30.01.2015 Views

Reservoir properties and petrophysical modelling of carbonate sand ...

Reservoir properties and petrophysical modelling of carbonate sand ...

Reservoir properties and petrophysical modelling of carbonate sand ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Geological Society, London, Special Publications published online June 27, 2012 as doi: 10.1144/SP370.6<br />

<strong>Reservoir</strong> <strong>properties</strong> <strong>and</strong> <strong>petrophysical</strong> <strong>modelling</strong><br />

<strong>of</strong> <strong>carbonate</strong> s<strong>and</strong> bodies: outcrop analogue study<br />

in an epicontinental basin (Triassic, Germany)<br />

DENIS PALERMO 1,2 *, THOMAS AIGNER 1 , BJOERN SEYFANG 1,3 & SERGIO NARDON 2<br />

1 Department <strong>of</strong> Geosciences, Centre <strong>of</strong> Applied Geosciences,<br />

University <strong>of</strong> Tübingen, Sigwartstr. 10, D-72076 Tübingen, Germany<br />

2 Eni S.p.A. – Exploration & Production Division, via Emilia 1,<br />

20097 San Donato Milanese, Italy<br />

3 Present address: Centre Scientifique et Technique Jean Feger (CSTJF),<br />

Avenue Larribau, F-64018 Pau Cedex, France<br />

*Corresponding author (e-mail: denis.palermo@eni.it)<br />

Abstract: This paper represents the second part <strong>of</strong> an integrated study that is focussed on the<br />

development <strong>and</strong> distribution <strong>of</strong> reservoir bodies <strong>and</strong> <strong>properties</strong> in epeiric <strong>carbonate</strong> systems. It<br />

is based on outcrop analogue data from Triassic ‘layer-cake’ <strong>carbonate</strong>s in the South German<br />

Basin, which were deposited along an epicontinental, very gently inclined <strong>carbonate</strong> ramp.<br />

The reservoir facies consists <strong>of</strong> skeletal <strong>and</strong> oolitic <strong>carbonate</strong> grainstones (F max 23%, K max<br />

700 mD), which are organized in a pronounced hierarchy <strong>of</strong> stratigraphic cycles. Based on outcrops,<br />

cores, gamma ray (GR) logs <strong>and</strong> thin sections, a high-resolution, geocellular 3D facies<br />

model was generated, which covers the area <strong>of</strong> a Middle East giant gas field (25 × 36 km). The<br />

spatial distribution <strong>of</strong> reservoir <strong>properties</strong> was systematically investigated on different scales.<br />

The lateral distribution <strong>of</strong> reservoir <strong>properties</strong> remains in the same order <strong>of</strong> magnitude for hundreds<br />

<strong>of</strong> metres, within in the same stratigraphic position. However, on a kilometre scale, facies bodies,<br />

diagenetic trends <strong>and</strong> thus reservoir <strong>properties</strong> show gradual lateral changes. Vertically, in contrast,<br />

<strong>properties</strong> change commonly on a decimetre scale <strong>and</strong> are largely controlled by stratigraphic<br />

cycles. Petrophysical <strong>modelling</strong> enhanced the underst<strong>and</strong>ing <strong>of</strong> key factors <strong>and</strong> processes controlling<br />

both reservoir quality <strong>and</strong> quantity.<br />

<strong>Reservoir</strong> heterogeneity remains a significant issue in<br />

reservoir <strong>modelling</strong> <strong>and</strong> prediction <strong>of</strong> field performance.<br />

Outcrop analogues are ideal for establishing<br />

the possible geometries <strong>and</strong> property distributions<br />

at the inter-well scale. Quantitative data on <strong>carbonate</strong><br />

rock bodies remain scarce (e.g. H<strong>and</strong>ford 1988;<br />

Burchette et al. 1990; Grant et al. 1994; Harris &<br />

Kowalik 1994; Borgomano et al. 2002; Grammer<br />

et al. 2004; Kostic & Aigner 2004; Ruf & Aigner<br />

2004; Rankey et al. 2006; Qi et al. 2007; Aigner<br />

et al. 2007; Palermo et al. 2008). Qualitative outcrop<br />

studies <strong>of</strong> geometry (e.g. Burchette et al. 1990;<br />

Gawthrope & Gutteridge 1990; Azerêdo 1998) <strong>and</strong><br />

quantitative <strong>petrophysical</strong> studies carried out on<br />

outcrops at the appropriate scale (e.g. Kittridge<br />

et al. 1990; Senger et al. 1991; Eisenberg et al.<br />

1994; Cavallo & Smosna 1997; Jennings 2000;<br />

Savary & Ferry 2004; Pranter et al. 2005, 2006)<br />

made useful contributions to the available data.<br />

Outcrop analogue studies <strong>of</strong> this nature are<br />

important in order to condition subsurface reservoir<br />

models to real data <strong>and</strong> concepts, thus significantly<br />

improving field appraisal work <strong>and</strong> development<br />

planning. The present work has resulted from a<br />

joint ENI E&P–University Research Consortium<br />

on the ‘Geometry <strong>of</strong> Carbonate Objects’. Within<br />

this project, the Triassic Upper Muschelkalk <strong>carbonate</strong>s<br />

in the South-German Basin were studied<br />

as an analogue to the ‘layer-cake’-type reservoir<br />

systems <strong>of</strong> the Middle East. The Upper Muschelkalk<br />

was deposited on a gently inclined <strong>carbonate</strong> ramp,<br />

filling an epicontinental basin, <strong>and</strong> therefore represents<br />

an analogue to an important type <strong>of</strong> ‘nonreefal’<br />

skeletal <strong>and</strong> oolitic <strong>carbonate</strong> s<strong>and</strong> reservoirs<br />

(e.g. Khuff, Hanifa, <strong>and</strong> Arab in the Middle East).<br />

A series <strong>of</strong> excellent quarries <strong>and</strong> natural outcrops<br />

in Southern Germany has been used to investigate<br />

the geometries within high-energy shoal<br />

water deposits along the margin <strong>of</strong> the Upper<br />

Muschelkalk Basin. Close outcrop spacing allows<br />

for lateral tracing <strong>of</strong> beds <strong>and</strong> mapping <strong>of</strong> lateral<br />

facies transitions. A hundred years <strong>of</strong> detailed<br />

From: Garl<strong>and</strong>, J., Neilson, J. E., Laubach,S.E.&Whidden, K. J. (eds) 2012. Advances in Carbonate Exploration<br />

<strong>and</strong> <strong>Reservoir</strong> Analysis. Geological Society, London, Special Publications, 370,<br />

http://dx.doi.org/10.1144/SP370.6 # The Geological Society <strong>of</strong> London 2012. Publishing disclaimer:<br />

www.geolsoc.org.uk/pub_ethics


Geological Society, London, Special Publications published online June 27, 2012 as doi: 10.1144/SP370.6<br />

D. PALERMO ET AL.<br />

logging <strong>and</strong> mapping has resulted in a wellconstrained<br />

litho- <strong>and</strong> biostratigraphic framework<br />

(e.g. Wagner 1913; Vollrath 1938, 1955, 1957,<br />

1958, 1970; Skupin 1969; Bachmann 1973; Aigner<br />

1985; Ulrichs & Mundlos 1987, 1990; Hagdorn &<br />

Simon 1988; Ockert 1988; Geyer & Gwinner 1991;<br />

Braun 2003). This well-established stratigraphy provides<br />

the opportunity to transform the so far largely<br />

descriptive picture into a genetic, process-based<br />

analysis <strong>of</strong> the sedimentary bodies <strong>and</strong> packages.<br />

A detailed outcrop, core <strong>and</strong> well logging programme<br />

was carried out (e.g. GR logs, cores, porosity<br />

<strong>and</strong> permeability), allowing an analysis <strong>of</strong> the<br />

sedimentology, <strong>petrophysical</strong> characteristics, internal<br />

architecture <strong>and</strong> reservoir <strong>properties</strong>, within a<br />

3D geocellular model. Palermo et al. (2010) have<br />

recently reported on the facies <strong>modelling</strong> aspects<br />

<strong>of</strong> this project. This companion paper focusses on<br />

reservoir <strong>properties</strong> <strong>and</strong> <strong>petrophysical</strong> <strong>modelling</strong>.<br />

Geological setting<br />

During the Triassic, rifting <strong>and</strong> disintegration <strong>of</strong> the<br />

Pangean supercontinent caused the western extension<br />

<strong>of</strong> the Tethys Ocean. This regional crustal extension<br />

induced the subsidence <strong>of</strong> a complex network <strong>of</strong><br />

grabens <strong>and</strong> troughs in Western <strong>and</strong> Central Europe<br />

(e.g. Ziegler 1990) <strong>and</strong> formed the Triassic basins<br />

including the South German Muschelkalk Basin<br />

(Fig. 1). This basin extended roughly from Pol<strong>and</strong><br />

to the North Sea <strong>and</strong> from the Alpine foothills to<br />

Denmark. During the Middle Triassic (≏240–<br />

231 Ma), the Muschelkalk Basin was covered by an<br />

epicontinental, semi-enclosed marginal sea that was<br />

separated by the Vindelician/Bohemian High from<br />

the open Tethys ocean in the SE. Temporary connections<br />

existed only through three narrow shifting seaways<br />

(Ziegler 1990; Dercourt et al. 1993). In the<br />

early Anisian a relative sea-level rise induced the<br />

Lower Muschelkalk transgression from the open<br />

Tethys ocean via the Silesian–Moravian <strong>and</strong> East-<br />

Carpathian Gates into the Germanic Basin with<br />

fully marine conditions. The open communication<br />

became restricted during the later Anisian, resulting<br />

in hypersaline conditions <strong>and</strong> formation <strong>of</strong> evaporites<br />

<strong>of</strong> the Middle Muschelkalk. Finally the fully marine<br />

<strong>carbonate</strong>s <strong>of</strong> the Upper Muschelkalk were deposited<br />

during a transgression in the uppermost Anisian, connecting<br />

the Germanic Basin with the open Tethys<br />

realm via the three gates. The Upper Muschelkalk<br />

succession (Middle/Late Ladinian) represents one<br />

overall transgressive/regressive third-order sea-level<br />

cycle (Aigner 1985). The transgressive part is represented<br />

by an overall fining-upward sequence <strong>of</strong><br />

crinoidal/shelly shoal water <strong>carbonate</strong>s grading into<br />

Fig. 1. Palaeogeography <strong>of</strong> Central Europe in the Middle Triassic modified after Ziegler (1990), Hagdorn (1991) <strong>and</strong><br />

Palermo et al. (2010) (AAPG # 2010, reprinted by permission <strong>of</strong> the AAPG, whose permission is required for further use).


