Anda di halaman 1dari 46

International Journal of Coal Geology 101 (2012) 36–81

Contents lists available at SciVerse ScienceDirect

International Journal of Coal Geology


journal homepage: www.elsevier.com/locate/ijcoalgeo

Review article

Coalbed methane: A review


Tim A. Moore ⁎
Cipher Coal Consulting Ltd., 173 Clifton Terrace, Sumner, Christchurch, 8081, New Zealand
Department of Geological Sciences, University of Canterbury, Christchurch, New Zealand

a r t i c l e i n f o a b s t r a c t

Article history: The commercial extraction of methane from coal beds is now well established in a number of countries
Received 29 December 2011 throughout the world, including the USA, Australia, China, India and Canada. Because coal is almost pure
Received in revised form 19 May 2012 carbon, its reservoir character is fundamentally different to conventional gas plays. Coalbed methane (CBM)
Accepted 20 May 2012
forms as either biogenically- or thermogenically-derived gas. The former occurs in ‘under mature’ (b 0.5%
Available online 14 July 2012
vitrinite reflectance) coals and is the result of bacterial conversion of coal into CO2 or acetate, which is then
Keywords:
transformed by archaea into CH4. Thermogenic gas is formed as part of the coalification process and is purely a
Coalbed methane chemical devolatilization that releases CH4. Methane is primarily stored in coal through adsorption onto the
Coal seam gas coal surface; thus it is pore surface area that determines the maximum gas holding potential of a reservoir (as
Reservoir opposed to pore volume in a conventional reservoir). Although macro-, meso-, and micropores are present in
Biogenic the coal matrix, it is thought that the micropores are where most methane adsorption occurs. In many of the
Thermogenic micropores, the methane molecule may actually stretch, minutely, the pore and thus with de-gassing of the
Gas saturation reservoir, could result in matrix shrinkage, allowing opening of the fracture (cleat) system in the coal and thus
Production
enhancing permeability. The organic composition of the coal is paramount in determining porosity and perme-
Exploration
ability character and thus maximum gas holding capacity. In general, the higher the vitrinite content the higher the
Diffusion
Permeability gas holding potential (and ultimately the amount of desorbed gas) and permeability (all other factors being the
Desorption same). There are other organic component/gas property relationships but these seem to be specific to individual
Adsorption basins, or even seams.
Microbiology Characterising a CBM reservoir during an exploration programme is a challenge but the two most vital
Methanogens measures to determine are permeability and % gas saturation. Permeability will largely determine gas (and
Pores water) flow rate, dictating how commercial a prospect might be. Gas saturation, determined from desorption
Organic composition and adsorption measurements, also influences gas rate and the ultimate recoverability of gas from a reservoir.
Modelling of gas flow from the reservoir is highly dependent on knowledge of these parameters. Designing a
successful pilot well programme and ultimately production wells will rely mostly on the permeability and % gas
saturation character. Certification of resources and reserves, which is also very important to CBM companies as
they explore and develop their permits, depends heavily on accurate estimates of reservoir character; primarily
seam continuity, % gas saturation and permeability.
© 2012 Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
1.1. Previous compilations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
2. Producing fields of the world . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
3. Gas type, quality and measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.1. Biogenic gas and the microbes that create it . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.2. Generation of thermogenic gas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.3. Gas quality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
3.4. Desorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.5. Desorption accuracy and use . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.6. Adsorption isotherms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

⁎ Tel.: +64 21 246 6641.


E-mail address: tmoore@ciphercoal.com.

0166-5162/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
doi:10.1016/j.coal.2012.05.011
T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81 37

4. Geological framework . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
4.1. General rank effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
4.2. Impact of original depositional environment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
4.3. Role of structure and stress . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
4.4. Influence of external effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
5. Reservoir characterisation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
5.1. Pores . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
5.2. Sorption and diffusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
5.3. Matrix shrinkage and swelling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
5.4. Permeability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
5.4.1. Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
5.4.2. Fracture/cleat . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
5.4.3. Induced fracturing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
5.5. Gas saturation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
6. Coal composition–gas property relationships . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
6.1. Pores and permeability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
6.2. Maximum gas holding capacity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
6.3. Desorbed gas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
7. Production, modelling and water . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
7.1. Production profiles and modelling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
7.2. Co-produced water . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
8. Key reservoir measurements during exploration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
8.1. The importance of sampling protocol and mitigation of uncertainty . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
9. Enhanced production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
10. The impact of CBM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
10.1. Project viability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
10.2. From resource to reserve certification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
11. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
12. Useful terms and conversions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

1. Introduction properties are to what are the most important measurements to


take during exploration, to how to complete production wells, and
The approach taken in this review aims to satisfy an apparent need even how to produce the gas and certify the reserves. Just how different
for a current and brief summary of coalbed methane for both students CBM is from conventional gas resources should not be underestimated.
and new colleagues in the industry. Coalbed methane (CBM) is also So there is no confusion, let me say what this paper will not
known as coal seam gas (CSG), coal seam methane (CSM), and coal address. The more theoretical aspects of methane generation from
seam natural gas (CSNG)—all of which are identical for the purposes coal will not be covered in depth and nor will methane in coal mining.
of this paper. It is recognised, however, that any gas coming from Methane has long plagued the coal industry (Raymond, 1986). This
coal is not just pure methane. But the term ‘coalbed methane’ will aspect of methane from coal has been covered in other places, including
be used in this paper because it conveys the two principal words of Bibler et al. (1998), Boyer et al. (1990), Karacan et al. (2011), Sloss
utmost interest for this forum: coal and methane. This paper is slanted (2005) and United Nations (2010). Various methods are used to identify
towards the practical and is intended to be broad rather than deep. and extract methane in order to allow mining with the least cost to
The tone will be less formal than pure research papers in order to human life. Although canaries were once used to detect methane and
reach a wider than normal audience. However, it is intended that have earned their place in pop culture (e.g. Sumner, 1980), technologi-
the scientific foundations of gas in coal are clearly laid out and cal sophistication has progressed substantially in the detection and
referenced so that the reader can dip deeper where individual interest drainage of gas from coal mines. Extraction and potential monetization
may alight. of methane from current or abandoned coal mines (commonly termed
The paper will be organised with two major bounding milestones: ‘coal mine methane’ [CMM]) is, as yet, unquantified but surely has
exploration and production—but in the opposite order. By looking at significant potential, both commercially and in mitigation of green-
the production of some CBM basins first we can gain a perspective house gases (GHG). However, concepts on methane in coal as explained
of the scale and potential that CBM can manifest itself in the world's here are readily transferable from ‘virgin’ seams to those being or
markets. Towards the end of the paper, critical reservoir properties having been mined. The paper will also not deal in any kind of depth
used in the assessment of an exploration programme will be reviewed. of the engineering aspects of vertical or horizontal well completions
The middle part of the paper will explore how a coal reservoir works although the reader is referred to articles by Gentzis (2009), Gentzis
and what controls gas occurrence. Finally, two brief reviews will be et al. (2009a, 2009b), Han et al. (2009), Nie et al. (2012) and references
given, one on enhanced CBM and the other on the impacts of CBM. within.
It is worthwhile to note here that CBM is considered an ‘uncon- Any review on a subject as deeply complicated as methane in coal is
ventional’ gas resource. Though somewhat of an arbitrary designation doomed to disappoint someone. To grasp the entirety of coalbed methane
it is unconventional only because it isn't classically how oil and gas would be impossible within the constraints of a journal article. Thus it is
are extracted from clastic and limestone reservoirs. It is true, however, best to view this review as a narrative, and as in any narrative it is limited
that gas in coal is held in a fundamentally different way. Since gas to the point of view and experience of its author. It is necessarily partial
is held in such a different way in coal, than in conventional reser- and incomplete. Its author will uniquely determine its path; another
voirs, this influences everything from what the significant reservoir author would take a different path.
38 T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81

Finally, a note on the units of measure that are used throughout emissions and modelling, among other subjects. This was the first
this paper. The petroleum industry and people in the USA tend to stand-alone book that took such an encompassing view of CBM and,
use imperial measurements whereas others tend to use metric. This importantly, discussed microbial enhancement of CBM (Scott, 1999),
paper adopts the metric system for most units (for example gas charge probability analysis of CBM properties (Zuber, 1999), gas saturation
is expressed in m 3/ton) though for gas flow and resource/reserve es- in coal reservoirs (Khavari-Khoransani and Michelsen, 1999) and
timates the imperial system is used (for example MCF [thousand pore space in coals specifically applied to CBM reservoirs (Radlinski
cubic feet]). There are various reasons for this but mostly because and Radlinska, 1999).
those units are generally accepted and known throughout the CBM in- By the mid 1990s CBM production from high-rank coals in the USA
dustry. The author has found that most people in the CBM industry (i.e. the Black Warrior and San Juan basins) began to level off (see
can use either metric or imperial when it comes to gas contents but gen- Section 2) which accelerated exploration and development in
erally always refer to flow rates and reserves exclusively in imperial. In low-rank coals such as those in the Powder River basin of Wyoming
any case, the last section of this paper (Section 12) gives some useful and Montana. These developments in turn catalysed research into
conversions. mechanisms of gas formation in low-rank coals. The special volume
edited by Flores (2008) was the first internationally available set of
1.1. Previous compilations papers dealing exclusively with CBM in low-rank coal and thus
addressing some of the issues of the genesis and implications of bio-
It would be remiss not to identify previous CBM compilations and genic gas. The major difference between the first set of papers edited
reviews. I will admit that most of this review will be English by Flores (1998) and this later set was the rise in CBM production
language-centric in its references and I apologise for the limitation. from what would be termed ‘non-mature’ coals. Many of the papers
Also, there is no doubt that other regionally produced proceedings in this volume address the methanogenic consortium in coals and
addressing CBM are not listed here, but this in no way reduces their describe laboratory experiments to determine their biokinetics. Though
significance. much work is still to be done, this set of papers heralds an important
The oil crisis of the 1970s in the USA spurred a focused interest in step in defining and understanding a type of methane generation that
potential gas resources derived from coal beds. Rightmire et al. (1984) a decade ago was barely recognised in the petroleum industry. An
give a good summary of reports that were commissioned by the U.S. important finding revealed in this set of papers was that methane can
Department of Energy on thirteen basins in the USA. Rightmire et al. be generated in stages and is not simply biogenic or thermogenic (see
(1984) make the case that these reports, republished together in Section 3.1) (Flores et al., 2008).
1984 as part of an AAPG Special Publication prompted a deepening Other compilations of papers exist and the reader is directed to four
interest in development of CBM in the USA. other special volumes, the first of which was edited by Lyons (1998)
One of the first major compilations of research on gas in coal, how- and focuses on the reserves and geological controls of Appalachian
ever, was published almost a decade later (Law and Rice, 1993) and CBM. The second is the volume by Collett and Barker (2003), which
covered virtually all aspects of hydrocarbon generation in coal. Most discusses coalbed methane properties from the Ferron coals of Utah.
papers in the volume delve into the detailed foundation, origins and The third compilation, by Karacan et al. (2009), discusses aspects of
influences on CBM. Notably, the paper by Levine (1993) outlines the CO2 sequestration and methane recovery in coal beds and the last, by
aspects and processes of coalification of organic material and the Golding et al. (2010), has papers covering topics from Appalachian
pathways to formation of CBM. Similarly, the paper by Rice (1993) basin resources to controls on coal cleating.
clarifies the origins of coal gas and its formation, as understood at In addition to the above compilations, two institutions in North
the time. Flores (1993) also outlines the depositional setting of gas America have made significant contributions to research into CBM.
producing sequences, notably presaging the advent of significant pro- The Gas Technology Institute in Illinois (formally known as the Gas
duction from the Powder River Basin. Other papers in the compilation Research Institute [GRI]) has produced countless reports and presenta-
deal with natural fractures in coal (Close, 1993), the mechanisms of tions on all aspects of gas in coal and the reader is encouraged to visit
gas sorption (Yee et al., 1993) and methane in underground mining their web site and browse their publications (www.gastechnology.
(Diamond, 1993) to name just a few. If I were to recommend one set org). Similarly, the University of Alabama has run the International
of papers to read to a colleague new to CBM, it would be this volume. Coalbed Methane Symposium for decades and much new research
Another compendium of papers was published by the Geological and understanding has come from those proceedings (www.coalbed.
Society (Gayer and Harris, 1996) and covers various aspects of methane ua.edu).
resources (e.g. the calculation of resources), mainly in Europe but it also There are also some non-aligned, independent, broad-interest and
contains brief discussions of plays in the USA. In addition, there are review CBM papers and the reader is directed to Ayers (2002), Bell
several papers that report on the coal character, such as mineralisation, (2006), Busch and Gensterblum (2011), Bustin and Clarkson (1998),
in relation to reservoir properties. In almost all cases these papers focus Liu et al. (2011), Pashin (1998), and Scott (2002). Finally, although
on high-rank coals as CBM producers. A geographic summary of CBM the paper by Laxminarayana and Crosdale (2002) is primarily about
characteristics for Russia, Germany, Ukraine and the UK characterises sorption character of selected Indian coals, it does discuss and present
the coal reservoir in terms of such things as resource size and the poten- some fundamental, and certainly quite broadly applicable, gas–coal
tial for development. The paper by Levine (1996) is an early work on relationships.
coal shrinkage in relation to gas production and permeability changes;
an aspect of a CBM reservoir which is still not fully understood. 2. Producing fields of the world
Later, an evaluation of CBM as a danger and a resource was collated
by Flores (1998). These papers addressed aspects of underground coal Wherever there are coal basins, it is inevitable that there will be at
mine outbursts and emissions, as well as sorption characteristics, per- least some methane. But in many cases, monetising that gas is prob-
meability, stratigraphy and resource assessment. The composition of lematic because of unfavourable reservoir properties, barriers to
coal derived methane gas was studied (Clayton, 1998)—a subject market, or a total lack of markets (think basins in remote jungles, or
returned to a decade later, in a slightly different way, in a volume of high latitude arctic or Antarctic regions). There are, however, some
special papers by Flores (2008). stellar examples where the right geology, geography and economics
A broad reaching book edited by Mastalerz et al. (1999) covered have combined to produce a vibrant industry.
many aspects of CBM including regulatory regimes, resource assess- Even when there is abundant production, it is not easy to get accurate
ments, controls on CBM occurrence, reservoir evaluation, methane data on gas volumes. As most of us realise, hydrocarbon reserves directly
T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81 39

influence a country's economic standing. Thus, there is always potential amount of water produced. Over 678 million barrels of water was
for ‘gaming’ of the numbers. With this caveat in mind, and in order to produced in 2006 alone and since 1984 a total of 6.5 billion barrels of
give an indication of the scale and importance of coal-derived methane water have been produced (Fig. 4). Much of this water is treated and
on world economics, some examples are given below. discharged (see Section 7.2).
Four countries will be discussed: the USA, Australia, China and With its huge high-rank coal reserves and numerous underground
India. These countries are currently producing and selling CBM as mines, gas has always been a mining hazard in the coal fields of Australia.
pipeline gas, or utilising for electrical generation directly on or near A number of early trials at development of CBM by major oil companies
site. In all cases, the gas is used domestically. Canada is also a fairly failed but by 1996 production began with significant ramp up occurring
major producer of CBM although will not be discussed here because in 2001 (Fig. 5). Practically all of the gas comes from either the Bowen or
the USA and Australia already give ample examples of major fields; the Surat basins in Queensland, however some development and pro-
China and India are briefly mentioned to highlight emerging CBM duction occur in the Sydney Basin in New South Wales (Fig. 6). Although
markets. Other countries on the verge of commercial CBM production, a significant amount of Australia's production comes from coal of high-
are Indonesia and Russia but will not be discussed in this paper. rank, the Surat basin, which is mostly subbituminous in rank, does con-
The USA currently has the largest CBM production in the world tribute significantly to overall gas production. Finally, there have been
with over 1.91 TCF of gas sold in 2009 (Fig. 1; see also Pashin, 2011). exploration activities in the brown coals of Victoria, but no production
The majority of CBM production has come out of three basins: The has occurred to date.
Black Warrior (Alabama), San Juan (New Mexico, Utah, Colorado) Reliable production data for CBM in China is elusive, therefore
and the Powder River (primarily Wyoming) (Fig. 2). The fourth values reported here should only be taken as indicative at best. Simi-
basin, the Raton, has also produced CBM but is smaller and more larly, gaining information on the number and types of wells is virtually
restricted than the other USA basins thus will not be discussed. impossible, but the general feeling in the industry is that thousands of
In the USA commercial production began in the Black Warrior development wells have been drilled, but commercial utilisation of
Basin in the early 1980s but by 1989 the San Juan Basin was the the gas is localised as compared to the large integrated network of
major producer (Fig. 3). Since the mid 1990s, annual production in CBM pipelines seen in the USA, Australia and Canada. What is known
the Black Warrior Basin has stabilised around 115 BCF. CBM in the is that practically all of the production either comes from the
San Juan Basin peaked in 1997 with an annual production of high-rank coals in the Ordos or the Qinshui basins (Fig. 7) although
597 BCF, but has since declined and, as of 2009, was approximately some CBM production can be found almost anywhere in China
432 BCF per annum. Although most of the significant production in where there is coal and people, which covers much of the country.
the San Juan Basin began in the late 1980s, there was an early and Data reported in the general press has CBM production in 2006 at
spectacular well that began in the mid 1950s—the San Juan 32–7 well just over 1 BCF and by 2010 at 51 BCF (Fig. 8). Although most certainly
reportedly produced methane for over 40 years, though generally at inaccurate and underestimating total production, these values probably
rates around 100 MCF/day (Murray, 1996). Although both the San do reflect the rate of increase in large CBM production projects in China
Juan and Black Warrior basins are in long term decline at present, over the last five years. The other certainty is that CBM in China, because
they are expected to produce significant methane for a decade or two of both the size of its coal resources (estimated to be in the trillions of
to come. tons) and its voracious energy needs, will likely increase to rival other
Production through the 1980s in the USA saw double digit annual producing countries in the world.
percentage increases which fell to less than 5% by 1999 (because of Like China, India has a huge population and growing economy and
the decline in production in the Black Warrior and San Juan basins). is looking for ways to maximise its energy resources. Therefore, India
However, by 2000 the Powder River basin ramp-up (Fig. 3) resulted has had an aggressive initiative for the last few years to explore and
in significant annual increases in the overall USA CBM production. develop its CBM resources through international tendering processes.
Although initial production in the Powder River basin started as early Commercial production has started and is believed to have been just
as 1984, it took over 15 years before the right combination of technology over 2 BCF in 2009 (Fig. 9). Again, the data most likely under represents
and economics resulted in previously unheard of development rates. the actual CBM being produced and sold.
Over seven years, the number of wells drilled in the Powder River A final note on CBM production; over the last few years there has
basin went from a few dozen to over 500 per annum (Montgomery, been a race to feed LNG plants with methane from CBM prospects.
1999) and in 2001 up to 3655 wells were drilled (Ayers, 2002). Although This is mainly occurring in Australia where it is likely that the first pur-
the annual CBM production in the Powder River basin has dropped in pose built, relatively large scale, plant will take coal seam-derived gas in
the last two years, the field is still expected to grow and produce signif- the next 5 years. A similar race is also ensuing within Indonesia. The
icant methane for a decade or two. The basin has already produced carrot for all the companies is the relatively high export price paid for
4.3 TCF of gas since 1984 (Wyoming Oil and Gas Commission, 2011). A LNG; but in order for this to work economically, a large resource has
defining attribute of the Powder River basin CBM production is the to be able to be developed and sustained over a large number of years.

3. Gas type, quality and measurement

CBM is primarily of interest because of its energy potential; therefore


this review considers the gas itself up front. Reviews of origins of gas in
coal and the composition or ‘quality’ of the gas have been conducted by
Clayton (1998), Hunt (1979), Rice (1993), Rice and Claypool (1981) and
Whiticar et al. (1986). The reader is directed to these papers, and the
references within, for an additional perspective and an in-depth exam-
ination of gas formation and gas composition in coal beds. To start out
with, however, we can consider that there are two primary origins of
CBM: biogenic and thermogenic (Fig. 10). The gas in the coal seam
Fig. 1. Coalbed methane (CBM) production in the USA from 1989 to 2009. Source: U.S.
may contain varying proportions of methane relative to other gases,
Energy Information Administration, Washington, DC; http://tonto.eia.doe.gov/dnav/ referred to as gas quality (see Section 3.3 for further discussion on gas
ng/hist/rngr52nus_1a.htm. quality).
40 T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81

Fig. 2. Map showing locations of USA basins discussed in this paper.

From a practical perspective, purely biogenic gas plays are almost papers in Flores, 2008 as well as Midgley et al., 2010, Penner et al.,
always lower in gas content than thermogenically-derived CBM. Gas 2010 and Strąpoć et al., 2008a).
contents are rarely above 4 to 6 m 3/t in biogenically derived A number of recent studies have categorised microbial taxa found
resources though this is not a hard and fast rule. High-rank coals can in coal-derived water and coal matrix that have been deeply (500 to
be in excess (though rarely) of 20 m3/t. Remember too that the transi- 1000 m) buried (e.g. Green et al., 2008; Hendry et al., 2007; Li et al.,
tion from biogenic to thermogenic is not instantaneous and there will 2008; Midgley et al., 2010; Penner et al., 2010; Shimizu et al., 2007;
no doubt be mixing of biogenic and thermogenic gases (see next Section Strąpoć et al., 2008a). Usually the microbial assemblage found in
below). And finally, within a single basin, the shallow coal beds may be coal will consist of hundreds of taxa but it is generally believed that
biogenic plays whereas the deeper coals may be thermogenic (e.g. there is only a smaller subgroup, aptly termed methanogens, which
Hackley et al., 2009). metabolise the methane (Strąpoć et al., 2008a). This paper won't
detail the specific taxa found in coals, but three areas will be discussed:
3.1. Biogenic gas and the microbes that create it 1. General taxa present, 2. The methanogenic process of making methane
and 3. Indicative environmental conditions needed for microbial growth.
It is now recognised that life below the surface of the earth is vast In terms of the taxa present, creation of biogenic methane from
(Gold, 1992) and could account for half of Earth's biomass (http:// coal requires the cooperation of organisms from two out of the
www.deepscience.org/contents/dark_life.shtml). Mostly, life below our three Domains of Life—bacteria and archaea (note that the three
lawns consists of microscopic organisms (microbes) (Fig. 11) which Domains [Superkingdoms] of Life are: Eukarya, Bacteria and Archaea
play a huge role not only in decay of plant material (Rogoff et al., 1962; (Woese et al., 1990). Plants and animals (humans among them) lie
Waksman and Stevens, 1929) but also in the generation of methane in within the Eukarya branch of life). Collectively, the organisms involved
peat and low-rank coal (Strąpoć et al., 2008a; Wang et al., 1996). are referred to as methanogenic consortia because there are literally
When methane is derived from microbes, it is said to be biogenically hundreds of species of both bacteria and archaea that have their own
derived. It is interesting to read Rogoff et al. (1962), which is an early special role in making methane (Green et al., 2008). Bacteria begin the
review of the microbiology of coal. The questions and doubts raised in process and archaea complete it (Fig. 12). There is a developing body
that paper are now just beginning to be addressed and answered (see of research on biogenic gas formation; however it is recognised that
there are many more details still to be understood. The following
discussion is gleaned from a few papers (Green et al., 2008; Midgley et
al., 2010; Shimizu et al., 2007; Strąpoć et al., 2008a; Ulrich and Bower,

Fig. 3. Coalbed methane production profile for the Black Warrior basin, USA from 1980
to 2009, San Juan basin, USA from 1989 to 2009 and the Powder River Basin, USA 1985
to 2010. Source: Energy Information Association http://www.eia.gov/dnav/ng/hist/
rngr52sal_1a.htm, and Wyoming Oil and Gas Conservation Commission: http://wogcc. Fig. 4. Co-produced water from the Powder River basin from 1985 to 2010. Source:
state.wy.us/. Wyoming Oil and Gas Conservation Commission: http://wogcc.state.wy.us/.
T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81 41

increasing the temperature even a little (22 to 38 °C) increased labo-


ratory methane production by 300%. Similarly, decreasing the pH
from 7.4 to 6.4 resulted in a 680% increase in CH4 laboratory produc-
tion. Finally, when particle size was decreased (and thus surface area
increased) methane production increased by 200%. Rice and Claypool
(1981) also suggest that for methanogenic microbes to thrive, the
formation waters have to be low in sulphur. Recently, Papendick et
al. (2011) found that increases in both hydrogen content and surface
area may encourage microbial activity producing more methane.
These early laboratory results seem to indicate that reservoir condi-
tions of water pH, abundant surface area and temperature have to
Fig. 5. CBM Production in Australia from 1996 to 2009. Source: http://www.
be just right before the methanogenic process can be of relevance in
australiamines.atlas.gov.au and http://www.ga.gov.au. gas production. Variation in these reservoir conditions may explain
why some basins are prolific biogenic gas producers (Powder River
Basin, Wyoming [Flores et al., 2008]; Surat Basin, Queensland [Li et al.,
2008]), and others seem to be quite variable in gas contents (Huntly
2008; Wang et al., 2010a; Whiticar et al., 1986) and greatly oversim- Coalfield, New Zealand [Mares and Moore, 2008]) or have essentially
plifies the process of gas formation; however, it should give the reader no biogenically-sourced gas at all (Hanoi basin, Vietnam; as indicated
at least a “mostly true” (at this point in time!) understanding of how from Dart Energy's relinquishment; Dart Energy Ltd., 2012).
bacteria and archaea work to form methane. A review of this same sub- Biologically-produced methane can form at almost any time during
ject can also be found in Gilcrease and Shurr (2007). a coal's life cycle. However, if coals are exposed to high heat with
It is believed that bacteria begin the initial fragmentation of the resultant rank increases, it is thought that the methanogenic consortium
coal macro-molecule through two main processes: fermentation may be killed off (Aravena et al., 2003; Strąpoć et al., 2008a, 2008b). It is
and/or anaerobic oxidation. The fermentation of polyaromatics and thought that the consortia of microbes needed to produce methane can
carbonylic acids results in acetate whilst ketones are fermented to survive at temperatures greater than 55 °C (Scott et al., 1994). This does
H–CO2 compounds. In contrast, different bacterial species anaerobically not, apparently, preclude the reoccurrence of methanogenic methane
oxidise the aromatic and aliphatic structures into CO2. Additionally, formation after heating, if the coal is uplifted and favourable subsurface
through a process of homoactogenesis, some H–CO2 compounds may environmental conditions are restored. An important criteria in the
be metabolised into acetate by bacteria. With the fermentation and re-formation of biogenic gas is that the coals are re-inoculated, through
anaerobic oxidative processes complete, archaea take over. A number water recharge, with methanogenic consortia (Clayton, 1998; Rice, 1993;
of studies have found that archaea methanogens create CH4 through Ulrich and Bower, 2008). Shimizu et al. (2007) found methanogenic
either an acetoclastic or CO2-reduction methanogenesis process (e.g. archaea in a thermogenic gas play in Japan and interpret that they may
Green et al., 2008; Hendry et al., 2007; Penner et al., 2010 among be contributing to the overall gas volume of the seam. Similarly, Ayers
others). (2002), Faiz et al. (2003), Flores et al. (2008), and Smith and Pallasser
The environmental factors limiting bacteria and archaea in a bio- (1996) have all shown evidence that there has been more than one epi-
genic system are not well understood. It is known that archaea are sode of biogenically produced methane in the San Juan, Sydney, Bowen
strictly anaerobic micro-organisms (Green et al., 2008) and are con- and Powder River Basins, respectively. In contrast, Strąpoć et al. (2007)
sidered extremophiles as some species thrive in high salt and heat found that high-rank coals in Kentucky were not secondarily enriched
conditions. In laboratory experiments, Green et al. (2008) have in biogenic gas because the present geological conditions, structural in
shown that the quantity of biogenically-derived CH4 is influenced nature, have prevented hydrologic recharge, and thus re-inoculation of
by three factors: temperature, pH and surface area. They show that the seam with methanogenic consortium. It is conceivable however,
that in high-rank coals (e.g. semi-anthracites), the organic fraction has
been transformed into material not readily useable by the microbes
(Rice and Claypool, 1981).
Understanding the controls on biogenic processes in the formation
of methane in coal reservoirs is important as it may provide an alter-
native energy source that potentially could become ‘renewable’ as
long as there is a coal seam. Already a number of laboratories are
beginning to experiment with how to ‘inoculate’ a coal reservoir with
the right microbial consortium with the expressed purpose of genera-
tion of methane in real time (e.g. Martini et al., 2005; Ulrich and
Bower, 2008).

3.2. Generation of thermogenic gas

Simply put, when coal attains a rank of approximately 0.5–0.6%


vitrinite reflectance (high volatile bituminous), thermogenic gas
begins to evolve (Clayton, 1998). Gas evolution results from a combina-
tion of time, heat and pressure. These factors cause devolatilisation and
production of compounds and gases such as methane, carbon dioxide,
nitrogen, hydrogen sulphide as well as higher hydrocarbon chain
gases (e.g. ethane and propane). Onset of thermogenic gas generation
will have variable timing, in relation to coal rank, depending on organic
composition and specific basin and burial history (Whiticar, 1994).
Based on bulk organic composition as well as formation tempera-
Fig. 6. Coal basins in Australia. ture, theoretical volumes of gas, and gas type can be predicted (Hunt,
42 T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81

Fig. 7. Coal basins in China discussed in this paper.

