You are on page 1of 35

Outcrop-based reservoir AUTHORS

Anthony Scott  University of Aberdeen,


characterization of Department of Geology and Petroleum Geol-
ogy, King’s College, Aberdeen, United Kingdom;
a kilometer-scale present address: Statoil, Reservoir Modeling,
Sandslihaugen 30, Bergen, Norway;
sand-injectite complex ajsc@statoil.com
Anthony Scott currently holds a position as a
Anthony Scott, Andrew Hurst, and Mario Vigorito senior sedimentologist in the Reservoir Char-
acterization and Geomodeling Group at Statoil
(2009–2012). He obtained a B.Sc. degree from
the University of Durham and an M.Sc. degree
ABSTRACT and a Ph.D. from the University of Aberdeen.
His current research interests are in the evo-
This study documents that Danian-aged sand remobilization lution of fluid-flow phenomena, sediment re-
of deep-water slope-channel complexes and intrusion of fluid- mobilization of deep-water depositional systems,
ized sand into hydraulically fractured slope mudstones of the and large-scale clastic intrusions.
Great Valley sequence, California, generated 400-m (1312 ft)– Andrew Hurst  University of Aberdeen,
thick reservoir units: unit 1, parent unit channel complexes for Department of Geology and Petroleum Geol-
shallower sandstone intrusions; unit 2, a moderate net-to-gross ogy, King’s College, Aberdeen, United Kingdom;
interval (0.19 sand) of sills with staggered, stepped, and multi- ahurst@abdn.ac.uk
layer geometries with well-developed lateral sandstone-body Andrew Hurst is professor of production geosci-
connectivity; unit 3, a low net-to-gross interval (0.08 sand) of ence in the Department of Geology and Petro-
exclusively high-angle dikes with good vertical connectivity; and leum Geology, University of Aberdeen, a post he
unit 4, an interval of extrusive sandstone. Unit 2 was formed has held since 1992. Previously, he worked for
Statoil and Unocal in a 13-yr industry career.
during a phase of fluidization that emplaced on an average Sand injectites have been a major area of his
0.19 km3 (0.046 mi3) of sand per cubic kilometer of host sed- research since 1998, and he leads the Aberdeen-
iment. Probe permeametry data reveal a positive relationship based Sand Injectites Research Group. He was
between sill thickness and permeability. Reservoir quality is coeditor of AAPG Memoir 87 (2007), Relevance
reduced by the presence of fragments of host strata, such as the of Sand Injectites to Hydrocarbon Exploration
incorporation of large rafts of mudstone, which are formed by and Production.
in-situ hydraulic fracturing during sand injection. Mudstone Mario Vigorito  University of Aberdeen,
clasts and clay- and silt-size particles generated by intrusion- Department of Geology and Petroleum Geol-
induced abrasion of the host strata reduce sandstone perme- ogy, King’s College, Aberdeen, United Kingdom;
ability in multilayer sills (70 md) when compared to that in present address: Statoil, Exploration, Forus Vest,
Stavanger, Norway; mavig@statoil.com
staggered and stepped sills (586 and 1225 md, respectively).
Post-injection cementation greatly reduces permeability in Mario Vigorito joined Statoil in 2009 and, since
then, has worked as a sedimentologist in the
high-angle dikes (81 md). This architecturally based reservoir exploration on the Norwegian continental shelf
zonation and trends in reservoir characteristics in dikes and sills (North Sea Basin). From 2005 to 2009, he held a
form a basis for quantitative reservoir modeling and can be position as a lecturer in geology and petroleum
used to support conceptual interpretations that infer injectite geology at the University of Aberdeen and pre-
viously worked as a surveyor for the Italian
Geological Survey and as a consulting sedimen-
tologist for oil companies. He has work and
research experience in a variety of oil provinces
(North Sea [United Kingdom] and Norwegian Sea
Copyright ©2013. The American Association of Petroleum Geologists. All rights reserved.
[Norway], Sacramento and San Joaquin basins
Manuscript received December 9, 2011; provisional acceptance February 28, 2012; revised manuscript
received April 3, 2012; final acceptance May 14, 2012.
[California], and western Mediterranean), basins,
DOI:10.1306/05141211184

AAPG Bulletin, v. 97, no. 2 (February 2013), pp. 309–343 309


and tectonic settings, with a strong focus on architecture in situations where sands in low net-to-gross in-
carbonate, siliciclastic, and unconventional plays tervals are anticipated to have well-developed lateral and
(e.g., sand intrusions and fractured basement)
vertical connectivity.
and basin-flow analysis. He obtained an M.Sc.
degree and a Ph.D. in carbonate sedimentology
from Università degli Studi di Napoli “Federico II,”
Italy. INTRODUCTION

When translating static geologic models into dynamic reservoir-


ACKNOWLEDGEMENTS engineering models that are designed to evaluate recovery
This study was conducted as part of the Sand mechanism and field development strategy, geologically de-
Injectites Phase 2 at the University of Aberdeen, fined flow units are required, and their spatial distribution is
funded by Department of Energy and Climate mapped. This is particularly relevant to sand-injectite reser-
Change (DECC) (formerly the UK Department of voirs because they are typified by complex geometries (Duranti
Trade and Industry), Dong, Lundin, Marathon,
and Hurst, 2004; Hamberg et al., 2007) and architectural hi-
Statoil, and Total. We thank our colleagues in the
Injected Sands Group for their support. The erarchies (Vigorito et al., 2008; Hurst et al., 2011) and are more
interpretations presented here are solely those permeable than the strata into which they were emplaced
of the authors. We thank Ian D. Bryant and (Duranti et al., 2002; Scott et al., 2009). They form by sand
two anonymous reviewers, and AAPG Editor, remobilization and injection of fluidized sand from an over-
Stephen E. Laubach, for the peer-review com- pressured parent unit (e.g., channels that sourced the fluidized
ments that enhanced the manuscript. We also
sand) into a hydraulically fractured seal lithology (e.g., mud-
thank Frances P. Whitehurst for editing the final
manuscript. We would like to thank Statoil U.K. stone) (Cosgrove, 2001; Jolly and Lonergan, 2002; Cartwright,
and, in particular, Mariner Petroleum Technol- 2010), thereby creating a network of sills and high- and low-
ogy, for financial support in the publication of angle dikes (Taylor, 1982; Hurst et al., 2011). These networks
this manuscript. form conduits for aqueous and hydrocarbon fluids (Jenkins,
The AAPG Editor thanks the following reviewers 1930; Jonk et al., 2003, 2005a, b; Mazzini et al., 2003a, b) that
for their work on this paper: Ian D. Bryant
focus flow between reservoirs separated by hundred-meter–
and two anonymous reviewers.
thick low-permeability seal lithology (Huuse et al., 2005).
Sandstone intrusions may pose a risk to the seal capacity
DATASHARE 46 in strata that overlie conventional traps (Hurst et al., 2003a;
Appendices 1–16 are accessible in electronic Cartwright et al., 2007; Hurst and Cartwright, 2007), but their
version on the AAPG Website (www.aapg.org mapping (Huuse et al., 2005; Frey-Martínez et al., 2007) and
/datashare) as Datashare 46. the definition of a new trapping style, termed “intrusive traps”
(Hurst et al., 2005), led to the successful deliberate exploration
of sand injectites for hydrocarbons (de Boer et al., 2007).
Sandstone intrusions commonly form large-volume reservoirs
and contain substantial reserves of hydrocarbon in many oil
fields (Table 1). They may contain millions of cubic meters of
sandstone (5 × 104 m3 [1.77 × 106 ft3], Hurst et al., 2003a;
50–60 × 106 m3 [17.66–21.19 × 108 ft3] intrusion–1, Huuse
and Mickelson, 2004), crosscut hundreds of meters of strati-
graphic section (Jolly et al., 1998; Vigorito et al., 2008), and
are regionally developed (Jackson, 2007).
Sandstone intrusions can cause unexpected (in the context
of stratiform depositional reservoirs) fluid-flow effects during
production—for example, deleterious effects such as unpre-
dicted, early water breakthrough (Briedis et al., 2007) and
helpful effects such as fieldwide enhanced vertical permeability

310 Reservoir Characterization of a Sand-Injectite Complex


Table 1. Examples of Fields where Hydrocarbons are Reservoired in Post-Depositional Remobilized Sandstone and Sandstone Injectite Complexes*

Reserves
Field/Structure Reservoir Interval and Age Depositional Environment Reservoir Style and Scale of Intrusion (× 106 BOE**) Key References
Harding (UK) Mid-Eocene Deep-water basin-floor setting Intrareservoir shales crosscut by intrusions; 280 Dixon et al. (1995);
crestal intrusions Bergslien (2002)
Gryphon (UK) Balder Formation (early Eocene) Deep-water slope setting Intrareservoir shales crosscut by intrusions; 117 Purvis et al. (2002);
crestal intrusions Lonergan et al. (2007)
Chestnut (UK) Alba/Chestnut sequence Deep-water slope and Remobilized depositional sandstones; conical 25 Huuse et al. (2005)
(upper Eocene) basin-floor setting intrusions in reservoir underburden
Volund (Norway) Sele, Balder, Horda formations Deep-water slope setting Saucer-shape sandstone intrusion NA** Szarawarska et al. (2010)
(early Eocene)
Balder (Norway) Lista, Hermod, Balder formations Deep-water basin-floor setting Low-angle dikes at channel margins; 186 Briedis et al. (2007);
(Paleocene–early Eocene) crestal intrusions; conical intrusions Wild and Briedis (2010)
Nini area (Denmark) Våle, Lista, Balder formations Deep-water slope setting Low-angle dikes at margins of remobilized NA Svendsen et al. (2010)
(late Paleocene–early Eocene) sandstone; intrachalk sands; fault-zone
injectites
Alba (UK) Alba Formation (upper Eocene) Deep-water slope setting Low-angle dikes at channel margins; 460 Duranti and Hurst (2004);
crestal intrusions Fretwell et al. (2007);
Huuse et al. (2007)
Sleipner Øst (Norway) Ty Formation (lower Paleocene) Deep-water basin-floor setting Intrareservoir shales crosscut by intrusions NA Satur and Hurst (2007)
Danica (UK) Sele, Balder, Horda formations Deep-water slope setting Low-angle dikes at channel margins; NA Szarawarska et al. (2010)
(early Eocene) crestal intrusions
Jotun (Norway) Heimdal Formation (Paleocene) Deep-water basin-floor setting Intrareservoir shales crosscut by intrusions; 203 Guargena et al. (2007)
crestal intrusions
Hamsun (Norway) Sele, Balder formations Deep-water slope and Sandstone-injectite complex of low-and NA De Boer et al. (2007)
(early Eocene) basin-floor setting high-angle dikes and sills
Scott et al.

