Sei sulla pagina 1di 39

Factors controlling prolific AUTHORS

Keith W. Shanley  7 Lantana, Littleton,


gas production from Colorado 80127; shanleypk@msn.com

low-permeability sandstone Keith W. Shanley is a consulting geologist with


more than 22 years experience in exploration,
development, and research. He has published
reservoirs: Implications for numerous papers dealing with sequence
stratigraphy and reservoir architecture and has
resource assessment, prospect served as editor of several publications. He
received his B.A. degree in geology from Rice
development, and risk analysis University and his M.Sc. degree and his Ph.D.
in geology from the Colorado School of Mines.
His current research interests include sequence
Keith W. Shanley, Robert M. Cluff, and John W. Robinson stratigraphy and reservoir architecture, the
integration of petrophysics, and risk analysis.
Robert M. Cluff  The Discovery
ABSTRACT Group, Inc., 1560 Broadway, Suite 1470,
Denver, Colorado 80202;
Low-permeability reservoirs from the Greater Green River basin of
bobcluff@discovery-group.com
southwest Wyoming are not part of a continuous-type gas accumu-
Bob Cluff is a geologist with 28 years
lation or a basin-center gas system in which productivity is depen-
petroleum exploration and development ex-
dent on the development of enigmatic sweet spots. Instead, gas fields
perience. Bob’s research interests include the
in this basin occur in low-permeability, poor-quality reservoir rocks
integration of geology with petrophysics and
in conventional traps. We examined all significant gas fields in the the evaluation of nonconventional reservoirs.
Greater Green River basin and conclude that they all occur in con- Bob received his B.S. in geology from the
ventional structural, stratigraphic, or combination traps. We illus- University of California-Riverside, his M.S. from
trate this by examining several large gas fields in the Greater Green the University of Wisconsin-Madison, with
River basin and suggest that observations derived from the Greater additional studies at the University of Illinois-
Green River basin provide insight to low-permeability, gas-charged Urbana-Champaign, University of Colorado-
sandstones in other basins. We present evidence that the basin is Denver, and Metropolitan State College Denver
neither regionally gas saturated, nor is it near irreducible water sat- in geology, math and physics. Bob founded
uration; water production is both common and widespread. Low- The Discovery Group in 1987.
permeability reservoirs have unique petrophysical properties, and John W. Robinson  22 North Ranch
failure to fully understand these attributes has led to a misunder- Road, Littleton, Colorado 80127;
standing of fluid distributions in the subsurface. An understanding robinson_ jw@msn.com
of multiphase, effective permeability to gas as a function of both John W. Robinson is a consulting geologist in
varying water saturation and overburden stress is required to fully Denver and has 30 years of experience in
appreciate the controls on gas-field distribution as well as the con- exploration and development. In 1999, he and
trols on individual well and reservoir performance. Low-permeability coauthor Peter McCabe received the AAPG
gas systems such as those found in the Greater Green River basin do Wallace Pratt Award for the best paper in the
not require a paradigm shift in terms of hydrocarbon systems as some 1997 AAPG Bulletin. He received B.S. and M.S.
have advocated. We conclude that low-permeability gas systems degrees in geology from San Diego State
similar to those found in the Greater Green River basin should be University and a Ph.D. in geology from the
evaluated in a manner similar to and consistent with conventional Colorado School of Mines. His research
interests are in fluvial sedimentology and
hydrocarbon systems.
multidisciplinary reservoir studies.

Copyright #2004. The American Association of Petroleum Geologists. All rights reserved.
Manuscript received June 1, 2003; provisional acceptance August 26, 2003; revised manuscript received
March 17, 2004; final acceptance March 25, 2004.

AAPG Bulletin, v. 88, no. 8 (August 2004), pp. 1083 – 1121 1083
To date, resource assessments in the Greater Green River basin
ACKNOWLEDGEMENTS have assumed a widespread, continuous-type resource distribution.
The ideas expressed in this paper have emerged Failure to recognize some of the fundamental elements of low-
over many years of trying to better under- permeability reservoirs has led to an underappreciation of the risks
stand tight-gas reservoirs and a near-constant associated with exploration and development investment decisions
interchange of ideas among the authors. K.W. in these settings and likely a significant overestimation of available
Shanley acknowledges the contribution of Lee resource levels.
T. Shannon (Anadarko Production Co.) to
many of the ideas expressed in this paper.
Conversations with Lee regarding the nature
INTRODUCTION
of tight gas date back to 1981 when both the
senior author and Lee Shannon were gradu-
With demand for efficient, environmentally clean sources of energy
ate students at the Colorado School of Mines.
Those conversations have continued to the increasing, attention is sharply focused on natural gas resources. In
present. In addition, K.W. Shanley benefited North America, much of this attention is directed at those un-
from several people at BP, where exposure to conventional resources (e.g., Law and Curtis, 2002), commonly re-
tight-gas data sets had a profound impact. ferred to as basin-centered or continuous gas accumulations. This
Peter Jordan, Mike Short, Kate Hadley, and focus reflects (1) the very large resource estimates assigned to these
Hunter Rowe provided management support targets, (2) the perception that they are associated with high prob-
of joint projects between business units and able success rates ( > 50% chance of technical recovery), and (3) the
the Technology Group. Alan Skorpen, Todd perception that the driving forces required to bring these resources
Stephenson, and Mike Bowman provided guid- to market are improvements in drilling and completion technolo-
ance and support within the Technology Group.
gies, higher product prices, and improved land access. Because of
Insights into low-permeability reservoirs were
the large volume of gas resources attributed to these types of
gained through discussions with many, includ-
accumulations, they are of interest not only to the upstream portions
ing Jack Thomas, Robert Lengerich (now de-
ceased), Dennis Cox, Bill Hanson, and Kirk Hird. of the oil and gas industry but also to the gas-transmission industry as
In terms of understanding the risks associated well as public policy entities charged with developing national energy
with tight-gas decisions, conversations with Peter policy and strategies for appropriate land-use management.
Carragher were instrumental. Since leaving BP, Natural gas is found in both conventional traps and as con-
conversations with Alan Byrnes (Kansas Geo- tinuous gas accumulations. Conventional traps are reasonably well
logical Survey), Sue Cluff (The Discovery Group), understood; they may be structural, stratigraphic, or a combination
Bob Coskey (Platte River Associates), John of the two and are associated with seals that trap gas migrating
Dacy (Core Laboratories, Houston, Texas), under buoyancy. Continuous-gas accumulations, by contrast, are
Tom Feldkamp (Kerr McGee, Denver, Colo- relatively poorly understood. They are thought to lack seals or traps,
rado), Jack Thomas (Consultant and AAPG, the role of buoyancy is thought to be diminished, and they are com-
Tulsa, Oklahoma), and Mihai Vasilache (SCAL
monly referred to as unconventional resources, although the mean-
Inc., Midland, Texas) have been instrumental.
Alan Byrnes provided data on relative per- ing of the term unconventional varies with time and perspective.
meability, and Dan Soeder (U.S. Geological Continuous-gas accumulations are regional in extent and are thought
Survey, Baltimore, Maryland) and Shirley of as large single fields having spatial dimensions equal to or ex-
Dutton (Texas Bureau Economic Geology, ceeding those of plays (Schmoker, 1995). Coalbed methane is per-
Austin, Texas) provided photographs of haps the most widely known and best understood type of continuous
various pore types. Discussions with Peter gas accumulation.
McCabe and Don Gautier (U.S. Geological Sur- In many of the Rocky Mountain basins of the western United
vey, Denver, Colorado) and AAPG reviewers States, natural gas accumulations in low-permeability sandstones
Lee Billingsley, Lee Krystinik, Kent Bowker, and are associated with widespread gas shows while drilling, a lack of
John Lorenz helped to further focus this paper. associated water production, and, as a result, have come to be re-
garded as continuous accumulations. Although gas shows are com-
mon, the productivity of wells drilled in these settings varies
dramatically. This has led to the suggestion that there are localized
areas with more favorable rock properties, sweet spots, or that varia-
tions in drilling and completion technology account for the pro-
ductivity variability.

1084 Gas Production from Low-Permeability Sandstone Reservoirs


This paper critically examines concepts surrounding (e.g., Masters, 1979, 1984; McPeek, 1981; Law and
basin-centered gas systems and challenges the hypoth- Dickenson, 1985; Spencer, 1987, 1989; Law et al.,
esis that many of the gas accumulations associated with 1989; Law and Spencer, 1993, in press; Surdam, 1997a;
low-permeability sandstones in the Rocky Mountain re- Law, 2000, 2002a, b; Schmoker, 2002). The term basin-
gion are continuous-type accumulations. Gas accumula- centered has come to include gas systems variously re-
tions in many of these low-permeability reservoirs occur ferred to as deep-basin gas systems, tight-gas systems,
in conventional traps, and considerations of reservoir, and continuous-type gas systems. Although in this
trap geometry, and seal are critical elements. Detailed paper, we take issue with the notion of basin-centered
investigation of multiphase permeability at in-situ con- gas, we use the term to refer to those systems that, upon
ditions has reinforced our view that petrophysical be- initial review, appear to meet the criterion established
havior of low-permeability reservoirs deviates substan- by earlier researchers (e.g., Schmoker, 2002; Law and
tially from traditional concepts. This understanding of Spencer, in press). Early descriptions of basin-centered
trap and petrophysics explains phenomena such as the gas had a concise list of attributes that characterized
apparent juxtaposition of water-bearing and gas-bearing these gas systems (e.g., Masters, 1979); however, more
strata, the lack of well-defined gas-water contacts, the recent work has recognized greater complexity and an
lack of widespread produced water, subtle variations in ever-increasing number of exceptions (e.g., Schmoker,
gas/water ratios, and dramatic variations in well pro- 1995, 2002; Law, 2002a, 2003; Law and Spencer, in
ductivity over short distances, all of which are common press). In many basin-centered accumulations, source
in basin-centered or continuous-type gas accumulations. rocks are thought to be in close physical proximity to
If these low-permeability gas accumulations are in con- reservoir rocks, and structural and stratigraphic traps,
ventional traps, as we suggest, current assessments of in the sense of conventional hydrocarbon systems, are
such areas may have substantially overestimated the thought to be of little importance (e.g., National Petro-
natural gas resource. Low-permeability settings similar leum Council [NPC], 1980; Law and Spencer, 1993,
to those found in the Rocky Mountain region contain in press; Surdam, 1997a; Schenk and Pollastro, 2002).
substantial amounts of gas; however, the fundamental Table 1 summarizes the attributes commonly associ-
elements of low-permeability reservoirs lead to a greater ated with basin-centered gas systems.
degree of risk associated with exploration and develop- Commercial production of natural gas from these
ment of this type of resource than has been widely rec- basin-centered gas accumulations is generally associated
ognized to date. with areas described as sweet spots that have improved
To properly frame this discussion, we examine pre- productivity and/or permeability (e.g., Billingsley and
vious work on low-permeability gas systems, in partic- Reinert, 1994). Surdam (1997a, p. 289), for example,
ular the assumptions that led to the development of characterized sweet spots ‘‘. . .as those reservoir rocks
basin-centered gas concepts. We review some of the that are characterized by porosity and permeability
underlying petrophysical concepts and offer additional values greater than the average values for tight sands at
insights based on our own work. We examine the con- a specific depth interval.’’ Because these reservoir rocks
trols on existing fields in the Greater Green River basin have low matrix permeability, it is widely perceived
(GGRB) (Figure 1) and review several large gas fields as that commercial production is strongly dependent on
well as a recent exploration program in the GGRB. the presence of open natural fractures and the ability
Finally, we discuss the implications of this work for to connect these natural fracture systems through hy-
resource assessments and risk analysis. draulic fracture stimulation (e.g., Spencer, 1989; Law
and Spencer, 1993; Schmoker, 1995; Surdam 1997a;
Popov et al., 2001). This has resulted in exploration
REVIEW OF LOW-PERMEABILITY BASIN- strategies and research programs oriented to finding
CENTERED CONCEPTS improved rock quality in the vicinity of natural frac-
tures within supposed gas-saturated portions of sedi-
Present Status and Understanding mentary basins.

Over the last 25 yr, basin-centered or continuous-type Development of Concepts


gas accumulations have come to be regarded as a dis-
tinct hydrocarbon system with unique characteristics Regions of widespread apparent gas saturation in low-
distinct from conventional oil and gas accumulations permeability reservoirs received attention during the

Shanley et al. 1085


1086 Gas Production from Low-Permeability Sandstone Reservoirs
Table 1. Summary of Characteristics Commonly Associated with Low-Permeability, Basin-Centered Gas Accumulations

Geographic area tens to hundreds of square miles


commonly in the more central, deeper portions of sedimentary basins
located in widespread gas-saturated regions
much larger than conventional oil and gas traps
Resource size very large in-place resource
low overall recovery factor
Relationship to water generally lack downdip water contacts; buoyancy is not a significant factor
generally located downdip of pervasive water-saturated rocks
water production is generally absent to very low
Trap boundaries structural and stratigraphic traps, in the conventional sense, are thought
to be of limited importance
Reservoir pressure overpressure and underpressure both common
Source rocks in close physical proximity to reservoir rocks
Reservoir permeability generally less than 0.1 md

late 1960s and early 1970s, when nuclear-based explo- for a stratigraphic or structural barrier between the
sives were used to fracture-stimulate low-permeability water and gas zones.’’ Although the updip and down-
reservoirs in the Piceance and San Juan basins. An ad- dip portions of the stratigraphic sections described by
ditional nuclear stimulation was contemplated on the Masters (1979) are in broad lithostratigraphic conti-
Pinedale anticline, in the GGRB of southwest Wyo- nuity at the scale of geologic formations, more recent
ming; however, disappointing results in the Piceance work in these basins has shown that individual fluvial
and San Juan basin projects coupled with health and sandstones and shallow-marine parasequences are not
safety concerns caused the Pinedale project to be can- in stratal continuity across the entire basin. As a result,
celled (e.g., Martin and Shaughnessy, 1969; Shaugh- considerable stratigraphic complexity separates areas
nessy and Butcher, 1974; Law and Spencer, 1993). of water saturation from areas of hydrocarbon satura-
The first clear description of a low-permeability gas tion. To explain the juxtaposition of water-bearing
province in terms that are commonly associated with strata that were thought to lie updip of gas-saturated
basin-centered systems is by Masters (1979), who de- reservoirs, Masters (1979) introduced the concept of a
scribed the deep, gas-saturated Cretaceous sandstone water block in which the relative permeability to gas
reservoirs of western Alberta, the San Juan basin in New would dramatically deteriorate at higher water satura-
Mexico, and Wattenberg field in the Denver basin of tions, rendering the reservoir rock incapable of
Colorado (the term basin-centered likely originated producing gas (Figure 2). The water block described
with Rose et al., 1986). Masters (1979) noted that the by Masters (1979) essentially formed the updip seal on
reservoirs in these basins had relatively low porosity large basin-centered gas accumulations. These ideas
and permeability (7 –15%, 0.15 –1.0 md), had mod- concerning the development of a water block continue
erate water saturations (34 – 45%), and were located in to the present time (e.g., Masters, 1984; Law and Dick-
the deeper portions of their respective basins. Masters enson, 1985; Spencer, 1985, 1989; Surdam 1997a; Law
(1979) also interpreted water-bearing strata structur- and Spencer, in press). Importantly, Masters’ (1979)
ally updip of gas-bearing strata. To describe the transi- description of water block introduced the geologic com-
tion from gas to water, Masters (1979, p. 156) stated munity to the potential importance of understanding
that ‘‘. . .the water-saturated section grades impercep- effective permeability and multiphase flow and indicated
tibly through a transition zone 5 to 10 mi wide into a that a strong petrophysical understanding would be re-
gas-saturated zone,’’ and that ‘‘. . .there is no evidence quired to decipher these low-permeability reservoirs.

Figure 1. Regional map of the Greater Green River basin (GGRB), southwest Wyoming, outlining general structure and well
locations. Locations of fields and key wells described in Table 3 and in the section titled ‘‘Trapping Mechanisms in Low-Permeability
Gas Accumulations’’ are shown with numbered circles and squares. Form-line contours drawn on top of the Mowry, contour interval
2000 ft (610 m).

