<<

Reconstructing the Upper Permian sedimentary facies distribution of a tight gas fi eld in Central Europe on the basis of a modern analog fi eld study in the Panamint , western U.S.

Anna Alexandra Vackiner1,*,†, Philipp Antrett1,†, Frank Strozyk1, Harald Stollhofen2, Stefan Back1, and Peter Kukla1 1Geological Institute, Energy and Mineral Resources Group, RWTH Aachen University, Wüllnerstaße 2, 52062 Aachen, Germany 2North Bavarian Center of Earth Sciences, 24 FAU Erlangen-Nürnberg University, Schlossgarten 5, 91054 Erlangen, Germany

ABSTRACT to (2) distributary fl uvial deposits sic, and Cretaceous (Lohr et al., 2007). Much of toward the basin center and (3) ephem- the original Permian structural and stratigraphic Comparison of modern deposits in the eral lake deposits in the deepest basin area. grain, including the location of Permian depo- Panamint Valley, western , to (4) Eolian accumulation and preserva- centers and associated eolian reservoirs, was core and geophysical data from a Permian tion is mainly concentrated on hanging-wall rearranged. Therefore, most Permian fault-con- (Rotliegend, Germany) tight gas fi eld allows locations. However, additional dune deposits trolled paleohighs do not match the present-day for improved understanding of the inter- are proposed above overlapping step faults structural highs (Vackiner et al., 2011); however, action of tectonics and sedimentary processes and on footwalls of synsedimentary active their identifi cation is crucial for the understand- during Rotliegend . The Panamint faults. (5) Sandfl ats occur on the upwind and ing of Upper Rotliegend II facies and Valley was selected for a modern analog of downwind margins of the dune fi eld. These accommodation space generation. Well infor- the subsurface Rotliegend Basin because predictions are calibrated to core and - mation in the study area is only available from both study sites are characterized by (1) elon- physical well log data. present-day structural highs below the Zech- gated grabens with large-scale bounding stein salt, while present-day structural lows are fault zones resulting from synsedimentary INTRODUCTION commonly undrilled. Subsurface information of transtensional tectonics; (2) fault-controlled these sites is thus often limited to seismic data. paleotopography as key controlling param- Studying the sedimentary and tectonic com- The few cored wells alone do not allow a satis- eter for the facies distribution, plexity of Permian (Rotliegend ) tight factory regional interpretation and/or extrapola- including alluvial fans, dunes, wet and damp gas fi elds in Central Europe requires integrated tion of the sedimentary facies. interdune sandfl ats, and ephemeral approaches from fi eld-based analog studies , Because of the limitations of geophysical deposits; and (3) local sediment provenance laboratory analysis, seismic and well data inter- data, the modern analog study presented herein from sedimentary and volcanic rocks. The pretation, and structural modeling. One of these took place in the northern part of Panamint analysis of satellite images and fi eld data tight gas fi elds in the focus of recent research Valley (Lake Basin) in Inyo County, east- from the Panamint Valley enabled the devel- is located in northwestern Germany, east of the ern , United States. It is located in a opment of a conceptual model involving Dutch Groningen gas fi eld, on the eastern foot- zone of active transtensive deformation along topography, synsedimentary faulting, and wall of the Ems Graben at ~4200 m depth (Figs. the North American plate boundary (Stewart, activity as controlling factors for the 1 and 2). The tight character of the reservoir 1988), an area of complex intracontinental sediment facies distribution. The application is attributed to extensive formation of quartz deformation (Lee et al., 2009). The north- of the model to the reconstructed Rotliegend overgrowth, pressure solution, and authigenic south–trending Panamint Valley represents one paleotopography of the German subsurface fi brous illite (Gaupp and Solms, 2005). The of the active grabens of the Basin and Range study site allows for prediction of the facies reservoir rocks, which are formed by hetero- Province (Smith, 1976; Fig. 3). The regional distribution prior to the Triassic–Cretaceous geneous fl uvio-eolian facies, were deposited tectonic setting strongly infl uences the hydro- tectonic overprinting. As a consequence, we at the southwestern margin of the Southern thermal and structural characteristics of the expect a sediment facies succession from Permian Basin (Fig. 1; Vackiner et al., 2011). region (Jayko et al., 2008). Large alluvial fans (1) alluvial fan deposits along the hang- These rocks are overlying the Carboniferous developed simultaneously with modern tec- ing walls of the basin-bounding fault zones basement and patchy andesitic to basaltic Rot- tonics. The fans source a shallow ephemeral dry liegend volcanic rocks. lake in the central basin. In the North Panamint The prediction of location and distribution of Valley, an active dune fi eld is located on allu- *Corresponding author: [email protected] -aachen.de sandstone reservoirs in the fl uvio-eolian deposits vial material between the dry lake and the cur- †Present address: Wintershall Holding GmbH, is challenging, particularly due to the multiphase rently active fan. The study site further includes Friedrich-Ebert-Straße 160, 34119 Kassel, Germany tectonic overprinting during the Triassic, Juras- volcanic rocks at the base of the sedimentary

Geosphere; October 2012; v. 8; no. 5; p. 1129–1145; doi:10.1130/GES00726.1; 16 fi gures; 1 supplemental fi le. Received 27 May 2011 ♦ Revision received 3 May 2012 ♦ Accepted 29 May 2012 ♦ Published online 18 September 2012

For permission to copy, contact [email protected] 1129 © 2012 Geological Society of America

Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/8/5/1129/3342812/1129.pdf by guest on 30 September 2021 Vackiner et al.

E 0° E 10° N 55° Mid North High Ringk Øbing Fyn High

N 55° SPBSPB

ELEL London Brabant Mass GHGH

E 0° 200 km

Lambert Conformal if Conic Projection E 10°

Alluvial fan conglomerates and fluvial depositional environments Faults and structural lineaments

Aeolian sandstone depositional environments Location of study area

Sandflat to depositional environments

Basin centred dry lake deposit

Figure 1. Map outlining maximum extent of the depositional area of the Southern Permian Basin (SPB) during the of the Permian late Upper Rotliegend II (modifi ed from Ziegler, 1982; Legler, 2005). The star shows the location of the German subsurface study site (displayed in detail in Fig. 3). NL— Low; GH—Groningen High; EL–Ems Low.

succession. Consequently, the Panamint Valley described in two separate subsections. The fi rst the central part of the Southern Permian Basin provides a sedimentary facies, tectonic setting, is a discussion of the subsurface study site, Ems (Gast and Gaupp, 1991). and basement type similar to those of the Rot- Graben in northwestern Germany. The second The Ems Graben in the central Southern liegend subsurface study site. A particular focus subsection focuses on the modern analog study Permian Basin underwent sedimentary transten- of this analog study was to delineate the pos- site in the Panamint Valley, western U.S. sional tectonics during deposition of the Upper sible topographic control on the sedimentary Rotliegend II, while subsequent phases of tec- facies distribution, and to study the reactive vol- Subsurface Study Site, Ems Graben, tonic activity, e.g., rifting in the North Sea dur- canic infl uence on the sedimentary system in the Northwestern Germany ing earliest Triassic until Late Jurassic to Early German study site. Even though the Panamint Cretaceous time (Ziegler, 1990), overprinted Valley’s present-day confi guration represents The subsurface study area is located at the the Rotliegend structural highs (Vackiner et al., only a snapshot in its development history, it boundary of the Ems Graben at the southwest- 2011). The reconstructed graben structure in the provides valuable high-resolution insights into ern margin of the Southern Permian Basin, study area is characterized by two bounding, the facies distributions, the facies architecture in and is characterized by a U-shaped, mostly north-south–trending strike-slip to normal fault a regional context, and the interaction of con- north-south–trending Zechstein salt wall, zones with offsets of as much as 250 m in the temporaneous sedimentary systems, unavailable situated above an asymmetrical Late Permian west (Fig. 2, FZ-1) and as much as 150 m in from geophysical subsurface data and punctual graben (Vackiner et al., 2011). During deposi- the east (Fig. 2, FZ-2). To the north, the eastern information of wells and cores. tion of the Rotliegend, the Southern Permian fault zone ceases, and the asymmetrical gra- Basin represented an intracontinental basin ben changes into a half-graben (Vackiner et al., GEOLOGICAL SETTING of ~1700 km length and 300–600 km width, 2011; Fig. 2). A third fault zone in the graben extending from the eastern UK to and center exhibits an Upper Rotliegend II fault- In the following, the geological settings the Czech Republic (Plein, 1993; McCann, controlled paleorelief of ~100–150 m. of the study areas in northwestern Germany 1998; Fig. 1). During deposition of the Upper The deepest part of the graben most likely and in the Panamint Valley, western U.S., are Rotliegend II, a perennial saline lake occupied represents an area where ephemeral dry lakes