Geological Society, London, Special Publications published online June 27, 2012 as doi: 10.1144/SP370.6<br />

CARBONATE SAND BODIES IN GERMANY<br />

muddy <strong>carbonate</strong>s <strong>and</strong> marlstones. The upper regressive<br />

section is represented by an overall coarseningupward<br />

sequence, grading from the muddy sediments<br />

around maximum transgression into shelly/oolitic<br />

shoal <strong>carbonate</strong>s <strong>and</strong> muddy backshoal sediments.<br />

The boundary between the Muschelkalk <strong>and</strong> the<br />

overlying, mainly siliciclastic Keuper is considered<br />

as maximum regression (Aigner et al. 1999). Generally,<br />

five major facies belts can be distinguished<br />

in the Upper Muschelkalk: (a) a narrow zone <strong>of</strong><br />

coastal siliciclastics; (b) an irregular backshoal zone<br />

dominated by (partly dolomitized) peloidal mud<strong>and</strong><br />

wackestones; (c) a high energy belt <strong>of</strong> skeletal/<br />

oolitic pack- to grainstones; (d) shallow-ramp sediments<br />

dominated by skeletal storm sheets; <strong>and</strong> (e) a<br />

deeper ramp characterized by mud- <strong>and</strong> marlstones.<br />

The facies belts are modified by subtle structural<br />

trends that define subtle palaeo-highs <strong>and</strong> palaeolows,<br />

which are well known in the regional geology<br />

<strong>and</strong> can also be recognized by thickness variations<br />

(isopach maps) in various stratigraphic levels (e.g.<br />

Geyer & Gwinner 1991). These are probably reactivated<br />

basement blocks, inducing slight variations in<br />

differential subsidence.<br />

Study area, methods <strong>and</strong> database<br />

The SW German Hohenlohe area provides numerous<br />

quarries <strong>and</strong> natural outcrops in the Upper<br />

Muschelkalk (Fig. 2). The shoal water <strong>carbonate</strong>s<br />

represent a major target for the raw-material industry<br />

in this region. Therefore, the closest spacing <strong>of</strong><br />

quarries can be observed in areas where the clean<br />

<strong>and</strong> porous skeletal/oolitic grainstone bodies show<br />

their thickest development. Furthermore, natural<br />

exposures along river gorges allow a lateral tracing<br />

<strong>of</strong> individual bodies <strong>and</strong> property transitions.<br />

The database consists <strong>of</strong> 50 measured outcrop<br />

sections, supplemented by six cores <strong>and</strong> wireline<br />

logs (1787 m). Thirty-one <strong>of</strong> these sections (711 m)<br />

were integrated from previous work (Aigner 1985;<br />

Kostic 2001; Ruf 2001; Braun 2003; Allgöwer<br />

2006; Dmitrieva 2006; Looser 2006; Seyfang<br />

2006). For the investigation <strong>of</strong> diagenetic <strong>and</strong> <strong>petrophysical</strong><br />

<strong>properties</strong> 568 cylinder-shaped plugs were<br />

extracted from rock samples <strong>and</strong> cores. The plugging<br />

focused on potential reservoir rocks, areas<br />

with visual porosity or dolomite content. The<br />

plugs have a diameter <strong>of</strong> four centimetres <strong>and</strong><br />

Fig. 2. Composite map <strong>of</strong> the study area indicating the data points <strong>and</strong> the thickness <strong>of</strong> the Upper Muschelkalk;<br />

modified after Palermo et al. (2010).


Geological Society, London, Special Publications published online June 27, 2012 as doi: 10.1144/SP370.6<br />

D. PALERMO ET AL.<br />

were cut to a depth between 4.5 cm <strong>and</strong> 3 cm.<br />

Additionally, four rock columns (1 m × 0.2 m ×<br />

0.2 m) were derived from a block (2.1 m ×<br />

1.1 m × 1.0 m) in order to investigate the distribution<br />

<strong>of</strong> permeability within an individual reservoir<br />

body on different scales.<br />

Porosity <strong>and</strong> density measurements were carried<br />

out with two devices in two separate steps:<br />

(a)<br />

(b)<br />

The determination <strong>of</strong> net volume (netVol) <strong>and</strong><br />

net density (without pores) was carried out with<br />

a calibrated helium pycnometer.<br />

Gross volume (grossVol) <strong>of</strong> the sample was<br />

measured with a powder pycnometer. This<br />

device calculates the rock density <strong>and</strong> the percentage<br />

porosity (F) with the formula:<br />

F = (grossVol − netVol) × 100/grossVol.<br />

Permeability was measured with a gas minipermeameter.<br />

Eight measurements were carried<br />

out on every sample, to provide the mean horizontal<br />

<strong>and</strong> the mean vertical permeability. The device is<br />

controlled by special automatic s<strong>of</strong>tware, which<br />

adjusts the contact pressure, regulates the flow rate<br />

<strong>and</strong> calculates the permeability from the gas-flow<br />

rate. The measuring error <strong>of</strong> the mini-permeameter<br />

is around 0.5 mD.<br />

For the reconstruction <strong>of</strong> the diagenetic history<br />

<strong>and</strong> the deeper underst<strong>and</strong>ing <strong>of</strong> reservoir <strong>properties</strong>,<br />

451 stained thin sections were investigated.<br />

For a systematic semi-quantitative analysis <strong>of</strong><br />

several parameters (e.g. amount <strong>of</strong> components,<br />

mud content, cement generations, pore types) the<br />

comparison charts <strong>of</strong> Flügel (2004) were applied.<br />

The petrography <strong>and</strong> cement-types were investigated<br />

with transmission light <strong>and</strong> cathodoluminescence<br />

microscopy. The pore types were categorized<br />

<strong>and</strong> analysed by using mainly the concepts <strong>of</strong><br />

Choquette <strong>and</strong> Pray (1970) <strong>and</strong> Lucia (1983). In<br />

order to quantify the results with the porosity <strong>and</strong><br />

permeability measurements, the thin sections were<br />

commonly derived from the trim ends <strong>of</strong> the plugs.<br />

The 3D <strong>petrophysical</strong> <strong>modelling</strong> was carried out<br />

using industry-st<strong>and</strong>ard s<strong>of</strong>tware within a high resolution<br />

3D facies model described in Palermo<br />

et al. (2010).<br />

Facies <strong>and</strong> diagenesis<br />

Facies types<br />

Based on lithology, Dunham texture, particle size,<br />

sorting, sedimentary structures <strong>and</strong> environmentindicative<br />

allochems such as for example ooids,<br />

oncoids <strong>and</strong> black pebbles 16 different lith<strong>of</strong>acies<br />

types were distinguished in the Upper Muschelkalk<br />

<strong>of</strong> the study area. These have been described in<br />

Palermo et al. (2010) <strong>and</strong> thus do not need to be<br />

further discussed here. The facies types are interpreted<br />

to be deposited within a shallow epeiric <strong>carbonate</strong><br />

ramp setting, characterized by a very gentle<br />

depositional gradient (between 0.0028 to 0.38).<br />

The <strong>carbonate</strong> ramp can be subdivided into three<br />

major subenvironments (cf. Burchette et al. 1990):<br />

(1) low energy inner ramp or backshoal; (2) highenergy<br />

mid-ramp with <strong>carbonate</strong> s<strong>and</strong>s in shoreline<br />

detached shoals; <strong>and</strong> (3) low energy foreshoal<br />

<strong>and</strong> outer ramp with mud-dominated successions<br />

(Fig. 3). Figure 4 shows a generalized depositional<br />

model with the main reservoir facies types <strong>and</strong><br />

their petrographic composition.<br />

Diagenetic history<br />

The diagenetic history <strong>of</strong> the Upper Muschelkalk<br />

grainstones has been subject to several previous<br />

studies (e.g. Bachmann 1973; Kostic 2001; Braun<br />

2003; Seyfang 2006). Bachmann (1973) differentiated<br />

two major diagenetic stages with several<br />

cementation <strong>and</strong> leaching phases:<br />

(1) Within the early diagenetic stage under<br />

shallow burial, isopachous circum-granular<br />

crust cement (A1 cement; Fig. 5.1) consolidated<br />

the original sedimentary grains. This<br />

process was followed by leaching <strong>of</strong> the<br />

aragonitic components <strong>and</strong> the precipitation<br />

<strong>of</strong> secondary isopachous crust cement (A2<br />

cement; Fig. 5.2) <strong>of</strong> similar appearance,<br />

grown exclusively on the fringes inside the<br />

dissolved pores.<br />

(2) The late diagenetic stage was refined by observations<br />

<strong>of</strong> Braun (2003), who recognized<br />

an additional phase <strong>of</strong> late diagenetic dog<br />

tooth cement (B1 cement) grown in irregular<br />

rims <strong>of</strong> isopachous prismatic spar, which<br />

was followed by the precipitation <strong>of</strong> equant<br />

drusy calcite-spar (B2 cement; Fig. 5.3).<br />

Finally, a minor amount <strong>of</strong> scattered dolomite<br />

cement precipitated in remnant separate vug<br />

pore space.<br />

The present investigation <strong>of</strong> 442 thin sections<br />

from different locations <strong>and</strong> stratigraphic intervals,<br />

using st<strong>and</strong>ard petrographic, cathodoluminescence<br />

<strong>and</strong> fluorescence microscopy, resulted in slight revisions<br />

<strong>of</strong> some aspects <strong>of</strong> the so far established late<br />

diagenetic history (Table 1).<br />

Discussion: controlling factors <strong>of</strong> diagenesis<br />

Facies. Facies types <strong>and</strong> their primary composition<br />

seem to have a major impact on diagenesis<br />

<strong>and</strong> reservoir development since both leaching<br />

<strong>and</strong> dolomitization are selective with respect to<br />

the mineralogy <strong>of</strong> components. For instance, the


Geological Society, London, Special Publications published online June 27, 2012 as doi: 10.1144/SP370.6<br />

CARBONATE SAND BODIES IN GERMANY<br />

Fig. 3. Facies types with key attributes, depositional environment <strong>and</strong> colour code.<br />

semi-quantitative analysis <strong>of</strong> 442 thin sections<br />

(Figs 5 & 6) shows a negative correlation between<br />

mud content <strong>and</strong> the intensity <strong>of</strong> early diagenetic<br />

A-type cements (correlation coefficient, 20.7; st<strong>and</strong>ard<br />

deviation, 0.19). However, apart from the<br />

depositional energy <strong>and</strong> mud content, the growth<br />

intensity <strong>of</strong> A1 cements is also controlled by the<br />

type <strong>of</strong> associated components. For example, A1<br />

cements are <strong>of</strong>ten pore filling in ooid dominated<br />

grainstones, while they from only thin seams<br />

around crinoid ossicles. In general, however, the<br />

relationship between primary facies <strong>and</strong> diagenesis<br />

is complex (see below).<br />

Stratigraphy. The systematic semi-quantitative<br />

investigation <strong>of</strong> thin sections in their stratigraphic<br />

context showed that diagenetic alteration<br />

is also controlled by stratigraphic cycles. This is


Geological Society, London, Special Publications published online June 27, 2012 as doi: 10.1144/SP370.6<br />

D. PALERMO ET AL.<br />

Fig. 4. Facies distribution <strong>and</strong> petrographic composition, based on the semi-quantitative analysis <strong>of</strong> 442 thin sections.<br />

particularly important for the prediction <strong>of</strong> systematic<br />

changes <strong>of</strong> reservoir <strong>properties</strong> within individual<br />

facies bodies. An example for these variations<br />

within the reservoir facies <strong>of</strong> one regressive<br />

medium-scale hemicycle is depicted in Figure 6.<br />

In particular, the mud content <strong>and</strong> primary A<br />

cements change systematically within the cycle.<br />

Generally, the sum <strong>of</strong> mud content <strong>and</strong> A cements<br />

constitutes a more or less constant petrographic<br />

volume fraction together (arithmetic mean <strong>of</strong> mud<br />

content <strong>and</strong> A cement, 40.9 Vol%; st<strong>and</strong>ard deviation,<br />