Cecile (Denmark) Våle Formation (late Deep-water slope setting Dome-shape sandstone injection; crestal NA Hamberg et al. (2007)
Paleocene–early Eocene) intrusions
*Reserve estimates are compiled from Hurst et al. (2005, table 1) and may not represent present-day estimates.
**BOE = barrels of oil equivalent; NA = not applicable.
311
and pressure communication in a field with re- sections in the northern San Francisco area, Sacra-
gionally developed intra- and interreservoir mud- mento Basin, and San Joaquin Basin (Thompson
stones (Hurst et al., 2007; Satur and Hurst, 2007). et al., 1999, 2007; Huuse et al., 2004; Vigorito et al.,
Therefore, incorporating their sedimentologic and 2008; Scott et al., 2009). Basin-scale overpressure
petrophysical properties into 3-D reservoir models linked to eastward subduction has been measured
is important (Briedis et al., 2007). This process has in oil and gas fields in both the San Joaquin and
been inhibited by the lack of outcrop analog data, Sacramento basins (Berry, 1973; McPherson and
which limits accurate subsurface reservoir char- Garven, 1999). The occurrence of sand injectites
acterization and geologic modeling (Purvis et al., within these basins may reflect tectonically induced
2002; Briedis et al., 2007); in particular, thickness, basin-scale fluid overpressure. The PGIC is found
detailed architecture, textures, and porosity and per- within the upper Mesozoic strata of the Great Val-
meability (e.g., vertical-to-horizontal ratios; Hurst ley sequence (GVS). The GVS filled the forearc
and Buller, 1984; Duranti et al., 2002) are poorly basin from the east during the Cretaceous by ero-
constrained. As a consequence, spurious reservoir sion of abruptly exhumed plutonic rocks in the
models may be generated, which in turn leads to southern Sierra Nevada magmatic arc from the
ambiguous and inaccurate reserve estimates and early to mid-Maastrichtian, and from the late to
inappropriate field developments. the end of the Maastrichtian, sediment was derived
Here, we provide an outcrop-based reservoir from the west from exhumation of the Franciscan
characterization of the kilometer-scale Panoche giant complex (Dickinson, 1976; Ingersoll, 1979, 1983;
injection complex (Figure 1A), which extends over DeGraff-Surpless et al., 2002; Mitchell et al., 2010).
an area of approximately 400 km2 (154 mi2) in the These petrotectonic provinces formed concurrently
San Joaquin Basin, central California. Reservoir ar- during Pacific plate subduction beginning in the
chitecture and characteristics are investigated so Late Jurassic until the transformation of the sub-
that an enhanced appreciation is possible of their duction zone into a transform margin during the late
significance in numerical geologic models. The fo- Cenozoic (Atwater and Molnar, 1973; Dickinson
cus of this article is on the relationship between and Seely, 1979).
sandstone intrusion distribution, sandstone-body The PGIC extends across the Panoche and Mo-
connectivity, architecture, microtexture, and poros- reno formations of the GVS (Figure 1B). The Pa-
ity and permeability. This study provides a compre- noche Formation is a sandstone-rich succession that
hensive database whereby readers can access data evolves from submarine turbidite-fan deposits to
on reservoir architecture (geometry, lithologic char- amalgamated channelized-fan complexes in its up-
acter, and net to gross [N/G]) and characteristics permost part (Ingersoll, 1979, 1983). In the upper-
(qualitative and quantitative petrography, and po- most parts, the sandstone units exhibit evidence
rosity and permeability) of sandstone intrusions and of channeling (50–100 m [164–328 ft] thick and
remobilized sandstones (depositional units mod- 500–2500 m [1640–8202 ft] long), suggesting de-
ified by sand fluidization) and locate this within position in a proximal fan environment (Ingersoll,
the injectite complex and regional stratigraphy 1979). The Moreno Formation unconformably over-
(Appendices 1–16; see AAPG Datashare 46 at lies the Panoche Formation and consists of four
www.aapg.org/datashare). members. The Dosados Member is an approxi-
mately 60-m (197-ft)–thick sequence of isolated
base-of-slope channel deposits (10–40 m [33–131 ft]
thick and 100–200 m [328–656 ft] wide) crosscut by
GEOLOGIC SETTING sandstone dikes. The Tierra Loma Shale Member is
an approximately 350-m (1148-ft)–thick sequence
Reservoir-scale sand-injectite outcrops including of organic-rich, dark-brown mudstone deposited
the Panoche giant injection complex (PGIC) are on the lower to middle slope under suboxic con-
reported from California in different stratigraphic ditions (McGuire, 1988), and the overlying Marca

312 Reservoir Characterization of a Sand-Injectite Complex


Figure 1. Regional petrotectonic provinces, lithostratigraphy, and architecture of the Panoche giant injectite complex (PGIC; Vigorito et al.,
2008). (A) Petrotectonic provinces and major faults cropping out in central California (modified from Ingersoll, 1979; Nilsen and Dibblee,
1981). The PGIC crops out on the eastern flank of a monocline of the Great Valley sequence along the western margin of the Central Valley of
California. Faults: H = Hayward; SA = San Andreas. (B) Lithostratigraphy and architecture of the northern PGIC (modified from Vigorito and
Hurst, 2010). (C) Geologic map of the Panoche Hills (modified from Bartow, 1996). Study sites are located with boxes. Locations of logged
sections (in Figure 3) are marked. See Appendices 1 to 10 for text, maps, and stratigraphic sections of the study areas. MG = Moreno Gulch;
NoC = No Name Canyon; Mc = Marca Canyon; Ca = Capita Canyon; Ro = Rosseta Canyon; Do = Dosados Canyon; Es = Escapardo Canyon;
RA = Right Angle Canyon; WT = West Tumey; Fm = Formation; Mbr = Member; Cima Sst. L. = Cima Sandstone lentil; En. = environment;
Res. = reservoir unit.

Shale Member is an approximately 100-m (328-ft)– sea level highstands, the base of which is a flooding
thick pale-gray diatomaceous mudstone deposited surface. At the top of the sequence is the Dos Palos
on the upper slope during times of upwelling during Shale Member, which is an approximately 175-m

Scott et al. 313


Figure 2. Outcrop local-
ities, reservoir units, and
sample sites (see Table 2
for sample characteristics).
Reservoir units (1–4) of the
northern and transitional
PGIC are marked. (A) No
Name Canyon (NoC). Here,
a stepped sill measuring
more than 6 m (20 ft) thick
with a multilayered lower
margin interbedded with
slope turbidites extends
more than 440 m (1444 ft)
laterally and connects to
remobilized slope-channel
sandstones located approxi-
mately 200 m (656 ft) below
by 0.5- to 1.0-m (1.6–3.3 ft)–
wide thickness sandstone
dikes. (B) Marca Canyon
(Mc). Multilayer sills with
individual thickness of
0.01 to 0.3 m (0.033–0.98 ft)
are found along the same
stratigraphic horizon as
thin-bedded upper slope
siltstones in unit 2. (C)
Capita Canyon (Ca). A high-
angle dike crosscuts more
than 500 m (1640 ft) of
stratigraphic section in
unit 1, thereby connecting
proximal-fan channelized
sandstone positioned at
different stratigraphic lev-
els. (D) Dosados Canyon
(Do). A low-angle dike
(5–12 m [16–39 ft] thick)
crosscuts units 2 and 3.
(E) Escapardo Canyon (Es).
A stepped sill approximately
1.0 to 2.0 m (3.3–6.6 ft)
thick extends for 180 m
(591 ft) and connects to
remobilized isolated slope-
channel sandstones in the
underlying Dosados Mem-
ber by high-angle dikes. See
Appendix 11 (AAPG Data-
share 46, www.aapg.org
/datashare) for raw data.
PUB = parent unit belt;
IB =intrusive belt; SEB =
extrusive belt.

314 Reservoir Characterization of a Sand-Injectite Complex


Figure 2. Continued.

(574-ft)–thick mudstone with minor siltstones de- seep carbonates (Schwartz et al., 2003; Minisini and
posited on the upper slope to the shelf edge and Schwartz, 2006; Vigorito et al., 2008).
intercalated with the Cima Sandstone Lentil sand- Overall, the stratigraphic section from the
stones that consist of sandstone extrudites and cold- lowermost Panoche Formation to the uppermost

Scott et al. 315


316 Reservoir Characterization of a Sand-Injectite Complex
Moreno Formation records a shoaling from a basin- lected, their positions and orientations were recorded
floor setting in the middle Maastrichtian to a slope and were then air dried for 24 hr to remove for-
and shelf setting in the Danian (Anderson and Pack, mation fluids.
1915; McGuire, 1988). The occurrence of geneti- Permeability data were acquired using a steady-
cally related sand extrudites dates injection to the state probe permeameter (Halvorsen and Hurst,
Danian (65–63 Ma) (Vigorito et al., 2008). 1990; Sutherland et al., 1993; Hurst and Goggin,
1995; Hurst et al., 1995) to determine the perme-
ability of a small volume of sandstone intrusion and
compare this with its petrographic microtexture
METHOD (Hurst and Rosvoll, 1991; Antonellini and Aydin,
1994; Appendices 13–14). Oriented samples were
The methods of outcrop and laboratory data ac- taken from the same locations as probe permea-
quisition were similar for all studied localities and meter measurements. Samples selected for thin
samples (Appendix 1; see AAPG Datashare 46 sections were impregnated with blue araldite dye
[www.aapg.org/datashare]). Geographic and strati- to highlight porosity and were then mounted on
graphic positions of sandstones were located using glass slides measuring 75 × 23 to 46 mm (3.0 ×
high-resolution satellite images followed by field 0.91–1.8 in.) wide, ground to a thickness of ap-
observations (Appendices 2–10; see AAPG Data- proximately 30 mm, and prepared without glass
share 46 [www.aapg.org/datashare]). Sandstone- cover slips. Mineralogy, grain- and pore-size distri-
body geometry was mapped, photographed, and bution (Appendix 15), and visual (Øv) and minus-
logged. The stratigraphic intervals were identified cement (Øm) porosity estimates were determined by
by logging vertical sections along canyons including point counting (∼200 counts per thin section) and
Moreno Gulch (MG), No Name Canyon (NoC), were compared with sill thickness (Appendix 16).
Capita Canyon (Ca), Rossetta Canyon (Ro), and Here, visual porosity is a measure of the void spaces
Right Angle Canyon (RA) (Figure 1C). From these highlighted by blue dye in thin section and is a
sections, the net-to-gross (N/G) ratio and proportion fraction of the point-counted volume of blue dye
of parent sandstone units, sandstone intrusions, and over the total volume as a percentage between 0
sandstone extrudites were calculated (Appendix 11). and 100%, and minus-cement porosity is the per-
Estimates of the N/G of high-angle dikes were made cent of both void space (blue dye) and volume of
using image-analysis software (Appendix 12). De- cements and is determined by the ratio of point-
tailed sampling of architectural elements was con- count totals for void space + carbonate cement +
ducted at outcrops including Marca Canyon (Mc), quartz cement to the total point counts. Selected
Dosados Canyon (Do), and Escapardo Canyon (Es). thin sections were polished and carbon coated. An
Where samples of sandstone intrusion were col- Oxford Instruments Limited cathodoluminescence

Figure 3. The lithostratigraphy, thickness, net-to-gross (N/G) ratio, and proportions of architectural elements in reservoir units (1–4) of
the northern and transitional Panoche giant injectite complex. Because the base of the Marca Shale Member is a flooding surface, it has
been used as a datum for correlating logged sections. Each unit has a different proportion of undeformed depositional units, remobilized
parent sandstone units, sandstone sills, low-angle dikes, sandstone extrudites, and host mudstones: Unit 1 is characterized by remobilized
depositional sandstones in the Panoche Formation and Dosados Member, which acted as parent units that source fluids and sand to the
overlying intrusive complex (Vigorito et al., 2008). Unit 1 is characterized by high N/G ratios; unit 2 is dominated by sandstone sills (4–
30%) and low-angle dikes (<1–17%) in the Dosados and Tierra Loma members and has high N/G ratios (0.04–0.30 sandstone); unit 3 is
dominated exclusively by high-angle dikes in the upper Tierra Loma Shale Member, Marca Shale Member, and the Dos Palos Shale Member
and has consistently low N/G ratios (<0.01–0.05, see Appendix 12 for raw data); and unit 4 is defined by extrusive sandstone associated with
carbonate cold-seep communities in the Dos Palos Shale Member (Schwartz et al., 2003). MG = Moreno Gulch; NoC = No Name Canyon;
Mc = Marca Canyon; Ca = Capita Canyon; Ro = Rosseta Canyon; Do = Dosados Canyon; Es = Escapardo Canyon; RA = Right Angle Canyon;
WT = West Tumey; PUB = parent unit belt; IB = intrusive belt; EB = extrusive belt; LAD = low-angle dikes; MLS = multilayer sill; STS =
staggered sill; SLT = thin-bedded slope turbidites; STP = stepped sill; RMB = remobilized slope channel; Fm = Formation; Mbr = Member.

