saturations and reservoir pressures were used to derive the relative permeability curves from
solution of the radial flow equation. Absolute permeability was assumed to be static in the derivation and effective
permeability to water and gas were converted to relative permeability using this fixed absolute permeability. The simulated
cases were presented previously (Clarkson et al. 2009). In Fig. 13a, the output from a commercial tank-model simulator was
matched with a spreadsheet tank-model (described in Clarkson and McGovern 2005) the relative permeability curves input
into both models were straight-lines. Using the radial-flow equation to solve for effective permeability, the straight-line
SPE 131791 21

relative permeability curves were reproduced this is because the tank model has no associated transient flow. In Fig. 13b,
the output from a numerical simulator was analyzed in a similar way note that a transient flow period is modeled with the
numerical simulator and that the derived relative permeability curves differ from the model input values during the transient
flow period during boundary-dominated flow, the derived and input relative permeability curves (straight-lines) appear to be
in reasonably good agreement with some deviation due to model error. We will continue to investigate these effects and
report the results in a future paper.

a)1 b) 1
0.9 0.9
B-D Transient
0.8 0.8
0.7 0.7
0.6 0.6
Krg or Krw

krg or krw
0.5 0.5
0.4 0.4
0.3 0.3
0.2 0.2
0.1 0.1
0 0
0% 20% 40% 60% 80% 100% 0% 20% 40% 60% 80% 100%
Sw Sw

Fig. 13 Derivation of relative permeability curves using output from pseudo-steady state (tank) model (a) and numerical
simulator (b). The modified Clarkson et al. method works well for pseudo-steady state conditions (a), as evidenced by
the reproduction of the (input) straight-line relative permeability curves. For the numerically-simulated case (b), the
method fails to reproduce the straight-line curves during transient flow but works well for boundary-dominated (b-d)
flow. Solid lines are derived curves, and dashed are input into the simulator (blue = water relative permeability, red =
gas relative permeability). Normalized water saturation is used in the plots.

CBM reservoirs are fractured reservoirs, with the bulk of the fluid flow to the well occurring through the fractures straight-
line relative permeability curves are expected for fractured reservoirs yet CBM reservoir relative permeability curves derived
from various sources to date do not show this characteristic shape. There are many controls on the shape of derived relative
permeabilities in coals; if production analysis is used to derive the curves, in addition to transient flow, gravity segregation,
multi-layer effects, and non-static absolute permeability may all affect the shape of the derived relative permeability curves.
For example, in simulated cases (using tank models), we have found that absolute permeability changes have a profound effect
on the shape of the derived relative permeability curves. To illustrate this, we generated production profiles using the same
tank model inputs as in Fig. 13a (Table 7 and Fig. 10 of Clarkson et al. 2009), except the absolute permeability was allowed to
grow exponentially with a decrease in reservoir pressure (Fig. 14a). The relative permeability curves were derived using the
approach above (Fig 14b), with base permeability corresponding to the maximum effective permeability (kg) calculated from
the last time step. Note that we have varied permeability and not fracture porosity in this example changes in fracture
porosity will also affect fluid saturations (we will include these changes in future studies). The derived curves are very
different from the straight-line relative permeability curves input into the model the conclusion is that that the relative
permeability curves may indeed be straight-line curves (as observed in fractured reservoirs), but unique CBM reservoir
properties (in this case non-static absolute permeability) may cause the derived curves to differ from actual. Interestingly, the
derived curves in this simple example are qualitatively similar to relative permeability curves derived by Clarkson from field
data in the Fruitland coal fairway. If absolute permeability is allowed to decrease exponentially, derived relative permeability
curves (using maximum effective permeability, which is kw at initial conditions) bend in the opposite direction (concave down
for gas, concave up for water). The shape of the derived relative permeability curves may therefore be used to infer the
direction of absolute permeability changes, if the actual relative permeability curves are assumed to be straight-lines. If the
derived relative permeability curves are used in simulation history-matching, then absolute permeability should not be altered
during the deletion period covered by the curves as it is already built-in to the relative permeability curves. We plan to do
many more sensitivities of this nature to establish additional controls on the shapes of derived relative permeability curves.
Admittedly there may be many contributing factors to the non-linear nature of derived CBM relative permeability curves.
22 SPE 131791

a) b) 1
100 0.9
0.8
10 0.7
0.6

Krg or Krw
k/ki

1 0.5
0.4

0.1 0.3
0.2
0.1
0.01
0
100 200 300 400 500 0% 20% 40% 60% 80% 100%
Reservoir Pressure (psia) Sw

Fig. 14 Derivation of relative permeability curves using output from pseudo-steady state (tank) model with exponential
permeability growth (a). When absolute permeability growth occurs, the modified Clarkson et al. method yields
relative permeability curves that are very different than the straight-line curves that are input into the model (b). In Fig
14b, solid lines are derived curves, and dashed are input into the simulator (blue = water relative permeability, red =
gas relative permeability). Normalized water saturation is used in (b).

CBM Completion and Stimulation Efficiency Analysis. A variety of wellbore architectures, completion and stimulation
designs have been used to extract CBM (Palmer 2010). Until recently, the most popular approach to CBM exploitation was to
use vertical wells completed in single- or multiple coal seams and stimulated with hydraulic fracturing, or openhole (natural or
cavity completions). As with absolute permeability estimation discussed above, conventional well test techniques can be
used, after accounting for unique CBM reservoir behavior, to estimate hydraulic-fracture properties (half-length and
conductivity) or effective wellbore radius in the case of cavity-completed wells. Mavor and Saulsberry (1996) noted that type-
curves developed for hydraulically-fractured wells or non/slightly-stimulated wells completed in conventional reservoirs may
be used for this purpose, and provided several examples. Recently, Clarkson (2009) and Clarkson et al. (2009) noted that
hydraulically-fractured well type-curves developed for conventional or tight gas reservoir production analysis may be used for
CBM, provided that CBM reservoir characteristics are accounted for. For CBM wells exhibiting multi-phase flow, Clarkson
et al. 2009 recommended that type-curve methods be used in parallel with reservoir simulation to improve the uniqueness of
the analysis results. As an example application of production type-curve analysis, the (dry) Horseshoe Canyon well example
given in Fig. 9 was also analyzed using type-curves designed for hydraulically-fractured wells (Fig. 15) the resulting
fracture half-length (assumed infinite conductivity) is ~ 1 ft, which is consistent with the stimulation method of nitrogen
injection with no proppant, designed to slightly-stimulate wells, or at least reduce the effects of drilling-induced damage.
Radial flow type-curves were also used for analysis (Fig. 10).

Pratikno-Blasingame Type-Curves

10.00

1.00
qDd

0.10

0.01
0.001 0.01 0.1 1 10
tDd

Fig. 15 Match of Horseshoe Canyon CBM well data using type-curves designed for hydraulically-fractured wells (Pratikno and
Blasingame 2003).

In the past decade, horizontal wells (single- and multi-laterals) have become a popular method for exploiting some CBM
reservoirs. Many of the wells completed have been openhole, natural completions to estimate the effective (contributing)
well-bore length, welltest and advanced production analysis methods may be used. Clarkson et al. (2009) provide examples of
the use of straight-line and type-curve methods for analyzing dry and two phase (gas+water) horizontal (non-stimulated) CBM
SPE 131791 23

wells. Very recently, some companies have been using multi-fractured horizontal wells to exploit low-permeability CBM
reservoirs, consistent with what is currently being done to develop shale gas reservoirs. To our knowledge, no examples yet
appear in the literature of the analysis of CBM wells exploited in this manner. We anticipate that well test and production
analysis techniques currently being developed for shale gas reservoirs may be used to analyze CBM reservoirs exploited with
multi-fracd horizontal wells.
We have started generating simulation cases of multi-fractured horizontal CBM wells to determine if straight-line and type-
curve methods may be used to analyze them, analogous to tight gas or shale gas wells (see Clarkson and Beierle 2010). We
provide one of our initial test cases below. A simulated multi-transverse fracture (cased hole) dry (single-phase gas) CBM
example is analyzed (Fig. 16) we used the same well geometry/hydraulic fracture geometry and conductivity as in a case
previously analyzed as a tight gas example (Case 1 in Clarkson and Beierle 2010) for comparison purposes, although we
realize that the modeled hydraulic-fracture half-lengths (450 ft) may not be realistic. In this example, we follow the analysis
procedures described in Clarkson and Beierle (2010) the reader is referred to that work for detailed analysis procedures.

TABLE 4 INPUT PARAMETERS FOR SIMULATED


HORIZONTAL WELL (SINGLE-PHASE GAS, CBM)
Input Parameter Parameter Value
THICKNESS (FT) 50
BULK DENSITY (G/CM3) 1.33
CLEAT POROSITY (%) 0.5
GAS GRAVITY 0.6
VERTICAL PERMEABILITY (MD) 0.01
HORIZONTAL PERMEABILITY (MD) 0.1
INITIAL RESERVOIR PRESSURE 350
(PSIA)
RESERVOIR TEMPERATURE (F) 70.6
LANGMUIR VOLUME (SCF/TON, 250
IN-SITU) Fig. 16 Snapshot of simulator gridblock pressure illustrating
LANGMUIR PRESSURE (PSIA) 661 the early linear flow regime analyzed with straight-
line analysis.
DRAINAGE AREA (ACRES) 150
WELLBORE DIAMETER (IN.) 7.875
HORIZONTAL WELL-LENGTH (FT) 2000
TOTAL FRACTURE HALF-LENGTH
(FT) 2250
# HYDRAULIC FRACTURES 5
INDIVIDUAL HYDRAULIC
FRACTURE HALF-LENGTH (FT) 450
HYDRAULIC FRACTURE SPACING 500
FLOWING BOTTOMHOLE
PRESSURE (PSIA) 50

The first step in the analysis is to identify flow regimes using derivative or other diagnostic methods. The radial and linear
flow derivatives derived from the simulated production data are given in Fig. 17. In this horizontal well example, only the
early-linear and boundary-dominated flow regimes are evident from derivative analysis. The hydraulic fracture tips are
evidently too close to the reservoir boundary to observe any of the other expected flow regimes such as early-radial, late-
linear, or late-radial flow (Clarkson and Beierle 2010).
24 SPE 131791

a) Radial Derivative Plot b) Linear Derivative Plot

1.0E+06 1.0E+04

d(m(p)/q)/d(tca*)^0.5
1.0E+05
d(m(p)/q)/dlntca*

B-D flow
1.0E+04
Early linear flow Early linear flow

1.0E+03 1.0E+03
1.0E+00 1.0E+01 1.0E+02 1.0E+03 1.0E+04 1.0E+00 1.0E+01 1.0E+02 1.0E+03 1.0E+04
tca* (Days) tca* (Days)

Fig. 17 Radial and linear derivative plots illustrating flow regimes for simulated horizontal hydraulically-fractured CBM well
(infinite-conductivity fracture). A commercial numerical simulator was used to generate the production data.