Shanley et al. 1087


McPeek (1981) and Spencer (1985, 1987, 1989)
noted the apparent discrepancy between reservoir per-
meability measured in the laboratory (lower) and per-
meability calculated from well performance (higher).
They suggested that natural fractures were critical
to establishing economically significant production
and offered that these fractures might arise from the
thermal generation of natural gas. Law and Dickenson
(1985) and Law (2002b), among others, noted that
water production is relatively uncommon in many low-
permeability accumulations and suggested that vast
areas of gas-saturated rock are at irreducible water sat-
uration reflecting active gas generation and expulsion
of formation water. Spencer (1989) and Law and Spen-
cer (1993) further elaborated on the early ideas of water
block and suggested that because of the very low per-
meabilities associated with these gas accumulations, the
lack of water production, and the presumption of large
areas at irreducible water saturation, gas could not mi-
grate by buoyancy. As a result, they suggested that gas
accumulations would be widely distributed through
the reservoir-bearing interval with no discrete gas-water
contacts. Suggestions of large, gas-saturated regions
coupled with the idea that buoyancy is an ineffective
process led many authors to regard basin-center gas
provinces as comprising a unique but poorly under-
Figure 2. The concept of water block (Masters, 1979) has
stood petroleum system.
been used to explain how, within lithologically continuous
units, downdip gas-bearing strata could be trapped by updip Building on these paradigms, recent research has
water-bearing strata. In this model, water effectively provides focused on the origin, geometry, and detection of areas
the updip seal. with anomalous subsurface pressure, the detection of
areas of increased natural fracture density, and under-
standing the impact of source rocks on accumulation
Detailed work in the GGRB of southwest Wyo- type (e.g., Surdam, 1995, 1997a, b; Surdam et al., 1995,
ming led Law et al. (1980), McPeek (1981), Law and 1997a, b, 2000; Law, 2000, 2002b; ARI, 2001; Boswell
Smith (1986), Law (1984), Law and Dickenson (1985), et al., 2002a, b).
Spencer (1985, 1987, 1989), Pawlewicz et al. (1986),
Spencer and Mast (1986), and Law and Spencer (1989)
to suggest apparent relationships between thermal mat-
uration of coal-bearing strata, elevated subsurface tem- PETROPHYSICAL ATTRIBUTES OF LOW-
peratures, the development of widespread areas of both PERMEABILITY RESERVOIRS
anomalous pressure and gas saturation, and the devel-
opment of low-permeability reservoir rock. A series of Understanding the differences in petrophysical behavior
conceptual models were developed by Law and Dick- between traditional reservoirs and low-permeability res-
enson (1985) and were more recently modified by Law ervoirs is critical to predicting and understanding res-
(2000, 2002a, b) and Law and Spencer (in press). These ervoir performance in low-permeability gas systems.
models suggest that connate waters are expelled dur- Low-permeability sandstone reservoirs in the United
ing burial and diagenesis, and that maturation of source States (e.g., Dutton et al., 1993, 1995; Byrnes, 1997)
rocks in close proximity to low-permeability reservoirs are not dominated by ‘‘immature, muddy sandstones
results in elevated subsurface temperatures, abnormally with large volumes of diagenetically reactive detrital
high formation pressure, and regions of widespread, clay matrix, but rather are generally clean sandstones
present-day gas saturation. deposited in high-energy depositional settings whose

1088 Gas Production from Low-Permeability Sandstone Reservoirs


Table 2. Common Rock Property Values for Low-Permeabilty Sandstone Reservoirs in the Rocky Mountain Region

Depositional environment deep-water, paralic, fluvial


Porosity 5–10%
Stressed porosity 5–10%
Permeability (K a, Klinkenberg corrected) 0.1– 10 md
Stressed permeability (K a, Klinkenberg corrected) 0.001 –10 md
Stressed effective permeability (K eg, Klinkenberg corrected) 0.00001 –0.1 md
Water saturation (producible range) < 50%
Critical gas saturation 40 –70%
Irreducible water 20 –40%
Pore type slot pores dominate permeability
Pore-throat description (Port R35)* micro- to nanopore throats at R35 ( < 0.5 mm)
K/Phi 0.2– 1.0
Hg injection displacement pressure (Hg air) 80 –300 psia
*For a description of Port R35, see Hartmann and MacMillan (1992) and Hartmann and Beaumont (1999).

intergranular pores have been largely occluded by authi- md = 10  6 darcy). (2) The second type is sandstone
genic cements (mainly quartz and calcite)’’ (Dutton with highly altered primary pores occluded with authi-
et al., 1993, p. 5). Despite a complex burial history, genic cements (quartz or calcite). The majority of pores
original depositional fabric continues to exert a strong and pore throats are reduced to narrow slots connecting
control on productivity; those rocks that had more significant secondary pores created by grain dissolution
favorable porosity and permeability structure during (Dutton et al., 1993). On a volumetric basis, most of
deposition tend to remain the more attractive reser- the porosity in this pore structure occurs in secondary
voir targets despite a complex burial and uplift history. pores; however, flow paths are provided by the slot
Low-permeability reservoirs are characterized by high pores. This pore structure is by far the most common in
capillary entry pressures, high irreducible water sat- low-permeability reservoirs and dominates those reser-
urations, low to moderate porosity, and low permeabil- voirs with between 1- and 10-md permeability (Figure 3).
ity (e.g., Arastoopout and Chen, 1987; Byrnes, 1997). (3) The third type is muddy sandstone in which the
Table 2 summarizes petrophysical parameters com- intergranular volume is largely filled with detrital clay
monly found in many low-permeability gas accumula- matrix, and porosity is dominated by microporosity.
tions in the Rocky Mountain region. The vast majority This pore structure is common in many rocks with less
of low-permeability gas reservoirs require hydraulic- than 1 md permeability. This characterization contrasts
fracture stimulation to produce gas at economically ac- with earlier studies that suggested low-permeability
ceptable rates. reservoirs were dominated by grain-supported primary
porosity identical to more conventional reservoirs, but
Porosity and the Effect of Overburden Stress with pore throats plugged by authigenic clay minerals.
The effect of increasing overburden stress on po-
Low-permeability reservoirs are a unique rock type with rosity in reservoir-quality low-permeability sandstones
distinct properties that differentiate them from more is small (Figure 4). Byrnes (1997), for example, found
traditional reservoir rocks (e.g., Soeder and Randolph, that routine helium porosity values measured at am-
1987; Spencer, 1989; Soeder and Chowdiah, 1990; bient conditions tend to be within 95% of those mea-
Dutton et al., 1993, 1995; Byrnes, 1997, 2003; Cluff sured under in-situ conditions. Laboratory studies in-
et al., in press). Soeder and Randolph (1987) and Dutton dicate that the majority of pore volume compression
et al. (1993) suggested a threefold rock-type classifica- occurs with the first 2000 psi increase in net overburden
tion for low-permeability reservoirs. (1) The first type with progressively less additional decrease in pore vol-
is sandstone with open intergranular pores and pore ume with increasing stress (Figure 4). The minor re-
throats plugged by authigenic clay minerals. This pore sponse of pore volume to changes in confining stress in
structure is common in reservoirs with between 10- these low-permeability sandstones reflects the well-
and 100-md permeability (md = microdarcy = 0.001 cemented, rigid framework common to most of these

Shanley et al. 1089


Figure 3. Photomicrograph and scanning electron microscope images illustrating slot-type pores and pore throats commonly found
in low-permeability reservoirs. The slot-type pore network commonly consists of secondary, solution-derived pores that are connected
by narrow, sheetlike slots. At overburden stress, these narrow slots compress significantly, reducing permeability. (A) Frontier
Formation, Amoco Shute Creek 1, 10,779.8 ft (3285.6 m), 100, plane polarized light; (B) Williams Fork Formation, MWX 3, 5830 ft
(1777 m), 1400; (C) Travis Peak Formation, SFE 2, 8275.3 ft (2522 m), 100, plane polarized light; (D) Travis Peak Formation, SFE 2,
8275.3 ft (2522 m), 100, fluorescent light. Photographs for (B), (C), and (D) are provided courtesy of D. J. Soeder, U.S. Geological
Survey.

rocks and the fact that the slot pores that do compress ability (K a), or absolute permeability (K abs) (in this
under stress make up a minor portion of the overall article we will use K g to collectively refer to K g, K a,
pore volume. or K abs, unless otherwise noted). Although estimates
of K g are routinely used to guide decision making, they
Permeability, Relative Permeability, and the Effect of do not occur in natural, subsurface settings (Byrnes,
Overburden Stress 1997). In low-permeability sandstone reservoirs, in-situ,
high-pressure permeabilities relative to a gas phase
Routine core analysis consists of permeability measure- (K eg-at stress) range from 10 to 10,000 times less than
ments taken at relatively low pressure (approximately routine gas-permeability values (K g) measured in a
0–300 psia, commonly referred to as ambient condi- laboratory. This decrease is largely caused by the com-
tions) and under single-phase conditions (100% gas sat- bined effects of gas slippage (Klinkenberg correction),
uration, 0% brine saturation). These data are commonly confining stress, and partial brine saturation and its
referred to as routine gas permeability (K g), air perme- effect on effective permeability.

1090 Gas Production from Low-Permeability Sandstone Reservoirs


Figure 4. Plot of porosity and
permeability at varying levels
of net overburden stress illus-
trating the effect of overburden
stress on porosity and perme-
ability for 35 core samples. For
each core sample, three levels
of net overburden are shown:
ambient conditions, 2000 psia
overburden, and 4000 psia over-
burden. Samples with an ambient
permeability of between 1 and
10 md experience a reduction
in permeability by a factor of
approximately 0.3 –0.5. Sam-
ples with an ambient perme-
ability of less than 1 md may
experience a reduction in per-
meability of as much as two
orders of magnitude to less than
0.01 md. These data are from
the Frontier Formation in the
Moxa arch region of southwest
Wyoming.

Slippage, or Klinkenberg corrections (e.g., Bass, reduction in permeability as a function of increasing


1989), accounts for the difference in gas behavior at low stress occurs in those sands with the highest initial val-
pressure, such as a laboratory, vs. high pressure, such as ues of porosity and permeability. In low-permeability
the subsurface. These corrections can reduce routine sandstones, however, the greatest response to increasing
permeability measurements (K g ) by as much as a factor overburden stress occurred where porosity and perme-
of 3 for samples with routine permeabilities (K g) less ability were dominated by slot pores and pore throats.
than 1 d (Byrnes, 1997). Permeability measurements Byrnes and Keighin (1993, reported in Byrnes, 1997)
that have been adjusted for slippage effects are com- reported that pore throats in low-permeability sand-
monly referred to as equivalent liquid permeability, stones could decrease by 50–70% with increasing over-
K liq, K 1, or K inf. Slippage effects are reasonably well burden stress. Because of the inverse relationship be-
understood and will not be considered further. tween stress and permeability, overpressured settings
The differences in reservoir behavior between tra- are characterized by improved permeability relative to
ditional sandstone reservoirs and low-permeability res- normally or even subnormally pressured settings. This
ervoirs are largely caused by the combined effects of improved permeability reflects the reduction in effec-
overburden stress and partial brine saturation; we will tive stress because of increased formation pressures.
discuss each in some detail. It is well established that Although the vast majority of permeability data are
low-permeability reservoirs experience a dramatic de- routine measurements of K g, reservoir performance is
crease in permeability (K g) as a function of increasing governed by effective permeability at reservoir condi-
overburden stress (e.g., Thomas and Ward, 1972; Byrnes tions, and an understanding of relative permeability
et al. 1979; Jones and Owens, 1980; Dutton et al., 1993; behavior is critical. Effective permeability (K ei) is de-
Byrnes, 1997, 2003) (Figure 4). Figure 4 illustrates that fined as the permeability of a porous medium to a fluid
this effect is particularly pronounced for reservoirs with in the presence of one or more additional fluids (K ei;
a K g (at ambient conditions) of 0.5 md or less. In a study where i refers to water = K ew; gas = K eg; or oil = K eo).
of stress-dependent permeability, Davies and Davies Strictly speaking, relative permeability (K rw, K rg, K ro,
(1999) compared unconsolidated high-porosity and where the subscripts ‘‘w,’’ ‘‘g,’’ and ‘‘o’’ refer to water,
high-permeability sands with low-permeability gas sand- gas, and oil, respectively) is the ratio of the effective
stones. In unconsolidated sand reservoirs, the greatest permeability to a fluid at a given saturation to the fluid

Shanley et al. 1091


permeability at 100% saturation (e.g., Arps, 1964). In tion, there is less than 2% relative permeability to either
practice, there is little consistency in the literature as fluid phase, and critical water saturation and irreducible
to whether relative permeability is referenced to K g water saturation occur at very different water sat-
(100% gas saturation) or K liq (100% brine saturation), uration values. In these reservoirs, the lack of water
and in many cases, K rg is referenced to K g, whereas production cannot be used to infer irreducible water
K rw is referenced to K liq. The relative permeability of a saturation (Cluff et al., in press). In 1994, A. Byrnes (2002,
pore system depends on the nature of the fluids and personal communication) coined the term ‘permeability
their saturation, wettability, and the geometry of the jail’ to describe the saturation region across which there is
pore system. The distribution of each phase at a par- negligible effective permeability to either water or gas.
ticular saturation controls the flow of that phase irre- Failure to fully understand these relationships has led
spective of any others that are present (e.g., Almon to widespread misunderstanding as to how hydrocarbon
and Thomas, 1991; Dullien, 1992). systems are manifested in low-permeability reservoirs.
In low-permeability reservoirs, the impact of partial These effective permeability relationships suggest
brine saturation and overburden stress on reservoir the following: (1) The lack of water production from
performance is dramatic. It is not unusual in low- low-permeability reservoirs does not imply that the
permeability reservoirs for effective permeability to gas rocks are at irreducible saturation; it simply implies that
at conditions of overburden stress (K eg-at stress) to be one the water saturation is less than the critical water sat-
to three orders of magnitude less than routine perme- uration. There is a large range of water saturations over
ability (K g). Similarly, effective permeabilities to brine which the low-permeability structure of the reservoir
(K ew-at stress) are such that for many low-permeability may preclude any significant permeability to either gas
reservoirs, water is essentially immobile even at high or water. (2) The steepness of the relative permeability
water saturations. The relative permeability behavior to gas curve is such that small changes in brine saturation
of low-permeabilty reservoirs is such that traditional can result in significant changes in relative permeability.
concepts surrounding critical water saturation (the wa- In the subsurface, this means that wells can be drilled
ter saturation at which water ceases to flow), critical gas very close to economic relative permeability, and yet not
saturation (the gas saturation at which gas begins to produce significant gas or, in a few cases, produce some
flow), and irreducible water saturation (the water sat- formation water. (3) Regions where water saturations
uration at which further increases in capillary pressure exceed 50% are unlikely to have significant permeability
produce little to no additional decrease in water sat- to gas. (4) The effective permeability behavior of these
uration) must be reexamined. Investigations of relative low-permeability rocks should be regarded as a funda-
permeability to gas in low-permeability reservoirs illus- mental, distinguishing rock property; it cannot be al-
trate that gas permeability decreases rapidly at water tered or overcome by drilling and completion technol-
saturations above 40 –50% (e.g., Thomas and Ward, ogy. (5) Because of these effective permeability
1972; Byrnes et al., 1979; Jones and Owens, 1980; relationships, well tests have to be carefully examined
Walls, 1981; Walls et al., 1982; Smith, 1984; Byrnes, before test results can be attributed to variability in rock
1997, 2003; Cluff et al., in press), such that there is a permeability. Well tests that fail to produce fluids may
wide range of water saturation over which both gas and be derived from reservoirs with identical porosity and
water are virtually immobile (Figures 5, 6). Figure 5, permeability (K g) as those where well tests produced
for example, offers a comparison of traditional reser- significant volumes of gas. (6) Because low-permeability
voir behavior with low-permeability reservoir behav- reservoirs have little to no effective permeability to gas
ior. In a traditional reservoir, if we consider 2% relative at high water saturations, the gas that may be present at
permeability as a benchmark (i.e., Cluff et al., in press), these higher water saturations should not be considered
it is clear that there is relative permeability in excess of a resource. Failure to appreciate the unique behaviors of
2% to one or both fluid phases across a wide range of low-permeability reservoir rock with respect to effective
water saturation. In traditional reservoirs, critical water permeability can lead to serious misconceptions and fail-
saturation and irreducible water saturation occur at sim- ure to appropriately understand subsurface information.
ilar values of water saturation. Under these conditions,
the absence of widespread water production common- Capillary Pressure
ly implies that a reservoir system is at, or near, irre-
ducible water saturation. In low-permeability reservoirs, Low-permeability reservoirs are commonly character-
however, we find that over a wide range of water satura- ized by high to very high capillary pressures at relatively

1092 Gas Production from Low-Permeability Sandstone Reservoirs


Figure 5. Schematic illustration of capillary pressure and relative permeability relationships in traditional and low-permeability
reservoir rocks. Critical water saturation (S wc), critical gas saturation (S gc), and irreducible water saturation (S wirr) are shown. In
traditional reservoirs, irreducible water saturation and critical water saturation are similar. In low-permeability reservoirs, however,
irreducible water saturation and critical water saturation can be dramatically different. In traditional reservoirs, there is a wide range
of water saturations at which both water and gas can flow. In low-permeability reservoirs, there is a broad range of water saturations
in which neither gas nor water can flow. In some very low-permeability reservoirs, there is virtually no mobile water phase even at
very high water saturations.