Scott et al. 317


318 Reservoir Characterization of a Sand-Injectite Complex
detector on a LINK Analytical AN101055S electron- equivalent to EB. In log section, the upper contact
dispersive system was used to determine the chem- of unit 1 is defined by the top of the shallowest
ical grain mineralogy and morphology, pore shape parent sandstone unit encountered in the PUBb. In
and fill, cement composition, and microtextures reality, remobilized units may occur shallower in
under backscattered-electron detector mode. the section but were simply not encountered in the
Permeability maps were constructed by mea- line of the logged sections. The lower contact of
suring permeability on samples collected along and unit 1 is not defined because evidence of remobi-
across intrusions on orthogonal grids. A preferred lization and injection in the Panoche Formation
orientation to gridding was applied by weighing data occurs more than 500 m (1640 ft) below the base
points located along one axis with respect to those of the logged sections. The upper contact of unit 2
located on another axis to ensure that no gridding is defined by the top of the shallowest sandstone
in the surrounding low-permeability host strata ex- sill encountered in the IBa. The upper contact of
ists. Gridding was preferentially applied along the unit 3 is defined by the first occurrence of the
vertical and horizontal axes in high-angle dikes and sandstone extrudite and is equivalent to the upper
sills, respectively. Smoothing (spline) functions were contact of the IBb. The upper contact of unit 4 is
used to remove angular contours and to eliminate defined by the upper contact of the sandstone ex-
noise. trudite that is interbedded with the Cima Sand-
stone Lentil unit and coincides with the upper con-
tact of the EB.
RESERVOIR ARCHITECTURE
Unit 1
A genetically linked parent unit belt (PUB), intru-
sive belt (IB), and extrusive belt (EB) is present This unit contains depositional sandstones of the
throughout the PGIC (Figure 1B; Vigorito and upper Panoche Formation and the lower Moreno
Hurst, 2010). The PGIC is divided into three geo- Formation (Dosados Member) (Figures 2, 3). De-
graphical sectors: the northern (MG to RA), transi- positional sandstones are locally deformed in the
tional (RA to West Tumey [WT]), and southern upper Panoche Formation and are more intensely
(Tumey Gulch) sectors (Figure 1C). Based on the deformed in the Dosados Member and, together
PGIC lithostratigraphy of Vigorito and Hurst and with low-angle dikes, they constitute the dominant
on logged sections measured along several canyon proportion of sandstone. In the upper Panoche For-
sections in the northern and transitional sectors mation, high-angle dikes are only occasionally pre-
(Figure 2), we define the architecture and quantify sent (Figure 4A), where they crosscut stratigraphy,
the reservoir characteristics of four units (Figure 3): thereby connecting depositional sandstones that
unit 1 equivalent to PUBa and PUBb, unit 2 equiv- were originally separated by mudstones. The con-
alent to IBa, unit 3 equivalent to IBb, and unit 4 tact between dikes and depositional sandstones is

Figure 4. The external geometry and internal sedimentary structure of remobilized depositional sandstones and sandstone intrusions
of unit 1; lithologies represented in the photographs are mudstone (dark brown) and sandstone (light brown). (A) A high-angle dike
approximately 600 m (197 ft) long crosscuts three vertically staked sandstone units separated by mudstone within the upper Panoche
Formation: (i) the dike (dashed black line) connects depositional sandstones separated by thick (30 m [98 ft]) host mudstones at different
stratigraphic levels; (ii) line-drawing interpretation of (i). The dike splits and rejoins along its course. Upper Panoche Formation, Capita
Canyon (Ca). (B) A closer view of a 1.0- to 2.0-m (3.3–6.6-ft)–thick high-angle dike from Capita Canyon. (C) Detail of the sharp, planar
dike margin (dashed white line). Fluid-escape structures occur at the margin. (D) Remobilized sandstones in unit 1: (i) sandstone bodies
displaying winglike geometries at their lateral margins (white arrows) separated by a raft of mudstone (gray arrow); (ii) line drawing of
(i). High-angle dikes (black arrows) are connected to the upper margins of the parent unit, resulting in very good vertical connectivity.
Dosados Member, Moreno Gulch (MG). (E) Internal structure of remobilized depositional sandstones: (i) structureless sandstones
crosscut by a high-angle dike. The sandstone body (0.03 m [0.098 ft] thick) has a sharp contact; (ii) interpretation of (i). PUB = parent unit
belt; IB = intrusive belt; EB = extrusive belt.

Scott et al. 319


320 Reservoir Characterization of a Sand-Injectite Complex
Figure 6. Detail of multilayer sills in Figure 5E. (i) Individual sills are separated by thin (0.05–0.2 m [0.16–0.66 ft] thick) blocks of
mudstone. (ii) Line-drawing interpretation showing that the sills locally amalgamate into a composite sill with a cumulative thickness of
approximately 1 m (3.3 ft).

commonly sharp (Figure 4B, C). In the Dosados sandstones (cf. black intrusions in Figure 4Dii).
Member, channelized sandstones are extensively Internally, the depositional sandstones are crosscut
remobilized where, at the margins and sometimes by dikes (Figure 4Ei, ii).
from the top of the channelized unit, low-angle
dikes (30–40° relative to bedding) form 5- to 15-m
(16–49-ft)–thick winglike projections (white ar- Unit 2
rows, Figure 4Di, ii), resulting in isolated rafts of
host mudstone (gray arrow). These dikes enhance This unit is a 100- to 290-m (328–951-ft)–thick
the lateral and vertical connectivity of channelized mudstone-rich interval characterized by sandstone

Figure 5. Unit 2. (A) Staggered sills defined by sharp planar margins that are laterally and vertically offset by high-angle dikes: (ii) in-
terpretation of (i). Tierra Loma Shale Member, Marca Canyon (Mc). (B) A stepped composite sill (steps, >0.3 m [0.98 ft] in height; black
arrow). This sill arrangement is shown in Hurst et al. (2007, their figures 18, 19). Tierra Loma Shale Member, Escapardo Canyon (Es). (C)
Overview of a stepped sill with an erosional (scalloped; Figure 5D) upper margin and multilayered lower margin interbedded with thin-
bedded slope turbidites. In places, the sill is more than 6 m (19.7 ft) thick and, at outcrop, extends more than 440 m (1444 ft) laterally.
Tierra Loma Shale Member, No Name Canyon (NoC). (D) Convex-upward sandstone intrusions (white line) that cut into the overlying host
mudstone, termed “scallops” (SC; sensu Hurst et al., 2005): (i) a similar projection is shown in Vigorito et al. (2008, their figure 3B); (ii) line
drawing of a scallop; (iii) detail of the discordant margin. The SC are characterized by a sandstone intrusion that forms steep (70–80°),
discordant, and erosional flanks and an upper crest that cuts as much as 2 m (6.6 ft) (white line) of overlying host mudstone. (E) Multilayer
sills with individual thickness of 0.01 to 0.3 m (0.033–0.98 ft). Individual sills are separated by thin (0.05–0.2 m [0.16–0.66 ft] thick) rafts of
host mudstone. Tierra Loma Shale Member, Marca Canyon (Mc). Black boxes locate the permeability maps shown in Figure 12.

Scott et al. 321


Figure 7. Unit 3. (A) A high-angle dike that crosscuts approximately 200 m (656 ft) of the Marca Shale Member. The black box locates
the permeability map shown in Figure 16A. The locality is the No Name Canyon (NoC). (B) High-angle dikes crosscutting the Marca Shale
Member (Marca Canyon [Mc]). (C) A high-angle (80–90°), large thick (0.5 m [1.6 ft] wide) dike crosscutting the Marca Shale Member in
Marca Canyon. (D) Detail of the steep high-angle dike in C. The smaller dike on the left deviates from the main dike at an angle of
approximately 30°.

sills within the Dosados and Tierra Loma mem- voir units 2 and 3. The sills have three geometries,
bers (Figures 2, 3); the average N/G ratio is 0.19. from stratigraphically deeper to shallower, stag-
The greatest proportions of sandstone-rich reser- gered (Figure 5A), stepped (Figure 5B, C, D), and
voir facies are sills (average, 13%) and low-angle multilayer (Figure 5E).
dikes (average, 5%). Low-angle dikes are locally Staggered sills have intermediate thickness
present and tend to depart at high angles (>40°) (<1.0 m [3.3 ft]) and extend laterally <10 m (33 ft)
from remobilized channelized sandstones in unit 1 (Figure 5A). Furthermore, upsection stepped com-
and become lower angle (5–17°) upward in reser- posite sills that are thicker (as much as 11.0 m

322 Reservoir Characterization of a Sand-Injectite Complex


Figure 8. Extrusive sandstones in unit 4, the Dos Palos Shale Member, Marca Canyon (Mc). (A) Extrusive sandstones: (i) a high-angle dike
terminating at the base of a layer of extruded sandstone (dashed black line); (ii) interpretation showing the high-angle dike connected to
the base of the sandstone extrudite. (B) Internal sandstone extrusion geometry. (i) mounded morphology with internal subvertical
laminae that represent the vents of sand volcanoes with surrounding draped low-angle bedded sandstone, which commonly display soft-
sediment deformation; (ii) interpretation showing steeply inclined laminae dipping away from a central conduit. IB = intrusive belt; EB =
extrusive belt.

[36 ft]) and more laterally continuous (>400 m 2-m (3.3–6.6-ft)–thick upper main sill and sev-
[1312 ft] long), with stepped, erosive upper and eral lower centimeter-thick (0.01–0.1 m [0.033–
lower margins (Figure 5B, C, D), occur. The sill 0.33 ft]) multilayered sills. The main sill is con-
complex is composed of an approximately 1- to nected to dikes (black lines, Figure 5B). High-angle

Scott et al. 323


324
Table 2. Outcrop Localities (Canyon Sections) and Architectural Elements Sampled in the Northern Panoche Giant Injection Complex*

No Name Canyon (NoC)


A, <1 (31)% B, 2 (22)% C, 15 md D, 30 (30)%; E, 15 (15)%; F, 27 md G, <1 (1)% H, 3 (3)% I, 1 md J, <1–5
1937 md 384 md (5–19)%;
Reservoir Characterization of a Sand-Injectite Complex

4–13 md
HAD** HAD HAD LAD** LAD HAD HAD HAD MLS** MLS

K, 3–30 L, 20 md M, 5 (5)%; N, 5–6 O, 3 (3)%; P, 3 (3)%; Q, 13 (13)%; R, 20 (20)% S, 4 (4)%; U, 85 md


(16–30)%; 34 md (6–8)%; 16 md 25 md 130 md 46 md
1780–4894 md 48–52 md
STP** HAD HAD HAD HAD HAD HAD HAD HAD HAD

V, 56 md W, <1 (19)%;
3 md
HAD MLS

Marca Canyon (Mc)


A, 0.5 (0.5)%; B, 1.5–4 C, <1 D, 7 (14)%; G, 2 (2)%; H, 4 (4)%; E, <1–13 (3–8)%; F, 2–10 (4–10)%; I, 3 (26)% J, 3 (20)%;
3 md (25–34)%; (14–16)%; 11 md 19 md 9 md 14–196 md 11–117 md 21 md
3–17 md 3–17 md
HAD PFC** PFC HAD HAD HAD MLS MLS MLS MLS

Capita Canyon (Ca)


A, 2 (20)%; B, <1 (35)%; C, 4 (26)%; D, 2 (28)%; E, 1 (32)%; F, 5 (26)%; G, <1 (1)%; H, <1 (28)%; I, 3 (33)%; J, 5 md
50 md 52 md 33 md 63 md 42 md 63 md 1 md 4 md 66 md
PFC PFC PFC PFC PFC PFC PFC PFC PFC HAD

K, <1 (21)%; L, 25 md M, 1 md N, 4 (20)%; O, <1 (10)%; P, <1 (719)%; Q, 4 (18)%; R, 52 md S, 6 md T, 1 md


5 md 122 md 5 md 4–6 md 16 md
HAD HAD HAD HAD HAD HAD HAD HAD HAD RLC**

U, 22 (31)%; V, 14 (14)%; W, 9 md X, 32 (32)%; Z, 14 md AA, 2 md BB, 21 md CC, 182 md DD, 3 (9)%; EE, 10 (10)%;
1147 md 296 md 4068 md 1 md 76 md
LAD HAD HAD STS** HAD HAD HAD MLS HAD HAD

FF, 3–34 md GG, 43–72 md HH, 10 md


HAD HAD HAD
Dosados Canyon (Do)
A, 5–14 (5–14)%; B, 3 (3)%; C, 3 (3)%; D, 6 (11)%; E, 5–8 (5–8)%; F, 4 (4–18)%; G, 4–10 H, 7–26 I, 2 (2)%; J, 5–10
45–231 md 268 md 65 md 65 md 91 md 458 md (4–10)%; (7–26)%; 8 md (10–12)%;
32–105 md 466 md 270 md
HAD HAD HAD HAD HAD STS STS STS STS STS