17.3; n ¼ 442 thin sections). As displayed<br />

in the column ‘Cumulative Vol %’ <strong>of</strong> Figure 6,<br />

upwards decreasing mud content is compensated<br />

by a systematic increase in B-type cements. Both<br />

A-type <strong>and</strong> B-type cement combined show a clear<br />

upward increasing fracture towards the mediumscale<br />

regressive maximum.<br />

Maximum values <strong>of</strong> interparticle <strong>and</strong> mouldic<br />

porosity are commonly located around the regressive<br />

maximums <strong>of</strong> the medium-scale cycles in shoal<br />

facies associations, following an upward decreasing<br />

fraction <strong>of</strong> matrix mud (Fig. 7b). However, intense<br />

late diagenetic B cements can partially plug the<br />

pore space in the upper portion the regressive hemi<br />

cycles, shifting the maximum porosity <strong>and</strong> permeability<br />

values towards the middle part (Fig. 6).<br />

The upward decreasing mud content reflects the<br />

vertically increasing depositional energy <strong>of</strong> a prograding<br />

shoal complex <strong>and</strong> seems to be a major<br />

controlling factor in the reservoir <strong>properties</strong>. The<br />

<strong>carbonate</strong> mud in the lower portion <strong>of</strong> the regressive


Geological Society, London, Special Publications published online June 27, 2012 as doi: 10.1144/SP370.6<br />

CARBONATE SAND BODIES IN GERMANY<br />

Fig. 5. Thin section photographs. (1) Thin section Crailsheim, parallel Nichols, stained: blue indicates porosity.<br />

Pel-oolitic grainstone: the peloids are surrounded by A1 cement (A1) forming typical circum-granular crusts <strong>of</strong> fibrous<br />

to bladed calcite crystals. Note the preserved primary interparticle porosity between the grains (Ip). The black opaque<br />

inclusions within the peloids are probably pyrite crystals (Py). (2) Thin section Sattelorf-Bölgental, parallel Nichols,<br />

stained: blue indicates porosity. Crinoidal grainstone: shape-preserving micritic envelope <strong>of</strong> a dissolved aragonitic<br />

bioclast (M). A1 cement (A1) forms a more regular seam <strong>of</strong> fibrous to bladed crystals compared with the more irregular<br />

A2 cement (A2) grown inside the biomouldic pore (Bm). (3) Thin-section Steinbächle, parallel Nichols. Bioclastic<br />

grainstone: a mosaic <strong>of</strong> drusy, equant B2 cement (B2) overgrows the earlier generation <strong>of</strong> B1-‘dog tooth’ cement (B1).<br />

(4) Thin-section Crailsheim, parallel Nichols, stained: blue indicates porosity. Oolitic grainstone: the dolomitization <strong>of</strong><br />

ooids <strong>and</strong> coated grains occurs preferentially by replacing the cortical layers with a sub- to anhedral crystal mosaic (D).<br />

Note the undisturbed primary A1 cement (A1) surrounding the grains. Note the preserved primary porosity (Ip). (5a, b)<br />

Thin-section Steinbächle, parallel Nichols (1); crossed Nichols (2), stained: blue indicates porosity. Oolitic grainstone:<br />

patchy saddle dolomite cement with typical curved crystal faces (S) <strong>and</strong> sweeping extinction (E). (6) Thin section<br />

Satteldorf Bölgental, parallel Nichols, stained: blue indicates porosity. Crinoidal grainstone: limonitic crusts <strong>of</strong> leached<br />

saddle dolomites (LC), <strong>and</strong> partly leached dolomite crystals (LD) with rusty limonitic fractures. Remaining A1 cement<br />

has grown in the mould <strong>of</strong> an early diagenetic leached possibly aragonitic coated grain-core (A1). The cortical layers<br />

were dolomitized <strong>and</strong> subsequently leached during the second leaching phase (L2).<br />

hemicycles reduced the primary porosity <strong>and</strong> the<br />

circulation <strong>of</strong> diagenetic fluids, which enhanced<br />

the porosity in later diagenetic stages. However, in<br />

cases with intense B cementation at the top <strong>of</strong> the<br />

regressive hemicycles, the best reservoir <strong>properties</strong><br />

tend to be located below the regressive maxima<br />

(Fig. 6).<br />

<strong>Reservoir</strong> <strong>properties</strong><br />

In order to quantify the reservoir <strong>properties</strong>, 442 thin<br />

sections <strong>and</strong> 570 porosity <strong>and</strong> permeability plugs<br />

from outcrop samples <strong>and</strong> cores were analysed.<br />

Porous intervals are generally restricted to discrete<br />

reservoir bodies, composed <strong>of</strong> high-energy shoal<br />

<strong>and</strong> shoal fringe facies types. However, the internal<br />

<strong>petrophysical</strong> characterization <strong>of</strong> these reservoir<br />

bodies turned out to be complex. During diagenesis,<br />

several cementation <strong>and</strong> leaching phases led to a<br />

stepwise alteration <strong>of</strong> the pore-space, which resulted<br />

in a complex relationship between reservoir <strong>properties</strong><br />

<strong>and</strong> reservoir-facies types. Nevertheless,<br />

owing to the selective character <strong>of</strong> the diagenetic<br />

processes, the initial sediment composition <strong>and</strong>


Geological Society, London, Special Publications published online June 27, 2012 as doi: 10.1144/SP370.6<br />

D. PALERMO ET AL.<br />

Table 1. Overview <strong>of</strong> the diagenetic history <strong>and</strong> the impact on pore-space<br />

Diagenetic stage Diagenetic phase Impact on pore space<br />

Early marine, shallow<br />

Circum granular crust Protection <strong>and</strong> reduction <strong>of</strong> primary porosity<br />

burial diagenesis<br />

cement (A1 cement)<br />

First leaching phase Creation <strong>of</strong> biomouldic porosity owing to<br />

selective leaching <strong>of</strong> aragonite<br />

Circum granular crust<br />

cement (A2 cement)<br />

Slight reduction <strong>of</strong> complete plugging<br />

<strong>of</strong> biomouldic porosity<br />

Late marine, deeper<br />

burial cementation<br />

Dog tooth cement<br />

(B1 cement)<br />

Slight reduction <strong>of</strong> complete plugging<br />

<strong>of</strong> biomouldic <strong>and</strong> primary porosity<br />

Patchy calcite spar<br />

cement (B2 cement)<br />

Slight reduction <strong>of</strong> complete plugging<br />

<strong>of</strong> biomouldic <strong>and</strong> primary porosity<br />

Late diagenetic, deep<br />

burial dolomitization<br />

Patchy dolomitization<br />

<strong>and</strong> recrystallization<br />

Slight reduction <strong>of</strong> complete plugging<br />

<strong>of</strong> biomouldic <strong>and</strong> primary porosity,<br />

dolomitization <strong>of</strong> ooids, coated grains<br />

<strong>and</strong> crinoidal columnar plates<br />

Uplift <strong>and</strong> meteoric alteration Second leaching phase Creation <strong>of</strong> Oo- <strong>and</strong> crino-mouldic porosity<br />

mineralogy remains one <strong>of</strong> the most important controlling<br />

factors. The investigated Upper Muschelkalk<br />

reservoir bodies have average reservoir<br />

<strong>properties</strong> <strong>of</strong> 11.5% porosity (st<strong>and</strong>ard deviation,<br />

4.9) <strong>and</strong> 30.8 mD permeability (st<strong>and</strong>ard deviation<br />

72.4) with a correlation coefficient <strong>of</strong> 0.41.<br />

Pore-types<br />

Lucia (1983) developed commonly used concepts<br />

for reservoir characterization <strong>of</strong> <strong>carbonate</strong> rocks by<br />

subdividing the pore-space into the three major poretype<br />

classes: (a) touching vug porosity; (b) separate<br />

vug porosity; <strong>and</strong> (c) interparticle porosity. The<br />

Upper Muschelkalk is dominated by separate vug<br />

porosity (SV, 60%). Touching vug (TV) <strong>and</strong> interparticle<br />

porosity (IP, 20%) constitute the minor<br />

part, but are important factors controlling permeability.<br />

However, the touching vugs observed in<br />

the Upper Muschelkalk are smaller (up to several<br />

millimetres; Fig. 8) than those described by Lucia<br />

(1983).<br />

Using the fabric-selective concept for the classification<br />

<strong>of</strong> <strong>carbonate</strong> pores after Choquette <strong>and</strong> Pray<br />

Fig. 6. Diagenesis <strong>and</strong> the resulting petrography are controlled by stratigraphic cycles. Generally, mud content <strong>and</strong><br />

A cements constitute a more or less constant petrographic volume fraction, as displayed in the column ‘Cumulative<br />

Vol %’; upwards decreasing mud content is compensated by a systematic increase <strong>of</strong> A cement.


Geological Society, London, Special Publications published online June 27, 2012 as doi: 10.1144/SP370.6<br />

CARBONATE SAND BODIES IN GERMANY<br />

Fig. 7. (a) Cross section through two reservoir bodies (yellow/orange) along the depositional gradient, highlighting<br />

facies distribution, sedimentology <strong>and</strong> reservoir <strong>properties</strong>. Within the individual facies bodies, the lateral porosity <strong>and</strong><br />

permeability distribution is mostly marked by gradational transitions. (b) Cross section highlighting the qualitative<br />

distribution <strong>of</strong> cement types <strong>and</strong> mud content. The vertical sections depict the cumulative petrographic composition<br />

<strong>of</strong> the shoal bodies, combined with <strong>petrophysical</strong> <strong>properties</strong>. Both <strong>properties</strong> show laterally gradational transitions <strong>and</strong><br />

are strongly influenced vertically by stratigraphic cycles.<br />

(1970), biomouldic porosity comprises the largest<br />

pore-fraction. Most <strong>of</strong> the investigated samples<br />

belong to one <strong>of</strong> the following combined pore-types:<br />

(1) biomouldic (BM, 27%);<br />

(2) biomouldic <strong>and</strong> interparticle (BM + IP, 42%);<br />

(3) biomouldic, interparticle <strong>and</strong> Oomouldic<br />

(BM + IP + OM, 23%);<br />

(4) biomouldic, <strong>and</strong> oomouldic (BM + IP + OM,<br />

8%).<br />

Porosity <strong>and</strong> permeability relationships<br />

Pore types after Choquette <strong>and</strong> Pray (1970) v.<br />

porosity <strong>and</strong> permeability (Fig. 9a). Generally, the<br />

samples with pure mouldic porosity show less


Geological Society, London, Special Publications published online June 27, 2012 as doi: 10.1144/SP370.6<br />

D. PALERMO ET AL.<br />

Fig. 8. Thin section photographs, stained: blue indicates porosity, parallel Nichols: typical pore-types in the Upper<br />