1979; Fig. 13). However, laboratory experiments indicate that kinetic may be two fold: Firstly, their holding capacity is greater (mostly
models, in mathematical isolation, probably underestimate the amount because of lack of moisture; see Section 4.1) and secondly,
of gas derived from organic matter (Butala et al., 2000). Zhang et al. devolatilisation of coal will kinematically produce more gas than bio-
(2008) suggest specific kinetic and/or laboratory predictive models genic processes. This does not mean that high-rank coals will always
must be confined to individual basins. In their research, vitrinite reflec- have more gas than low-rank coals; de-gassing because of pressure
tance is related to gas generation volume and type (Fig. 14). They found, loss or elevated temperatures, and thus lowering of the gas saturation
using samples from the Piceance basin that with increasing vitrinite of a reservoir may occur because of uplift, tectonic fracturing and
reflectance, CO2 generation decreased substantially whilst CH4 and migration of gases or other factors (see Section 5.5 for the definition
higher hydrocarbons increased. This supports a previous study by Behar and controls on gas saturation).
et al. (1995) that found good experimental evidence that gas yield
increased with rank in coals. Other researchers, however, have suggested 3.3. Gas quality
that gas generation as modelled in laboratory experiments tends to be
less than what has been observed in actual coal beds (Clayton, 1998). In CBM, ‘gas quality’ refers to the types of gases inherently present
What is clear, however, is that coal beds that are high enough in in the reservoir, that is, the composition of the gas. The word quality is
rank to produce and retain thermogenic gas tend to have higher used as it is the various proportions of the mixed gases that determine
total gas contents than coals with purely biogenically-derived gas. Of the value of the gas as a commodity. Typically, gases in a CBM reservoir
course, the reasons for higher rank coals generally having more gas can be divided into two groups: ‘productive gases’ (CH4, C2, C3, etc.)
and ‘inert gases’ (CO2, N2, H2S). Productive gases are those that add
economic value through their heating content upon combustion and
inert gases simply do not combust and thus take away value (on a
volume basis). Note that C1, C2, C3 are short-hands commonly used in
the gas business and academic circles to refer to methane, ethane and

Fig. 8. CBM production in China from 2006 to 2010. Source: http://www.kczywm.cn/


kuangyexingqing/200810/23-16_2.html, http://www.kczywm.cn/kuangyexingqing/
200810/23-16_3.html, http://www.kczywm.cn/kuangyexingqing/200810/23-16.html,
http://www.gov.cn/jrzg/2009-12/22/content_1494088.htm, and Hongmei Li, personal Fig. 9. CBM production in India from 2000 to 2009. Source: Mr Sudhansu Adhikari,
communication, 2011. personal communication, 2010.
T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81 43

Fig. 10. Schematic showing biogenic and thermogenic gas generation in relation to rank, moisture, vitrinite reflectance and coalification stages.

propane, respectively (Whiticar, 1994). Most low-rank CBM reservoir 1979). Hydrogen sulphide or H2S is relatively rare in CBM plays, howev-
gases are ‘dry’, as opposed to ‘wet’ gas; the definition of dry or wet gas er, even small quantities can have a profound effect on how the gas is
is if it has less than 1.3 or >4 l/100 m 3 of condensable liquids (Hunt, handled. The gas is incredibly toxic, with even a fraction of a percent
being fatal, as well as extremely corrosive, which has implications for
casing and pipelines. Thus, there is usually zero tolerance for H2S in
any gas stream. The terms ‘sweet’ and ‘sour’ refer to gases that are low
and high, respectively, in H2S (Hunt, 1979). H2S is derived primarily
through either the devolatilisation process of coalification (Fig. 14;
Hunt, 1979) or through bacterial sulphate reduction (Bustin, 2004). Bac-
terial sulphate reduction can occur through natural processes or, if

Fig. 11. Scanning Electron Microscope photographs of methanogenic microbes from Fig. 12. Simplified pathway of secondary biogenic methane production in coal. See text
coal, sourced from the Powder River Basin. Courtesy of Luca Technology. for further explanation and also Strąpoć et al. (2008a).
44 T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81

upwards through multiple seams (Faiz et al., 2007a). Furthermore, in


the Bulli seam within the same basin, ethane proportions vary signifi-
cantly in relation to the depth of the coal seam (Faiz et al., 2003).
Creedy and Pritchard (1983) and Rice (1993) also found a relationship
between depth and gas quality, though this doesn't appear to be univer-
sal across basins. An interesting association in the San Juan Basin exists
where coals seams below the topographically higher recharge areas
have the highest CH4 concentrations (and importantly low CO2; Scott
et al., 1994). This has been postulated to be a result of water differentially
flushing the more mobile CO2 down dip, effectively concentrating the
CH4 (Cui et al., 2004). Some production companies do note that CO2
will increase at the expense of CH4 with continued degassing of the
seam; in part this might be the result of different desorption kinetics
between CO2 and CH4 (Cui et al., 2004).
Care does have to be taken when analysing for gas quality as some
gases may be introduced or created during the sampling process or
even be altered from the containment vessel itself (van Holst et al.,
2010). For example, if canisters are not constructed out of material
which are inert and non reactive, it has been shown that significant
H2 may be produced (Faraj and Hatch, 2004). Nitrogen can also be
an introduced contaminant (as part of atmospheric air) during gas
quality analysis (Kim 1973). There is no doubt that N2 exists naturally
within some higher rank coals (Creedy and Pritchard, 1983; Hunt,
1979) but in some cases O2 may react with the coal itself and be pref-
erentially adsorbed, with the effect of leaving excess N2 in the gas
mixture (Creedy and Pritchard, 1983; Havton, 2003; Jin et al., 2010).
Fig. 13. Timing of gas type generation in relation to temperature for humic type organic It is important that this relic N2 is not mistaken as being inherent
matter. Also showing approximate H/C ratios with burial stages. Modified from Hunt within the seam.
(1979) and Zhang et al. (2008).
3.4. Desorption
things go wrongly, through use of bacterially contaminated drilling
or fracturing fluids introduced into the reservoir; in the latter case, The amount of actual gas (also referred to as ‘gas charge’) in a coal
cleaning up the reservoir can be time consuming and expensive. bed can be estimated through taking a sample of fresh coal and
In general, the gases found within a CBM reservoir are greatly desorbing gas from it over a period of time. When a sample of coal is
influenced by rank (see Section 4.1). Because of the nature of biogenic brought to the surface and placed in a canister some of the gas will be
processes, low-rank coals contain mostly CH4 and usually only small released from the coal. At depth, reservoir pressure, which is mainly
quantities of CO2. However, in some plays there may be spatial variation from hydrostatic head, keeps gas molecules physically adsorbed onto
in CH4 and CO2 because of the kinetics of differential gas transport. This the surface of the coal micropores (see also Section 5.2 and Busch and
variation can have economic ramifications over the production life of a Gensterblum, 2011 for more detail on sorption properties). The zero
basin (Cui et al., 2004). hydrostatic pressure at the Earth's surface allows the gas to desorb out
Some CBM plays may have a multitude of gases present whose of the coal and thus be measured. This process is called desorption.
relative concentrations vary greatly laterally and vertically within the Some procedures dictate that desorption should take place at the equiv-
basin. For example, in the Sydney basin CO2 concentrations increase alent reservoir temperature that the sample came from (e.g. American
Standard Testing Material [ASTM] D7965-10, 2010) whilst other proce-
dures place less emphasis on that requirement, relying more on math-
ematical corrections to simulate reservoir temperature (e.g. AS 3980–
1999, Anonymous, 1999; Saghafi et al., 1995; Williams, 1996).
Some service companies are now offering in situ gas measurements
that determine the critical desorption pressure (CDP) without the need
for core. As this technology is just being commercially trialled, the reader
is directed to www.welldog.com and www.weatherfords.com for addi-
tional information on these procedures. See also Lamarre and Pope
(2007) and Carlson et al. (2008) for a discussion. It is also possible to
use well cuttings for desorption (ASTM, 2010; Newell, 2007), though
these will yield less certain results.
Some of the earliest modern desorption procedures were developed
by the U.S. Bureau of Mines in the 1970s and early 1980s (e.g. Diamond,
1978; Diamond and Levine, 1981; Kim, 1973, 1977; Kim and Douglas,
1973; McCulloch et al., 1975). The basic concept is that coal samples
are brought quickly to surface (preferably using a wire-line coring
system), placed in a leak proof canister of roughly the same size as the
sample, and over a period of time the gas that is evolved is measured.
Since those early studies, there have been improvements (e.g. Barker
et al., 2002, Crosdale et al., 2005; McLennan et al., 1995; Moore and
Fig. 14. Changes in relative volumes of generated gases to vitrinite reflectance; from Butland, 2005; Stricker et al., 2006; Yee et al., 1993; also see Diamond
experimental pyrolysis data. Modified from Zhang et al. (2008). and Schatzel, 1998 for an earlier review), but the essential steps have
T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81 45

remained the same. The most recent standard for desorption to be pub- regression is typically used for calculation of lost gas, some researchers
lished is that of the ASTM (ASTM D7965-10, 2010). The basic equipment advocate a polynomial regression (Waechter et al., 2004a, 2004b),
needed for desorption are canisters, and appropriately sized manome- which may be more applicable in high permeability coals. The effect
ters to measure the desorbed gas (see Fig. 15). Various spread sheets of using a polynomial regression can be seen in Fig. 16b, which increases
for recording gas desorption measurements are given in McLennan et the estimated lost gas to 0.53 m 3/t. Lost gas estimation is one of the
al. (1995), Barker et al. (2002) and Barker and Dallegge (2005). areas that contributes to the overall uncertainty in gas desorption
In most desorption procedures, there are three measurements, or results. Estimation of lost gas is especially important in underground
calculations, that are made: mines which are gassy. Prediction of mine gas outbursts requires char-
acterisation of initial desorption rate and accurate measurement of lost
1. Lost gas,
gas is integral to that assessment (Beamish and Crosdale, 1998;
2. Measured gas, and
Williams and Weissmann, 1995). Note that gas outbursts are rarely
3. Residual gas.
recorded in low-rank coal, and appear to be predominantly high‐rank
Lost gas (sometimes referred to as Q1) is the calculated amount of coal phenomena (Beamish and Crosdale, 1998; Lama and Bodziony,
gas that has potentially been lost from the time the core leaves the 1998).
reservoir at depth to when it is placed within a canister at the surface. Measured gas, for the purposes of this paper, is defined according to
Depending upon the particular procedure, a number of closely spaced the ASTM standard (ASTM, 2010) as the gas which is directly desorbed
gas readings are made over the first hour or more and the regression from the canister and measured using a manometer. The Australian
of those gas volume measurements is plotted against the square root Standard (AS 3980–1999) refers to this as “Q2” (Anonymous, 1999;
of time (Bertard et al., 1970). The Y-intersect of that line estimates the Beamish and Crosdale, 1998). The cumulative gas over this period is
amount of lost gas. For example, in Fig. 16a a linear regression line plotted against the square root of time (e.g. Fig. 17). The shape of the
indicates a lost gas of about 0.26 m 3/t. Caution should be taken when curve can be significantly different between coal beds and even be-
doing lost gas estimation as even a slight interpreted change of slope tween coal samples from the same coal bed. The length of time, that is
can lead to significant differences in lost gas amount. Although a linear the number of measurements taken for measured gas, differs between
individual operators. If an accurate measurement of residual gas is
needed (see below), then desorbed gas must be measured until there
is essentially zero gas being desorbed (a standard being that any reading
must be less than 1% of cumulative gas desorbed to date). This may take
some time in low-rank coals, on the order of months in some cases. That
length of time may be unacceptable to some users and therefore a
number of other procedures have been developed. For example, in
the case of coal mine degassing it is essential that results be obtained
almost immediately thus ‘fast’ desorption procedures are used (e.g. AS
3980–1999, Anonymous, 1999; ASTM D7569-10, ASTM, 2010). Some
operators may use a compromise between ‘fast’ and ‘slow’ desorption
which captures enough information to allow desorption rate to be
estimated but accommodates the need for relatively rapid gas volume
estimates for commercial and operational decisions.
It is important to note that low-rank coals are susceptible to produc-
ing their own gas, especially if there has been contamination within the
canister before sealing. Thus, some care and caution must be taken when
desorbing coals for long periods of time. As has already been stated in
Section 3.3, reactions with certain kinds of canister materials can pro-
duce gas unrelated to the original gas within the coal (Faraj and Hatch,
2004).
After a sample is either ‘fully’ desorbed or when the operator
decides that enough has been measured to achieve the desired goals
of the sampling, the coal is removed from the canister and a subsplit is
tested for residual gas (or Q3, in the Australian Standard; Anonymous,
1999). The apparatus for testing residual gas has been described else-
where (Anonymous, 1999), but is essentially a method for crushing
coal to approximately 250 μm size particles, which has gas tight seals
and a gas lead from which the gas evolved can be removed for measure-
ment (Fig. 18). It is important to note that residual gas is only truly
‘residual’ if the measured gas has more or less been fully desorbed. In
either the ‘fast’ desorption method (Anonymous, 1999) or with canisters
that have been decommissioned early, ‘residual’ gas will only represent
the remaining gas and not necessarily gas which might be difficult to
access for CBM recovery. However, when total gas volume is reported,
it is given as the sum total of the lost gas (Q1), measured gas (Q2), and
residual gas (Q3) (note that in the Australian Standard ‘measured gas’
[Qm] is equivalent to the ‘total’ gas of ASTM). Before the gas volumes
can be reported, however, the volume of coal in each canister must be
measured and coal quality parameters such as moisture and ash yield
determined (see ASTM D7569-10). It is crucial that when viewing
Fig. 15. Photograph of typical PVC desorption canister and manometer used in the desorption results the analysis basis is always noted (for example,
measurement of desorbed gas from core. ‘as-received’ or ‘dry ash-free’).
46 T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81

Fig. 16. Example of (A) a linear regression indicating 0.26 m3/t of lost gas (Q1) and (B), on the same data, a second order polynomial regression line indicating 0.53 m3/t of lost gas.
Which is the correct way of estimating lost gas is still a debated topic.

Fig. 17. Example of measured gas (Q2) in a desorption isotherm test.


T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81 47

As most exploration geologists know, desorption data is often not


available, yet, some estimate of GIP is still required for commercial
assessments. When this is the case, an alternative method of assess-
ment is described in Moore and Friederich (2010) and Moore et al.
(2009).
Gas in-place values should always be reported on an as-received
basis or more importantly not on a dry ash-free basis, which is far
from the real reservoir character. Reporting GIP on a dry ash-free
basis will overstate the actual amount of gas present. Caution is also
needed in how GIP is interpreted from a gas producibility perspective.
Some resources may have a high gas in-place but relatively low gas
saturation making economic rates of recovery unlikely because by
the time the reservoir is de-watered there will insufficient pressure
to flow the small amount of gas (see Section 5.5 on gas saturation);
the marrying of high gas in-place with low gas saturations can be
especially prevalent in geologically young, low-rank coals. Note that
recovery rates and GIP are two very different measurements. GIP
Fig. 18. Photograph of ring mill manometer used in the pulverisation of coal and just gives an estimate of how much gas is predicted to be in the res-
measurement of residual gas (Q3).
ervoir whereas the amount of economic recoverable gas is a function
of permeability, gas saturation, gas price, area actually assessable for
development etc.; all of these aspects are covered in later sections
of this paper. Lastly, desorption contents are not to be used as an
3.5. Desorption accuracy and use indication for flow rate of a well for all the reasons given above.
With the exception of one study (Bodden and Ehrlich, 1998) there
The accuracy of desorption measurements is hard to estimate, for the have been no studies linking desorption properties with individual
mere fact that a sample can't be desorbed twice. As mentioned previously, well flow rates, though this is an area of study which is sorely needed
lost gas is unquestionably a source of error in total gas estimates. To a de- from a predictive perspective.
gree, the uncertainty can be quantified for lost gas from running differing
regression line scenarios. Similarly, residual gas can be assigned a vari- 3.6. Adsorption isotherms
ance, if two or more samples are taken from a single canister (but of
course this will not assess instrument error, only partially the sample var- As previously stated (and covered in more detail in Section 5.2,
iability). However, the bulk of the total gas will usually be the measured Busch and Gensterblum, 2011 and Rice, 1993), methane is adsorbed
gas and this cannot be ‘re-desorbed’. Thus there are no truly quantifiable on to the internal surfaces of coal (and thus surface area becomes of
ways to determine error and repeatability. Note too, that the amount of extreme importance in the sorption properties of coal and distin-
desorbed gas will be highly sensitive to the amount of inorganic material guishes it from clastic reservoirs). We've just covered, above, how
and moisture in the coal as well as the accuracy of the measurement of to measure the amount of methane that occurs within a coal seam.
the bulk volume of coal in the canister. Mavor and Nelson (1997) and However, the amount of methane found within a coal may not be
Nelson et al. (2000) do make some estimates on the reliability of desorp- the amount it can hold and thus a gas adsorption test, conducted in
tion measures and they estimate a factor of ∓15%. To put that into per- the laboratory was developed for coal. It is especially important to
spective, if a sample had 5 m3/t of gas (at a stated coal quality), the remember that in order to understand reservoir character, it is abso-
error factor on that measurement alone would mean the value could be lutely essential that both the amount of gas present and the amount
within the range of 4.25 to 5.75 m3/t. In a different approach Mares et the reservoir is capable of holding are estimated. Currently these
al. (2009a), working on a 5 m coal seam with 10 samples, determined two values are estimated through desorption and adsorption testing.
that in order to gain a relatively high level of accuracy (in approximating The outcome of those tests allows the percent gas saturation to be
the mean) at least four samples were required. More of these sorts of calculated (see Section 5.5).
studies should be conducted to determine the natural variance on differ- An adsorption isotherm test measures the maximum gas holding
ent types of coal beds. capacity of a sample of coal at any particular pressure at a stable tem-
Finally, just a word of caution about using desorption results for gas perature. Selection of the temperature is usually done based upon an
in-place analysis. Single, or even a few, desorption data points are noto- estimate of what the reservoir temperature is from the depth where the
riously susceptible to measurement error having a wide range of varia- sample was taken. One study indicates that the results from adsorption
tion within a single seam. This variation may be a result of organic tests are particularly sensitive to the temperature at which they
composition alone (again, see Mares et al., 2009a for an example) or are conducted (e.g. Bustin and Bustin, 2008) whilst another study
slight changes in coal quality. The way the desorbed gas was measured (Crosdale et al., 2008) indicates that temperature can vary widely before
and the reporting basis also should be noted as inconsistencies can there is any significant difference in adsorption capacity. Obviously, this
result in an inability to compare between data sets. Desorption data is is another area needing additional investigation. Another important
often used to estimate gas in-place (GIP): criterion is that the adsorption test is conducted at equilibrium moisture,
which, in theory, approximates the moisture levels found within the
GIP ¼ A  CT  d  GC ð1Þ reservoir. In the study conducted by Crosdale et al. (2008) it was found
that coal samples which had dried out—even at a small amount—gave
where,
higher maximum gas holding capacity results at most pressures than
samples that had not dried out. This stresses the importance of analysing
GIP gas in-place only fresh samples that have not lost even small amounts of moisture.
A area or distribution of coal being estimated There have been numerous studies that outline the various methods
CT coal thickness (e.g. gravimetric, volumetric) and procedures for adsorption testing
d density (e.g. Busch and Gensterblum, 2011; Kim, 1977; Krooss et al., 2002;
GC gas content. Levine et al., 1993; Yee et al., 1993) and thus the reader is referred to
48 T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81

those papers and references within. Although there are no standardised analysis—if these values are out of line, then both results should be
methods (e.g. ASTM, Standards of Australia, International Standards queried. Ideally, vitrinite reflectance is also measured on each adsorp-
Organisation) as yet, the adsorption testing community has conducted tion isotherm sample. Finally, ash yield (an estimate of inorganic
repeatability tests (Crosdale et al., 2005; Mavor et al., 2004). It is worth- material in the coal) can indicate a reason for lower (or higher)
while nonetheless to explain—however briefly—the fundamentals of than expected gas holding capacity in one sample relative to others.
interpreting the results from an adsorption isotherm test. The adsorption test results are usually tabulated and then
Fig. 19 shows an example of a typical adsorption isotherm result graphed. In the graph in Fig. 19 the x-axis defines pressure (in psi,
sheet. Although the eye is immediately drawn to the graph, probably kPa or MPa) and the y-axis the gas holding capacity (in standard
the most important information to register is the coal quality data. cubic feet per ton [scf/t] or in cubic meters per ton [m 3/t]). It is gen-
The graph cannot be properly assessed unless the rank (e.g. volatile erally accepted that in coal the gas holding capacity as related to pres-
matter), moisture and ash yield are known. As will be discussed sure follows what is known as a Langmuir curve, although there are
(Section 4.1), rank has a profound influence on the holding capacity other models that predict holding capacity (Langmuir, 1918; and
of the sample (higher rank, in general, equating to higher gas holding see Harpalani et al., 2006 and references therein for a fuller explana-
capacity). The moisture value is a good check to see if indicative rank tion). The measured data points, which are unique to each sample, are
is commensurate with the rank as indicated from the volatile matter then used to derive the Langmuir volume and pressure which when

Fig. 19. Example of a common adsorption isotherm report sheet. See discussion in text and note the volatile matter and ash parameters as well as the Langmuir volume and pressure
used to calculate the curve from the empirically derived data points.
T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81 49

placed within the Langmuir equation define the shape of the curve. For
example, the data points shown in Fig. 19 have a Langmuir volume and
pressure of 7.36 m 3/t and 8.7 MPa and using the Langmuir equation:

V ¼ ðVL  pÞ=ðp þ PL Þ ð2Þ

where,

V maximum gas adsorption capacity at any particular


pressure
VL Langmuir volume (m 3/t, scf/t, etc.)
PL Langmuir pressure (kPa, MPa, psi, etc.)
p pressure (kPa, MPa, psi, etc.) Fig. 20. Example of two types of adsorption isotherm curves, one high rank (from the
Pocahontas coal seam in the Appalachian basin) and a low rank coal from the Powder
River basin. Note that the high rank coal has an initial steep slope and then flattens
allows the curve to be generated not just through the data points but whereas the low rank coal tends to have a more steady slope.
extrapolated to higher pressures with a fair amount of confidence.
Adsorption results may be reported on an ‘as-received’ (ar) or ‘dry
ash free’ (daf) basis or both. In all cases the dry ash free basis results Although not as hard to estimate as desorption, the repeatability of
will indicate a higher gas holding capacity than the as-received results adsorption tests is still problematic to determine. There are many
for any given pressure because moisture has been mathematically possible sources of measurement error including handling procedures,
removed thus theoretically allowing more sites for methane adsorp- incorrect moisture levels, methods used to compute gas density, as
tion. Inorganic material will also be ‘removed’ in the daf calculation, well as poor temperature and pressure calibration (Busch and
which, in general, would also increase the gas holding capacity at Gensterblum, 2011; Mavor et al., 2004). Although some inter-
any particular pressure. So, which to use and when? Viewing adsorp- laboratory experiments have been conducted (e.g. Crosdale et al.,
tion isotherms on a dry ash free basis remains one of the best ways to 2005) it is still hard to isolate what the inherent variability of an adsorp-
understand rank and organic compositional controls on gas holding tion isotherm test is. This may be the result that there are no standards
capacity. In contrast, when trying to understand reservoir properties for adsorption determination of methane in coal and each laboratory
or estimating gas‐in-place, adsorption capacity should always be does things a little differently. It is also not as simple as running repeat
viewed or calculated to an as-received basis. It is unrealistic to assign tests on the same sample—quite possibly the microstructure of coal will
a dry ash free basis adsorption value to a reservoir as it will be neither change with each successive test (see the concerns of Crosdale et al.,
‘dry’ nor ‘ash’ free. It would be especially unwise to use adsorption 2005). Of the trials that have been conducted, the general feeling is
values on a dry ash free basis for gas‐in-place estimates (or within that the uncertainty of adsorption isotherm tests ranges from 7 to 20%
flow models) as this could seriously inflate the apparent gas resource (Crosdale et al., 2005; Mavor et al., 2004). These uncertainties are im-
values, especially in low-rank coals where moisture values are very portant to remember when making statements about the reservoir.
high. The last few sentences should be reread a few times by anyone who
Interpreting an adsorption curve in regard to a sample's maximum uses these data to make commercial decisions.
gas holding capacity is not difficult. If the reservoir is under normal
pressure, then a constant pressure gradient is assumed (0.98 MPa/m). 4. Geological framework
If estimating gas content from the pressure gradient, it is important to
remember that it is not the depth from the surface to the target coal Often taking the time to understand the geological framework of a
seam, but the depth from the top of the water table (which can be 10s CBM reservoir is undervalued and thus neglected. Characterisation of
of meters below the land surface) that should be used. The true reser- geological influences on a CBM reservoir may be subtle, but are no
voir pressure can be determined from direct permeability measure- less important or worthwhile than understanding the vagaries of
ments of the reservoir through drill stem tests (DST), injection fall off the market the gas may be sold into.
tests (IFO) or draw-down/recovery tests (DDR) (for a brief explanation Think of it like listening to classical music. If you can recognise
of permeability testing see Section 5.4.2). Beethoven from Haydn within a few bars, then the subtleties of
As an example lets determine maximum gas holding capacity for a where the music is going can be predicted. Similarly, it is like knowing
sample which was collected from a depth of 410 m. The resulting that, in general, fluvial paleoenvironments can produce dissected and
adsorption curve is given in Fig. 19. If the hydrostatic head is 400 m discontinuous coal reservoirs as opposed to the generally more
then the pressure can reasonably be assumed to be 3.92 MPa (4×
0.98 MPa) (or 3920 kPa) at that depth. Drawing a vertical line perpen-
dicular from the x-axis at 3.92 MPa until it intersects with the adsorption
curve will give, in theory, a maximum gas holding capacity of 6 m3/t. As
will be discussed further (Sections 5 and 6.2), there are many influences
that could affect the results of adsorption analyses and make it unwise
to rely on just one adsorption isotherm to estimate maximum holding
capacity (see also Mares et al., 2009a).
Different samples can have quite distinct adsorption curve shapes.
A ‘classic’ shape (usually associated with higher rank coals) is a curve
that has an initially steep slope but levels out quickly and remains
relatively flat even at high pressures; lower rank coals tend to have a
flatter curve throughout all pressures (Fig. 20); though it should be
noted that there are exceptions to the rule at all ranks. For example,
within a single seam, and using samples of similar composition, the Fig. 21. Nine adsorption isotherm test results from a single 5 m coal seam. Note the
shape of the adsorption curve can vary greatly (Fig. 21) and this can differences in the shape of the curves; ash yield and organic composition are all roughly
have ramifications on inferred producibility (see Section 5.5). the same. Modified from Mares et al. (2009a).
50 T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81

continuous nature of coastal plain-derived coals (for more details on scientists. It is generally defined as the level of organic metamorphism
the shape of coal bodies, see Ferm et al., 1979). These kinds of ‘rules’ or the level of coalification. In simple terms, it is the degree to which
or ‘properties’ allow one to have a better understanding of both the organic matter has been ‘cooked’, one way or another, and transformed
behaviour of music and CBM reservoirs making prediction of trends, in both physical and chemical terms from one state to the next in an
in either field, easier. irreversible process.
To help in understanding how the geology affects both the coal There are many parameters used to estimate rank. The coal geolo-
and gas there are some useful conceptual guides to call upon. gist may refer to organic material as being lignite or subbituminous or
Warwick (2005) outlines a system that identifies four relatively linear bituminous whilst the coal petrographer will use vitrinite reflectance;
stages starting from peat accumulation through preservation-burial, the chemist may use carbon content whilst the coal quality technician
diagenesis-coalification and finally the coal and hydrocarbon forma- may rely on moisture or volatile matter. All of these measures estimate
tion phase. Although complex, there are some fundamental tenets rank and can roughly be calibrated to each other (Fig. 22) but not all
that can be relied upon within this system—for example, that matura- are equally adept at characterising rank changes across the whole
tion only goes forward, not backwards, that volatiles are driven off, coalification range (Fig. 23). By far the most accepted single parameter
not taken on. used to determine rank is vitrinite reflectance, though even this mea-
sure can be influenced into giving falsely low or high readings because
4.1. General rank effects of original and secondary processes acting on the coal (Newman and
Newman, 1982).
Coal rank is arguably the most important factor influencing methane The processes of coalification are extremely complex and in almost
generation in coal beds. It is important to note that coal quality is a com- all cases boundaries between states, or stages, of coalification cannot
bination of rank, grade and type. Rank is explained in more detail below be precisely defined. This is a result of both the extreme heterogeneity
but grade and type are also essential to take into account in order to of coal (Mukhopadhyay and Hatcher, 1993; Taylor et al., 1998) as well
understand the character of a coal reservoir. Coal grade usually refers as the myriad of influences that organic matter can be subjected to
to its level of purity, that is, what are the relative amounts of organic during burial. For a detailed consideration of coalification, the reader
and inorganic materials present within a particular coal. Different coal is directed to summaries by Levine (1993) and Taylor et al. (1998) as
types are distinguished based on the composition of the organic mate- well as references therein. Because of its importance to CBM, a brief
rial, which is ultimately related to the original vegetation type, which description is given below drawing upon the above references as well
formed the original peat. The effects on the coal reservoir by both coal as a few others.
grade and type are further discussed in Section 6. Coal is composed of plant remains, inorganic matter and water.
It has already been presaged in Section 3 that low-rank coal will However these components can be broken down even further, to the
have only biogenic gas whereas higher rank coals could have either or basic elements of C, H, O and N, plus the odd impurity (namely, the
both thermogenic or biogenic gas (see Section 5.5). “Rank”, however, inorganic matter, which forms—by definition—a minority proportion
can be hard to define and mean different things to different types of within the coal). Although the base components are few, they generate

Fig. 22. Correlation of different rank parameters, all approximate.