K, 8 (9)%; L, 14 (15)%; M, 53–57 md N, 5–32 O, 11–28 P, 18 (19)%; Q, 18 (18)%; R, 4–5 (4–5)%; S, 367 md T, 13 (13)%;
56 md 294 md (5–32)%; (28–31)%; 220 md 302 md 131–302 md 131 md
190–1655 md 727 md
STS STS STS STS LAD LAD LAD LAD LAD LAD

U, 6305 md W, 27 (27)%; X, 40 md Y, 17 md Y, 30 md Z, 75 md AA, 40 md BB, 298 md CC, 3–6 DD, 16


4948 md (3–6)%; (16)%;
20–38 md 268 md
LAD LAD LAD STS HAD HAD HAD HAD HAD HAD

EE, 179 md FF, 2 (2)%;


24 md
HAD HAD

Escapardo Canyon (Es)


A, 1 (22)%; B, 0 (22)%; C, 26 (26)%; D, 13 (13)%; E, 12 (12)%; F, 18 (18)%; G, 27 (27)%; H, 432 md I, 12 (12)% J, 432 md
19 md 3 md 1331 md 266 md 2132 md 287 md 1517 md
PFC PFC STP STP STP STP STP STP STP STP

K, 19–31 L, 634 md M, 228 md


(20–31)%;
260–3048 md
STP STP STP
*In each canyon, the samples are assigned a letter. The numbers outside the brackets represent visual porosity (%) and permeability (md), whereas minus-cement porosity (%) is given within the brackets. Locations of samples of
Scott et al.

sandstone are labeled with the corresponding letter along canyon sections in Figure 2, allowing the reservoir characteristics of the sample to be located within the injectite complex and regional stratigraphy.
**HAD = high-angle dike; LAD = low-angle dike; MLS = multilayer sill; STP = steeped sill; PFC = proximal-fan channelized sandstones; RLC = remobilized lower slope channelized sandstones; STS = staggered sill.
325
Table 3. Overview of the Mean Mineral Composition and Petrographic Characteristics of the Panoche Giant Injection Complex Sandstones*

Element Remobilized Parent Units Sandstone Intrusions Sandstone Extrudites

Sample number (n) 28 121 7


Grain size
mm 0.2 0.2 0.2
phi 2.4 2.4 2.4
Detrital composition (%)
Qn, Fn, Ln Q42F40L18 Q43F45L12 Q47F42L12
Mica 1.8 1.7 1.6
Mud 3.5 12.9 7.9
Glauconite 0.9 0.8 0.4
Cements (%)
Quartz 1.7 1.1 2.4
Carbonate 11.7 1.7 20
Mud matrix (%) 5.0 5.3 0.8
Minus-cement porosity (%) 16.0 12.9 23.5
*Mineral composition and minus-cement porosity are given in percentage of whole-rock volume. Relative percentages of quartz (Q), feldspar (F), and lithics (L) exclude
detrital mudstone clasts.

dikes crosscut the stratigraphy and connect to the High-angle sandstone dikes range in thickness
lower and upper margins of the sill (black lines, from less than 0.1 to more than 0.6 m (0.33–1.97 ft),
Figure 5C). Close to the upper contact of unit 2, crosscut host mudstones at 80 to 90° (relative to
multilayer sills are located along or near the strat- bedding), and are laterally continuous, in places
igraphic horizon where thin-bedded, finer grained crosscutting more than 200 m (656 ft) of strati-
depositional sandstone units are found (Figure 2). graphic section, thereby producing excellent ver-
The sills tend to be thin (0.01–0.3 m [0.033– tical but limited lateral connectivity (Figure 7A, B).
0.98 ft] thick) and laterally continuous (10–100 m The dikes are en echelon in the lower 60 to 80 m
[33–328 ft] wide), emanate from a high-angle dike (197–263 ft) and bifurcate in the upper 140 to
(black lines, Figure 5E), and are separated by 160 m (459–525 ft) (Figure 7C, D). In the upper
mudstone units. Individual sills propagate along 100 m (328 ft), the dikes eventually taper out or
bedding horizons and locally amalgamate into ap- terminate beneath sandstones in unit 4.
proximately 1-m (3.3-ft)–thick units (Figure 6).
Unit 4
Unit 3
An interval of extruded sandstone of as much as
Unit 3 is a 330- to 440-m (1083–1444-ft)–thick 30 m (98 ft) thick that overlies an unconformity
mudstone-dominated interval of exclusively high- within the Dos Palos Shale Member comprises
angle sandstone dikes within the Tierra Loma, Marca unit 4 (Figures 2, 3). Because unit 4 is defined by
Shale, and Dos Palos Shale members (Figures 2, 3). the presence of sandstone, the N/G ratio is typi-
Net-to-gross ratio is very low (average, 0.02) when cally 1. High-angle dikes connect to the base of unit
derived from logged sections; the N/G determined 4 (Figures 2, 8Ai, ii), thus establishing connectivity
from the image analysis of high-angle dikes in out- among sandstones in units 1 to 4.
crop cross sections indicates a significantly higher Sandstones in unit 4 can be as much as 30 m
N/G (average, 0.08). Locally, thick (5–12 m [16– (98 ft) thick and extend laterally for at least 100 m
39 ft]) low-angle dikes originating in unit 2 crosscut (328 ft), but in Marca Canyon, the sandstone units
this unit, for example, the dike studied in Do (black are locally thin (∼5 m [16 ft] thick) and laterally
box, Figure 2D). continuous (>30 m [98 ft] wide) (Figure 8Ai, ii).

326 Reservoir Characterization of a Sand-Injectite Complex


Scott et al.

Figure 9. Visual and minus-cement porosity distribution of sandstone intrusions. (A) Total sill and dike population. (B) Total sill population. (C) Staggered sills. (D) Stepped sills.
(E) Multilayer sills. (F) Total dike population. (G) High-angle dikes. (H) Low-angle dikes. av = average; sd = standard deviation. See Appendix 14 (AAPG Datashare 46 at www.aapg
327

.org/datashare) for sample data.


Table 4. Comparison of Low-, Moderate-, and High-Permeability (Injected) Sandstones, Showing Mean Values*

Low (Poor) Permeabilities Moderate (Good to Very Good) High (Excellent)


Petrographic Variable (<0.1–100 md) Permeabilities (100–1000 md) Permeabilities (>1000 md)
Grain size
mm 0.2 0.2 0.2
phi 2.4 2.4 2.3
Mudstone clast (%)
Cements (%) 13 11 9
Quartz 2 <1 <1
Carbonate 1 <1 0
Mud matrix (%) 8 4 1
Visual porosity (%) 4 14 28
Minus-cement porosity (%) 7 15 29
Mean permeability (md) 30 266 2457
Intrusive element (%) 54% are high-angle dikes 53 and 35% are low-angle 31 and 69% are low-angle
dikes and sills, respectively dikes and sills, respectively
Sample number (n) 53 40 13
*The mineral compositions are given in percentages of whole-rock volume. Consult Appendix 14 for sample data.

Mounding is common and locally contains steeply (Figure 10B). The pore network is well connected
inclined laminae (Figure 8Bi) that dip away from (range of length, 0.1–0.3 mm [0.004–0.012 in.]
central conduits, the sandstone-filled vents of sand wide), and large pore-throat diameters (0.05–
volcanoes. 0.1 mm [0.002–0.004 in.] wide) with high visual
porosity (12–19%; Figure 9C, D) and a broad dis-
tribution of pore-throat diameters (dashed line;
RESERVOIR CHARACTERISTICS Figure 11A) are predominant. Individual quartz
grains are pervasively crosscut by fractures and are
Description sometimes fragmented into arrays of smaller clasts
(Figure 10B). Fine to medium sand-size mudstone
Sandstones were investigated from all reservoir units clasts are common constituents of multilayer sills
with sampling focused on units 2 and 3 (Figure 2; (Figure 10C) along with clay- and silt-size quartz
Table 2). All sandstones are arkosic to lithic arkosic grains that occlude pores (Figure 10D). This pro-
in composition (Table 3). A broad range of visual duces poor sorting and a poorly connected pore
porosity is present in the sandstone intrusions (2– network with small pore-throat diameters (average,
33%; Figure 9A). Permeability is variable (sd = 0.04 mm [0.002 in.] wide), low visual porosity (5%;
884 md). High-permeability data (>1000 md) are Figure 9E), and narrow pore-throat diameters (solid
predominantly from sills, whereas lower permeabil- line, Figure 11A). The abundance of silt-size and finer
ities (<0.1–100 md) are mainly from high-angle dikes grained clay particles in multilayer sills (Figure 10C,
(Table 4). D), when compared with similar grains in stepped
Forty-nine thin sections were described. The to- sills (Figure 10A, B), explains the lower perme-
tal sill population has mean visual and minus-cement ability of more than 2 orders of magnitude in the
porosities of 13 and 15%, respectively (Figure 9B). multilayer sills.
Staggered and stepped sills tend to be moderately Permeability measurements were taken on 57
sorted and loosely packed (Figure 10A) and have samples, and a positive relationship between sill
open pore space in which some grains appear to thickness and permeability is noted (Table 5). Ac-
float without obvious contact with adjacent grains quisition of permeability data from three localities

328 Reservoir Characterization of a Sand-Injectite Complex


Figure 10. Microtextures of sandstone sills from unit 2. Porosity is green to blue and black in plane-polarized light and backscattered-
electron microscopy modes, respectively. (A) A high-permeability sandstone (4893 md) from a stepped sill (location in Figure 12B). Note
the presence of fine to medium sand-size mudstone clasts (Md; black grains). Plane-polarized light. (B) Backscattered-electron mi-
croscopy image of the pore structure in a stepped sill that includes fractured grains. (C) A tightly packed, poorly sorted grain fabric from a
low-permeability sandstone (24 md) from a multilayer sill (location in Figure 12C). Grains of quartz (Qz) and Md are surrounded by finer
silt-size quartz grains and fine sand-size Md (black grains). Plane-polarized light. (D) Backscattered-electron microscopy displaying pores
between Qz and feldspar (F) grains in a multilayer sill. Clay-size particles occlude pores.

with sills of differing thickness shows that stepped against permeability, a modest positive correlation
sills have high permeability (green to blue areas; exists (r 2 = 0.69; Figure 14).
Figure 12A, B) relative to multilayer sills (red
shades, Figure 12C). Note the lower permeability Sandstone Dikes
sandstone present toward the sill margins (286–
299 md; red and orange areas, Figure 12A). The total Seventy-two thin sections were characterized. The
sill population has a mean of 629 md (Figure 13B), total dike population has mean visual and minus-
and the mean permeability of the staggered and cement porosities of 7 and 11%, respectively (Fig-
stepped sills is 586 (Figure 13C) and 1225 md ure 9F). High-angle dikes are tightly packed and
(Figure 13D), respectively. Multilayer sills have poorly sorted, with elongate grains displaying sub-
a mean permeability more than seven times less vertical alignment (dashed circles, Figure 15A). Frac-
(70 md; Figure 13E). When thickness is plotted tured quartz and feldspar grains are ubiquitous

Scott et al. 329


Figure 11. Pore-throat diameter distributions in sandstone sills (A) and dikes (B) from point counting versus cumulative percent visual
porosity. (A) The mean porosity within stepped sills (dashed black line; average, 0.63-mm [0.03-in.] pore diameter) is greater than in
multilayer sills (solid black line; average, 0.04-mm [0.002-in.] pore diameter). (B) The mean porosity in low-angle sandstone dikes
(dashed black line; average, 0.10-mm [0.004-in.] pore diameter) is larger than in high-angle sandstone dikes (solid black line; average,
0.05-mm [0.002-in.] pore diameter). See Appendix 15 (AAPG Datashare 46 at www.aapg.org/datashare) for data.