Muschelkalk reservoir bodies. Pore-types after Lucia (1983): touching vugs, TV; separate vugs, SV; interparticle<br />

porosity, IP. Pore-types after Choquette & Pray (1970): biomouldic porosity, BM; oomouldic porosity, OM. (a) Oolitic<br />

grainstone with preserved primary interparticle porosity. (b) Bioclastic grainstone showing two pore generations <strong>of</strong><br />

biomouldic porosity. Upper left corner: leached dolomite recrystallization. Middle right corner: early diagenetic<br />

biomouldic pore with several cement generations. (c) Bioclastic pack- to grainstone with touching vugs owing to<br />

leached dolo-cement. The dissolution enlarged touching vugs <strong>of</strong> the Upper Muschelkalk reservoir bodies are much<br />

smaller (several millimetres) than those described by Lucia (1983). Furthermore, remnants <strong>of</strong> collapsed biomouldic<br />

porosity are visible in the upper middle part. (d) Oolitic grainstone. Typical oomouldic separate vug porosity, enhanced<br />

by a large touching vug <strong>of</strong> leached dolo-cement <strong>and</strong> small amounts <strong>of</strong> interparticle porosity. (e) Bioclastic grainstone:<br />

oomouldic porosity <strong>of</strong> leached dolomite recrystallizations combined with early diagenetic biomouldic separate vug<br />

porosity <strong>and</strong> large interparticle pores. (f) Bioclastic grainstone: early diagenetic biomouldic separate vug porosity.<br />

permeability than samples where both, mouldic- <strong>and</strong><br />

interparticle porosity are combined. Regarding the<br />

pore type distribution within the Lucia classes,<br />

samples with biomouldic, oomouldic <strong>and</strong> interparticle<br />

porosity tend to class 2, while samples with biomouldic<br />

<strong>and</strong> interparticle porosity plot slightly more<br />

towards Lucia’s class 1.<br />

<strong>Reservoir</strong> rock types <strong>and</strong> classes using Lucia (1983)<br />

classification (Fig. 9b). Several studies on the<br />

Upper Muschelkalk <strong>carbonate</strong>s have shown that<br />

porosity–permeability relationships show strong<br />

scattering (e.g. Braun 2003; Kostic & Aigner<br />

2004; Ruf & Aigner 2004; Dmitrieva 2006;<br />

Seyfang 2006). However, Lucia (1999) was able<br />

to categorize porosity–permeability relationships<br />

in discrete <strong>petrophysical</strong> classes by the definition<br />

<strong>of</strong> rock-fabric types, which are a combination <strong>of</strong><br />

both Dunham texture <strong>and</strong> pore-type. His rockfabric<br />

<strong>petrophysical</strong> classes were mainly developed<br />

for non-vuggy <strong>carbonate</strong>s with interparticle porosity.<br />

As the fraction <strong>of</strong> vuggy porosity within the<br />

investigated reservoir bodies amounted to 80%,<br />

the Lucia classes could only be applied to a<br />

certain extent. Most <strong>of</strong> the porosity <strong>and</strong> permeability<br />

values plot in class 2 <strong>of</strong> Lucia, followed by<br />

class 1. The values above 3 mD permeability<br />

show a relatively linear trend within a certain<br />

range <strong>of</strong> scatter. The values plot relatively close<br />

together between Lucia classes 1 <strong>and</strong> 2. As


Geological Society, London, Special Publications published online June 27, 2012 as doi: 10.1144/SP370.6<br />

CARBONATE SAND BODIES IN GERMANY<br />

Fig. 9. Porosity <strong>and</strong> permeability relationships keyed to (a) pore type according to (a) Choquette & Pray (1970),<br />

(b) Lucia (1983), (c) approximated pore-throat size measured in thin sections <strong>and</strong> (d) pore size.<br />

expected, samples with only separate vugs have<br />

generally lower permeability. When touching-vug<br />

porosity is present, permeability is higher. The<br />

highest porosity <strong>and</strong> permeability values occur<br />

when all pore types are present (Table 2).<br />

However, biomouldic pores compose the largest<br />

part <strong>of</strong> porosity <strong>and</strong> occur, independent <strong>of</strong> Dunham<br />

texture, in all reservoir rock types. This fact implies<br />

that the distribution <strong>of</strong> biomouldic porosity could<br />

be an additional key controlling factor for the<br />

reservoir <strong>properties</strong>.<br />

Pore-throat size v. poro-perm (Fig. 9c). The porethroat<br />

diameters presented in this study have been<br />

measured in 2D thin sections <strong>and</strong> represent therefore<br />

just a limited approximation to the real 3D<br />

pore-throat distribution. Nevertheless, Figure 9c<br />

depicts a clear increase <strong>of</strong> both permeability <strong>and</strong><br />

porosity with larger pore-throats (Table 4). The<br />

remaining data scatter could be due to the fact that<br />

pore-throat diameters were measured in 2D.<br />

Pore size v. porosity <strong>and</strong> permeability (Fig. 9d).<br />

Despite considerable scattering, permeability generally<br />

increases with an increasing pore-size (Table 2).<br />

Relationship facies association: reservoir<br />

<strong>properties</strong><br />

As can be observed in many <strong>carbonate</strong> systems,<br />

facies <strong>and</strong> reservoir <strong>properties</strong> <strong>of</strong> the Upper<br />

Muschelkalk do not show a direct relationship.<br />

This is mainly due to the susceptibility <strong>of</strong> <strong>carbonate</strong>s<br />

to diagenetic alteration, but also to the higher<br />

complexity <strong>of</strong> primary sedimentary parameters in<br />

addition to measuring inaccuracies. Nevertheless,<br />

the primary lith<strong>of</strong>acies plays a fundamental role<br />

for the distribution <strong>of</strong> reservoir <strong>properties</strong> in the<br />

Upper Muschelkalk, since porous shoal bodies are<br />

exclusively formed <strong>of</strong> high energy shoal <strong>and</strong> shoalfringe<br />

facies types. Furthermore, the diagenetic<br />

leaching <strong>and</strong> dolomite cementation phases proceeded<br />

very selectively according to the mineralogy<br />

<strong>of</strong> the components, a parameter that is in turn mainly<br />

facies controlled.<br />

Therefore, despite considerable data scattering, a<br />

certain relationship between facies association <strong>and</strong><br />

reservoir <strong>properties</strong> can be observed.<br />

Shoal-fringe facies association (pel-ooidal packto<br />

grainstone, oncoidal pack- to grainstone, intraclastic<br />

packstone). The facies types forming the<br />

shoal-fringe facies association show a strong scattering,<br />

in particular within the range below 3 mD<br />

permeability <strong>and</strong> 3% porosity (Fig. 10a, Table 3). A<br />

possible explanation for the strong data scatter<br />

could be a scale-dependent measuring inaccuracy,<br />

which occurs specifically with the mini-permeameter.<br />

Vugs <strong>and</strong> large bioclastic moulds may act as<br />

direct flow-connection through the plug-sample.<br />

Intraclastic packstones have low porosities, predominantly<br />

in the form <strong>of</strong> large vugs. This facies


Geological Society, London, Special Publications published online June 27, 2012 as doi: 10.1144/SP370.6<br />

D. PALERMO ET AL.<br />

Table 2. Correlation coefficients, st<strong>and</strong>ard deviation <strong>and</strong> arithmetic mean between porosity/permeability <strong>and</strong><br />

selected parameters (pore size, pore throat size <strong>and</strong> pore types), based on the semi-quantitative analysis <strong>of</strong><br />

thin sections<br />

Correlation coefficient<br />

Poro (%) Micro-pores (%)<br />

Poro/micro Mean 12.0 2.2 20.07<br />

n ¼ 222 St<strong>and</strong>ard deviation 4.6 1.8<br />

Perm H Meso-pores (%)<br />

Perm H/meso Mean 32.5 6.5 0.17<br />

n ¼ 222 St<strong>and</strong>ard deviation 58.1 3.8<br />

Perm H Macro-pores (%)<br />

Perm H/macro Mean 32.5 3.2 0.55<br />

n ¼ 222 St<strong>and</strong>ard deviation 58.1 3.8<br />

Poro (%) TV (%)<br />

Poro/TV Mean 12.0 2.6 0.63<br />

n ¼ 222 St<strong>and</strong>ard deviation 4.6 2.7<br />

Poro (%) SV (%)<br />

Poro/SV Mean 12.0 7.1 0.47<br />

n ¼ 222 St<strong>and</strong>ard deviation 4.6 3.0<br />

Poro (%) IP (%)<br />

Poro/IP Mean 12.0 2.2 0.50<br />

n ¼ 222 St<strong>and</strong>ard deviation 4.6 3.1<br />

Perm H (mD) TV (%)<br />

Perm H/TV Mean 32.5 2.6 0.58<br />

n ¼ 222 St<strong>and</strong>ard deviation 58.1 2.7<br />

Perm H (mD) SV (%)<br />

Perm H/SV Mean 32.5 7.1 0.00<br />

n ¼ 222 St<strong>and</strong>ard deviation 58.1 3.0<br />

Perm H (mD) IP (%)<br />

Perm H/IP Mean 32.5 2.2 0.30<br />

n ¼ 222 St<strong>and</strong>ard deviation 58.1 3.1<br />

Poro (%) Pore-throats (mm)<br />

Poro/pore-throats Mean 12.0 177.8 0.57<br />

n ¼ 168 St<strong>and</strong>ard deviation 4.6 94.6<br />

Perm H (mD) Pore-throats (mm)<br />

Perm H/pore-throats Mean 32.5 177.8 0.52<br />

n ¼ 168 St<strong>and</strong>ard deviation 58.1 94.6<br />

type shows the most extreme scatter combined<br />

with the poorest reservoir <strong>properties</strong>. In contrast,<br />

peloidal- <strong>and</strong> oncoidal pack- to grainstones have<br />

additionally small amounts <strong>of</strong> regularly distributed<br />

primary porosity <strong>and</strong> commonly smaller vugs.<br />

The limited reservoir <strong>properties</strong> (F max ¼ 13%;<br />

K max ¼ 12 mD) <strong>of</strong> the shoal-fringe facies types<br />

could be explained by the combination <strong>of</strong> a high<br />

mud content, occluding the primary interparticle<br />

pore space <strong>and</strong> the dominance <strong>of</strong> mostly calcitic,<br />

relatively leaching-resistant components (e.g. oncoids,<br />

peloids, intraclasts), <strong>and</strong> preventing the creation<br />

<strong>of</strong> vuggy porosity (Fig. 10a).<br />

Mid-shoal facies association (crinoidal pack- to<br />

grainstone, poorly sorted pack- to grainstone, amalgamated<br />

packstone). These facies types constitute<br />

the largest part <strong>of</strong> the reservoir bodies <strong>and</strong> show relatively<br />

similar porosities <strong>and</strong> permeability trends<br />

Within this group, amalgamated packstones have<br />

the best correlation between porosity <strong>and</strong> permeability<br />

(Fig. 10b, Table 4). Amalgamated packstones<br />

are dominated by biomouldic porosity; the high<br />

mud content plugs the potential interparticle porespace.<br />

Therefore, permeability is probably due to the<br />

occurrence <strong>of</strong> solution-enlarged touching vugs. Consequently,<br />

diagenetic leaching is the most important<br />

controlling factors on pore space within this facies<br />

type. In contrast, crinoidal pack- to grainstones have<br />

additionally substantial amounts <strong>of</strong> interparticle porosity.<br />

They show the overall best reservoir <strong>properties</strong><br />

(F max ¼ 25%; k max ¼ 710 mD), but also a strong<br />

scattering <strong>of</strong> data points. The scattering within the<br />

porosity <strong>and</strong> permeability relationship is thus mainly<br />

controlled by the interplay <strong>of</strong> different pore-systems<br />

(interparticle <strong>and</strong> vuggy). Nevertheless, also scaledependent<br />

measuring inaccuracy (see above) can be<br />

observed within the poorly sorted pack- to grainstone<br />

facies, where vugs <strong>of</strong> large shells are common, dispersing<br />

the values towards higher permeability.