1130 Geosphere, October 2012

Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/8/5/1129/3342812/1129.pdf by guest on 30 September 2021 Panamint Valley—Field analog for a European Permian tight gas fi eld

Topography term fault activity, the northern Panamint Valley Well locations FZFZ 2 x 300 is bordered by as high as 1800 m to Overlying salt dome the east along the Panamint Valley fault zone and decreases seismic 250 resolution by as high as 800 m to the west along the 11 FZ-1-4FZ-1-4 Ash Hill fault. The elevation difference between 1 Fault zones 200 Faults with the ephemeral dry lake, located at ~400 m above indicated dip direction 150 mean sea level, and the peaks is as 10 much as 1350 m. FZ-3FZ-3 100 The inner part of the Panamint Valley is a 12 relatively shallow with an estimated 50 maximum sedimentary basin fi ll of <1000 m at 13 6 its northern end (Blakely et al., 1999; Blakely 3/3a FZ-1FZ-1 4 0 and Ponce, 2001). The Panamint Valley is one 2 5 of a continuous chain of basins linked by the Owens system during glacial and pluvial 7 periods (Jayko et al., 2008). 8 The sedimentary facies of the Panamint Val- ley comprise different types of alluvial fans, a mud-dominated ephemeral dry lake surface, and eolian dune, interdune, and sandfl at deposits (Fig. 4; Supplemental File1). The ephemeral dry lake located in the center of the valley is bounded by alluvial fans descending from the mountain 9 ranges in the north, east, and west. Toward the 2 km FZ-FZ- 2 north, the dry lake are overlain by a FZ-4FZ-4 transition of sandfl at and eolian dune deposits that partially cover the northern alluvial fans at Figure 2. Late Upper Rotliegend II paleotopography of the subsurface study area in north- ~700–830 m elevation. The present-day loca- western Germany. Synsedimentary active Upper Rotliegend II faults are plotted in black. tion of the dunes is due to a northward upfan Plot is based on isopach maps (Vackiner et al., 2011). Drilling rig symbols depict wells. migration from an original position closer to the dry lake at an average rate of 0.8 m/yr (Prestud Anderson and Anderson, 1990). The eastern and occurred during the Upper Rotliegend II depo- Glennie, 1990a; Rieke et al., 2001). In contrast, western alluvial fans interfi nger with patchy sition (Vackiner et al., 2011). This assumption the source of major fl uvial sediment input was andesitic to basaltic volcanic rocks. is supported by the identifi cation of a polygonal located in the Variscan hinterland toward the in three-dimensional (3D) seismic data, south (Glennie, 1990b; Plein, 1993). The preser- DATA AND METHODS interpreted to represent faults and that vation of eolian dunes was governed by tectonic had their origin in giant desiccation polygons on subsidence (Kocurek, 2003). The available data from the subsurface ephemeral dry lakes (Antrett et al., 2012). study area in northwestern Germany comprise The main tight gas reservoir interval is Modern Analog Study Site, Panamint 3D seismic refl ection data in time and depth, 80–170 m thick and unconformably overlies Valley, Western United States wireline log data, and core material. Key strati- the top of the Upper Carboniferous. It con- graphic horizons were interpreted on pre- sists of fl uvio-eolian sediments deposited ca. The Panamint Valley is located in the cen- depth-migrated 3D seismic data, covering an 259–260 Ma. The Ameland lake-level highstand tral part of the Basin and Range Province and area of 293 km2 (Vackiner et al., 2011). At the (ca. 260 Ma; Gast, 1991; Legler and Schneider, developed as isolated basin in a pull-apart sys- target interval, the seismic data set has an aver- 2008) coincided with the onset of sedimentation tem (Burchfi el et al., 1987). The general struc- age vertical resolution of ~40 m. In addition, in the study area. Following deposition of the res- ture of the basins is a rhomb-shaped graben or information from 14 wells, including digital ervoir interval, the study area underwent several half-graben bordered by strike-slip faults (Aydin wireline logs from 7 wells and core data from lake-level fl uctuations (e.g., Legler and Schnei- and Nur, 1985; Price and Cosgrove, 1990). The 4 wells, was used for the sedimentary facies der, 2008). Core data of the Upper Rotliegend Panamint Valley is bounded by three major fault reconstruction and for stratigraphic correlation. II sediments in the study area revealed that these systems that are still active (Figs. 3 and 4): the Dipmeter logs of three wells (wells 2, 3, and 3a) are of fl uvio-eolian origin, including braided north-south–trending, slightly oblique right- and one formation microimaging/scanning log stream, sheetfl ood, eolian dune, and wet to dry lateral strike-slip–dominated Ash Hill fault in (FMI/FMS at well 3a) were used for analyzing interdune deposits of the Wustrow and Bahnsen the west (Densmore and Anderson, 1997), the Members (ca. 260–259 Ma, early Wuchiapin- low-angle north-northwest–south-southeast– 1Supplemental File. Sedimentary facies distribu- gium; Vack iner et al., 2011). The majority of the trending normal dip-slip detachment Panamint tion in the Panamint Valley. The .kmz fi le can be eolian sediment was supplied from local sedi- Valley fault zone in the east (Burchfi el et al., viewed using Google EarthTM. If you are viewing the PDF of this paper or reading it offl ine, please visit ment sources, such as Carboniferous highs and 1987), and the northwest-striking dextral Hunter http://dx.doi.org/10.1130/GES00726.S1 or the full- patchy andesitic to basaltic volcanic rocks, via Mountain fault zone in the north (Smith, 1976; text article on www.gsapubs.org to view the Supple- eastern trade (Gast, 1988; McCann, 1998; Blakely and Ponce, 2001). In response to long- mental File.

Geosphere, October 2012 1131

Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/8/5/1129/3342812/1129.pdf by guest on 30 September 2021 Vackiner et al.

tal geological maps (Jennings, 1975; Jennings et al., 2010) (Fig. 4). Based on fi eld work, the EurekaValley 20 km following sedimentary facies were identifi ed N37°12′ and mapped: alluvial fans of different angles N and braided stream systems, eolian sediment bodies such as dunes, interdune, and sandfl at deposits, and mudfl at sediments of the ephem- eral dry lake. Fault interpretations were adapted from Jennings (1975) and Jennings et al. (2010) and compared to the measurements of high- resolution Lidar data, our own fi eld observations (Fig. 5), and satellite image interpretations. The composition of sediment samples from dune ′ N36°48 Furnace Creek Fault Zone and clays of the dry lake surface was ana- Inyo Mountains lyzed by thin-section analyses and X-ray dif- fractometry (XRD).

Hunter Mountain CottonwoodFault Mountains RESULTS

We discuss here the results of the analysis of the subsurface study site in northwestern Ger- Owens Valley Tuc many and summarize the results of the fi eld ki Mountain work in the Panamint Valley. N36°24′ Subsurface Study Site Germany Darwin Death

Panamint V The 3D seismic interpretation of the sub- Valley surface study site of the Ems Graben, northern Germany, reveals considerable differences in alley sediment thicknesses between hanging-wall and Pa namint Mountains footwall settings interpreted to record tectonic Deat activity of the graben during the deposition of h V the Upper Rotliegend II studied interval (Fig. 2). a lley We observed step faults, pull-apart structures, N36° Fault Zone and relay ramps that induced a complex fault- sediment interaction. The lower parts of core data from wells 2 and 3 show coarse-grained, grain-supported structureless conglomerates with polymict com- ponents of various sizes, interbedded with fi ne- to medium-grained, small-scale cross-bedded (Fig. 6). This facies is interpreted to represent deposits from braid-dominated allu- vial fans, with conglomerates being deposited in deposits while sandstones accu- Garlock Fault mulated as intrachannel fi lls. On top of these fans, sediments show no internal geometry W117°30′ W117° and consist of breccias to conglomerates with Figure 3. Regional tectonic setting of the Basin and Range Province, eastern California, angular to rounded grains of as much as 5 cm in western United States. Faults are plotted in red. Faults and dipping directions of faults are diameter. These deposits are interpreted to have based on fi eld observations, Jennings et al. (2010), and Jennings (1975). (Satellite image— originated from hyperconcentrated gravitational 2010 Google and DigitalGlobe.) mass fl ows (Fig. 6A). Frequently interbedded sandstones with bimodal fi ne and medium grain sizes and cross-stratifi cation are either remains structural dips and dip directions of foresets in tion from satellite images (Google Earth) (Fig. of eolian dune deposits or fl uvially reworked the eolian successions. Thin sections from core 4), light detection and ranging (Lidar) data pro- eolian units. material were analyzed for the mineral content vided by GeoEarthScope (Southern and East- On top of the alluvial fan intervals, the sedi- and cement mineralogy. ern California project, SoCal_Panamint target, mentary facies changes toward a section domi- The large-scale tectonic setting of the Pana- http://opentopo.sdsc.edu/gridsphere/gridsphere nated by eolian deposition (in cores of wells 2, mint Valley was studied by combining informa- ?cid=datasets), and U.S. Geological Survey digi- 3, and 3a) with dry (dry sandfl at, Fig. 7; eolian