Shanley et al. 1093


Figure 6. Plot of water satura-
tion and relative permeability to
gas (at overburden stress) for low-
permeability sandstone reservoirs
from several Upper Cretaceous
reservoir intervals in the Greater
Green River basin of southwest
Wyoming. Two data sets have
been merged in this figure. Data
from A. Byrnes are single-point
measurements of relative perme-
ability to gas at irreducible brine
saturation at estimates of specific
overburden stress (typically sam-
ple depth  0.5). Additional data
are multipoint measurements
of relative permeability to gas
at varying water saturations at
4000 psi overburden stress. These
data represent a wide range of
rock types; however, despite this
variation, sandstones with water
saturations greater than approx-
imately 50% have significantly
reduced relative permeability to
gas. Byrnes data set provided
courtesy of A. P. Byrnes, Kansas
Geological Survey.

moderate wetting-phase saturations. In many cases, determined by an understanding of pore and pore-
wetting-phase saturations of 50% are associated with throat geometry and its effect on multiphase flow. This
capillary pressures in excess of 1000 psia, suggesting that has led to a rock-catalog approach to reservoir descrip-
a large number of pore throats are less than 0.1 mm in tion, where macroscopic core observations are com-
diameter and are of the micro- to nanoscale (terminol- bined with petrographic and petrophysical information
ogy of Hartmann and Beaumont, 1999). The commonly (both wireline and laboratory measurements) to gain
used concept of irreducible saturation has limited utility, insights to flow capacity and structure (e.g., Sneider
particularly in low-permeability reservoirs (e.g., Dullien, et al., 1981, 1983, 1984; Hartmann and Coalson, 1990;
1992). In many low-permeability sandstone reservoirs, Hartman and MacMillan, 1992; Gunter et al., 1997;
wetting-phase saturation continues to decrease with in- Hartmann and Beaumont, 1999). Fundamental to this
creasing capillary pressure (e.g., Dullien 1992; Cluff, approach is the assumption that an understanding of
2002). A more pragmatic approach involves calibration pore geometry directly relates to permeability (K g) and,
of capillary pressure to wetting-phase saturation at par- by implication, effective (K eg) and relative permeability
ticular gas-column heights (e.g., Byrnes, 1997; Cluff, (K rg, K rw) (e.g., Wells and Amaefule, 1985). Capillary
2002). Comparison of capillary pressure data from a pressure measurements are commonly made to aid in
wide range of low-permeability gas reservoirs in the understanding pore geometry and pore-throat distribu-
GGRB suggests that gas-column heights required to tion (e.g., Yuan and Swanson, 1986; Spencer, 1989;
achieve acceptable levels of effective permeability to Vavra et al., 1992; Dutton et al., 1993). Although there
gas or to achieve typical irreducible saturation levels is a substantial body of literature that relates perme-
range from 300 –450 to more than 1000 ft (90 –140 to ability and relative permeability to capillary pressure,
more than 300 m). the notion of a relatively straightforward relationship
Understanding fluid flow from the rock matrix between capillary pressure and multiphase effective per-
into the wellbore or hydraulic fracture system is largely meability has come under scrutiny (e.g., Rose, 1999).

1094 Gas Production from Low-Permeability Sandstone Reservoirs


Figure 7. Core data from the Lewis Sandstone taken from two different wells in the Greater Green River basin; samples were
selected for having similar porosity and permeability (K g). These data illustrate the differences in relative permeability to gas at
overburden for rocks that have similar capillary pressure functions and, therefore, calculated pore geometry. A rock-catalog approach
would result in these three samples being classified as a single rock type, yet their performance is highly variable, particularly at
water saturations greater than 25%. FWL = free-water level, NOB = net-overburden stress.

Our own data suggest that although pore geometry and, capillary pressure measurements are normally recorded
hence, the rock-catalog approach may satisfactorily pre- at ambient confining stress, whereas we have already
dict routine gas permeability (K g), at ambient condi- discussed the extreme stress dependence of permeabil-
tions, there can be considerable uncertainty in the predic- ity in these reservoirs. Finally, it should be recognized
tion of both effective permeability to gas and relative that the capillary pressure data collected from samples
permeability to gas at conditions of overburden stress largely reflect the pore and pore-throat distribution at
particularly in low-permeability reservoirs. Figure 7, for the time of laboratory measurement. This may not be
example, illustrates that for the range of water satura- the pore and pore-throat distribution at the time of gas
tions that are commonly of interest in low-permeability migration and charge. In basins such as the GGRB that
reservoirs, there can be considerable variation in effective have experienced considerable relative uplift and where
permeability to gas for reservoirs with similar calculated gas migration and charge commenced before maximum
pore geometry and capillary pressure as measured at burial was reached, it is likely that simple extrapola-
ambient conditions. Although additional work is re- tions of capillary pressure data to hydrocarbon-column
quired, we suggest that as pore geometry becomes in- heights and saturations will be misleading. These factors
creasingly complex and dominated by slot pores and may explain the poor correlation that generally occurs
sheetlike pore throats that are not well described by between predictions of low-permeability reservoir be-
simple radii or cannot be regarded as a bundle of capil- havior based on rock-catalog solutions vs. estimates of
laries, fluid-flow properties deviate from classical theory, reservoir performance based on production logging
especially at increased overburden stress. Furthermore, (e.g., Al-Qarni et al., 2001).

Shanley et al. 1095


Fractures may impact the placement of hydraulic fractures via
stimulation methods. Furthermore, horizontal drilling
The role of natural fractures in contributing to im- programs in the GGRB that have deliberately targeted
proved deliverability of natural gas in low-permeability, potential fractures as a means of boosting gas produc-
basin-centered accumulations is poorly understood. In tion have failed to establish positive results (e.g., Bil-
many low-permeability reservoirs, it has been suggested lingsley, 1994; Branagan, 1994; Iverson and Surdam,
that gas-flow rates cannot be accounted for with the 1995; Blakeney-DeJarnett et al., 2001). Natural frac-
available permeability to gas at in-situ conditions (e.g., tures may enhance gas production if encountered in a
Spencer, 1989; Ammer et al., 2000), thereby requiring structurally high position within a trap. If, however,
the contribution of significant fracture permeability. fractures are encountered toward the base of a trap or
Observations such as these, coupled with the inherently outside of a trap, they are unlikely to improve gas pro-
low matrix permeability, have led to the widespread duction and may account for increased rates of water
working hypothesis that sweet spots are heavily depen- production.
dent on the development of a relatively dense network
of natural fractures (e.g., McPeek, 1981; Spencer, 1985, Petrophysical Summary and Implications for
1987, 1989; Law and Spencer, 1993, in press; Ammer Trapping Mechanisms
et al., 2000; Boswell et al., 2002b; Law, 2002b). Al-
though the argument in favor of natural fractures is The widespread occurrence of gas shows while drilling
well understood, actual documentation of natural frac- and limited examples of produced water have led to a
tures controlling deliverability is limited. In the Pice- model of basin-centered gas in which large portions of a
ance basin and the San Juan basin, researchers have basin are interpreted as being at irreducible water sat-
collected data supporting a positive correlation between uration. Within this gas-saturated region, vast resources
improved gas deliverability and the local occurrence are thought to exist, and productivity is conceptually
of natural fractures (e.g., Northrop, 1986; Finley and governed by the occurrence of sweet spots. With the
Lorenz, 1989, 1995; Lorenz and Finley, 1989; Lorenz expectation of little to no water and a large area at
et al., 1989; Raynolds and Pasternack, 1994; Harstad irreducible saturation, two essential strategies have
et al., 1995; Harstad and Teufel, 1998a, b; Al-Hadrami evolved. First, there has been a focus on drilling and
and Teufel, 2000). In other basins, such as the GGRB, completion technology in the supposed gas-saturated
the role of natural fractures is much less clear. In gen- region. This reflects an underlying assumption that lack
eral, studies in the GGRB have failed to establish a of commercial production reflected poor drilling prac-
positive correlation between improved gas deliverability tices or ineffective stimulation. Second, there is the goal
and the occurrence of natural fractures (e.g., Iverson and of finding better reservoir rock in the supposed gas-
Surdam, 1995; Shaver and Towler, 1997; Sturm et al., saturated part of the basin, particularly areas of in-
2000, 2001; Blakeney-DeJarnett et al., 2001, Horn and creased natural fractures. With the perspective that
Schrooten, 2001). Many of these studies conclude that basin-centered regions are a unique type of hydrocar-
primary depositional fabric is the predominant control bon system, there has been comparatively little focus
on gas deliverability. Our own work has reached similar on understanding trapping elements and other more
conclusions. In cases where geologic and petrophysical conventional aspects of the hydrocarbon system.
models of pore volume have been used as input to res- The most significant differences between conven-
ervoir simulation and history matching, there has been tional reservoirs and low-permeability reservoirs lie in
little need to modify matrix values of porosity and per- the low-permeability structure itself, the response to
meability (that is, the introduction of fracture perme- overburden stress, and the impact that the low-
ability) to achieve satisfactory results. We suggest that permeability structure has on effective permeability re-
fractures associated with widely spaced joint sets, on a lationships under conditions of multiphase saturation
scale of several meters to tens of meters, are likely; (Cluff et al., in press). Because of the effective per-
however, dense fracture networks resulting in locally meability structure of most low-permeability reservoirs,
improved subsurface permeability are largely undocu- there is a large range of water saturations over which
mented in the subsurface of the GGRB. Several studies both water and gas are essentially immobile. A lack of
have concluded that whereas the contribution of nat- water production (or recovery from a test) should not be
ural fractures to improved gas production is unclear, used to infer that the rocks are at, or near, irreducible
they do contribute to increased rates of water flow and water saturation nor should these regions be regarded

1096 Gas Production from Low-Permeability Sandstone Reservoirs


as water free. Instead, low-permeability reservoirs rocks ical laws described by buoyancy and capillary pressure.
should be regarded as having insufficient permeability to Instead, buoyancy and capillary response must be con-
either gas or water over a wide range of water satura- sidered in light of burial and uplift history, overburden
tions. Because these transition zones are commonly stress, and multiphase effective permeability. Examina-
manifested by wells that fail to produce either gas or tion of effective permeability to gas at varying water
water, drilling activity rarely extends sufficiently far saturations and at stress for rocks with similar pore
downdip to encounter the free-water level associated geometry suggests that the prediction of reservoir per-
with a given gas accumulation. Our work suggests that formance based on the similarity of pore geometry and
a water phase is indeed present; however, in many capillary pressure alone should be reexamined. Our
areas, because of the effective permeability structure data show that as the overall rock quality degrades, the
of the rocks, much of the water has limited mobil- variability in effective permeability can be dramatic.
ity (e.g., Breit and Skinner, 2002; Cluff et al., in press). The suggestion that low-permeability production sweet
Where water production in these low-permeability spots simply occur wherever slightly better rock prop-
reservoirs occurs, it is commonly associated with slight erties can be located in a supposed gas-saturated portion
improvements in reservoir quality or a loss in struc- of a basin is, in our opinion, without support. It is clear
tural position. Figure 8 illustrates the differences en- that low-permeability reservoirs are commonly com-
countered in drilling a low-permeability reservoir vs. a prised of a variety of rock types that are intercalated
more traditional one. both vertically and horizontally. Some of these rock
Much of the literature suggests that basin-centered types have reservoir properties more akin to conven-
gas regions have little to no water production; how- tional reservoirs, whereas others are dominated by the
ever, we have not found that to be the case. Consider- low-permeability reservoir properties discussed in this
ation of gas-water phase behavior at reservoir temper- paper. Interbedding of these rock types commonly
ature and pressure (dew-point considerations) indicates results in confusing fluid relationships.
that for the vast majority of Cretaceous reservoirs in the The petrophysical characteristics of low-permeabil-
GGRB, the amount of water that can be accounted for ity reservoirs described here are not confined to Cre-
as condensation in the gas stream ranges from less than taceous strata of the Rocky Mountain region. Byrnes and
0.15 to 1.5 bbl water/mmcf gas. Water production in Castle (2000) compared reservoir properties in Paleo-
excess of 0.15–1.5 bbl water/mmcf gas reflects the zoic low-permeability reservoirs of the Appalachian
production of free formation water. Figure 9 is a regional basin with Cretaceous low-permeability reservoirs of
map of the GGRB illustrating water production across Wyoming (largely from the GGRB) and found them to
the basin. Close examination of many of the large field be virtually indistinguishable in terms of reservoir prop-
areas clearly indicates that there are large numbers of erties. As a result, it is likely that low-permeability res-
wells producing formation water, and that in many ervoirs in general will exhibit reservoir behavior similar
cases, the water content increases in a downdip direc- to those found in the GGRB.
tion. Figure 10 is a cumulative frequency plot illustrat-
ing the distribution of wells and overall gas production
as a function of water/gas ratio. If water of condensation
is benchmarked at 0.15 bbl water/mmcf gas, only 16% TRAPPING MECHANISMS IN LOW-
of the gas wells in the GGRB have gas/water ratios less PERMEABILITY GAS ACCUMULATIONS:
than 0.15 bbl water/mmcf gas. These wells account for EXAMPLES FROM THE GGRB
approximately 25% of the total gas production in the
GGRB. If water of condensation is benchmarked at The GGRB of southwest Wyoming is a prolific gas
1.5 bbl water/mmcf gas, approximately 35% of the province. As of October 2002, the basin has produced
gas wells in the GGRB have gas/water ratios less than approximately 10.5 tcf gas, 193 million bbl oil (11.7 tcf
1.5 bbl water/mmcf gas. These wells account for ap- equivalent), and 84 million bbl water from approxi-
proximately 56% of the total gas production in the mately 7150 wells. Natural gas production occurs from
GGRB. Irregardless of the benchmark, water produc- reservoirs ranging in age from Pennsylvanian to Ter-
tion is both common and widespread in the GGRB. tiary, with most of the gas production occurring from
The petrophysical data discussed here suggest that low-permeability reservoirs of Late Cretaceous age.
low-permeability gas reservoirs are not a unique petro- Many of the low-permeability reservoirs are slightly to
leum system that somehow functions outside the phys- moderately overpressured (with respect to hydrostatic

Shanley et al. 1097


Figure 8. Schematic illustration highlighting relationships between capillary pressure, relative permeability, and position within a
trap, as represented by map and cross section views. In both cases (A) and (B), the map illustrates a reservoir body that thins and
pinches out in a structurally updip direction. The figure in (A) illustrates relationships found for a reservoir with traditional rock
properties. In conventional reservoirs, water production extends downdip to a free-water level (FWL). In the middle part of the
reservoir, both gas and water are produced, with water decreasing updip. The updip portion of the reservoir is characterized by water-
free production of gas. The figure in (B) illustrates relationships found for a reservoir with low-permeability reservoir properties. In low-
permeability reservoirs, significant water production is restricted to very low structural positions near the FWL. In many cases, the
effective permeability to water is so low that there is little to no fluid flow at or below the FWL. Above the FWL, a wide region of little to
no fluid flow exists. Farther updip, water-free gas production is found. Because of the wide region with little to no fluid flow, once
drilling encounters the wide transition zone with virtually no fluid production, drilling uncommonly extends downdip to a true FWL.

gradients), with pressure gradients (referenced to the tions suggest that gas migration from Cretaceous and
surface) ranging from approximately 0.43 to as much younger sources commenced approximately 40–50 Ma
as 0.85 psia/ft (9.793 to as much as 19.229 MPa/km). (e.g., Hanson et al., 1994, in press; Burch and Cluff,
From the perspective of petroleum systems, the GGRB 1997; Coskey, in press), and that relative uplift of
is a charge-rich basin with multiple source rock pack- between 2000 and more than 8000 ft (610 and more
ages, including both marine and nonmarine source rocks. than 2440 m) has occurred. Initial gas production in the
Reservoirs in the GGRB include deep-water turbidites GGRB was established in the early 1950s, and
and paralic and alluvial sandstones. Burial reconstruc- exploration and production operations continue to be

1098 Gas Production from Low-Permeability Sandstone Reservoirs


Figure 8. Continued.

active today. Because of widespread gas production gas fields involve stratigraphic traps contributing some
from low-permeability reservoirs, the GGRB has a rich 30% of the gas production, and 21% of the fields occur
and comprehensive database and has emerged as a in combination traps contributing approximately 20%
classic basin-centered gas province. of the gas production. Wells drilled outside of primary
In an attempt to better understand significant gas traps may record gas shows while drilling that are re-
accumulations in the GGRB, we examined all fields lated to either (1) encountering extremely small strat-
producing from Tertiary and Cretaceous reservoirs igraphic traps, generally caused by thin, isolated fluvial
with an expected ultimate recovery (EUR) greater than sandstones in largely continental strata, (2) encounter-
50 bcf (Figure 1; Table 3) (EUR from Barlow et al., ing low-saturation, residual gas in reservoirs that are in
1994). We concluded that 100% of these fields occur permeability jail (i.e., high water saturation), in which
in conventional traps. The cumulative production rep- the process of drilling destroys the pore system al-
resented by the gas fields outlined in Table 3 accounts lowing gas to be liberated, or (3) encountering gas in
for more than 90% of all gas production from low- the pore system of shales that are juxtaposed with
permeability reservoirs in the GGRB (production vol- sandstone reservoirs.
umes from Barlow et al., 1994). Approximately 38% of To illustrate the range of trap styles, we briefly
the gas fields involve structural traps accounting for review four fields and discuss the results of a recent
50% of the gas production from the GGRB, 41% of the exploration program in the GGRB. It is not our intent

Shanley et al. 1099


1100 Gas Production from Low-Permeability Sandstone Reservoirs
Figure 10. Cumulative frequency curve describing the distribution of gas wells and cumulative gas production as a function of
water/gas ratio expressed in barrels of water/million cubic feet of gas for gas wells producing from Tertiary –Cretaceous reservoirs in
the GGRB. Approximately 7888 gas wells contribute to this plot. Phase considerations suggest that the maximum volume of water that
can be accounted for by condensation in the gas stream ranges from approximately 0.15 to 1.5 bbl water/mmcf gas. Wells producing
at water/gas ratios in excess of these maximum values are clearly producing free formation water. If a benchmark of 1.5 bbl water/
mmcf gas is considered, approximately 35% of the gas wells in the GGRB have gas/water ratios less than 1.5 bbl water/mmcf gas.
These wells account for approximately 56% of the gas production in the GGRB.

to provide a detailed field study of each of these fields; Jonah field is a fault-bounded, structural trap located in
for that, the reader is referred to the literature. Instead, the northwestern part of the GGRB (Figures 1, 11). The
our intent is to use these fields as typical examples of field produces gas from slightly overpressured, lentic-
the types of traps that are both common and prolific ular fluvial sandstones of the Upper Cretaceous (Maas-
gas producers and that illustrate the conventional nat- trichtian) Lance Formation (e.g., Robinson et al., 1995,
ure of these low-permeability gas fields. We suggest 1996; Montgomery and Robinson, 1997; Robinson,
that the insights gained from the GGRB can provide 2000; Hanson et al., in press; Dubois et al., in press).
valuable analogs for other low-permeability reservoirs Through October 2003, Jonah field has produced ap-
both within and outside the Rocky Mountain region. proximately 891.1 bcf gas, 9.0 MMBO, and 2.8 million
bbl water from approximately 530 completed wells;
Jonah Field daily production as of October 2003 was approximately
705 mmcf gas and 6.3 million bbl oil and 4.3 million
The initial wells in what is now Jonah field were drilled bbl water (WOGCC). Gas-in-place estimates vary; how-
in 1986 and 1987; however, it was not until 1992 that ever, Dubois et al. (in press) has suggested as much
significant development and production began to occur. as 8.3 tcf gas in place.