The next step in the analysis is the use of flow regime-specific Cartesian plots to extract reservoir and hydraulic fracture
properties (Fig. 18). The early-linear flow regime was analyzed along with boundary-dominated flow (Table 5). Note that the
flowing material balance plot is non-linear for much of the well-life; the radial derivative plot suggests however that boundary-
dominated flow is reached.

a) Linear Flow Plot b) Flowing Material Balance


2.0E+04
6.0E-02

1.5E+04
Normalized Rate, scf/D/psi2/cp
[m(pi)-m(pwf)]/qg

4.0E-02
1.0E+04

5.0E+03 2.0E-02

0.0E+00
0 20 40 60 0.0E+00
0 500 1000 1500
Superposition Time Normalized Cumulative Production, MMscf

Fig. 18 Straight-line analyses of simulated horizontal hydraulically-fractured CBM well (infinite-conductivity fracture): a) linear
flow analysis and b) flowing material balance analysis. A commercial analytical simulator was used to generate the
production data.

TABLE 5 RESULTS OF STRAIGHT-LINE ANALYSIS FOR SIMULATED HYDRAULICALLY-FRACTURED HORIZONTAL CBM WELL
(INFINITE-CONDUCTIVITY FRACTURE)

Flow Regime Extracted Parameter Extracted Value

EARLY-LINEAR TOTAL FRACTURE HALF-LENGTH (FT)* 3024

BOUNDARY DOMINATED DRAINAGE AREA (ACRES) 147

(* total hydraulic fracture half-length is the sum of all individual hydraulic fractures)
SPE 131791 25

The derived (from linear flow analysis) total hydraulic fracture length is substantially greater than the input value (2250 ft).
In the tight gas (no adsorption) case analyzed in Clarkson and Beierle (2010), using the same well/hydraulic fracture inputs as
this CBM case, the total hydraulic fracture half-length is within 5% of the sum of the individual hydraulic fracture lengths
input into the simulator; we believe that the large discrepancy for the CBM case may be due to desorption. Although we have
used a desorption compressibility in our calculations, the calculations are referenced to initial pressure in the simulation, the
near-fracture drawdown may be creating additional desorption that would lead to a negative skin effect we continue to
investigate the source of the discrepancy.
Because only two flow regimes (early-linear and boundary-dominated) are evident in this hydraulically-fractured
horizontal well example, an attempt was made to match the data on a linear-flow type-curve (Wattenbarger et al. 1998) as
discussed previously (ex. Clarkson et al. 2009), desorption is accounted for in the type-curve analysis. Both the match and the
forecast appear reasonable (Fig. 19). We note that because multi-fractured horizontal wells can contain several transitional
flow regimes (ex. early-radial, late-linear), it is not always possible to use type-curves designed for vertical wells, as was done
in this case, to match the horizontal wells. In the type-curve match, a total xf of 2600 ft was used, which is within ~ 15% of the
simulator input and closer than for the straight-line analysis.

a) Linear Flow Type-Curves b) Linear Flow Forecast

100 100000

10
10000
qg (Mscf/D)
ye/xe qD

1
1000

0.1

100
0 200 400 600 800 1000
0.01
tca* (days)
0.0001 0.001 0.01 0.1 1 10 100
tDye

Fig. 19 Type-curve match of simulated horizontal hydraulically-fractured well (infinite-conductivity fracture) case: a) match to
linear-flow type-curve and b) resulting forecast.

The last step in the analysis is the derivation of individual hydraulic fracture lengths using allocated rates from production
logging along with multi-stage analytical simulation. The analytical simulator developed for multi-layer CBM analysis
(Clarkson 2009) is used to match the stage contributions along with the total commingled stage production (Clarkson and
Beierle 2010), constrained by the total hydraulic fracture length derived from production data analysis. In this simulated case,
each hydraulic fracture is of the same length. Matching of individual stage (# stages = # fractures in this case) production was
achieved by adjusting individual hydraulic fracture lengths while keeping the total hydraulic fracture length (sum of individual
stages) constant. The results of commingled and individual-stage analysis are given in Fig. 20.

a) Total Commingled Production b) Stage Rate Contributions

Rate (Mscf/D)
100000 10000000
0 100 200 300 400 500

1000000 1
10000

100000 2
Cumg (Mscf)
qg (Mscf/D)

1000

10000 3
Model
Stage

Actual
100
1000 4

5
10 100
0 20 40 60 80 100
Producing Time (days)
6

Actual qg Model qg Actual Cumg Model Cumg

Fig. 20 Match of commingled-stage (a) and individual hydraulic fracture stage (b) production data. The individual fracture
stage match occurs at a producing time of 10 days. Note that the number of fractures = number of stages in this case.
26 SPE 131791

We will continue to test this production analysis methodology against simulated and field data.

Use of Surveillance Techniques to Optimize Recovery. In the above sections, we discussed current methods to extract
reservoir property and well completion and stimulation efficiency information from CBM wells. Modern surveillance
methods are needed to fully optimize field development. Examples of modern surveillance methods that are currently being
used for CBM development, along with a brief discussion of their use, are provided below. The surveillance data discussed
below are measured periodically or continuously throughout the well life, and do not include other critical data usually only
measured once during the well life, such as gas contents, adsorption isotherms, well logs, initial pressure transient testing and
pre-fracture stimulation testing.

1. Continuous monitoring of flowing pressures (tubing and annulus), line pressures and flow rates and temperatures
using SCADA (supervisory control and data acquisition) on individual wells. These data can be used, with reservoir
pressure estimates (see below), for:
Calculating backpressure coefficient a simple plot of q/m(p) or q/P2 (Eq. 18) can be used to track well
delivery over time which can in turn be used to diagnose problems with completions (ex. plugging of liners
with coal fines) or track changes in effective permeability to gas (ex. Gierhart et al. 2006)
Calculating effective permeability to gas and water over well life (Clarkson et al. 2008a) used for
predicting future well performance
Minimum lifting rate calculations for wellbore fluids used to determine if the well can lift wellbore fluids
under current operating conditions
Effective horsepower calculations used to compare to nameplate horsepower to establish efficiency
Advanced production analysis used to determine reservoir and stimulation properties, which can be
mapped for use in field simulation
Simulation history-matching SCADA data can be used to calibrate reservoir simulation models, which in
turn can be used in development planning
Compression optimization - a calibrated wellbore/reservoir model can be used to generate production
forecasts under various backpressure (horsepower) scenarios; economics can be used to select the optimal
horsepower configuration (Clarkson and McGovern 2005)
2. Pressure observation wells (POW) continuous monitoring of coal seam pressures through isolation of coal seams
with bridge-plug/ packer assemblies. These data can be used with other SCADA data for the purposes discussed
above. Pressure observation wells can be scattered throughout a field to assist with the optimization of field
development by field area it is expected that reservoir properties will differ in various portions of the field, so the
location and density of POW wells will be dictated by reservoir heterogeneity. In addition to the applications
discussed above, POWs can be used for group or individual well material balance calculations (see Jensen and Smith
1997).
3. Production logging production logs (including spinner and temperature surveys), run periodically throughout well
life, can be used to track contributions from individually-stimulated zones in vertical wells completed in multiple coal
seams, from sections in openhole vertical or horizontal wells or individual hydraulic fracturing stages in multi-
fractured horizontal wells. If the production logs are run periodically throughout the well life, commingled
production along with the stage or layer-rate allocations can be history-matched to extract individual stage/layer
reservoir and stimulation properties (Spivey 2006, Clarkson 2009, and Clarkson and Beierle 2010). Production logs
can also be run prior to shutting-in wells for individual coal seam/zone pressure buildup analysis to assist with
pressure transient analysis (Bastian et al. 2005).
4. Time-lapse pressure transient analysis as described in a previous section, time-lapse PTA may be used to provide
estimates of reservoir pressure and changes in effective permeability (gas and/or water) during depletion. These
measurements can be made in lieu of or independently of production analysis. Further, changes in fracture
stimulation properties (ex. hydraulic fracture conductivity, or effective length) may be monitored to establish if re-
fracturing is necessary.
5. Wellhead gas sampling if the coalbed gas is comprised of significant amounts of other gaseous components in
additional to methane, then the potential for produced gas compositional changes exists due to relative adsorption
effects (Clarkson et al 2008a). Period sampling of wellhead gases for gas compositional analysis may be necessary to
allow for calibration of multi-component adsorption models, which in turn can be used to predict gas composition as
a function of reservoir pressure. When combined with reservoir simulation, the gas compositions can then be
projected as a function of time and used for facilities design and economic analysis. In enhanced recovery (ECBM)
or CO 2 sequestration operations, wellhead gas sampling is also necessary to establish whether breakthrough has
occurred (Koperna et al. 2009).
6. Surface deformation monitoring both tiltmeter and Interferrometric Synthetic Aperture Radar (InSAR) have
recently been used (Koperna et al. 2009) in an attempt to detect subsurface movement of CO 2 as part of a CO 2-
ECBM/storage pilot in the Pump Canyon area of the San Juan Basin. Both tiltmeter and InSAR can detect surface
SPE 131791 27

deformation which can be used to infer activity in the subsurface, for example monitoring of the CO2 plume during
injection operations. Although tiltmeters have traditionally been applied in the oil and gas industry to map hydraulic
fractures, the recent application of the technology for tracking fluid movement could be quite useful for both primary
and ECBM operations.

An additional surveillance method worthy of note, with potential increased application to coalbed methane, is microseismic
monitoring of hydraulic-fracturing treatments. Because of the difficulty in predicting hydraulic fracture geometry a-priori
using existing hydraulic fracture models, it is anticipated that microseismic monitoring of CBM well frac treatments will
increase, particularly if use of multi-fractured horizontal well technology increases. This section has focused on surveillance
techniques applied periodically or continuously throughout CBM well-life, and microseismic surveys are generally performed
only once during initial completion operations. It is possible, however, that microseismic surveys may also be used to monitor
ref-frac treatments or pay adds in horizontal wells.
With complexity of completion and stimulation methods for CBM exploitation expected to increase, there will be an
increase in demand for cost-effective surveillance technology to assist with well-performance monitoring and effective field
development in primary and enhanced recovery scenarios. The following is a discussion of possible future development
scenarios for CBM, technical challenges that currently exist, and future technical challenges associated with CBM
development.