Geological Society, London, Special Publications published online June 27, 2012 as doi: 10.1144/SP370.6<br />

CARBONATE SAND BODIES IN GERMANY<br />

Fig. 10. (a) Porosity–permeability cross-plot depicting the reservoir <strong>properties</strong> <strong>of</strong> shoal fringe facies types. Strong<br />

scattering towards low permeability is mainly interpreted as scale-dependent measuring inaccuracy, since vugs <strong>of</strong> large<br />

shells may act as direct flow connection through the plug sample. (b) Porosity–permeability cross-plot depicting the<br />

reservoir <strong>properties</strong> <strong>of</strong> mid-shoal facies types. Strong scattering towards low permeability within the poorly sorted packto<br />

grainstone facies is interpreted as scale-dependent measuring inaccuracy. Amalgamated packstones are dominated by<br />

vuggy porosity <strong>and</strong> show an excellent fit into the main cluster, whereas crinoidal pack- to grainstones tend to a stronger<br />

scattering, but overall better reservoir <strong>properties</strong>. The stronger scattering can be explained by two different interacting<br />

pore-systems owing to a combination <strong>of</strong> interparticle <strong>and</strong> vuggy porosity. (c) Porosity–permeability cross-plot<br />

depicting the reservoir <strong>properties</strong> <strong>of</strong> inner-shoal facies types. Facies types <strong>of</strong> the mid-shoal facies association have a<br />

higher fraction <strong>of</strong> oomouldic porosity <strong>and</strong> fewer solution enlarged touching vugs, which results in a scattering towards<br />

higher porosities.<br />

Inner-shoal facies association (oolitic grainstone,<br />

well-sorted pack- to grainstone). These facies<br />

types, in particular oolitic grainstones, show a<br />

tendency towards higher porosities (Fig. 10c,<br />

Table 5) <strong>and</strong> comparably low permeability. Vugs<br />

<strong>and</strong> larger bioclastic moulds which interconnect


Geological Society, London, Special Publications published online June 27, 2012 as doi: 10.1144/SP370.6<br />

D. PALERMO ET AL.<br />

Table 3. Correlation coefficients, arithmetic mean <strong>and</strong> st<strong>and</strong>ard deviations <strong>of</strong> shoal-fringe facies types<br />

Facies type Porosity Permeability Correlation coefficient<br />

Oncoidal WP Mean 7.6 4.4 0.34<br />

n ¼ 14 St<strong>and</strong>ard deviation 3.8 6.6<br />

Pelo-ooidal P Mean 5.5 7.4 0.14<br />

n ¼ 7 St<strong>and</strong>ard deviation 3.1 11.1<br />

Intraclastic P Mean 3.9 0.6 20.25<br />

n ¼ 12 St<strong>and</strong>ard deviation 2.9 0.8<br />

the pores are rare in this grain-dominated, commonly<br />

well sorted facies association. Thus, this<br />

facies association is characterized by relatively<br />

high amounts <strong>of</strong> oomouldic porosity, which is,<br />

however, not accompanied by higher permeability<br />

values. If biomouldic pores are present, they are<br />

commonly surround by thick cement crusts. These<br />

cements are commonly not affected by selective<br />

diagenetic alteration <strong>and</strong> are an additional reason<br />

for reduced pore connectivity. As observed in thin<br />

sections, solution-enlarged touching vug porosity<br />

is present but not very abundant, which could be<br />

the result <strong>of</strong> the more stable pore-framework preventing<br />

the circulation <strong>of</strong> late diagenetic fluids <strong>and</strong><br />

associated porosity creation.<br />

Distribution <strong>of</strong> reservoir <strong>properties</strong><br />

on different scales<br />

The final aim <strong>of</strong> a reservoir characterization is to<br />

predict the spatial distribution <strong>of</strong> reservoir <strong>properties</strong><br />

on a field scale. In this outcrop study, the geometries<br />

<strong>of</strong> the facies bodies show a remarkable lateral<br />

continuity on a kilometre-scale (for details see<br />

Palermo et al. 2010). In contrast to one-dimensional<br />

well data, the outcrops have the advantage that the<br />

individual facies bodies <strong>and</strong> the presence <strong>of</strong> porosity<br />

can be traced laterally in a qualitative way, since<br />

millimetre- to centimetre-sized mouldic macro<strong>and</strong><br />

meso-pores can be observed easily with a<br />

h<strong>and</strong> lens. In order to calibrate these observations,<br />

plug measurements <strong>and</strong> thin sections were taken<br />

along vertical sections. Different scales <strong>of</strong> reservoir<br />

heterogeneities in one selected shoal body were<br />

investigated in a hierarchical way (for details see<br />

Seyfang 2006).<br />

Decimetre scale (tens <strong>of</strong> centimetres)<br />

In order to investigate the distribution <strong>of</strong> permeability<br />

on a decimetre scale, four rock columns<br />

from the central part <strong>of</strong> a major shoal body were<br />

cut from a 2.1 × 1.1 m-sized block from the<br />

quarry Neidenfels. The investigated shoal portion<br />

consists <strong>of</strong> the poorly sorted pack- to grainstone<br />

facies <strong>and</strong> shows a variable mud content. The twodimensional<br />

permeability distribution was measured<br />

with a mini-permeameter by covering each<br />

rock column (1 × 0.2 m) with a grid <strong>of</strong> 7 × 50<br />

data points spaced at 2 cm intervals (Fig. 11). The<br />

porosity values <strong>of</strong> the investigated samples range<br />

between 15 <strong>and</strong> 23%. The values were interpolated<br />

with the s<strong>of</strong>tware Surfer TM using a kriging algorithm<br />

<strong>and</strong> a logarithmic colour scheme. Sedimentary<br />

structures are not recognizable at this scale (Fig.<br />

12) <strong>and</strong> the facies appeared as a moderately bioturbated,<br />

massive unit; therefore the interpolation<br />

was carried out without preferential direction. The<br />

resulting permeability is rather patchily distributed,<br />

which can be partly related to unidirectional algorithm.<br />

Nevertheless, the values show a clear upward<br />

increasing trend in all four rock columns. The values<br />

follow a small-scale regressive hemicycle with an<br />

upward decreasing fraction <strong>of</strong> matrix mud. The permeability<br />

ranges in the order <strong>of</strong> 10 mD (green) in the<br />

Table 4. Correlation coefficients, arithmetic mean <strong>and</strong> st<strong>and</strong>ard deviations <strong>of</strong> mid-shoal facies types<br />

Facies type Porosity Permeability Correlation coefficient<br />

Crinoidal PG Mean 14.2 63.2 0.26<br />

n ¼ 55 St<strong>and</strong>ard deviation 4.0 123.2<br />

Poorly sorted PG Mean 10.5 28.2 0.51<br />

n ¼ 103 St<strong>and</strong>ard deviation 5.1 56.0<br />

Amalgamated P Mean 11.8 38.7 0.60<br />

n ¼ 63 St<strong>and</strong>ard deviation 3.4 77.7


Geological Society, London, Special Publications published online June 27, 2012 as doi: 10.1144/SP370.6<br />

CARBONATE SAND BODIES IN GERMANY<br />

Table 5. Correlation coefficients, arithmetic mean <strong>and</strong> st<strong>and</strong>ard deviations <strong>of</strong> inner-shoal facies types<br />

Facies type Porosity Permeability Correlation coefficient<br />

Oolitic G Mean 12.9 7.9 0.45<br />

n ¼ 33 St<strong>and</strong>ard deviation 3.7 13.2<br />

Well sorted PG Mean 13.1 25.9 0.42<br />

n ¼ 38 St<strong>and</strong>ard deviation 4.7 49.6<br />

lower part <strong>and</strong> reaches up to several 100 mD (orange<br />

to red) in the top part (Fig. 12). Thin sections document<br />

that this trend is combined with an upwards<br />

increase in interparticle porosity.<br />

Metre scale<br />

For the determination <strong>of</strong> the metre-scale heterogeneities<br />

in permeability, a 2.1 × 1.1 m-sized block from<br />

the quarry Neidenfels was investigated for sedimentary<br />

structures. The positions <strong>of</strong> the above-described<br />

rock columns within the block are marked in<br />

Figure 13. The individual rock columns appeared as<br />

massive units, whereas the entire block reveals<br />

some subtle, irregular low-angle cross-beds. They<br />

consist <strong>of</strong> crudely graded <strong>and</strong> shell-rich laminae<br />

above an erosive base <strong>and</strong> are intercalated into moderately<br />

bioturbated poorly sorted pack- to grainstones.<br />

Generally, slight differences in texture, sorting <strong>and</strong><br />

mud content seem to follow the orientation <strong>of</strong> the<br />

cross bedding on this larger scale. The photograph<br />

<strong>of</strong> the block (Fig. 13) depicts the predominant inclination<br />

<strong>of</strong> the sets to the left, traced with yellow lines.<br />

The heterogeneities in permeability are mainly<br />

due to variations in the amount <strong>of</strong> mouldic porosity<br />

<strong>and</strong> the shape <strong>of</strong> the leached components. Portions<br />

with leached shells make a better contribution<br />

to permeability than those with ooids <strong>and</strong> crinoidal<br />

columnar plates. The shells are <strong>of</strong>ten connected by<br />

solution-enlarged touching vugs <strong>and</strong> build an interconnected<br />

framework with primary interparticle<br />

pores. Based on the observed sedimentary structures,<br />

the interpolation <strong>of</strong> permeability values<br />

between the four measured columns was carried<br />

out with a directional kriging algorithm, characterized<br />

by an elevated horizontal range. The resulting<br />

permeability distribution is displayed as an overlay<br />

with logarithmic colour gradient (Fig. 13). An<br />

important result is the systematic upward increase<br />

in permeability that follows a small-scale regressive<br />

hemicycle <strong>and</strong> the upward decreasing mud content.<br />

This systematic vertical trend results in a laterally<br />

continuous highly permeability zone in the upper<br />

portion <strong>of</strong> the block. Note also that the lowpermeability<br />

lower part is laterally continuous, as<br />

well as some high permeability streaks in the<br />

middle part, which seem to be controlled by the<br />

sedimentary structures. The permeability distribution<br />

shows that the highly permeable areas are<br />

<strong>of</strong>ten vertically interconnected.<br />

Tens <strong>of</strong> metres scale. For the determination <strong>of</strong> porosity<br />

<strong>and</strong> permeability trends within the shoal body on<br />

a tens <strong>of</strong> metre scale, two vertical sections were<br />

investigated along a 25 m-long outcrop wall that<br />

was previously mapped for facies. The outcrop<br />

wall panel, log positions <strong>and</strong> the corresponding<br />

porosity <strong>and</strong> permeability histograms are depicted<br />

in Figure 14. The reservoir body can be subdivided<br />

into four layers (A–D) consisting <strong>of</strong> alternating<br />

well-sorted pack- to grainstone (Lft 2e) <strong>and</strong><br />

poorly sorted bioclastic pack- to grainstone (Lft 2e)<br />

Fig. 11. Measurements <strong>of</strong> the two-dimensional permeability distribution with the mini-permeameter. Each rock<br />

column is covered (1× 0.2 m) with a grid <strong>of</strong> 7× 50 data points.