Figure 9. Regional map of the Greater Green River basin of southwest Wyoming illustrating water production, in the form of water/
gas ratio for Cretaceous reservoirs. Data from approximately 6500 wells were used to compute barrels of water per million cubic feet
of gas. Water production volumes more than 1.5 bbl water/mmcf gas reflect the production of formation water. Water/gas ratios are
overlain on a form-line structure contour map on the Mowry Shale.

Shanley et al. 1101


Table 3. Summary of Fields with Expected Ultimate Recoveries Greater than 50 bcf Gas in the Greater Green River Basin,
Southwest Wyoming*

Production Discovery EUR


Field Name Interval Year (bcf gas) Trap Description Map Number

Barlow et al., 1994; GRI 94/0363


Arch Mesaverde 1959 97 stratigraphic-unconformity 1
Baggs South Lewis 1954 64 structural and stratigraphic 2
Barrel Springs Mesaverde 1966 63 stratigraphic-updip pinch-out 3
Baxter basin S Dakota 1922 101 structural-faulted anticline 4
Baxter basin S Frontier 1922 76 structural-faulted anticline 4
Big Piney-Labarge Frontier 1959 541 structural and stratigraphic 5
Big Piney-Labarge Mesaverde 1957 62 structural-updip thrust fault, 5
lateral strike-slip fault
Big Piney-Labarge Ft. Union 1925 174 structural-stratigraphic 5
Birch Creek Dakota 1957 141 stratigraphic 6
Birch Creek Frontier 1957 230 stratigraphic 6
Bird Canyon Frontier 1973 58 stratigraphic 7
Blue Forest-Lincoln Dakota 1983 311 stratigraphic-updip pinch-out 8
Road-Monument Butte-Swan
Brady Dakota 1973 62 structural-anticline 9
Bruff Dakota 1980 256 structural-stratigraphic 10
Bruff Frontier 1961 447 structural-stratigraphic 10
Canyon Creek Mesaverde 1954 266 anticline 11
Chimney Butte Frontier 1957 64 stratigraphic 12
Church Buttes Dakota 1946 496 anticline 13
Church Buttes Frontier 1978 79 structural-anticline 13
Coal Gulch Mesaverde 1977 80 stratigraphic across broad arch 14
Desert Springs Lewis 1958 236 updip stratigraphic pinch-out 15
Dripping Rock Mesaverde 1984 85 updip stratigraphic pinch-out 16
Dry Piney Frontier 1957 75 structural-anticline 17
Echo Springs Mesaverde 1977 472 stratigraphic across broad arch 18
Fabian Ditch Dakota 1980 120 structural-stratigraphic 19
Fabian Ditch Frontier 1976 68 structural-stratigraphic 19
Fontenelle Frontier 1973 124 stratigraphic 20
Green River Bend Frontier 1958 275 stratigraphic 21
Hay Reservoir Lewis 1976 160 updip stratigraphic pinch-out 22
Hiawatha Ft. Union 1926 324 anticline 23
Hiawatha West Lewis 1930 54 combination: structural-unconformity 23
Hogsback Frontier 1955 459 faulted anticline 24
Jonah Lance 1992 1200 fault-bounded structural trap 25
Lincoln Road Frontier 1977 72 stratigraphic 26
Lost Soldier Frontier 1916 51 structural-anticline 27
Luckey Ditch Dakota 1986 224 combination: structural-stratigraphic 28
Mesa Frontier 1976 83 stratigraphic 29
Nitchie Gulch Frontier 1961 128 combination: structural-stratigraphic 30
(draped across north plunge RSU)**
Patrick Draw Mesaverde 1959 84 updip stratigraphic pinch-out 31
Pine Canyon Dakota 1964 72 combination: structural-stratigraphic 32
(draped across north plunge RSU)
Pinedale Lance 1955 – 1997 1200 anticline 33

1102 Gas Production from Low-Permeability Sandstone Reservoirs


Table 3. Continued

Production Discovery EUR


Field Name Interval Year (bcf gas) Trap Description Map Number

Powder Wash Ft. Union 1931 286 structural-anticline 34


Standard Draw Mesaverde 1979 408 stratigraphic across broad arch 35
Sugarloaf Mesaverde 1953 84 structural-anticline 36
Table Rock Mesaverde 1954 257 structural-anticline 37
Table Rock/Table Rock SW Lewis 1954 66 structural and stratigraphic 37
Tip Top Frontier 1951 518 faulted anticline 38
Trail Unit Mesaverde 1958 94 structural-anticline 39
Wamsutter Mesaverde 1958 110 stratigraphic 40
West Side Canal Lewis 1964 186 combination: structural-stratigraphic 41
Whiskey Butte Frontier 1975 133 stratigraphic 42
Wild Rose Mesaverde 1974 185 stratigraphic across broad arch 43
Wilson Ranch Frontier 1972 53 stratigraphic 44
Subtotal –plays listed 11,614
Total all K-T plays in GRI 94/0363 12,638 Jonah and Pinedale added to Barlow
et al. (1994) GRI 94/0363 report
92%
*Fields largely taken from Barlow et al. (1994) with supplemental data from Cardinal and Stewart (1979) and Miller et al. (1992). Recent field developments at Jonah
and Pinedale postdate Barlow et al. (1994). Conservative EUR estimates for these fields have been added.
**RSU = Rock Springs uplift.

Hanson et al. (in press) describe the details of the gas ratio map for Jonah field (Figure 12). This map
Jonah trap. Lateral seals are provided by the west fault highlights the role of faults in both defining the gas
and the south Jonah fault, and a top seal is provided by accumulation at Jonah as well compartmentalizing the
an upper Lance shale that extends well beyond the field reservoir. Cluff and Cluff (in press) relate the downdip
boundaries. The west fault is an en echelon, strike-slip decrease in well performance to increasing water sat-
fault set that is near vertical and extends to the Pre- uration and, therefore, decreasing relative permeability
cambrian basement. Fault throws are up to approxi- to gas. Both Hanson et al. (in press) and Coskey (in press)
mately 300 ft (90 m). The west fault juxtaposes the suggest gas migration and charge occurred prior to com-
normally pressured Lance Formation on the west with plete burial. Coskey (in press) further suggests that the
the overpressured, commercially productive Lance For- overpressures observed today most likely reflect uplift
mation on the east. The south Jonah fault is a complex during the last 40 m.y. Similar conclusions regarding
zone of deformation containing multiple fault ele- the origin of overpressure in low-permeability settings
ments, characteristic of a dominantly strike-slip fault. were advanced by Katahara and Corrigan, 2002.
Despite the lack of significant vertical offset across Shanley (2001, in press) described the reservoir rocks
these faults, they form the lateral seals on the Jonah in Jonah field and the porosity-permeability relationships
accumulation, even where there is no observable ver- for cores taken both within Jonah field as well as imme-
tical displacement (Hanson et al., in press). Hanson et al. diately outside the field limits. Cores from nonproduc-
(in press) suggest that Jonah was charged through tive wells immediately outside the field have reservoir
buoyancy-driven migration that resulted in both grav- rock properties identical to those found within the field,
ity segregation as well as abrupt differences in gas negating the idea that productive fields have substantially
productivity between adjacent fault blocks internal to better rock properties than the surrounding host rock.
the field. Buoyancy-driven migration is further sup-
ported by the suggestion of a vertical decrease in water Hay Reservoir
saturation within the producing Lance Formation
(Dubois et al., in press). Cluff and Cluff (in press) studied Hay reservoir was discovered in 1977 and underwent
the petrophysics of Jonah field and presented a water/ significant development and production beginning in

Shanley et al. 1103


Figure 11. Structure contour map on the base of the lower Tertiary Fort Union Formation based on three-dimensional seismic data
at Jonah field, Sublette County, Wyoming. The outline of the 3-D survey is shown. The updip fault trap that is critical to the existence
of Jonah field is clearly illustrated in this map (map provided courtesy of D. Dubois, EnCana Oil & Gas, U.S.A.). The location of the
Corona 7-24 is also shown. A core taken in the Corona 7-24 recovered virtually identical reservoir properties to cores taken within the
field boundaries. Contour interval = 100 ft. Color bar is shown in the upper right corner.

September 1978 (Figure 1). The field produces gas the field. Mapping of movable hydrocarbon pore vol-
from overpressured, deep-water submarine fan depos- ume in the major lobes (F and G of Van Horn and
its in the Upper Cretaceous (Maastrichtian) Lewis Shannon, 1985, 1989) indicates that the most produc-
Shale that stratigraphically pinch out in an updip struc- tive wells are located toward the updip portions of
tural direction (e.g., Van Horn and Shannon, 1985, the sandstone lobe, with production quality decreas-
1989; Shannon, 1992; Anderson, 1994) (Figure 13). ing in a downdip direction. Reservoir porosity ranges
Through September 2003, Hay reservoir has produced from 5 to 12%, and permeability (K g) ranges from
approximately 149.3 bcf gas, 2.4 million bbl oil, and 0.01 to 0.35 md (Shannon, 1992; Anderson, 1994).
297 million bbl water from approximately 59 complet- In an attempt to address the interrelationship between
ed wells; daily production as of September 2003 is ap- capillary pressure, relative permeablity, and produc-
proximately 11.4 mmcf gas and 132 bbl oil, down from ible limits, Van Horn and Shannon (1985, 1989) con-
peak daily production rates of approximately 34.1 mmcf structed pseudo-capillary pressure and relative perme-
gas and 577 bbl oil in October 1993 (WOGCC). ability curves from wireline-derived water saturations.
Van Horn and Shannon (1985, 1989) recognized Although this work predated the understanding of
several discrete reservoir-bearing submarine fan lobes, permeability jail that exists today, their data clearly
each of which pinches out in an updip direction; two of indicated decreasing water saturations in an updip di-
these lobes, the F and G, are the principal reservoirs in rection. Similar observations regarding water saturation

1104 Gas Production from Low-Permeability Sandstone Reservoirs


Figure 12. Water/gas
ratio map for Jonah field
expressed in barrels of
water per thousand cubic
feet of gas. Dew-point
calculations at Jonah sug-
gest that approximately
0–1.5 bbl water can be
accounted for as water
of condensation. Water/
gas values in excess of
1.5 bbl/mmcf gas are in-
dicative of produced for-
mation water.

profiles were made at Hay reservoir by Henry et al. duction. Our work suggests that only the upper 750 ft
(1999) and by D. Wheeler (2003, personal communica- (230 m) of gas column may have effective permeability
tion) in an integrated study of Hay reservoir incorpo- to gas. Doelger and Morton (1999) observed the same
rating a high-quality three-dimensional seismic survey. downdip decrease in both production quality and gas
Consideration of EUR, hydrocarbon pore volume, and saturation seen by Van Horn and Shannon (1985, 1989);
water saturation indicates that the present downdip however, they interpreted this to reflect limited hy-
limit to production is approximately coincident with a drocarbon charge. We suggest that the decrease in pro-
water saturation of between 40 and 45% (D. Wheeler, duction quality cannot be used to comment on the effec-
2003, personal communication). This decrease in EUR tiveness of hydrocarbon charge; instead, the downdip
is well understood in light of Figure 6, which illustrates decrease in gas saturation and production quality re-
the change in relative permeability as a function of water flects the interplay of relative permeability and water
saturation in low-permeability reservoirs. Wells drilled saturation in a low-permeability gas reservoir. Figure 13
in a downdip position at Hay reservoir are still in the gas illustrates the trapping elements of Hay reservoir.
column; however, they lack effective permeability to
gas. Van Horn and Shannon (1985, 1989) suggested a Standard Draw-Echo Springs
gas column of approximately 2500 ft (760 m) at Hay
reservoir with only the upper 900 ft (270 m) having The Standard Draw-Echo Springs-Coal Gulch field area
sufficient effective permeability to gas to allow for pro- is located in the Washakie basin portion of the GGRB

Shanley et al. 1105


Figure 13. Map and cross section illustrating the nature of the updip stratigraphic trap in the principal reservoir at Hay reservoir.
(A) Structure map (250-ft [75-m] contour interval) overlain with an isopach map of gross reservoir sandstone (25-ft [7.5-m] contour
interval on bold contours) based on the integration of both three-dimensional seismic data and well control (courtesy of D. Wheeler).
Productive wells are focused near the southwestern limit of the sand, which pinches out in an updip direction. (B) Cross section with
a basal datum extending across Hay reservoir. Notice that the highest cumulative production occurs near the updip pinch-out of the
reservoir sand. To the northeast, reservoir thickness increases; however, production falls as water saturation increases, causing a
precipitous drop in effective permeability to gas.