T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81 51

Fig. 23. Comparison of different measurement parameters that indicate rank. Note that vitrinite reflectance is the best estimator of rank over the whole rank range.

an extraordinary range of variation when combined. This is perhaps not 5. Graphitization


surprising considering the vast array of plant material (which changes a. Organic material composed of almost all carbon (Bustin et al.,
through time) that goes into making a peat bog, and eventually a coal. 1995; Schopf, 1966),
Organic material goes from being immature to mature when its b. Capacity to hold gas may continue to increase,
hydrocarbons are mobilised by increased temperature and pressure; c. Methane production can continue into the anthracite rank
i.e. the burial process starts to change things chemically, and thus range, but by the time organic matter is graphitised methane is
physically. However, chemical maturation is not the start of physical virtually zero and any remnants of gas have long been driven off,
and chemical changes in coal. From the start of peat accumulation d. Moisture increases slightly as a result of an increase in porosity
organic material is progressively modified on its journey from a mixture (Taylor et al., 1998),
of plants and water into the solid material called coal (see Lüttig, 1986 e. Aromatic lamellae re-orient from the semi-anthracite to meta-
for a review of how plants are transformed into peat and beyond). anthracite rank stages.
The geochemical–biochemical transformation of plant substances
can be described as occurring over five steps, or coalification stages As the rank of a coal increases, maximum gas holding capacity also
(Levine, 1993; and Fig. 10): increases. This very general association is related to both a loss in
moisture with increasing rank as well as an increase in porosity. How-
1. Peatification
ever, the relationship between rank and gas properties is neither
a. Plant materials in peat are humified and macerated, straightforward nor universal. For example, there is no doubt that
b. Compaction of material begins (see Clymo, 1984, 1987; Dehmer, rank is a primary influence on the maximum gas holding capacity of
1995; Moore and Hilbert, 1992), coal. It is generally thought that all other variables being equal (e.g.
c. Primary biogenic methane develops whilst the mire is still accu- ash yield, organic composition) the more mature a coal, the more gas
mulating peat (Hornibrook et al., 1997; Wang et al., 1996). it can hold (Hildenbrand et al., 2006; Kim, 1977; Fig. 24a). However, it
2. Dehydration is not just a simple case of rank ‘A’ will correspond to gas holding capac-
a. Significant loss of water with orders of magnitude loss of porosity ity ‘B’ anywhere in the world. Within basins, the above trend may hold,
(Moore and Hilbert, 1992; Shearer and Moore, 1996), but there is little correlation between basins (Bustin and Clarkson, 1998)
b. Organic material enters lignite through subbituminous rank (because of differences in coal type, grade, and botanically inherited
range, holding capacity of methane increases substantially, microstructure). For example, gas holding capacity increases with
allowing storage of secondary biogenic gas in the coal, increasing rank as shown in Fig. 25 which shows samples from a single
c. Continued development of biogenic gas, if conditions are right basin, from coal seams of similar organic composition. (Note that there
(see Section 3.1). is a reversal in holding capacity for the two highest rank samples,
3. Bituminisation because of very low ash yield content of the sample from 682 m; see
a. Mobilisation of hydrocarbons (enters the ‘oil window’) (Saxby Section 6 for more discussion.) When different coals from different
and Shibaoka, 1986; Wilkins and George, 2002), basins are plotted together, it is clear that rank and maximum gas hold-
b. Initiation of generation of thermogenic methane, occurring any ing capacity have little universal correlation (see Bustin and Clarkson,
time and throughout the subbituminous A to high volatile A 1998; Fig. 24b).
bituminous stages (Hunt, 1991), Moisture content is very sensitive to increasing rank at the early
c. Maximum gas holding capacity increases substantially, though, stages of coalification (Fig. 23). In most cases, moisture content decreases
mobilisation of bitumens may also clog pores, decreasing the with depth (Sivek et al., 2010). With this in mind, the predominant
capacity. reason for higher gas holding capacity is considered to be a result of
4. Debituminisation less moisture competing for methane adsorption sites at higher ranks
a. Maximum generation of thermogenic methane and other oil (Bustin and Clarkson, 1998; Crosdale et al., 2008; Joubert et al., 1974;
substances, Ozdemir and Schroeder, 2009). Thus in a lignite, although it has abun-
b. Exits the oil window, dant porosity and therefore sites for methane adsorption, any gas pro-
c. Moisture values at their nadir. duced (biogenically) would have few places to adsorb because of high
52 T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81

Fig. 24. Adsorption isotherms related to rank. (A) The oft cited isotherm graph from Kim (1977) relating coal rank to maximum gas holding capacity (at 0 °C; red numbers are
approximate mean–maximum vitrinite reflectance). Although rank–gas relationships shown are in general correct, they are far from universally observed (see text). (B) Adsorption
isotherms from different coal ranks from different basins around the world (Goodman et al., 2004; Hildenbrand et al., 2006; Laxminarayana and Crosdale, 1999). Mean–maximum
vitrinite reflectance is noted to the right of each isotherm. Note that there is no systematic trend of higher rank with higher gas holding capacity.

moisture (generally >30%). Even slight changes in moisture content will down hole (from 0.53% at 327 m to 0.69% at 799 m) and although
significantly affect gas holding capacity (Fig. 26). the maximum gas holding capacity will increase with hydrostatic
Desorbed gas contents can also be correlated with rank. Fig. 27 pressure, rank certainly plays an important role in the higher gas
illustrates rank and measured gas down hole in three drill holes; all charge. The same relationship can be seen in other basins (e.g. Scott
wells are from the same formation but drill hole ‘A’ has been affected et al., 2007; Fig. 28).
by a localised shallow-seated heat source (Herudiyanto, 2006; Nas, The type of gas that occurs in a coal bed is also rank dependent—in
1994) whereas drill holes ‘B’ and ‘C’ have been less affected. Drill under mature coals, methane is generated only biogenically because
holes ‘B’ and ‘C’ have little rank change down hole and the increase the heating (or ‘cooking’) process (bituminisation) has not been
in desorbed gas is probably related to the increased capacity of the sufficient to begin a chemical change and thus, mobilisation of the
coal to hold gas from the increased hydrostatic pressure. However, organic constituents. Although isotopic evidence shows that biogenic
it is equally clear that in drill hole ‘A’ there is an increase in rank gas can recharge high-rank coals after uplift (Ayers, 2002; Flores et al.,

Fig. 25. Adsorption isotherms from a drill hole demonstrating relationship of rank with maximum gas holding capacity. Depths are in meters below surface and mean maximum
vitrinite reflectance is within parentheses. Ash yield (in red) given in percent (air dried basis). All seams have vitrinite concentrations between 84 and 94%. Adsorption isotherms
conducted at 35 °C and reported on an as received basis.
T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81 53

Fig. 26. Relationship between maximum gas holding capacity for methane and moisture
content. Modified from Levy et al. (1997).

2008), it is the thermogenic gas which is chemically produced from


major physico-chemical changes that occur during the bituminisation
and de-bituminisation stages of coalification (see Section 3.2).
Gas composition or ‘quality’ is also affected by coalification and is
closely tied to the origin of the gas (biogenic/thermogenic). In many, if
not most, cases of low-rank coal the gas is predominantly CH4 (see
Section 3.1 for discussion on how CH4 is derived biogenically). However,
although high-rank coal gas may also be predominantly CH4, often there
are other higher carbon chains such as C2H4, C2H6, C3H8, C4H10 (i.e. eth-
ene, ethane, propane and butane). This is the result of coalification
induced geochemical changes and evolution of gases from the coal, as
well as lipid vs humic sources. There is also a higher probability that
Fig. 28. Relationship between depth and gas content in an Australian coal basin. Note
high-rank coals may contain greater concentrations of inert gases, that there is a rough trend but much scatter most likely related to differences in gas
control from local geological conditions and coal composition. Modified from Scott et
al. (2007).

especially CO2 (e.g. the Sydney basin; Faiz et al., 2007b). The reasons
for this are not well understood, but it may be that heat sources (e.g. in-
trusives, hydrothermal fluids, shallow seated magma, etc.) which have
increased the rank of the coal and most likely its gas content too, have
also contributed volcanically-derived CO2. Therefore CO2 present in
high-rank coals may not be derived from the coal but could originate
from the association with a heat source, which could also be the driver
for increased rank.
Finally, permeability is affected in interesting and complicated
ways as coal increases in rank. Most of all, permeability to water
and gas is stress dependent, but at a given stress variability can be
related to the fracture network in the coals. All other factors being the
same (e.g. organic composition, % ash yield etc.), as rank increases the
fractures in coal (also termed cleats) related permeability tends to
increase (Law, 1993; Wang, 2006). This relationship may not be as
straightforward as it first appears (Laubach et al., 1998; Su et al., 2001)
because of the interplay between fracture and matrix permeability
(described in Section 5.4). In contrast to permeability derived from frac-
tures, matrix permeability is mostly reported to decrease with increas-
ing rank (Balan and Gumrah, 2009; Wang and Ward, 2009) and this is
seen primarily as a function of decreasing pore sizes (Durucan et al.,
2009; Rodrigues and Lemos de Sousa, 2002; see also Section 5.1).
Increase in fracture permeability with rank may not be a completely
causal relationship; with increased rank usually comes increased expo-
sure to tectonic influences that tend to produce higher cleat frequency.
Lower matrix permeability with increased rank may also be derived
Fig. 27. Down hole gas contents from three drill holes, all within 5 km of each other. Note
from mobilisation of volatiles which then fill pores, as opposed to actual
that drill hole ‘A’ has been rank affected, vitrinite reflectance values are around 0.5% at the
top and 0.69% at the bottom of the well. Vitrinite reflectance values in the other two drill loss of the pore structure (Schneider, 1986; Taylor et al., 1998;
holes range from 0.33 to 0.44% with only minor rank increase down hole. Teichmüller, 1989).
54 T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81

4.2. Impact of original depositional environment

The depositional setting in which peat forms will naturally affect its
spatial geometry—i.e. seam thickness and splitting character (Ayers,
2002; Cameron et al., 1990; Cohen, 1984; Esterle and Ferm, 1994;
Ferm and Staub, 1984; Ferm et al., 1979; Fielding, 1984; Flores, 1993;
Flores and Hanley, 1984; Roehler, 1988). Additionally, vegetation type
plays an integral part in determining differences in composition both
between coal beds and within coal beds (facies variation). Domed peat
mires in SE Asia are an example of how this works. It has been well
documented that the type of surface vegetation in mires is correlated
with proximity to streams—big trees at the edge (where nutrients
from flood events can reach) to lower stature vegetation in the centre
(where vegetation tends to be more nutrient starved; Anderson, 1964,
1983). These differences are translated into different peat types with
lateral and vertical variability (Cameron et al., 1990; Esterle and Ferm,
1994; Esterle et al., 1992; Moore and Hilbert, 1992; Moore et al., 1996,
2000). There are other models for peat (e.g. Diessel, 1982; Teichmüller,
1989) but a common feature in all models is vertical and lateral variation
in vegetation and consequent peat types (and finally, with burial, coal
lithotypes). At a given rank the coal lithotypes and their grade will influ-
ence gas generation and holding capacity, coal strength, and fracture
permeability networks, which are all important reservoir parameters.
However, to make things even more complicated, vegetation type
has a fourth dimension: time. As plants have evolved, their chemical
makeup has changed, thus different ages of coal are unique in their
properties and this translates into unique properties of coal (see
Collinson and Scott, 1987; DiMichele et al., 1987; Greb et al., 2006; Fig. 29. Minimum effective stress influence on (A) permeability and (B) gas and water
Shearer et al., 1995). As will be discussed in Section 6, differences in production in the Cedar Cove area, Black Warrior Basin, Alabama. Modified from Bell
organic composition can play an integral part in determining gas prop- (2006) and Sparks et al. (1995).

erties within and between seams.


In general, conventional oil and gas structural exploration models
are not always appropriate for CBM. For example, targeting structural
4.3. Role of structure and stress highs, such as anticlines, has had little success within the context of
producing gas from a coal reservoir. In the Powder River Basin, however,
The response of coal beds to stress can also best be described as com- it has been found that in some cases, sandstones capping anticlines can
plex. Perhaps one reason for this is that coal is compressible, relative to act as reservoirs and these features can be developed as gas plays in
conventional sandstone reservoirs (Ayers, 2002), and controls on that their own right (Ayers, 2002). In addition, the Peat and Scotia CBM fields
compressibility are numerous and interrelated, including rank, organic in Australia also behave more like a conventional play in that higher gas
composition and mineral matter content. contents, and even ‘free’ gas have been reported at the top of anticlines
The first and most important type of stress that is encountered in (Draper and Boreham, 2006). In other cases, structural highs are found
CBM is that of lithostatic pressure; that is, the deeper you go the more to have low gas saturation, possibly because the tensional nature of
pressure there is from the overlying strata. It has been well documented the strata has allowed gas to escape through fractures. In contrast,
that permeability decreases drastically with depth (increased stress) the synclines may have good gas saturation but low permeability
most likely because the fractures are closed (Enever et al., 1999; resulting in low gas flow for wells. In these types of plays, the limbs of
Somerton et al., 1975). By way of a scale, some coals can have a perme- the synclines/anticlines may actually be the best methane producers
ability of hundreds of millidarcies (mD) at, say, 100 m depth but by and this may be the result of a combination of gas saturation, permeabil-
1000 m the permeability may be as low as 1 or 2 mD (more on that in ity and reservoir compartmentalisation (Kinnon and Esterle, 2008;
Section 5.4). Kinnon et al., 2010).
The horizontal stress regime that basins or sub-basins are subjected Finally, present and historic stress of a basin will result in faults.
to can also have major ramifications for CBM development. Horizontal Regardless of fault type (i.e. normal, reverse etc.) the net effect will
pressure tends to close up fractures. Therefore basins that have low be compartmentalisation of the coal reservoir. This may be good, or
principal horizontal stress (PHS) are more amenable to artificial this may be bad. On the one hand, closely spaced faults may reduce
(induced) fracturing of the reservoir. the effective reserve area of a well or wells, but on the other hand,
The effective stress in a basin has also been shown to affect perme- the appropriate spacing of faults can limit de-watering required and
ability (Bell, 2006; Sparks et al., 1995). Effective stress is the difference allow quick gas to surface, possibly enhancing the economics of a
between reservoir or seam pressure and the opening pressure prospect. Analysis of fault frequency must always take into account
measured in the borehole or well (Esterle et al., 2006). As the minimum their possible fractal nature and therefore, whatever the data density,
effective stress in the basin increases, permeability can decrease from a proportion will no doubt be undetectable.
nearly 10 mD at 1 MPa to less than 1 mD at 8 MPa (Sparks et al.,
1995) (Fig. 29a). The decreasing permeability will affect both gas and 4.4. Influence of external effects
water flow rates (Fig. 29b). This effect should not be underestimated
when considering basinal controls on permeability. However, Bustin External effects on CBM prospects are potentially a huge topic and
(1997) did find that vitrinite content (and especially vitrinite occurring therefore cannot be covered in a review paper such as this. Discussion
as bands) in some Australian coal seams had a profound effect on per- herein will be confined to how volcanic intrusives influence a coal,
meability, sometimes greater than effective stress. and thus CBM potential.
T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81 55

In most cases direct cross cutting of igneous intrusions through (though there was an increase again at low volatile bituminous rank)
coal seams will have minimal regional effect on coal as it has been and micropores increased with rank (Fig. 30). Other studies however
shown that rank change is usually spatially restricted in these circum- have not necessarily confirmed this relationship (Bustin and Clarkson,
stances (e.g. Cooper et al., 2007). Murchison and Raymond (1989) 1998; Radovic et al., 1997). In a Turkish coal bed, Gürdal and Yalçin
showed that intrusions in Scotland affected the rank of the coals to a (2001) found that pore surface area increased along with rank up
larger regional extent when the timing of those intrusions were into until about 1.0% vitrinite reflectance but then started to decrease
seams that were not fully mature. again with further rank increase. Ozdemir and Schroeder (2009) also
Shallow igneous intrusives are a different story when it comes to found that pore sizes in general decreased with carbon content (i.e. in-
affecting regional rank trends. For example, the Sangatta coal deposit creased rank). Mastalerz et al. (2009), in contrast, did report an increase
of East Kalimantan, Indonesia has been affected by a well-recognised in pore volume with increasing rank, but this was a rank increase in-
shallow heat source (Fukasawa et al., 1987; Herudiyanto, 2006). The duced from an igneous dike. Thus, in this case, the pore volume is likely
resultant rank increase from subbituminous to bituminous is responsi- to be more related to coke structures rather than ‘normal’ coalification
ble for the level of current mining in the area (Nas, 1994) as well as CBM changes.
exploration and development (Dart Energy Ltd., 2011)—coals of similar Pore volume is particularly difficult to measure in the laboratory
age in East Kalimantan are lower in rank and have lower gas content. and the resulting measurements are notorious for their inaccuracies.
The Bukit Asam deposit in Sumatra has been similarly affected from a Mercury impregnation has classically been performed but, because
relatively shallow heat source (Amijaya and Littke, 2006; Susilawati of lack of repeatability in coal, is rarely used now. Instead, estimates
and Ward, 2006). Gurba and Weber (2001) have also reported in- of pore volume are conducted using small angle scatter techniques
creased gas contents in the Gunnedah Basin, Australia in association (e.g. SAX or SANS; Radlinski et al., 2004), which give more quantitative
with igneous intrusive sills. assessments. Typical porosity values used for gas flow modelling usually
Shallow volcanic intrusives can also affect gas type. For example, range from 1 to 5% (Close, 1993; Zarrouk, 2008; Zarrouk and Moore,
the Sydney Basin in Australia is thought to have increased CO2 content 2009) although some SAXS and SANS results indicate porosity between
in the gas because of a shallow heat source (Faiz et al., 2007b). In a 16 and 25% (Mares et al., 2009b). This discrepancy serves to emphasise
different mechanism, large pockets of CO2 in the Yaojie coalfield that the understanding of the nature of pores in coal still has a way to go.
(China) have been attributed to large scale basinal faulting that has
allowed gases from underlying carbonates to migrate into the coal (Li 5.2. Sorption and diffusion
et al., 2011).
Sorption is the term that describes the physico-chemical nature of
5. Reservoir characterisation how methane is stored within coal. In general there are three states
(Ceglarsk-Stefańska and Brzóska, 1998; Ceglarsk-Stefańska and Zarębska,
5.1. Pores 2002; Rice, 1993):

1. Free state—where methane molecules are within fractures and


A fundamental property which makes coal such a special and
macropores
unique material—and has implications for coal properties ranging
2. Adsorbed state—where methane molecules have a physical
from not only methane holding capacity but also activated carbon
adsorption within micropore complexes
and liquefaction (Radovic et al., 1997)—is it's porosity. The pores are
3. Absorbed state—where the methane is ‘dissolved’ within the
so plentiful that just 1 cm3 of coal can contain pores with an internal
molecular structure of the coal.
surface area of 3 m 2 (Mares et al., 2009b; Radlinski et al., 2004; another
study reported surface areas as much as 115 m 2/g of coal; Şenel et al., Methane in a free state is self explanatory, i.e. gas is present in
2001), the size of a very large dining table. Now envision a 5 m thick open voids unattached to coal material. Adsorption happens when a
coal seam over an area of a square kilometre, not an unusually large molecule of methane attaches itself in a loose physico-chemical
seam. This seam contains around 5 million m 3 of coal and could have state to the surface area of a pore. In contrast, absorption is when a
internal pore surface areas of more than 1500 km2, about the size of single molecule of CH4 is chemically held within the molecular struc-
Rhode Island. With this in mind it is easier to understand how coal ture of coal. Milewska-Duda et al. (2000) describe a multiple sorption
can hold very large quantities of gas, adsorbed onto its very large pore model whereby a molecule of CH4 is both absorbed into and adsorbed
surface area. on to the surface area of submicropores. Others have speculated that
It is generally accepted that there are actually two key parts to the molecule itself, because of the elastic nature of coal, can modify
coal's porosity system, large fractures together with the much smaller
pores (Clarkson and Bustin, 1999; Cui et al., 2004; Harpalani and
Chen, 1997; Rodrigues and Lemos de Sousa, 2002). Within the smaller
pore system it is recognised that three size classes of pores may exist
—macro, meso and micropores (Şenel et al., 2001). Macropores have
been defined as being > 50 nm in diameter whereas mesopores are
between 2 and 50 nm and micropores b2 nm (Gan et al., 1972).
Interestingly, Yao et al. (2008) report that pores follow a fractal distri-
bution being primarily influenced by the presence of natural surface
area irregularities such as original plant pores and mineral matter. It
has been argued that the size distribution of pores is even more com-
plex than a ‘simple’ bi- or trimodal distribution with porosity being
influenced by a range of types and sizes of micro-fractures (Gamson
et al., 1996; Li et al., 2003). Understanding the “dual porosity” system
is important because it is core to understanding gas desorption, max-
imum gas holding capacity and flow.
Rank influences both pore volume and pore frequency (Gan et al.,
1972). Levine (1996) plotted macropores and micropores in relation Fig. 30. Relationship between % total pore volume between macropores and micropores
to carbon content and found that in general macropores decreased versus rank. Modified from Levine (1996).
56 T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81

the structure of the coal (Milewska-Duda et al., 2000; Ozdemir et al., of travel. Thus, for example, in high-rank coals that have abundant
2003; St. George and Barakat, 2001). Wang et al. (2010b) state that fractures, diffusion rates are higher as there is less distance for the
methane will significantly change the volume of both micro- and gas to travel.
macropores. Gases diffuse at different rates; CH4 tends to diffuse more slowly
Most methane in coal, however, is thought to be adsorbed, as op- through coal than, CO2, for example, because of differences in water
posed to free or absorbed methane (Milewska-Duda et al., 2000). solubility of the gas (Busch and Gensterblum, 2011; Cui et al.,
This means that pore surface area rather than pore volume is the 2004). Although it has been hypothesized that diffusion should most-
most crucial factor in terms of gas holding capacity (Moore, 2012). ly occur perpendicular to the bedding within a coal seam (Cui et al.,
This is a fundamental difference between how gas is held in conven- 2004), the physics of diffusion dictate that CH4 will move, through
tional versus coal reservoirs. Brownian motion (Philibert, 2005), from areas with high concentra-
It has been reported that only micropores will play a significant tions of methane to areas with low concentrations. When permeabil-
role in methane adsorption; macro- and mesopores most likely act ity or the hydraulic head gradient is low, then diffusion will be the
primarily as transport conduits (Ceglarsk-Stefańska and Brzóska, primary method for methane migration. Some studies have shown,
1998; Ceglarsk-Stefańska and Zarębska, 2002; Laubach et al., 1998; through modelling, that diffusion rates will change with time as a res-
St. George and Barakat, 2001). Although different theories exist as ervoir is being produced (Pillalamarry et al., 2011).
to how exactly methane is adsorbed onto the surface area of the
pores, it is likely that they occur as single layers, although double 5.3. Matrix shrinkage and swelling
layers of molecules have also been hypothesized (Zhou and Zhou,
2009). The adsorption capacity of pores is also dependent on gas During CBM production (i.e. during the de-watering stage), cleats
type—in most cases because of its molecular make up (that is, shape loose water buoyancy and close up, resulting in lower permeability
of the molecule), CO2 is sorbed preferentially over CH4 (Harpalani et than in the reservoir's initial state, which will have the net effect of
al., 2006). A well recognised retardant to gas adsorption within reducing gas flow. The effect that matrix shrinkage has on the coal
pores is the presence of water (Hildenbrand et al., 2006; Ozdemir is to re-open up the cleat system allowing greater permeability
and Schroeder, 2009; see also Section 4.1) which is one of the main (Gray, 1987). Thus, understanding the mechanism of coal matrix
reasons why high-moisture, low-rank coals contain only small con- shrinkage and swelling is fundamental to understanding permeability
centrations of methane. changes in a coal reservoir, and therefore will be discussed separately
Whilst sorption character informs us about how methane is held from, and before, permeability; though there will, of course, be some
within coal, diffusion theory tells us how it moves from one place to overlap in the discussion. It should also be stated that, although the
another (from one pore to the next—until the methane molecule phenomena of shrinkage and swelling in coal have been the subject
reaches a cleat ‘superhighway’). It is Fick's Law that describes how of many recent papers (e.g. Balan and Gumrah, 2009; Mazumder et
gas diffusion takes place between the pores (Fick, 1855). Fick's Law al., 2012; Pan and Connell, 2011a; Pan and Connell, 2012; St. George
also can apply to how fluids diffuse through water, though in this and Barakat, 2001; Wang et al., 2010b), it is still not well characterised
paper we deal only with gas diffusion. Initially Fick's Law was hard or understood.
to verify (Philibert, 2005) but with time was supported through ex- A coal reservoir is a dynamic system and it has been demonstrated
perimental data. The formula for Fick's Law is: that over the life of a well, shrinkage may cause significant changes
(such as increases in permeability; Gu and Chalaturnyk, 2006;
F ¼ D dC=dx ð3Þ Harpalani and Chen, 1995, 1997; Wang 2006; Wang and Ward,
2009). Matrix shrinkage occurs when gas is desorbed out of the coal
where, matrix; fundamentally it is related to the mechanical elastic proper-
ties of a coal (Durucan et al., 2009; Levine, 1996; Milewska-Duda et
F diffusion flux al., 2000). Remember that it is thought that the majority of the meth-
D diffusion coefficient ane is held in micropores (Section 5.1), and that these micropores can
dC/dx concentration gradient. be so small (~ 5.7 Å 3 with a maximum linear dimension of ~ 1.5 Å)
that the methane molecule adsorbed (and possibly absorbed) onto
The diffusion coefficient, D, is derived by experiment and is de- (and into) the surface may actually stretch and distorts the pore.
pendent on the properties of different gases (some have high D Thus, when gas is desorbed the micropores shrink, returning to
which means that they diffuse rapidly whilst others have low D their original shape (St. George and Barakat, 2001). Clearly the com-
which means that they diffuse more slowly). In CBM, D can be esti- bined effect of millions of micropores in a very small area of coal
mated from desorption time, which is related to the diffusion coeffi- shrinking will have incremental but profound effects. There has also
cient (Lama and Nguyen, 1987; Zarrouk, 2008). been some studies which indicate that loss of moisture in a reservoir
What Fick's Law boils down to are three basic concepts: Firstly, could result in matrix shrinkage as well (Ozdemir and Schroeder,
diffusion is directly proportional to the concentration gradient; in 2009; Ozdemir et al., 2003).
other words, the steeper the concentration gradient, the faster the If the shrinkage is great enough, then this could counteract any
rate of diffusion. One can think of it like a woman's perfume (or a decrease in permeability from the loss of pore pressure (from
man's cologne): The perfume may originally been sprayed in one dewatering) and cleat closure. The amount that a coal may shrink is
area, but quickly disperses to areas distant from the original source. probably related to many things including coal rank and organic com-
However, the distance reached from the original source is dependent position (Wang and Ward, 2009). Whether matrix shrinkage can
on the original concentration. If the concentration is high then the overcome cleat closure during dewatering might be a function of
perfume might meet you from several meters away, say from even gas saturation. Wang (2006) found that if the pressure where gas de-
around a street corner, whilst lower concentrations will require sorption takes place in a reservoir is relatively high (such as in high
more intimate proximity for detection. Secondly, diffusion is propor- gas saturated reservoirs) then cleat closure is relatively small before
tional to the amount of pore surface area—the more pores the less significant gas desorption occurs with concomitant matrix shrinkage.
solid the gas has to diffuse through and thus the faster the rate of dif- In this scenario, shrinkage would have the net effect of increasing the
fusion. Once again, it is the surface area of pores that has a major im- permeability. However, in reservoirs where desorption pressure is
pact on the behaviour of CBM. Thirdly, speed of diffusion is inversely relatively low (i.e. a lot of water has to be pumped from the coal to
proportional to distance—the greater the distance the slower the rate reach critical desorption pressure), compared to the original reservoir
T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81 57

pressure (as in under saturated coal seams) then cleat closure would other things, of the remnant inherent plant porosity and decreases
be so significant that by the time desorption and matrix shrinkage with rank.
started to occur, the reservoir would have already become too tight,
in permeability terms, for matrix shrinkage to counteract. 5.4.2. Fracture/cleat
Matrix swelling is not seen as an issue during normal CBM produc- Naturally occurring fractures in coal (i.e. cleats) are the single most
tion (i.e. straight desorption of gas off the reservoir). However, it may important physical attribute governing gas flow in a CBM reservoir.
be an important factor if theoretical gas enhancements are used, such The ‘how, where and why’ of cleat formation is still a matter that is
as CO2 flooding, to push out methane and speed up production. Thus a debated (see Close, 1993; Laubach et al., 1998 for extensive reviews of
brief review of matrix swelling will be given here. Carbon dioxide is cleating in coal). There are two main processes that are usually cited,
generally considered to be the most viable gas to use for enhanced however, for cleat formation: endogenetic or exogenetic (Close,
coalbed methane (ECBM); it is well known that CO2 has a much 1993). The former is the result of compaction and contraction of the
slower break out time than, say N2 (Zarrouk and Moore, 2009) and coal from desiccation and the coalification process, whilst the latter is
has a high adsorption value in coal compared to CH4 (Durucan et al., from tectonic stress and strain placed upon a seam as it is being buried.
2009; Mastalerz et al., 2004; Stanton et al., 2001). It is also considered Cleats in coal almost always occur as two mutually perpendicular
a method for reducing GHG through sequestration of CO2. sets of fractures (i.e. having an orthogonal geometry). The ‘face cleat’
It has long been recognised that coal can swell when injected with is the dominant fracture system whereas the ‘butt cleat’ is less laterally
certain gases (e.g. see Anderson et al., 1956). As already discussed, continuous and almost always terminates where it intersects a face
when methane is released in coal, the molecules that are stretching cleat (Fig. 31). Cleats are perpendicular (or nearly so) to bedding
the micropores, and thus the matrix in a cumulative effect, allow and face cleats will typically strike perpendicular to the axes of folds
the structure to relax and, for reasons described above, shrinkage and be parallel to normal fault systems (Cameron, 1995). There are
occurs. Thus, when injected with any number of gases (i.e. CO2, CH4, many instances where cleat directionality does not follow what
H2S, N2) these have been shown to have, not unsurprisingly, the might be inferred from the tectonics of the basin (e.g. see Laubach et
opposite effect of gas desorption, and causes the matrix of coal to al., 1998). In particular, directly adjacent to shear zones, the coal
swell (Bustin, 2004). The swelling doesn't occur isometrically in all cleat direction can be altered to reflect the localised stress–strain
directions, but interestingly seems to have the greatest swelling relationships.
perpendicular to bedding (Day et al., 2008; Levine, 1996; Pan and Because of the importance of cleats in CBM production (and min-
Connell, 2011a). In general lower rank coals may be more affected by ing), much work has been conducted on understanding the controls
swelling when injected with CO2 than higher rank equivalents on cleat frequency and distribution. Cleat frequency has generally
(Majewska et al., 2010). Balan and Gumrah (2009) specifically specu- been noted to decrease as a function of bed thickness (Harpalani
late, however, that dry coals, even if fully gas saturated, are most affected and Chen, 1997). In a detailed study on a small set of Bowen basin
by swelling whereas water and gas saturated seams are least affected. coals, Dawson and Esterle (2010) found four classes of cleats. This
Note that ‘dry’ coals are those CBM reservoirs which are not water satu- classification may be broadly applicable to other areas, at least in con-
rated and require virtually no de-watering to initiate gas flow. The cept (see also Anna, 2003). The four types of cleats found by Dawson
Horseshoe Canyon CBM play in Alberta, Canada is a notable example and Esterle (2010) were (Fig. 32):
of a ‘dry’ CBM play. The Powder River Basin (USA) would be considered
the opposite of a dry CBM play. Composition of the coal is also thought to 1. Master cleats
affect swelling properties; coal types dominated by vitrinite and liptinite 2. Single vitrain layer cleats
swelled more than inertinite and clay-rich coals coal types, evidently 3. Multiple vitrain layer cleats, and
because of higher rates of CO2 dissolution (Karacan, 2003). 4. Durain layer cleats.