Geological Society, London, Special Publications published online June 27, 2012 as doi: 10.1144/SP370.6<br />

D. PALERMO ET AL.<br />

Fig. 12. Rock-columns from the central part <strong>of</strong> a major shoal body were derived for a 2.1× 1.1 m sized slab which was<br />

previously mapped for sedimentary structures. Permeability shows a patchy distribution but also a clear upward<br />

increasing trend in all four rock columns.<br />

facies types. The well-sorted pack <strong>and</strong> pack- to<br />

grainstones (A) <strong>of</strong> the bottom part have about<br />

10% porosity <strong>and</strong> low permeability, which can be<br />

explained by comparably high mud content, occluding<br />

parts <strong>of</strong> potential primary interparticle porespace.<br />

In contrast, the upper well-sorted grainstone<br />

sheet (C) has lower mud content <strong>and</strong> higher<br />

amounts <strong>of</strong> interparticle porosity, resulting in better<br />

porosity <strong>and</strong> permeability values. The poorly sorted<br />

bioclastic pack- to grainstones also show textural<br />

differences, but are additionally controlled by the<br />

amount <strong>of</strong> biomodic porosity <strong>of</strong> leached shells.<br />

Generally, the porosities in the investigated poorly<br />

sorted pack- to grainstone beds (B, D) range<br />

between 11.3 <strong>and</strong> 18.2%. The moderate permeability<br />

in bed B ranges between 7.6 <strong>and</strong> 32.9 mD, while<br />

the average permeability in the upper bed (D)<br />

varies between 90.1 <strong>and</strong> 202.0 mD. Thus, lateral<br />

changes in porosity <strong>and</strong> permeability within an individual<br />

facies body remain in the same order <strong>of</strong><br />

magnitude within a similar stratigraphic position.<br />

These changes seem to correspond generally to the<br />

vertical textural changes that seem to follow the<br />

observed stratigraphic cycles. Moreover, also the<br />

lateral differences in porosity <strong>and</strong> permeability<br />

within the individual layers seem to be accompanied<br />

by slight lateral variations in the Dunham texture.<br />

The general occurrence <strong>of</strong> laterally continuous<br />

reservoir bodies is commonly controlled by<br />

medium-scale cycles, since porous units occur<br />

preferentially around their regressive maxima. In<br />

contrast, internal porosity <strong>and</strong> permeability variations<br />

are <strong>of</strong>ten controlled by small-scale cycles.<br />

The best reservoir <strong>properties</strong> are found around<br />

small-scale regressive maximum. This effect is<br />

most likely the result <strong>of</strong> stratigraphically controlled<br />

variations in mud content, <strong>and</strong> additional complex<br />

diagenetic overprints <strong>of</strong> cementation <strong>and</strong> selective<br />

leaching (see previous section).<br />

Regarding the entire flow unit, the lateral pattern<br />

<strong>of</strong> the permeable layers along the same stratigraphic<br />

cycles indicates that the observed centimetre-scale<br />

heterogeneities owing to sedimentary structures<br />

play a subordinate role.


Geological Society, London, Special Publications published online June 27, 2012 as doi: 10.1144/SP370.6<br />

CARBONATE SAND BODIES IN GERMANY<br />

Fig. 13. A 2.1×1.1 m sized block sampled from a major shoal-body in quarry Neidenfels with subtle low-angle cross<br />

bedding (1) <strong>and</strong> the resulting permeability distribution (2). Permeability was interpolated from the rock columns along<br />

the major sedimentary structures. Note the systematic upward increase <strong>of</strong> permeability following a small-scale<br />

regressive hemicycle <strong>and</strong> the resulting lateral continuity <strong>of</strong> the highly permeable zone in the upper portion <strong>of</strong> the block.<br />

Kilometre scale<br />

Three sections through two shoal bodies, with a<br />

lateral spacing <strong>of</strong> 1 km each, were investigated for<br />

the distribution <strong>of</strong> facies, diagenetic changes <strong>and</strong> <strong>petrophysical</strong><br />

<strong>properties</strong> along the depositional gradient.<br />

The cross-section in Figure 7a displays the geometries<br />

<strong>of</strong> facies-bodies combined with porosity <strong>and</strong><br />

permeability values <strong>of</strong> medium- <strong>and</strong> small- scale<br />

cycles. Within the individual facies bodies, the<br />

porosity <strong>and</strong> permeability distribution shows gradual<br />

lateral changes. The cross-section <strong>of</strong> Figure 7b<br />

depicts the gradational lateral changes <strong>of</strong> mud<br />

content <strong>and</strong> cement-types, which are held as key<br />

indicators for the relationship between the diagenetic<br />

overprint <strong>and</strong> the resulting reservoir <strong>properties</strong>. The<br />

overall distribution <strong>of</strong> petrographic trends <strong>and</strong> reservoir<br />

<strong>properties</strong> shows similar patterns. Both <strong>properties</strong><br />

show laterally gradational transitions <strong>and</strong> are<br />

strongly influenced by stratigraphic cycles. The vertical<br />

distribution <strong>of</strong> reservoir <strong>properties</strong> is mainly<br />

controlled by medium-scale cycles, whereas the<br />

smaller internal variations seem to be controlled by<br />

small-scale cycles.<br />

Petrophysical <strong>modelling</strong><br />

Three-dimensional geological <strong>modelling</strong> s<strong>of</strong>tware<br />

<strong>of</strong>fers a wide range <strong>of</strong> different possibilities for the<br />

distribution <strong>of</strong> <strong>petrophysical</strong> <strong>properties</strong>. However,<br />

in contrast to the subsurface, outcrop analogue<br />

studies allow a direct determination <strong>of</strong> the spatial<br />

distributions <strong>of</strong> reservoir <strong>properties</strong>. Furthermore,<br />

outcrop data can be used to constrain different conditioning<br />

factors <strong>and</strong> algorithms that are applied<br />

in the <strong>modelling</strong> process. As documented above,<br />

the lateral distribution <strong>of</strong> reservoir <strong>properties</strong>


Fig. 14. Outcrop wall panel with stratigraphic cycles, sedimentology <strong>and</strong> the reservoir <strong>properties</strong>. Lateral changes in porosity <strong>and</strong> permeability within the individual facies bodies<br />

remain in the same order <strong>of</strong> magnitude within a similar stratigraphic position. The red arrows indicate a common trend in the Dunham texture (mud content) <strong>and</strong> the reservoir property<br />

values, following the stratigraphic cycles. (Dunham Textures: G, grainstone; PG, pack- to grainstone; P, packstone; W, wackestone; M, mudstone. Particle size: L, lutite; S, siltite;<br />

A, arenite; R, rudite.)<br />

D. PALERMO ET AL.<br />

Geological Society, London, Special Publications published online June 27, 2012 as doi: 10.1144/SP370.6


Geological Society, London, Special Publications published online June 27, 2012 as doi: 10.1144/SP370.6<br />

CARBONATE SAND BODIES IN GERMANY<br />

Fig. 15. Fence diagram <strong>of</strong> the 3D facies model (200× vertical exaggeration), which was used as the trend for the spatial<br />

distribution <strong>of</strong> the reservoir <strong>properties</strong>.<br />

within a facies body, in the same stratigraphic position,<br />

remains in the same order <strong>of</strong> magnitude for<br />

hundreds <strong>of</strong> metres. On a kilometre scale, facies<br />

bodies, reservoir <strong>properties</strong> <strong>and</strong> diagenetic trends<br />

show gradual lateral changes, whereas the mostly<br />

stratigraphic cycle-controlled vertical differences<br />

vary commonly on a decimetre scale.<br />

These observations within this particular setting<br />

were considered for the construction <strong>of</strong> the<br />

geological model. The detailed, deterministic geomodel<br />

<strong>of</strong> Palermo et al. (2010) was used as a main<br />

input for geostatistical data analysis <strong>and</strong> the distribution<br />

<strong>of</strong> reservoir <strong>properties</strong>. It covers an area<br />

on the scale <strong>of</strong> a Middle East giant gas-field<br />

(25 × 36 km) <strong>and</strong> provides both (a) high resolution<br />

sequence stratigraphic reservoir layering <strong>and</strong> (b) a<br />

detailed facies distribution (Fig. 15). It is composed<br />

<strong>of</strong> 619 layers <strong>and</strong> 3.5 million cells. This high vertical<br />

resolution provides the possibility to model detailed<br />

variations <strong>of</strong> reservoir <strong>properties</strong> within individual<br />

reservoir bodies. Furthermore, these observations<br />

<strong>and</strong> the data distribution suggest a deterministic<br />

approach for <strong>modelling</strong> reservoir <strong>properties</strong>. With<br />

respect to the palaeogeographic positions <strong>and</strong> the<br />

resulting differences in facies distribution, the<br />

well spacing in this study is commonly similar or<br />

smaller than the minimum dimension (commonly<br />

dip width) <strong>of</strong> the major reservoir bodies (cf. Kerans<br />

& Tinker 1997).<br />

Input data<br />

The investigation <strong>of</strong> thin sections showed that preservation<br />

<strong>and</strong> creation <strong>of</strong> pore space is restricted to<br />

high-energy shoal <strong>and</strong> shoal-fringe facies types<br />

<strong>and</strong> is modified by the selective diagenetic history.<br />

Therefore, the primary facies types can be clearly<br />

subdivided into reservoir <strong>and</strong> non-reservoir facies.<br />

Furthermore, structural influences (e.g. fractures)<br />

on the reservoir <strong>properties</strong> are generally negligible.<br />

These observations <strong>and</strong> the s<strong>of</strong>tware-dependent<br />

necessity for a constant sampling rate led to the following<br />

steps in preparing the dataset for 3D <strong>modelling</strong>:<br />

after the insertion <strong>of</strong> the plug measurements as<br />

point values, all portions <strong>of</strong> reservoir facies without<br />

corresponding plug-data were set to ‘undefined’<br />

(2999.2). Subsequently, the porosity <strong>and</strong> permeability<br />

values <strong>of</strong> all non-reservoir facies types were<br />

set to zero. The step <strong>of</strong> data preparation was furthermore<br />

used for a quality control <strong>of</strong> the dataset. The<br />

upscaling process transfers porosity <strong>and</strong> permeability<br />

plug measurements to the grid cells neighbouring


Geological Society, London, Special Publications published online June 27, 2012 as doi: 10.1144/SP370.6<br />

D. PALERMO ET AL.<br />

Table 6. Results from the sample-variogram analysis for porosity <strong>and</strong> permeability, which are the main input<br />

parameters for <strong>petrophysical</strong> distribution algorithms (sequential Gaussian simulation <strong>and</strong> kriging)<br />