(Figure 1). Field discoveries occurred during the pe- uous and are considered here to be a single field.
riod 1977 – 1978. Although Standard Draw, Echo Standard Draw-Echo Springs is an updip, stratigraphic
Springs, and Coal Gulch remain distinct production trap that is draped in a strike direction across the broad
units, they are geographically and geologically contig- Wamsutter arch (Figure 14). The field produces gas

1106 Gas Production from Low-Permeability Sandstone Reservoirs


from overpressured, shoreface, and barrier-island de- bodies ranging from approximately 10 to more than
posits (sometimes referred to as the Almond bar), as 40 ac (0.04 to more than 0.16 km2) in areal extent. The
well as from isolated fluvial sandstones (main Almond trapping element for the main Almond is provided by
or upper Almond) in the Upper Cretaceous (Maas- the isolated nature of the sand bodies themselves as well
trichtian) Almond Formation (e.g., Burch and Cluff, as a basinward facies change from coastal-plain deposits
1997; Horn and Schrooten, 2001; Horn et al., 2001). to shoreface and barrier-island deposits; the contact
Through October 2003, the Standard Draw-Echo between these two depositional facies is a transgressive
Springs field area had produced approximately 975.7 surface. Toward the base of the main Almond, the
bcf gas, 17.3 million bbl oil, and 4.2 million bbl water Almond interfingers with higher net/gross, coarser grain
from approximately 521 wells; daily production is ap- fluvial deposits of the Ericson. It is not uncommon for
proximately 155.9 mmcf gas, 1990 bbl oil, and 2233 bbl well tests of the laterally extensive reservoir bodies of
water. Expected ultimate recovery for the field area the Ericson to encounter significant water production.
exceeds 1.4 tcf gas. We interpret this water production to reflect the more
Burch and Cluff (1997) demonstrate a close rela- conventional reservoir properties associated with coars-
tionship between EUR, gas in place, storage capacity, er grained strata resulting in better effective permeabil-
and net sand thickness (Figure 14). With regard to the ity to water as well as a lack of trapping geometry in the
upper Almond shoreface and barrier-island reservoir, higher net/gross Ericson fluvial sandstones.
isopach maps of net sandstone as well as storage capac-
ity clearly delineate the field boundaries and demon- Swan-Blue Forest
strate the updip pinch-out of reservoir-quality sand-
stone. In general, wells with the greatest productive Swan-Blue Forest field is located on the Moxa arch in
capacity are located in areas with thick reservoir- the western part of the GGRB. The Moxa arch is a
quality sandstone along the crest of the Wamsutter south-plunging anticline extending 120 mi (190 km)
arch. Reservoir porosity ranges from 7 to 15%, and per- from the LaBarge platform in the north to the Uinta
meability ranges from 0.01 to 1.5 md. The downdip Mountains to the south (Figure 1). Swan-Blue Forest
extent of the upper Almond reservoir is controlled by was initially discovered in 1970 (Walters, 1992); how-
the extent of reservoir-quality rock. Burch and Cluff ever, significant field development did not occur until
(1997) noted that the downdip extent of reservoir- the mid-1980s. The primary reservoir in the field is the
quality rock equates to a horizontal surface, suggesting Lower Cretaceous Dakota and Muddy sandstones, in-
a paleo – gas-water contact and early gas charge. Basin terpreted as a northeast-southwest–trending transition-
modeling by Hanson et al. (1994) suggest gas charge al marine sandstone associated with the Shell Creek
between 40 and 50 Ma. The shoreface and barrier- transgression (Figure 15); Dolson et al. (1991) inter-
island trend can be mapped north and south of Stan- preted the Dakota and Muddy as a lowstand shoreface
dard Draw-Echo Springs; however, a change from deposit. The Dakota and Muddy reservoir is over-
kaolinite to illite cements and increasing water satura- pressured (0.53 psi/ft; 11.988 MPa/km), with almost no
tion result in a decrease in K g and K eg. These observa- associated water production. Through July 2002, Swan-
tions suggest continued diagenesis and loss of reservoir Blue Forest has produced approximately 312 bcf gas,
quality adjacent to the field following initial gas charge. 8.2 million bbl oil, and 194 million bbl water from ap-
This further supports the necessity of a conventional proximately 77 producing Dakota wells (data from IHS
trap for both emplacement and production of gas from Energy, Rocky Mountain Production Data CD). Aver-
these low-permeability reservoirs. age well performance amounts to 4.1 bcf gas, 106 mil-
In addition to the upper Almond reservoir, sig- lion bbl condensate, and 2.5 million bbl water; the best
nificant production at Standard Draw-Echo Springs has wells in the reservoir will ultimately recover in excess
developed beneath and to the west of the shoreface of 20 bcf gas.
sandstones in isolated fluvial sandstones of the main A map of the reservoir-bearing interval indicates
Almond (Horn and Schrooten, 2001; Horn et al., 2001). that the northeast-southwest –trending Dakota and
These reservoirs average approximately 0.75 bcf/well Muddy can be traced along depositional strike for a dis-
and enjoy favorable economics, especially given the tance of more than 12 mi (19 km) with a width of ap-
infrastructure of the upper Almond reservoir. The main proximately 4 mi (6 km). The seaward edge of the sand
Almond is a low net/gross fluvial system in which res- body is fairly straight, whereas the landward edge is
ervoirs are located in stratigraphically isolated sand more irregular. The reservoir-bearing sandstones parallel

Shanley et al. 1107


1108 Gas Production from Low-Permeability Sandstone Reservoirs
structural strike and pinch out to the northwest into shows while drilling and extensive efforts to stimulate
marine shales, making Swan-Blue Forest a classic com- the wells, significant production was not established in
bination trap in which the shoreface reservoir is draped any of the three wells. Data released on this drilling
across a prominent arch (Figure 15). The trap style ob- program indicate that all three wells encountered
served at Swan-Blue Forest is common for many of the reservoir-quality rock of similar porosity and perme-
fields on the Moxa arch. Reservoir quality in Swan-Blue ability to that found in the Jonah field (Bowker and
Forest is generally better than many low-permeability, Cotner, 1995). Furthermore, detailed analysis of vit-
basin-centered accumulations with porosity ranging rinite profiles indicated a similar burial history for the
from 15 to 18% and permeability ranging from 50 to three wells as is derived for Jonah field, suggesting that
125 md, with intervals exceeding 475 md (unstressed productivity at Jonah field is not related to enhanced
measurements). Because of the improved reservoir local gas generation or some anomalous thermal con-
quality, many original completions were natural, with dition. Available data suggest that although the wells
little to no additional stimulation or treatment. encountered gas shows, suggesting some degree of gas
charge, and comparable reservoir rock to that found in
Major Oil Company Exploration Program — Northern producing fields, they did not encounter a trapping con-
Greater Green River Basin figuration sufficient to cause a large accumulation to
exist. The gas shows are attributed to having encoun-
During the middle 1990s, a major oil company em- tered small, subeconomic stratigraphic traps in the
barked on a bold test of the basin-centered gas concept. nonmarine rocks of the Lance and Mesaverde intervals
After careful review of the available material on basin- as well as intervals with high water saturation that gave
center gas and studying available data from Jonah field up gas shows as the pore structure was destroyed dur-
(at that time an emerging field), the major company ing drilling. Tests of these intervals either resulted in
determined that a large part of the northern GGRB was depletion on test or encountered a lack of productivity,
gas saturated. The company secured leasehold interests most likely because of low effective permeability to
through participation in lease sales as well as farm-outs gas.
in a large area of Sublette County in the northwestern
part of the GGRB and drilled three wells: Billy 33-16
(NWSE Sec. 16, T31N, R111W, TD 12,355 ft), Ross
33-7 (NWSE Sec. 7, T30N, R110W, TD 12,120 ft), ASSESSMENTS OF GAS
and Paradise 23-31 (NESW Sec. 31, T31N, R109W, RESOURCES IN LOW-PERMEABILITY,
TD 12,270 ft). Because of the apparently pervasive gas ROCKY MOUNTAIN SANDSTONES
saturation, well locations were biased by considerations
of infrastructure as opposed to considerations of trap Resource assessments of natural gas attempt to quantify
and seal (Bowker and Cotner, 1995). All three wells a state of knowledge regarding a geologic province, at
penetrated the Mesaverde Formation and drilled equiv- a particular point in time, and provide guidance for
alent stratigraphic sections to the producing intervals at public policy issues such as national energy policy and
the nearby Jonah field (Figure 1). Despite scattered gas land-use planning. The Rocky Mountain region is broadly

Figure 14. Maps illustrating the nature of the updip stratigraphic trap that controls the vast majority of production from the upper
Almond reservoir in the Standard Draw-Echo Springs trend (modified from Burch and Cluff, 1997). (A) Inset from a regional structure
map across the Wamsutter arch illustrating the location of the Standard Draw-Echo Springs area along the crest of the arch. The north
and south limits of the field occur at a structural elevation of approximately  3100 ft (  945 m) subsea. (B) Isopach map of net
sandstone thickness in the upper Almond with greater than 10% effective porosity. Contour interval is 4 ft (1.2 m). This map clearly
identifies the trend of the upper Almond parasequences and illustrates both the basinward (structurally updip) pinch-out into marine
shales and siltstones, as well as the landward (structurally downdip) pinch-out into coastal-plain deposits. (C) Isopach map of gas in
place for the upper Almond reservoir-bearing interval. Note the abrupt truncation at both the north and south ends of the field
caused by high water saturation and low effective permeability to gas. These limits may reflect a paleo –gas-water contact. Contour
interval is 4 mmcf gas/ac. These maps illustrate a classic stratigraphic trap geometry in which reservoir-quality sandstones pinch out
in a basinward direction that is presently in an updip structural position. Additional drilling into the deeper main Almond has
occurred throughout much of the Standard Draw-Echo Springs area. Production from these reservoirs is found in stratigraphically
isolated fluvial- and estuarine-channel deposits.

Shanley et al. 1109


Figure 15. Map and
cross section of Swan-Blue
Forest field located on
the Moxa arch of south-
west Wyoming. (A) Iso-
pach map of reservoir
thickness overlain on a
structure map. Wells pro-
ducing from this interval
are highlighted; majority
of the remaining wells
are producing from the
shallower Frontier Sand-
stone. This map illustrates
the stratigraphic nature
of the trap. To the south-
west of this field, wells
drilled at elevations sim-
ilar to the base of pro-
duction have encountered
significant water in the
Dakota interval. (B) Cross
section of the Dakota
interval illustrating the
facies change to the
southeast of Swan-Blue
Forest that controls the
distribution of reservoir-
quality sand.

characterized by low-permeability reservoirs, wide- inces. There are two primary assessment techniques
spread gas shows while drilling, and limited produced currently in use for these basin-centered areas. One as-
water. Gas production is generally viewed as being con- sessment technique is to derive an understanding of gas
trolled by sweet spots in relatively homogenous, gas- in place through consideration of hydrocarbon pore
saturated host rock as opposed to being dominated by volume. This is broadly the approach followed by the
conventional traps similar to more traditional oil and Department of Energy and the National Energy Tech-
gas provinces. These observations have supported a view nology Laboratory (DOE and NETL) (Boswell et al.,
that large portions of these basins are gas saturated, 2002b). The other approach is to first determine wheth-
which necessitated an evaluation methodology differ- er a petroleum system contains a continuous gas (i.e.,
ent from those used in conventional oil and gas prov- basin-centered) accumulation. If gas accumulations are

1110 Gas Production from Low-Permeability Sandstone Reservoirs


thought to be continuous, then a more statistical view In an effort to be more scientifically rigorous and to
of the basin is taken in which cells are aggregated and provide guidance for public policy, the U.S. Geological
integrated with concepts related to production rate, Survey undertook and completed a national assess-
drilling success, drainage area, and likelihood of a sweet ment of petroleum resources, including unconvention-
spot occurring to produce an estimate of the gas re- al, continuous-type gas accumulations (Gautier et al.,
source. This is broadly the approach followed by the 1995, 1996). Within the category of continuous-type
U.S. Geological Survey in their assessments of contin- accumulations, resource estimates were based on sub-
uous gas provinces. In low-permeability reservoirs such division of an area into cells in which cell sizes reflected
as those found in many of the Rocky Mountain basins, the median drainage area for producing wells in the
the consequences of using either approach have in- play type under investigation. The cells were assigned a
evitably led to very large estimates of either gas in distribution of potential gas recovery volumes at vary-
place or technically recoverable gas (see Table 4 for a ing probabilities of success. For many of the continuous-
summary of resource estimates). type accumulations for which resource assessments
One of the earliest efforts to quantify the resource were done, the average probability of success (of finding
associated with low-permeability gas accumulations technically recoverable gas) exceeded 65% (e.g., Gau-
in the Rocky Mountain region was by the Supply- tier et al., 1995, 1996; Attanasi, 1998; Dyman et al.,
Technical Advisory Task Force (1973), on behalf of 1998). Economic scenarios were then constructed to
the Federal Power Commission. This effort suggested determine how the recoverable resource varied with
large volumes of gas in the GGRB (240 tcf gas), the changing cost of supply considerations (e.g., Attanasi,
Piceance basin (149 tcf gas), and the Uinta basin (207 1998). The suggestions of this work were clear. The
tcf gas). Kuuskraa et al. (1978) estimated a resource of quantity of recoverable gas was thought to be very
241 tcf gas in place in approximately eight Rocky large and varied directly with product price as well as
Mountain basins, with some 91 tcf gas in place in the improvements in drilling and completion technology.
GGRB alone. Also in 1978, the Federal Energy Reg- In the Rocky Mountain region, the U.S. Geolog-
ulatory Commission reviewed previous estimates of ical Survey suggested a mean recoverable resource of
low-permeability resources (Supply-Technical Advisory 160.5 tcf gas, 568 million bbl oil, and 1829 million bbl
Task Force, 1978) and augmented them with approxi- of natural gas liquids (NGL) across four sedimentary
mately 63 tcf gas in place in the San Juan basin, bringing basins in unconventional, continuous-type accumula-
the total resource to some 663 tcf gas in place in the tions hosted in sandstone reservoirs. The U.S. Geological
Rocky Mountain region. Survey suggested an average project success rate in ex-
In 1980, the NPC estimated the resource in tight- cess of 60%, leading to a risked resource of approxi-
gas reservoirs (defined as reservoirs with less than mately 96 tcf gas, 341 million bbl oil, and 1097 million
1.0 md permeability) at approximately 256 tcf gas in bbl NGL. This was followed in 1996 by an assessment
place across six basins, with 136.1 tcf gas in place in of the Wind River basin (Johnson et al., 1996) that
the GGRB alone. The NPC (1980) estimated econom- identified a mean in-place resource of approximately
ically recoverable gas to range from 35.2 to 50.2 tcf 995 tcf gas. More recently, the U.S. Geological Survey
gas in these same six basins (base case and advanced has conducted additional detailed geologic studies and
case at $2.50/mmcf ) with 3.1 –12.3 tcf recoverable gas new assessments of several key basins, including those
estimated for the GGRB. An early assessment of low- basins with large unconventional resource potential.
permeability resources in the Piceance basin by the These studies suggest that continuous-type sandstone
U.S. Geological Survey (Johnson et al., 1987) suggested reservoirs contain mean, undiscovered resources of ap-
423 tcf gas in place, and a similar assessment for the proximately 80.6 tcf gas and 2500 million bbl NGL in
GGRB (Law et al., 1989) suggested 73 tcf gas to be the Green River basin of southwest Wyoming, 18.8 tcf
recoverable under the current conditions and 5063 tcf gas and 33.4 million bbl NGL in the Uinta and Piceance
gas in place for the GGRB as a whole. More recent basins, and 26.2 tcf gas and 144.4 million bbl NGL in the
efforts to describe the low-permeability resources in San Juan basin (Kirschbaum et al., 2002a, b; Ridgley
the GGRB include a revision of previous work by the et al., 2002). The U.S. Geological Survey has extended
NPC, which suggested an undiscovered resource of their assessment of continuous-type accumulations out-
56.21 tcf gas (NPC, 1992), and the Gas Technology side the Rocky Mountain region. In the Appalachian
Institute (GTI), which estimated an undiscovered re- basin, for example, Milici et al. (2003) has suggested
source of 55.9 tcf gas (Doelger and Morton, 1999). mean, undiscovered gas resources of approximately

Shanley et al. 1111


Table 4. Summary of Various Resource Assessments Focused on Low-Permeability, Basin-Centered Gas Accumulations

Report Date Report/Assessment/Reference Region Resource Assessment*

1973 Federal Power Commission; three Rocky Mountain basins 596 tcf gas in place in three basins
Supply-Technical Advisory
Task Force (1973)
Greater Green River = 240 tcf
Piceance basin = 149 tcf
Uinta basin = 207 tcf
1978 Kuuskraa et al. (1978) eight Rocky Mountain basins 241 tcf gas in place across eight basins
Greater Green River = 91 tcf
Piceance basin = 36 tcf
Uinta basin = 50 tcf
San Juan basin = 15 tcf
Wind River basin = 3 tcf
Big Horn basin = 24 tcf
Douglas Creek arch = 3 tcf
Denver basin = 19 tcf
1978 FERC 1978 modification to FPC; four Rocky Mountain basins 663 tcf gas in place in four basins
Supply-Technical Advisory
Task Force (1973)
Greater Green River = 240 tcf
Piceance basin = 150 tcf
Uinta basin = 210 tcf
San Juan basin = 63 tcf
1980 National Petroleum six Rocky Mountain basins 256 tcf gas in place in six Basins;
Council (1980) 35.2 – 50.2 tcf recoverable
Greater Green River = 136.1 tcf;
3.1 – 12.3 tcf recoverable
Piceance basin = 49.1 tcf; 12.8 tcf
recoverable
Uinta basin = 20.5 tcf; 12.2 – 14.7 tcf
recoverable
San Juan basin = 3.3 tcf; 0 – 1.5 tcf
recoverable
Wind River basin = 33.7 tcf; 7 – 8.7 tcf
recoverable
Denver basin = 13.3 tcf; 0 tcf
recoverable
1987 U.S. Geological Survey; Piceance basin 423 tcf gas in place
Johnson et al. (1987)
1989 U.S. Geological Survey; Greater Green River basin 73 tcf – current technology
Law et al. (1989)
5063 tcf gas in place
1992 National Petroleum Greater Green River basin 56.2 tcf undiscovered resource
Council (1992)
1995 – 1996 U.S. Geological Survey National four Rocky Mountain basins 160 tcf gas, 568 MMBO, 1829 million
Assessment; Gautier et al. (1995, bbl NGL technically recoverable
1996); estimates shown here resource in five basins
do not include any resources
attributed to fractured shale
or coal-bed gas resources

1112 Gas Production from Low-Permeability Sandstone Reservoirs


Table 4. Continued

Report Date Report/Assessment/Reference Region Resource Assessment*

Greater Green River = 5 plays, 119.2 tcf


mean resource; play success
rate = 50 – 70%
Uinta and Piceance basins = 6 plays,
16.7 tcf mean resource; play success
rate = 20 – 88%
San Juan basin = 3 plays, 21.0 tcf mean
resource; play success rate = 50 – 60%
Denver basin = 2 plays, 3.07 tcf,
226 MMBO mean resource; play
success rate = 50 – 60%
1996 U.S. Geological Survey – Wind River basin mean in-place resource of 995 tcf, F95 of
Johnson et al. (1996) 603 tcf, F5 of 1530 tcf in 22 identified
play opportunities
1999 Gas Technology Institute Greater Green River basin 55.9 tcf undiscovered resource
2000 Gas Research Institute Rocky Mountain region new-field potential = 206 tcf
2001 Potential Gas Committee four basin areas = Greater 69.6 tcf gas most likely
Green River basin, Uinta resource in four basins
and Piceance basins, Wind
River basin, San Juan basin
estimates shown here are total Greater Green River = 17.6 tcf
most likely resources between most likely
0- and 15,000-ft (0- and 4570-m)
drilling depth. PGC does not
specifically address low-
permeability resources
Uinta and Piceance basins = 31.0 tcf
most likely
San Juan basin = 12.8 tcf most likely
Wind River basin = 8.2 tcf most likely
2002 DOE and NETL; Boswell Greater Green River basin 3013 tcf gas in place in Greater
et al., 2002b and Wind River basin Green River basin in six play intervals
1332 tcf gas in place in Wind River basin
in seven play intervals
2002 U.S. Geological Survey – Kirschbaum three basin areas = Greater 125.4 tcf gas, 2678 million bbl NGL mean
et al. (2002a, b), Ridgley Green River basin (southwest recoverable resource in three basins
et al. (2002) Wyoming), Uinta and Piceance
basins, San Juan basin
estimates shown here do not Greater Green River = 80.6 tcf, 2500 million
include any resources attributed bbl NGL mean resource in seven
to fractured shale or coal-bed assessment units
gas resources
Uinta and Piceance basins = 18.8 tcf,
33.4 million bbl NGL mean resource in
seven assessment units
San Juan basin = 26.2 tcf, 144.4 million
bbl NGL mean resource in five
assessment units

*NGL = natural gas liquid.