5.4. Permeability Note that both ‘vitrain’ and ‘durain’ refer to macroscopically recog-
nizable coal types in coal seams. For an explanation of these see
Whereas porosity informs us about the storage capacity of gas in Stopes (1919) and Taylor et al. (1998). When banding is referred to
coal, permeability defines the level of transportability of that gas. in this paper, it refers to the degree (i.e. the proportion) to which
Even if a coal seam has high gas volume (i.e. ‘charge’) low permeability vitrain bands make up a coal type (i.e. lithotype).
can result in uneconomic gas production rates. Permeability in a coal is
affected by many things, including rank and basin wide stress (see
Sections 4.1 and 4.3). However it is vital to understand that there are
two broad types of permeability: matrix and fracture. How these two
types of permeability interact with each other determines, to a very
large degree, the gas production profile for a well (Levine, 1996;
Wang, 2006; Wang and Ward, 2009).

5.4.1. Matrix
Simply put, matrix permeability is the permeability of the coal
itself, not taking into account natural fractures that may crosscut and
subdivide the coal. The pore system within the coal matrix ultimately
determines matrix permeability (see Section 5.1 and Adeboye and
Bustin, in press). Estimating matrix permeability is performed in the
laboratory usually on core samples that have been collected specifically
for that purpose. Besides giving a better understanding of the coal reser-
voir from a scientific sense, matrix permeability can be used in flow
models to refine estimates of diffusion.
Typically, matrix permeability will be orders of magnitudes lower Fig. 31. Schematic of face and butt cleats; view is looking down on the bedding plane.
than fracture permeability in the same coal. It is generally thought Note the more continuous nature of face cleats and the termination of butt cleats at the
that matrix permeability is higher in lower rank coals because, among face cleats.
58 T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81

The permeability of many CBM reservoirs is anisotropic, i.e. perme-


ability is greater in one orientation than another, and this is thought
to be the influence of differential in situ stress on cleats (Paul and
Chatterjee, 2011). The directionality of permeability is important in
optimising well placement. Anisotropy of permeability has mostly
been ascribed to the greater continuity of the face cleat, which occurs
in one direction, over that of the butt cleat, which occurs in a nearly
perpendicular direction. The specific nature of the connectivity
between the butt and face cleats however may prove to be as important
in controlling permeability anisotropy as any other property (Law,
1993).
Cleat aperture (i.e. the width of the opening) is another important
property that has an immense impact on reservoir permeability.
Obviously, direct study of cleat aperture under reservoir conditions is
difficult. However, Laubach et al. (1998) attempted to understand the
relationship between cleat spacing, aperture size and permeability.
From hundreds of measurements in the San Juan and Black Warrior
basins they proposed the relationship shown in Fig. 33. This figure
shows that, at any particular aperture size, the effective permeability
increases with increasing cleat frequency (see again the above discus-
sion about coal type controlling cleat frequency and height). Although
Fig. 32. Schematic of cleat system in a coal face (view is perpendicular to bedding, i.e. replete with caveats on it use, this diagram does at the very least
opposite from Fig. 31, which is parallel to bedding). (A) Cleats confined to a single demonstrates how cleat frequency, aperture size and permeability
vitrain layer, (B) cleats within a package of multiple vitrain layers, bounded top and may relate to each other (though not necessarily in the absolute terms
bottom by durain layers, (C) master cleats which cross both vitrain and durain layers, given in the figure) in other basins.
(D) cleats that occur only in durain layers, (E) large master cleat/joint systems that
cross all coal and rock parting. From Dawson and Esterle (2010).
The rank of a coal undoubtedly plays an important role in cleat
development. Both Ammosov and Eremin (referenced in Close, 1993)
and Ting (1977) cite an increase in cleat frequency from lignite into
low volatile bituminous coal with a subsequent decrease above that
In terms of patterns of cleating, Dawson and Esterle (2010) found rank into anthracite. There are obviously other contributing controls
that cleat spacing was inversely proportional to cleat height (i.e. the that affect cleat spacing, as has already been discussed. The effect of
less frequent the cleats, the greater the length of any individual rank on permeability, in which cleat frequency plays a major role, has
cleat). In an example of this, vitrain layers occurring as isolated lenses already been discussed in Section 4.1.
(type 2 above) or groups of lenses (type 3 above) had the most The most predictable relationship with permeability is depth
frequent cleat spacing but were the shortest (see also Close, 1993). (though from basin to basin the exact permeability corresponding to
In contrast, durain layers had more widely spaced cleats (type 4 a depth will be quite different because of differing coal type, tectonics
above) with significantly more vertical and lateral continuity than and basin burial histories). Almost without exception, the deeper you
types 2 and 3. Finally, master cleats (type 1 above), which behave go, the less permeable the coal reservoir will be. Coal beds from the
more like conventional rock joints (Dawson and Esterle, 2010), are Carboniferous in the Upper Silesian Coal Basin (USCB) in Poland
very infrequent in their spacing but attain the greatest height. This (Fig. 34) exemplify this relationship. Similar relationships have also
relationship between cleat spacing and height with coal type is some- been noted in coal basins in Australia (Esterle et al., 2006; Krelle,
thing that can greatly improve our understanding of why and how 2011) (Figs. 35 and 36). Note that at any particular depth there can
permeability changes from coal bed to coal bed as well as horizontally still be considerable variability in permeability. In the example given
and laterally within a single coal. At present our predictive capabilities in Fig. 35, even at depths of 400 m the permeability varies from tenths
are still in need of refinement. to tens of a mD. The governing reasons behind these kinds of variations

Fig. 33. Relationship between cleat aperture, number of cleats (per inch) and predicted permeability. Developed for the San Juan basin. Modified from Laubach et al. (1998).
T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81 59

Fig. 36. In situ coal bed permeability versus depth in a coal seam from the Permian age
Bowen basin (Australia). Permeability has been segregated into geological domains
(i.e. geographic areas of similar geology and gas relationships) and two areas of
‘high’ and ‘low’ permeability have been identified. This has allowed a much tighter
relationship between depth and permeability. See Esterle et al. (2006) for further
explanation.

Fig. 34. Relationship between permeability and depth for some coals in the Upper
Silesian Coal Basin, Poland. Modified from McCants et al. (2001). within coal. However, localised shear zones can override and dramati-
cally change the permeability. This kind of tectonism has generally
been thought to increase permeability (Karacan and Okandan, 2000;
Li and Ogawa, 2001). Equally important in terms of permeability is the
are not well understood but have to be in part related to coal type (see style of shearing. When the coal is deformed in a brittle or cataclastic
discussion above on cleat and coal type) as well as the difficulty in manner, permeability can be enhanced 3–8 times above the surround-
obtaining reliable permeability measurements. A study by Esterle et ing reservoir (Li and Ogawa, 2001). This type of shearing tends to
al. (2006) demonstrates that sometimes a clearer relationship between open up existing cleat apertures and increase their interconnectedness,
depth and permeability can be gained through selectively partitioning thus increasing permeability (Li et al., 2003). In contrast, when shears
data into sub-populations with different trends that often relate to behave in a more ductile or mylonitic manner the resulting permeabil-
areas of similar geological character (Fig. 36). ity has been demonstrated to be greatly reduced (Gurba et al., 2001; Li
As already stated, a basin's stress–strain relationship (past and pres- and Ogawa, 2001). With mylonitic deformation the coals tend to be
ent) will have a huge impact upon the orientation of the cleat system tightly compressed and cleat apertures collapse (Li et al., 2003). These
types of shear zones have also been related to gas outburst zones in
underground mines (Gurba et al., 2001).
It is beyond the scope of this paper to describe how in situ perme-
ability is measured. However, because of its importance, a brief mention
is warranted. There are many techniques to measure permeability of a
coal seam in the bore hole (and they go by many names) but the basic
types are:
• Injection fall-off test (IFO),
• Drawdown recovery test (DDR),
• Drill stem test (DST), and
• Diagnostic fracture injection test (DFIT).
The specifics of various tests and combination of tests are defined
and discussed in Clarkson and Bustin (2011), Koenig and Schraufnagel
(1987), Ramurthy et al. (2002), and Wang and Wu (2011). All of
these tests monitor changes in water flow and/or pressure as the reser-
voir is injected or drawn down and allowed to recover. However, it is
generally accepted that different types of permeability testing will
yield different results, even for the same reservoir and in the same
drill hole.

5.4.3. Induced fracturing


The permeability of coal seams cannot be increased (Close, 1993).
Coal seams, like other gas reservoirs, can be artificially fractured in
order to increase connectivity of those fractures. However, because
Fig. 35. In situ coal bed permeability versus depth in a Permian coal basin in Australia of both the reactive nature of coal (i.e. its high carbon content will
(location is not given because of confidential nature of data). behave very differently to fracking polymers than conventional
60 T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81

reservoirs) and its ductile nature (i.e. coal's high compressibility) samples were taken. A normal hydrostatic pressure gradient is usually
hydraulic fracturing of a coal seam may not be as straight forward used (see Section 3.5) unless a direct measure of pressure has been
as fracking a sandstone or other conventional gas reservoirs. Note made at a few different depths.
that there is some controversy around what is the correct spelling The importance of gas saturation is illustrated in Fig. 37. In this
for this word: frack, frac, fracking or fraccing. For the purpose of this example, measured gas was determined from a sample at a depth that
paper, frack and fracking are used. would equate to the hydrostatic pressure Ps and was only 50% of the
In basins that are under a compressive stress regime, proppant maximum possible holding capacity at that pressure (Fig. 37a). Before
(fine sand particles, either natural or artificially made of ceramic gas can be produced, the reservoir would have to be depressurised
spheres or other materials) may have to be used in order to keep (i.e. de-watered) to a pressure of Pp. Only upon reaching pressure Pp,
the fractures open after fracking. The generally ductile nature of coal, commonly referred to as the critical desorption pressure, would gas
however, can mean that even fracking and proppant may not be enough begin to flow. If Pa is the pressure at which the well would be aban-
to keep the newly developed (and expensive) fracture systems open. doned (because of economic reasons, i.e. it costs more to run the well
The key to fracking is finding the balance between expenditure on than the return on revenue from the gas produced) then the difference
the frack itself and the resultant gas flow. A good understanding of between the gas level at Pp minus Pa is the estimate of the total produc-
the stress regime, permeability and structure of the coal seam is ible gas. At 85% gas saturation (Fig. 37b), one can see the significant
crucial to the success of any fracking operation. increase in the amount of hypothetically producible gas. This has huge
implications in the economics of a CBM play, both in the time it takes
5.5. Gas saturation to get a well to produce gas (think delayed return on the capital invest-
ments) and the cost of water disposal during the depressurisation
The degree of gas saturation is arguably the second most important phase. Note too that the shape of the curve can have an impact on
parameter (after permeability) in a CBM reservoir. Without sufficient de-watering times. For example, the curve for the low-rank coal
gas saturation, gas flow will not be economic. A general rule of thumb shown in Fig. 20 may be less sensitive to changes in the level of gas
in the CBM industry is that reservoirs need to have greater than ~ 70% saturation than the high-rank coal.
gas saturation in order to be economic. Having said that, some reser- Gas saturation can vary vertically within a single seam. Mares et al.
voirs with low gas saturations can and have been economically devel- (2009b) showed that, even within a single 5 m seam, the gas saturation
oped. For example, initial CBM developments in the Powder River varied from 66% to over 100% (Fig. 38). This is related to the degree of
Basin were in areas that were quite low in gas saturation (Flores et heterogeneity of the seam that ultimately will relate back to coal com-
al., 2005) but because production wells were shallow and inexpensive position (see Section 6). Many researchers believe that a coal cannot
and pipelines near to market, operators could profitably extract gas. be oversaturated. If gas saturation values are greater than 100%, as in
In order to calculate gas saturation, both the level of gas charge as the case of the Mares et al. (2009b) study, then the argument is that it
well as the maximum holding capacity of any particular coal at any par- must be analytical error. The salient point of the study by Mares et al.
ticular depth and temperature must be known (see also Sections 3.5 (2009b) is that, without the proper sampling density (see Section 8.1
and 3.6). All measurements must also be on the same analytical basis for a brief discussion of uncertainty), it is easy to over or under-
(e.g. as-received or dry ash-free). A simple way to envision the interplay estimate the average gas saturation of a reservoir even at a single loca-
of gas charge and maximum gas holding capacity in relation to the tion, within a single seam.
percent gas saturation is to think of a set of cups filled to different levels. There are other factors beyond organic composition affecting gas
The size of the cup represents the maximum holding capacity and the saturation. Gas saturation can be decreased in at least three possible
amount of liquid in the cup represents the gas charge. These two ways. Firstly, the coal seam can be buried to such a depth that temper-
together determine the percent to which the cup is filled. For example, ature is high enough to reverse, i.e. decrease, the gas holding capacity of
if a 100 ml cup is filled with 50 ml of liquid it would equate to 50% full the coal, expelling gas (see Figs. 39 and 40). This, in and of itself, will still
(or in CBM terms, 50% saturated). The complication comes in when dif- result in the coal being fully saturated. If that same reservoir is uplifted,
ferent coal seams, coal types within single seams, and different depths with concomitant recovery of its original gas holding capacity, the gas
all combine to result in a plethora of different holding capacities saturation will be lowered (unless the reservoir is somehow recharged)
(Fig. 21). A coal reservoir may hold, for example, 100 BCF of gas, but if (see also Ayers, 2002; Bustin and Clarkson, 1998; Khavari-Khoransani
that represents only 50% gas saturation, then the amount of recovery and Michelsen, 1999; Scott et al., 1994). Some experimental data, how-
would be low (see discussion of this below). In contrast, a 50 BCF coal ever, suggests that only reservoirs subjected to relatively high pressures
reservoir at near 100% gas saturation may yield more recoverable gas and temperatures at depth will lose holding capacities (Bustin and
(all other parameters being equal) than the under-saturated 100 BCF Bustin, 2008).
example. Secondly, gas saturation can also be lowered if uplift is sufficient
The formula used to calculate gas saturation is simply: enough to allow the gas to escape to the surface. This is common
throughout the Carboniferous of Europe and the UK (for example see
g ¼ 1−ðða−dÞ=aÞ ð3Þ Hildenbrand et al., 2006; Kędzior, 2009, 2011). This, in part, might be
related to the higher permeability at shallow depths, which more easily
where, allow gas leakage (as possibly seen in the Sydney Basin, Australia;
Bustin and Clarkson, 1998). Thirdly, there is some evidence that, in
g % gas saturation hydrodynamically active basins/reservoirs, the movement of water
a the maximum holding capacity for a particular depth can strip methane (Cui et al., 2004; Qin et al., 2005) and/or inhibit bio-
d the measured, or desorbed, gas volume (‘gas charge’) of a genic gas production (Bustin and Bustin, 2008).
sample at a specific depth. It is now believed that gas saturation can be increased, primarily
through biogenic recharge, if a reservoir is located at an effective
When calculating the percent gas saturation it is important that depth and geography for introduction of the necessary microbial con-
both the measured gas and the maximum holding capacity are sortium (Kędzior, 2011). Biogenic gas generation supplementing ther-
expressed on the same coal quality basis. Some operators express mogenic gas in a high-rank coal has been hypothesized for a coal bed
these measurements on an as-received basis whilst others report in Japan (Shimizu et al., 2007). Similarly, other basins are thought to
them on a dry ash-free basis. A key factor in estimating gas saturation have had secondary biogenic gas generation when a seam, previously
is the correct determination of the pressure at the depth from which deeply buried, has been uplifted and re-inoculated with methanogen
T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81 61

Fig. 37. Comparison of effect of two different gas saturations on de-pressurisation and predicted ultimate gas production. (A) 50% gas saturation and (B) 85% gas saturation. See text
for further explanation.

consortium (see Ayers, 2002; Faiz et al., 2003; Flores et al., 2008; telovitrinite and the volume of micropores. In contrast, in a study of
Pashin, 2010; Smith and Pallasser, 1996). New Zealand Eocene age coals, Mares et al. (2009b) cite a negative cor-
relation between telovitrinite and macroporosity. Chalmers and Bustin
6. Coal composition–gas property relationships (2007) found that ‘reactive’ inertinite had similar pore architecture to
vitrinite and thus similar methane holding potential. (Note that
Examples of the role of coal composition on gas properties have telovitrinite is a submaceral of vitrinite but represents the relatively
already been presented, a few times in this paper. Although coal non decayed remains of recognizable plant remains whilst inertinite is
rank is a, if not the, major factor affecting a CBM reservoir, it is the the most carbon rich maceral group and represents both plant remains
original plant components that form the foundation of how gas which have been oxidised as well as fungal remains; ICCP, 1998, 2001).
behaves in coal. Only three gas reservoir properties will be discussed Coals with higher non-reactive inertinite have a greater abundance of
in relation to composition, though many more properties are affected macro- and mesoporosity as opposed to microporosity (Adeboye and
(e.g. see Busch and Gensterblum, 2011). It is also recognised that Bustin, in press; Clarkson and Bustin, 1996; Lamberson and Bustin,
many studies find no relationship between coal composition and gas 1993). Interestingly, Chalmers and Bustin (2007) also found that in
reservoir properties. liptinite-rich coals (e.g. ‘boghead’ coals) methane is held as solution
gas within the pores as opposed to being adsorbed on pore surfaces.
6.1. Pores and permeability This is an example of a depositional environment, albeit at one end of
the depositional system spectrum, influencing pore type distribution
A fundamental attribute of a CBM reservoir, is its pore character in a CBM reservoir. Zhang et al. (2010) also cite a correlation between
(and thus sorption character), and this is indisputably connected to micropore distribution and coal facies, though more work on this
its organic composition (Ceglarsk-Stefańska and Zarębska, 2002), needs to be conducted.
which, in turn, is a result of the original peat and depositional setting If the pore system is influenced by the organic composition of coal
(see Section 4.2). then it is reasonable to assume that permeability (both fracture and
It is well established that micropore volumes are associated with matrix) is also influenced (given that permeability is the level of con-
coals with high vitrinite content (Bustin and Clarkson, 1998; Clarkson nectivity between the voids i.e. the pores). Adeboye and Bustin (in
and Bustin, 1996; Unsworth et al., 1989). Within an isorank series, press) argue that matrix permeability is greatest in high inertinite
Chalmers and Bustin (2007) found that bright lithotypes (i.e. vitrinite‐ coals where macro- and meso-porosity predominate over micropo-
rich) in Australian coals generally had greater pore volume than dull rosity. Similarly, Wang and Ward (2009) demonstrate from laboratory
coals. Specifically they cite a positive correlation between the maceral experiments that high vitrinite coal samples (with all other coal quality
62 T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81

Fig. 38. Profiles of (A) measured gas and maximum holding capacity and (B) total percent gas saturation in a 5 m subbituminous coal seam. Note the variability in both measured
gas and maximum holding capacity even over a relatively small coal bed. How many samples would be needed to accurately approximate the mean? (see Mares et al., 2009a for an
answer).

parameters being about equal) have the highest total permeability at lithotypes rich in vitrinite tend to have an increased frequency of
any given pressure as well as having the highest rate of increase in per- cleats and thus have a correspondingly higher permeability than
meability as pressure decreases (Fig. 41). It is also noted that ash yield coals with lower vitrinite (Ayers, 2002; Smyth and Buckley, 1993).
may have a profound effect on permeability. The combination of the composition of the coal and the cleat patterns
Finally, as already discussed in Section 5.4.2, coal lithotypes (which has been termed ‘fabric’ by Bustin (1997) and no doubt the interplay
are differentiated by their macro-and microscopic composition) affect between these two attributes determines desorption, diffusion and
both the degree of cleat spacing as well as cleat continuity. Coal permeability in a coal reservoir.

Fig. 39. Time schematic of biogenic and thermogenic gas generation as related to burial and effect on gas saturation. (A) Peat is formed and buried, primary biogenic gas gives way to
secondary biogenic with depth. (B) Continued burial and rank increase have the effect of moisture loss and increase in maximum holding capacity, gas saturation is variable
depending on local conditions being conducive to biogenic gas generation. (C) With burial and temperature rise (and time) thermogenic gas begins to be generated, supplementing
biogenic gas. Eventually, burial temperature will be too great and all biogenic gas will cease. Thermogenic gas will be at its maximum and thus gas saturation, and if geological con-
ditions are right, could be 100%. (D) Burial reaches a depth where temperature is so great that maximum holding capacity actually decreases (as per ideal gas law); gas is expelled
but gas saturation remains the same. Maximum depth and maximum rank are achieved. (E) With uplift, temperature decreases and maximum gas holding capacity increases. How-
ever, since some gas has been expelled, gas saturation gets concomitantly lower. (F) In some cases if uplift is substantial and coal seams subcrop at the surface, biogenic enhance-
ment may result in re-saturation of coal seams.
T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81 63

Fig. 40. Graphical representation of Fig. 39. See Fig. 39 for explanation.

6.2. Maximum gas holding capacity Laxminarayana and Crosdale (1999, 2002) and Vessey and Bustin
(1999) among others. These latter studies don't always arrive at exactly
If organic composition affects the pore volume of a coal reservoir, the same conclusions as each other though, with Laxminarayana and
then the maximum gas holding capacity will also be affected. An early Crosdale (1999) suggesting that organic composition had less influence
study by Lamberson and Bustin (1993) found the highest gas holding at higher ranks and Chalmers and Bustin (2007) concluding the opposite.
capacities associated with vitrinite-rich coal types in the Gates As more studies are conducted, however, there does seem to be a
Formation of western Canada; conversely the lowest gas holding capac- fairly consistent positive association between vitrinite abundance and
ity was in the high inertinite samples. These results were later supported maximum gas holding capacity (Fig 42; Bustin and Clarkson, 1998).
by the work of Chalmers and Bustin (2007), Crosdale et al. (1998), Conversely, higher inertinite will decrease the gas holding capacity
(Fig. 43). Other variables such as submaceral type, inorganic content
and the overall fabric of the coal will no doubt make any correlation
far from perfect.
The effect of inorganic matter on the gas holding capacity of coal is
well documented. Lamberson and Bustin (1993) found that surface
area, and thus gas holding capacity, decreased with increased inorganic
material in the coal. Crosdale et al. (1998) also concluded that inorganic
material acted as a non-adsorbent material to coal gases, thus high ash
yield coal has lower gas holding capacity. Similarly, Laxminarayana and
Crosdale (1999, 2002) found that Langmuir volume decreased as ash
yield increased (Fig. 44). This sort of relationship can also be seen in
Fig. 25, where the highest gas holding capacity sample is actually a
slightly lower rank, as evidenced by vitrinite reflectance, than the
sample with the next highest holding capacity. The reason for this
‘switch’ is that the slightly lower rank coal sample (they all have roughly
the same organic composition, i.e. high vitrinite) has less than 1% ash
yield whilst the other sample has a ash yield of about 6% (air dried
Fig. 41. Influence of percent vitrinite of a sample on permeability. Note that at all pressures,
basis). Finally, inorganic material may also further decrease available
the higher the percent vitrinite, the higher the permeability. Modified from Wang and methane adsorption sites if secondary diagenetic and epigenetic min-
Ward (2009). erals are precipitated (Pitman et al., 2003).
64 T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81

Fig. 44. Relationship between % ash yield and Langmuir volume. Modified from
Laxminarayana and Crosdale (1999).

seams tend to be slightly lower rank based on vitrinite reflectance. A


straight comparison between measured gas and liptinite content
(Fig. 45) shows a lot of scatter but still is convincing enough to indicate
Fig. 42. Relationship between percent total vitrinite (raw coal) and maximum gas
holding capacity. Modified from Bustin and Clarkson (1998).
some sort of trend. Examination of a single well (Fig. 46) affords a better
comparison and clearly indicates an association between measured gas
and liptinite. Hemza et al. (2009) and Mares (2009) also found an asso-
ciation between hydrogen (liptinite being hydrogen-rich) and measured
6.3. Desorbed gas gas content. In biogenic plays (as many of these cited above are) liptinite
may act as a better feedstock for the microbial consortium generating
Unlike adsorption studies, there is a paucity of published data methane.
relating organic composition to the amount of measured gas (i.e. In coals where vitrinite is dominant, the relationship between coal
desorbed gas). In part this is a result of the expense and, in most composition and measured gas contents is harder to reconcile. In the
cases, the commercial sensitivity surrounding obtaining desorption Powder River Basin, there is some indication that coal with more
data in a CBM play. It is also easier to collect adsorption samples and abundant vitrain will contain more gas at a given pressure or depth
conduct laboratory experiments than undertake full scale, meaningful (Moore et al., 2002; Fig. 47a). Esterle et al. (2006) also found a similar
studies on core samples that have been desorbed following rigorous relationship in Australian medium volatile rank coals (Fig. 47b). How-
procedures. ever, these associations may not be causal and other factors such as
In one of the few studies, Scott et al. (2007) characterised the pre- permeability and gas migration may be at work. Butland and Moore
dominately low-rank (mean vitrinite reflectance from 0.40 to 0.54%) (2008), Mares (2009), and Mares and Moore (2008) also found
Juandah and Taroom Coal Measures (Australia) in terms of measured vitrain abundance to be related to increased gas volume in the Eocene
gas content and organic composition. They found a general increase age, subbituminous Kupakupa coal seam of New Zealand, but the
in liptinite content of the coal beds down section from the Juandah overlying seam (the Renown) displayed the opposite relationship.
into the Taroom Coal Measures and then a decrease in liptinite content This is curious and if all the data for both the Kupakupa and Renown
at the bottom of the Taroom Coal Measures. This variability in liptinite coal seams in the Mares (2009) study are considered only by coal
content was found to have a direct correlation with gas content. Addi- type, the dominant relationship is that there is a negative association be-
tionally, the lowest gas contents, at the top of the Juandah Coal Measures, tween measured gas and vitrain banding (Fig. 48; Mares and Moore,
are associated with the highest inertinite contents; albeit these coal 2008). Although perhaps counterintuitive, this might be explained

Fig. 43. Relationship between % total inertinite (whole coal basis) and maximum gas holding capacity. From the Xishanyao Formation, South Junggar Basin, China. Isotherms
conducted at 30 °C by GeoGas Pty. Ltd. (Australia). Vr = random vitrinite reflectance; Vr and maceral analysis conducted by Coal & Organic Petrology Services Pty Ltd. (Australia).
T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81 65

Fig. 45. Measured gas content compared to liptinite content in coal seams from the
Juandah and Taroom Coal Measures, Australia. Modified from Scott et al. (2007).

through the higher number of micropores, and thus surface area found
in coal types with less vitrain banding (Mares et al., 2009b).
Other studies have shown other correlations such as a positive
relationship between residual gas content and gelovitrinite (and a
corresponding negative relationship with telovitrinite; note that
gelovitrinite is a submaceral of the vitrinite group. It is generally amor-
phous in nature, and represents the decay products of plant material;
ICCP, 1998) or a positive relationship between lost gas and fusinite
(Fig. 49; Butland and Moore, 2008). In an early study, Ettinger et al. Fig. 47. Relationship between degree of vitrain banding and gas content in (A) a coal
(1966) found higher gas volumes related to an increase in inertinite bed from the Powder River basin, Wyoming (cited in Moore et al., 2002), and (B) in
content. The predictive power of these associations should be viewed medium volatile rank coal from Australia. Modified from Esterle et al. (2006).
with some caution, however, as these relationships are not strong and
are clearly not universal because of the differences in microstructure
and organic composition. 7. Production, modelling and water
As would be expected, coals with high inorganic content have low
measured gas (Wang, 2006; Warwick et al., 2008). This follows the 7.1. Production profiles and modelling
same trend that inorganic material has on maximum gas holding
capacity. Inorganic material cannot adsorb gas and thus the higher A CBM reservoir is different from a conventional gas play, thus it is
the ash yield of a coal, the less gas it can hold. This has been observed not surprising that production profiles between the two are also very
in many basins including the Sydney basin (Faiz et al., 2007b), the Gulf different. The fundamental distinction is that, in a conventional gas
Coast (Texas; Warwick et al., 2008) and most coal basins in New play, gas flow rates start high and, in most cases, almost immediately
Zealand (Fig. 50; Butland and Moore, 2008). However, Mares and begin a fairly steep decline with the final phase of the well being
Moore (2008) found that ash yield had to reach a threshold of 10% marked by water production. In contrast, in a CBM play, water volumes
before a reasonably strong negative correlation with measured gas start high with low gas flow rates but as the reservoir is progressively
was obtained. Below that threshold, no correlation could be discerned dewatered (i.e. depressurized) gas rates continue to rise to a peak
and variation was most likely related to organic composition and ulti- rate, months or years after dewatering started in the well (Fig. 51). As
mately pore size distribution (see previous discussion in Section 6.1
on pore size and organic composition and Mares and Moore, 2008;
Mares et al., 2009b). It should be noted that these relationships hold
for these coals and may not necessarily be applicable to other coals;
more studies are sorely needed.