1132 Geosphere, October 2012

Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/8/5/1129/3342812/1129.pdf by guest on 30 September 2021 Panamint Valley—Field analog for a European Permian tight gas fi eld

lt Zone 5 km N

Panamint Valley Fau ′

W117°22 shown. eld observation points are

rlay and fi Ash Hill Fault Hill Ash

′ Jennings et al. (2010) and based on light detec-

26

117°

W

Hunter Mountain Fault

N36°30

′ ′ ′

2

36°2 36°18

N36°26

N N N36°30 Faults Photo location of Fig. 5 A–C Location depicted on Fig. 5 D–E Photo Location of Figs. 7 and 8 Photo Locations of Fig. 6 Photo Location of Fig. 9 Photo Location of Fig. 13 GPS points taken at photo, measurement or sample location Braided-stream-dominated alluvium High angle debris-flow-dominated alluvial fans Lower angle alluvial fans, mass-flows inactive, less deep channels Inactive alluvium, deep fluvial channels interbedded with ancient lake level highstands Mudflat, dry lake fine grained aeolian / Very Mudflat - sandflat, intercalated braided steam deposits Sandflat with wind ripples Dunes Fault photo locations: Sedimentary facies photo locations: Projection: WGS 84 Satellite images © 2010 Google and DigitalGlobe LIDAR © EarthScope 5 km N

Panamint Valley Fault Zone ′

117°22

W Ash Hill Fault Hill Ash ′

17°26

W1

Hunter Mountain Fault

′ ′ ′

2

36°2 36°18

N36°26

N N N36°30 Stars indicate locations of photos in Figures 4–13. Right: Sedimentary facies distribution in the Panamint Valley. 4–13. Right: Sedimentary facies distribution in the Panamint Stars indicate locations of photos in Figures Figure 4. Overview of the Panamint Valley area. Faults are plotted in red. Left: Fault map of the Panamint Valley; faults after Valley; Left: Fault map of the Panamint plotted in red. Faults are area. Valley 4. Overview of the Panamint Figure ove (satellite image—2010 Google and DigitalGlobe). Lidar tion and ranging (Lidar) data (Earthscope, http://www.earthscope.org)

Geosphere, October 2012 1133

Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/8/5/1129/3342812/1129.pdf by guest on 30 September 2021 Vackiner et al.

BC A N N

C E M E N T E D F A U L T S C A R P

D E B A S A L T I C L A V A F L O W D E P O S I T N N

III II IV

L AN D S L I D E

I N C I S E D F L U V I A L C H A N N E L I

B R A I D E D S T R E A M D O M I N A T E D A L L U V I A L F A N

Figure 5. Photos of outcrops of faults of the Panamint Valley fault zone. (The photo locations are in Fig. 4.) (A–C) Normal fault of the east- ern fault zone; coordinates are WGS 84, UTM (World Geodetic System, Universal Transverse Mercator) zone 11S, 467388m east 4028610m north, 1479 m above mean sea level, average strike/dip: 150°/65°. Fault scarp with cementation along offset is depicted in A. In B, the fault scarp is highlighted in red. C is a close-up of A and B showing cementation along the normal fault. Fault striations and cementation already grinded by fault movement are visible. (D) View on fault zone on the northeastern part of Panamint Valley; picture taken from UTM zone 11S, 462787m east 4034029m north, 577 m above mean sea level. (E) Synsedimentary normal faults are highlighted in red. I—alluvial fan with deep incised channel cut by a normal fault, fault location UTM zone 11S, 463095m east 4033894m north, 578 m above mean sea level; II, III—several fault scarps belonging to the same fault zone; IV—fault scarp with on the hanging wall.

Figure 6. Comparison of alluvial fan deposits observed from core material from the tight gas fi eld in northwestern Germany to field observations in Panamint Valley; photo locations are in Figure 4. (A) Sheetfl ood and braid-dominated alluvial fans with gravel bar and cross-stratified channel deposits. The core depicted in the middle might represent fi rst dune succes- sions. (B, C) View on main fan-head channel passing into a distributive network of channels and distribu- tive network of -fi lled shallow channels sourced from alluvial fans and fi nally passing into the dry lake.

1134 Geosphere, October 2012

Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/8/5/1129/3342812/1129.pdf by guest on 30 September 2021 Panamint Valley—Field analog for a European Permian tight gas fi eld

Figure 7. Comparison of close lateral changes in eolian sedi- ment from core data of the sub- surface study area in Germany to the fi eld analog in Panamint Valley. Photo location depicted in Figure 4. 1—damp sandfl at deposit, 2—eolian dune base deposit, 3—pond deposit, 4— lake or pond margin deposit; 1A–4A show deposits in core material of the subsurface area in northwestern German, and 1B–4B are the Panamint Valley correlatives.

dune base and eolian dune, Fig. 7) to wet (inter- paleofootwall positions (FZ-1) may be related and overlain by the next dune, a defl ation lag, dune mudfl at, Figs. 7, 8, and 9; interdune pond to subtle paleorelief, the formation of sub- or a wet deposit. or lake, Fig. 10; fl uvial and pond or lake margin, seismic-scale footwall collapse compartments, Interdune units exhibit a variety of facies Fig. 8) sedimentary successions. The fl uvio- or increased moisture content at the sediment types, including damp to wet sandfl at, mudfl at, eolian–dominated intervals reach thicknesses surface (Vackiner et al., 2011). We estimate the lake margin, and pond deposits (Figs. 8 and 10). of as much as 150 m in wells at the footwall of maximum initial dune heights to ±20 m, assum- Eolian mudfl ats (Fig. 10) are mainly composed FZ-2, and only 50 m at the footwall of FZ-1. ing that (1) ~60% of the original dune heights of (>50%; Amthor and Okkerman, 1998) The evaluation of dipmeter logs showed that the were eroded during sedimentation (Allen and with lensoid to aligned concentrations of - footwall of FZ-1 is mainly composed of low- Allen, 1990) and (2) a depth compaction coeffi - stone and very fi ne to fi ne-grained sandstones. dipping sandfl at deposits and wet deposits pro- cient of 0.27 km–1 for fi ne- to medium-grained They are characterized by convolute bedding ducing an irregular log signature. FMI/FMS and sandstones can be applied (Sclater and Christie, and ball-and-pillow structures, and are the most dipmeter log analysis in sandstone–dominated 1980). Typically, dip angles increase gradually abundant interdune deposits. Their depositional successions of the FZ-2 footwall (Fig. 11) indi- from values of <5° (dry sandfl at) over 5°–15° environment is mainly shallow subaquatic with cate upward-increasing dip angles (blue pattern) (dune base) to 15°–35° (dune). Dune tops are blown-in eolian sands (George and Berry, 1993). in dune-dominated areas with prevailing west- characterized by erosional truncation (Fig. 9) The pond or lake interdune deposits consist of ward dips, partly superimposed by an irregular pattern of damp sandfl at deposits. Measured west to west-southwest dip directions are in line with the prevailing wind direction from the east to east-northeast described previously (e.g., Gast, 1988; Rieke et al., 2001). Eolian strata of dune origin form the princi- pal reservoir rocks. The interpretation of core material reveals that individual dune sets show maximum preserved thicknesses of 3 m and a pervasive cross-bedding with a considerable spread in paleotransport indicators. Therefore, the dune bodies are interpreted to represent iso- lated barchans or barchanoid dunes with amal- gamated dune ridges (aklé dunes; Fig. 9). Initial dune set thicknesses are diffi cult to determine. Prevailing barchanoid dune forms suggest lim- ited sediment supply and/or restricted avail- ability of accommodation, which prevented the Figure 8. Comparison of wet sandfl at and distal fl uvial deposits observed from core mate- evolution of larger, more stable dune forms such rial from the tight gas fi eld in northwestern Germany to fi eld observations in Panamint as transverse dunes (Mountney, 2006), or that Valley; photo location depicted in Figure 4. (A) Wet sandfl at deposit with intercalated clay the eolian system built above preservation space in the core material of well 3. (B) Wet sandfl at deposit as an interdune environment in the and had little potential for being incorporated Panamint Valley. (C) Fluvial sand-fi lled channel with clay rip-up clasts in the core material into the rock record (Kocurek and Havholm, of well 2. (D) Fluvial sand-fi lled channel in the Panamint Valley. (E) Close up (area in photo 1993). The preservation of dune deposits in of D) of cross-stratifi cation in the channel , microconglomeratic to sandy sediment.