Discussion: Future Possibilities for CBM Development and Associated Technical Challenges
Over the past two decades, we have seen the development of new CBM plays that were not conceivable at the time that the
San Juan Basin Fruitland coal fairway, the most prolific CBM reservoir in the world (to date), was first being developed in
1980s. For example, the low-rank coals of the Powder River Basin, once considered not to be a target because of low gas
content and excessive early water production, became a major exploitation target in the late 1990s/early 2000s. The dry,
multi-seam, low-rank, Horseshoe Canyon coals of Western Canada were not considered a target because initial permeability
tests with water injection/falloff testing were not encouraging only after technology adapted from the Appalachian Basin
(high-rate N 2 injection) was applied did it become apparent that commercial CBM rates could be achieved. Finally, a new
frontier of CBM play is developing, thanks to adaptation of technology: deep, low-permeability coals. An example of this is
the deep (> 5000 ft), dry Mannville coals in Western Canada, which are currently being tested using multi-fractured horizontal
well technology adapted from shale gas development (Tuffs 2009). In the Corbett Creek area, in shallow (< 4000 ft) wet
portions of the Mannville play, multi-lateral wells are now being used to exploit the play (Finn et al. 2009). Multi-lateral
technology for CBM exploitation had been previously suggested by Maricic et al. (2005).
Early CBM development was contemporaneous with Devonian shale development in North America, and some CBM
technology was adapted specifically from that developed initially from Devonian shale development (ex. N2 fracs currently
employed in Horseshoe Canyon CBM play). It is expected that the current focus on shale and tight gas development in North
America will similarly result in technology that can be readily adapted for CBM development, particularly as deeper, lower-
permeability coals are being targeted. Completion technology is not the only technology expected to be adapted the
following is a list of technologies currently in-use, or being developed, specifically for shale and tight gas plays in North
America, that may find increased or regular application for the exploitation of CBM reservoirs:

Microseismic currently in regular use for shale gas and tight gas plays to establish induced hydraulic-fracture
geometry and optimize completions, this technology may be more regularly used for CBM development, particularly
if use of multi-fractured horizontal well technology continues to grow. New developments in this area (ex. treatment
well or surface microseismic) could similarly benefit CBM.
New hydraulic-fracture models the creation of complex fracture networks associated with high-rate water fracs in
the Barnett shale play (ex. Mayerhofer et al. 2008) have necessitated investment in technology to predict complex
fracture geometries due in part to the existence of a healed natural fracture network, as well as other reservoir
heterogeneities and in-situ stress state (magnitude and orientation). It has long been known that hydraulic-fracturing
in coals (Fig. 21) can create a complex fracture geometry (Palmer et al. 1993), as observed in mine-back studies, yet
there has been an absence of technology for predicting fracture geometry (and distribution of conductivity) in
naturally-fractured coals. Palmer et al. 2009a,b discuss the mechanisms leading to enhanced permeability in
naturally-fractured reservoirs during hydraulic fracturing operations. It is hoped that the current focus on the
development of new technology for predicting complex fracture growth for shale gas will lead to advances that can
also be used for CBM, although it is recognized that there are important differences in CBM and shale reservoir
properties.
28 SPE 131791

Fig. 21 Illustration of complex fracturing associated with CBM hydraulic fracture stimulation (plan view). Modified from
Palmer et al. (1993).

Time-lapse (4D) seismic this surveillance method is currently being tested in a tight gas reservoir in Western
Canada the concept is to shoot seismic surveys periodically during initial stimulation of multi-fractured horizontal
wells and then analyze velocity/anisotropy effects created by the hydraulic fracturing. The technology is being tested
as a way to infer hydraulic fracture geometry when microseismic cannot be used (for example due to a lack of
observation wells). Continued surveys taken during depletion may be used to detect dynamic changes in the reservoir
(ex pore pressure changes), which in turn can be used for development planning. Applicability of this surveillance
method to CBM remains uncertain.

Although CBM development is expected to continue to benefit from technologies/evaluation methods initially developed
for shale and tight gas, there are some CBM-specific knowledge and technology gaps requiring further R&D that are unlikely
to be filled from shale/tight gas research alone. These include:

Relative permeability estimation. Many CBM plays exhibit 2-phase flow of gas and water through the natural
fracture system, which can be modeled using Darcys Law by inclusion of relative permeability curves. It is expected
that derived (from field or core data) relative permeability curve shape will be coal-specific, and is likely a function
several CBM properties such as pressure and desorption-dependent permeability. There is a paucity of published
relative permeability curves (from core or other sources) for CBM, making it difficult to ascertain the fundamental
controls. Further, given the complications associated with core-derived relative permeability testing (obtaining
representative samples etc.), new methods for developing relative permeability from field data should be researched.
Stress- and desorption-dependent permeability prediction. Although there have been a number of analytical models
developed for predicting permeability changes in coal during depletion or gas injection (see Palmer 2009), the models
developed to date make simplifying assumptions that should continue to be tested. Gu and Chalaturnyk (2006)
summarize some of the key limiting assumptions and simplifications of the models developed to that point, including:
isotropic permeability, linear elasticity (continuum model), ignoring interaction between stress and strain when
applying superposition and ignoring mechanical/hydraulic aperture differences. Those authors demonstrate that some
of the assumptions are limiting by comparing predictions with a discontinuum medium coupled model
(geomechanical simulator). Few of the models have been tested in even the simplest of CBM development scenarios
(primary depletion). Further testing of these models is required against actual field data, preferably in the simple
pressure depletion scenario, prior to their use for predicting permeability changes in the more complicated ECBM
scenario. Use of coupled geomechanical-flow models (ex. Gu and Chalaturnyk 2006) for this purpose warrants
further investigation, particularly if the models can be calibrated with field data (ex. time-lapse PTA); these more
rigorous models, if properly calibrated, can be used to assess under what conditions the use of the simpler analytical
models may be appropriate.
Wellbore stability prediction. The increase in use of single- and multi-lateral horizontal wells in a variety of coal
basins with different coal properties subject to a variety of in-situ stress conditions will necessitate additional
geomechanical testing and modeling to assist with the prediction of wellbore stability, permeability changes during
drilling and fines generation. Deisman et al. (2008) performed a series of innovative geomechanical tests supporting
well design for Canadian CBM reservoirs unconfined compressive strength and triaxial tests, were performed using
a variety of stress paths. The testing was performed specifically to investigate coal strength, permeability changes
with stress, fines generation and particle size distributions of coal that had failed the latter has implications for
slotted liner design in the laterals. Deisman et al. (2009) discuss the very important issue of scaling up
geomechanical properties (strength and deformation) of coal from small samples to field scale, including the impact
SPE 131791 29

of fracture and joints. Empirical and new numerical methods (Synthetic Rock Mass) were investigated. In addition
to scale, the geomechanical properties of coal can be expected to be a function of a variety of coal properties
including coal rank, and organic matter content, so continued investigations of CBM geomechanical properties will
be required for CBM reservoirs being exploited with horizontal well technology to allow for improvements in drilling
and completion efficiency.
Petrophysical methods for in-field estimation of adsorbed gas content and permeability. Presently there is no
petrophysical method for directly measuring in-situ adsorption, which is the dominant storage mechanism in coals.
Currently, lab testing of multiple reservoir samples is required to relate adsorption to coal physical properties, which
in turn can be used to calibrate logs to predict adsorbed gas content levels for intervals with no laboratory
measurements (Mavor 1996). As mentioned above, a technique based upon in-field RAMAN spectroscopy was
shown to be able to estimate critical desorption pressure (Lamarre and Pope 2007; the estimation of in-situ gas
content, however, still requires the use of an adsorption isotherm, measured in the lab. Others have investigated the
possibility of using low-field NMR to directly quantify adsorbed gas on coal. After gathering NMR data on coal
exposed to methane at varying pressure, Guo et al. (2007) interpreted the portion of the T 2 spectrum at very small
relaxation times (< 1 ms) to be due to adsorbed methane (Fig. 22). There is therefore the possibility that NMR may
be used to estimate in-situ adsorption. It is likely that some calibration to core data will still be required.

Methane in
porosity?

Adsorbed
methane?

Fig. 22 Use of low-field NMR to detect the adsorbed phase (methane) in coal. Note that it is assumed that the portion of
the spectrum at relaxation times < 1 ms is attributable to the adsorbed phase. Modified from Guo et al. (2007).

Petrophysical logs have been used recently to infer CBM permeability. For example, a software program called
LogFAC has been used to detect fractured coals and areas of enhanced permeability by using conventional log suites
to calculate movable fluid volume in the coal cleat system (Rozak et al. 2002). Image logs, such as FMI, can provide
evidence of fracturing from which permeability levels can be inferred, but a standard approach for direct estimation
of permeability from well-logs remains elusive.
Advanced production analysis methods. The increased use of complex wellbore geometries (ex. multi-lateral wells)
and stimulation methods (multi-stage hydraulic fracturing in vertical and horizontal wells) will necessitate improved
methods for quantitative production analysis, to enable estimation of critical reservoir properties such as kh and
hydrocarbon-in-place, and stimulation properties such as hydraulic fracture half-length and fracture conductivity.
Although reservoir simulation may currently be used for this purpose, these models are time-consuming and data-
intensive to setup, and there is a need for simpler analytical methods to allow for quicker evaluation of CBM
reservoir properties, that could be used to constrain simulation models. These analytical methods must account for
CBM reservoir properties, such as 2-phase flow (gas+water), desorption and non-static absolute permeability, as
discussed in Clarkson et al. (2009). It is anticipated that improved methodologies for integrating surveillance data,
such as microseismic, production logs (or distributed temperature surveys), and tracer logs, with production analysis
will be required to improve the uniqueness of analysis, particularly when multi-stage stimulations are used. The latter
is the subject of a recent paper, although tight gas reservoirs were analyzed (Clarkson and Beierle 2010).
Sweetspot identification. Deeper and lower permeability (generally due to stress sensitivity) coals continue to be
explored for and tested. As a result, there is going to be a need for continued development of tools for identifying
enhanced productivity areas. One approach, recently tried by EnCana in the Mannville CBM play of Western Canada
(Tuffs 2009), is to use 3-D seismic for Gaussian curvature analysis (Fig. 23). Once areas of anomalous curvature are
identified using 3-D seismic, the seismic is calibrated to FMI and production logs. It is hoped that additional
techniques for enhanced permeability identification can be developed, without necessarily relying on 3-D seismic
data, which is expensive to acquire for CBM exploration alone.
30 SPE 131791

Fig. 23 Example of the use of 3-D seismic (Gaussian Curvature) to identify areas of enhanced natural fracturing in a coalbed
methane reservoir. Source: Tuffs 2009.