Fig. 46. Comparison of measured gas content and liptinite in a single well. Increasing Fig. 48. Relationship between the degree of vitrain banding and measured gas content
sample number related to increasing depth. Modified from Scott et al. (2007). in two New Zealand coals. Modified from Mares and Moore (2008).
66 T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81

Fig. 49. Examples of maceral associations with desorbed gas contents. (A) Gelovitrinite
vs residual gas (from the Greymouth coal field, New Zealand) and (B) fusinite vs lost
gas (from the Ohai coal field, New Zealand). See text for further explanation. Modified
from Butland and Moore (2008).

discussed in Section 5, these differences are a result of how gas is held in


coal versus a conventional gas reservoir.
Schematically, a CBM production profile is thought of as having Fig. 51. Schematic of water and gas production from (A) conventional gas well from a
three stages: a. dewatering, b. stable production and c. decline clastic reservoir and (B) CBM production profile showing the different phases of
production.
(Fig. 51b). The time required for dewatering is highly dependent on a
combination of the percent gas saturation and degree of permeability
(all mechanical things, such as pump size and quality of performance factors, including permeability, well spacing and well maintenance
being equal). The time between initial pumping and stable (peak) pro- frequency.
duction will have a huge consequence on the economics of a well. The The difference between a ‘good’ well and a ‘bad’ well, in terms of
rate of decline following peak production can be influenced by many production, is almost purely related to the rate and timing of gas

Fig. 50. Relationship between ash yield and gas content in four cores from New Zealand. Modified from Butland and Moore (2008).
T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81 67

flow. The worse well would take a long time to produce gas with wells will also have a profound effect on the shape of a production pro-
relatively low peak gas flow rates. A good well, of course, would be file. Even wells in the same reservoir (that is, the same seam), in close
the opposite. The shape of curves, even among good wells will vary sig- proximity (hundreds of meters), with virtually the same coal seam
nificantly between basins. For example, Fig. 52 shows some production properties can have hugely different gas flow profiles (Fig. 53). This
profiles from the San Juan (USA), Bowen (Australia) and Powder River kind of variability is something seen in every CBM field under produc-
(USA) basins. In this example the profiles differ for a variety of reasons tion and is the result of the complex interactions between multitudes
but fundamentally because of rank (the Powder River is subbituminous of reservoir properties (e.g. pressure, saturation, permeability). All
whereas the San Juan and Bowen Basins are generally high to med vol- these highlight the difficulty of predicting gas flow prior to actual
atile bituminous), gas charge (the Powder River generally has less than production.
3 m 3/t, the San Juan 5–9 m 3/t and the Bowen 4 to 16 m 3/t), net coal In the science community, it has become somewhat of a cliché to
thickness (the Powder River has generally 10s of meters of thickness say ‘garbage in, garbage out’, and no one would argue with the under-
whilst the San Juan and Bowen have 5–20 m maximum), and well lying truth of the statement. When it comes to modelling gas flow
design (both the Powder River and San Juan are vertical wells whereas through coal, however, there is probably no more apt a saying. Most
the Bowen is horizontal). The number and maturity of surrounding CBM exploration companies will have flow models for their prospects

Fig. 52. Three different production profiles from the (A) San Juan [USA], (B) Bowen [Australia] and (C) Powder River basins. Note the different y- and x‐axis scales (flow rate and
time, respectively) between the three different graphs.
68 T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81

Fig. 53. Example of three flow profiles from the Powder River Basin (Wyoming) illustrating the variability of production even in a small area. Wells are in Section 35 (1 mi2), Town-
ship 44, Range 72. Data from http://wogcc.state.wy.us/.

if only to feed into economic models that are essentially mandatory changes (e.g. shrinkage, changes in permeability etc.) that take place in
for reserve certification and/or taking a company into the public the reservoir as production proceeds over time. These coupled models
investment arena. What is usually produced is a single flow profile are still relatively specialised and are reported on by Deisman et al.
of the reservoir which is then extrapolated over the entire prospect
and that is what the economics are calculated on. In more sophisticated Table 1
simulations low, best, and high cases are run, but even in these, input Essential data input values for gas flow modelling.
parameters are usually still not that well constrained. In green field Parameter Unit Example inputs
(i.e. virgin to CBM production and essentially data barren) situations,
Well type Vertical
so little is known that input parameters for flow simulations can often Coal seam depth m 500.00
be tweaked to give a large number of possible outcomes. In fields Coal seam thickness m 15.00
where production has commenced, history matching of model outputs Reservoir temperature °C 42.00
to the actual flow profile removes a large amount of ambiguity sur- Hydrostatic head m 475.00
Adsorption gas volume m3/t 2.00
rounding reservoir behaviour (but certainly not all) and makes gas
Gas saturation % 80
rates more predictable. It is in these more mature types of basins Measured gas (gas charge) m3/t 1.60
where the effect of, for example, permeability changes, during produc- Absolute cleat permeability mD 20.00
tion become obvious (Palmer, 2009). Langmuir volume m3/t 3.58
The key inputs for a flow model in a CBM reservoir are given in Langmuir pressure kPa 1366.12
Initial water saturationa % 100
Table 1 (note the parameters which are usually estimated, all others Cleat porositya % 1
can be measured). As stated above it is wise to give a range for many Well diameter in. 6
of the parameters to reflect the uncertainty inherent in any particular Fracture half-length, xfa m 75
play. There are a variety of modelling packages that are available for Well spacinga Acre 80
Drainage areaa km2 0.4
estimating gas flow rates in coal seams including COMET, SIMED II,
Bottom hole pressurea kPa 4655
ECLIPSE, F.A.S.T.CBM and TOUGH II. This section does not cover Abandoned pressurea kPa 700
methods of gas flow modelling and the reader is directed to other pub- Ash content (%) 5
lications that have compared some of these programmes (Law et al., Moisture content (%) 7
2005; Wei et al., 2007; Zarrouk, 2008). Some of the more recent Coal density (cm3/g) 1.30

modelling packages couple gas flow dynamics and the geomechanical a


Usually estimated, other parameters can be measured.
T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81 69

(2008, 2010), Gu and Chalaturnyk (2006), Wei et al. (2007), and refer-
ences within. Some modifications to models have also been proposed
which would take into account the individual sorption and diffusion
character of individual gases (e.g. Clarkson, 2003). As can probably be
gathered, there are considerable refinements on flow models which
are still underway.
Regardless of which modelling package is used, a productive exer-
cise is to run sensitivities around key reservoir parameters. Fig. 54
shows the sensitivity around two key parameters already discussed:
gas saturation and permeability. As is obvious, even relatively slight
(modelled) changes in the values of those parameters result in signifi-
cant impacts on the production profile in terms of peak rate and cumu-
lative gas. Although a vertical well scenario is given in the example of Fig. 55. Sensitivity analysis of gas flow to permeability changes in a horizontal well,
Fig. 54, a horizontal well will show the same types of sensitivities consisting of two lateral wells of 1000 m length intersecting a vertical well in a single
(Fig. 55). Note that it is not the absolute values shown on the x- and CBM reservoir of 6 m thickness.

y-axes which are important in these examples, but the sensitivity of


gas flow to those key parameters. the coal beds (i.e. if it is a ‘confined’ or ‘non-confined’ reservoir—see
below), together with the permeability and gas saturation, among
7.2. Co-produced water other variables. A confined coal reservoir is one where the layers
above and below the coal seam are so impermeable that water move-
In almost all cases, methane cannot be produced from a coal seam ment across those layers is not possible (these layers are termed an
without producing a significant amount of water first. There is no ‘aquitard’). For example, the coal layers in the Black Warrior basin are
doubt that water produced from CBM wells can have an effect, both thought to be confined aquifers (Pashin, 2007) whereas coals in the
in regard to depletion of ground water resources as well as impacts Powder River basin are generally thought to be unconfined aquifers.
on the environment from its disposal. In the latter case, dewatering is not restricted to just the coal seam
The volume of water produced from CBM wells is highly variable, and can encompass a number of other rock units in the stratigraphy
from virtually zero to greater than 2000 barrels a day (the amount of until an aquitard is encountered.
water coming from a CBM well is often related to pump size and the
desire of the operator to regulate water take so as not to damage the
reservoir). In a typical CBM production well, the water is pumped to a
centralised tubing and pump arrangement and as water drawdown
proceeds, the gas rises up between the outside of that tubing and
inside of the casing (Fig. 56). The ultimate amount of water pumped
from any particular well is a function of the hydrodynamic nature of

Fig. 54. Sensitivity analysis where all variables are held the same except for (A) percent
gas saturation and (B) permeability to demonstrate their effect on gas flow (note: (A) Fig. 56. Schematic diagram showing an open hole completion with a screen (to protect
and (B) come from two different real world data sets). the pump) for a typical CBM well.
70 T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81

A number of studies have shown that the extent of the recharge shallow aquifers and into streams (McBeth et al., 2003). In the Pow-
area and local geology can have a profound effect on the reservoir der River basin, Jackson and Reddy (2007) found that Na, alkalinity,
pressure, water chemistry and production performance of wells SAR and pH downstream from unlined impoundment ponds were
(Anna, 2003; Ayers, 2002; Myers, 2009; Pashin, 2007; Scott, 1993). significantly higher than background levels. In addition, Healy et al.
In the Black Warrior basin, recharge plumes were found to control (2011) concluded that if co-produced water passed through the
water chemistry; water pumped from wells that are within the local alluvial sediments, the water quality became even worse. The
recharge plumes can be safely discharged at the surface with little processes such as gypsum dissolution, cation exchange and pyrite
or no treatment whereas away from the recharge plumes, the water oxidation are suspected as the culprits in decreasing the water quality.
chemistry is such that it needs to be reinjected into other strata The land surrounding impoundment ponds (or wherever the
(Pashin, 2007). The San Juan basin is also thought to have similar co-produced water has contact with the soil) can also be impacted.
recharge areas that penetrate into the Fruitland Formation at depth There is a general increase in both the salinity and sodicity of the
(Ayers, 2002). However, other studies have indicated that inflow soil, and as discussed above, SAR values also will increase (Stearns
into the formation is quite limited (less than 5 km from outcrop; et al., 2005). There also tends to be more invasive, exotic plant species
Snyder et al., 2003), although microbes and nutrients are thought to that inhibit these sites; ones that are more salt tolerant (Bergquist et
still be able to move into the deeper parts of the basin (Zhou et al., al., 2007; Stearns et al., 2005).
2005). Recharge rates are of special importance in CBM development The treatment of CBM water to make it safe to discharge is a recent
as the number of wells and the amount of water are often very large. focus of research. Part of the issue is the sheer volume of water that
Anna (2003) found that the drawdown in some parts of the Ferron often needs to be treated. Various options of utilising treated (and
Sandstone coals extended to within a short distance from the outcrop. in some cases untreated) waters have been summarised in a report
In the Powder River basin, Myers (2009) found that the ≥50,000 by All-Consulting (2003). In some cases, a CBM company will set up
current wells that are presently in the basin could take up to 40% of a separate business just dealing with water (see, for example, www.
the water resources available in the target coal zone and even if pumping arrowenergy.com.au). Research such as reverse osmosis and integrated
stopped now, it would take 45 years to recover to an acceptable level of membrane systems is in the forefront of mass water treatment
recharge and over 200 years to fully recover. Obviously the manage- (Chalmers et al., 2010). Site specific solutions, such as using locally
ment of water resources needs care, foresight and rigorous study. sourced zeolites to reduce Na levels may also offer alternatives (Taulis
The geochemistry of the co-produced waters has been the focus of and Milke, 2009).
many studies, particularly surrounding the Powder River basin. This is To conclude, co-produced water is inevitable for almost all CBM
unsurprising considering the number of wells drilled and the volume developments. Where the fields are large and volume of water sub-
of water discharged as well as the speed at which it has occurred. stantial, the challenges will be significant. Some of the challenges
Regardless of which basin, just about all CBM co-produced waters will be perceived but many, if not most, will be real and CBM operators
have similar chemistries (Van Voast, 2003). In general, the waters will have to plan their water management systems carefully.
are dominated by sodium (Na) and bicarbonate and devoid of calcium
(Ca), magnesium (Mg) and sulphate (except when marine-derived 8. Key reservoir measurements during exploration
sediments are proximal to the coal deposits). This distinct geochemical
signature evolves from the process of biochemical reduction of sul- The various influences on and characteristics of a CBM reservoir
phate, enrichment of bicarbonate and the precipitation of Ca and Mg have been discussed. But what is the essential data that needs to be
(Van Voast, 2003). Co-produced waters that are closest to recharge collected during exploration? There are many evaluation methods
areas are relatively fresh but with depth the geochemical signature and tools that can be used (for example, see Clarkson and Bustin,
will become more prominent. The basic geochemical signature tends 2011). Which should be used depends on the goals of the exploration
to occur regardless of the coal formation, geographic position or age. programme.
This observation has been borne out by other studies in other CBM fields Essentially, exploration is an exercise to reduce (or at least define)
(Anna, 2003; Cheung et al., 2009; Dahm et al., 2011; Rice, 2003; Rice et uncertainty and thus risk. Collecting data during exploration is always
al., 1989; Rice et al. 2000; Taulis and Milke, 2007). However, Dahm et al. an effort and is always costly but the quantitative information gathered
(2011) found that slight variations of the geochemical pattern (though helps to decrease uncertainty (see also Bratvold and Begg, 2008;
still very recognisably a ‘CBM water chemistry signature’) can be Bratvold et al., 2002). Taking it as a priori that gathering primary data
influenced by the type of coal deposit, its rank and the proximity to is crucial to a CBM exploration programme, let us define what are likely
fresh water recharge. to be the two main goals:
In a particularly extensive study, Dahm et al. (2011) looked at the
1. Define gas in-place to a sufficient certainty to allow reserve certifi-
water quality from over 3000 CBM production wells from four basins
cation over a maximised area (see Section 10.2), and
in the Rocky Mountain region. The study assessed water quality for a
2. Gain information on the reservoir, which will give maximum success
multitude of parameters but the major ones were total dissolved
to a pilot programme, with success being defined as representative
solids (TDS), chlorine, bicarbonate, fluorine, iron, sodium adsorption
gas flow rates being obtained.
ratio (SAR) and temperature. The results were compared to the US
standards for drinking water, livestock and irrigation. For drinking To obtain these goals, three crucial pieces of data need to be
water, almost all of the wells were outside the standard whereas for collected from the target reservoir:
livestock use some basins were mostly compliant and some weren't.
a. Rock (coal) volume,
Water temperature was also a major issue for livestock. SAR levels
b. Level of gas saturation (which requires both desorption and
are an issue for most basins, especially the Powder River. Waters
adsorption analyses), and
with high SAR, if used for irrigation, will result in salt accumulation,
c. Permeability.
effectively killing off native plant species that are not salt tolerant (see
below). Interestingly, other than SAR, the Powder River co-produced Firstly, determining the volume of a reservoir is a basic geological
water meets standards more closely than water produced in other undertaking. The accuracy to which reservoir volume can be deter-
basins. mined is dependent on the data density and the inherent heterogeneity
Co-produced waters are often held in impoundment ponds and of the coal seam (Moore et al., 2009). In carrying out this task, it is essen-
many of those ponds are unlined. The issue with unlined impound- tial that the original defining well logs or measured sections should be
ment ponds is the possible leaching of the co-produced waters into sought and examined.
T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81 71

Secondly, gas saturation is absolutely essential to determine, for profitability for an operator who may derive economic benefit through
the reasons already discussed in this paper. The two measures needed sequestration of CO2. This technique is commonly referred to as
to estimate saturation are gas charge, measured through desorption ‘enhanced coalbed methane’ (ECBM) and is in many ways equivalent
procedures, and characterisation of maximum gas holding capacity to the goals of enhanced oil recovery (EOR) in conventional reservoirs.
through an adsorption isotherm test of coal samples. Numerous studies have examined the processes by which CO2
The third and final crucial reservoir property—permeability—can be adsorption occurs within coal (e.g. Mazumder and Wolf, 2008;
quite elusive to obtain in an exploration programme—but more often Ozdemir, 2009; Pan and Connell, 2007). A review of the kinetics and
than not, worth the effort (and expense). Probably the best, most reliable sub-molecular processes of ECBM has recently been completed by
permeability measurements are obtained in either pilot or production Busch and Gensterblum (2011). Models for ECBM and CO2 sequestra-
wells where the well bores are cased off, stable and long-term tests can tion are also given in Palmer and Mansoori (1998), Pan et al. (2010),
be conducted without the pressure of rig time and concomitant expense. and Shi and Durucan (2005). Mazzotti et al. (2009) reviewed ECBM
During an exploration drill hole, however, there is no such luxury. There and specifically outlined pilots and micro-pilot projects in Canada,
is no one best method to obtain accurate permeability measurements in Japan, the USA, China and Poland that have occurred over the last
a coal seam (see Section 5.4.2), especially during exploration drilling. 15 years.
Distilled down to the very basics, an ECBM project involves at least
8.1. The importance of sampling protocol and mitigation of uncertainty two wells: a CBM production well, which has had some level of deple-
tion, and a second injection well at a distance from the production
Without proper sampling, an exploration programme runs the risk well (the distance being defined by the permeability attributes of
of being invalidated or at the very least will not decrease uncertainty. the reservoir, i.e. anywhere from a couple of hundred to 1000 m or
In assessing reservoir properties for CBM, there are two rules to more). In theory, CO2 would be injected at some rate into the reservoir.
follow in order to avoid invalid and/or ambiguous results: The carbon dioxide, preferentially replacing the methane adsorbed onto
the surface of the coal would create a methane front that would increase
1. Samples need to be fresh, collected as close to in situ, as possible,
flow within the production well. Over time, the CO2 used to push out the
and
methane would ‘break through’ into the production well and at that
2. Multiple samples for each and every analysis needs to be taken.
point the well would be shut in (e.g. Ross et al., 2009; Sinayuç and
Coal properties, especially of low-rank coal, are affected by loss in Gümrah, 2009; Zarrouk and Moore, 2009). As would be expected, pilot
moisture and oxidation with even the briefest exposure to air, there- scale projects indicate that reality can be slightly more complicated.
fore samples have to be collected in as close to in situ conditions as Two pilot projects illustrate the challenges of ECBM using CO2. In
possible. This means that, in almost all cases, samples will need to the first, the Yubari project (testing took place between May 2004 and
be collected from coal core at the time of arrival at the surface. October 2007) in the Ishikari coalfield of Japan attempted to inject su-
Adsorption samples are especially critical to take immediately and percritical carbon dioxide into a water-saturated 6 m coal seam at a
then protect appropriately from the atmosphere, as even a small amount depth of about 900 m (Fujioka et al., 2010; Shi et al., 2008). Original
of moisture loss may result in significant over estimation of maximum rates of injection were anticipated to be between 16 and 25 t/day of
gas holding capacity (Crosdale et al., 2008). CO2 but the actual rate was only ~14% of that (i.e. between 2.3 and
Because of the inherent heterogeneity of coal, considerable uncer- 3.5 t/day). With such slow rates, the injection was temporarily halted
tainty will be associated with any measured parameter. Uncertainty and N2 was injected in an effort to reduce what was suspected as swell-
can be reduced through identification of the major variables that ing in the coal. When they later re-injected, CO2, rates were better, but
have the greatest impact. The most significant variables may change quickly fell off again.
from basin to basin or coal reservoir to coal reservoir. The collection Laboratory studies seem to support that it was matrix swelling
of more data for these crucial parameters may reduce the uncertainty, causing the lower than expected CO2 flow rates in the Yubari project.
moving what is known about a deposit from largely ambiguous, Kiyama et al. (2011) used coal from the same seam and mimicked the
through a range of possibilities to clear direction. conditions of the Yubari field test in the laboratory. They found in-
The quantification of uncertainty, however, needs to be communicat- creased strain rates, which indicate swelling of the coal. An injection
ed throughout the decision making process (Bratvold and Begg, 2008; of N2 in the laboratory sample was found to increase the permeability,
Zabalza-Mezghani et al., 2004). For example, original gas-in‐place but re-injection of CO2 again caused increased strain rates (swelling)
should always be expressed in probabilistic terms (Moore et al., 2009), and a concomitant decrease in permeability. It seems fairly conclusive
giving a best, median and low case. A single number, such has been that swelling caused by CO2 in the Yubari project reduced the perme-
reported for some counties (e.g. Hadiyanto and Stevens, 2005; Stevens ability of the coal. Swelling caused by CO2 adsorption with related re-
et al., 2001) gives no idea of the uncertainty surrounding a value and duction in effective permeability has also been observed in numerous
may give a false impression of certainty where little exist. other laboratory studies (Liu et al., 2011; Siemons and Busch, 2007;
Siriwardane et al., 2009; Wu et al., 2011). The implication of the
9. Enhanced production Yubari pilot is that the process of CO2 injection into coal differs from
injection into clastic rock and, because of the reactivity of the coal,
Over the last decade numerous studies have looked at carbon capture may not be feasible i.e. ECBM with CO2 may not be feasible in all
and storage technology (CCS) in conventional and unconventional reser- coal seams. This type of reaction by the coal seam, however, may
voirs. The motivation for this has mostly been the concern regarding the not necessarily be all bad news. Pan and Connell (2011b) modelled
concept of anthropogenically-driven climate change and a consequent interbedded coal and rock and the results seem to indicate that coal
desire to reduce CO2 emissions. Coal beds can hold a disproportionately facilitates the storage and migration pathway of the carbon dioxide
high volume of CO2 compared to CH4 because of the molecular shape of in the clastics, enhancing the capability of the rock layers to hold CO2.
carbon dioxide (Bromhal et al., 2005; Mastalerz et al., 2004; Ottiger et al., In the second field trial, conducted over a number of years in the
2008; Stanton et al., 2001; Zarrouk and Moore, 2009) and thus have San Juan basin (USA), a different result ensued. Injection of CO2
been a focus of much of this research (Bachu et al., 2000; Damen et al., began in 1995 with an initial rate of about 250 t/day. Like the Yubari
2005; Gunter et al., 1997; Wong et al., 1998; Wong et al. (2007); Yu et project, an abrupt decrease in injectivity was observed almost imme-
al., 2007). CCS studies have not only looked at how CO2 can be stored diately (Ozdemir, 2009; Reeves, 2001). However, after this decline,
in coal but also employed to expel methane from the seam. This could there was a slow, but steady increase in carbon dioxide injection rate,
potentially increase methane recovery rates as well as improve returning to a maximum of about 150 t/day. During the five years the
72 T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81

pilot operated, there was only minimum breakthrough of CO2 into the The NPV and %IRR in a CBM project are most sensitive (but not
production well. limited) to the following top parameters:
These differing results (and these are just two field trials) highlight
the heterogeneous nature of coal seams and their differing responses 1. Gas price
to ECBM with CO2. In addition, it is not at all certain that the CO2 will 2. Drilling and completion costs
be retained within the coal seam for any length of time as the current 3. Maximum gas flow rate (i.e. MCF/day)
state of modelling has a high degree of uncertainty (Korre et al., 2009). 4. Total well reserve (i.e. BCF per well)
Moreover, the economics of CO2 injection into coal seams may or may 5. Production per year of the whole field
not make it attractive to industry (Robertson, 2009; Wong et al., 2010)
and is something that, for now, will need to be assessed on a case- The costs of drilling and infrastructure are probably the only key
by-case basis. variables that a company has the highest influence over. During
exploration and pilot programmes, drilling and completion costs will
10. The impact of CBM be high; this is mostly related to the ‘one-off’ nature of the wells and
equipment. When development and production begin in earnest, the
For the practicing, as well as the more academic geologists, it is volume of capital requirements can help to obtain lower rates. However,
important to understand the drivers and impacts of CBM develop- realistically estimating what those rates might be five years out from
ments. Without such an understanding, the usefulness of research, development is very difficult and uncertain. Also, the type of develop-
and especially application of that research to tangible projects, is ment will have an impact on the economics. For example, vertical
diminished. The modern geologist (both practicing and academic) wells tend to be cheaper but may have lower flow rates than horizontal
should exercise his or her multidisciplinary skills and embrace at wells (when a one to one comparison is made in the same formation);
least some understanding on the economies of CBM. horizontal wells (generally selected to overcome low permeability)
As an example, the current slow down in development of CBM pros- are expensive and require high gas content coals to be economic.
pects in the USA is directly linked to low gas price, but the same gas Although an attractive option in many cases because of normally higher
price may be found in other countries (especially, though not limited flow rates, horizontal wells cannot always justify their cost (Ayers,
to, developing countries), yet in those areas (think China, India and 2002). Estimating well costs are extremely time dependent and quite
Indonesia, among others) there are vibrant exploration programmes. influenced by location in the world. In a mature basin drilling costs for
In Australia, where gas prices have historically been low, CBM develop- a 500 m vertical well including all bottom completion for a single
ment is still growing (Fig. 5). seam might be as little as $US200,000 but in other places may cost
The bulk of this paper has dealt with how a CBM reservoir works over $US550,000 (without any stimulation). A horizontal well with
and Section 6 discussed what data are needed to assess those proper- two 1000 m laterals can be in the range of $US1.8 million to $US3.5
ties in an exploration programme. The quality and quantity of those million depending on geographic location and stability of the formation.
data directly impact CBM prospectively. In the science of statistics it All sets of costs include down hole equipment.
is vital to know what parameters are most important to an outcome; When a field is put into production, and history matching of gas pro-
in CBM development, it is also vital to know how the data are collected duction profiles is possible, more realistic assessments of the total
impacts on the viability of prospect. This knowledge illuminates and reserve can be made than in the exploration or pilot production phases.
informs and thus is important to discuss. In this section, the major However, during exploration when only gas flow models are available,
influences on the viability of a CBM project are discussed and this di- even slight changes into the input parameters can have major conse-
rectly and importantly links to all other previous sections of this paper. quences on both the predicted peak rate (and timing) and well reserve
(see Section 7.1). Like gas price, there is little a company can do to
10.1. Project viability change reservoir character: it is what it is. The difference comes when
imaginative and innovative engineering solutions are used to coax the
In terms of viability, CBM plays are more like mining projects than sometimes reluctant gas from the coal bed.
conventional petroleum prospects (Atkins, 2003). Typically, there are Finally, understanding sensitivities on a prospect's viability is essen-
five stages that a CBM project may go through: tial both from the technical as well as the economic sense. Table 2 gives
some examples showing the sensitivities of gas price, costs, peak flow
1. Initial identification of a prospect through a desk top study, rate and well reserve. These are meant only for illustrative purposes
2. Exploration drilling and sampling (see Section 8), and it is acknowledged that different modellers would come out with
3. Pilot wells, different, though probably not widely different, results. One issue not
4. Development drilling ahead of full commercialisation, and discussed but should not be underestimated, is the influence that
5. Full production with further drilling to maintain gas deliverability.
Table 2
The viability of a CBM project will constantly be re-evaluated Sensitivity around selected variables and effect on NPV and %IRR.
throughout this stream of work, but usually the decision to go to
Variable NPV $US millions %IRR
full development will occur after stage 3.
There are many parameters used in the assessment of the viability Gas price ($US/GJ)
$6.50 $80.6 20.3%
of a CBM project, both technical and economic. In an economic sense
$7.00 $104.3 22.9%
the two most widely used parameters are net present value (NPV) $8.00 $151.4 27.7%
and percentage of internal rate of return (%IRR). The NPV predicts if
the project will make money over its entire life, taking into account Peak flow rate (MCFD)
inflation, changes in gas price as well as capital and operational costs. 550 $77.5 18.2%
650 $96.6 21.3%
In theory, any project with a NPV above zero is worthwhile—but of
800 $116.0 25.5%
course, in a capital constrained environment where multiple projects
for investment are available, the logical choice is the one with the Well reserve (BCF)
highest NPV. Simplistically, the %IRR (also called ‘discounted cash 0.80 $53.5 19.9%
1.00 $104.3 22.9%
flow’ or ‘rate of return’ [ROR]) is the amount of profit over and above
1.50 $165.2 28.5%
capital investment.
T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81 73