Geosphere, October 2012 1135

Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/8/5/1129/3342812/1129.pdf by guest on 30 September 2021 Vackiner et al.

Figure 9. Comparison of eolian dune deposits observed from core material from the tight gas fi eld in northwest- ern Germany to fi eld observations in Panamint Valley; photo location depicted in Figure 4. (A) Panamint dunes located on incised channels of older alluvium. (B) Barchanoid dune deposits of core 3 with chang- ing dipping angles between the cross- stratifi cations. (C) Barchanoid to aklé dune deposits of the Panamint Valley. (D) Stacked eolian dune set of core 3. The fi rst eolian dune set with dip angles of ~20° is cut by erosional trun- cation and overlain by a lag deposit and a new dune set with slightly upward-increasing dip angles above. (E, F) Panamint Valley eolian dunes after wet weather. Eolian stacking is visible in active dune set.

1136 Geosphere, October 2012

Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/8/5/1129/3342812/1129.pdf by guest on 30 September 2021 Panamint Valley—Field analog for a European Permian tight gas fi eld

True Dip

Figure 10. Comparison of interdune mudfl at and dry lake mudfl at deposits observed from core material from the tight gas fi eld in northwestern Germany to fi eld observations in Panamint Valley; photo location depicted in Figure 4. (A) Mudfl at deposit of core 3; ~1-mm- thick interbedded sandstone layers in interval dominated by clay. (B) Mudfl at with abun- dant postsedimentary dewatering structures, mostly convolute bedding (core 3). (C) Mudfl at in the Panamint Valley with desiccation cracks on the surface.

100% clay; they are structureless and red, due of 2 m (well 3) to 120 m (well 7) were drilled to the absence of the reducing effect of hydro- on the footwalls of the western (FZ-1) and the carbon migration (Chan et al., 2000). The sedi- eastern fault zone (FZ-2). Similar vesicular ments in the transition from dune to interdune andesitic and basaltic Upper Rotliegend II lava pond and/or lake deposits (Fig. 6) are composed fl ows were described by Geißler et al. (2008) of fi ne- to medium-grained sand with interca- from local grabens at the southern margin of the lated clays. They are interpreted to originate Southern Permian Basin. An example of a vesic- from the progradation of the lee sides of eolian ular textured lava fl ow is shown in Figure 13. dunes into ponds or lakes. Ripple laminations Lava tops are heavily brecciated due to rework- along marginal pond and/or lake deposits are ing, e.g., by mass fl ows. Therefore, a high common. Wet sandfl at deposits (Fig. 8) are very amount of volcaniclastic material is observed fi ne to fi ne-grained, poorly sorted sandstones in the sediments directly overlying the vol- and siltstones with clay contents of 20%–50% canic rocks. Thin-section analysis of the eolian (Amthor and Okkerman, 1998). In contrast, dune deposits above the volcanic rocks showed damp sandfl at deposits (Fig. 8) are slightly quartz grains discontinuously coated by early coarser grained, fi ne- to medium-grained sand- diagenetic chlorite and illite rims, which origi- stones with clay contents of <20% (Amthor and nate from weathered volcanic material (Fig. 12). Okkerman, 1998). Both facies types are char- Due to the fact that the seismic resolution at acterized by the occurrence of discontinuous, the depth of the target interval in northwestern irregular to wavy argillaceous adhesion ripples Germany only allows for basic interpretations and small-scale contortions (<0.2 m ampli- and well information in the study area is only tude). They are often accompanied by irregular, available from present-day structural highs, lensoid to aligned concentrations of sandstone (Mountney and Jagger, 2004; George and Berry, 1993). Thin-section analysis (Fig. 12) of the eolian dune and interdune deposits showed Figure 11. Exemplary resistivity based dip- quartz, feldspar, lithoclasts, and clay minerals meter log section and gamma ray (GR) log comprising chlorite and illite as main compo- section of German well 3a. Blue pattern nents. Sandstone classifi cation (after Pettijohn, (upward-increasing dip angle with uniform 1963) reveals that the sandstone is a litharenite. dip direction) indicates eolian dune succes- 0GR [API] 200 The detrital quartz grains are partially coated by sions, which were deposited under winds N discontinuous illite and minor chlorite rims. blowing from the east. Green pattern (uni- WE At the base of the sedimentary succession, form dip angle) indicates sandfl at deposits S vesicular textured andesitic to basaltic volcanic with unimodal dip direction. SD—standard Bedding after SD removal lava fl ows with extremely variable thicknesses deviation.

Geosphere, October 2012 1137

Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/8/5/1129/3342812/1129.pdf by guest on 30 September 2021 Vackiner et al.

Figure 12. (A) Photomicrograph of recent dune sample of the Panamint Valley dunes (location in Fig. 4). (B–F) Thin-section photos of samples from wells 3 and 3a of the northwestern German (GER) tight gas fi eld. Qz—quartz; Fsp–feldspar; IC—cutaneous illite; Ab—albite; IM— illite meshwork.

Figure 13. Comparison of volcanic lava fl ows observed from core mate- rial from the tight gas fi eld in north- western Germany to fi eld observations in Panamint Valley; photo location depicted in Figure 4. (A) Basaltic to andesitic lava fl ow in the core material of well 3. (B) Basaltic to andesitic lava fl ow in Panamint Valley, present on the footwall and on the hanging wall (intercalated into alluvial fan) of the Panamint Valley fault zone.

1138 Geosphere, October 2012

Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/8/5/1129/3342812/1129.pdf by guest on 30 September 2021 Panamint Valley—Field analog for a European Permian tight gas fi eld

conclusions about lateral continuity or distribu- tion of sedimentary facies or volcanic rocks can only be roughly interpolated and extrapolated. During the core interpretation and the work on isopach maps, uncertainties of the relative posi- tions of sediment bodies in relation to paleo- topography emerged. The alluvial fan deposits were observed in Upper Rotliegend II paleo- footwall positions of basinward downstepping fault zones. In addition, the main gas-bearing reservoir rocks, which are eolian dune deposits, conformably overlie these alluvial fan deposits. Accommodation of eolian dunes is thus depen- dent on wind direction and topography, similar to the controls on accommodation in the modern

Panamint Valley dune fi eld. Lateral extent and WGS— ledepot.com/maps/; internal architecture of the dune fi eld cannot be determined solely on the basis of existing sub- surface data.

Panamint Valley

The relative timing of fault activity with respect to sediment deposition as well as the ′

fault-sediment interaction of the Panamint Val- 117°18 ley (Figs. 4 and 5) can be determined based on plots. le fi eld and satellite observations. Sediments of either alluvial fans or eolian dunes cover most faults of the Panamint Valley. Faults in the east further offset active channels of alluvial fans, indicating recent synsedimentary fault activity (Figs. 5D, 5E). Enhanced cementation associ- ′

ated with fault planes (Figs. 5A–5C) and obvi- 117°22 ous sulfur smells suggest modern fl uid circula- tion in hydraulically active faults. Precipitation of euhedral calcite of as much as 0.5 cm diameter can be observed along fault scarps, but they are also present along certain stratigraphic layers, like limestone and mudstone or dolomite intervals, in the closer vicinity of faults. ′

To the north, east, and west of the Panamint 117°26 Valley, extensive alluvial fans descend from the footwall blocks of the outermost basin-bound- ing faults (Figs. 4 and 6). The alluvial fans cover the major faults, like the Ash Hill fault in the west and the Panamint Valley fault zone in the east, that, in turn, interact with the active

channel network of the fan apron. In the central ′ part of the basin, the alluvial fans partly cover the dry lake surface. The lack of varnish 117°30 implies recent along all allu- vial fans. Two types of alluvial fan systems can be related to variable slope angles in the north-

eastern and the northwestern basin (Blair and http://www.gpsfi Topo; Southwest USA (map source: Valley map of the Panamint Topographic 14. Left: Figure Geodetic System) showing alluvial walking tracks 1–8. Right: profi World McPherson, 1994). Alluvial fans dominated

by sheetfl ood and braided distributary stream ′ processes are mainly located in the northwest of Panamint Valley, e.g., along the northern 117°34