We end our discussion with some comments regarding the need to improve CBM development sustainability by
considering additional methods for energy recovery besides primary depletion of natural gas. It is clear that increasingly
complex well bore architecture and completion methodologies will be required to yield economic recovery of hydrocarbons
from low-permeability CBM reservoirs. These techniques are designed to maximize surface area contacted with the reservoir.
It is hoped that the use of multi-lateral/multi-fractured horizontal well technology for enhanced recovery (ECBM)/CO 2
sequestration operations will be investigated. Different combinations of wellbore architecture and stimulation treatments
could be evaluated where injectors and producers take advantage of the increased reservoir contact afforded by these
technologies. Use of complex wellbores for ECBM was suggested by Koperna and Riestenberg (2009). Although it is
expected that efficiency of ECBM/CO2 sequestration could be improved over conventional vertical well injector/producers, we
realize that these operations could be economically challenged, depending upon the combination of capital investment required
and operating expense, the cost of CO2 injection (CO2 purchase, cost of compression etc.) and commodity price obtained. In
addition to investigating the use of these technologies for enhanced recovery, additional ways to extract energy should be
considered, including in-situ (underground) coal gasification. Underground coal gasification is a process whereby coal is
gasified through injection of oxidants through wellbores to produce various gaseous products (also through wellbores),
including carbon dioxide, hydrogen, carbon monoxide, methane and hydrogen sulphide. Conceptually, the greenhouse gases
could be reinjected or used for ECBM or enhanced oil recovery (EOR). The end-products could be used for power generation
or commodity sale. Details of the process are beyond the scope of the current work the reader is referred to
www.lincenergy.com.au for additional details. The process is complex, requirements for sites strict, and there are
environmental issues to consider. The concept is illustrated below (Fig. 24) using vertical well (a) as well as multi-laterals (b)
for process implementation. It is expected that multi-lateral wells would be more efficient for this process.

a) b)

Fig. 24 Example of the underground coal gasification process using vertical wells (a) as injectors and producers and horizontal
wells (b). Source: Pana and Richardson 2009.
SPE 131791 31

Conclusions
After several decades of exploitation of coalbed methane, much has been learned about how to characterize this reservoir type,
although clearly there is still room for improvement in certain areas. The current work has summarized advancements in the
assessment of key reservoir properties such as gas content and permeability, although not all topics were covered. For
example, geomechanical property assessment, critical for hydraulic fracture and wellbore geometry design, was not covered in
detail; the reader is referred to recent work (ex. Deisman et al. 2008) on this subject.
The future of coalbed methane exploitation will include development of deeper, lower quality CBM reservoirs,
necessitating use of advanced drilling and completion methods. As a result, new procedures for reservoir evaluation will need
to be advanced to improve, for example, interpretation of pressure/rate transient signatures associated with complex
wellbore/completion designs. Although it is expected that CBM development will continue to benefit from advancements in
technology related to exploitation of other unconventional reservoirs such as shale and tight gas, CBM-specific technology
will need to be developed, particularly when it comes to implementation of enhanced recovery operations.
In the future, CBM development may also become linked to other methods for energy extraction access to unmineable
coal through technologies that provide high reservoir contact may allow for other energy extraction methods, such as
underground coal gasification, to be considered, thereby maximizing usage of the infrastructure already in place for CBM
development. We hope that this kind of synergy is being considered, with due consideration for the environment, and
encourage research and development along these lines.

Nomenclature
Field Variables
Bg = gas formation volume factor, reservoir volume to surface volume
Bw = water formation volume factor, reservoir volume to surface volume
C = backpressure coefficient (Eq. 18), Mscf/(D*psi2)
cd = Desorption compressibility, psi1
D = inertial or turbulent flow factory, D/Mscf
Fc = Fracture conductivity, md-ft
h = Formation thickness, ft
k = Absolute (formation) permeability, md
kg = Effective permeability to gas, md
krg = Relative permeability to gas, dimensionless
kw = Effective permeability to water, md
krw = Relative permeability to water, dimensionless
m(p) = Real gas pseudopressure, psi2/cp
M(p) = Multiphase flow potential, psi/cp
n = exponent in backpressure equation (Eq. 18), dimensionless
p = Pressure, psia
pL = Langmuir pressure constant, psia
pR = Volumetric average reservoir pressure, psia
pwf = Flowing bottomhole pressure, psia
qg = Gas production (surface) flowrate, MSCF/D
qw = Water production (surface) flowrate, STB/D
re = Drainage radius, ft
rw = Wellbore radius, ft
s = Skin factor, dimensionless
Sw = Water saturation, dimensionless, fraction
tca* = Material balance pseudotime, accounting for desorption, days
VL = Langmuir volume constant, scf/ton
xf = Fracture half-length, ft

Dimensionless Variables
qD = Dimensionless rate
tD = Dimensionless time
Greek Variables
g = Gas viscosity, cp
w = Water viscosity, cp
= Fracture porosity, fraction
c = Coal bulk density, g/cm3
32 SPE 131791

Subscripts
a = Pseudo
c = Material balance
ca = Material balance pseudo
D = Dimensionless variable
Dd = Dimensionless decline variable
g = Gas
R = Reservoir
sc = Standard conditions
w = Water
wf = Sandface
Superscripts
* = altered variable

References
Al-Khalifah, A-J.A., Horne, R.N. and Aziz, K. 1987. In-Place Determination of Reservoir Relative Permeability Using Well
Test Analysis. Paper SPE 16774 presented at the 62nd Annual Fall Technical Conference and Exhibition of the Society of
Petroleum Engineers, Dallas, Texas, 2730 September.
Arri, L.E., Yee, D., Morgan, W.D., and Jeansonmne, M.W. 1992. Modeling Coalbed Methane Production With Binary Gas
Sorption. Paper SPE 24363 presented at the SPE Rocky Mountain Regional Meeting, Casper, Wyoming, 1821 May.
DOI: 10.2118/24363-MS.
Australian Standards. 1991. Guide to Desorbable Gas Content of Coal, Direct Method: (AS 3980/1991). Standards
Association of Australia, Standards House, 80 Author St., North Sydney, NSW.
Bastian, P.A., Wirth, O.F.R., Wang, L., and Voneiff, G.W. 2005. Assessment and Development of the Dry Horseshoe Canyon
CBM Play in Canada. Paper SPE 96899 presented at the SPE Annual Technical Conference and Exhibition, Dallas, TX,
912 October.
Barenblatt, G.I, Zheltov, I.P and Kochine, I.N. 1960. Basic Concepts in the Theory of Seepage of Homogeneous Liquids in
Fissured Rocks: Journal of Applied Mathematics and Mechanics (USSR), Vol. 24, No. 5, pp. 1286-1303, June 1960.
Barree, R.D., Barree, V.L., and Craig, D.P. 2007. Holistic Fracture Diagnostics. Paper SPE 107877 SPE Rocky Mountain
Petroleum Technology Conference, Denver, Colorado, 1618 April.
Bell, G., 2005. Gas Content Variation of Horseshoe Canyon Coals: An Analytical Review of Measurment Results. Canadian
Soceity of Petroluem Geology, Gussow Conference, Canmore Alberta.
Bhanja, A.K., and Srivastava, O.P. 2008. A New Approach to Estimate CBM Gas Content from Well Logs. Paper SPER
115563, Asia Pacific Oil and Gas Conference, Perth, October 2008.
Brace, W.F., Walsh, J.B. and Frangos, W.T. 1968. Permeability of Granite under High Pressure: Journal of Geophysical
Research, 73, 2225.
Bourbie, T. and Walls, J. 1982. Pulse Decay Permeability: Analytical Solution and Experimental Test: Society of Petroleum
Engineers Journal, 22, 71922.
Bustin, R.M. 1997. The Importance of Fabric and Composition on stress Sensitivity of Permeability in Some Coals of the
Northern Sydney Basin, Australia: Relevance to Coalbed Methane Exploitation: American Association of Petroleum
Geologist Bulletin, v. 81 pp. 1894-1908.
Bustin, R.M. 2005. Coalbed Methane Geology and Engineering: Application to Exploration and Development, CBM
Solutions, Calgary, Alberta, 250 p.
Bustin, R.M. 2009. Role of Moisture in the Adsorption and Desorption of Gas from Coal: Geological Society of America
Abstracts with Programs, Vol. 41, No. 7, p. 550.
Bustin, A.A.M. and Bustin, R.M. 2009. Gas in Box: How Much Producabile Gas is in the Horseshoe Canyon, CSUG, Calgary,
Nov. 2009.
Bustin, A..M.M., Bustin, R.M, and X. Cui. 2008. Impact of Shale Fabric on Production, SPE Paper number 114167-MS.
Bustin, R.M., and Clarkson, C.R. 1999. Free Gas in Matrix Porosity: a Potentially Substantial Resource in Low Rank Coals
Coalbed Methane Symposium Proceedings, Tuscaloosa, Alabama, May 1999, p. 197-214.
Chikatamarla, L, Cui, X. and Bustin, R.M. 2004. Implications of Volumetric Swelling/Shrinkage of Coal in Sequestration of
Acid Gases, Proc. Int. Coalbed Methane Symposium (2004).
Chikatamarla, L., Bustin, R.M. and Cui, X. in press. CO2 Sequestration into Deep Coal Beds: Insights from Laboratory
Experiments and Numerical Modeling, AAPG Special Publication "Carbon Dioxide Sequestration in Geological Media -
State of the Science: AAPG Studies in Geology # 59 (editors: M.Grobe, J.C. Pashin, and R.L. Dodge).
Clarkson, C.R., 2003. Application of a New Multicomponent Gas Adsorption Model to Coal Gas Adsorption Systems; SPE
Journal Vol. 8, P. 236-251.
SPE 131791 33