(dashed circles, Figure 15B). Where pores are oc- shades, Figure 16A, C). By contrast, a very thick,
cluded by carbonate cement, a disconnected pore low-angle winglike dike that originates near the top
network is created with small pore-throat diameters of unit 2 and extends upward into unit 3 is char-
(0.02 mm [0.0008 in.]), low porosity (Øv = 5.0%, acterized by high permeability (4947–6304 md;
average; Figure 9G), and a narrow distribution of purple shades, Figure 16B). The total dike popu-
pore-throat diameters (solid line, Figure 11B). lation has a low average permeability (202 md;
Low-angle dikes are typically moderately sorted Figure 13F), but high- and low-angle dike popula-
and have loose grain packing with high visual po- tions have significantly different average perme-
rosity (Figure 15C, D); this creates a highly con- ability, 81 and 529 md, respectively (Figure 13G, H).
nected pore network with large pore-throat diame-
ters (>0.15 mm [0.005 in.]), moderate porosity (Øv =
11%, average; Figure 9H), and a broad pore-throat DISCUSSION
diameter distribution (dashed line, Figure 11B).
One-hundred and fifteen permeability mea- Creation of quantitative static models that are re-
surements were made by sampling along and across alistic representations of the reservoir architecture
dikes. Detailed sampling of dikes at three outcrop and pay distribution is strongly dependent on the
localities shows that high-angle dikes have perva- interpretation and integration of seismic and bore-
sive low permeability throughout (<1–122 md; red hole data. This process is enhanced by the use of

Table 5. Comparison of Thin Multilayer (<0.1 m [0.33 ft] thick) and Thick Stepped (>0.1 m [0.33 ft]) Sandstone Sills*

Petrographic Parameter Sandstone Sills (<0.1 m [0.33 ft]) Sandstone Sills (>0.1 m [0.3 ft])
Average grain size (mm) 0.16 0.20
Mean visual porosity (%) 7 16
Permeability (md) 121 911
Microtextures Abundant silt-size grains, tight Infrequent silt-size grains, loose
packing, and poor sorting packing, and well sorted
*The thin sills contain tightly packed and poorly sorted grains, resulting in a low visual porosity and low permeability. Consult Appendix 19 for sample data.

330 Reservoir Characterization of a Sand-Injectite Complex


Figure 12. Permeability maps of sandstone sills in unit 2; the lithology represented is mudstone (gray), and sandstone permeability is
represented by a red-to-purple (low-to-high) color scale. (A) A thick stepped sill characterized by high-permeability sandstones (1561–
3047 md; green to blue areas). The gridding method used was natural neighbor (anisotropy of 0.1 and angle of 5°). (B) A very thick
stepped sill with a multilayered lower margin characterized by very high permeability sandstones (1780–4893 md; green to purple
areas). Note the low-permeability sandstone toward the sill margins (4–36 md; red and orange areas). The gridding method used was
natural neighbor (anisotropy of 0.1 and angle of 5°). (C) Multilayer sills characterized by low-permeability sandstones (10–20 md; red
areas). Note the area of higher permeability (58–117 md; yellow to green areas). The gridding method used was point krigging.

knowledge from subsurface and outcrop examples ervoir units 2 and 3. Lateral and vertical variations
that are geologically similar. As sandstone intrusions in sandstone geometry and N/G occur, and in
create cross-stratal permeable networks in other- general, the abundance of depositional sandstones
wise low-permeability strata, they are a significant and their proximity to the base of the sill zone, the
challenge when modeling their distribution and be- lithostatic equilibrium surface (Vigorito and Hurst,
havior during hydrocarbon production. The data 2010), determine the N/G of the sandstone intru-
we present from the PGIC form a basis for subsur- sions. Identification of unit 2 is important in sub-
face analogy and quantitative modeling (Figure 17; surface modeling because the main variations in
Table 6) that can enhance conceptual interpreta- reservoir volume occur in unit 2 where sandstone-
tions that infer vertical connectivity in situations rich features (N/G of as much as 0.30 sandstone)
where dikes create fieldwide pressure communica- that are commonly identified in the subsurface such
tion (Briedis et al., 2007). as sills and saucer-shape and conical intrusions oc-
cur (Hurst et al., 2003b; Szarawarska et al., 2010).
Reservoir Architecture At a smaller scale, sandstone-body geometry
varies within unit 2 where sill geometry is orga-
The tripartite organization of parent units (unit 1), nized stratigraphically (Figure 5). When examining
intrusive units (units 2 and 3), and sandstone ex- borehole data, the challenge is to identify different
trusions (unit 4) is used as the basis for defining a classes of sill, to differentiate between sills and de-
reservoir architecturally based zonation (Figures 1– positional sandstones, and to differentiate even be-
3). The intrusive unit is further subdivided into res- tween sill and dike margins simply because so little

Scott et al. 331


Figure 13. Permeability distribution in sandstone intrusions. (A) Total sill and dike population. (B) Total sill population. (C) Staggered sills.
(D) Stepped sills. (E) Multilayer sills. (F) Total dike population. (G) High-angle dikes. (H) Low-angle dikes. See Appendix 14 (AAPG Datashare
46 at www.aapg.org/datashare) for sample data. kh = horizontal permeability; kv = vertical permeability; n = number of samples; av =
average; sd = standard deviation; cv = coefficient of variation; kh = horizontal permeability; kv = vertical permeability.

rock is sampled. Diagnostic features of sand injec- orito and Hurst (2010; Table 1), but comparison
tion such as feeder dikes (black lines, Figure 5), with other sand injectites is compromised by the
upward erosive surfaces (white line, Figure 5D), less extensive exposure elsewhere. However, in sev-
clasts of host mudstone derived from the injectite eral cases, the tripartite organization is identified
margins (Figure 6), and meter-scale stepped ero- despite substantial differences in sandstone content.
sional margins (black arrow, Figure 5B, D) can be Our experience with the interpretation of subsur-
used collectively to differentiate sandstone intru- face sand injectites is that it is similar to the archi-
sions from depositional units. tecture in the PGIC, although the observation of
The general relevance of the tripartite organi- unit 1 (Figure 4) is typically poor (de Boer et al.,
zation to other sand injectites is discussed by Vig- 2007; Wild and Briedis, 2010).

Figure 14. Permeability (md) plotted


against sill thickness (m). Line of best fit is
plotted as a linear trend line. See Appendix
16 (AAPG Datashare 46 at www.aapg.org
/datashare) for data.

332 Reservoir Characterization of a Sand-Injectite Complex


Figure 15. Microtextures in dikes from units 2 and 3. (A) A tightly packed and poorly sorted texture with faint subvertical grain
alignment (dashed circles) in low-permeability sandstone (50 md) from a high-angle dike; Capita Canyon (location in Figure 16A).
(B) Backscattered-electron microscopy of a poorly sorted sandstone in a high-angle dike with fractured quartz and feldspar grains
(dashed circles) and pores that are pervasively carbonate cemented. (C) A moderately to loosely packed grain fabric in a high-permeability
sample (4947 md) from a low-angle dike; Dosados Canyon (Do) (location in Figure 16B). Large pores (>0.15 mm [0.006 in.] in diameter) are
common. (D) A loosely packed grain fabric in a low-angle dike; Dosados Canyon (Do). Md = mudstone clasts.

Reservoir Characteristics and Kawamura, 2002; Diggs, 2007). An alternative


origin is their elutriation from the remobilized par-
In sills, permeability has a moderate positive log- ent sandstone units (Duranti and Hurst, 2004).
normal correlation with intrusion thickness (Fig- Many of the clay-size particles were likely derived
ure 14). These correlate with variations in grain from shear-induced abrasion of host mudstone clasts,
sorting, packing, and abundance of clay- and silt- as fluidized sand entered the hydraulically fractured
size particles (Figure 10), which result in varia- strata (cf. Scott et al., 2009).
tions in visual porosity and pore-size distribution In contrast, the origin of the silt-size grains in
(Figures 9C, D, E; 11). Origins of clay-size par- multilayer sills is more obvious and restricted to a
ticles in sills include entrainment at the injectite single origin. The silt-size grains likely originated
margins as fluidized sand penetrated host mud- from the depositional turbidite beds located along
stone and disintegration of mudstone rafts through or near the stratigraphic horizon at which the mul-
shear-induced abrasion during injection (Kamakami tilayer sills intruded. The occurrence of thin-bedded

Scott et al. 333


Figure 16. Permeability maps of sandstone dikes from a stratigraphically deep (Panoche Formation) high-angle dike (A), a low-angle
dike (B), and a high-angle dike (C) (location in Figure 2). Lithologies represented are mudstone (light gray) and depositional sandstone
(gray); sandstone permeability is represented by the color scale. (A) Low-permeability sandstones characterize the entire dike (10–
20 md; red areas); Panoche Formation. (B) A thick low-angle dike in the Tierra Loma Shale Member, which originates in unit 2 and
crosscuts unit 3. (C) A high-angle dike characterized by low-permeability sandstones (40–80 md; green areas); Marca Shale Member.
PUB = parent unit belt.

sandy and silty turbidites appears to have had an pared to that in the staggered and stepped sills
important function in promoting the formation (Figures 12A, B; 13C, D).
of multilayer sills (see maps A–E in Figure 2), as Carbonate cement contributes to the occur-
these depositional units represent heterogeneities rence of very low visual porosity and permeability
within an otherwise mudstone-rich background of (Figures 9G; 13G; 15A, B) and a narrow pore-
unit 3. The incorporation of silt-size grains would size distribution in high-angle dikes (solid line,
occur through hydraulic mixing of fine to me- Figure 11B). Jonk et al. (2005a, b) emphasize the
dium sand-size injected grains with depositional importance of sandstone-intrusion geometry, di-
silt beds, thereby resulting in a lower permeability in mensions, and proximity to host strata in control-
the multilayer sills (Figures 12C; 13E) when com- ling carbonate cementation and suggest that cations

334 Reservoir Characterization of a Sand-Injectite Complex


Scott et al.

Figure 17. Conceptual three-dimensional (3-D) diagrams that synthesize the salient reservoir architectural elements described in this study. (A) Panoche giant injectite complex (PGIC)
reservoir geometry and volumetrics: (i) an interpretation of the 3-D geometry showing differing reservoir connectivity among architectural elements in units 1, 2, 3, and 4; (ii) estimates
of sandstone-intrusion volume in units 2 and 3 (Table 6). (B) Summary of the 3-D geometry and trends in reservoir characteristics of architectural elements in the PGIC: (i) low-angle
dike, (ii) staggered sill, (iii) stepped sill, (iv) multilayer sill, and (v) high-angle dike. PUB = parent unit belt; IB = intrusive belt; EB = extrusive belt; N/G = net to gross.
335
336
Reservoir Characterization of a Sand-Injectite Complex

Table 6. Summary of the Reservoir Characteristics and Their Trends of Architectural Elements

Dimensions** Porosity Characteristics† Permeability Characteristics††


Architectural Petrographic Trends in Reservoir
Element* y t Characteristics*** Øv Øm k kv/kh Characteristics

HAD <10–500 m 0.01–8 m f-s-s to m-s-s, Md, Mm, Average, 5%; Average, 5%; Average, 81 md; 1.4 Lateral continuity of low
(33–1640 ft) (0.03–26 ft) CC, Tgp, Pgs, VGA sd, 5% sd, 5% sd, 282 md porosity and permeability
LAD 10–100 m <1–15 m f-s-s to m-s-s, Mgp, Average, 11%; Average, 13%; Average, 529 md; 1.2 Lateral continuity of high
(33–330 ft) (3.3–49 ft) Pgs to Mgs sd, 7% sd, 9% sd, 1389 md porosity and permeability
STS 1–10 m <1 m (3.3 ft) f-s-s to m-s-s, Tgp to Average, 12%; Average, 14%; Average, 586 md; 1.3 Lateral continuity of moderate
(3.3–33 ft) Mgp, Mgs, HGA sd, 11% sd, 11% sd, 1058 md porosity and permeability
STP 10–400 m <1–12 m f-s-s to m-s-s, Md, Mgp Average, 19%; Average, 22%; Average, 1224 md; 1.8 Lateral continuity of high
(33–1312 ft) (3.3–39 ft) to Lgp, Ggs, HGA sd, 10% sd, 22% sd, 1300 md porosity and permeability
MLS 10–50 m 0.05–0.3 m s-s to m-s-s, Md, Mm, Average, 5%; Average, 6%; Average, 70 md; 1.4 Lateral continuity of very low
(33–164 ft) (0.16–0.98 ft) Tgp, Pgs sd, 4% sd, 5% sd, 64 md porosity and permeability
*HAD = high-angle dike; LAD = low-angle dike; STS = staggered sill; STP = stepped sill; sd = standard deviation; MLS = multilayer sill.
**y = lateral extension; t = thickness.
***f-s-s = fine sand-size grains; m-s-s = medium sand-size grains; s-s = silt-size grains; Md = mudstone clasts; Mm = mud matrix; CC = carbonate cement; Tgp = tight grain packing; Pgs = poor grain sorting; VGA = vertical grain
alignment; Mgp = moderate grain packing; Mgs = moderate grain sorting; HGA = horizontal grain alignment; Lgp = loose grain packing; Ggs = good grain sorting.