differing tax regimes have on the viability and ultimately the impact of coal beds needs to be established along with some rank and coal quality
CBM development (Godfrey et al., 2010). data. Within the prospective category, levels of uncertainty can also be
defined through low, best and high estimates. These can be done either
10.2. From resource to reserve certification deterministically (i.e. just straight calculations) or probabilistically (see
Moore and Friederich, 2010; Moore et al., 2009). Prospective resources
A second area where the types, quality and quantity of data have may be quite a bit less than the GIP depending on the local geological
an impact on CBM viability is in the certification of resources and conditions, level and standard of data as well as the generosity of the
reserves. This is a case where science meets—on an equal basis—the certification company. For example, it might be determined that a
economics of a prospect. It behoves the scientist to understand how certain proportion of the GIP is too deep and thus “unrecoverable” (in
the certification process works as much as it behoves the financial the current technological and economic environment).
analysts to understand the technical aspects of a CBM reservoir. It is common to report GIP, prospective and contingent resources
The value in a CBM resource lies not only with selling gas, but also on a probabilistic basis. Three numbers are often cited as: P10, P50
with determining the quantity of energy potentially available for and P90. In all mathematical and statistical realms P10 is meant to
development which is then read as value. A resource or reserve booking denote, “there is only a 10% chance of a value being less than the number
is like having a house of value ‘X’ that you are able to borrow against. In given” and P90 means “there is a 90% chance that a number will be less
a publically held company, the certification can be lodged with the stock than the number given”. However, in the PRMS it is important to note
exchange with which the company is listed and this acts essentially as that this is reversed, for example: P90 becomes the low estimate, and
part of the valuation of the company, though this valuation is quite is defined as “there is a 90% chance that a value will be higher than
‘liquid’ and at any particular point in time will be based on the mood the number given”. P50 has the same meaning in both the statistical
of the market and local and world economies. This discussion will and PRMS systems, which is “there is a 50% chance that a number
focus on certifiable resources and reserves; methods for gas‐in-place could be higher or lower than the number given”.
(GIP) assessments are covered by Boyer and Qingzhao (1998), Mavor Contingent resources require all the data that prospective resources
and Nelson (1997), Meneley et al. (2003), Moore and Friederich need plus gas data and permeability measurements from the reservoirs
(2010), Moore et al. (2009) and Scott et al. (1995). within the permit area. Gas desorption and adsorption data (for calcula-
As might be imagined, because of the potentially huge financial tion of gas saturation) are required before most certification companies
implications associated with resource/reserve certification, the proce- will give any bookings at the contingent level. Exceptions to this might
dures are tightly defined. The Society of Petroleum Engineers' (SPE) be made if there are local mines near or in the permit area that have
work on defining and standardising how to assess CBM reserves and known and documented gas shows or if adjacent permits are already
resources is largely accepted by most CBM and certification compa- producing gas, demonstrating that the reservoir is likely to be produc-
nies as well as stock exchanges. The SPE system is detailed in the tive. The volume of certified contingent resources is also likely to be
Petroleum Resources Management System (PRMS) and a discussion substantially less than prospective resources depending on the level
is given in Anonymous (2011) and Barker (2008). Essentially, there and amount of data as well as the complexity of the geological condi-
are three resource/reserve categories that represent levels of certainty tions in any particular basin. In addition, gas flow models will also be
and commerciality (Fig. 57). From least to most certain and commer- used to help estimate what the recovery factor may be for individual
cial the categories are: reservoirs.
In order to achieve certification at the reserve level, commercial
1. Prospective resources
flow rates must be demonstrated. The amount of reserves certified
2. Contingent resources
can be dependent on several factors including the extent of wells
3. Reserves
with gas data throughout the permit, number of permeability tests
Prospective resources, being the least certain, require the least conducted, and demonstrable extent of the reservoir, among other
amount of data to substantiate them. Usually, just the presence of factors. Within the reserves classification, different levels of uncer-
tainty can be indicated. For example, a “proven reserve” designates
the drainage area of one well (that is, the producing well) and has
the highest level of certainty; “probable reserves” usually radiate out-
ward in equal directions either two to three offsets from the proven
reserve area (note that an ‘offset’ is equal to the gas drainage area
of one well); finally “possible reserves” have the least certainty and
are the most distant from the producing well, possibly extending out-
wards to seven offsets or more from the well with the proven reserves.
The size of the assignment of probable and possible reserves is usually
at the discretion of the certification company. In a less mature basin
where reservoir behaviour has a higher degree of uncertainty a certifi-
cation company will be more conservative in their estimates of reserves.
In more mature fields, where a number of operators have producing
wells, reserves might be booked without a flowing well directly associ-
ated with the permit under certification.
Note that the term 1P is equal to just the proven reserve, whilst 2P
is equal to the proven plus probable reserves and 3P is equal to the
proven plus probable plus possible reserves. A usual requirement in
obtaining a reserve booking is the ability to demonstrate the existence
of a gas market, either through an existing gas sales agreement at the
time of certification or through a convincing argument that shows that
there is an unequivocal need for gas in the immediate region and a
means to deliver it. Hypothetically, all contingent resources should
or could convert into reserves but this is rarely the case as additional
Fig. 57. Resource and reserve classification (see Anonymous, 2011; Barker, 2008). data gathered from pilot production wells has the usual effect of
74 T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81

betraying the natural enthusiasm of reservoir engineers and hypothetical permit may provide to a company. It is important to
geologists. note that the values used in Table 3 are for illustrative purposes
As might be gathered from the discussion above, well placement, in only. The range in value in this example is from $23.5 to $400 million
a geographic sense within a permit, can be hugely relevant to booking of with a best case of $119 million. The values given are indicative of a
reserves. As an example, Fig. 58a shows the position of three strati- development where two or three core holes and one to three pilot
graphic wells confirming the presence of coal. In many cases these wells have been drilled. In this scenario, given the present cost of
wells might have existed prior to the presence of a CBM permit. Addi- wells, even the low end of market capitalisation would return good
tional drilling of wells specifically for gas and permeability testing has value for the drilling investment.
the effect of converting the prospective resources into contingent Finally, it is worthwhile to note that the impact of most CBM projects
(Fig. 58b). Conversion to reserves can be a bit trickier; as previously is assessed based on the 2P and 3P reserves (as opposed to conventional
stated, a flowing well is almost always needed. Also, the positioning of gas projects where 1P is the primary assessor). An example illustrating
the pilot within the permit can also be important. Pilot wells, if possible, the level of 2P reserves and market valuation is illustrated by the
should be placed so that offset assignment of probable and possible growth of Arrow Energy Ltd. (Australia) which rose from zero reserves
reserves is within the permit boundaries. Strategically placed observa- and a market valuation of $AU20 million in the year 2000 to 719 BCF of
tion wells completed in the target seam(s) can record pressure changes 2P reserves and a market valuation of $AU1.7 billion in 2007 to over
from the other wells and thus confirm connectivity (which can increase 6000 BCF of 2P reserves and a market value of $AU3.5 billion in 2010.
the amount of reserves) (Fig. 58c). What is not desirable is when, for
example, a pilot well is situated near a permit boundary and some of 11. Summary
the reserve/resource bookings would then be ‘lost’ (Fig. 58d).
The impact of maximising a CBM resource and reserve booking The presence of methane in coal seams has long been recognised,
can be huge. As an example, Table 3 gives scenarios regarding the first as a hazard and but more lately as an economically harvestable
value that different levels of reserves and resources from the same commodity. Early exploitation of CBM viewed the reservoir in much

Fig. 58. One example (of a possible many) of how well type and placement will determine certification of resources and reserves. (A) Three stratigraphic drill holes establish there
are coals and thus prospective resources may be booked. (B) Wells specifically drilled to test for gas volume (and to calculate gas saturation) and permeability allow contingent
resources to be booked over some of the permit. (C) Pilot well which demonstrates commercial flow allows reserves to be booked. Observation well establishes connectivity of
the reservoir and allows probable reserves to be extended. (D) If the pilot well is located too close to the permit boundary, then the full potential of booked reserves may not
be realised and could actually benefit the neighbouring permit.
T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81 75

Table 3
Hypothetical valuation of certified reserves and resources; shown for illustrative purposes only.

Category Certified (BCF) $US value/GJ energy Unrisked estimated market valuation

Low Best High Low Best High

Proven 2 $2.00 $4.00 $5.00 $4,000,000 $8,000,000 $10,000,000


Probable 30 $0.20 $0.70 $3.00 $6,000,000 $21,000,000 $90,000,000
Possible 150 $0.05 $0.40 $1.40 $7,500,000 $60,000,000 $210,000,000
Contingent 300 $0.02 $0.10 $0.30 $6,000,000 $30,000,000 $90,000,000
Total $23,500,000 $119,000,000 $400,000,000

GJ = gigajoule.

the same manner as a clastic (or limestone) gas reservoir would be relationships are never universal from basin to basin or even seam
viewed. This proved a mistake as coal, being overwhelmingly carbon to seam.
based, behaves in a fundamentally different way than clastic gas • Characterisation of the reservoir in terms of gas‐in-place, permeability
reservoirs. This review has only scratched at the surface of CBM fun- and gas saturation is crucial to insure both a successful pilot
damentals; however, the following crucial points have been covered: programme as well as resource certification.
• There will be some level of uncertainty in all the aspects of the char-
• Significant commercial quantities of CBM are being extracted within acterisation of a CBM reservoir and this uncertainty needs to be
several countries throughout the world, including the USA, Australia, recognised as well as quantified. Ideally, this quantification of uncer-
Canada, China and India. Many more countries (Indonesia, the UK, tainty will be conducted throughout all decision-making processes.
Russia etc.) are actively exploring and developing their CBM resources • Often the impact of a CBM prospect can be obtained prior to gas sales
and are likely to have saleable gas in the next few years. through certification of its resources and/or reserves. Maximisation of
• Initial CBM developments were in higher rank coals; understandably the impact of a CBM project is dependent on data type, quality and
these coals can generate and contain more gas, returning a higher quantity. In addition, the importance of the spatial distribution of
profit margin. Gas from these coals is termed thermogenic because those data within the permit should not be underestimated.
methane is generated after a certain rank is obtained (about 0.5–
0.6% vitrinite reflectance).
• Since the early 1990s it has been recognised that ‘under mature’ (i.e. 12. Useful terms and conversions
b0.5% vitrinite reflectance) coals can also contain commercial quanti-
ties of gas. These gases are generated biogenically through microbial
processes. Coals of subbituminous rank are the primary target for
biogenic CBM development (as opposed to lignite); although the vol- SCF standard cubic feet
umes of gas are commonly lower than thermogenically-derived CBM MCF thousand standard cubic feet
reservoirs, if the low-rank coals are thick and not too deep, the eco- MMCF million standard cubic feet
nomics may allow commercialisation. BCF billion standard cubic feet
• The fundamental measures needed to characterise a reservoir, regard- TCF trillion standard cubic feet
less of rank or gas type (i.e. biogenic or thermogenic), are gas quality MCFD thousand standard cubic feet per day
(i.e. the amount of CH4, CO2 etc.), the gas charge (i.e. desorbed gas), GJ gigajoule
the maximum gas holding capacity (derived from the adsorption iso- Bbls barrels
therm) and the permeability. mD millidarcies
• Permeability is by far the most important attribute controlling gas nm nanometer
flow in a coal seam reservoir, and is influenced by depth, stress Å angstrom
regime within the basin and also, fundamentally, the organic compo- 1m 3.3 ft
sition of the coal. 1 m3 35.31 SCF
• The second most important attribute is the level of gas saturation, 1 MCF 28.3168 m 3
which is determined through desorption and adsorption measure- 1 MMCF 28,316.8 m 3
ments. Coal reservoirs can be under saturated for a number of reasons 1 Bbls 158.98 l
including uplift after deep burial (where holding capacity and gas 1 Bbls 0.158987 m 3
charge are decreased but with uplift holding capacity is increased 1 MCF CH4 1.05 GJ
again but the gas charge is not), gas escape if the reservoir is close to 1 BCF CH4 1.05 PJ
surface or through removal of gas in a hydrodynamically active system. 1 t of CO2 556.2 m 3
• Coals of both high and low-rank my have secondary gas recharge if 1m 1 billon nm
reservoir conditions are conducive and this will have the net effect 1 nm 10 Å
of increasing gas saturation.
• Pores are the defining feature of a coal reservoir, but unlike conven-
tional gas plays where pore volume determines a rock's ability to Acknowledgements
hold gas, in coal it is the pore surface area (because methane is (pri-
marily) adsorbed onto the surface of the coal). The pore surface area More than a few people have helped to gain information and/or
is directly related to the size distribution of pores, with most methane made valuable comments at the right time and place. These colleagues
being held in micropores. are, in no particular order: Romeo M. Flores, Ray Williams, Greg
• Rank influences many attributes of a coal reservoir including gas type, Twombly, Steve Stepanek, Peter Warwick, Lesile ‘Jingle’ Ruppert,
gas quality, pore distribution, maximum gas holding capacity and per- Hongmei Li, R. Marc Bustin, John Hattner, Nathan Shahan, Michael H.
meability. Trippi, Sudhansu Adhikari, Chris M. Nelson, Nick Davies and Marina
• Coal composition can also have a significant influence on reservoir S. Manning (for the musical similes). Also, thanks to two of my students,
properties, but because of the highly variable nature of coal, these Tennille E. Mares and Carol I. Butland, who inevitably taught me more
76 T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81

than I them. Both Sadiq J. Zarrouk and Xingjin Wang have been my Bertard, C., Bruyet, B., Gunther, J., 1970. Determination of desorbable gas concentration
coal (direct method). International Journal of Rock Mechanics and Mining Science
teachers on all things gas modelling and have supplied some of the 7, 43–65.
figures. Thanks to the dynamic duo of Joan S. Esterle, who keeps me Bibler, C.J., Marshall, J.S., Pilcher, R.C., 1998. Status of worldwide coal mine methane
grounded in real-world data and Catherine R. Moore, who reminds emissions and use. International Journal of Coal Geology 35, 283–310.
Bodden, W.P., Ehrlich, R., 1998. Permeability of coals and characteristics of desorption
me of the uncertainty in all things measured. Appreciation goes to tests: implications for coalbed methane production. International Journal of Coal
Luca Technology for providing photomicrographs of methanogen Geology 35, 333–347.
microbes. Arrow Energy Ltd., Dart Energy Ltd. and Ephindo Energy Boyer II, C.M., Qingzhao, B., 1998. Methodology of coalbed methane resource assess-
ment. International Journal of Coal Geology 35, 349–368.
Ltd. are also acknowledged because much of this paper was written Boyer II, C.M., Kelafant, J.R., Kuuskraa, V.A., Manger, K.C., 1990. Methane Emissions
whilst in the service of these three companies; they not only provided from Coal Mining: Issues and Opportunities for Reduction. US EPA, Washington
me with intellectual stimulus and practical exploration and production DC. EPA/400/9-90/008: 86 pp.
Bratvold, R.B., Begg, S.H., 2008. I would rather be vaguely right than precisely wrong: a
experience, but also allowed for many happy hours on the plane, unen-
new approach to decision making in the petroleum exploration and production in-
cumbered by phone calls, emails and meetings. Enormous thanks to dustry. AAPG Bulletin 92, 1373–1392.
Jane C. Shearer who edited an early version of the paper. Thanks to Bratvold, R.B., Begg, S.H., Campbell, M., 2002. Would you know a good decision if you
Özgen Karacan for his patient and careful inputs and to Timothy saw one? SPE Paper 77509, Society of Petroleum Engineers Annual Technical Con-
ference and Exhibition, San Antonio, Texas. 11 pp.
Horscroft for initiating the paper. Finally, two anonymous reviewers Bromhal, G.S., Sams, W.N., Jikich, S., Ertekin, T., Smith, D.H., 2005. Simulation of CO2 se-
helped to improve the manuscript. questration in coal beds: the effects of sorption isotherms. Chemical Geology 217,
201–211.
Busch, A., Gensterblum, Y., 2011. CBM and CO2-ECBM related sorption processes in
coal: a review. International Journal of Coal Geology 87, 49–71.
References Bustin, R.M., 1997. Importance of fabric and composition on the stress sensitivity of
permeability in some coals, Northern Sydney Basin, Australia: relevance to coalbed
Adeboye, O.O., Bustin, R.M., in press. Variation of gas flow properties in coal with probe methane exploitation. AAPG Bulletin 81, 1894–1908.
gas, composition and fabric: examples from Western Canadian Sedimentary Basin. Bustin, R.M., 2004. Acid gas sorption by British Columbia coals: implications for perma-
International Journal of Coal Geology, http://dx.doi.org/10.1016/j.coal.2011.06.015. nent disposal of acid gas in deep coal seams and possible co-production of meth-
ALL-Consulting, 2003. Handbook on Coal Bed Methane Produced Water: Management ane. Final report funding agreement 2000‐16. University of British Columbia,
and Beneficial Use Alternatives. Ground Water Protection Research Foundation, Vancouver. 126 pp.
U.S. Department of Energy, National Petroleum Technology Office, Bureau of Bustin, A.M.M., Bustin, R.M., 2008. Coal reservoir saturation: impact of temperature and
Land Management, Washington DC. 322 pp. pressure. AAPG Bulletin 92, 77–86.
American Standard Testing Material (ASTM), 2010. Determination of Gas Content of Coal Bustin, R.M., Clarkson, C.R., 1998. Geological controls on coalbed methane reservoir ca-
—Direct Desorption Method, D7569-10. ASTM International, West Conshohocken, PA. pacity and gas content. International Journal of Coal Geology 38, 3–26.
12 pp. Bustin, R.M., Ross, J.V., Rouzaud, J.N., 1995. Mechanisms of graphite formation from
Amijaya, H., Littke, R., 2006. Properties of thermally metamorphosed coal from the kerogen: experimental evidence. International Journal of Coal Geology 28, 1–36.
Tanjung Enim Area, South Sumatra Basin, Indonesia with special reference to the Butala, S.J., Medina, J.C., Taylor, T.Q., Bartholomew, C.H., Lee, M.L., 2000. Mechanisms
coalification path of macerals. International Journal of Coal Geology 66, 271–295. and kinetics of reactions leading to natural gas formation during coal maturation.
Anderson, J.A.R., 1964. The structure and development of the peat swamps of Sarawak Energy & Fuels 14, 235–259.
and Brunei. Journal of Tropical Geography 18, 7–16. Butland, C.I., Moore, T.A., 2008. Secondary biogenic coal seam gas reservoirs in New
Anderson, J.A.R., 1983. The tropical peat swamps of western Malaysia. In: Gore, A.J.P. Zealand: a preliminary assessment of gas contents. International Journal of Coal
(Ed.), Ecosystems of the World 4B, Mires, Swamp, Bog, Fen and Moor. Elsevier, Am- Geology 76, 151–165.
sterdam, The Netherlands, pp. 181–199. Cameron, M.J., 1995. Cleat Structure, Waikato Coal Region. Department of Geology.
Anderson, R.B., Hall, W.K., Lecky, J.A., Stein, K.C., 1956. Sorption studies on American University of Auckland, Auckland, p. 101.
coals. The Journal of Physical Chemistry 60, 1548–1558. Cameron, C.C., Esterle, J.S., Palmer, C.A., 1990. The geology, botany and chemistry of se-
Anna, L.O., 2003. Groundwater flow associated with coalbed gas production, Ferron lected peat-forming environments from temperate and tropical latitudes. Interna-
Sandstone, east-central Utah. International Journal of Coal Geology 56, 69–95. tional Journal of Coal Geology 12, 105–156.
Anonymous, 1999. Guide to the Determination of Gas Content of Coal—Direct Desorp- Carlson, F.M., Hartman, R.C., Obluda, J.C., Miller, M.T., 2008. Field validation of new
tion Method. Standards Association of Australia. AS 3980–1999, 33 pp. methodology for resource assessment of undersaturated CBNG reservoirs. Interna-
Anonymous, 2011. Guidelines for Application of the Petroleum Resources Management tional Coalbed & Shale Gas Symposium. University of Alabama, Tuscaloosa, Ala-
System. Society of Petroleum Engineers. 221 pp. bama. Paper 819, unpaginated.
Aravena, R., Harrison, S.M., Barker, J.F., Abercrombie, H., Rudolph, D., 2003. Origin of Ceglarsk-Stefańska, G., Brzóska, K., 1998. The effect of coal metamorphism on methane
methane in the Elk Valley coalfield, southeastern British Columbia, Canada. Chem- desorption. Fuel 77, 645–648.
ical Geology 195, 219–227. Ceglarsk-Stefańska, G., Zarębska, K., 2002. The competitive sorption of CO2 and CH4
Atkins, B., 2003. Coal Bed Methane—From Resource to Reserves. Focus. Gaffney, Cline & with regard to the release of methane from coal. Fuel Processing Technology
Associates, pp. 1–2. 77–78, 423–429.
Ayers Jr., W.B., 2002. Coalbed gas systems, resources, and production and a review of Chalmers, G.R.L., Bustin, R.M., 2007. On the effects of petrographic composition on
contrasting cases from the San Juan and Powder River basins. AAPG Bulletin 86, coalbed methane sorption. International Journal of Coal Geology 69, 288–304.
1853–1890. Chalmers, S., Kovse, A., Stark, P., Facer, L., Smith, N., 2010. Treatment of coal seam gas
Bachu, S., Brulotte, M., Grobe, M., Steward, S., 2000. Suitability of the Alberta subsurface water. Water—Journal of Australian Water Association 37, 28–33.
for carbon-dioxide sequestration in geological media. Alberta Geological Survey, Cheung, K., Sanei, H., Klassen, P., Mayer, B., Goodarzi, F., 2009. Produced fluids and shal-
Earth Sciences Report 00‐11. 86 pp. low groundwater in coalbed methane (CBM) producing regions of Alberta, Canada:
Balan, H.O., Gumrah, F., 2009. Assessment of shrinkage-swelling influences in coal trace element and rare earth element geochemistry. International Journal of Coal
seams using rank-dependent physical coal properties. International Journal of Geology 77, 338–349.
Coal Geology 77, 203–213. Clarkson, C.R., 2003. Application of a new multicomponent gas adsorption model to gas
Barker, G.J., 2008. Application of the PRMS to coal seam gas. 2008 Society of Petroleum adsorption systems. Society of Petroleum Engineers Journal 8, 236–250 SPE-
Engineers Asia Pacific Oil & Gas Conference and Exhibition, SPE Paper 117124, 78146-PA. doi: 78110.72118/78146-PA.
Perth, Australia. 13 pp. Clarkson, C.R., Bustin, R.M., 1996. Variation in micropore capacity and size distribution
Barker, C.E., Dallegge, T.A., 2005. Zero-headspace coal-core gas desorption canister, re- with composition in bituminous coal of the Western Canadian Sedimentary Basin.
vised desorption data analysis spreadsheets and a dry canister heating system. U.S. Fuel 75, 1483–1498.
Geological Survey Open File Report 2005‐1177, Reston, Va. 9 pp. Clarkson, C.R., Bustin, R.M., 1999. The effect of pore structure and gas pressure upon the
Barker, C.E., Dallegge, T.A., Clark, A.C., 2002. USGS coal desorption equipment and a transport properties of coal: a laboratory modeling study. 2. Adsorption rate
spreadsheet for analysis of lost and total gas from canister desorption measure- modeling. Fuel 78, 1345–1362.
ments. U.S. Geological Survey Open-File Report 02‐496, Reston, Va. 13 pp. Clarkson, C.R., Bustin, R.M., 2011. Coalbed methane: current field-based evaluation
Beamish, B.B., Crosdale, P.J., 1998. Instantaneous outbursts in underground coal mines: methods. SPE Reservoir Evaluation and Engineering 14, 60–75.
an overview and association with coal type. International Journal of Coal Geology Clayton, J.L., 1998. Geochemistry of coalbed gas—a review. International Journal of Coal
35, 27–55. Geology 35, 159–173.
Behar, F., Vandenbroucke, M., Teermann, S.C., Hatcher, P.G., Leblond, C., Lerat, O., 1995. Close, J.C., 1993. Natural fractures in coal. In: Law, B.E., Rice, D.D. (Eds.), Hydrocarbons
Experimental simulation of gas generation from coals and a marine kerogen. from Coal. American Association of Petroleum Geologists, Tulsa, Oklahoma, pp.
Chemical Geology 126, 247–260. 119–132.
Bell, J.S., 2006. In-situ stress and coal bed methane potential in Western Canada. Bulle- Clymo, R.S., 1984. The limits to peat bog growth. Philosophical Transactions of the
tin of Canadian Petroleum Geology 54, 197–220. Royal Society of London, Series B 303, 605–654.
Bergquist, E., Evanelista, P., Stohlgren, T.J., Alley, N., 2007. Invasive species and coal bed Clymo, R.S., 1987. Rainwater-fed peat as a precursor of coal. In: Scott, A.C. (Ed.), Coal
methane development in the Powder River Basin, Wyoming. Environmental Mon- and Coal-bearing Strata: Recent Advances. Geological Society Special Publication,
itoring and Assessment 128, 381–394. London, U.K., pp. 17–23.
T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81 77