Clarkson, C.R. 2009. Case Study: Production Data and Pressure Transient Analysis of Horseshoe Canyon CBM Wells. The
Journal of Canadian Petroleum Technology 48 (10): 27-38.
Clarkson, C.R., and Beierle, J.J. 2010. Integration of Microseismic and Other Pos-Fracture Surveillance with Production
Analysis: A Tight Gas Study. Paper SPE 131786 presented at the 2010 SPE Unconventional Gas Conference, Pittsburgh,
PA, 23-25 February.
Clarkson, C.R., and Bustin, R.M. 1997. Variation in Permeability with Lithotype and Maceral Composition of Cretaceous
Coals of the Canadian Cordillera. The International Journal of Coal Geology 33: 135-151.
Clarkson, C.R. and Bustin, R.M. 2000. Binary Gas Adsorption/Desorption Isotherms: Effect of Moisture and Coal
Composition upon Component Selectivity: International Journal of Coal Geology, vol. 42, p. 241-272.
Clarkson, C.R., and McGovern, J.M. 2003. A New Tool for Unconventional Reservoir Exploration and Development. Paper
0336 presented at the International Coalbed Methane Symposium, Tuscaloosa, Alabama, 59 May.
Clarkson, C.R., and McGovern, J.M. 2005. Optimization of Coalbed Methane Reservoir Exploration and Development
Strategies Through Integration of Simulation and Economics. SPEREE 8 (6): 502519. SPE-88843-PA.
Clarkson, C.R., Jordan, C.L., Gierhart, R.R., and Seidle, J.P. 2007. Production Data Analysis of CBM Wells. Paper SPE
107705 presented at the Rocky Mountain Oil and Gas Technology Symposium, Denver, CO, 16-18 April.
Clarkson, C.R., Jordan, C.L., Gierhart, R.R., and Seidle, J.P. 2008a. Production Data Analysis of CBM Wells SPEREE 11 (2):
311325. SPE-107705-PA.
Clarkson, C.R., Jordan, C.L., Ilk, D. and Blasingame, T.A. 2009. Production Data Analysis of Fractured and Horizontal CBM
Wells. Paper SPE 125929 presented at the 2009 SPE Eastern Regional Meeting held in Charleston, West Virginia, 23-25
September.
Clarkson, C.R., Pan, Z., Palmer, I. and Harpalani, S. 2008b. Predicting Sorption-Induced Strain and Permeability Increase with
Depletion for CBM Reservoirs. Paper SPE 114778 presented at the SPE Annual Technical Conference and Exhibition,
Denver, CO, 2124 September.
Conway, M.J., Mavor, M.J., Saulsberry, J., Barree, R.B. and Schraufnagel, R.A. 1995. Multi-Phase Flow Properties for
Coalbed Methane Wells: A Laboratory and Field Study. Paper SPE 29576 Joint Rocky Mountain Regional Meeting and
Low-Permeability Reservoirs Symposium, Denver, CO, 20-22 March.
Cox, D.O., Young, G.B.C., and Bell, M.J. 1995. Well Testing in Coalbed Methane (CBM) Wells: An Environmental
Remediation Case. Paper SPE 30578 presented at the SPE Annual Technical Conference and Exhibition,
Dallas, TX, 22-25 October.
Cui, X., and Bustin, R.M. 2006, Controls of Coal Fabric on Coalbed Gas Production and Compositional Shift in Both Field
Production and Canister Desorption Test: Journal Society of Petroleum Engineers. March 2006, P. 111-117.
Cui, X., Bustin, R.M., and Laxminarayana, C. 2007. Impacts of Adsorption-Induced Coal Swelling on Permeability and Acid
Gas Sequestration into Coal Seams: Insights from Experimental and Numerical studies, Journal of Geophysical Research,
vol. 112.
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 (2009) 9, 208223
Cussler, E.L., 1997. Diffusion Mass Transfer in Fluid Systems, 2nd edition. Cambridge University Press, Cambridge, 580 pp.
Deisman, N., Gentzis, T., and Chalaturnyk, R.J. 2008. Unconventional Geomechanical Testing on Coal for Coalbed Reservoir
Well Design: The Alberta Foothills and Plains. The International Journal of Coal Geology 75: 15-26.
Deisman, N., Mas Ivas, D., Darcel, C., and Chalaturnyk, R.J. 2009. Empirical and Numerical Approaches for Geomechanical
Characterization of Coal. The International Journal of Coal Geology, In Press.
Diamond, W.P., Schatzel, S.J. 1998. Measurement of Gas Content of Coal: a Review: International Journal of Coal Geology,
vol. 35, p. 311-331.
Dicker, A.I., Smits, R.M. 1988. A Practical Approach for Determining Permeability from Laboratory Pressure-Pulse decay
measurements. SPE Paper 17578 Presented at the SPE International Meeting in Petroleum Engineering, 14 November,
Tianjin, China.
Durucan, S., and Edwards, J.S. 1986. The Effects of Stress and Fracturing on Permeability of Coal, Mining Science and
Technology 3 (1986), pp. 205216.
EIA. 2009. Coalbed methane reserves and production; US Energy Information Administration,
http://tonto.eia.doe.gov/dnav/ng/ng_enr_cbm_a_EPG0_r51_Bcf_a.htm.
Enever, J.R.E, and Hennig, A. 1997. The Relationship Between Permeability and Effective Stress for Australian Coals and Its
Implications with Respect to Coalbed Methane Exploration and Reservoir Modelling. Paper 9722 presented at the
International Coalbed Methane Symposium, Tuscaloosa, Alabama, 1217 May.
Fetkovich, M.J., Guerrero, E.T., Fetkovich, M.J. and Thomas, L.K. 1986. Oil and Gas Relative Permeabilities from Rate-Time
Performance Data. Paper SPE 15431 presented at the SPE Annual Technical Conference and Exhibition, New Orleans,
Louisiana, 5-8 October.
Fetkovich, M.J. 1980. Decline Curve Analysis Using Type Curves. Journal of Petroleum Technology 32 (6): 10651077;
Trans., AIME 269. SPE-4629-PA.
Finn, C.M., Seely, D., and Martin, J. 2009. Maximum Reservoir Contact (MRC) Wells for Coalbed Methane Exploitation.
Presented at the Canadian Society of Unconventional Gas Conference, Calgary, Alberta, Canada, 1820 November.
34 SPE 131791

Flores, R.M., 1998. Coalbed Methane: From Hazard to Resource. International Journal of Coal Geology, v. 35, p.3-25.
Gasem, K., Mohammed, S. and Robinson, R.L. 2009. Modeling Coalbed Methane Adsorption and CO 2 Sequestration. in
Encyclopedia of Chemical Processing, S. Lee Editor., Taylor and Francis.
Gibbs, J.W. 1961. On the Equilibrium of Heterogeneous Substances in J.W. Gibbs, ed., The Scientific Papers of J.W. Gibbs,
New York NY: Dover Publications, v. 1, p. 55349.
Gierhart, R., Gips, G., and Seidle, J. 2006. San Juan Basin Fruitland Coal Pressure Dependent Permeability Observations.
Paper presented at the SPE Advanced Technology Workshop on Unconventional Gas, Keystone, CO, 1921 April.
Gierhart, R.R., Clarkson, C.R., and Seidle, J.P. 2007. Spatial Variation of San Juan Basin Fruitland Coalbed Methane Pressure
Dependent Permeability: Magnitude and Functional Form. Paper IPTC 11333 presented at the International Petroleum
Technology Conference, Dubai, U.A.E., 46 December.
Goodman, A.L., Busch, A., Bustin, R.M, Chikatamarla, L., Day, S., Duffy, G.J, Fitzgerald, J.E, Gasem, K.A.M, Gensterblum,
Y., Hartman, C., Jing, C., Krooss, B.M., Mohammed, S., Pratt, T., Robinson Jr., R.L., Romanov, V., Sakurovs, R.,
Schroeder, K., and White, C.M. 2007. Inter-Laboratory Comparison II: CO2 Isotherms Measured on Moisture-
Equilibrated Argonne Premium Coals at 55 C and up to 15 MPa: International Journal of Coal Geology 72, Issues 3-4,
22 November 2007, Pages 153-164 .
Gu, F., Chalaturnyk, R.J. 2006. Numerical Simulation of Stress and Strain Due to Gas Sorption/Desorption and Their Effects
on In Situ Permeability of Coalbeds. The Journal of Canadian Petroleum Technology 45 (10): 52-62.
Gumma, S., and Talu, O. 2003. Gibbs Dividing Surface and Helium Adsorption: Adsorption, 2003; 9:1728.
Guo, R., Mannhart, K., and Kantzas, A. 2007. Characterizing Moisture and Gas Content of Coal by Low-Field NMR. The
Journal of Canadian Petroleum Technology 46 (10): 49-54.
Guo, R., Mannhart, K., and Kantzas, A. 2008. Laboratory Investigation on the Permeability of Coal During Primary and
Enhanced Coalbed Methane Production. The Journal of Canadian Petroleum Technology 47 (10): 27-32.
Gray, I., 1987. Reservoir Engineering in Coal Seams: Part 1- The Physical Process of Gas Storage in Coal Seams: Society of
Petroleum Engineering Reservoir Engineering, p. 28-34
Hager, C.J., and Jones, J.R. 2001. Analyzing Flowing Production Data With Standard Pressure Transient Methods. Paper SPE
71033 presented at the SPE Rocky Mountain Petroleum Technology Conference, Keystone, Colorado, 2123 May.
Ham, Y. and Kantzas, A. 2008. Measurement of Relative Permeability of Coal: Approaches and Limitations. Paper SPE
114994 presented at the CIPC/SPE Gas Technology Symposium 2008 Joint Conference, Calgary, Alberta, Canada, 1619
June.
Harpalani, C., and Chen, G. 1995. Influence of Gas Production Induced Volumetric Strain on Permeability of Coal, Geotech.
Geol. Engng., 15, 303-325.
Harpalani, S. and Pariti, U.M. 1993. Study of Coal Sorption Isotherms Using a Multicomponent Gas Mixture. Paper 9356,
presented at the 1993 International Coalbed Methane Symposium, The University of Alabama/Tuscaloosa (May 17-21,
1993) Vol. I, pp. 151-160.
Harpalani, S., Singh, K., and Zutshi, A. 2006. CO2/N2 Flow Behavior of Deep Coal-Gas Reservoirs. Paper presented at the 41st
U.S. Symposium on Rock Mechanics: 50 Years of Rock Mechanics Landmarks and Future Challenges, Golden,
Colorado, 17-21 June.
Hawkins, J.M., Schraufnagel, R.A., Olszewski, A.J., ResTech Pittsburgh, 1992. Estimating Coalbed Gas Content and Sorption
Isotherm Using Well Log Data. SPE Paper 24905-MS, SPE Annual Technical Conference and Exhibition, Washington,
D.C.4-7 October 1992.
Hongguan Yu, Guo, W., Cheng, J., and Hu, Q. 2008. Impact of Experimental Parameters for Manometric Equipment on CO2
Isotherms Measured: Comment on Inter-laboratory Comparison II: CO2 Isotherms Measured on Moisture-Equilibrated
Argonne Premium Coals at 55C and up to 15 MPa by Goodman et al. (2007). International Journal of Coal Geology. 74
(2008) 250258
Hopkins, C.W., Frantz, J.H., Flumerfelt, R.W., and Spivey, J.P. 1998. Pitfalls of Injection/Falloff Testing in Coalbed Methane
Reservoirs. Paper SPE 39772 presented at the SPE Permian Basin Oil and Gas Recovery Conference, Midland, TX, 2527
March.
Hsieh, P.A., Tracy, J.V., Neuzil, C.E., Bredehoeft, J.D., and Silliman, S.E. 1981. A Transient Laboratory Method for
Determining the Hydraulic Properties of Tight Rocks: I. Theory. International Journal of Rock Mechanics and Mining
Sciences, 18, 24552.
Jensen, D., and Smith, L.K. 1997. A Practical Approach to Coalbed Methane Reserve Prediction Using a Modified Material
Balance Technique. Paper 9765 presented at the Intl. Coalbed Methane Symposium, Tuscaloosa, Alabama, 1216 May.
Jikich, S., McLendon, R.T., and Smith, J.H. 2009. Permeability Variations in an Upper Freeport Coal Core Due to Changes in
Effective Stress and Sorption. Paper 124348 presented at the SPE Annual Technical Conference and Exhibition, New
Orleans, Louisiana, 4-7 October.
Jochen, V. A. and Lee, W.J. 1993. Reservoir Characterization of an Openhole Cavity Completion Using Production and Well
Test Data Analysis. Paper 26917 presented at the SPE Eastern Regional Conference and Exhibition, Pittsburgh, PA, 2-4
November.
Jones, S.C. 1997. A Technique for Faster Pulse-Decay Permeability Measurements in Tight Rocks. SPE Reservoir
Evaluation & Engineering (SPEREE) 12: 1925. SPE 28450-PA.
SPE 131791 35