Øv = average visual porosity of element; Øm = average minus-cement porosity of element.
††
k = average permeability; kv/kh = average ratio of vertical-to-horizontal permeability.
kilometer of GRV;

**RU = reservoir unit; t = average thickness (m); l = length of outcrop exposure of reservoir unit (m); w = inferred extension of Panoche giant injection complex into the San Joaquin Basin (m); N/G = average net to gross of reservoir
partly derived from the host mudstones contributed

IPV per cubic

0.001 km3
0.02 km3
to the formation of carbonate cement. The high

Table 7. Injected Sandstone and Pore-Volume Estimates in Units 2 and 3 Based on Averaged Sedimentary Log Data in the Panoche Giant Injection Complex (cf. Figure 2)*
angle of discordance between the sandstone intru-
sions and bedding in the host strata may create a
contact that enhances fluid flux along the bedding
planes (i.e., higher permeabilities along bedding
compared to across bedding). Enhanced flux from

6.33 × 107
0.83 × 107

unit; Øv = average visual porosity of injected sandstones within the reservoir unit (%); GRV = gross rock volume (m3); ISV = intrusive sand volume (m3); IPV = intrusive pore volume (m3).
*Estimates assume that the the Panoche giant injection complex extends 11.1 km (6.9 mi) into the San Joaquin Basin (i.e., the same dimensions as the northern and transitional outcrop belt).
IPV**
the surrounding host mudstones into a dike would
carry with it cations that are conducive to carbonate
cementation. It is reasonable that a continued influx
of cations would pervasively reduce sandstone per-
meability in high-angle dikes (Figure 16A, C).

kilometer of GRV
ISV per cubic

0.19 km3
0.02 km3
Reservoir Volumetrics

The PGIC constitutes large volumes of sandstone


intrusion within a predominantly mudstone-rich
(low N/G) background. By using average N/G

5.43 × 108
1.10 × 108
and thicknesses derived from the logged sections,

ISV**
the volumes of sandstone intrusion in reservoir
units 2 and 3 in the northern and transitional PGIC
are estimated as 0.19 km3 (0.046 mi3) and 0.02 km3
(0.005 mi3), respectively, of intruded sandstone per

2.85 × 109
5.54 × 109
GRV**
cubic kilometer of gross rock volume (Table 7).
In the subsurface, sandstone intrusions are likely
to have much greater volume than the PGIC; for
example, in the North Sea, the Paleogene sand-
injectite play has hydrocarbon reserves estimated
Øv (%)**

to be approximately 2.5 billion bbl of oil equivalent


12
7
(BOE) (Hurst et al., 2005), all of which tend to
be reservoired only in the upper parts of injectite
complexes (figure 15 in Jonk, 2010; figure 2 in
N/G**
0.19
0.02

Vigorito and Hurst, 2010). Hence, the total vol-


ume of these subsurface sandstone intrusions is
much greater. The upper part corresponds to units
2 and 3, but in the North Sea Paleogene, unit 2 is
1000
1000
w**

typically considerably thicker relative to unit 3 and


is more sandstone-rich than the PGIC. To the best
of our knowledge, sand extrudites (unit 4) have not
16,000
16,000

been positively identified as significant reservoirs in


l**

any subsurface examples (Hurst et al., 2006).


Mudstone distribution within PGIC sandstone
intrusions has a major influence on reservoir qual-
179
346
t**

ity. Large rafts of mudstone are common in sills


and reduce their N/G (Figure 6Ai, ii, B, C). Some
of the large clasts are in situ and formed during
RU**

the hydraulic fracturing of the host mudstone


2
3

Scott et al. 337


(Cosgrove, 2001); these large clasts commonly solved using seismic data can be reconciled by ref-
have sandstone-filled microfractures that contrib- erence to outcrop analogs that show the abundance
ute to hydrocarbon storage and enhance intraclast of small-scale sandstone intrusions and by consid-
permeability. Dikes and sills both have varied con- eration of production data that show substantially
tents of fine- to coarse-grained mudstone parti- that more hydrocarbon is present than was pre-
cles (Md; Figures 10A, C; 15A) probably derived dicted from preproduction reservoir models.
from hydraulic fracturing and erosion of host strata
(Archer, 1984; Diggs, 2007; Macdonald and Flecker,
2007; Scott et al., 2009). IMPLICATIONS FOR SUBSURFACE
Volumetric models are based on maps of N/G, RESERVOIR MODELS
porosity, and saturation, and we demonstrate spa-
tial variations in N/G and porosity, which are sum- The PGIC allows elucidation of several problems
marized in Figure 17. However, when making res- facing subsurface characterization of sand-injectite
ervoir volumetric estimates, several major challenges reservoirs. Many of these problems are directly as-
exist. Although significant elements of sills are ap- sociated with the laterally and vertically discontin-
proximately conformable to bedding, they step and uous character of sandstone intrusions and the dif-
erode upward (Figure 5Di, ii), features that are ficulty of detecting and mapping reservoir geometry
foreign to depositional sandstones with approxi- using subsurface data. Although data from PGIC
mately stratiform geometries. In the PGIC, even have been used to support sand-injectite reservoir
the most sandstone-rich parts of the sill-dominated models (e.g., Briedis et al., 2007), we are unaware
unit 2 rarely exceed N/G = 0.3 sandstone; how- of examples where outcrop data are used to pop-
ever, the levels of lateral and vertical connectivity ulate subsurface models or where the distribution
and storability are probably far better (Figures 5, of sandstones is recreated using inferences that are
17) than most depositional sandstones with similar based on relationships documented in this study
N/G. The Balder field (Norway) is an example and observed in the PGIC (Smyers and Peterson,
where higher reserves and better-than-expected 1971; Vigorito et al., 2008; Vigorito and Hurst,
hydraulic continuity between thin sandstones are 2010). In light of our outcrop observations, it ap-
encountered (Briedis et al., 2007). pears inevitable that populating subsurface models
Because independent interpretation of sand- with data derived only from the subsurface (e.g.,
stone distribution in Balder field demonstrated sche- Fretwell et al., 2007) will invariably make a sub-
matically how continuity of pore pressure was possi- stantial underestimation of reservoir volume and
ble if sand-injectite architecture is invoked (figure 9 connectivity.
in Hurst et al., 2005). Despite no obvious interwell Our study demonstrates a basis for quantita-
correlation among sandstones, fieldwide pressure tive use of sand-injectite outcrop-analog data in
communication during production prevails in sand- subsurface models. If key characteristics of sand
injectite fields (Briedis et al., 2007; Guargena et al., injectites can be recognized, such as the paleo–sea
2007; Satur and Hurst, 2007). This reflects the high floor at the time of sand injection (unit 4; Figure 8),
level of connectivity between sandstone intrusions the base of unit 2 (base sill zone and lithostatic
even in conditions where the N/G of the reservoir equilibrium surface; Figures 2, 3), and the geom-
interval is very low (<0.01). etry of individual sandstone intrusions, then a res-
As many sandstone intrusions in oil fields are ervoir zonation can be inferred. If the stress regime
neither cut by boreholes nor detected on seismic at the time of sand injection is known, the orien-
data (Briedis et al., 2007; Huuse et al., 2007), the tation of intrusions in units 1 and 3 can be predicted
proportion of reservoir sandstone actually present (Boehm and Moore, 2002). Recognizing the ori-
is likely to be underrepresented in models based entation of intrusions in unit 3 is relevant because
only on subsurface data. Estimating the proportion this is commonly the part of a sand injectite first
of unidentified pay simply because it cannot be re- encountered during drilling.

338 Reservoir Characterization of a Sand-Injectite Complex


An example illustrating the problem of using (Purvis et al., 2002; Briedis et al., 2007; Fretwell
subsurface data as the sole input for reservoir models et al., 2007). Fine-scale reservoir heterogeneities
of sand injectites is the misinterpretation of small- (dimensions, <<0.1 m [0.3 ft]) recognized in PGIC
scale sandstone intrusions. Individually small (sub- similar to those known from the subsurface are typ-
seismic-scale) sandstone intrusions are commonly ically coarsened into grid cells with volumes between
abundant and volumetrically significant parts of 0.6 and 3 × 104 m3 (0.22 and 1.06 × 106 ft3), which
large injectites, for example, the numerous minor is recognized to be a major limitation to the validity
intrusions associated with sills (Figures 5E, 6A), of full-field models (Purvis et al., 2002; Briedis et al.,
that, when observed in subsurface data, could easily 2007; Guargena et al., 2007). The problem of av-
be misinterpreted as a depositional unit (the sill) eraging fine-scale permeability heterogeneity in
with minor sandstone intrusions instead of as a reservoir models is exemplified by the occurrence
small part of a very large sand-injectite complex. of abundant small-scale (<0.1-m [0.33-ft] thick-
The Paleocene Ty Formation in the Sleipner Øst ness) sandstone intrusions that locally enhance the
field (Satur and Hurst, 2007; Figure 6) is an exam- permeability of otherwise low-permeability inter-
ple where dikes that individually constitute tiny vals by many orders of magnitude (Figures 12, 16).
volumes of sandstone but are laterally extensive If, in full-field simulation models, reservoir vol-
enhance vertical sweep efficiency. ume is increased to facilitate vertical connectivity
A common challenge to all subsurface studies (transmissibility), unreasonably high storage will
of sand injectites is that sandstone intrusions and, be generated in the models, and the localized ef-
specifically, dikes crosscut stratigraphy. Any bio- fect of injectites on fluid flow during production
stratigraphic, sequence-stratigraphic, and chrono- will be misrepresented. As well as producing a
stratigraphic boundary can lie above, within, or misleading distribution of hydrocarbons, the full-
below a specific sandstone intrusion (de Boer et al., field models will be compromised or invalidated
2007; Szarawarska et al., 2010). Reservoir correla- with respect to their value for the prediction of
tions based on conventional stratigraphic methods well production rates, production profiles, and
will not capture the geometry of sandstone intru- field lifetime.
sions or place them in an appropriate spatial frame- When sand injectites are part of mudstone-
work (Hurst et al., 2005), and geosteering wells dominated units (unit 3, cf. Figure 7B [low N/G]
using biostratigraphy may consequently be mis- and Figure 8A [very low N/G]), their presence
leading when applied to reservoir correlation. Un- may be rarely detected in boreholes and may be
expectedly good reservoir performance may be elusive on seismic data but may have a deleterious
attributable to the misidentification of thick sand- impact on seal capacity (Hurst et al., 2003a, b).
stone intrusions as depositional sandstone units by Where such units are evaluated as intervals into
creating improved lateral and vertical continuity which gases (for example, CO2) or cuttings may
with other sandstone units. Disappointing reser- be sequestered, a failure to assess the influence of
voir performance and volumetrics may occur where sand injectites will have potentially disastrous ef-
lateral extent and thickness distribution are over- fects if the low N/G of the sandstones is not asso-
estimated. For example, if a well was located at the ciated with a high level of connectivity.
northwestern end of the main sandstone body in
Figure 5E and a depositional origin was inferred,
the reservoir quality and volume will be underesti- CONCLUSIONS
mated. If, however, a well was located toward the
southeast, reservoir volume would be overestimated. Despite approaching two decades of studies on the
Upscaling from a geologic model to a full-field broad significance of sand injectites to hydrocarbon
reservoir-simulation model presents some very spe- exploration and production, before this study,
cific challenges, in particular, preserving high verti- outcrop studies that published porosity and per-
cal permeability in low (<0.01%) N/G stratal units meability data from the architectural elements of