Cohen, A.D., 1984. The Okefenokee Swamp: a low sulphur end-member of a shoreline- Faiz, M., Stalker, L., Sherwood, N., Saghafi, A., Wold, M., Barclay, S., Choudhury, J.,
related depositional model for coastal plain coals. In: Rahmani, R.A., Flores, R.M. Barker, W., Wang, I., 2003. Bio-enhancement of coal bed methane resources in
(Eds.), Sedimentology of Coal and Coal-bearing Sequences. Special Publication of the southern Sydney Basin. Australian Petroleum Production and exploration Asso-
the International Association of Sedimentologists, Oxford, pp. 231–240. ciation 43, 595–610.
Collett, T.S., Barker, C.E., 2003. Coalbed methane in the Ferron coals. International Jour- Faiz, M.M., Saghafi, A., Barclay, S.A., Stalker, L., Sherwood, N.R., Whitford, D.J., 2007a.
nal of Coal Geology 56, 1–202. Evaluating geological sequestration of CO2 in bituminous coals: the southern Syd-
Collinson, M.E., Scott, A.C., 1987. Implications of vegetational change through the geo- ney Basin, Australia as a natural analogue. International Journal of Greenhouse Gas
logical record on models for coal-forming environments. In: Scott, A.C. (Ed.), Coal Control 1, 223–235.
and Coal-bearing Strata: Recent Advances. Geological Society Special Publication, Faiz, M., Saghafi, A., Sherwood, N., Wang, I., 2007b. The influence of petrological prop-
London, pp. 67–85. erties and burial history on coal seam methane reservoir characterisation, Sydney
Cooper, J.R., Crelling, J.C., Rimmer, S.M., Whittington, A.G., 2007. Coal metamorphism Basin, Australia. International Journal of Coal Geology 70, 193–208.
by igneous intrusion in the Raton Basin, CO and NM: implications for generation Faraj, B., Hatch, H., 2004. Mechanism of hydrogen generation in coalbed methane de-
of volatiles. International Journal of Coal Geology 71, 15–27. sorption canisters: causes and remedies. Gastips 10, 15–19.
Creedy, D.P., Pritchard, F.W., 1983. Nitrogen and carbon dioxide occurrence in UK coal Ferm, J.C., Staub, J.R., 1984. Depositional controls of mineable coal bodies. Special Pub-
seams. International Journal of Mining Engineering 1, 71–77. lication of International Association of Sedimentologists 7, 275–289.
Crosdale, P.J., Beamish, B.B., Valix, M., 1998. Coalbed methane sorption related to coal Ferm, J.C., Staub, J.R., Baganz, B.P., Clark, W.J., Galloway, M.C.C.B., Hohos, E.F., Jones Jr., T.L.,
composition. International Journal of Coal Geology 35, 147–158. Mathew, D., Pedlow III, G.W., Robinson, M.J., 1979. The shape of coal bodies. In: Ferm,
Crosdale, P.C., Saghafi, A., Williams, R., Yurakov, E., 2005. Inter-laboratory comparative J.C., Horne, J.C. (Eds.), Carboniferous Depositional Environments in the Appalachian
CH4 isotherm measurement on Australian coals. In: Beeston, J.W. (Ed.), Bowen Region. University of South Carolina, Columbia, South Carolina, pp. 605–619.
Basin Symposium 2005. The Geological society of Australia, Yeppon, Queensland, Fick, A., 1855. On liquid diffusion. Philosophical Magazine 10, 30–39.
Australia, pp. 273–277. Fielding, C.R., 1984. ‘S’ or ‘Z’ shaped coal seam splits in the coal measures of County
Crosdale, P.J., Moore, T.A., Mares, T.E., 2008. Influence of moisture content and tempera- Durham. Proceedings of the Yorkshire Geological Society 45, 85–89.
ture on methane adsorption isotherm analysis for coals from a low-rank, Flores, R.M., 1993. Coal-bed and related depositional environments in methane gas-
biogenically-sourced gas reservoir. International Journal of Coal Geology 76, 166–174. producing sequences. In: Law, B.E., Rice, D.D. (Eds.), Hydrocarbons from Coal.
Cui, X., Bustin, M., Dipple, G., 2004. Differential transport of CO2 and CH4 in coalbed American Association of Petroleum Geologists, Tulsa, Oklahoma, pp. 13–37.
aquifers: implications for coalbed gas distribution and composition. International Flores, R.M., 1998. Coalbed methane: from hazard to resource. International Journal of
Journal of Coal Geology 88, 1149–1161. Coal Geology 35, 1–379.
Dahm, K.G., Guerra, K.L., Xu, P., Drewes, J.E., 2011. A composite geochemical database Flores, R.M., 2008. Microbes, methanogenesis, and microbial gas in coal. International
for coalbed methane produced water quality in the Rocky Mountain Region. Envi- Journal of Coal Geology 76, 1–185.
ronmental Science and Technology 45, 7655–7663. Flores, R.M., Hanley, J.H., 1984. Anastomosed and associated coal-bearing fluvial de-
Damen, K., Faaij, A., Bergen, v.-F., Gale, J., Lysen, E., 2005. Identification of early oppor- posits: Tongue River Member, Paleocene Fort Union Formation, northern Powder
tunities for CO2 sequestration—worldwide screening for CO2-EOR and CO2-ECBM River Basin, Wyoming. In: Rahmani, R.A., Flores, R.M. (Eds.), Sedimentology of
projects. Energy 30, 1931–1952. Coal and Coal-bearing Sequences. Special Publication of the International Associa-
Dart Energy Ltd., 2011. First gas in Sangatta West. http://www.dartenergy.com.au/page/ tion of Sedimentologists, Oxford, pp. 85–103.
Investor_Relations/ASX_Announcements; 1 April 2011, Brisbane, 3 pp. Flores, R.M., McGarry, D.E., Stricker, G.D., 2005. CBNG development: confusing coal
Dart Energy Ltd., 2012. Interim report for the half-year ended 31 December 2011. http:// stratigraphy and gas production in the Powder River Basin. Canadian Society of Pe-
www.dartenergy.com.au/page/Investor_Relations/ASX_Announcements/ 9 March troleum Geologists Gussow Conference, Canmore, Canada, Abstracts with Pro-
2012, Brisbane, 27 pp. grams. unpaginated.
Dawson, G.K.W., Esterle, J.S., 2010. Controls on coal cleat spacing. International Journal Flores, R.M., Rice, C.A., Stricker, G.D., Warden, A., Ellis, M.S., 2008. Methanogenic path-
of Coal Geology 82, 213–218. ways of coal-bed gas in the Powder River Basin, United States: the geological fac-
Day, S., Fry, R., Sakurovs, R., 2008. Swelling of Australian coals in supercritical CO2. In- tor. International Journal of Coal Geology 76, 52–75.
ternational Journal of Coal Geology 74, 41–52. Fujioka, M., Yamaguchi, S., Nako, M., 2010. CO2-ECBM field tests in the Ishikari coal
Dehmer, J., 1995. Petrological and organic geochemical investigation of recent peats basin of Japan. International Journal of Coal Geology 82, 287–298.
with known environments of deposition. International Journal of Coal Geology Fukasawa, H., Sunaryo, R., Napitupulu, P.H., 1987. Hydrocarbon generation and migra-
28, 111–138. tion in the Sangatta area, Kutei Basin, Sixteenth Annual Convention. Proceedings
Deisman, N., Gentzis, T., Chalaturnyk, R.J., 2008. Unconventional geomechanical testing Indonesian Petroleum Association. Indonesian Petroleum Association, Jakarta, In-
on coal for coalbed reservoir well design: the Alberta Foothills and Plains. Interna- donesia, pp. 123–139.
tional Journal of Coal Geology 75, 15–26. Gamson, P., Beamish, B., Johnson, D., 1996. Coal microstructure and secondary miner-
Deisman, N., Mas Ivas, D., Darcel, C., Chalaturnyk, R.J., 2010. Empirical and numerical alization: their effect on methane recovery. In: Gayer, R., Harris, I. (Eds.), Coalbed
approaches for geomechanical characterization of coal seam reservoirs. Interna- Methane and Coal Geology. : Special Publication, 109. Geological Society, London,
tional Journal of Coal Geology 82, 204–212. pp. 165–179.
Diamond, W.P., 1978. Evaluation of the methane gas content of coalbeds; Part of a com- Gan, H., Nandi, S.P., Walker Jr., P.L., 1972. Nature of the porosity in American coals. Fuel
plete coal exploration program for health and safety and resource evaluation. Second 51, 272–277.
International Coal Exploration Symposium, Denver, Co., October 1978. unpaginated. Gayer, R., Harris, I., 1996. Coalbed methane and coal geology. The Geological Society,
Diamond, W.P., 1993. Methane control for underground coal mines. In: Law, B.E., Rice, Special Publication, 109. 343 pp., London.
D.D. (Eds.), Hydrocarbons from Coal. American Association of Petroleum Geolo- Gentzis, T., 2009. Stability analysis of a horizontal coalbed methane well in the Rocky
gists, Tulsa, Oklahoma, pp. 237–267. Mountain Front Ranges of southeast British Columbia, Canada. International Jour-
Diamond, W.P., Levine, J.R., 1981. Direct Method Determination of the Gas Content of nal of Coal Geology 77, 328–337.
Coal: Procedures and Results. U.S. Bureau of Mines, Washington, D.C. 36 pp. Gentzis, T., Deisman, N., Chalaturnyk, R.J., 2009a. Effect of drilling fluids on coal perme-
Diamond, W.P., Schatzel, S.J., 1998. Measuring the gas content of coal: a review. Inter- ability: impact on horizontal wellbore stability. International Journal of Coal Geol-
national Journal of Coal Geology 35, 311–331. ogy 78, 177–191.
Diessel, C.F.K., 1982. An appraisal of coal facies based on maceral characteristics. Aus- Gentzis, T., Deisman, N., Chalaturnyk, R.J., 2009b. A method to predict geomechanical
tralian Coal Geology 4, 474–484. properties and model well stability in horizontal boreholes. International Journal
DiMichele, W.A., Phillips, T.L., Olmstead, R.G., 1987. Opportunistic evolution: abiotic of Coal Geology 78, 149–160.
environmental stress and the fossil record of plants. Review of Palaeobotany and Gilcrease, P.C., Shurr, G.W., 2007. Making microbial methane work: the potential for
Palynology 151, 151–178. new biogenic gas. World Oil 228, 37–48.
Draper, J.J., Boreham, C.J., 2006. Geological controls on exploitable coal seam gas distri- Godfrey, P., Ee, T., Hewitt, T., 2010. Coal bed methane development in Indonesia: gold-
bution in Queensland. APPEA Journal 46, 343–366. en opportunity or impossible dream? Journal of Energy & Natural Resources Law
Durucan, S., Ahsan, M., Shi, J.-Q., 2009. Matrix shrinkage and swelling characteristics of 28, 233–264.
European coals. Energy Procedia 1, 3055–3062. Gold, T., 1992. The deep, hot biosphere. Proceedings of the National Academy of Sci-
Enever, J., Casey, D., Bocking, M., 1999. The role of in-situ stress in coalbed methane ence of the USA 89, 6045.
production. In: Mastalerz, M., Glikson, M., Golding, S.D. (Eds.), Coalbed Methane: Golding, S.D., Rudolph, V., Flores, R.M., 2010. Asia Pacific Coalbed Methane Symposium:
Scientific, Environmental and Economic Evaluation. Kluwer Academic Publishers, selected papers from the 2008 Brisbane symposium on coalbed methane and CO2-
Dordrecht, pp. 297–303. enhanced coalbed methane. International Journal of Coal Geology 82, 1–298.
Esterle, J.S., Ferm, J.C., 1994. Spatial variability in modern tropical peat deposits from Goodman, A.L., Busch, A., Duffy, G.J., Fitzgerald, J.E., Gasem, K.A.M., Gensterblum, Y., Krooss,
Sarawak, Malaysia and Sumatra, Indonesia: analogues for coal. International Jour- B.M., Levy, J., Ozdemir, E., Pan, Z., Robinson Jr., R.L., Schroeder, K., Sudibandriyo, M.,
nal of Coal Geology 26, 1–41. White, C.M., 2004. An inter-laboratory comparison of CO2 isotherms measured on Ar-
Esterle, J.S., Moore, T.A., Shearer, J.C., 1992. Comparison of macroscopic and microscop- gonne premium coal samples. Energy & Fuels 18, 1175–1182.
ic size analyses of organic components in both coal and peat. 26th Newcastle Sym- Gray, I., 1987. Reservoir engineering in coal seams: Part 1. The physical process of gas
posium: Advances in the study of the Sydney Basin, Newcastle, NSW, Australia, storage and movement in coal seams. SPE Reservoir Engineering 2, 28–34.
April 3–5, 1992 143–149. Greb, S.F., DiMichele, W.A., Gastaldo, R.A., 2006. Evolution and importance of wetlands
Esterle, J.S., Williams, R.J., Sliwa, R., Malone, M., 2006. Variability in gas reservoir pa- in earth history. In: Greb, S.F., DiMichele, W.A. (Eds.), Wetlands Through Time.
rameters that impact on emissions estimations for Australian black coals. Final Re- Geological Society of America, Boulder, Co: Special Paper, 399, pp. 1–40.
port ACARP Project C13071, Brisbane. 36 pp. Green, M.S., Flanegan, K.C., Gilcrease, P.C., 2008. Characterisation of a methanogenic
Ettinger, I., Eremin, I., Zimakov, B., Yanovskaya, M., 1966. Natural factors influencing coal consortium enriched from a coalbed methane well in the Powder River Basin,
sorption properties: I. Petrographic and sorption properties of coals. Fuel 45, 267–275. U.S.A. International Journal of Coal Geology 76, 34–45.
78 T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81

Gu, F., Chalaturnyk, R.J., 2006. Numerical simulation of stress and strain due to gas Kim, A.G., 1973. The composition of coalbed gas. Report of Investigations 7762. U.S. Bu-
sorption/desorption and their effects on in situ permeability of coalbeds. Journal reau of Mines, Washington DC. 8 pp.
of Canadian Petroleum Technology 45, 1–11. Kim, A.G., 1977. Estimating methane content of bituminous coalbeds from adsorption
Gunter, W.D., Gentzis, T., Rottenfusser, B.A., Richardson, R.J.H., 1997. Deep coalbed data. Report of Investigation 8245. U.S. Bureau of Mines, Washington DC.
methane in Alberta, Canada: a fuel resource with the potential of zero greenhouse Kim, A.G., Douglas, L.J., 1973. Gases desorbed from five coals of low gas content. Report
gas emissions. Energy Conservation Management 38 (Suppl.), S217–S222. of Investigations 7768. U.S. Bureau of Mines, Washington DC. 9 pp.
Gurba, L.W., Weber, C.R., 2001. Effects of igneous intrusions on coalbed methane po- Kinnon, E.C.P., Esterle, J.S., 2008. Geological Controls on Gas Flow Pathways in Coal
tential, Gunnedah Basin, Australia. International Journal of Coal Geology 46, Seams, 2008 Coalbed and Shale Gas Symposium, Paper 0803. Birmingham,
113–131. Alabama.
Gurba, L.W., Gurba, A., Ward, C., Wood, J., Filipowski, A., Titheridge, D., 2001. The im- Kinnon, E.C.P., Golding, S.D., Boreham, C.J., Baublys, K.A., Esterle, J.S., 2010. Stable iso-
pact of coal properties on gas drainage efficiency. In: Doyle, Moloney (Ed.), Geolog- tope water quality analysis of coal bed methane production gases from the
ical Hazards—The Impact to Mining, pp. 215–220. Bowen Basin, Australia. International Journal of Coal Geology 82, 219–231.
Gürdal, G., Yalçin, M.N., 2001. Pore volume and surface area of the Carboniferous coals Kiyama, T., Nishimoto, S., Fujioka, M., Xue, Z., Ishijima, Y., Pan, Z., Connell, L.D., 2011.
from the Zonguldak basin (NW Turkey) and their variations with rank and maceral Coal swelling strain and permeability change with injecting liquid/supercritical
composition. International Journal of Coal Geology 48, 133–144. CO2 and N2 at stress-constrained conditions. International Journal of Coal Geology
Hackley, P.C., Guevara, E.H., Hentz, T.F., Hook, R.W., 2009. Thermal maturity and organ- 85, 56–64.
ic composition of Pennsylvanian coals and carbonaceous shales, north-central Koenig, R.A., Schraufnagel, R.A., 1987. Application of the slug test in coalbed methane
Texas: implications for coalbed gas potential. International Journal of Coal Geology testing. Society of Petroleum Engineers Reprint Series No. 35, 79–89.
77, 294–309. Korre, A., Shi, J.-Q., Imrie, C., Durucan, S., 2009. Modelling the uncertainty and risks as-
Hadiyanto, Stevens, S., 2005. Coalbed methane prospects in lower rank coals of Indone- sociated with the design and life cycle of CO2 storage in coalbed reservoirs. Energy
sia. In: Prihatmoko, S., Digdowirogo, S., Nas, C., van Leeuwen, T., Widjajanto, H. Procedia 1, 2525–2532.
(Eds.), Indonesian Mineral and Coal Discoveries. Indonesian Association of Geolo- Krelle, A., 2011. Geological Controls on Coal Seam Gas Domains—A Case Study in the
gists, Bandung, Indonesia, pp. 152–162. Moranbah Coal Measures at Saraji East. Department of Earth Sciences. The Univer-
Han, J.-Z., Sang, S.-X., Cheng, Z.-Z., Huang, H.-Z., 2009. Exploitation technology of pres- sity of Queensland, Brisbane.
sure relief coalbed methane in vertical wells in the Huainan coal mining area. Min- Krooss, B.M., van Bergen, F., Gensterblum, Y., Siemons, N., Pagnier, H.J.M., David, P.,
ing Science and Technology 19, 25–30. 2002. High-pressure methane and carbon dioxide adsorption on dry and
Harpalani, S., Chen, G., 1995. Estimation of changes in fracture porosity of coal with gas moisture-equilibrated Pennsylvanian coals. International Journal of Coal Geology
emission. Fuel 74, 1491–1498. 51, 69–92.
Harpalani, S., Chen, G., 1997. Influence of gas production induced volumetric strain on Lama, R.D., Bodziony, J., 1998. Management of outburst in underground coal mines. In-
permeability of coal. Geotechnical and Geological Engineering 15, 303–325. ternational Journal of Coal Geology 35, 83–115.
Harpalani, S., Prusty, B.K., Dutta, P., 2006. Methane/CO2 sorption modeling for coalbed Lama, R.D., Nguyen, V.U., 1987. A model for determination of methane flow parameters
methane production and CO2 Sequestration. Energy & Fuels 20, 1591–1599. in coal from desorption tests. Proceedings of the Twentieth International Sympo-
Havton, S., 2003. Technical note: Oxidation of Coal in Desorption Canisters. sium on the Application of Computers and Mathematics in the Mineral Industries:
Unpublished Petroleum Report PR2948. Ministry of Economic Development, Mining, volume 1, pp. 275–282. SAIMM, Johannesburg.
Crown Minerals, Wellington, New Zealand. 4 pp. Lamarre, R.A., Pope, J., 2007. Critical-gas-content technology provides coalbed-
Healy, R.W., Bartos, T.T., Rice, C.A., McKinley, M.P., Smith, B.D., 2011. Groundwater methane-reservoir data. Journal of Petroleum Technology 59, 108–113.
chemistry near an impoundment for produced water, Powder River Basin, Wyo- Lamberson, M.N., Bustin, R.M., 1993. Coalbed methane characteristics of Gates Forma-
ming, USA. Journal of Hydrology 403, 37–48. tion coals, northeastern British Columbia: effect of maceral composition. AAPG Bul-
Hemza, P., Sivek, M., Jirásek, J., 2009. Factors influencing the methane content of coal letin 77, 2062–2076.
beds in the Czech part of the Upper Silesian Coal Basin, Czech Republic. Interna- Langmuir, I., 1918. The adsorption of gases on plane surfaces of glass, mica and plati-
tional Journal of Coal Geology 79, 29–39. num. Journal of the American Chemical Society 40, 1361–1403.
Hendry, P., Faiz, M., Li, D., Gong, S., Fuentes, D., 2007. Microbial analysis of some New Laubach, S.E., Marrett, R.A., Olson, J.E., Scott, A.R., 1998. Characteristics and origins of
Zealand coal samples. CSIRO, Sydney, Australia, Report to Solid Energy NZ Ltd. 69 pp. coal cleat: a review. International Journal of Coal Geology 35, 175–207.
Herudiyanto, 2006. Possible explanation of the increasing coal rank at Sangatta. Inter- Law, B.E., 1993. The relationship between coal rank and cleat spacing: implications for
nal Report. Geological Resource Centre, Geological Agency of Indonesia, Bandung. the prediction of permeability in coal. 1993 International Coalbed Methane Sym-
16 pp. posium, The University of Alabama, Tuscaloosa, Alabama (USA), pp. 435–441.
Hildenbrand, A., Krooss, B.M., Busch, A., Gaschnitz, R., 2006. Evolution of methane sorp- Law, B.E., Rice, D.D., 1993. Hydrocarbons in coal. American Association of Petroleum
tion capacity of coal seams as a function of burial history—a case study from the Geologists. AAPG Studies in Geology 38 394 pp., Tulsa, Oklahoma.
Campine Basin, NE Belgium. International Journal of Coal Geology 66, 179–203. Law, D.H.-S., van de Meer, L.G.H., Gunter, W.D., 2005. Comparison of numerical simula-
Hornibrook, E.R.C., Longstaffe, F.J., Fyfe, W.S., 1997. Spatial distribution of microbial tors for greenhouse gas storage in coalbeds, Part IV: history match of field micro-
methane production pathways in temperate zone wetland soils: stable carbon pilot test data. In: Wilson, M., Morris, T., Gale, J., Thambimuthu, K. (Eds.), Green-
and hydrogen isotope evidence. Geochimica et Cosmochimica Acta 61, 745–753. house Gas Control Technologies. Elsevier Ltd., pp. 2239–2242.
Hunt, J.M., 1979. Petroleum Geochemistry and Geology. Freeman & Co., San Francisco. Laxminarayana, C., Crosdale, P.J., 1999. Role of coal type and rank on methane sorption char-
617 pp. acters of Bowen Basin, Australia coals. International Journal of Coal Geology 40, 309–325.
Hunt, J.M., 1991. Generation of gas and oil from coal and other terrestrial organic mat- Laxminarayana, C., Crosdale, P.J., 2002. Controls on methane sorption capacity of Indian
ter. Organic Geochemistry 17, 673–680. coals. AAPG Bulletin 86, 201–212.
ICCP, 1998. The new vitrinite classification (ICCP System 1994). Fuel 77, 349–358. Levine, J.R., 1993. Coalification: the evolution of coal as source rock and reservoir rock
ICCP, 2001. The new inertinite classification (ICCP System 1994). Fuel 80, 459–471. for oil and gas. In: Law, B.E., Rice, D.D. (Eds.), Studies in Geology Series 39–77 Tulsa.
Jackson, R.E., Reddy, K.J., 2007. Geochemistry of coalbed natural gas (CBNG) produced Levine, J.R., 1996. Model study of the influence of matrix shrinkage on absolute perme-
water in Powder River Basin, Wyoming: salinity and sodicity. Water, Air, and Soil ability of coal bed reservoirs. In: Gayer, R., Harris, I. (Eds.), Coalbed Methane and
Pollution 184, 49–61. Coal Geology: Geological Society Special Publication, 109, pp. 197–212.
Jin, H., Schimmelmann, A., Mastalerz, M., Pope, J., Moore, T.A., 2010. Coalbed gas de- Levine, J.R., Jonson, P.W., Beamish, B.B., 1993. High-pressure microgravimetry provides
sorption in canisters: consumption of trapped atmospheric oxygen and implica- a viable alternative to volumetric method in gas sorption studies on coal. Proceed-
tions for measured gas quality. International Journal of Coal Geology 81, 64–72. ings of the 1993 International Coalbed Methane Symposium. University of Ala-
Joubert, J.I., Grein, C.T., Bienstock, D., 1974. Effect of moisture on the methane capacity bama, Tuscaloosa, Al, pp. 187–195.
of American coals. Fuel 53, 186–191. Levy, J.H., Day, S.J., Killingley, J.S., 1997. Methane capacities of Bowen Basin coals relat-
Karacan, C.Ö., 2003. Heterogenous sorption and swelling in a confined and stressed ed to coal properties. Fuel 76, 813–819.
coal during CO2 injection. Energy & Fuels 17, 1595–1608. Li, H., Ogawa, Y., 2001. Pore structure of sheared coals and related coalbed methane.
Karacan, C.Ö., Okandan, E., 2000. Fracture/cleat analysis of coals from Zonguldak Basin Environmental Geology 40, 1455–1461.
(northwestern Turkey) relative to potential of coalbed methane production. Inter- Li, H., Ogawa, Y., Simada, S., 2003. Mechanism of methane flow through sheared coals
national Journal of Coal Geology 44, 109–125. and its role on methane recovery. Fuel 82, 1271–1279.
Karacan, C.Ö., Larsen, J.W., Esterle, J.S., 2009. CO2 sequestration in coals and enhanced Li, D., Hendry, P., Faiz, M., 2008. A survey of the microbial populations in some Austra-
coalbed methane recovery. International Journal of Coal Geology 77, 1–242. lian coalbed methane reservoirs. International Journal of Coal Geology 76, 14–24.
Karacan, C.Ö., Ruiz, F.A., Cotè, M., Phipps, S., 2011. Coal mine methane: a review of cap- Li, W., Cheng, Y.-P., Wang, L., 2011. The origin and formation of CO2 gas pools in the
ture and utilization practices with benefits to mining safety and to greenhouse gas coal seam of the Yaojie coalfield in China. International Journal of Coal Geology
reduction. International Journal of Coal Geology 86, 121–156. 85, 227–236.
Kędzior, S., 2009. Accumulation of coal-bed methane in the south-east part of the Liu, J., Chen, Z., Elsworth, D., Qu, H., Chen, D., 2011. Interactions of multiple processes
Upper Silesian Coal Basin (southern Poland). International Journal of Coal during CBM extraction: a critical review. International Journal of Coal Geology
Geology 80, 20–34. 87, 175–189.
Kędzior, S., 2011. The occurrence of a secondary zone of coal-bed methane in the Lüttig, G., 1986. Plants to peat: the process of humification. In: Fuchsman, C.H. (Ed.),
southern part of the Upper Silesian Coal Basin (southern Poland): potential for Peat and Water. Elsevier, Amsterdam, pp. 9–19.
methane exploitation. International Journal of Coal Geology 86, 157–168. Lyons, P.C., 1998. Introduction. International Journal of Coal Geology 38, 1–2.
Khavari-Khoransani, G., Michelsen, J.K., 1999. Coal bed gas content and gas Majewska, Z., Majewski, S., Ziętek, J., 2010. Swelling of coal induced by cyclic sorp-
undersaturation. In: Mastalerz, M., Glikson, M., Golding, S.D. (Eds.), Coalbed Meth- tion/desorption of gas: experimental observations indicating changes in coal
ane: Scientific, Environmental and Economic Evaluation. Kluwer Academic Pub- structure due to sorption of CO2 and CH4. International Journal of Coal Geology
lishers, Dordrecht, pp. 207–231. 83, 475–483.
T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81 79