Kamal, M.M., and Six, J.L. 1989. Pressure Transient Analysis of Methane Producing Coalbeds. Paper SPE 116688 presented
at the 64th Annual Technical Conference and Exhibition, San Antonio, TX, 8-11 October.
King, G.R., Ertekin, T., and Schwerer, F.C. 1986. Numerical Simulation of the Transient Behavior of Coal-Seam
Degasification Wells, SPE Reservoir Formatopm Evaluation April 1986: 165-183.
Kissell, F.N, McCulloch, C.M., Elder, C.H., 1973. The Direct Method of Determining Methane Content of Coalbed for
Ventilation Design. U.S. Bureau of Mines Report of Investigation, R1 7776, 17 p.
Koperna, G.J., Jr., Oudinot, A.Y., McColpin, G.R., Liu, N., Heath, J.E., Wells, A., and Young, G.B. 2009. CO2-
ECBM/Storage Activities at the San Juan Basins Pump Canyon Test Site. Paper SPE 127073 presented at the SPE
Annual Technical Conference and Exhibition, New Orleans, Louisiana, 4-7 October.
Koperna, G.J., Jr., and Riestenberg, D. 2009. Carbon Dioxide Enhanced Coalbed Methane and Storage: Is There Promise?
Paper SPE 126627 presented at the SPE International Conference on CO2 Capture, Storage and Utilization, San Diego,
California, 2-4 November.
Kuuskraa, V.A., 2009. Worldwide Gas Shales and Unconventional Gas: A Status Report. United Nations Climate Change
Conference, Natural Gas, Renewable and Efficiency Pathways to a Low Carbon Economy, Copenhagen, Dec. 2009.
Lamarre, R.A., and Pope, J. 2007. Critical-Gas-Content Technology Provides Coalbed-Methane-Reservoir Data. Paper SPE
103539 published in the Journal of Petroleum Technology, Technology Today Series, November.
Langmuir, I., 1916. The constitution and fundamental properties of solids and liquids: Journal of the American Chemical
Society, 38: 2221-2295.
Laubach, S.E., Marrett, R.A., Olson, J.E., and Scott, A.R. 1998. Characteristics and Origin of Coal Cleat: A Review.
International Journal of Coal Geology 35: 175-207.
Lin, W., Tang, G.-Q., Kovscek, A.R. 2008. Sorption-Induced Permeability Change of Coal During Gas-Injection Processes.
SPE Reservoir Evaluation & Engineering (SPEREE) August 2008: 792802. SPE 109855-PA.
Manik, J., Ertekin, T., and Kohler, T.E. 2002. Development and Validation of a Compositional Coalbed Simulator. The
Journal of Canadian Petroleum Technology 41 (4): 39.
Maricic, N., Mohaghegh, S.D., and Artun, E. 2005. A Parametric Study of the Benefits of Drilling Horizontal and Multilateral
Wells in Coalbed Methane Reservoirs. Paper 96018 presented at the SPE Annual Technical Conference and Exhibition,
Dallas, TX, 9-12 October.
Mavor, M.J.: Coalbed Methane Reservoir Properties. A Guide to Coalbed Methane Reservoir Engineering, Gas Research Inst.
Report GRI-94/0397, Chicago (1996).
Mavor, M.J. and Gunter, W.D. 2006. Secondary Porosity and Permeability of Coal vs. Gas Composition and Pressure. SPE
Reservoir Evaluation & Engineering (SPEREE) April 2006: 114125. SPE 90255-PA.
Mavor, M.J., and Nelson, C.R.: Coalbed Reservoir Gas-In-Place Analysis, Gas Research Inst. Report GRI-97/0263, Chicago
(1997).
Mavor, M.J., and Robinson, J.R. 1993. Analysis of Coal Gas Reservoir Interference and Interference Test. Paper SPE 25860
presented at the Rocky Mountain Regional/Low Permeability Reservoirs Symposium, Denver, CO, 12-14 April.
Mavor, M.J., and Saulsberry, J.L. 1996. Testing Coalbed Methane Wells. A Guide to Coalbed Methane Reservoir Engineering,
Gas Research Inst. Report GRI-94/0397, Chicago (1996).
Mavor, M.J. and Vaughn, J.E. 1998. Increasing Coal Absolute Permeability in the San Juan Basin Fruitland Formation. SPE
Reservoir Evaluation & Engineering (SPEREE) June 1998: 201206. SPE 39105-PA.
Mavor, M.J., Close, J.C, and Pratt, T.J. 1991. Summary of the Completion Optimization and Assessment Laboratory (COAL)
Site, Gas Research Institute Topical Report No. GRI-91/0377, Chicago, Illinois (December 31, 1991).
Mavor, M.J., Owen, L.B., Pratt, T.J. 1990. Measurement and Evaluation of Coal Sorption Isotherm Data. 65 th Annual
Technical Conference and Exhibition of the Society of Petroleum Engineers. Society of Petroleum Engineers, New
Orleans, LA, p. 157170, SPE 20728.
Mavor, M.J., Pratt, T.J., and Britton, R.N. 1994. Improved Methodology for Determining Total Gas Content,
Vol. I: Canister Gas Desorption Data Summary, Gas Research Institute Topical Report No. GRI-93/0410, Chicago,
Illinois (May, 1994)
Mayerhofer, M.J., Lolon, E.J., Warpinski, N.R., Cipolla, C.L. and Walser D. 2008. What is Stimulated Reservoir Volume?
Paper SPE 119890 presented at the SPE Shale Gas Production Conference, Fort Worth, TX, 1618 November.
Mazumder, S., and Wolf, K.H. 2008. Differential Swelling and Permeability Change of Coal in Response to CO2 Injection for
ECBM. Paper SPE 114791 presented at the SPE Asia Pacific Oil & Gas Conference and Exhibition, Perth, Australia, 20-
22 October.
McLennan, J.D., P.S. Schafer, and T.J. Pratt. 1995. A Guide to Determining Coalbed Gas Content, Gas
Research Institute Report No. GRI-94/0396, Chicago, Illinois, 1995.
Mitra, A., Harpalani, S. and Kumar, A. 2008. CO2/N2 in Deep Coals and its Impact on Coal Permeability. Paper presented at
the 42nd U.S. Rock Mechanics Symposium and 2nd U.S.-Canada Rock Mechanics Symposium, San Francisco, California,
29 June2 July.
Mora, C. A., and Wattenbarger, R.A. 2009. Analysis and Verification of Dual Porosity and CBM Shape Factors. Journal of
Canadian Petroleum Technology, vol. 48., p.17-21.
36 SPE 131791

Palmer, I. 2009. Permeability Changes in Coal: Analytical Modeling. The International Journal of Coal Geology 77: 119-
126.
Palmer, I. 2010. Coalbed Methane Completions: Worldview. Submitted to The International Journal of Coal Geology.
Palmer, I., and Mansoori, J. 1996. How Permeability Depends on Stress and Pore Pressure in Coalbeds: a New Model, SPE
36737 (1996).
Palmer, I., Cameron, J., and Ponce, J. 2009a. Natural Fractures Influence Shear Stimulation Direction. Oil and Gas Journal
107 (12): 37-43.
Palmer, I., Cameron, J., and Ponce, J. 2009b. Additional Keys Give Stimulation Insights. Oil and Gas Journal 107 (13): 45-
51.
Palmer, I., Lambert, S.W. and Spitler, J.L.: Coalbed Methane Well Completions and Stimulations, Hydrocarbons from Coal,
AAPG, Tulsa (1993).
Palmer, I., Mavor, M., and Gunter, B. 2007. Permeability Changes in Coal Seams During Production and Injection. Paper
0713 presented at the International Coalbed Methane Symposium, Tuscaloosa, Alabama, 57 May.
Pan, Z. and Connell, L. 2008. Comparison of Adsorption Models in Reservoir Simulation of Enhanced Coalbed Methane
Recovery and CO2 Sequestration in Coal: International Journal of Greenhouse Gas Control, vol. 3, P 77-89.
Pana, C. and Richardson, R. 2009. Underground Coal Gasification Process - A Potential Source of Energy in Alberta.
Presented at the Canadian Society of Unconventional Gas Conference, Calgary, Alberta, Canada, 1820 November.
Perrine, R.L. 1956. Analysis of Pressure Buildup Curves, Drilling and Production Practice, American Petroleum Institute
(1956): 482-509.
Pini, R., Ottiger, S., Storti, G. and Mazzottim. 2009. Pure and Competitive Adsorption of CO2, CH4 and N2 on Coal for
ECBM. Energy Proceida, vol., 1, p. 705-1710.
Pini, R., Ottiger, S., Burlini, L. Storti, G. and Massotti, M. 2009. Role of Adsorption and Swelling on the Dynamics of Gas
Injection in Coal Journal of Geophysical Research, V. 114, B04203.
Pratikno, H., and Blasingame, T.A. 2003. Decline Curve Analysis Using Type-Curves Fractured Wells. Paper SPE 84287
presented at the SPE Annual Technical Conference and Exhibition, Denver, Colorado, 58 October.
Ramurthy, M., Marjerisson, D.M., and Daves, S.B. 2002. Diagnostic Fracture Injection Test in Coals to Determine Pore
Pressure and Permeability. Paper SPE 75701 presented at the Gas Technology Symposium, Calgary, Alberta, Canada, 30
April 2 May.
Robertson, E.P., and Christiansen, R.L. 2007. Modeling Laboratory Permeability in Coal Using Sorption-Induced-Strain Data.
SPE Reservoir Evaluation & Engineering (SPEREE) June 2007: 260-269. SPE 97068-PA.
Ross, D.J.K. and Bustin, R.M. 2007. Impact of Mass Balance Calculations on Microporous Shale Gas Reservoirs. Fuel, vol.
85, p 2596-2706.
Rouquerol, F., Rouquerol, J., and Sing, K. 1999. Adsorption by Powders and Porous Solids. Acad. Press, London, 1999.
Rozak, A.T., Bustin, R.M., Strashok, G.W., Beaton, A., Richardson, R., and Hunter, T. 2002. Application of LogFAC* to
Coalbed Methane Exploration in Western Canada: A Case History from Ardley Coals near Red Deer, Alberta. Paper SPE
75676 presented at the Gas Technology Symposium, Calgary, Alberta, Canada, 30 April 2 May.
Ruckenstein, R., Vaidyanthan, A.S., and Youngquist G.R. 1971. Sorption by Solids with Bidisperse Pore Structures. Chemical
Engineering Sciences, 26: 1305-1318.
Sakurovs, R., Day, S., and Weir, S., 2009. Causes and Consequences of Errors in Determining Sorption Capacity of Coal for
Carbon Dioxide at High Pressure. International Journal of Coal Geology. Vol 77m o 16021.
Schlachter, G. 2007. Using Wireline Formation Evaluation Tools to Characterize Coalbed Methane Formations. Paper SPE
111213 presented at the SPE East Kentucky Symposium, Lexington, Kentucky, 18-21 October.
Schraufnagel, R.A. and P.S. Schafer, 1996. The Success of Coalbed Methane. In: Saulsberry, J.L., Scahfer, P.S., Schraufnagel,
R.A. (Editors) A Guide to Methane Reservoir Engineering Gas Research Institute (GRI Reference Number GRI-94/0397),
Chicago, Illinois. P. 1.1- 1.10.
Schwerer, F.C. and Pavone, A.M. 1984. Effect of Pressure-Dependent Permeability on Welltest Analysis and Long-Term
Production of Methane from Coal Seams. Paper SPE/DOE/GRI 12857 presented at the SPE/DOE/GRI Unconventional
Gas Recovery Symposium, Pittsburgh, PA, 13-15 May.
Seidle, J.P., 1992. Application of Matchstick Geometry to Stress Dependent Permeability in Coals. Paper SPE 24361
presented at the 1992 SPE Rocky Mountain Regional Meeting held in Casper, Wyoming, 18-21 May.
Seidle, J.P., Kutas, G.M., and Krase, L.D. 1991. Pressure Falloff Tests of New Coal Wells. Paper SPE 21809 presented at the
Rocky Mountain Regional and Low-Permeability Reservoirs Symposium, Denver, CO, 15-17 April.
Shedid, S.A. and Rahman, K. 2009. Experimental Investigation of Stress-Dependent Petrophysical Properties and Reservoir
Characterization of Coalbed Methane (CBM). Paper SPE 120003 presented at the SPE Asia Pacific Oil and Gas
Conference and Exhibition, Jakarta, Indonesia, 4-6 August.
Shi, J.Q., and Durucan, S. 2003. Changes in Permeability of Coalbeds During Primary Recovery, Proc. International Coalbed
Methane Symposium (2003).
Shi, J. Q., and Durucan, S. 2004. Drawdown Induced Changes in Permeability of Coalbeds: A New Interpretation of the
Reservoir Response to Primary Recovery, Transport in Porous Media, 56, 1-16.
SPE 131791 37