Shanley et al. 1113


57.6 tcf gas and 833 million bbl NGL in continuous- much of the gas was not economic under the assump-
type accumulations associated with sandstone and tions presented. The implication of this work, however,
shale reservoirs; in excess of 75% of these resources are was clear. Because these were perceived to be contin-
assigned to low-permeability sandstone reservoirs. uous gas accumulations, improvements in drilling and
The U.S. Geological Survey is not alone in the completion technology or product price would allow
very large resource estimates assigned to these low- additional technically recoverable gas to become eco-
permeability targets in the Rocky Mountain region. The nomic, and the nation was richly endowed with tech-
Gas Research Institute (GRI) has estimated a new field nically recoverable gas in low-permeability strata. Most
gas potential in low-permeability reservoirs in the Rocky forecasting groups in the United States (including the
Mountain region to exceed 206 tcf gas (Cochener, Department of Energy and Energy Information Agency)
2000). The Potential Gas Committee (PGC) does not have assumed that a very large volume of unconven-
specifically address gas resources in low-permeability tional, basin-centered natural gas is available at relatively
sandstone reservoirs; however, they have suggested high probabilities of success. The starting volumes for
a most likely resource of 18.4 tcf gas in the GGRB, many of these econometric models are similar to or
30.6 tcf gas in the Uinta and Piceance basins, 12.2 tcf based on the mean resource value suggested by the U.S.
gas in the San Juan basin, and 6.7 tcf gas in the Wind Geological Survey derived from its assessment efforts
River basin (PGC, 2003), much of which is thought (e.g. EIA, 2003).
to be contained in low-permeability, so-called basin- Our work in the GGRB, however, suggests a very
centered accumulations. A recent review by the GTI different set of conclusions regarding resource estimates.
suggested that in 1998, production from tight-gas sands The widespread occurrence of gas shows while drilling
in the United States comprised almost 70% of gas pro- and the lack of produced water does not imply a vast
duction from all unconventional gas resources and ac- area of gas-saturated rock. Many of these gas shows
counted for as much as 19% of total gas production from reflect sandstones with minor, residual gas saturation
all sources (Prouty, 2001). This same study suggested (i.e., high water saturation) and negligible effective
that total producible tight-gas resources in the United permeability to gas in which drilling has destroyed the
States exceed 625 tcf gas, with an economically recov- pore system liberating minor amounts of gas as shows.
erable resource of almost 185 tcf gas. Recognizing the Other gas shows record the presence of a large number
difficulties inherent to resource estimation related to of very small stratigraphic traps. We suggest, based on
economic considerations, the DOE and NETL recently the geometry of the fluvial sand bodies, that many of
announced a gas-in-place estimate for the GGRB of these small traps are likely to contain less than 0.1 bcf
3013 tcf gas and a gas-in-place estimate for the Wind gas in place and may well contain less than 0.05 bcf gas
River basin of 1332 tcf gas (Boswell et al., 2002a, b). in place. The geologic conditions that cause this have
Virtually every effort to quantify gas resources in low- already been described. Because of relative permeabil-
permeability formations, particularly in the Rocky ity considerations, unless reservoirs are charged to
Mountain region has concluded that very large natural water saturations typically less than 40 – 50%, there is
gas resources remain to be exploited. insufficient effective permeability to gas to allow for
As a result of these very large resource estimates, gas production under virtually any economic scenario.
low-permeability, basin-centered resources are expected Finally, our work suggests that virtually all significant
to become a significant part of the nation’s future gas fields in the GGRB occur in conventional traps.
energy portfolio. The GTI, for example, suggests that The only fundamental distinction between the con-
year-end gas reserves in low-permeability rocks will ventional traps in the GGRB and those that are de-
more than double by the year 2015 (Cochener, 2000), scribed in conventional oil and gas provinces is that the
and that annual natural gas consumption could rise to reservoir rock in the GGRB has low permeability.
32.7 tcf by 2015, accounting for 28% of the United Resource assessments predicated on the notion
States’ energy consumption (Prouty, 2001). of a continuous gas accumulation have likely over-
estimated the size of natural gas resources in low-
Implications of this Work to Resource Assessments permeability reservoirs in the GGRB by a substantial
amount. Our work illustrates that (1) the distribution
The initial economic modeling of Attanasi (1998) sug- of gas fields is controlled by an understanding of con-
gested that although there might be large technically ventional traps; (2) natural gas located in reservoirs at
recoverable gas resources in low-permeability strata, high water saturations (i.e., permeability jail) cannot

1114 Gas Production from Low-Permeability Sandstone Reservoirs


be recovered because of effective permeability consid- al fabric, pore geometry, and routine vs. effective
erations, unless the pore system is destroyed (i.e., drilled permeability, all under conditions of overburden
and crushed), and the gas is liberated; (3) as in all con- stress, are required to predict low-permeability res-
ventional oil and gas provinces, there are large areas that ervoir performance. A better understanding of the
lack significant trapping geometries and therefore are relationship between rock fabric and gas produc-
unlikely to hold significant volumes of gas; and (4) sim- tivity requires careful investigations into multiphase
ilar to all oil and gas provinces worldwide, there are a permeability under conditions of varying water
large number of very small gas fields, too small to sup- saturation and net-overburden stress, as well as an
port full-cycle exploration and development. In this re- analysis of capillary pressure and net-overburden
gard, we suggest that the gas resources in the GGRB have stress. The lack of widespread water production
a field-size distribution similar (in concept) to con- does not imply that vast areas of a sedimentary basin
ventional oil and gas regions. are at irreducible water saturation. Instead, it im-
An understanding of the potential resources in the plies a complex, effective permeability-to-gas rela-
GGRB, as well as other basins thought to contain basin- tionship. Water production, at levels well in excess
center gas resources, cannot rest on the continuous-type of condensation from the produced gas stream,
evaluation strategy. Models that incorporate discovery occurs in many gas fields in the GGRB.
process, field-size distributions, and an appreciation for 3. Low-permeability reservoir systems such as those
the elements needed to assess source, migration, reser- found in the GGRB are not examples of basin-center
voir, trap, and seal are needed to more clearly under- or continuous-type accumulations, nor are they a
stand the resource potential of the GGRB and other unique type of petroleum system as some have sug-
low-permeability regions. Our work should not be taken gested. We suggest that the only truly continuous-
to suggest that exploration and production opportuni- type gas accumulations are to be found in hydrocar-
ties in these low-permeability settings are lacking. On bon systems in which gas entrapment is dominated
the contrary, substantial portions of these basins re- by adsorption. Perhaps the most clearly understood
main lightly explored, and these regions have attractive and well-documented examples include coal-bed
source, reservoir, and seal considerations that warrant methane, some oil-prone source rocks, and some
continued exploration. The understanding of a diverse organic-rich shales.
set of trap styles that can occur in basins such as the 4. Resource assessments of these regions have assumed
GGRB should provide ample opportunity for explora- that a continuous, gas accumulation exists across a
tion and development. large area locally interrupted by the development of
sweet spots. Our work suggests, however, that this
viewpoint is at odds with the reservoir characteristics
CONCLUSIONS AND FUTURE DIRECTIONS of low-permeablity reservoirs, and we find little sup-
port for this model. Much of the gas that is in low-
1. Exploration efforts in low-permeability settings must permeability reservoirs is in subeconomic traps
be deliberate and focus on fundamental elements of (isolated, small point-bar sandstones, for example)
hydrocarbon traps. We suggest that significant gas or at gas saturation levels too low to have significant
reserves will not be extracted simply through im- effective permeability. Changes in drilling technol-
proved completion and drilling technology or by im- ogy, completion technology, and product price are
proved product prices. Improvements in completion not going to change effective permeability. Signif-
and drilling technology will allow well-identified icant production is dependent on the presence and
geologic traps to be fully exploited, and improve- identification of conventional traps. Existing re-
ments in product price will allow smaller accu- source estimates, therefore, are likely to have been
mulations or lower rate wells to exceed economic substantially overestimated.
thresholds. These elements are true in virtually every 5. Resource assessments in these low-permeability
petroleum province. basin-centered regions must recognize the reservoir
2. Petrophysics is a critical technology required for properties inherent to these rocks and should inte-
understanding low-permeability reservoirs. Low- grate the necessary concept of source, reservoir, trap,
permeability reservoirs behave in a fundamentally seal, migration, and charge. As a result, resource as-
different manner than more traditional reservoirs. sessments of these low-permeability, basin-centered
Understanding the relationships between deposition- areas should be done in a manner consistent with the

Shanley et al. 1115


assessment of conventional oil and gas systems. Field- Ammer, J. R., G. L. Covatch, G. P. Sames, T. H. Mroz, and W. J.
Gwilliam, 2000, Technology for increased production from
size distributions, an understanding of the discovery low-permeability gas reservoirs: An overview of U.S. DOE’s
process, analog studies, and risk assessment of the Gas Program — Successes and future plans: 2000 Society of
independent variables central to petroleum systems Petroleum Engineers Rocky Mountain Regional/Low Perme-
ability Reservoirs Symposium, Denver, Colorado, March 12 –
must be incorporated; they should no longer be
15, 2000, SPE Paper 60310.
assessed in a purely statistical manner. The available Anderson, G. P., 1994, A field study of the Hay reservoir area, an
data clearly indicate that the accumulations are analog for exploration of Lewis reservoirs, in S. A. Sonnenberg,
neither random nor simply the occurrence of some- compiler, First Biennial Conference — Natural gas in the
western United States: Lakewood, Colorado, Rocky Mountain
what improved rock quality, nor simply the result Association of Geologists and Colorado Oil and Gas Associ-
of improved drilling and completion technology. ation, October 17 – 18, 1994, Extended Abstracts.
6. The nature of the petroleum systems within basins Arastoopout, H., and S. T. Chen, 1987, Analysis of flow of gas and
water in a fractured and non-fractured low-permeability
such as the GGRB is not well understood nor well reservoir under production: 1987 Society of Petroleum
documented. Early models suggested that organic Engineers Annual Technical Conference, Dallas, Texas.,
deposits intercalated with reservoirs are the source September 27 – 30, 1987, SPE Paper 16948.
ARI (Advanced Resources International, Inc.), 2001, Federal lands
rocks, and that generation rates in excess of leakage
analysis, natural gas assessment, southern Wyoming and north-
(diffusion) are responsible for the overpressures western Colorado — The Greater Green River basin study:
observed in many of these types of basins today. http://fossil.energy.gov:7778/programs/oilgas/publications/fla/
(accessed October 2003).
Although these models are appealing, due to sim-
Arps, J. J., 1964, Engineering concepts useful in oil finding: AAPG
plicity, they do not accurately address many aspects Bulletin, v. 48, p. 157 – 164.
of the petroleum system that are probably in effect. Attanasi, E. D., 1998, Economics and the 1995 National Assess-
The role of uplift is poorly described in these basins, ment of United States Oil and Gas Resources: United States
Geological Survey Circular 1145, 35 p.
particularly as it relates to the generation of over- Barlow, J. A. Jr., M. J. Doelger, and T. J. McCutcheon, 1994,
pressures and the exsolution of gas. Understand- Accessibility to the Greater Green River basin gas supply,
ing of this in a quantitative sense could provide southwestern Wyoming: Gas Research Institute Report 94/
0363, GRI Contract No. 5093-212-2683.
insights to gas charge and trap timing. Permeability
Bass, D. M. Jr., 1989, Properties of reservoir rocks, in H. B. Bradley,
jail is an emerging concept that offers insights to ed., Petroleum engineering handbook: Society of Petroleum
low-permeability reservoir behavior and elegantly Engineers, chapter 26, p. 26-1 – 26-33.
explains many of the observations made in basin- Billingsley, R. L., 1994, Case history: Amoco horizontal program,
Wamsutter area, Green River basin, Wyoming, in Horizontal
centered areas. However, understanding how to well drilling and completions in the Greater Green River basin:
integrate this concept with more traditional rock- Gas Research Institute (GRI), Department of Energy (DOE),
typing methods and applying those learnings to sub- and the Independent Petroleum Association of Mountain States
(IPAMS) sponsored Consortium for Emerging Gas Resources in
surface data sets is still emerging. It is likely that the Greater Green River Basin, November 7, 1994, p. 77 – 102.
integration of these concepts will advance our petro- Billingsley, R. L., and L. W. E. Reinert, 1994, The Wamsutter
physical understanding well beyond the more tradi- sweetspot: A continuing enigma in tight formation gas (abs.):
AAPG Annual Meeting Program, v. 3, p. 104 – 105.
tional rock-typing methods that have dominated the
Blakeney-Dejarnett, B., F. H. Lim, L. F. Krystinik, and M. L. Bacon,
scene for many years. 2001, Greater Green River Basin Production Improvement
Project: Final report of the joint Department of Energy/Union
Pacific Resources Project Contract DE-AC21-95MC31063,
CD-ROM.
Boswell, R. M., A. S. B. Douds, H. R. Pratt, K. R. Bruner, K. K. Rose,
REFERENCES CITED J. A. Pancake, V. A. Kuuskraa, and R. L. Billingsley, 2002a,
Assessing technology needs of ‘‘sub-economic’’ gas resources
Al-Hadrami, H. K., and L. W. Teufel, 2000, Influence of in Rocky Mountain basins: Gas Tips, v. 8, no. 3, p. 4 – 12.
permeability anisotropy and reservoir heterogeneity on opti- Boswell, R. M., A. S. B. Douds, H. R. Pratt, K. K. Rose, J. A. Pancake,
mization of infill drilling in naturally fractured tight-gas K. R. Bruner, V. A. Kuuskraa, and R. L. Billingsley, 2002b,
Mesaverde sandstone reservoirs, San Juan basin: 2000 Society Assessing technology needs of sub-economic resources: Phase 1.
of Petroleum Engineers Rocky Mountain Regional/Low Greater Green River and Wind River basins, Fall 2002, in
Permeability Reservoirs Symposium, Denver, Colorado, Natural gas resources of the Greater Green River and Wind River
March 12 – 15, 2000, SPE Paper 60295. basins of Wyoming: National Energy Technology Laboratory/
Almon, W. R., and J. B. Thomas, 1991, Pore system aspects of Department of Energy CD 2002/1176, Final Version, February
hydrocarbon trapping, in H. J. Gluskoter, D. D. Rice, and R. B. 2003.
Taylor, eds., Economic geology, U.S.: Geological Society of Bowker, K. A., and B. C. Cotner, 1995, Lance and Mesaverde of the
America, The Geology of North America, v. P-2, p. 241 – 254. Hoback basin: Preliminary results from recent drilling, in
Al-Qarni, A. O. et al., 2001, From reservoir specifics to stimulation Geological characterizations of the Cretaceous gas plays in the
solutions: Oilfield Review, v. 12, p. 42 – 60. Greater Green River basin: Gas Research Institute (GRI),