Ash Hill fault zone, and have slopes of ~4° ′ ′ ′

(Fig. 14). The northwestern Panamint Valley is 36°30 36°26 36°22

Geosphere, October 2012 1139

Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/8/5/1129/3342812/1129.pdf by guest on 30 September 2021 Vackiner et al.

dominated by late Pleistocene to recent alluvial (distal) margins of the dune fi eld. Such sand- fans were mechanically eroded and chemically fan deposits characterized by gravel bar and fl ats form during long-lasting strong winds and weathered by a network of braided channels swale microtopography that consist of silt, fi ne- periods when sand supplies are limited (Fry- (Fig. 13). The high amount of volcaniclastic to coarse-grained sand, , cobbles, and berger and Dean, 1979, 1983; Kocurek, 1988). material found in the alluvial fans indicates that rare (Jayko, 2009). The braided chan- A gradual increase in dip angles from 0°–5° at the volcanic rocks served as one of the major nel network comprises coarse-grained, mas- the dry sandfl at, to 10°–15° at the dune bases, local sediment sources. Furthermore, quartz sive gravel bar deposits and fi ne- to medium- to >15° at the dunes was observed. Defl ation grains from the dune succession on top of the grained, cross-stratifi ed intrachannel deposits. lags accompany or directly overlie the sandfl ats. alluvial fans and in front of the volcanic rocks In the northeastern Panamint Valley, debris Interdune deposits comprise muddy to fi ne- show discontinuous chlorite and illite coatings fl ow–dominated alluvial fans with enhanced grained sandy damp to wet sandfl ats (Fig. 8), that probably originate from weathered volcanic slope angles of as much as ~10° (Figs. 6 and 14) mudfl ats, lake margins (Fig. 7), and ponds (Figs. material. characterize the Panamint Valley fault zone. At 7 and 10). Dune marginal deposits (Fig. 7) are The described sedimentary facies the intersection point of the fan trench with the composed of fi ne-grained sand with interfi nger- architecture and the tectonic setting of the Pana- general alluvial fan surface, the confi ned fl ow ing clay. In many cases, a progradation of the mint Valley provide a 3D temporal snapshot that along fan-head channels changes into a distrib- lee side of eolian dunes into the ponds or lakes is in many aspects similar to the sedimentologi- utive fl ow along a network of sand-fi lled shal- was observed. Ripple lamination along marginal cal and structural setting of the subsurface study low channels (Fig. 6). The northeastern Pana- lake or pond deposits is common. The damp and site in Germany during the Upper Rotliegend mint Valley is dominated by late Pleistocene to wet sandfl at deposits (Fig. 8) were identifi ed II deposition. On the basis of the modern ana- Holocene alluvial deposits that are essentially as siltstones or very fi ne to fi ne-grained poorly log study in the Panamint Valley, a comparative unconsolidated (Jayko, 2009). A comparison of sorted sandstones. It can be assumed that these geological model of the Upper Rotliegend II in the eastern and the western fl anks of the Pana- sediments were deposited under the infl uence the subsurface study area was established. In the mint Valley shows that (1) the occurrence of of a shallow groundwater at ~3 m depth following, this model and a detailed sedimentary sheetfl ood and braid-dominated alluvial fans but were also affected by ephemeral fl ooding facies analysis at macroscale and microscale and is associated with lower topography and higher events (Fryberger et al., 1988; Meadows and its relationship to a fault-controlled morphology rates of Tertiary to Quaternary volcanic , 1993). The eolian sedimentation is fur- are discussed for both study sites. rocks and Paleozoic shales and carbonates, and ther infl uenced by the saline groundwater and (2) the occurrence of debris fl ow–dominated by the adhesion of wind-transported eolian DISCUSSION alluvial fans is associated with higher relief and grains to an episodically damp sediment sur- reduced erosion of Paleozoic carbonates, Ter- face. Small current ripples or horizontal lamina- The geological observations of the fi eld and tiary granitoids, and Tertiary to Quaternary vol- tions are common . the subsurface study site suggest that the Pana- canic rocks. Incision depths of >10 m of most The XRD measurements of the dune sand mint Valley with its complex synsedimentary, channels imply that the present erosion rate of samples show quartz, feldspar, and calcite transtensional fault zone activity and related the fans is higher than their accumulation rate. as main components. Accessories are illite/ sediment facies architecture can be used as a The northern Panamint Valley is characterized muscovite, chlorite, kaolinite, and amphibole. fi eld analog for the subsurface study site in by middle to Late Pleistocene (referred to as Thin-section analysis further revealed high per- Germany. The transtensional tectonic regime inactive) alluvial fan deposits (Jayko, 2009), centages of lithoclasts (Fig. 12), a composition that formed the Panamint Valley and that could which are cemented to variable degrees. indicating that the local sediment sources of the be reconstructed for the subsurface study site The Panamint Dunes are located on inactive sand dunes were the Quaternary alluvial fans, the in northwestern Germany during deposition of alluvial deposits north of the dry lake centered Tertiary volcanic rocks, the Mesozoic to Tertiary the Upper Rotliegend II is interpreted to have between the three major fault zones (Figs. 4 and granitoids, metasediments, and the Paleozoic infl uenced sedimentary processes, ultimately 8A). The dune fi eld, including dry sandfl ats with dolomites to limestones. XRD measure ments of causing a distinct sedimentological pattern. In wind ripples north and south of the dunes, cov- samples from the lake surface revealed dolomite, the following, we discuss the Panamint Valley’s ers an area of ~2.1 km × 4.5 km. The prevail- quartz, illite/muscovite, calcite, and feldspar as fault-controlled topography as an important ing wind direction is from the south and infre- the main components. Accessory minerals are controlling factor for the sedimentary facies quently from the northeast; this coincides with chlorite, kaolinite, hematite, thernadite, and distribution within the valley, and consider wind the dominant orientation measured for dune amphiboles. Swelling clays were identifi ed by directions and sediment provenance. The line and ripple crests. The dunes reach a maximum glycol dehydration of the clay fraction. of arguments used for the surface example will height of ~30 m and are classifi ed as star dunes to Patchy basaltic to andesitic volcanic rocks then be applied similarly to the subsurface study aklé dunes, while barchanoid and aklé dunes cover large parts of the mountain ranges sur- site (Figs. 15 and 16). In the following, the most are dominating toward the margins of the dune rounding the Panamint Valley (Fig. 13). This probable subsurface sedimentary facies distri- fi eld. The fault-induced relief of the Hunter main volcanic unit (ca. 14 Ma) is introduced by bution will be discussed. Mountain fault zone that borders the Panamint basaltic to andesitic fl ows and associated debris- On the basis of the fault-controlled paleo- Valley to the north acts as a trap for wind-blown fl ow deposits, which include minor amounts of topog raphy map (Fig. 15A), we reconstructed sediments, which are then deposited as eolian interlayered basaltic lava (Andrew and Walker, the alluvial fan deposits, which were observed at dunes and sandfl ats. The existence of overlap- 2009). The relatively thin layers were formed the base of the wind-blown sediments in the core ping step faults and synsedimentary fault activ- by lava fl ows (Andrew and Walker, 2009) that material. In the Panamint Valley, alluvial fans ity causes continuous sedimentation along fault followed a preexisting topography, and are sourced from the northwestern range are sheet- zones, leading to an easier upwind migration therefore localized along the major channel sys- fl ood and braid-dominated alluvial fans (Figs. 4 of dune sands to footwall positions. Sandfl ats tems on the hanging walls of faults. Lava fl ow and 14). Volcanic rocks and carbonate-cemented occur on the upwind (proximal) and downwind deposits that were also observed in the alluvial metasediments form the main sediment sources.

1140 Geosphere, October 2012

Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/8/5/1129/3342812/1129.pdf by guest on 30 September 2021 Panamint Valley—Field analog for a European Permian tight gas fi eld . Sedimen- ntial facies edimentary uvial transport directions ats. tary facies sketch. (A) Upper Rotliegend II thickness map of the subsurface tight gas study area in northwestern Germany. (B) S in northwestern Germany. Rotliegend II thickness map of the subsurface tight gas study area tary facies sketch. (A) Upper and pote interpretations of well log and core with respect Valley Panamint facies distribution and determination extracted from Rotliegend II isopach map. (C) Potential wind and fl distribution on seismic and isopach maps plotted onto Upper indicate potential accumulations of eolian dunes and sandfl Figure 15. Transferring the sedimentary facies outlines of the Panamint Valley to the northwestern German subsurface study site Valley the sedimentary facies outlines of Panamint Transferring 15. Figure

Geosphere, October 2012 1141

Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/8/5/1129/3342812/1129.pdf by guest on 30 September 2021 Vackiner et al.