Shu, D.M., Lakshmanan, C.C., and White, N. 1994. Estimation of In-Situ Coal Permeability from Slug and Packer Tests.
Paper SPE 28664, Unsolicited Manuscript.
Singh K.P., Kakati, M.C. 2000. Simple Models for Estimating Helium Densities of Coals. Chemical Engineering Journal; 76;
67 71.
Sinkinson, E., and Turner, H. 1926. Adsorption of Carbon Dioxide by Coal. Industrial and Engineering Chemistry, vol. 8 p.
602-605.
Sircar S. 1985. Excess Properties and Thermodynamics of Multicomponent Gas Adsorption. Journal of the Chemical Society,
Faraday Transactions, 1985; 81; 1527 40.
Smith, D.M. and Williams, F.L., 1981. A New Technique for Determining the Methane Content of Coal. Proceedings Intersoc.
Ener. Conversion Eng. Conf. p. 1267-1272.
Somerton et al., 1975. Effect of Stress on Permeability ofCcoal, International Journal of Rock Mechanics and Mining Sciences
Geomechanics Abstract 12 (56) (1975), pp. 129145.
Somerton, W. H., I. M. Soylemezoglu, and R. C. Dudley, 1974, Effect of Stress on Permeability of Coal: U.S. Bureau of
Mines Open-File Report, v. 45-74, p. 56.
Sparks, D.P., McLendon, T.H., Saulsberry, J.L., and Lambert, S.W. 1995. The Effects of Stress on Coalbed Reservoir
Performance, Black Warrior Basin, U.S.A. Paper SPE 30734 presented at the SPE Annual Technical Conference and
Exhibition, Dallas, TX, 22-25 October.
Spivey, J.P. 2006. Estimating Layer Properties for Wells in Multilayer Low-Permeability Gas Reservoirs by Automatic
History Matching Production and Production Log Data. Paper SPE 100509 presented at the SPE Gas Technology
Symposium, Calgary, Alberta, Canada, 1517 May.
Starzewski P, Grillet Y. 1989. Thermochemical Studies of Adsorption of He and CO2 on Coals at Ambient Temperature.
Fuel 68; 375 9.
Tang, G.Q., Jessen, K., and Kovscek, A.R. 2005. Laboratory and Simulation Investigation of Enhanced Coalbed Methane
Recovery by Gas Injection. Paper SPE 95947 presented at the SPE Annual Technical Conference and Exhibition, Dallas,
TX, 9-12 October.
Trimmer D, Bonner B, Heard HC, Duba A (1980) Effect of Pressure and Stress on Water Transport in Intact and Fractured
Gabbro and Granite. Journal of Geophysical Research, 85, 705971.
Tuffs, B. 2009. Mannville CBM in Central Alberta. Presented at the Canadian Society of Unconventional Gas Conference,
Calgary, Alberta, Canada, 1820 November.
Warren, J.E., and Root, P.J. 1963. The Behavior of Naturally Fractured Reservoirs; Transactions of the SPE of AIME, Vol. 3,
pp. 245-255, September 1963.
Wattenbarger, R.A., El-Banbi, A.H., Villegas, M. E. and Maggard, J.B. 1998. Production Analysis of Linear Flow Into
Fractured Tight Gas Wells. Paper SPE 39931 presented at the SPE Rocky Mountain Regional/Low-Permeability
Reservoirs Symposium, Denver, Colorado, 58 April.
Wei, X.R., Wang, G.X., Massarotto, P., Golding, S.D., and Rudolph, V. 2007. A Review on Recent Advances in the
Numerical Simulation for Coalbed-Methane-Recovery Proces. SPEREE December 2007: 657666. SPE-93101-PA.
White, C.M., Smith, D.H, Jones, K.L., Goodman, A.L., Jikich, S.A., LaCount, R.B., DuBose, S.B., Ozdemir, E., Morsi, B.
I. and Schroeder, K.T. 2005. Sequestration of Carbon Dioxide in Coal with Enhanced Coalbed Methane RecoverysA
Review. Energy and Fuels., v. 19. P. 659-724.
Wold, M.B. and Jeffrey, R.G. 1999. A Comparison of Coal Seam Directional Permeability as Measured in Laboratory Core
Tests and in Well Interference Tests. Paper SPE 55598 presented at the SPE Rocky Mountain Regional Meeting, Gillette,
Wyoming, 15-18 May.
Wold, M.B., Choi, S.K., Koenig, R.A. and Davidson, S.C. 1996. Anisotropic Seam Response to Two-Phase Fluid Injection
Into a Coalbed Methane Reservoir Measurement and Simulation. Paper SPE 36984 presented at the SPE Asia Pacific
Oil and Gas Conference, Adelaide, Australia, 28-31 October.
Yamada SE, Jones AH (1980) A review of pulse technique for permeability measurements. Society of Petroleum Engineers
Journal, 20, 3578.
Yi, J., Akkutlu, I.Y., and Deutsch, C.V. 2008. Gas Transport in Bidisperse Coal Particles: Investigation for an Effective
Diffusion Coefficient in Coal Beds. Journal of Canadian Petroleum Technology, Vol. 47, no. 10. P. 20-26.
Zahner, R. 1997. Application of Material Balance to Determine Ultimate Recovery of a San Juan Fruitland Coal Well. Paper
SPE 38858 presented at the SPE Annual Technical Conference and Exhibition, San Antonio, TX, 5-8 October.
Zuber, M.D., Sparks, D.P. and Lee, W.J. 1990. Design and Interpretation of Injection/Falloff Tests for Coalbed Methane
Wells. Paper SPE 20569 presented at the 65th Annual Technical Conference and Exhibition, New Orleans, Louisiana, 23-
26 September.
38 SPE 131791

TABLE 2 Summary of CBM Welltest Studies (Since 1990)

Author Year Tests Analyzed Summary


Discussed problems with traditional pre-fracture-
stimulation injection-falloff test interpretation, including
effects of exceeding formation pressure, testing
Zuber et al. 1990 Pre-frac IFOT multiple coal seams and stress-dependence of
porosity and permeability proposed new test
procedures. Coals from the Black Warrior Basin were
analyzed.
Discussed the effects of free gas and gas desorption
on pre- and post-fracture stimulation injection-falloff
test results, and developed a correlation for required
Seidle et al. 1991 Pre-and post-frac IFOT shut-in time to reach radial flow. The effects of stress-
dependence of porosity and permeability were not
discussed. Coals from the San Juan Basin were
analyzed.
In this study, DST, post-cavitation (single-well)
flow/buildup tests and multi-well interference tests
were conducted. Both conventional (using approach
of Perrine, 1956) and multi-phase flow potential
approaches were used and compared for post-
DST, FBU AND cavitation FBU, where it was observed that the multi-
Mavor and Robinson 1993 MULTIWELL phase flowwith
consistent potential approach
reservoir yieldedMulti-well
simulation. results
INTERFERENCE
interference tests yielded consistent results with the

The welltests were


This work has become the standard

In this study,
single-well simulated 2-phase CBM data, created for
tests.
various flow/shut-in
for multi-phase flow times, was analyzed using
CBM analysis.
conventional
performed at analysis methods
the GRI test site inas well
the SanasJuan
usingBasin
the
modified
(Fruitlandpseuodpressure
coal). function and approach
suggested by Kamal and Six (1989). Neither approach
DST, FBU, AND yielded results in exact agreement with simulator
Jochen and Lee 1993 MULTIWELL inputs, but what was more surprising was that the
INTERFERENCE conventional analysis approach yielded values closer
than the Kamal and Six approach. As with the Mavor
and Robinson study, welltests from the GRI test site
were performed, with somewhat different estimates of
absolute permeability. A zone of altered (higher)
permeability zone around cavitated wells was inferred.
In this study, slug test-derived permeability and skin
values were shown to be non-unique. Using a field
example from the Rock Creek site in the Black Warrior
SLUG/PRESSURE-
Basin, the authors mention that slug tests performed
Shu et al. 1994 BUILDUP PACKER
by withdrawing
effects. fluids should
Pressure-buildup consider
packer tests2-phase flowon
performed
TESTS
core holes at a colliery displayed s-shaped buildups
that were difficult to analyze.
In this study, several types of welltests were applied to
the Marie Lee-Blue Creek coals at the Rock Creek
production pilot site in the Black Warrior Basin.
Welltests performed included slug, water and gas
injection-falloff, production/shut-in and tank tests. The
influence of multi-phase flow on test results was
investigated. Water injection tests were found to
SLUG, IFOT (gas and provide a reasonable lower-limit estimate absolute
Conway et al. 1995 permeability. The authors
is performed under recommended
multi-phase conditions,that
a if testing
water), FBU, TANK
flow/(long-term) buildup test should be performed to
yield gas- and water-effective permeability information,
followed by
absolute a water injection-falloff
permeability. They also suggest
test to estimate
that field

estimates of relative permeability can be obtained


analogously to procedures derived for solution gas-
drive reservoirs (Al-Khalifa et al. 1987).
SPE 131791 39

In this study, coal permeability was quantified as a


function of stress late-time falloff data was analyzed
from IFOT tests in the Cedar Cove area, and slug tests
Sparks et al. 1995 IFOT/SLUG were analyzed
Warrior basin. in the Oakstress
Minimum Grovevalues
area ofinthe Black
both areas

were obtained from acid-breakdown tests.