Results from sample variograms Shoal fringe Mid-shoal Inner shoal<br />

Porosity (F)<br />

Variogram type Gaussian Gaussian Gaussian<br />

Nugget (r<strong>and</strong>om values) 27% 27% 29%<br />

Vertical range 0.6 m 1.5 m 1.8 m<br />

Major direction 2938 2888 2158<br />

Range 3991 m 4702 m 1930 m<br />

B<strong>and</strong>width 2211 m 2612 m 1101 m<br />

Number <strong>of</strong> lags 10 15 15<br />

Search radius 9000 m 10230 m 6580 m<br />

Minor direction 2038 1988 1258<br />

Range 2029 m 2843 m 1732 m<br />

B<strong>and</strong>width 1624 m 1004 m<br />

Number <strong>of</strong> lags 10 15 15<br />

Search radius 6000 m 6292 m 5070 m<br />

Permeability (K)<br />

Variogram type Gaussian Gaussian Spherical<br />

Nugget (r<strong>and</strong>om values) 53% 24% 36%<br />

Vertical range 1.1 m 0.8 m 3.2 m<br />

Major direction 1168 2238 1498<br />

Range 920.7 m 4552.3 m 4721 m<br />

B<strong>and</strong>width 1121.7 m 1975 m 2585<br />

Number <strong>of</strong> lags 15 20 20<br />

Search radius 4254 m 12844.4 m 14070<br />

Minor direction 268 1258 598<br />

Range 728 m 1732 m 2568<br />

B<strong>and</strong>width 1121 m 1610 m 1858<br />

Number <strong>of</strong> lags 25 15 17<br />

Search radius 4096 8310.4 m 8000<br />

the wells. In contrast to porosity, permeability tends<br />

to change exponentially <strong>and</strong> requires a different averaging<br />

method for the upscaling process:<br />

(1) Porosity: arithmetic average <strong>of</strong> point data<br />

upscaled on neighbour cells.<br />

(2) Permeability: geometric average (e.g. Warren<br />

& Price 1961; Jensen 1991) <strong>of</strong> point data<br />

upscaled on neighbour cells.<br />

However, owing to the extraordinary high vertical<br />

resolution (average 0.1 m) <strong>of</strong> the 3D grid, which<br />

corresponds to the sampling rate <strong>of</strong> the plug<br />

measurements, the upscaling had no considerable<br />

impact on the quality <strong>of</strong> the model.<br />

Geostatistical data analysis<br />

Variograms are geostatistical models for the 3D distribution<br />

<strong>of</strong> reservoir heterogeneity <strong>and</strong> fundamental<br />

input parameters for the most common <strong>modelling</strong><br />

algorithms, for example sequential Gaussian simulations<br />

<strong>and</strong> kriging. Variograms are defined as the<br />

semi-variance <strong>of</strong> the difference between field values<br />

at two locations across realizations <strong>of</strong> the field<br />

(Cressie 1993). In order to establish the necessary<br />

input parameters, a geostatistical variogram analysis<br />

(Gringarten & Deutsch 2001) was carried out on the<br />

complete dataset with the original log data spaced<br />

between 10 m <strong>and</strong> a few kilometres. Based on the<br />

observations described above, the reservoir <strong>properties</strong><br />

were analysed for each facies type <strong>and</strong> each<br />

association with various lag-sizes, ranges, <strong>and</strong> so<br />

on. A reasonable variogram model could only be<br />

found for the facies associations, by using relatively<br />

large bin sizes <strong>and</strong> large ranges that are in the order<br />

<strong>of</strong> the observed facies changes (Palermo et al.<br />

2010). The results <strong>and</strong> the used parameters are summarized<br />

in Table 6. The vertical semi-variance was<br />

determined within decimetre-sized lags <strong>and</strong> the<br />

resulting ranges for porosity result in 0.6 m for shoalfringe,<br />

1.6 m for mid-shoal <strong>and</strong> 1.8 m for inner-shoal<br />

facies associations. The lateral semi-variance <strong>of</strong> porosity<br />

results in Gaussian variogram models for all<br />

three facies associations with high r<strong>and</strong>om values<br />

(Nugget effect, after Eisenberg et al. 1994) slightly<br />

below 30% in all three cases. The lateral ranges are<br />

remarkably high <strong>and</strong> reach values in the major direction<br />

<strong>of</strong> 1930 m for the inner-shoal, 3991 m for shoal


Geological Society, London, Special Publications published online June 27, 2012 as doi: 10.1144/SP370.6<br />

CARBONATE SAND BODIES IN GERMANY<br />

fringe <strong>and</strong> 4702 m for mid-shoal facies associations.<br />

The semi-variance <strong>of</strong> the permeability is generally<br />

characterized by higher r<strong>and</strong>om values in the shoal<br />

fringe (53%) <strong>and</strong> the inner shoal (36%) facies types,<br />

whereas the r<strong>and</strong>om factor <strong>of</strong> the mid-shoal (24%)<br />

is comparably low. Compared with the porosity, the<br />

computed permeability ranges are significantly<br />

lower for the shoal fringe, similar for the mid-shoal<br />

<strong>and</strong> higher for the inner-shoal facies associations.<br />

The results <strong>of</strong> the variogram analysis for both porosity<br />

<strong>and</strong> permeability are summarized in Table 6.<br />

Effect <strong>of</strong> input data<br />

As described in the sections above, the major portion<br />

<strong>of</strong> the effective porosity within the system was<br />

created at an early stage <strong>and</strong> is mostly linked to<br />

either primary porosity or cycle-controlled early<br />

diagenetic leaching. In the upper described qualitative<br />

<strong>and</strong> quantitative outcrop observations, two major<br />

scales <strong>of</strong> lateral heterogeneities could be observed:<br />

(a) small-scale heterogeneities on centimetre to decimetre<br />

scale; <strong>and</strong> (b) large-scale, smooth transitions<br />

in the order <strong>of</strong> the lateral facies association extensions.<br />

A reasonable variogram model could only be<br />

found for the largest scale. The intermediate scale,<br />

tens to hundreds <strong>of</strong> metres <strong>and</strong> the centimetre scale<br />

did not provide enough data to make a sufficient<br />

variogram analysis. According to the previous<br />

observations, the remarkably high lateral porosity<br />

ranges can be related to the subtle gradational transitions<br />

<strong>of</strong> mud content along the depositional slope,<br />

reflecting the depositional energy <strong>of</strong> a flat <strong>and</strong> very<br />

gently inclined epeiric <strong>carbonate</strong> ramp.<br />

Owing to the continuous character <strong>of</strong> the porosity<br />

that can be traced visually along the outcrop<br />

walls, the predominantly Gaussian semi-variogram<br />

models <strong>and</strong> the resulting ranges were considered<br />

as the most probable scenario. However, the possibility<br />

<strong>of</strong> other scenarios cannot be generally<br />

excluded. Follow-up sampling as performed by,<br />

for example, Pranter et al. (2005) is planned to<br />

better quantify the small-scale heterogeneities.<br />

Conditioning <strong>and</strong> <strong>modelling</strong> algorithms<br />

In order to compare the impact <strong>of</strong> stratigraphy <strong>and</strong><br />

facies (Palermo et al. 2010) on the distribution <strong>of</strong><br />

reservoir <strong>properties</strong>, three different approaches<br />

were used for the <strong>petrophysical</strong> <strong>modelling</strong>.<br />

(1) Unconditioned. The grid used here is a relatively<br />

simple 3D cube that was subdivided<br />

into 180 conformable layers. Only the top<br />

<strong>and</strong> base <strong>of</strong> the model were defined by stratigraphic<br />

surfaces. Unrelated to facies or stratigraphy,<br />

the modelled reservoir <strong>properties</strong><br />

were directly interpolated between the data<br />

points using different algorithms.<br />

(2) Conditioned to stratigraphic cycles. In this<br />

approach, the <strong>petrophysical</strong> <strong>properties</strong> were<br />

conditioned to the zones based on stratigraphic<br />

cycles. Each reservoir bearing zone<br />

was subdivided into eight layers <strong>and</strong> the<br />

average vertical resolution was around 10 cm.<br />

Therefore, vertical variations <strong>of</strong> reservoir <strong>properties</strong><br />

within single reservoir units could also<br />

be displayed with a high level <strong>of</strong> detail.<br />

(3) Conditioned to stratigraphic cycles <strong>and</strong> facies<br />

associations. In this scenario the entire information<br />

from the geological model (Fig. 17)<br />

was used to model the spatial distribution <strong>of</strong><br />

reservoir <strong>properties</strong>. Each reservoir facies<br />

association was modelled individually, <strong>and</strong><br />

tight facies types were set to zero. To avoid<br />

unrealistically sharp contacts between the<br />

lateral facies boundaries the <strong>properties</strong> were<br />

slightly smoothened.<br />

Three different <strong>modelling</strong> algorithms were tested<br />

for the distribution <strong>of</strong> the reservoir <strong>properties</strong> in<br />

the upper described scenarios <strong>and</strong> compared with<br />

the outcrop observations (Table 7):<br />

(1) Moving average. This interpolation algorithm<br />

is based on the average values <strong>of</strong> the input data<br />

<strong>and</strong> weights the distribution according to the<br />

distance from data points without a spatial<br />

preference (Chou 1975).<br />

(2) Kriging. This is a deterministic interpolation<br />

method developed by Matheron (1963) <strong>and</strong><br />

D. G. Krige. The algorithm distributes a property<br />

with a spatial preference that has to be<br />

defined by an input variogram. The estimated<br />

variables are then calculated by the linear<br />

combination <strong>of</strong> the input data points. In contrast<br />

to the moving average, the kriging algorithm<br />

uses a weighting function <strong>and</strong> honours<br />

both distance <strong>and</strong> direction.<br />

(3) Sequential Gaussian simulation (SGS). This<br />

variogram based stochastic simulation technique<br />

is based on kriging <strong>and</strong> allows multiple<br />

equiprobable realizations <strong>of</strong> a property<br />

(Alabert 1987; Deutsch & Journel 1992).<br />

Kriging calculates the weighted mean <strong>and</strong><br />

the st<strong>and</strong>ard deviation at each estimated<br />

point. SGS, however, represents the variable<br />

as a r<strong>and</strong>om deviate from the Gaussian<br />

normal distribution, which represents the<br />

input data <strong>and</strong> trends better than most <strong>of</strong> the<br />

kriging algorithms.<br />

Porosity distribution with sequential<br />

Gaussian simulation (Fig. 16)<br />

Porosity – unconditioned. According to the r<strong>and</strong>omity<br />

<strong>of</strong> the algorithm, the distribution <strong>of</strong> porosity<br />

appears very patchy <strong>and</strong> the general geological


Table 7. Parameters <strong>and</strong> algorithms used for the spatial distribution <strong>of</strong> reservoir <strong>properties</strong><br />

Moving average Vertical range Point weighting Direction Distribution Trends<br />

Porosity 1 m Exponent: 4 Follow layers normal Facies<br />

Permeability 1 m Exponent: 4 Follow layers Logarithmic Facies, porosity<br />

Sequential Gaussian simulation Variogram settings Transformations Distribution Trends<br />

Porosity Parameters from variogram analysis Normal score Linear Facies<br />

Permeability Parameters from variogram analysis Normal score Logarithmic Facies, porosity<br />

Kriging Variogram settings Transformations Distribution Trends<br />

Porosity Parameters from variogram analysis Normal score Linear Facies<br />

Permeability Parameters from variogram analysis Normal score Logarithmic Facies, porosity<br />

D. PALERMO ET AL.<br />

Geological Society, London, Special Publications published online June 27, 2012 as doi: 10.1144/SP370.6


Geological Society, London, Special Publications published online June 27, 2012 as doi: 10.1144/SP370.6<br />

CARBONATE SAND BODIES IN GERMANY<br />

Fig. 16. Different porosity distributions with the Sequential Gaussian simulation algorithms conditioned to: (a) a<br />

simple layering – the porosity values are vertically scattered <strong>and</strong> form laterally simple bulls eye patterns; (b)<br />

stratigraphic cycles – vertical distribution matches better with the outcrop observations, although unconnected bulls<br />

eyes are still present; (c) stratigraphic cycles <strong>and</strong> facies – the smoothened distribution shows the best match with the<br />

outcrop continuities.<br />

pattern is not recognizable. Additionally, the reservoir<br />

bodies are relatively small with low lateral continuity.<br />

Both kriging <strong>and</strong> moving average provided<br />

better results than SGS in this scenario.<br />

Porosity – conditioned to stratigraphic cycles. The<br />

resulting reservoir bodies follow the stratigraphic<br />

trends <strong>and</strong> have acceptable dimensions. However,<br />

compared with the results produced with the<br />

moving average algorithm, many artefacts are<br />

visible, especially in the tight middle part.<br />

Porosity – conditioned to cycles <strong>and</strong> facies associations.<br />

The modelled porosities follow the stratigraphic<br />

trends <strong>and</strong> the outline <strong>of</strong> the facies bodies.<br />

Compared with the results with the moving average<br />

Fig. 17. Different permeability distributions with the Sequential Gaussian simulation algorithms conditioned to (a) a<br />

simple layering – the porosity values are vertically scattered <strong>and</strong> laterally isolated flow units, (b) stratigraphic cycles –<br />

more realistic vertical connectivity, but the flow units are laterally still isolated, <strong>and</strong> (c) stratigraphic cycles <strong>and</strong> facies –<br />

the smoothened distribution shows a good match with the outcrop observations.