A

B

Figure 16. Comparison between the simplifi ed sedimentary facies distribution of Panamint Valley (A) on the left and of the German sub- surface study area on the right (B).

As the elevation between the ephemeral dry lake elevation of ~250 m has been calculated by iso- At both study sites, dunes are sitting on sheet- and the mountain peaks in the northern Pana- pach analysis. On the basis of core interpreta- fl ood and braid-dominated alluvial fans in middle mint Valley is ~350 m in the west and 1350 m tion and due to comparable topography, it can slope position (Figs. 4 and 9). The prevailing in the east (Fig. 14), we focus on the western be interpreted that similar sheetfl ood and braid- wind direction and the topographic relief are of fl ank’s topography to compare with the recon- dominated alluvial fans developed in the Ger- primary interest for the analysis of the distribu- structed paleotopography of the subsurface man study site. The paleotopographies, from tion of wind-blown sediments. In the Panamint study site. The extent and pattern of the facies in which the alluvial fans originate, are Carbonif- Valley, the main wind direction is from the south. the subsurface study site in northwestern Ger- erous carbonates and Rotliegend volcanic rocks. Hunter Mountain, which rises 370 m above the many were transferred from the facies map of The general sediment source composition is dunes, represents a windward trap for eolian the Panamint Valley (Fig. 15B). For the German thus also in accord with our observations from sediment (Fig. 16). The well log, report, and subsurface study site, a minimum paleofootwall the Panamint Valley. core data set of the study area in northwestern

1142 Geosphere, October 2012

Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/8/5/1129/3342812/1129.pdf by guest on 30 September 2021 Panamint Valley—Field analog for a European Permian tight gas fi eld

Germany were used as basic information to ephemeral lakes, or was at least exposed to the 16) comparing the Panamint Valley sedimentary verify the locations of the reconstructed sedi- infl uence of a near-surface water table. This is facies distribution to the situation in the subsur- mentary facies distribution; the most probable supported by the distinct polygonal pattern in face study site in Germany offers the most prob- locations for dune and sandfl at accumulations this part of the basin, very similar in terms of able possible correlation of sediment dynamics. are depicted in Figure 15C. Comparable to the shape and size to the polygonal pattern on the modern analog study, dunes on the footwall of dry lake surface in the Panamint Valley (Antrett CONCLUSION FZ-2 in the subsurface study area developed et al., 2012). Due to the postulated high mois- as small barchanoid to aklé dunes under a uni- ture content of the sediment surface, eolian 1. The Panamint Valley’s heterogeneous sedi- modal east–east-northeast wind direction in a dunes might have accumulated and might be mentary facies comprise different types of allu- lee side trap (Figs. 15 and 16). The dunes of the preserved on the hanging wall of FZ-1. vial fans, sand dunes, mudfl ats, and sandfl ats. Panamint Valley and the subsurface site show At both study sites, patchy andesitic to basal- The distribution of these facies is controlled by a high degree of similarity in terms of dune tic lava fl ow units with vesicular texture occur in synsedimentary faults and topography, the local heights (i.e., 20–30 m), the amount of inter- the footwalls (Fig. 13). In the Panamint Valley, sediment sources, and the prevailing wind direc- dune deposits, and the dune sediment composi- these volcanic rocks are also intercalated with tion. The abrasion of quartz grain coatings, dune tion analyzed from thin sections. Furthermore, the alluvial fans in hanging-wall positions. Con- types and sizes, the presence or absence of des- at both study sites dune bases often interfi nger sequently, the hanging-wall volcanic rocks are ert varnish, and the incision depth of alluvial fan with aquatic interdune deposits, associated with eroded by the alluvial fan braided channel sys- channels can be used as proxies for estimating hydroplastic deformation structures (Fig. 7). tem and serve as active sediment sources (Fig. the sediment dynamics. Well records from the footwall of FZ-1 reveal 13). In the German subsurface study site, the 2. Core data analysis of the sedimentary the occurrence of sandfl ats, deposited on top occurrence of basaltic to andesitic lava fl ows is facies from the subsurface tight gas reservoir of fl uvial sandstones and conglomerates with only proven for footwall positions. Due to the in Germany compared to the Panamint Valley underlying monomict volcaniclastic breccias fact that volcanic lava fl ows are on the footwall reveals sheetfl ood and braid-dominated alluvial ranging to conglomerates and volcanic rocks. close to the fault zone, we propose that they fan deposits, partly stacked sand dune, sandfl at, Footwalls only exhibit sediment thicknesses of continue into graben positions. The input of and interdune deposits. Cored sediment thick- ~200 m, whereas hanging-wall sediments are as altered volcanic material from the footwall into nesses and facies reveal pronounced changes thick as ~450 m. The fault-induced paleorelief the deeper part of the basin provides a source across synsedimentary fault zones. is estimated as ~250 m during deposition of of reactive Al3+ and Si4+ and supports excessive 3. At both study sites, the presence of patchy the Upper Rotliegend II. Based on the distinct and early diagenetic development of alumino- basaltic to andesitic volcanic lava fl ow units facies distribution observed in the Panamint Val- silicate minerals (Jeans et al., 2000). Swelling strongly infl uenced the sediment composition ley, the isopach maps that reveal a paleorelief of clays (e.g., smectite) most likely originate from and supplied clays to the sedimentary systems. ~250 m, and the prevailing east–east-northeast the weathering of the volcanic rocks (Roen and 4. We developed a model comprising topog- wind direction, it was inferred that the studied Hosterman, 1982). raphy, synsedimentary faults, and wind direc- subsurface fault zone might have formed a wind- Thin-section analysis of both Panamint dune tions as key controlling factors of the sediment ward trap with a higher amount of conglomer- samples and the Upper Rotliegend II dune facies distribution at the analog study site and ates and sand accumulations in the hanging-wall deposits reveal discontinuous coating of quartz compared this to the subsurface area recon- position (Fig. 15C). The increased Upper Rot- and feldspar grains by detrital or synsedimen- structed to an Upper Rotliegend II setting prior liegend II sandstone thickness in the hanging tary early diagenetic chlorite and illite (Fig. 12), to multiphase tectonic overprinting. The loca- wall is also governed by the preservation space which is also provided by the weathering of vol- tion of eolian sandstones comprising dune and that cannot be determined from the comparison canic material and transported via the alluvial sandfl at deposits controlled by fault-induced to the modern analog, and thus represents the fan channel system. Furthermore, mechanical topography acting as an effective trap for eolian major uncertainty in the comparison. The pres- abrasion of clay coatings around quartz grains sand is particularly similar. ervation of dune deposits in hanging-wall posi- is observed in samples from both study sites. In 5. The fi eld analog observations concerning tions (FZ-1) may be related to the situation in stabilized dunes, subject to seasonal rainfall or the abrasion of quartz grain coatings, dune types a relative depression, which is less affected by intermittently fl ooded dunes, coatings can be and sizes, and the incision depth of alluvial fan erosion, and to increased moisture content at the very continuous (Winspear and Pye, 1995). In channels can be used to reconstruct the sediment dune bases. Considering accommodation and contrast to this, the coats are abraded by eolian dynamics of the tight gas reservoir during depo- preservation, the most probable realization of transport if dune sands are subsequently defl ated sition of the Upper Rotliegend II. At both study dune and sandfl at deposits is depicted in Figures and remobilized into active dunes (Walker, sites, the early postsedimentary chlorite and illite 15C and 16. Eolian sandstones might also have 1979; Ajdukiewicz et al., 2010). We therefore coatings around the detrital grains are abraded, reached and been preserved in the reconstructed propose that active dune sediment transport is indicating active transport of dune sands; this is basin center. responsible for the abrasion of the clay coatings supported at both study sites by the occurrence The ephemeral dry lake at the Panamint Val- around quartz grains at both study sites. of nonstable dune form deposits, such as barcha- ley (Fig. 4; Supplemental File [see footnote 1]) The main uncertainties in comparing the sedi- noid and aklé dune deposits, in the cores from is characterized by the occurrence of huge desic- ment distribution of the Panamint Valley to the the subsurface study site and in the Panamint ca tion polygons and is centered in the deepest German subsurface study site are represented by Valley. The Upper Rotliegend II paleotopog- part of the basin. The deepest part of the syn- the localization of the mobile eolian sands (Fig. raphy in the subsurface study site in Germany sedimentary Rotliegend asymmetric graben to 15C). Over time, the German study site might is not as pronounced as the current topography half-graben basin is located on the hanging wall be affected by a different rate of sediment sup- in the Panamint Valley; therefore, the different of FZ-1 (Figs. 2 and 15A). We assume that this ply, higher or lower water table, and small varia- sedimentary facies provide different basinward might have become preferentially occupied by tions in topography. However, the model (Fig. extensions. Due to shallower paleogradients at