Permeability
correlate withand
stress. The authors
production recommendedto
were demonstrated

performing IFOT tests at as low rates as possible to


reduce effects of stress.
This study presents a unique application of CBM
welltest analysis well testing was used to identify the
source of methane seepage observed in the vicinity of
coalbed methane wells in southwestern Colorado. The
welltests were designed to obtain permeability and
reservoir pressure information for reservoir simulation,
skin from producing wells and to determine if pressure
Cox et al. 1995 FBU interference occurred between producing CBM wells

Perrines (1956) method


PBU tests were performed on

and monitor wells.


monitor and producing wells.
was used to calculate effective permeability to gas and
water,
In this as wellaas
study, skin Perrines
multi-well approach
interference was used
test was chosen
to
quantify permeabilitypseudopressure
over the multi-phase anisotropy. approach
The test was
analyzed
because thewithwellsanalytical and numerical
exhibited primarily models.
single- phase
Permeability anisotropy
flow and relative waswas
permeability estimated to be 2.8:1
uncertain.
the principal directions of the permeability tensor were
not aligned
affected with cleat directions,
by larger-scale but apparently
discontinuities. were
Single phase
Wold et al. 1996 INTERFERENCE
test
interpretation
methods
resulted
in
interference
Pressure-buildup tests were performed in this study in
support of material
permeability balance
estimates analysis
that could ofbeFruitland
used coal
in the
(San Juan analysis
numerical Basin) wells. Late time testing
of multi-phase boundary effects
during cavity
were interpreted.
completion Of particular
provided wellbore note, this was
damage studyincluded.
Zahner 1997 FBU documented
The tests were the possible
conducted functional
on the form
Nipon (exponential)
0 Seam at
of absolute
Dawson permeability
River, growth Australia.
Bowen Basin, as a function of high
These
depletion in Fruitland
volatile bituminous coalare
coals wells a permeability
Permian in age.
growth function derived from pressure-buildup tests
was used in single-well numerical simulation modeling.
This study discussed the importance of injecting below
fracture pressure to obtain an accurate analysis of
injection-falloff tests, the impact of a non-stable
reservoir pressure,
communication withand
the importance of obtaining
reservoir. They good
stress the
Hopkins et al. 1998 IFOT
importance of estimating fracture closure pressure
prior to performing the injection falloff test so that the
tests can be designed not to exceed frac pressure.
Recommendations on testing prior to the injection-
falloff test are given.
As with Zahner (1997), this study quantified absolute
permeability growth in Fruitland coal wells using
Mavor and Vaughn 1998 DST, FBU welltest analysis. The results were used to calibrate
the Palmer-Mansoori equation for predicting absolute
permeability as a function of depletion.
This study quantified permeability anisotropy using 4-
well interference testing (central well as injector, outer
wells as observation) and core analysis. The
magnitude of the permeability values were consistent
Wold and Jeffrey 1999 INTERFERENCE, IFOT between field and core tests, but the maximum and
minimum at
oriented permeabilities
~ 90 to the obtained fromtest.
interference The tests
core tests were

were conducted on coals in the Dartbrook Mine, in the


Sydney coal basin, Australia.
40 SPE 131791

This study discussed welltesting techniques used in


the dry Horseshoe Coals of Western Canada. It was
noted that initial water injection-falloff tests indicated
very low permeability for the coals, yielding very
discouraging results, which was attributed to the water-
sensitive nature of the coals. They also noted that the
Bastian et al. 2005 IFOT, FBU stress-sensitive nature of the coals and non-Darcy flow
effects during injection caused injection tests with
nitrogen to yield difficult
interpretations. Conventional drawdown/buildup tests
and non-unique

were therefore suggested, in which individual coal


seams in the multi-seam coal package were isolated
and tested.
In this work, welltesting was used to develop a method
for predicting porosity and permeability changes as a
IFOT (gas and water),
Mavor and Gunter 2006 function of fracture pore pressure and gas content and
PBU
composition in support of enhanced coalbed methane
recovery (ECBM) and CO2 sequestration projects.
Gierhart et al. performed time-lapse pressure transient
analysis of flow/buildup tests, gathered for wells in the
Fruitland coal of the San Juan Basin, to quantify
absolute permeability (estimated from effective
permeability using relative permeability curves and
Gierhart et al. 2007 FBU
fluid saturation
depletion. As with
estimates)
the Zahner
changes
(1997),
as the
a function
functional
of

form of the permeability growth was noted to be


exponential the study analyzed wells from within and
external to the prolific Fruitland coal Fairway.
Clarkson analyzed single-seam pressure buildup data
gathered from the dry Horseshoe Coals of Western
Canada. Test results used to constrain inputs in for
Clarkson 2009 FBU multi-layer simulation and production data analysis.
Various reservoir heterogeneities were interpreted from
pressure-buildup derivatives including multi-layer,
composite and dual porosity behavior.
SPE 131791 41

TABLE 3 Summary of CBM Core-Derived PermeabilityStudies (Since 2005)

Author Year Tests Analyzed Summary


In this study, coalpack permeability to helium, carbon
dioxide and nitrogen was measured. Permeability of
Steady-state permeability the coalpack to nitrogen was shown to be fairly
Tang et al. 2005 to different pure gases at constant with pore pressure, whereas permeability
varying pore pressure decreased substantially with pressure for carbon
dioxide. Ground coal from the Powder River Basin was
pressed into cylindrical shapes (coalpack) for analysis.
In this study, permeability to methane, carbon dioxide
and flue gas (nitrogen/carbon dioxide mixtures) were
measured at varying effective stress (mean pore
pressure varied, while axial and confining stress was
maintained) and constant effective stress. A triaxial
cell was used to simulate in-situ conditions.
Steady-state, triaxially- Permeability to all gases decreased log-linearly with
stressed, permeability to changes in effective stress. Permeability to all gases
different pure gases and also decreased with a decrease in gas pressure at
Harpalani et al. 2006
flue gas (N2/CO2 constant effective stress, which was unexpected
mixtures) at varying because sorption-induced strain effects (at least for
effective stress levels pure CO2) weregas
with increasing expected to decrease
pressure. permeability
All gases exhibit similar

trends in permeability
reduction variation,
greatest in the pure CO with Injection of
permeability
2 case.

flue gas versus pure CO2 for ECBM operations is


suggested due to negligible coal swelling/permeability
loss in the former case.
In this study, permeability to nitrogen at variable net
stress and nitrogen, methane and carbon dioxide
permeabilities at variable pore pressure was measured.
The effects of sorption-induced matrix shrinkage were
Steady-state, hydrostatic quantified. An important observation is that sorption-
conditions, permeability induced strains measured at unconfined conditions may
Robertson and Christiansen 2007 to different pure gases at need to be corrected to apply to constrained core or
variable overburden and field permeability measurements. High volatile
pore pressures bituminous coals from the Gibson seam
of the Uinta-Piceance Basin as well as sub-bituminous
coals of the Powder River Basin were used coal
blocks were preserved in water prior to sampling for
analysis.
In this study, permeability to methane, nitrogen and
carbon dioxide was measured using cores that were
preserved in their native-state. Horizontal strain was
simultaneously measured. Cores were stressed
Steady-state, triaxially- triaxially in an attempt to represent in-situ conditions
stressed, permeability to reservoir temperature and moisture conditions were
Mitra et al. 2008 different pure gases at used. Permeability to N2 was the highest, followed by
constant horizontal and CH4 and CO2, followed by CH4 and N2. The ratio of
vertical stress permeability to gases did not match strain ratios

Coal cores were cut from coal


sorption-induced strain is not the sole control on
permeability variation.
In this study,
blocks, whichpermeability to methane,
were obtained carbon
from the Herrin dioxide
Formation
and nitrogen
of the was measured along with permeability
Illinois Basin.
associated
stress. Groundwith coal
gas mixtures at constant
and composite effective
coal cores were
Steady-state, hydrostatic tested after drying. Permeability reduction was
conditions, permeability greatest with CO2, followed by CH4 and N2. At a given
Lin et al. 2008 to different pure gases pore pressure, increasing CO2 relative to N2 in gas
and gas mixtures at mixtures reduces permeability; some nitrogen in the
constant effective stress injected gas will help to preserve permeability.
Permeability reduction in ground coal was greater than
with the composite cores. Coals from the Powder River
Basin were analyzed.
Steady-state, hydrostatic In this study, permeability, along with axial strain, was
Mazumder et al. 2008 conditions, real-time
permeability estimates
42 SPE 131791

while injecting carbon size was larger than typical, for the purpose of
dioxide into coal preserving coal heterogeneities. Measurements were
containing adsorbed performed on dry and moisture-equilibrated coals.
methane at constant pore Injection-production schemes were performed at
pressure and effective constant pore pressure and effective stress.
stress Differential swelling associated with CO2 injection was
quantified,Samples
porosity. were
as well as thefrom theon
impact Beringen coal mines
permeability and

in Belgium, Silezia coalfield in Poland and the Tupton


coalfields in the U.K.
In this study, permeability to helium and methane was
measured at variable confining pressures.
Permeability was also measured during displacement
of CHto
prior 4 by CO2. The
analysis. coal was equilibrated
Permeability withamoisture
to helium was strong
Steady-state, hydrostatic function of confining pressure and displayed
conditions, permeability hysteresis. Permeability to methane and carbon
Guo et al. 2008 to different gases at dioxide were smaller than helium, with CO2 yielding the
variable confining smallest values. Methane and carbon permeabilities
pressures were found to be a function of both net confining
pressure
with and matrix swelling and Ashrinkage
gas adsorption/desorption. sub-bituminous
associated

coal sample from the Mannville Group in Western


Canada was analyzed.
In this study, permeability to helium, methane and
carbon dioxide was measured at variable confining
pressures. Permeability estimated with helium at
increasing and decreasing confining pressures
Steady-state, hydrostatic displayed hysteresis. Permeability to all gases
conditions, permeability decreased with increasing confining pressure. A k1/2
Jikich et al. 2009 to different gases at vs. lnp relationship appeared to reasonably describe
variable confining helium and methane permeabilities measured at
pressures decreasing and increasing confining pressure,
respectively.
due to sorption-induced
The same strain
was not
effects. Bituminous
true for CO2, possibly

coals from the Upper Freeport formation of West


Virginia were analyzed.

Potrebbero piacerti anche