Scott et al. 339


sandstone intrusions did not exist. The key con- Atwater, T., and P. Molnar, 1973, Relative motion of the Pa-
cific and North American plates deduced from sea-floor
clusions of this study are as follows. spreading in the Atlantic, Indian, and South Pacific
oceans, in R. L. Kovach and A. Nur, eds., Proceedings
1. A tripartite organization of parent units (unit 1), of the conference on tectonic problems of the San An-
intrusive units (units 2 and 3), and sandstone dreas fault system: Stanford, California, Stanford Univer-
sity Publications in the Geological Sciences, p. 136–148.
extrusions (unit 4) has been identified and is the Bartow, J. A., 1996, Geological map of the west border of the
basis for defining distinct reservoir characteristics. San Joaquin Valley in the Panoche Creek–Cantua Creek
2. Evaluation of the intrusive complex reveals two area, Fresno and San Benito counties, California: U.S.
Geological Survey Miscellaneous Geologic Investiga-
reservoir units: a lower sill-dominated moderate
tions Map I-2430, scale 1:500,000, 4 sheets.
N/G (0.19) interval (unit 2) and an upper low Bergslien, D., 2002, Balder and Jotun: Two sides of the same
N/G (0.08) interval dominated by high-angle coin? A comparison of two Tertiary oil fields in the Nor-
bifurcating dikes (unit 3). wegian North Sea: Petroleum Geoscience, v. 8, p. 349–
363, doi:10.1144/petgeo.8.4.349.
3. Three distinct classes of sill are identified: stag- Berry, F. A. F., 1973, High fluid potentials in California Coast
gered, stepped, and multilayer geometries. Mud- Ranges and their tectonic significance: AAPG Bulletin,
stone clasts, clay-, and silt-size grains derived from v. 57, p. 1219–1245.
host strata generate lower visual porosity (aver- Boehm, A., and C. J. Moore, 2002, Fluidized sandstone intru-
sions as an indicator of Paleostress orientation, Santa Cruz,
age, 5%) in multilayer sills than in staggered and California: Geofluids, v. 2, p. 147–161, doi:10.1046/j
stepped sills (12–19%). .1468-8123.2002.00026.x.
4. Two dike classes are identified: winglike low- Briedis, N. A., D. Bergslien, A. Hjellbakk, R. E. Hill, and G. J.
Moir, 2007, Recognition criteria, significance to field per-
angle (30–40° relative to bedding) and high-
formance, and reservoir modeling of sand injections in the
angle (80–90°) intrusions. Carbonate cement Balder field, North Sea, in A. Hurst and J. A. Cartwright,
contributes to the occurrence of very low visual eds., Sand injectites: Implications for hydrocarbon ex-
porosity in high-angle dikes (5%) when com- ploration and production: AAPG Memoir 87, p. 91–102.
Cartwright, J. A., 2010, Regionally extensive emplacement
pared to low-angle dikes (11%). of sandstone intrusions: A brief review: Basin Research,
5. This study indicates that sandstone-intrusion v. 22, p. 502–516, doi:10.1111/j.1365-2117.2009
permeability is heterogeneous. Staggered and .00455.x.
stepped sills are seven times more permeable Cartwright, J. A., M. Huuse, and A. Aplin, 2007, Seal bypass
systems: AAPG Bulletin, v. 91, p. 1141–1166, doi:10
(average, 586 and 1225 md, respectively) than .1306/04090705181.
multilayer sills (70 md). Cosgrove, J. W., 2001, Hydraulic fracturing during the for-
6. Volumetric calculations indicate that unit 2 has mation and deformation of a basin: A factor in the dewa-
tering of low-permeability sediments: AAPG Bulletin,
the greatest volume of injected sand and could
v. 85, p. 737–748.
store as much as 6.33 × 107 m3 (2.24 × 109 ft3) de Boer, W., P. B. Rawlinson, and A. Hurst, 2007, Successful
of fluid. Mudstone raft and clast distribution has exploration of a sand-injectite complex: Hamsun prospect,
a major influence on N/G and volumes. Norway Block 24/9, in A. Hurst and J. A. Cartwright, eds.,
Sand injectites: Implications for hydrocarbon exploration
and production: AAPG Memoir 87, p. 65–68.
DeGraff-Surpless, K., S. A. Graham, J. L. Wooden, and M. O.
McWilliams, 2002, Detrital zircon provenance analysis of
REFERENCES CITED the Great Valley Group, California: Evolution of an arc-
forearc system: Geological Society of America Bulletin,
Anderson, R., and R. W. Pack, 1915, Geology and oil re- v. 114, p. 1564–1580, doi:10.1130/0016-7606(2002)
sources of the west border of the San Joaquin Valley 114<1564:DZPAOT>2.0.CO;2.
north of Coalinga, California: U.S. Geological Survey Bul- Dickinson, W. R., 1976, Sedimentary basins developed dur-
letin, v. 603, 220 p. ing evolution of Mesozoic–Cenozoic arc-trench system
Antonellini, M., and A. Aydin, 1994, Effect of faulting on fluid in western North America: Canadian Journal of Earth
flow in porous sandstones: Petrophysical properties: Sciences, v. 13, p. 1268–1287, doi:10.1139/e76-129.
AAPG Bulletin, v. 78, p. 355–377. Dickinson, W. R., and D. R. Seely, 1979, Structure and stra-
Archer, J. B., 1984, Clastic intrusions in deep-sea fan deposits tigraphy of forearc basins: AAPG Bulletin, v. 63, p. 2–31.
of the Rosroe Formation, Lower Ordovician, western Ire- Diggs, T. N., 2007, An outcrop study of clastic-injection struc-
land: Journal of Sedimentary Petrology, v. 54, p. 1197– tures in the Carboniferous Tesnus Formation, Marathon
1205. Basin, Trans-Pecos Texas, in A. Hurst and J. A. Cartwright,

340 Reservoir Characterization of a Sand-Injectite Complex


eds., Sand injectites: Implications for hydrocarbon explora- sandstones and their relationship to sedimentary struc-
tion and production: AAPG Memoir 87, p. 209–219. tures, in L. W. Lake, H. B. Carroll Jr., and T. C. Wesson,
Dixon, R. J., K. Schofield, R. Anderton, A. D. Reynolds, R. W. S. eds., Reservoir characterization: Salt Lake City, Utah,
Alexander, M. C. Williams, and K. G. Davies, 1995, Sand- Academic Press, p. 166–196.
stone diapirism and clastic intrusion in the Tertiary sub- Hurst, A., C. Halvorsen, and E. Siring, 1995, A rationale for
marine fans of the Bruce-Beryl embayment, quadrant 9, routine laboratory probe permeametry: The Log Ana-
UKCS, in A. J. Hartley and D. J. Prosser, eds., Character- lyst, v. 36, no. 5, p. 10–20.
ization of deep-marine clastic systems: Geological So- Hurst, A., J. A. Cartwright, and D. Duranti, 2003a, Fluidiza-
ciety (London) Special Publication 94, p. 77–94. tion structures produced by upward injection of sand
Duranti, D., and A. Hurst, 2004, Fluidization and injection in through a sealing lithology, in P. van Rensbergen, R. R.
the deep-water sandstones of the Eocene Alba Formation Hillis, A. J. Maltman, and C. K. Morley, eds., Subsurface
(U.K. North Sea): Sedimentology, v. 51, p. 503–531, doi:10 sediment mobilization: Geological Society (London) Spe-
.1111/j.1365-3091.2004.00634.x. cial Publication 216, p. 123–137.
Duranti, D., A. Hurst, C. Bell, S. Groves, and R. Hanson, Hurst, A., J. A. Cartwright, M. Huuse, R. Jonk, A. Schwab,
2002, Injected and remobilized Eocene sandstones from D. Duranti, and B. Cronin, 2003b, Significance of large-
the Alba field, UKCS: Core and wireline characteristics: scale sand injectites as long-term fluid conduits: Evi-
Petroleum Geoscience, v. 8, p. 99–107, doi:10.1144 dence from seismic data: Geofluids, v. 3, p. 263–274,
/petgeo.8.2.99. doi:10.1046/j.1468-8123.2003.00066.x.
Fretwell, P. N., W. G. Canning, J. Hegre, R. Labourdette, and Hurst, A., J. A. Cartwright, D. Duranti, M. Huuse, and M.
M. Sweatman, 2007, A new approach to 3-D geological Nelson, 2005, Sand injectites: An emerging global play
modeling of complex sand-injectite reservoirs: The Alba in deep-water clastic environments, in A. Doré and B.
field, United Kingdom, central North Sea, in A. Hurst Vining, eds., Petroleum geology: Northwest Europe and
and J. A. Cartwright, eds., Sand injectites: Implications global perspectives—Proceedings of the 6th Petroleum
for hydrocarbon exploration and production: AAPG Geology Conference: London, United Kingdom, The
Memoir 87, p. 119–127. Geological Society, p. 133–144.
Frey-Martínez, J., J. A. Cartwright, B. Hall, and M. Huuse, Hurst, A., J. A. Cartwright, M. Huuse, and D. Duranti, 2006,
2007, Clastic intrusion at the base of deep-water sands: Extrusive sandstones (extrudites): A new class of strati-
A trap-forming mechanism in the eastern Mediterra- graphic trap?, in M. R. Allen, G. P. Goffey, R. K. Morgan,
nean, in A. Hurst and J. A. Cartwright, eds., Sand injec- and I. M. Walker, eds., The deliberate search for the strat-
tites: Implications for hydrocarbon exploration and pro- igraphic trap: Geological Society (London) Special Pub-
duction: AAPG Memoir 87, p. 49–63. lication 254, p. 289–300.
Guargena, C. G., G. B. Smith, J. Wardell, T. H. Nilsen, and Hurst, A., M. Huuse, J. A. Cartwright, and D. Duranti, 2007,
T. M. Hegre, 2007, Sandstone injections at Jotun oil field, Sand injectites in deep-water clastic reservoirs: Are they
Norwegian North Sea: Modeling their possible effect on there and do they matter? in T. H. Nilsen, R. D. Shew,
hydrocarbon recovery, in A. Hurst and J. A. Cartwright, G. S. Steffans, and J. R. J. Studlick, eds., Atlas of deep-
eds., Sand injectites: Implications for hydrocarbon ex- water outcrops: AAPG Studies in Geology 56, 24 p.
ploration and production: AAPG Memoir 87, p. 81–89. (CD-ROM).
Halvorsen, C., and A. Hurst, 1990, Principles, practice and Hurst, A., A. Scott, and M. Vigorito, 2011, Physical character-
applications of laboratory minipermeametry, in P. F. istics of sand injectites: Earth-Science Reviews, v. 106,
Worthington, ed., Advances in core evaluation, accuracy p. 215–246, doi:10.1016/j.earscirev.2011.02.004.
and precision in reserves estimation: London, United Huuse, M., and M. Mickelson, 2004, Eocene sandstone intru-
Kingdom, Gordon & Breach, p. 521–549. sions in the Tampen Spur area (Norwegian North Sea
Hamberg, L., A. M. Jepsen, N. T. Borch, G. Dam, M. K. quad 34) imaged by 3-D seismic data: Marine and Petro-
Engkilde, and J. B. Svendsen, 2007, Mounded structures leum Geology, v. 21, p. 141–155, doi:10.1016/j.marpetgeo
of injected sandstones in deep-marine Paleocene reservoirs, .2003.11.018.
Cecile field, Denmark, in A. Hurst and J. A. Cartwright, Huuse, M., D. Duranti, N. Steinsland, C. G. Guaranga, P.
eds., Sand injectites: Implications for hydrocarbon ex- Prat, K. Holm, J. A. Cartwright, and A. Hurst, 2004,
ploration and production: AAPG Memoir 87, p. 69–79. Seismic characteristics of large-scale sandstone intrusions
Hurst, A., and A. T. Buller, 1984, Dish structures in some Pa- in the Paleogene of the South Viking Graben, U.K., and
leocene deep-sea sandstones (Norwegian sector, North Norwegian North Sea, in R. J. Davies, J. A. Cartwright,
Sea): Origin of the dish-forming clays and their effect S. A. Stewart, M. Lappin, and J. R. Underhill, eds., Seis-
on reservoir quality: Journal of Sedimentary Petrology, mic technology: Application to the exploration of sedi-
v. 54, p. 1206–1211. mentary basins: Geological Society (London) Memoir 29,
Hurst, A., and J. A. Cartwright, eds., 2007, Implications for p. 263–277.
hydrocarbon exploration and production: AAPG Mem- Huuse, M., J. A. Cartwright, R. Gras, and A. Hurst, 2005,
oir 87, 274 p. Kilometer-scale sandstone intrusions, in A. Doré and
Hurst, A., and D. Goggin, 1995, Probe permeametry: An over- B. Vining, eds., Petroleum geology: Northwest Europe
view and bibliography: AAPG Bulletin, v. 79, p. 463– and global perspectives—Proceedings of the 6th Petro-
473. leum Geology Conference: London, United Kingdom,
Hurst, A., and K. J. Rosvoll, 1991, Permeability variations in The Geological Society, p. 1577–1594.