1116 Gas Production from Low-Permeability Sandstone Reservoirs


Department of Energy (DOE), and the Independent Petro- prices: 2000 edition of the GRI baseline projection of U.S.
leum Association of Mountain States (IPAMS) sponsored Energy supply and demand to 2015: Gas Research Institute,
Consortium for Emerging Gas Resources in the Greater Green GRI-00/0005.
River Basin, November 6, 1995, p. 280 – 296. Coskey, R. J., in press, Burial history modeling of the Jonah field
Branagan, P., 1994, Reservoir modeling: Amoco horizontal well — area: An unusual and possibly unique gas accumulation in
Champlin 254 B-2H, in Horizontal well drilling and comple- the Green River basin, Wyoming, in J. R. Robinson and K. W.
tions in the Greater Green River basin: Gas Research Institute Shanley, eds., Jonah field: Case study of a giant tight-gas fluvial
(GRI), Department of Energy (DOE), and the Independent reservoir: AAPG Studies in Geology 52.
Petroleum Association of Mountain States (IPAMS) sponsored Davies, J. P., and D. K. Davies, 1999, Stress-dependent permeabil-
Consortium for Emerging Gas Resources in the Greater Green ity: characterization and modeling: 1999 Society of Petroleum
River Basin, November 7, 1994, p. 165 – 174. Engineers Annual Technical Conference, Houston, Texas, SPE
Breit, G. N., and C. C. Skinner, 2002, Produced waters database: Paper 56813.
United States Geological Survey, Digital Data Release, May Doelger, M. J., and D. K. Morton, 1999, Portfolio of emerging
2002: http://energy.cr.usgs.gov/prov/prodwat/index.html natural gas resources — Rocky Mountain basins: Section 1. Green
(accessed December 2002). River basin: Gas Research Institute, GRI-99/0169.1.
Burch, D. N., and R. M. Cluff, 1997, A volumetric analysis of Dolson, J., D. Muller, M. J. Evetts, and J. A. Stein, 1991, Regional
Almond Formation (Cretaceous, Mesaverde Group) gas paleotopographic trends and production, Muddy Sandstone
production in the Coal Gulch-Echo Springs — Standard Draw (Lower Cretaceous), central and northern Rocky Mountains:
field complex, Washakie basin, southwest Wyoming: 1997 Soci- AAPG Bulletin, v. 75, p. 409 – 435.
ety of Petroleum Engineers Rocky Mountain Regional Meeting, Dubois, D. P., P. J. Wynne, T. M. Smagala, J. L. Johnson, K. D.
Casper, Wyoming, May 18 – 21, 1997, SPE Paper 38368. Engler, and B. C. McBride, in press, Geology of Jonah field,
Byrnes, A. P., 1997, Reservoir characteristics of low-permeability Sublette County, Wyoming, in J. W. Robinson and K. W.
sandstones in the Rocky Mountains: Mountain Geologist, v. 34, Shanley, eds., Jonah field: Case study of a giant tight-gas fluvial
p. 39 – 51. reservoir: AAPG Studies in Geology 52.
Byrnes, A. P., 2003, Aspects of permeability, capillary pressure, and Dullien, F. A. L., 1992, Porous media: Fluid transport and pore
relative permeability properties and distribution in low- structure, 2d ed.: San Diego, California, Academic Press, 574 p.
permeability rocks important to evaluation, damage, and Dutton, S. P., S. J. Clift, D. S. Hamilton, H. S. Hamlin, T. F. Hentz,
simulation: Rocky Mountain Association of Geologists and W. E. Howard, M. S. Akhter, and S. E. Laubach, 1993, Major
Rocky Mountain Region of Petroleum Technology Transfer low-permeability sandstone gas reservoirs in the United States:
Council — Petroleum Systems and Reservoirs of Southwest Texas Bureau of Economic Geology Report of Investigations
Wyoming Symposium, September 19, 2003. No. 211, 221 p.
Byrnes, A. P., and J. W. Castle, 2000, Comparison of core Dutton, S. P., H. S. Hamlin, and S. E. Laubach, 1995, Geologic
petrophysical properties between low-permeability sandstone controls on reservoir properties of low-permeability sandstone,
reservoirs: Eastern U.S. Medina Group and western U.S. Frontier Formation, Moxa arch, southwestern Wyoming: Texas
Mesaverde Group and Frontier Formation: 2000 Society of Bureau of Economic Geology Report of Investigations No. 234,
Petroleum Engineers Rocky Mountain Low Permeability Res- 89 p.
ervoirs Symposium, Denver, Colorado, March 12 – 15, 2000: Dyman, T. S., J. W. Schmoker, B. E. Law, R. C. Johnson, T. S.
SPE Paper 60304. Collett, T. M. Finn, V. F. Nuccio, and J. L. Ridgley, 1998,
Byrnes, A. P., and C. W. Keighin, 1993, Effect of confining stress on Basin-centered and other continuous-type gas plays for
pore throats and capillary pressure measurements, selected resource assessment, in Developing a better understanding of
sandstone reservoir rocks (abs.): AAPG Annual Meeting basin centered gas plays: Gas Research Institute (GRI),
Program, v. 2, p. 82. Department of Energy (DOE), and the Independent Petro-
Byrnes, A. P., K. Sampath, and P. L. Randolph, 1979, Effect of leum Association of Mountain States (IPAMS) sponsored
pressure and water saturation on permeability of western tight Consortium for Emerging Gas Resources in the Greater Green
sandstones: Fifth Annual Department of Energy Symposium River basin, April 27, 1998, p. 33 – 53.
on Enhanced Oil and Gas Recovery and Improved Drilling EIA (Energy Information Administration), 2003, Annual energy
Technology, Tulsa, Oklahoma, August 22 – 24, 1979, p. 1 – outlook 2003 with projections to 2025: DOE/EIA 0383(2003):
15. www.eia.doe.gov/oiaf/aeo (accessed January 2003).
Cardinal, D. F., and W. W. Stewart, eds., 1979, Wyoming Oil and Finley, S. J., and J. C. Lorenz, 1989, Characterization and sig-
Gas Fields Symposium — Greater Green River basin: Casper, nificance of natural fractures in Mesaverde reservoirs at the
Wyoming, Wyoming Geological Association. multi-well experiment site: 1989 Society of Petroleum
Cluff, R. M., 2002, Reservoir properties of low-permeability gas Engineers Rocky Mountain Regional/Low-Permeability Reser-
sands, in Low permeability gas sands: Rocky Mountain Section voirs Symposium, Denver, Colorado, March 6 – 8, 1989, SPE
AAPG Meeting Short Course 4. Co-sponsored by Petroleum Paper 19007.
Technology Transfer Council (PTTC) and the Wyoming Finley, S. J., and J. C. Lorenz, 1995, Characterization of natural
Geological Association. Volume available through PTTC. fractures in Mesaverde core from the Multiwell experiment:
Cluff, S. G., and R. M. Cluff, in press, Petrophysics of the Lance Sandia Report SAND88-1800, UC-92, U.S. Department of
Sandstone reservoirs in Jonah field, Sublette County, Wyo- Energy Contract DE-AC04-94AL85000.
ming, in J. W. Robinson and K. W. Shanley, eds., Jonah field: Gautier, D. L., G. L. Dolton, K. I. Takahashi, and K. L. Varnes,
Case study of a giant tight-gas fluvial reservoir: AAPG Studies eds., 1995, 1995 National Assessment of United States Oil and
in Geology 52. Gas Resources — Results, methodology, and supporting data:
Cluff, R. M., K. W. Shanley, and A. P. Byrnes, in press, Permeability United States Geological Survey, Digital Data Series DDS-30
jail and implications for ‘‘basin-centered gas,’’ production, and (CD-ROM).
resource assessments (abs.): Rocky Mountain Section AAPG Gautier, D. L., G. L. Dolton, K. I. Takahashi, and K. L. Varnes,
Meeting, Denver, Colorado, August 2004. eds., 1996, 1995 National Assessment of United States Oil and
Cochener, J. C., 2000, The long-term trends in U.S. gas supply and Gas Resources — Results, methodology, and supporting data:

Shanley et al. 1117


United States Geological Survey, Digital Data Series DDS-30, the Almond Formation, Washakie basin, Wyoming: Society of
Release 2 (CD-ROM). Petroleum Engineers Rocky Mountain Regional/Low Perme-
Gunter, G. W., J. M. Finneran, D. J. Hartmann, and J. D. Miller, ability Reservoirs Symposium, Denver, Colorado, March 20 –
1997, Early determination of reservoir flow units using an 22, 1995, SPE Paper 29559.
integrated petrophysical method: 1997 Society of Petroleum Johnson, R. C., R. A. Crovelli, C. W. Spencer, and R. F. Mast, 1987,
Engineers Annual Technical Conference, San Antonio, Texas, An assessment of gas resources in low-permeability sandstones
October 5 – 8, 1997, SPE Paper 38679. of the Upper Cretaceous Mesaverde Group, Piceance basin,
Hanson, W. B., V. Carrillo, and S. Ritger, 1994, Evolution of the Colorado: U.S. Geological Survey Open-File Report 87-357.
Wamsutter arch, Wyoming foreland, U.S.A.: Influence on for- Johnson, R. C., T. M. Finn, R. A. Crovelli, and R. H. Balay, 1996,
mation of stratigraphic traps in Upper Cretaceous Almond An assessment of in-place gas resources in low-permeability
shoreline sandstones (abs.): AAPG Annual Meeting Program, Upper Cretaceous and lower Tertiary sandstone reservoirs,
v. 3, p. 163. Wind River basin, Wyoming: U.S. Geological Survey Open-
Hanson, W. B., V. Vega, and D. Cox, in press, Structural geology, File Report 96-264.
seismic imaging, and genesis of the giant Jonah gas field, Jones, F. O., and W. W. Owens, 1980, A laboratory study of low-
Wyoming, U.S.A., in J. R. Robinson and K. W. Shanley, eds., permeability gas sands: SPE Paper 7551, Journal of Petroleum
Jonah field: Case study of a giant tight-gas fluvial reservoir: Technology, v. 32, p. 1631 – 1640.
AAPG Studies in Geology 52. Katahara, K. W., and J. D. Corrigan, 2002, Effect of gas on poro-
Harstad, H., and L. W. Teufel, 1998a, Potential for infill drilling in a elastic response to burial or erosion, in A. R. Huffman and
naturally fractured tight-gas sandstone reservoir: 1998 Society G. L. Bowers, eds., Pressure regimes in sedimentary basins and
of Petroleum Engineers Rocky Mountain Regional/Low their prediction: AAPG Memoir 76, p. 73 – 78.
Permeability Reservoirs Symposium, Denver, Colorado, April Kirschbaum, M. A. et al., 2002a, Assessment of undiscovered oil
5 – 8, 1998, SPE Paper 39911. and gas resources of the Uinta – Piceance province of Colorado
Harstad, H., and L. W. Teufel, 1998b, Drainage efficiency in nat- and Utah, 2002: U.S. Geological Survey National Assessment of
urally fractured tight-gas sandstone reservoirs: 1998 Society of Oil and Gas Fact Sheet, FS-026-02: http://geology.cr.usgs.gov
Petroleum Engineers Gas Technology Symposium, Calgary, /pub/fact-sheets/fs-0026/index.html (accessed March 2002).
Alberta, Canada, March 15 – 18, 1998, SPE Paper 39974. Kirschbaum, M. A. et al., 2002b, Assessment of undiscovered oil
Harstad, H., L. W. Teufel, and J. C. Lorenz, 1995, Characterization and gas resources of the southwestern Wyoming, 2002: U.S.
and simulation of naturally fractured tight-gas sandstone Geological Survey National Assessment of Oil and Gas Fact
reservoirs: Society of Petroleum Engineers Annual Technical Sheet, FS-145-02: http://pubs.usgs.gov/fs/fs-145-02/fs-145
Conference, Dallas, Texas, October 22 – 25, 1995, SPE Paper -02.html (accessed December 2002).
30573. Kuuskraa, V. A., J. P. Brashear, T. M. Doscher, and L. E. Elkins,
Hartmann, D. J., and E. A. Beaumont, 1999, Predicting reservoir 1978, Enhanced recovery of unconventional gas, the program:
system quality and performance, in E. A. Beaumont and N. H. Volume II: U.S. Department of Energy, HCP/T2705-02, 537 p.
Foster, eds., Exploring for oil and gas traps — Treatise of Law, B. E., 1984, Relationships of source rocks, thermal matu-
petroleum geology: AAPG Treatise of Petroleum Geology, rity, and overpressuring to gas generation and occurrence in
chapter 9, p. 9-1 – 9-54. low-permeability Upper Cretaceous and lower Tertiary rocks,
Hartmann, D. J., and E. B. Coalson, 1990, Evaluation of the Morrow Greater Green River basin, Wyoming, Colorado, and Utah,
Sandstone in Sorrento field, Cheyenne County, Colorado, in in J. Woodward, F. F. Meissner, and J. L. Clayton, eds., Hy-
S. A. Sonnenberg, L. T. Shannon, K. Rader, W. F. von Drehl, drocarbon source rocks of the greater Rocky Mountain re-
and G. W. Martin, eds., Morrow sandstones of southeast gion: Rocky Mountain Association of Geologists Guidebook,
Colorado and adjacent areas: Rocky Mountain Association of p. 469 – 490.
Geologists Symposium Guidebook, p. 91 – 100. Law, B. E., 2000, What is a basin-centered gas system?: Rocky
Hartmann, D. J., and L. MacMillan, 1992, Petrophysics of the Mountain Association of Geologists, Basin-Center Gas Sym-
Wasatch Formation and Mesaverde Group, natural buttes pro- posium, Denver, Colorado, October 6, 2000, not paginated.
ducing area, Uinta basin, Utah, in T. D. Fouch, V. F. Nuccio, Law, B. E., 2002a, Concepts, geographic distribution, and
and T. C. Chidsey, eds., Hydrocarbon and mineral resources of exploration strategies, in Low permeability gas sands: AAPG
the Uinta basin, Utah and Colorado: Utah Geological Associ- Rocky Mountain Section Meeting Short Course 4, September
ation Guidebook, v. 20, p. 175 – 192. 11, 2002, Laramie, Wyoming, cosponsored by the Wyoming
Henry, M., M. Williams, S. Cook, and L. Tremper, 1999, Pros- Geological Association and the Petroleum Technology Trans-
pecting for subtle stratigraphic traps with 3-D seismic and well fer Council.
information: Examples from the Lewis Formation, Red Desert Law, B. E., 2002b, Basin-centered gas systems: AAPG Bulletin,
basin, Wyoming (abs.): AAPG Bulletin, v. 83, p. 1183. v. 86, p. 1891 – 1919.
Horn, B. W., and R. A. Schrooten, 2001, Development of the Echo Law, B. E., 2003, A review of basin-centered gas systems with a
Springs-Standard Draw field area: Using technology to enhance focus on the Greater Green River basin: Rocky Mountain
an infill program, Washakie basin, Wyoming, in D. S. Ander- Association of Geologists and Rocky Mountain Region of Petro-
son, J. W. Robinson, J. E. Estes-Jackson, and E. B. Coalson, leum Technology Transfer Council — Petroleum systems and
eds., Gas in the Rockies: Rocky Mountain Association of reservoirs of southwest Wyoming Symposium, Denver, Colo-
Geologists Guidebook, p. 171 – 188. rado, September 19, 2003.
Horn, B. W., T. A. Cross, J. A. Hornbeck, M. Vielma, and M. Law, B. E., and J. B. Curtis, 2002, Introduction to unconventional
Zavala, 2001, Stratigraphic controls on reservoir strata: A com- petroleum systems: AAPG Bulletin, v. 86, p. 1851 – 1852.
parison of fluvial and tidal reservoirs in the Almond Formation, Law, B. E., and W. W. Dickenson, 1985, Conceptual model for
Coal Gulch, Wamsutter, Echo Springs, and Table Rock fields, origin of abnormally pressured gas accumulations in low-
Washakie basin, Wyoming, in D. P. Stilwell, ed., Wyoming gas permeability reservoirs: AAPG Bulletin, v. 69, p. 1295 – 1304.
resources and technology: Wyoming Geological Association, Law, B. E., and C. R. Smith, 1986, Subsurface temperature map
52nd Field Conference Guidebook, p. 149 – 161. showing depth to 180j Fahrenheit in the Greater Green River
Iverson, W. P., and R. C. Surdam, 1995, Tight gas production from basin of Wyoming, Colorado, and Utah: U.S. Geological Survey