Geosphere, October 2012 1143

Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/8/5/1129/3342812/1129.pdf by guest on 30 September 2021 Vackiner et al.

the subsurface study site, the assumed ephem- Burchfi el, B.C., Hodges, K.V., and Royden, L.H., 1987, Jennings, C.W., 1975, Fault map of California with location eral dry lake, for example, may have fl ooded a Geology of Panamint Valley–Saline Valley pull-apart of volcanoes, thermal springs, and thermal wells: Cali- system, California: Palinspastic evidence for low-angle fornia Division of Mines and Geology Geologic Data much larger area. geometry of a Neogene range bounding fault: Journal Map 1, scale 1:750,000. 6. Our study shows that a thorough surface- of Geophysical Research, v. 92, no. B10, p. 10,422– Jennings, C.W., Bryant, W.A., and Saucedo, G., 2010, Fault 10426, doi:10.1029/JB092iB10p10422. activity map of California: California Geologic Data subsurface analog study enables the detailed Chan, M.A., Parry, W.T., and Bowman, J.R., 2000, Dia- Map Series Map 6, scale 1:750,000. interpretation, interpolation, and prediction of genetic hematite and manganese oxides and fault-related Kocurek, G., 1988, First-order and super bounding surfaces sedimentary facies in areas characterized by lim- fl uid fl ow in Jurassic sandstones, southeastern : in eolian sequences—Bounding surfaces revisited: American Association of Petroleum Geologists Bulle- Sedimentary Geology, v. 56, p. 193–206, doi:10.1016 ited subsurface information. The study further tin, v. 84, p. 1281–1310, doi:10.1306/A9673E82-1738 /0037-0738(88)90054-1. shows that paleosediment dynamics can be recon- -11D7-8645000102C1865D. Kocurek, G., 2003, Limits on extreme eolian systems: structed using a paleotopographic restoration . Densmore, A.L., and Anderson, R.S., 1997, Tectonic geo- of and Jurassic Navajo Sandstone examples, morphology of the Ash Hill fault, Panamint Valley, in Chan, M.A., and Archer, A.W., eds., Extreme depo- ACKNOWLEDGMENTS California: Basin Research, v. 9, p. 53–63, doi:10.1046 sitional environments: Mega end members in geologic /j.1365-2117.1997.00028.x. time: Geological Society of America Special Paper 370, The study is part of the Wintershall and RWTH Fryberger, S.G., and Dean, G., 1979, Dune forms and wind p. 43–52, doi:10.1130/0-8137-2370-1.43. Aachen University Tight Gas Initiative. We thank regime, in McKee, E.D., ed., A study of global sand Kocurek, G., and Havholm, K.G., 1993, Eolian sequence : Honolulu, Hawaii, University Press of the Pacifi c, stratigraphy; a conceptual framework, in Weimer, P., Wintershall Holding GmbH and GDF Suez E&P p. 137–169. and Posamentier, H.W., eds., Siliciclastic sequence Deutschland GmbH for providing the data and sup- Fryberger, S.G., Al-Sari, A.M., and Clisham, T.J., 1983, stratigraphy: Recent developments and applications: porting this project. The paper benefi tted in par- Eolian dune, interdune, sand sheet, and siliciclastic American Association of Petroleum Geologists Mem- ticular from fruitful discussions with Claudia Bärle, sediments of an offshore prograding sand sea, oir 58, p. 393–409. Harald Karg, Bernhard Siethoff, Wolfram Unver- Dhahran area, Saudi Arabia: American Association of Lee, J., Stockli, D.F., Owen, L.A., Finkel, R.C., and Kislit- haun, Petra Unverhaun, Wolf-Dieter Karnin, and Petroleum Geologists Bulletin, v. 67, p. 280–312. syn, R., 2009, Exhumation of the Inyo Mountains, Cali- Anton Irmen. We also thank Norbert Klitzsch and Fryberger, S.G., Schenk, C.J., and Krystinik, L.F., 1988, fornia: Implications for the timing of extension along the E.ON Energy Research Center, who gave us the Stokes surfaces and the effects of near-surface ground- the western boundary of the Basin and Range Province water-table on aeolian deposition: , v. 35, and distribution of dextral fault slip rates across the opportunity to conduct ground resistivity measure- p. 21–41, doi:10.1111/j.1365-3091.1988.tb00903.x. eastern California shear zone: Tectonics, v. 28, TC1001, ments. We also thank the United States National Gast, R.E., 1988, Rifting im Rotliegenden Niedersachsens: Die doi:10.1029/2008TC002295. Park Services, especially Richard Friese, for issu- Geowissenschaften Weinheim, v. 6, no. 4, p. 115–122. Legler, B., 2005, Faziesentwicklung im Südlichen Perm- ing a research and collecting permit for Death Val- Gast, R., 1991, The perennial Rotliegend saline lake in NW becken in Abhängigkeit von Tektonik, eustatischen ley National Park. We are grateful to Midland Valley Germany: Geologisches Jahrbuch, v. A119, p. 25–59. Meeresspiegelschwankungen des Proto-Atlantiks und Ltd. for providing academic software licenses for Gast, R., and Gaupp, R., 1991, The sedimentary record Klimavariabilität (Oberrotliegend, Nordwesteuropa): use at RWTH Aachen University. of the Late Permian saline lake in N.W. Germany, in Schriftenreihe der Deutschen Gesellschaft für Geowis- Renaut, R.W., and Last, W.M., eds., Sedimentary and senschaften, v. 47, p. 103. REFERENCES CITED paleolimnological records of saline lakes: Saskatoon, Legler, B., and Schneider, J.W., 2008, Marine ingressions into , Canada, National Hydrology Research the Middle/Late Permian saline lake of the Southern Ajdukiewicz, J.M., Nicholson, P.H., and Esch, W.L., 2010, Pre- Centre, p. 75–86. Permian Basin (Rotliegend, Northern Germany) pos- diction of deep reservoir quality using early diagenetic Gaupp, R., and Solms, M., 2005, Palaeo oil- and gasfi elds sibly linked to sea-level highstands in the Arctic rift process models in the Jurassic Nophlet Formation, in the Rotliegend of the North German Basin: Effects system: Palaeogeography, Palaeoclimatology, Palaeo- of : American Association of Petroleum upon hydrocarbon reservoir quality (Paläo-Öl- und ecology, v. 267, p. 102–114, doi:10.1016/j.palaeo.2008 Geologists Bulletin, v. 94, p. 1189–1227, doi:10.1306 Gasfelder im Rotliegenden des Norddeutschen Beckens: .06.009. /04211009152. Wirkungen der KW-Migration auf die Speicherqualitäts- Lohr, T., Krawczyk, M., Tanner, D.C., Samiee, R., Endres, Amthor, J.E., and Okkerman, J., 1998, Infl uence of early Entwicklung), in Gaupp, R., ed., Tight gas reservoirs— H., Oncken, O., Trappe, H., and Kukla, P.A., 2007, diagenesis on reservoir quality of Rotliegende Sand- Natural gas for the future: DGMK Research Report Strain partitioning due to salt; insights from interpreta- stone, northern Netherlands: American Association of 593-8, 237 p. tion of a 3D seismic data set in the NW German Basin: Petroleum Geologists Bulletin, v. 82, p. 2246–2265. Geißler, M., Breitkreuz, C., and Kiersnowski, H., 2008, Late Basin Research, v. 19, p. 579–597, doi:10.1111/j.1365 Andrew, J.E., and Walker, J.D., 2009, Reconstructing late Paleozoic volcanism in the central part of the southern -2117.2007.00338.x. Cenozoic deformation in central Panamint Valley, Cali- Permian Basin (NE Germany, W Poland): Facies dis- McCann, T., 1998, The Rotliegend of the NE German Basin: fornia: Evolution of slip partitioning in the Walker Lane: tribution and volcano-topographic hiati: International Background and prospectivity: Petroleum Geoscience, Geosphere, v. 5, p. 172–198, doi:10.1130/GES00178.1. Journal of Earth Sciences, v. 97, p. 973–989, doi: v. 4, p. 17–27, doi:10.1144/petgeo.4.1.17. Antrett, P., Vackiner, A.A., Kukla, P.A., Stollhofen, H., and 10.1007/s00531-007-0288-6. Meadows, N.S., and Beach, A., 1993, Structural and climatic Klitzsch, N., 2012, Impact of arid surface mega-cracks George, G.T., and Berry, J.K., 1993, A new lithostratigra- controls on facies distribution in a mixed fl uvial and on hydrocarbon reservoir properties: American Asso- phy and depositional model for the Upper Rotliegend aeolian reservoir; the Triassic Sherwood Sandstone in ciation of Petroleum Geologists Bulletin, v. 96, no. 7, of the UK Sector of the southern North Sea, in North, the , in North, C.P., and Prosser, D.J., eds., p. 1279–1299. C.P., and Prosser, D.J., eds., Characterization of fl uvial Characterization of fl uvial and aeolian reservoirs: Allen, P.A., and Allen, J.R., 1990, Basin analysis: Principles and aeolian reservoirs: Geological Society of London Geological Society of London Special Publication 73, and applications: Cambridge, Blackwell Sciences, 451 p. Special Publication 73, p. 291–319, doi:10.1144/GSL p. 247–264, doi:10.1144/GSL.SP.1993.073.01.15. Aydin, A., and Nur, A., 1985, The types and role of stepovers .SP.1993.073.01.18. Mountney, N.P., 2006, Eolian facies models, in Posamentier, in strike-slip tectonics, in Biddle, K.T., and Christie- Glennie, K.W., 1990a, Rotliegend sediment distribution; a H., and Walker, R.G., eds., Facies models revisited: Blick, N., eds., Strike-slip deformation, basin formation, result of Late Carboniferous movements, in Hardman, Society for Sedimentary Geology (SEPM) Memoir 84, and sedimentation: Society of Economic Paleontologists R.F.P., and Brooks, J., eds., Proceedings of tectonic p. 19–83. and Mineralogists Special Publication 37, p. 35–44. events responsible for Britain’s oil and gas reserves: Mountney, N.P., and Jagger, A., 2004, Stratigraphic evolution Blair, T.C., and McPherson, J.G., 1994, Alluvial fans and Geological Society of London Special Publication 55, of an margin aeolian system: The Permian Cedar their natural distinction from based on morphol- p. 127–138, doi:10.1144/GSL.SP.1990.055.01.06. Mesa Sandstone, SE Utah, USA: Sedimentology, v. 51, ogy, hydraulic processes, sedimentary processes, and Glennie, K.W., 1990b, Introduction to the petroleum geol- p. 713–743, doi:10.1111/j.1365-3091.2004.00646.x. facies assemblages: Journal of Sedimentary Research, ogy of the North Sea (third edition): London, Wiley- Pettijohn, F.J., 1963, Chemical composition of sandstones, v. A64, p. 450–489, doi:10.1306/D4267DDE-2B26 Blackwell, 416 p. excluding carbonate and volcanic sands: Representative -11D7-8648000102C1865D. Jayko, A.S., 2009, Surfi cial geologic map of the Darwin analyses, in Data of geochemistry (sixth edition): U.S. Blakely, R.J., and Ponce, D.A., 2001, Map showing depth Hills 30′ × 60′ Quadrangle, Inyo County, California: Geological Survey Professional Paper 440-S, p. 1–21. to pre-Cenozoic basement in the ground- U.S. Geological Survey Scientifi c Investigations Map Plein, E., 1993, Observations on Stephanian and Rotliegendes water model area, Nevada and California: U.S. 3040, scale 1:100,000. palaeogeography in the North German Basin, in Geology Geological Survey Miscellaneous Field Study Map Jayko, A.S., Forester, R.M., Kaufman, D.S., Phillips, F.M., and hydrocarbon potential of the Permian rocks of the MF-2361–3, scale 1:250,000. Yount, J.C., McGeehin, J., and Mahan, S.A., 2008, Late eastern North German Basin (in German): Geologisches Blakely, R.J., Jachens, R.C., Calzia, J.P., and Langenheim, Pleistocene lakes and wetlands, Panamint Valley, Inyo Jahrbuch, Reihe A, Allgemeine und Regionale Geologie V.E., 1999, Cenozoic basins of the Death Valley County, California: Geological Society of America Spe- BR Deutschland und Nachbargebiete, Tektonik, Stratig- extended terrane as refl ected in regional-scale gravity cial Paper 439, p. 151–184, doi:10.1130/2008.2439(07). raphie, Paläontologie, v. 131, p. 99–116. anomalies, in Wright, L.A., and Troxel, B.W., eds., Jeans, C.V., Wray, D.S., Merriman, R.J., and Fisher, M.J., Prestud Anderson, S., and Anderson, R.S., 1990, Debris- Cenozoic basins of the Death Valley region: Geologi- 2000, Volcanogenic clays in Jurassic and Cretaceous fl ow benches: Dune-contact deposits record paleo- cal Society of America Special Paper 333, p. 1–16, strata of and the North Sea Basin: Clay Min- sand dune positions in north Panamint Valley, Inyo doi:10.1130/0-8137-2333-7.1. erals, v. 35, p. 25–55, doi:10.1180/000985500546710. County, California: Geology, v. 18, p. 524–527, doi:

1144 Geosphere, October 2012

Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/8/5/1129/3342812/1129.pdf by guest on 30 September 2021 Panamint Valley—Field analog for a European Permian tight gas fi eld

10.1130/0091-7613(1990)018<0524:DFBDCD of Geophysical Research, v. 85, p. 3711–3739, doi: Walker, T.R., 1979, Red color in dune sand, in McKee, E.D., >2.3.CO;2. 10.1029/JB085iB07p03711. ed., A study of global sand seas: U.S. Geological Sur- Price, N., and Cosgrove, J., 1990, Analysis of geological struc- Smith, S.U., 1976, Late-Quarternary pluvial and tectonic vey Professional Paper 1052, p. 61–81. tures: Cambridge, Cambridge University Press, 520 p. history of Panamint Valley, Inyo and San Bernardino Winspear, N.R., and Pye, K., 1995, The origin and signifi - Rieke, H., Kossow, D., McCann, T., and Krawczyk, C., 2001, Counties, California [Ph.D. thesis]: Pasadena, Califor- cance of boxwork clay coatings on dune sand grains Tectono-sedimentary evolution of the northernmost nia Institute of Technology, 295 p. from the Nebraska Sand Hills, U.S.A: Sedimentary margin of the NE German Basin between uppermost Stewart, J.H., 1988, Tectonics of the Walker Lane belt, west- Geology, v. 94, p. 245–254, doi:10.1016/0037-0738 Carboniferous and Late Permian (Rotliegend): Geo- ern Great Basin: Mesozoic and Cenozoic deformation (94)00090-H. logical Journal, v. 36, p. 19–37, doi:10.1002/gj.873. in a zone of shear, in Ernst, W.G., ed., Metamorphism Ziegler, A.M., 1990, Phytogeographic patterns and conti- Roen, J.B., and Hosterman, J.W., 1982, Misuse of the and crustal evolution of the western United States: nental confi gurations during the Permian Period, in term “bentonite” for ash beds of Devonian age in the Rubey Volume VII: Englewood Cliffs, , McKerrow, W.S., and Scotese, C.R., eds., Palaeozoic Appalachian Basin: Geological Society of Amer- Prentice Hall, p. 683–713. palaeogeography and biogeography: Geological Soci- ica Bulletin, v. 93, p. 921–925, doi:10.1130/0016 Vackiner, A.A., Antrett, P., Stollhofen, H., Back, S., Kukla, P., ety of London Memoir 12, p. 363–379, doi:10.1144 -7606(1982)93<921:MOTTBF>2.0.CO;2. and Bärle, C., 2011, Syndepositional tectonic controls /GSL.MEM.1990.012.01.35. Sclater, J.G., and Christie, P.A.F., 1980, Continental stretch- and palaeo-topography of a Permian tight gas reservoir Ziegler, P.A., 1982, Geological atlas of Western and ing: An explanation of the post-mid Cretaceous sub- in NW Germany: Journal of Petroleum Geology, v. 34, Central Europe: Amsterdam, Elsevier Science Ltd., sidence of the central North Sea basin: Journal p. 411–428, doi:10.1111/j.1747-5457.2011.00512.x. 130 p.

Geosphere, October 2012 1145

Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/8/5/1129/3342812/1129.pdf by guest on 30 September 2021