Scott et al. 341


Huuse, M., J. A. Cartwright, A. Hurst, and N. Steinsland, Macdonald, D., and R. Flecker, 2007, Injected sand sills in a
2007, Seismic characterization of large-scale sandstone strike-slip fault zone: A case study from the Pil’sk Suite
intrusions, in A. Hurst and J. A. Cartwright, eds., Sand (Miocene), southeast Schmidt Peninsula, Sakhalin, in A.
injectites: Implications for hydrocarbon exploration and Hurst and J. A. Cartwright, eds., Sand injectites: Impli-
production: AAPG Memoir 87, p. 21–36. cations for hydrocarbon exploration and production:
Ingersoll, R. V., 1979, Evolution of the Late Cretaceous AAPG Memoir 87, p. 253–263.
forearc basin, northern and central California: Geological Mazzini, A., D. Duranti, R. Jonk, J. Parnell, B. T. Cronin, A.
Society of America Bulletin, v. 90, p. 813–826, doi:10 Hurst, and M. Quine, 2003a, Paleocarbonate seep struc-
.1130/0016-7606(1979)90<813:EOTLCF>2.0.CO;2. tures above an oil reservoir, Gryphon field, Tertiary, North
Ingersoll, R. V., 1983, Petrofacies and provenance of late Me- Sea: Geo-Marine Letters, v. 23, p. 323–339, doi:10.1007
sozoic forearc basin, northern and central California: /s00367-003-0145-y.
AAPG Bulletin, v. 67, p. 1125–1142. Mazzini, A., R. Jonk, D. Duranti, J. Parnell, B. T. Cronin, and
Jackson, C. A.-L., 2007, The geometry, distribution and de- A. Hurst, 2003b, Fluid escape from reservoirs: Implica-
velopment of clastic injectites in deep-marine deposi- tions from cold seeps, fractures and injected sands. Part
tional systems: Examples from the Late Cretaceous 1: The fluid-flow system: Journal of Geochemical Ex-
Kyrre Formation, Måløy slope, Norwegian margin, in A. ploration, v. 78–79, p. 293–296, doi:10.1016/S0375
Hurst and J. A. Cartwright, eds., Sand injectites: Impli- -6742(03)00046-3.
cations for hydrocarbon exploration and production: McGuire, D. J., 1988, Stratigraphy, depositional history, and hy-
AAPG Memoir 87, p. 37–48. drocarbon source rock potential of the Upper Cretaceous–
Jenkins, O. P., 1930, Sandstone dikes as conduits for oil migra- lower Tertiary Moreno Formation, central San Joaquin Ba-
tion through shales: AAPG Bulletin, v. 14, p. 411–421. sin, California: Ph.D. thesis, Stanford University, Stanford,
Jolly, R. J. H., and L. Lonergan, 2002, Mechanisms and con- California, 231 p.
trol on the formation of sand intrusions: Journal of the McPherson, B. J. O. L., and G. Garven, 1999, Hydrody-
Geological Society (London), v. 159, p. 605–617, doi:10 namics and overpressure mechanisms in the Sacramento
.1144/0016-764902-025. Basin, California: American Journal of Science, v. 299,
Jolly, R. J. H., J. W. Cosgrove, and D. N. Dewhurst, 1998, p. 429–466, doi:10.2475/ajs.299.6.429.
Thickness and spatial distribution of clastic dikes, north- Minisini, D., and H. Schwartz, 2006, An early Paleocene
west Sacramento Valley, California: Journal of Structur- cold-seep system in the Panoche and Tumey hills, cen-
al Geology, v. 20, p. 1663–1672, doi:10.1016/S0191 tral California, U.S.A., in A. Hurst and J. A. Cartwright,
-8141(98)00053-4. eds., Sand injectites: Implications for hydrocarbon ex-
Jonk, R., 2010, Sand-rich injectites in the context of short- ploration and production: AAPG Memoir 87, p. 185–
lived and long-lived fluid flow: Basin Research, v. 22, 197.
p. 603–621, doi:10.1111/j.1365-2117.2010.00471.x. Mitchell, C., S. A. Graham, and D. H. Suek, 2010, Subduc-
Jonk, R., D. Duranti, J. Parnell, A. Hurst, and A. E. Fallick, tion complex uplift and exhumation and its influence on
2003, The structural and diagenetic evolution of injected Maastrichtian forearc stratigraphy in the Great Valley
sandstones: Examples from the Kimmeridgian of NE Basin, northern San Joaquin Valley, California: Geologi-
Scotland: Journal of the Geological Society (London), cal Society of America Bulletin, v. 122, p. 2063–2078,
v. 160, p. 881–894, doi:10.1144/0016-764902-091. doi:10.1130/B30180.1.
Jonk, R., A. Hurst, D. Duranti, A. Mazzini, A. E. Fallick, and Nilsen, T. H., and T. W. Dibblee, 1981, Geology of the cen-
J. Parnell, 2005a, The origin and timing of sand injec- tral and northern Diablo Range, California: Introduc-
tion, petroleum migration and diagenesis: The Tertiary tion, in V. Frizzell, ed., Geology of the central and north-
petroleum system of the South Viking Graben, North ern Diablo Range, California: Pacific Section, SEPM Field
Sea: AAPG Bulletin, v. 89, p. 329–357, doi:10.1306 Trip Guidebook 19, p. 1–4.
/10260404020. Purvis, K., J. Kao, K. Flanagan, J. Henderson, and D. Duranti,
Jonk, R., J. Parnell, and A. Hurst, 2005b, Aqueous and pe- 2002, Complex reservoir geometries in a deep-water
troleum fluid flow associated with sand injectites: Basin clastic sequence, Gryphon field, UKCS: Injection struc-
Research, v. 17, p. 241–257, doi:10.1111/j.1365-2117 tures, geological modeling and reservoir simulation: Ma-
.2005.00262.x. rine and Petroleum Geology, v. 19, p. 161–179, doi:10
Kamakami, G., and M. Kawamura, 2002, Sediment flow and .1016/S0264-8172(02)00003-X.
deformation (SFD) layers: Evidence for intrastratal flow Satur, N., and A. Hurst, 2007, Sand-injection structures in
in laminated muddy sediments of the Triassic Osawa deep-water sandstones from the Ty Formation (Paleo-
Formation, northeast Japan: Journal of Sedimentary Re- cene), Sleipner Øst field, Norwegian North Sea, in A.
search, v. 72, p. 171–181, doi:10.1306/041601720171. Hurst and J. A. Cartwright, eds., Sand injectites: Impli-
Lonergan, J., C. Borlandelli, A. Taylor, M. Quine, and K. cations for hydrocarbon exploration and production:
Flanagan, 2007, The three-dimensional geometry of sand- AAPG Memoir 87, p. 113–117.
stone injection complexes in the Gryphon field, United Schwartz, H., J. Sample, K. D. Weberling, D. Minisini, and
Kingdom, North Sea, in A. Hurst and J. A. Cartwright, J. C. Moore, 2003, An ancient linked fluid-migration sys-
eds., Sand injectites: Implications for hydrocarbon ex- tem: Cold-seep deposits and sandstone intrusions in the
ploration and production: AAPG Memoir 87, p. 103– Panoche Hills, California, U.S.A.: Geo-Marine Letters,
112. v. 23, p. 340–350, doi:10.1007/s00367-003-0142-1.

342 Reservoir Characterization of a Sand-Injectite Complex


Scott, A. S. J., M. Vigorito, and A. Hurst, 2009, The pro- Alexander Island: British Antarctic Survey Bulletin, v. 51,
cess of sand injection: Internal structures and relation- p. 1–42.
ships with host strata (Yellowbank Creek injectite com- Thompson, B. J., R. E. Garrison, and C. J. Moore, 1999, A late
plex, California, U.S.A.): Journal of Sedimentary Cenozoic sandstone intrusion west of S. Cruz, California:
Research, v. 79, no. 8, p. 1–18, doi:10.2110/jsr.2009 Fluidized flow of water and hydrocarbon-saturated sedi-
.062. ments, in R. E. Garrison, I. W. Aiello, and J. C. Moore,
Smyers, N. B., and G. L. Peterson, 1971, Sandstone dikes and eds., Late Cenozoic fluid seeps and tectonics along the
sills in the Moreno Shale, Panoche Hills, California: Geolog- San Gregorio fault zone in the Monterey Bay region, Cal-
ical Society of America Bulletin, v. 82, p. 3201–3207, ifornia: AAPG Pacific Section, Volume and Guidebook
doi:10.1130/0016-7606(1971)82[3201:SDASIT]2.0 76, p. 53–74.
.CO;2. Thompson, B. J., R. E. Garrison, and C. J. Moore, 2007, A
Sutherland, W. J., C. Halvorsen, A. Hurst, C. A. Mcphee, G. reservoir-scale Miocene injectite near Santa Cruz, Cali-
Robertson, P. R. Whattler, and P. F. Worthington, 1993, fornia, in A. Hurst and J. A. Cartwright, eds., Sand injec-
Recommended practice for probe permeametry: Marine tites: Implications for hydrocarbon exploration and pro-
and Petroleum Geology, v. 10, p. 309–318, doi:10.1016 duction: AAPG Memoir 87, p. 151–162.
/0264-8172(93)90075-4. Vigorito, M., and A. Hurst, 2010, Regional sand-injectite ar-
Svendsen, J. B., H. J. Hansen, T. Stærmose, and M. K. Engkilde, chitecture as a record of pore-pressure evolution and
2010, Sand remobilization and injection above an active sand redistribution in the shallow crust: Insights from
salt diapir: The Tyr sand of the Nini field, eastern North the Panoche giant injection complex, California: Journal
Sea: Basin Research, v. 22, p. 548–561, doi:10.1111/j of the Geological Society (London), v. 167, p. 889–904,
.1365-2117.2010.00480.x. doi:10.1144/0016-76492010-004.
Szarawarska, E., M. Huuse, A. Hurst, W. de Boer, L. Liwei, S. Vigorito, M., A. Hurst, J. Cartwright, and A. Scott, 2008,
Molyneux, and P. Rawlinson, 2010, Three-dimensional Regional-scale shallow crustal remobilization: Processes
seismic characterization of large-scale sandstone intru- and architecture: Journal of the Geological Society (Lon-
sions in the lower Paleogene of the North Sea: Completely don), v. 165, p. 609–612, doi:10.1144/0016-76492007
injected versus in-situ remobilized saucer-shaped sand -096.
bodies: Basin Research, v. 22, p. 517–532, doi:10.1111 Wild, J., and N. Briedis, 2010, Structural and stratigraphic
/j.1365-2117.2010.00469.x. relationships of the Paleocene mounds of the Utsira high:
Taylor, B. J., 1982, Sedimentary dikes, pipes and related Basin Research, v. 22, p. 533–547, doi:10.1111/j.1365
structures in the Mesozoic sediments of southeastern -2117.2010.00479.x.

Scott et al. 343

You might also like