Mares, T.E., 2009. An Investigation of the Relationship Between Coal and Gas Properties Convention & Exhibition. Indonesian Petroleum Association, Jakarta, Indonesia.
in the Huntly coalfield, New Zealand. University of Canterbury, Christchurch, De- Paper IPA09-G-056, unpaginated.
partment of Geological Sciences. 394 pp. Mukhopadhyay, P.K., Hatcher, P.G., 1993. Composition of coal. In: Law, B.E., Rice, D.D.
Mares, T.E., Moore, T.A., 2008. The influence of macroscopic texture on biogenically- (Eds.), Hydrocarbons from Coal. American Association of Petroleum Geologists,
derived coalbed methane, Huntly coalfield, New Zealand. International Journal of Tulsa, Oklahoma, pp. 79–118.
Coal Geology 76, 175–185. Murchison, D.G., Raymond, A.C., 1989. Igneous activity and organic maturation in the
Mares, T.E., Moore, T.A., Moore, C.R., 2009a. Uncertainty of gas saturation estimates in a Midland Valley of Scotland. International Journal of Coal Geology 14, 47–82.
subbituminous coal seam. International Journal of Coal Geology 77, 320–327. Murray, D.K., 1996. Coalbed methane in the USA: analogues for worldwide develop-
Mares, T.E., Radlinski, A.P., Moore, T.A., Cookson, D., Thiyagarajan, P., Ilavsky, J., Klepp, J., ment. Coalbed methane and coal geology. In: Gayer, R., Harris, I. (Eds.), Coalbed
2009b. Assessing the potential for CO2 adsorption in a subbituminous coal, Huntly Methane and Coal Geology: Geological Society Special Publication, 109, pp. 1–12.
Coalfield, New Zealand, using small angle scattering techniques. International Jour- Myers, T., 2009. Groundwater management and coal bed methane development in the
nal of Coal Geology 77, 54–68. Powder River Basin of Montana. Journal of Hydrology 368, 178–193.
Martini, A.M., Nüsslein, K., Petsch, S.T., 2005. Enhancing microbial gas from unconven- Nas, C., 1994. Spatial Variations in the Thickness and Coal Quality of the Sangatta Seam,
tional reservoirs. Gastips 11, 3–7. Kutei Basin, Kalimantan. University of Wollongong, Indonesia.
Mastalerz, M., Glikson, M., Golding, S.D., 1999. Coalbed Methane: Scientific, Environ- Nelson, C.R., Hill, D.G., Pratt, T.J., 2000. Properties of Paleocene Fort Union Formation
mental and Economic Evaluation. Kluwer Academic Publishers, Dordrecht. 592 pp. Canyon seam coal at the Triton Federal coalbed methane well, Campbell County,
Mastalerz, M., Gluskoter, H., Rupp, J., 2004. Carbon dioxide and methane sorption in Wyoming. SPE/CERI Gas Technology Symposium, Calgary, Alberta Canada 3–5
high volatile bituminous coals from Indiana, USA. International Journal of Coal Ge- April, pp. 639–649.
ology 60, 43–45. Newell, K.D., 2007. Wellsite, laboratory, and mathematical techniques for determining
Mastalerz, M., Drobniak, A., Schimmelmann, A., 2009. Changes in optical properties, sorbed gas contents of coals and gas shales utilizing well cuttings. Natural Re-
chemistry, and micropore and mesopore characteristics of bituminous coal at the sources Research 16, 55–66.
contact with dikes in the Illinois Basin. International Journal of Coal Geology 77, Newman, J., Newman, N.A., 1982. Reflectance anomalies in Pike River coals: evidence of
310–319. variability in vitrinite type, with implications for maturation studies and “Suggate
Mavor, M., Nelson, C., 1997. Coalbed reservoir Gas-In-Place Analysis. Gas Research In- rank”. New Zealand Journal of Geology and Geophysics 25, 233–243.
stitute, Chicago, Illinois, USA. GRI-97/0263. Nie, R.-S., Meng, Y.-F., Guo, J.-C., Jia, Y.-L., 2012. Modeling transient flow behavior of a
Mavor, M.J., Hartman, C., Pratt, T.J., 2004. Uncertainty in sorption isotherm measure- horizontal well in a coal seam. International Journal of Coal Geology 92, 54–68.
ments. 2004 International Coalbed Methane Symposium. University of Alabama, Ottiger, S., Pini, R., Storti, G., Mazzotti, M., 2008. Measuring and modeling the compet-
Tuscaloosa, Alabama, Paper 0411. unpaginated. itive adsorption of CO2, CH4 and N2 on a dry coal. Langmuir 24, 9531–9540.
Mazumder, S., Wolf, K.H., 2008. Differential swelling and permeability change of coal in re- Ozdemir, E., 2009. Modeling of coal bed methane (CBM) production and CO2 seques-
sponse to CO2 injection for ECBM. International Journal of Coal Geology 74, 123–138. tration in coal seams. International Journal of Coal Geology 77, 145–152.
Mazumder, S., Scott, M., Jiang, J., 2012. Permeability increase in Bowen Basin coal as a Ozdemir, E., Schroeder, K., 2009. Effect of moisture on adsorption isotherms and ad-
result of matrix shrinkage during primary depletion. International Journal of Coal sorption capacities of CO2 on coals. Energy & Fuels 23, 2821–2831.
Geology 96–97, 109–119. Ozdemir, E., Morsi, B.I., Schroeder, K., 2003. Importance of volume effects to adsorption
Mazzotti, M., Pini, R., Storti, G., 2009. Enhanced coalbed methane recovery. Journal of isotherms of carbon dioxide on coals. Langmuir 19, 9764–9773.
Supercritical Fluids 47, 619–627. Palmer, I., 2009. Permeability changes in coal: analytical modeling. International Jour-
McBeth, I., Reddy, K.J., Skinner, Q.D., 2003. Chemistry of trace elements in coalbed nal of Coal Geology 77, 119–126.
methane product water. Water Research 37, 884–890. Palmer, I., Mansoori, J., 1998. How permeability depends on stress and pore pressure in
McCants, C.Y., Spafford, S., Stevens, S.H., 2001. Five-spot production pilot on tight spac- coal beds: a new model. SPE Reservoir Evaluation and Engineering 1, 539–544 SPE-
ing: rapid evaluation of a coalbed methane block in the Upper Silesian Coal Basin, 52607-PA.
Poland. The 2001 International Coalbed Methane Symposium. University of Ala- Pan, Z., Connell, L.D., 2007. A theoretical model for gas adsorption-induced coal swell-
bama, Tuscaloosa, pp. 193–204. ing. International Journal of Coal Geology 69, 243–252.
McCulloch, C.M., Levine, J.R., Kissell, F.N., Deul, M., 1975. Measuring the methane con- Pan, Z., Connell, L.D., 2011a. Modelling of anisotropic coal swelling and its impact on
tent of bituminous coalbeds. Report of Investigations 8043. US Bureau of Mines, permeability behaviour for primary and enhanced coalbed methane recovery. In-
Washington D.C. 22 pp. ternational Journal of Coal Geology 85, 257–267.
McLennan, J.D., Schafer, P.S., Pratt, T., 1995. A Guide to Determining Coalbed Gas Con- Pan, Z., Connell, L.D., 2011b. Impact of coal seam as interlayer on CO2 storage in saline
tent. Gas Research Institute (GRI), Chicago, Illinois. GRI-94-0396, 300 pp. aquifers: a reservoir simulation study. International Journal of Greenhouse Gas
Meneley, R.A., Calverley, A.E., Logan, K.G., Proctor, R.M., 2003. Resource assessment Control 5, 99–114.
methodologies: current status and future direction. AAPG Bulletin 87, 535–540. Pan, Z., Connell, L.D., 2012. Modelling permeability for coal reservoirs: a review of an-
Midgley, D.J., Hendry, P., Pinetown, K.L., Fuentes, D., Gong, S., Mitchell, D.L., Faiz, M., alytical models and testing data. International Journal of Coal Geology 92, 1–44.
2010. Characterisation of a microbial community associated with a deep, coal Pan, Z., Connell, L.D., Camilleri, M., 2010. Laboratory characterisation of coal reservoir
seam methane reservoir in the Gippsland Basin, Australia. International Journal permeability for primary and enhanced coalbed methane recovery. International
of Coal Geology 82, 232–239. Journal of Coal Geology 82, 252–261.
Milewska-Duda, J., Duda, J., Nodzeñski, A., Lakatos, J., 2000. Absorption and adsorption Papendick, S.L., Downs, K.R., Vo, K.D., Hamilton, S.K., Dawson, G.K.W., Golding, S.D.,
of methane and carbon dioxide in hard coal and active carbon. Langmuir 16, Gilcrease, P.C., 2011. Biogenic methane potential for Surat Basin, Queensland coal
5458–5466. seams. International Journal of Coal Geology 88, 123–134.
Montgomery, S.L., 1999. Powder River Basin, Wyoming: an expanding coalbed meth- Pashin, J.C., 1998. Stratigraphy and structure of coalbed methane reservoirs in the Unit-
ane (CBM) play. AAPG Bulletin 83, 1207–1222. ed States: an overview. International Journal of Coal Geology 35, 209–240.
Moore, T.A., 2012. How many holes does it take to fill the Albert Hall? The importance Pashin, J.C., 2007. Hydrodynamics of coalbed methane reservoirs in the Black Warrior
of porosity in a coalbed methane reservoir. 36th Annual Convention & Exhibition, Basin: key to understanding reservoir performance and environmental issues. Ap-
Indonesian Petroleum Association. Indonesian Petroleum Association, Paper plied Geochemistry 22, 2257–2272.
IPA12-G-092, Jakarta, Indonesia. unpaginated. Pashin, J.C., 2010. Variable gas saturation in coalbed methane reservoirs of the Black
Moore, T.A., Butland, C.I., 2005. Coal seam gas in New Zealand as a model for Indonesia. Warrior Basin: implications for exploration and production. International Journal
In: Prihatmoko, S., Digdowirogo, S., Nas, C., Leeuwen, T.v., Widjajanto, H. (Eds.), In- of Coal Geology 82, 135–146.
donesian Mineral and Coal Discoveries. Indonesian Association of Geologists, Pashin, J.C., 2011. Coalbed methane. American Association of Petroleum Geologists, En-
Bogor, Indonesia, pp. 192–200. ergy Minerals Division, Unconventional Energy Resources: 2011 Review: Natural
Moore, T.A., Friederich, M.C., 2010. A probabilistic approach to estimation of coalbed Resources Research, 20, pp. 282–286.
methane resources for Kalimantan, Indonesia. In: Basuki, N.I., Prihatmoko, S. Paul, S., Chatterjee, R., 2011. Determination of in-situ stress direction from cleat orien-
(Eds.), Kalimantan Coal and Mineral Resources. MGEI-IAGI Seminar, 29–30 tation mapping in coal bed methane exploration in south-eastern part of Jharia
March 2010, Balikpapan, Indonesia, pp. 61–71. Coalfield, India. International Journal of Coal Geology 88, 87–96.
Moore, T.A., Hilbert, R.E., 1992. Petrographic and anatomical characteristics of plant Penner, T.J., Foght, J.M., Budwill, K., 2010. Microbial diversity of western Canadian sub-
material from two peat deposits of Holocene and Miocene age, Kalimantan, Indo- surface coal beds and methanogenic coal enrichment cultures. International Jour-
nesia. Review of Palaeobotany and Palynology 72, 199–227. nal of Coal Geology 82, 81–93.
Moore, T.A., Shearer, J.C., Miller, S.L., 1996. Fungal origin of oxidised plant material in Philibert, J., 2005. One and a half century of diffusion: Fick, Einstein, before and beyond.
the Palangkaraya peat deposit, Kalimantan Tengah, Indonesia: implications for Diffusion Fundamentals 2, 1–10.
“inertinite” formation in coal. International Journal of Coal Geology 30, 1–23. Pillalamarry, M., Harpalani, S., Liu, S., 2011. Gas diffusion behavior of coal and its impact
Moore, T.A., Esterle, J.S., Shearer, J.C., 2000. The rôle and development of texture in coal on production from coalbed methane reservoirs. International Journal of Coal Ge-
beds: examples from Indonesia and New Zealand. In: Hadiyanto (Ed.), Southeast ology 86, 342–348.
Asian Coal Geology Conference. Ministry of Mines and Energy, Bandung, pp. Pitman, J.K., Pashin, J.C., Hatch, J.R., Goldhaber, M.B., 2003. Origin of minerals in joint
239–252. and cleat systems of the Pottsville Formation, Black Warrior basin, Alabama: impli-
Moore, T.A., Flores, R.M., Stanton, R.W., Stricker, G.D., 2002. The rôle of macroscopic cations for coalbed methane generation and production. AAPG Bulletin 87,
texture in determining coalbed methane variability in the Anderson-Wyodak 713–731.
coal seam, Powder River Basin, Wyoming. In: Robison, C.R. (Ed.), 18th Annual Qin, S., Song, Y., Tang, X., Fu, G., 2005. The mechanism of the flowing ground water
Meeting of the Society for Organic Petrology. The Society for Organic Petrology, impacting on coalbed gas content. Chinese Science Bulletin 50, 118–123.
Houston, Tx, pp. 85–88. Radlinski, A.P., Radlinska, E.Z., 1999. Coal bed gas content and gas undersaturation. In:
Moore, T.A., Mares, T.E., Moore, C.R., 2009. Assessing uncertainty of coalbed methane Mastalerz, M., Glikson, M., Golding, S.D. (Eds.), Coalbed Methane: Scientific, Environmen-
resources. Proceedings, Indonesian Petroleum Association, 33rd Annual tal and Economic Evaluation. Kluwer Academic Publishers, Dordrecht, pp. 329–365.
80 T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81

Radlinski, A.P., Mastalerz, M., Hinde, A.L., Hainbuchner, M., Rauch, H., Baron, M., Lin, J.S., Shimizu, S., Akiyama, M., Naganuma, T., Fuioka, M., Nako, M., Ishijim, Y., 2007. Molec-
Fan, L., Thiyagarajan, P., 2004. Application of SAX and SANS in evaluation of poros- ular characterization of microbial communities in deep coal seam groundwater
ity, pore size and surface area of coal. International Journal of Coal Geology 59, of northern Japan. Geobiology 5, 423–433.
245–271. Siemons, N., Busch, A., 2007. Measurement and interpretation of supercritical sorption
Radovic, L.R., Menon, V.C., Leon, Y., Leon, C.A., Kyotani, T., Danner, R.P., Anderson, S., on various coals. International Journal of Coal Geology 69, 229–242.
Hatcher, P.G., 1997. On the porous structure of coals: evidence for an Sinayuç, C., Gümrah, F., 2009. Modeling of ECBM recovery from Amasra coalbed in Zon-
interconnected but constricted micropore system and implications for coalbed guldak Basin, Turkey. International Journal of Coal Geology 77, 162–174.
methane recovery. Adsorption 3, 221–232. Siriwardane, H., Haljasmaa, I., McLendon, R., Irdi, G., Soong, Y., Bromhal, G., 2009. Influ-
Ramurthy, M., Marjerisson, D.M., Daves, S.B., 2002. Diagnostic Fracture Injection Test in ence of carbon dioxide on coal permeability determined by pressure transient
Coals to Determine Pore Pressure and Permeability. SPE Gase Technology Sympo- methods. International Journal of Coal Geology 77, 109–118.
sium. Society of Petroleum Engineers, Calgary, Alberta, Canada. 16 pp. Sivek, M., Jirásek, J., Sediáčková, L., Čáslavský, M., 2010. Variation of moisture content of
Raymond, R., 1986. Out of the Fiery Furnace. The Pennsylvania State University, Univer- the bituminous coals with depth: a case study from the Czech part of the Upper Si-
sity Park, Pennsylvania, The impact of metals on the history of mankind. 273 pp. lesian Coal Basin. International Journal of Coal Geology 84, 16–24.
Reeves, S., 2001. Geologic sequestration of CO2 in deep, unmineable coalbeds: an inte- Sloss, L.L., 2005. Coalbed methane emissions—capture and utilisation. IEA Clean Coal
grated research and commercial-scale field demonstration project. First National Centre, London, CCC/104. 56 pp.
Conference on Carbon Sequestration, May 14–17, 2001. Department of Energy Na- Smith, J.W., Pallasser, R.J., 1996. Microbial origin of Australian coalbed methane. AAPG
tional Energy Technology Laboratory, Washington DC. 12 pp., http://www.netl. Bulletin 80, 891–897.
doe.gov/publications/proceedings/01/carbon_seq/3a3.pdf. Smyth, M., Buckley, M.J., 1993. Statistical analysis of the microlithotype sequences in
Rice, D.D., 1993. Composition and origins of coalbed gas. In: Law, B.E., Rice, D.D. (Eds.), the Bulli Seam, Australia, and relevance to permeability for coal gas. International
Hydrocarbons from Coal. : AAPG Studies in Geology, 38. American Association of Journal of Coal Geology 22, 167–187.
Petroleum Geologists, Tulsa, Oklahoma, pp. 159–184. Snyder, G.T., Riese, W.C., Franks, S., Pelzmann, W.L., Gorody, A.W., Moran, J.E., 2003. Or-
Rice, C.A., 2003. Production waters associated with the Ferron coalbed methane fields, igin and history of waters associated with coalbed methane: 129I, 36Cl, and stable
central Utah: chemical and isotopic composition and volumes. International Jour- isotope results from the Fruitland Formation, CO and MN. Geochimica et
nal of Coal Geology 56, 141–169. Cosmochimica Acta 67, 4529–4544.
Rice, D.D., Claypool, G.E., 1981. Generation, accumulation, and resource potential of Somerton, W.H., Söylemezoḡlu, I.M., Dudley, R.C., 1975. Effects of stress on permeabil-
biogenic gas. AAPG Bulletin 65, 5–25. ity of coal. International Journal of Rock Mechanics and Mining Science &
Rice, D.D., Clayton, J.L., Pawlewicz, M.J., 1989. Characterization of coal-derived hydro- Geomechanics Abstracts 12, 129–145.
carbons and source-rock potential of coal beds, San Juan Basin, New Mexico and Sparks, D.P., McLendon, T.H., Saulsberry, J.L., Lambert, S.W., 1995. The effects of stress
Colorado, U.S.A. International Journal of Coal Geology 13, 597–626. on coalbed reservoir performance, Black Warrior Basin, U.S.A. Proceedings of the
Rice, C.A., Ellis, M.S., Bullock, J.H.J., 2000. Water co-produced with coalbed methane in Society of Petroleum Engineers Annual Technical Conference and Exhibition. SPE
the Powder River Basin. Wyoming: Preliminary Compositional Data. USGS Open- Paper 30743, Dallas, Texas, pp. 339–351.
File Report 00‐372, Denver, Colorado. 45 pp. St. George, J.D., Barakat, M.A., 2001. The change in effective stress associated with shrink-
Rightmire, C.T., Eddy, G.E., Kirr, J.N., 1984. Coalbed Methane Resources of the United age from gas desorption in coal. International Journal of Coal Geology 45, 105–113.
States. AAPG Studies in Geology, 17. American Association of Petroleum Geologists, Stanton, R.W., Flores, R.M., Warwick, P.D., Gluskoter, H., Stricker, G.D., 2001. Coal bed
Tulsa, Oklahoma. 375 pp. sequestration of carbon dioxide. First National Conference on Carbon Sequestra-
Robertson, E.P., 2009. Economic analysis of carbon dioxide sequestration in Powder tion, May 14–17, 2001. Department of Energy National Energy Technology Labora-
River Basin coal. International Journal of Coal Geology 77, 234–241. tory, Washington DC. CD-ROM Proceedings, paper 3A.3, 12 pp., http://www.netl.
Rodrigues, C.F., Lemos de Sousa, M.J., 2002. The measurement of coal porosity with dif- doe.gov/publications/proceedings/01/carbon_seq/3a3.pdf.
ferent gases. International Journal of Coal Geology 48, 245–251. Stearns, M., Tindall, J.A., Cronin, G., Friedel, M.J., Bergquist, E., 2005. Effects of coal-bed
Roehler, R.W., 1988. The Pintail coal bed and barrier bar G—a model for coal of barrier methane discharge waters on the vegetation and soil ecosystem in Powder River
bar-lagoon origin. Upper Cretaceous Almond Formation, Rock Springs Coal Field, Basin, Wyoming. Water, Air, and Soil Pollution 168, 33–57.
Wyoming. : U.S. Geological Survey Professional Paper, 1398. 60 pp. Stevens, S.H., Sani, K., Hardjosuwiryo, S., 2001. Indonesia's 337 tcf CBM resource: a low-
Rogoff, M.H., Wender, I., Anderson, R.B., 1962. Microbiology of coal. U.S. Bureau of cost alternative to gas, LNG. Oil & Gas Journal 99, 40–45.
Mines. Information Circular 8075. 85 pp. Stopes, M.C., 1919. On the four visible ingredients in banded bituminous coal. Proceed-
Ross, H.E., Hagin, P., Zoback, M.D., 2009. CO2 storage and enhanced coalbed methane ings of the Royal Society, Series B 90, 470–487.
recovery: reservoir characterisation and fluid flow simulations of the Big George Strąpoć, D., Mastalerz, M., Eble, C., Schimmelmann, A., 2007. Characterization of the or-
coal, Powder River Basin, Wyoming, USA. International Journal of Greenhouse igin of coalbed gases in southeastern Illinois Basin by compound-specific carbon
Gas Control 3, 773–786. and hydrogen stable isotope ratios. Organic Geochemistry 38, 267–287.
Saghafi, A., Williams, R.J., Roberts, D.B., 1995. Determination of coal gas content by Strąpoć, D., Picardal, F.W., Turich, C., Schaperdoth, I., Macalady, J.L., Lipp, J.S., Lin, Y.-S.,
quick crushing method. CSIRO, Sydney. 9 pp. Ertefai, T.F., Schubotz, F., Hinrichs, K.-U., Mastalerz, M., Schimmelmann, A., 2008a.
Saxby, J.D., Shibaoka, M., 1986. Coal and coal macerals as source rocks for oil and gas. Methane-producing microbial community in a coal bed of the Illinois Basin. Ap-
Applied Geochemistry 1, 25–36. plied and Environmental Microbiology 74, 2424–2432.
Schneider, W., 1986. Vergelung im 2. miozanen Flozhorizont der Lausitz - Genese und Strąpoć, D., Mastalerz, M., Schimmelmann, A., Drobniak, A., Hedges, S., 2008b. Variabil-
Erscheinung. Wiss. Techn. Geol. Wiss. 14, 735–744. ity of geochemical properties in a microbially dominated coalbed gas system from
Schopf, J.M., 1966. Definitions of peat and coal and of graphite that terminates the coal the eastern margin of the Illinois Basin, USA. International Journal of Coal Geology
series (graphocite). Journal of Geology 74, 584–592. 76, 98–100.
Scott, A.R., 1993. Composition and origin of coalbed gases from selected basins in the Stricker, G.D., Flores, R.M., McGarry, D.E., Stilwell, D.P., Hoppe, D.J., Stilwell, K.R., Ochs,
United States. Proceedings of the 1993 International Coalbed Methane Symposium, A.M., Ellis, M.E., Osvald, K.S., Taylor, S.L., Thorvaldson, M.C., Trippi, M.H., Grose, S.D.,
Birmingham, Alabama, pp. 207–222. Cockett, F.J., Shariff, A.A., 2006. Gas desorption isotherm studies in coals in the
Scott, A.R., 1999. Coal bed gas content and gas undersaturation. In: Mastalerz, M., Powder River Basin and adjoining basins in Wyoming and North Dakota. U.S. Geo-
Glikson, M., Golding, S.D. (Eds.), Coalbed Methane: Scientific, Environmental and logical Survey, Reston, Va, USGS Open File Report 2006‐1174. 273 pp.
Economic Evaluation. Kluwer Academic Publishers, Dordrecht, pp. 89–110. Su, X., Feng, Y., Chen, J., Pan, J., 2001. The characteristics and origins of cleat in coal from
Scott, A.R., 2002. Hydrogeological factors affecting gas content distribution in coal beds. Western North China. International Journal of Coal Geology 47, 51–62.
International Journal of Coal Geology 50, 363–387. Sumner, G.M., 1980. Canary in a coal mine. Police, Zenyatta Mondatta: A&M Records,
Scott, A.R., Kaiser, W.R., Ayers Jr., W.B., 1994. Thermogenic and secondary biogenic SP-3720, Track 4.
gases, San Juan basin, Colorado and New Mexico—implications for coalbed gas Susilawati, R., Ward, C.R., 2006. Metamorphism of mineral matter in coal from the
producibility. AAPG Bulletin 78, 1186–1209. Bukit Asam deposit, south Sumatra, Indonesia. International Journal of Coal Geolo-
Scott, A.R., Zhou, N., Levine, J.R., 1995. A modified approach to estimating coal and coal gy 68, 171–195.
gas resources: example from the Sand Wash Basin, Colorado. AAPG Bulletin 79, Taulis, M., Milke, M., 2007. Coal seam gas water from Maramarua, New Zealand: char-
1320–1336. acterisation and comparison to United States analogues. New Zealand Journal of
Scott, S., Anderson, B., Crosdale, P.J., Dingwall, J., Leblang, G., 2007. Coal petrology and Hydrology 46, 1–17.
coal seam gas contents of the Walloon Subgroup—Surat Basin, Queensland, Austra- Taulis, M., Milke, M., 2009. Sodium removal from Maramarua coal seam gas waters
lia. International Journal of Coal Geology 70, 209–222. using Ngakuru zeolites. Proceedings of the Water New Zealand Conference and
Şenel, I.G., Gürüz, A.G., Yücel, H., 2001. Characterization of pore structure of Turkish Expo. The New Zealand Water & Waste Association, Rotorua. unpaginated.
coals. Energy & Fuels 15, 331–338. Taylor, G.H., Teichmüller, M., Davis, A., Diessel, C.F.K., Littke, R., Robert, P., 1998. Organic
Shearer, J.C., Moore, T.A., 1996. Effects of experimental coalification on texture, compo- Petrology. Gebrüder Borntraeger, Berlin. 704 pp.
sition and compaction in Indonesian peat and wood. Organic Geochemistry 24, Teichmüller, M., 1989. The genesis of coal from the viewpoint of coal petrology. Inter-
127–140. national Journal of Coal Geology 12, 1–87.
Shearer, J.C., Moore, T.A., Demchuk, T.D., 1995. Delineation of the distinctive nature of Ting, F.T.C., 1977. Origin and spacing of cleats in coal beds. Journal of Pressure Vessel
Tertiary coal beds. International Journal of Coal Geology 28, 71–98. Technology 99, 624–626.
Shi, J.Q., Durucan, S., 2005. A model for changes in coalbed permeability during primary Ulrich, G., Bower, S., 2008. Active methanogenesis and acetate utilization in Powder
and enhanced methane recovery. SPE Reservoir Evaluation and Engineering 8, River Basin coals, United States. International Journal of Coal Geology 76, 25–33.
291–299 SPE-87230-PA. United Nations, 2010. Best practice guidance for effective methane drainage and use in
Shi, J.-Q., Durucan, S., Fujioka, M., 2008. A reservoir simulation study of CO2 injection coal mines. ECE Energy Series, 31. United Nations, New York. 92 pp.
and N2 flooding at the Ishikari coalfield CO2 storage pilot project, Japan. Interna- Unsworth, J.F., Flowler, C.S., Jones, L.F., 1989. Moisture in coal: 2. Maceral effects on
tional Journal of Greenhouse Gas Control 2, 47–57. pore structure. Fuel 68, 18–26.
T.A. Moore / International Journal of Coal Geology 101 (2012) 36–81 81

van Holst, J., Stalker, L., Le, Y., Sestak, S., 2010. Gas stable isotopes analysis—sample con- Proceedings of the Symposium on Geology in Longwall Mining, 12th–13th Novem-
tainment and ∂13C stability. Australian Organic Geochemistry Conference, Canber- ber 1996. University of New South Wales, Sydney, p. 11.
ra, ACT. unpaginated. Williams, R.J., Weissmann, J.J., 1995. Gas emission and outburst assessment in mixed
Van Voast, W.A., 2003. Geochemical signature of formation waters associated with CO2 and CH4 environments. Proceedings ACIRL Underground Mining Seminar. Aus-
coalbed methane. AAPG Bulletin 87, 667–676. tralian Coal Industry Research Laboratory, North Ryde, Australia. 12 pp.
Vessey, S.J., Bustin, R.M., 1999. Coalbed methane characteristics of the Mist Mountain Woese, C.R., Kandler, O., Wheelis, M.L., 1990. Towards a natural system of organisms:
Formation, Southern Canadian Cordillera: effect of shearing and oxidation. In: proposal for the domains Archaea, Bacteria, and Eucarya. Proceedings of the Na-
Mastalerz, M., Glikson, M., Golding, S.D. (Eds.), Coalbed Methane: Scientific, Envi- tional Academy of Science of the USA 87, 4576–4579.
ronmental and Economic Evaluation. Kluwer Academic Publishers, Dordrecht, pp. Wong, S., Foy, C., Gunter, B., Jack, T., 1998. Injection of CO2 for enhanced gas recovery:
367–384. coal bed methane versus oil recovery. Proceedings of the Fourth International Con-
Waechter, N.B., Hampton III, G.L., Shipps, J.C., 2004a. Overview of coal and shale gas ference on Greenhouse Gas Control technologies. Interlaken: Elsevier Science, pp.
measurements: field and laboratory procedures. Proceeding of the 2004 Interna- 175–180.
tional Coalbed Methane Symposium. The University of Alabama, Tuscaloosa, Ala- Wong, S., Law, D., Deng, X., Robinson, J., Kadatz, B., Gunter, W.D., Jianping, Y., Sanli, F.,
bama. 17 pp. Zhiqiang, F., 2007. Enhanced coalbed methane and CO2 storage in anthracitic coals
Waechter, N.B., Hampton III, G.L., Shipps, J.C., Seidle, J.P., 2004b. Comparison of lost gas —micro-pilot test at South Qinshui, Shanxi, China. International Journal of Green-
projections in coalbed methane. Rocky Mountain Section AAPG Poster Session, house Gas Control 1, 215–222.
Denver, CO. unpaginated. Wong, S., Macdonald, D., Andrei, S., Gunter, W.D., Deng, X., Law, D.H.-S., Ye, J., Feng, S.,
Waksman, S.A., Stevens, K.R., 1929. Contribution to the chemical composition of peat: Fan, S., Fan, Z., Ho, P., 2010. Conceptual economics of full scale enhanced coalbed
V. The role of microorganisms in peat formation and decomposition. Soil Science methane production and CO2 storage in anthracitic coals at South Quinshui
28, 315–340. basin, Shanxi, China. International Journal of Coal Geology 82, 280–286.
Wang, X., 2006. Influence of Coal Quality Factors on Seam Permeability Associated with Wu, Y., Liu, J., Elsworth, D., Siriwardane, H., Miao, X., 2011. Evolution of coal permeabil-
Coalbed Methane Production. University of New South Wales, Sydney, Australia. ity: contribution of heterogeneous swelling processes. International Journal of Coal
338 pp. Geology 88, 152–162.
Wang, X., Ward, C., 2009. Experimental investigation of permeability changes with Wyoming Oil and Gas Commission, 2011. http://wogcc.state.wy.us/2011.
pressure depletion in relation to coal quality. International Coalbed Methane Sym- Yao, Y., Liu, D., Tang, D., Tang, S., Huang, W., 2008. Fractal characterisation of
posium. University of Alabama, Tuscaloosa, Alabama. 31 pp. adsorption-pores of coals from North China: an investigation on CH4 adsorption
Wang, X., Wu, Q., 2011. Methodology review of interpreting permeability of CBM res- capacity of coals. International Journal of Coal Geology 73, 27–42.
ervoir across Asia Pacific region. The 3rd Asia Pacific Coalbed Methane Symposium, Yee, D., Seidle, J.P., Hanson, W.B., 1993. Gas sorption on coal and measurement of gas
Brisbane, Australia, May 3–6, 2011, Paper No. 091. 10 pp. content, in: Law, B.E., Rice, D.D. (Eds.), Hydrocarbons from Coal. American Associ-
Wang, Z., Zeng, D., Patrick, W.H., 1996. Methane emissions from natural wetlands. En- ation of Petroleum Geologists, Studies in Geology 38, Tulsa, Oklahoma, Studies in
vironmental Monitoring and Assessment 42, 143–161. Geology 38, pp. 203–218.
Wang, A., Qin, Y., Wu, Y., Wang, B., 2010a. Status of research on biogenic coalbed gas Yu, H., Zhou, G., Fan, W., Ye, J., 2007. Predicted CO2 enhanced coalbed methane recovery
generation mechanisms. Mining Science and Technology 20, 271–275. and CO2 sequestration in China. International Journal of Coal Geology 71, 345–357.
Wang, G.X., Wei, X.R., Wang, K., Massarotto, P., Rudolph, V., 2010b. Sorption-induced Zabalza-Mezghani, I., Manceau, E., Feraille, M., Jourdan, A., 2004. Uncertainty manage-
swelling/shrinkage and permeability of coal under stressed adsorption/desorption ment: from geological scenarios to production scheme optimization. Journal of Pe-
conditions. International Journal of Coal Geology 83, 46–54. troleum Science and Engineering 44, 11–25.
Warwick, P.D., 2005. Coal systems analysis: a new approach to the understanding of Zarrouk, S.J., 2008. Reacting Flows in Porous Media. VDM Verlag Dr Muller
coal formation, coal quality and environmental considerations, and coal as a source Aktiengesellshaft & Co, KG, Berlin.
rock for hydrocarbons. In: Warwick, P.D. (Ed.), Coal Systems Analysis: Geological Zarrouk, S.J., Moore, T.A., 2009. Preliminary model for enhanced coalbed methane
Society of America Special Paper, 387, pp. 1–8. Boulder, Co., USA. (ECBM) through CO2 injection: Huntly Coalfield, New Zealand. International Jour-
Warwick, P.D., Breland Jr., F.C., Hackley, P.C., 2008. Biogenic origin of coalbed gas in the nal of Coal Geology 77, 153–161.
northern Gulf of Mexico Coastal Plain, U.S.A. International Journal of Coal Geology Zhang, E., Hill, R.J., Katz, B.J., Tang, Y., 2008. Modeling of gas generation from the Cameo
76, 119–137. coal zone in the Piceance Basin, Colorado. AAPG Bulletin 92, 1077–1106.
Wei, X.R., Wang, G.X., Massarotto, P., Golding, S.D., Rudolph, V., 2007. A review of re- Zhang, S., Tang, S., Tang, D., Pan, Z., Yang, F., 2010. The characteristics of coal reservoir
cent advances in the numerical simulation for coalbed-methane-recovery process. pores and coal facies in Liulin district, Hedong coal field of China. International
SPE Reservoir Evaluation and Engineering 10, 657–666. Journal of Coal Geology 81, 117–127.
Whiticar, M.J., 1994. Correlation of natural gases with their sources. In: Magoon, L.B., Zhou, Y., Zhou, L., 2009. Fundamentals of high pressure adsorption. Langmuir 25,
Dow, W.G. (Eds.), The Petroleum System—From Source to Trap. : Memoir, 60. 13461–13466.
AAPG, Tulsa, pp. 261–283. Zhou, Z., Ballentine, C.J., Kipfer, R., Schoell, M., Thibodeaux, S., 2005. Noble gas tracing of
Whiticar, M.J., Faber, F., Schoell, M., 1986. Biogenic methane formation in marine and groundwater/coalbed methane interaction in the San Juan Basin, USA. Geochimica
freshwater environments—CO2 reduction vs acetate fermentation—isotope evi- et Cosmochimica Acta 69, 5413–5428.
dence. Geochem. Cosmochim. Acta 50, 693–709. Zuber, D.M., 1999. Coal bed gas content and gas undersaturation. In: Mastalerz, M.,
Wilkins, R.W.T., George, S.C., 2002. Coal as a source rock for oil: a review. International Glikson, M., Golding, S.D. (Eds.), Coalbed Methane: Scientific, Environmental and
Journal of Coal Geology 50, 317–361. Economic Evaluation. Kluwer Academic Publishers, Dordrecht, pp. 55–66.
Williams, R.J., 1996. Application of gas content test data to the evaluation of gas emis-
sion and hazards during longwall development and extraction. In: Ward, C.R. (Ed.),

Anda mungkin juga menyukai