1118 Gas Production from Low-Permeability Sandstone Reservoirs


Miscellaneous Field Studies Map MF 1504, scale 1:500,000, NPC (National Petroleum Council), 1980, Unconventional gas
1 sheet. sources: Tight-gas reservoirs: Parts I and II. Report of the Com-
Law, B. E., and C. W. Spencer, eds., 1989, Geology of tight-gas mittee on Unconventional Gas Sources, J. R. Bookout, Chairman.
reservoirs in the Pinedale anticline area, Wyoming, and at the NPC (National Petroleum Council), 1992, The potential for natural
multiwell experiment site, Colorado: U.S. Geological Survey gas in the United States: Source and Supply, v. 2, 330 p.
Bulletin, v. 1886. Pawlewicz, M. J., M. R. Lickus, B. E. Law, W. W. Dickenson, and
Law, B. E., and C. W. Spencer, 1993, Gas in tight reservoirs — An C. S. V. Barclay, 1986, Thermal maturity map showing
emerging major source of energy, in D. G. Howell, K. Wiese, subsurface elevation of 0.8 percent vitrinite reflectance in the
M. Fanelli, L. Zink, and F. Cole, eds., The future of energy gases: Greater Green River basin of Wyoming, Colorado, and Utah:
U.S. Geological Survey Professional Paper 1570, p. 233 – 252. U.S. Geological Survey Miscellaneous Field Studies Map MF
Law, B. E., and C. W. Spencer, in press, Basin-centered gas systems 1890, scale 1:500,000, 1 sheet.
and the Jonah field, in J. W. Robinson and K. W. Shanley, eds., PGC (Potential Gas Committee), 2003, Potential supply of natural
Jonah field: Case study of a giant tight-gas fluvial reservoir: gas in the United States (December 31, 2002): Potential Gas
AAPG Studies in Geology 52. Agency, Colorado School of Mines, March 2003, 316 p.
Law, B. E., C. W. Spencer, and N. H. Bostick, 1980, Evaluation of Popov, M. A., V. F. Nuccio, T. S. Dyman, T. A. Gognat, R. C.
organic maturation, subsurface temperature, and pressure with Johnson, J. W. Schmoker, M. S. Wilson, and C. Bartberger,
regard to gas generation in low-permeability Upper Cretaceous 2001, Basin-centered gas systems of the U.S.: U.S. Geological
and lower Tertiary strata in the Pacific Creek area, Sublette Survey Open-File Report 01-135, version 1.0 (CD-ROM).
County, Wyoming: Mountain Geologist, v. 17, p. 23 – 35. Prouty, J. L., 2001, Tight gas in the spotlight: Gas Tips, v. 7, no. 2,
Law, B. E., C. W. Spencer, R. R. Charpentier, R. A. Crovelli, R. F. p. 4 – 10.
Mast, G. L. Dolton, and C. J. Wandrey, 1989, Estimates of gas Raynolds, R. G., and I. Pasternack, 1994, Mesaverde Group sand-
resources in overpressured low-permeability Cretaceous and stone reservoir characteristics, San Juan basin, New Mexico and
Tertiary sandstone reservoirs, Greater Green River basin, Wyo- Colorado (abs.): AAPG Annual Meeting Program, v. 3, p. 242.
ming, Colorado, and Utah, in J. L. Eisert, ed., Gas resources of Ridgley, J. L., S. M. Condon, R. F. Dubiel, R. R. Charpentier, T. A.
Wyoming: Wyoming Geological Association, 40th Field Cook, R. A. Crovelli, T. R. Klett, R. M. Pollastro, and C. J.
Conference Guidebook, p. 39 – 61. Schenk, 2002, Assessment of undiscovered oil and gas
Lorenz, J. C., and S. J. Finley, 1989, Differences in fracture resources of the San Juan basin province of New Mexico and
characteristics and related production: Mesaverde Formation, Colorado, 2002: U.S. Geological Survey National Assessment
northwestern Colorado: SPE Paper 16809, Society of Petro- of Oil and Gas Fact Sheet, FS-147-02: http://pubs.usgs.gov/fs
leum Engineers Formation Evaluation, March 1989, p. 11 – 16. /fs-147-02/fs-147-02.html (accessed December 2002).
Lorenz, J. C., N. R. Warpinski, P. T. Branagen, and A. R. Sattler, Robinson, J. W., 2000, Discovery of Jonah field, Sublette County,
1989, Fracture characteristics and reservoir behavior of stress- Wyoming: Mountain Geologist, v. 37, p. 135 – 143.
sensitive fracture systems in flat-lying lenticular formations: Robinson, J. W., D. Delozier, and R. Finch, 1995, Integrated res-
SPE Paper 15244, Journal of Petroleum Technology, v. 41, ervoir description and analysis of the Lance Formation, Jonah
p. 615 – 622. field, Sublette County, Wyoming, in Geological characteriza-
Martin, W. B., and J. Shaughnessy, 1969, Project Wagon Wheel: tions of the Cretaceous gas plays in the Greater Green River
Wyoming Geological Association Symposium on Tertiary basin: Gas Research Institute/Department of Energy/Indepen-
Rocks of Wyoming, 21st Annual Field Conference Guidebook, dent Petroleum Association of Mountain States sponsored
p. 145 – 152. Consortium for emerging gas resources in the Greater Green
Masters, J. A., 1979, Deep basin gas trap, western Canada: AAPG River basin, p. 297 – 308.
Bulletin, v. 63, p. 152 – 181. Robinson, J. W., D. Delozier, and R. Finch, 1996, Integrated res-
Masters, J. A., 1984, Lower Cretaceous oil and gas in western ervoir description and analysis of the Lance Formation, Jonah
Canada, in J. A. Masters, ed., Elmworth — Case study of a field, Sublette County, Wyoming (abs.): AAPG Bulletin, v. 80,
deep basin gas field: AAPG Memoir 38, p. 1 – 34. p. 980.
McPeek, L. A., 1981, Eastern Green River basin: A developing giant Rose, P. R., J. R. Everett, and I. S. Merin, 1986, Potential basin-
gas supply from deep, overpressured Upper Cretaceous centered gas accumulation in Cretaceous Trinidad Sandstone,
sandstones: AAPG Bulletin, v. 65, p. 1078 – 1098. Raton basin, Colorado, in C. W. Spencer and R. F. Mast, eds.,
Milici, R. C., R. T. Ryder, C. S. Swezey, R. R. Charpentier, T. A. Geology of tight-gas reservoirs: AAPG Studies in Geology 25,
Cook, R. A. Crovelli, T. R. Klett, R. M. Pollastro, and C. J. p. 111 – 128.
Schenk, 2003, Assessment of undiscovered oil and gas Rose, W., 1999, Relative permeability ideas — Then and now
resources of the Appalachian basin province, 2002: U.S. (Richards to Leverett to Yuster, and beyond): 1999 Eastern
Geological Survey National Assessment of Oil and Gas Fact Regional Meeting, Society of Petroleum Engineers, Charleston,
Sheet, FS-009-03: http://pubs.usgs.gov/fs/fs-009-03/fs-009 West Virginia, October 20 – 22, 1999, SPE Paper 57442.
-03.html (accessed February 2003). Schenk, C. J., and R. M. Pollastro, 2002, Natural gas production in
Miller, T. S., F. J. Crockett, and S. H. Hollis, eds., 1992, Wyoming the United States: U.S. Geological Survey National Assessment
Oil and Gas Fields Symposium — Greater Green River basin of Oil and Gas Fact Sheet, FS-113-01: http://geology.cr.usgs.gov
and overthrust belt: Wyoming Geological Association, Casper, /pub/fact-sheets/fs-0113-01/index.html (accessed January 2002).
Wyoming, 364 p. Schmoker, J. W., 1995, Method for assessing continuous-type
Montgomery, S. L., and J. W. Robinson, 1997, Jonah field, Sublette (unconventional) hydrocarbon accumulations, in D. L. Gau-
County, Wyoming: Gas production from overpressured Upper tier, G. L. Dolton, K. I. Takahashi, and K. L. Varnes, eds.,
Cretaceous Lance sandstones of the Green River basin: AAPG 1995 National Assessment of United States Oil and Gas
Bulletin, v. 81, p. 1049 – 1062. Resources — Results, methodology, and supporting data: U.S.
Northrop, D. A., 1986, Insights into natural gas production from Geological Survey Digital Data Series DDS-30 (CD-ROM).
low-permeability reservoirs: Society of Petroleum Engineers Schmoker, J. W., 2002, Resource-assessment perspectives for un-
Gas Technology Symposium, Dallas, Texas, June 13 – 15, 1986, conventional gas systems: AAPG Bulletin, v. 86, p. 1993 – 1999.
SPE Paper 17706. Shanley, K. W., 2001, Fluvial reservoir description for a tight-gas

Shanley et al. 1119


giant: Jonah field, Green River basin, Wyoming, in J. W. analysis in Siberia Ridge field, Wyoming, in D. S. Anderson,
Robinson and K. W. Shanley, coconvenors, Tight-gas fluvial J. W. Robinson, J. E. Estes-Jackson, and E. B. Coalson, eds.,
reservoirs: A case history from the Lance Formation, Green Gas in the Rockies: Rocky Mountain Association of Geologists
River basin, Wyoming: AAPG Short Course No. 2, in asso- Guidebook, p. 145 – 170.
ciation with the AAPG Annual Meeting, 186 p. Supply-Technical Advisory Task Force, 1973, Task force report:
Shanley, K. W., in press, Fluvial reservoir description for a giant, National gas supply: National Gas Survey, v. II, Federal Power
low-permeability gas field: Jonah field, Green River basin, Commission, 662 p.
Wyoming, U.S.A., in J. R. Robinson and K. W. Shanley, eds., Supply-Technical Advisory Task Force, 1978, National gas survey,
Jonah field: Case study of a giant tight-gas fluvial reservoir: in Report to the Federal Energy Regulatory Commission by the
AAPG Studies in Geology 52. Supply-Technical Advisory Task Force on nonconventional
Shannon, L. T., 1992, Hay reservoir, in T. S. Miller, F. J. Crockett, natural gas resources: U.S. Department of Energy, Federal
and S. H. Hollis, eds., Wyoming Oil and Gas Fields Symposium— Energy Regulatory Commission, 109 p.
Greater Green River basin and overthrust belt: Casper, Wyo- Surdam, R. C., 1995, Sweetspot delineation in the Mesaverde Group,
ming, Wyoming Geological Association, p. 178 – 179. Greater Green River basin, in Geological characterizations of
Shaughnessy, J., and R. H. Butcher, 1974, Geology of Wagon the Cretaceous gas plays in the Greater Green River basin: Gas
Wheel Nuclear Stimulation Project, Pinedale field, Wyoming: Research Institute (GRI), Department of Energy (DOE), and the
AAPG Bulletin, v. 58, p. 2250 – 2259. Independent Petroleum Association of Mountain States (IPAMS)
Shaver, M. A., and B. F. Towler, 1997, Reservoir engineering for the sponsored Consortium for Emerging Gas Resources in the
Almond Formation in the Standard Draw gas field in southern Greater Green River basin, November 6, 1995, p. 158 – 207.
Wyoming: 1997 Society of Petroleum Engineers Rocky Surdam, R. C., 1997a, A new paradigm for gas exploration in
Mountain Regional Meeting, Casper, Wyoming, May 18 – 21, anomalously pressured ‘‘tight-gas sands’’ in the Rocky Moun-
1997, SPE Paper 38367. tain Laramide basins, in R. C. Surdam, ed., Seals, traps, and
Smith, R. D., 1984, Gas reserves and production performance of the petroleum system: AAPG Memoir 67, p. 283 – 298.
the Elmworth/Wapiti area of the deep basin, in J. A. Masters, Surdam, R. C., 1997b, A new exploration strategy for anomalously
ed., Elmworth — Case study of a deep basin gas field: AAPG pressured reservoirs in Rocky Mountain basins: Gas Tips, v. 3,
Memoir 38, p. 153 – 172. no. 4, p. 4 – 8.
Sneider, R. M., H. R. King, H. E. Hawkes, and T. K. Davis, 1981, Surdam, R. C., W. P. Iverson, and P. Yin, 1995, Gas reservoir sweet
Methods for detection and characterization of reservoir rock, spot detection and delineation in Rocky Mountain Laramide
deep basin gas area, western Canada: 56th Annual Conference Basins: Gas Research Institute, GRI-95/0443, Topical Report
of Society of Petroleum Engineers, San Antonio, Texas, SPE for GRI contract No. 5091-221-2146, October, 1995.
Paper 10072. Surdam, R. C., Z. S. Jiao, N. K. Boyd, P. Yin, S. Stookey, R. S. Mar-
Sneider, R. M., H. R. King, H. E. Hawkes, and T. K. Davis, 1983, tinsen, W. P. Iverson, F. P. Miknis, S. L. Kubichek, and T. L.
Methods for detection and characterization of reservoir rock, Dunn, 1997a, Multidisciplinary analysis of pressure chambers
deep basin gas area, western Canada: Journal of Petroleum in the Powder River basin, Wyoming and Montana, in A new
Technology, v. 35, p. 1725 – 1734. innovative exploitation strategy for gas accumulations within
Sneider, R. M., H. R. King, R. W. Hietala, and E. T. Connolly, pressure compartments: Gas Research Institute, GRI-97/0364,
1984, Integrated rock-log calibration in the Elmworth Final report for GRI contract No. 5089-260-1894, October
field — Alberta, Canada, in J. A. Masters, ed., Elmworth — 1997.
Case study of a deep basin gas field: AAPG Memoir 38, p. 205 – Surdam, R. C., Z. S. Jiao, and H. P. Heasler, 1997b, Anomalously
282. pressured gas compartments in Cretaceous rocks of the
Soeder, D. J., and P. Chowdiah, 1990, Pore geometry in high- and Laramide basins of Wyoming: A new class of hydrocarbon
low-permeability sandstones, Travis Peak Formation, east Texas: accumulation, in R. C. Surdam, ed., Seals, traps, and the
Society of Petroleum Engineers Formation Evaluation, De- petroleum system: AAPG Memoir 67, p. 199 – 222.
cember 1990, p. 421 – 430, SPE Paper 17729. Surdam, R. C., Z. S. Jiao, and N. G. K. Boyd, 2000, Detection and
Soeder, D. J., and P. L. Randolph, 1987, Porosity, permeability, and delineation of basin-centered gas accumulations occurring in
pore structure of the tight Mesaverde Sandstone, Piceance Rocky Mountain Laramide basins: Rocky Mountain Association
basin, Colorado: Society of Petroleum Engineers Formation of Geologists Basin-Centered Gas Symposium, Denver, Colo-
Evaluation, June 1987, p. 129 – 136, SPE Paper 13134. rado, October 6, 2000.
Spencer, C. W., 1985, Geologic aspects of tight gas reservoirs in the Thomas, R. D., and D. C. Ward, 1972, Effect of overburden pres-
Rocky Mountain region: SPE Paper 11647, Journal of Petro- sure and water saturation on gas permeability of tight sand-
leum Technology, v. 37, p. 1308 – 1314. stone cores: SPE Paper 3634, Journal of Petroleum Technology,
Spencer, C. W., 1987, Hydrocarbon generation as a mechanism for v. 24, p. 120 – 124.
overpressuring in the Rocky Mountain region: AAPG Bulletin, Van Horn, M. D., and L. T. Shannon, 1985, Stratigraphy of the
v. 71, p. 368 – 388. Lewis Formation: A tectonically influenced constructional
Spencer, C. W., 1989, Review of characteristics of low-permeability delta and submarine fan complex of the Late Cretaceous,
gas reservoirs in western United States: AAPG Bulletin v. 73, southwestern Wyoming (abs.): SEPM Annual Midyear Meet-
p. 613 – 629. ing Program, v. II, p. 91.
Spencer, C. W., and R. F. Mast, eds., 1986, Geology of tight-gas Van Horn, M. D., and L. T. Shannon, 1989, Hay reservoir field: A
reservoirs: AAPG Studies in Geology 25, 299 p. submarine fan gas reservoir within the Lewis Shale, Sweet-
Sturm, S. D. et al., 2000, Multi-disciplinary analysis of tight gas water County, Wyoming: Wyoming Geological Association
sandstone reservoirs, Almond Formation, Siberia Ridge field, 40th Annual Field Conference, p. 155 – 180.
Greater Green River basin: Gas Research Institute, GRI-00/ Vavra, C. I., J. G. Kaldi, and R. M. Sneider, 1992, Geological
0026, Final report for the Gas Research Institute contract applications of capillary pressure: A review: AAPG Bulletin,
No. 5094-210-3021, January 2000. v. 76, p. 840 – 850.
Sturm, S. D., L. W. Evans, B. F. Keusch, and W. J. Clark, 2001, Walls, J. D., 1981, Tight gas sands: Permeability, pore structure,
Almond Formation reservoir characterization and sweet spot and clay: 1981 Society of Petroleum Engineers/Department of

1120 Gas Production from Low-Permeability Sandstone Reservoirs


Energy Low Permeability Symposium, Denver, Colorado, May Wells, J. D., and J. O. Amaefule, 1985, Capillary pressure and
27 – 29, 1981, SPE Paper 9871. permeability relationships in tight gas sands: 1985 Society of
Walls, J. D., A. M. Nur, and T. Bourbie, 1982, Effects of pressure Petroleum Engineers/Department of Energy Low Permeability
and partial water saturation on gas permeability in tight sands: Gas Reservoirs Symposium, Denver, Colorado, May 19 – 22,
experimental results: SPE Paper 9378, Journal of Petroleum 1985, SPE Paper 13879.
Technology, v. 34, p. 930 – 936. WOGCC (Wyoming Oil and Gas Conservation Commission):
Walters, F. S., 1992, Swan field (Blue Forest unit, Monument Butte http://www.wogcc.state.wy.us. (accessed December 2003).
unit, Raptor unit and Lincoln Road), in T. S. Miller, F. J. Yuan, H. H., and B. F. Swanson, 1986, Resolving pore space
Crockett, and S. H. Hollis, eds., Wyoming Oil and Gas Fields characteristics by rate-controlled porosimetry: 1986 Society of
Symposium — Greater Green River basin and overthrust belt: Petroleum Engineers/Department of Energy Fifth Symposium
Wyoming Geological Association, Casper, Wyoming, p. 306 – on Enhanced Oil Recovery, Tulsa, Oklahoma, April 20 – 23,
308. 1986, SPE Paper 14892.

Shanley et al. 1121

Potrebbero piacerti anche