Geological Society, London, Special Publications published online June 27, 2012 as doi: 10.1144/SP370.6<br />

D. PALERMO ET AL.<br />

algorithm <strong>and</strong> kriging, the reservoir bodies modelled<br />

with a slightly smoothed sequential Gaussian Simulation<br />

show a better connectivity <strong>and</strong> more reasonable<br />

dimensions. Analysing this final scenario, the<br />

resulting net/gross ratio <strong>of</strong> the modelled Upper<br />

Muschelkalk area amounts to 7%, with a theoretical<br />

storage capacity <strong>of</strong> 419 490 000 m 3 .<br />

Permeability distribution with sequential<br />

Gaussian simulation (Fig. 17, Fig. 18)<br />

Permeability – conditioned to porosity. Like the<br />

porosity, the distribution <strong>of</strong> permeability appears<br />

patchy <strong>and</strong> the general geological pattern is not<br />

recognizable. The reservoir bodies are comparably<br />

small <strong>and</strong> highly compartmentalized.<br />

Permeability – conditioned to stratigraphic cycles<br />

<strong>and</strong> porosity. Comparable to the distribution <strong>of</strong> porosity,<br />

the resulting flow units follow the general<br />

stratigraphic trends. Although there are no artefacts<br />

in the middle part visible, the distribution <strong>of</strong> permeability<br />

remains patchy, especially in the upper part.<br />

Permeability – conditioned to cycles, facies associations<br />

<strong>and</strong> porosity. The modelled permeability<br />

follows the stratigraphic trends <strong>and</strong> the outline <strong>of</strong><br />

the facies bodies. Both dimensions <strong>and</strong> connectivity<br />

<strong>of</strong> the flow units seem to be more reasonable in<br />

this case.<br />

Modelling results<br />

The outcrop analogue studies <strong>of</strong> Palermo et al.<br />

(2010) could show that the shoal facies associations<br />

show remarkable lateral extensions <strong>of</strong> up to several<br />

tens <strong>of</strong> kilometres whereas vertical facies changes<br />

range within the order <strong>of</strong> decimetres. Another part<br />

<strong>of</strong> the study shows that lateral facies changes are<br />

characterized by very subtle <strong>and</strong> gradual transitions<br />

within the range <strong>of</strong> up to a few kilometres. Therefore,<br />

the Palermo et al. (2010) study suggests<br />

Fig. 18. Filtered permeable reservoir bodies (cut-<strong>of</strong>f 1 mD) distributed with sequential Gaussian simulation,<br />

conditioned to stratigraphic cycles <strong>and</strong> facies <strong>and</strong> processed with a smoothing algorithm. All three hierarchies <strong>of</strong><br />

stratigraphic cycles had an impact on quality <strong>and</strong> presence <strong>of</strong> the flow units: (a) large scale cycle – controls the lateral<br />

extension, retro- <strong>and</strong> progradation, (b) medium scale cycles – control the stratigraphic presence <strong>of</strong> a flow unit, <strong>and</strong><br />

(c) small-scale cycles – control the body internal vertical heterogeneities. Furthermore the distribution <strong>of</strong> palaeohighs<br />

<strong>and</strong> -lows is important for the general localization <strong>of</strong> the reservoir bodies.


Geological Society, London, Special Publications published online June 27, 2012 as doi: 10.1144/SP370.6<br />

CARBONATE SAND BODIES IN GERMANY<br />

encoding facies as a continuous rather than discrete<br />

property. The thin section observations <strong>and</strong> <strong>petrophysical</strong><br />

measurements confirm theses observations.<br />

Mud content in the matrix changes laterally <strong>and</strong> vertically<br />

in a gradual way <strong>and</strong> reflects the depositional<br />

energy <strong>of</strong> the epeiric system with a very gentle<br />

depositional gradient. The outcrop observations<br />

above show that mud content has a significant impact<br />

on the presence <strong>of</strong> primary porosity. However, early<br />

diagenetic mouldic porosity also seems to increase<br />

with increasing matrix mud content. The <strong>petrophysical</strong><br />

investigations on multiple scales showed two<br />

major scales <strong>of</strong> reservoir heterogeneities:<br />

(1) Small-scale heterogeneities on centimetre to<br />

several decimetre scales mostly controlled by<br />

sedimentary structures <strong>and</strong> diverse mouldic<br />

pore types from different components were<br />

observed. Similar small-scale heterogeneities<br />

have been described by several studies (e.g.<br />

Jennings et al. 1998; Pranter et al. 2005)<br />

However, the quantity <strong>of</strong> samples in this<br />

study was insufficient to perform a solid variogram<br />

analysis that could quantify the 3D lateral<br />

ranges <strong>of</strong> the small-scale heterogeneities.<br />

(2) On a large scale, relatively smooth transitions<br />

on the order <strong>of</strong> a few kilometres were<br />

observed <strong>and</strong> could be confirmed by the variogram<br />

analysis. However, the resulting lateral<br />

long range features are significantly higher<br />

compared with the dolomitized platform<br />

<strong>carbonate</strong>s investigated by Jennings et al.<br />

(1998), which can reach up to 800 m. Apart<br />

from the different depositional settings <strong>and</strong><br />

diagenetic histories, which make a direct<br />

comparison difficult, the possibility <strong>of</strong> additional<br />

nested heterogeneities as observed <strong>and</strong><br />

described by Pranter et al. (2005) cannot be<br />

excluded. At least one smaller heterogeneity<br />

scale is most likely hidden within the high<br />

r<strong>and</strong>om effect <strong>and</strong> could be an interesting<br />

subject for further investigations. Extensive<br />

lateral sampling is foreseen for a future<br />

study to investigate the presence <strong>of</strong> additional<br />

meso-scale heterogeneities.<br />

Conclusions<br />

This outcrop analogue study documents the reservoir<br />

<strong>properties</strong> in the Upper Muschelkalk formation<br />

as an outcrop analogue for epeiric <strong>carbonate</strong>s.<br />

Average porosity is 9–23%, <strong>and</strong> average permeability<br />

is 21–700 mD. The net/gross ratio <strong>of</strong><br />

the Upper Muschelkalk amounts to 7%, with a<br />

theoretical storage capacity <strong>of</strong> 419 490 000 m 3 .<br />

Laterally, units with high porosity <strong>and</strong> permeability<br />

show a kilometre-scale lateral continuity<br />

with subtle gradational transitions. The main part<br />

<strong>of</strong> the porosity in the shoal bodies has been<br />

created at an early stage <strong>and</strong> is mostly linked to<br />

either cycle-controlled (a) primary porosity triggered<br />

by a systematically changing mud content or<br />

(b) early diagenetic leaching around maximum<br />

regressions.<br />

Two major scales <strong>of</strong> lateral heterogeneities could<br />

be observed: (a) small-scale heterogeneities on<br />

centimetre to decimetre scale, mainly related to<br />

sedimentary structures <strong>and</strong> different mouldic pore<br />

types which are represented by high r<strong>and</strong>om<br />

values (30%) within the geostatistical variogram<br />

model; <strong>and</strong> (b) large-scale heterogeneities <strong>of</strong> up to<br />

4700 m lateral range. The remarkably high lateral<br />

porosity <strong>and</strong> permeability ranges can be partly<br />

related to subtle gradational transitions <strong>of</strong> mud<br />

content along the depositional slope, which reflect<br />

the depositional energy <strong>of</strong> a flat <strong>and</strong> very gently<br />

inclined epeiric <strong>carbonate</strong> ramp.<br />

Different <strong>petrophysical</strong> <strong>modelling</strong> scenarios<br />

could show the impact <strong>and</strong> necessity <strong>of</strong> geological<br />

constraints for the geostatistical distribution <strong>of</strong> porosity<br />

<strong>and</strong> permeability. The following factors help<br />

in predicting <strong>and</strong> <strong>modelling</strong> the spatial distribution<br />

<strong>of</strong> reservoir <strong>properties</strong> within the Upper Muschelkalk,<br />

which may be transferred to similar epeiric<br />

<strong>carbonate</strong> reservoirs (Fig. 15):<br />

(1) Facies associations. The reservoir <strong>properties</strong><br />

show a close relationship to the lith<strong>of</strong>acies<br />

<strong>and</strong> the associated matrix mud content.<br />

Porous facies types are restricted to the highenergy<br />

shoal facies, whereas both muddy<br />

inner <strong>and</strong> outer ramp facies are commonly<br />

tight since the muddy matrix protects the sediment<br />

from diagenetic fluids <strong>and</strong> associated<br />

mouldic porosity creation. <strong>Reservoir</strong> heterogeneities<br />

within the shoal bodies can be<br />

mainly related to the mud content <strong>and</strong> early<br />

diagenesis.<br />

(2) Statigraphic cycles. All three hierarchies <strong>of</strong><br />

stratigraphic cycles have an impact on quality<br />

<strong>and</strong> presence <strong>of</strong> the flow units: (a) large-scale<br />

cycle – controls the lateral extent, as well as<br />

retro- <strong>and</strong> pro- gradation <strong>of</strong> the reservoir<br />

bodies; (b) medium-scale cycles – control the<br />

stratigraphic presence <strong>of</strong> the reservoir facies<br />

associations; <strong>and</strong> (c) small-scale cycles –<br />

control the mud content <strong>and</strong> early mouldic<br />

porosity creation in facies types with primary<br />

porosity. which are considered as key drivers<br />

for systematic vertical changes in porosity<br />

<strong>and</strong> permeability within the reservoir bodies.<br />

(3) Palaeo-relief. Gross volume <strong>and</strong> dimensions<br />

<strong>of</strong> the reservoir bodies seem to be mainly controlled<br />

by the combination <strong>of</strong> both stratigraphic<br />

cycles <strong>and</strong> a subtle palaeo-relief,


Geological Society, London, Special Publications published online June 27, 2012 as doi: 10.1144/SP370.6<br />

D. PALERMO ET AL.<br />

Authors<br />

induced by slight differential subsidence <strong>of</strong><br />

inherited structural grains. In particular,<br />

palaeo-highs are important for the presence<br />

<strong>of</strong> reservoir facies.<br />

Denis Palermo studied geology at the University <strong>of</strong><br />

Tübingen <strong>and</strong> completed his MSc thesis in 2004 on<br />

<strong>carbonate</strong> reservoir sedimentology at the Nederl<strong>and</strong>se<br />

Aardolje Matschapij (NAM) in the Netherl<strong>and</strong>s.<br />

He then joined the Sedimentary Geology<br />

Group at the University <strong>of</strong> Tübingen to do a Ph.D.<br />

on outcrop analogue <strong>modelling</strong> <strong>and</strong> reservoir characterization<br />

in cooperation with Eni E&P. In<br />

2007 he joined Eni E&P as sedimentologist <strong>and</strong><br />

member <strong>of</strong> the Eni E&P <strong>carbonate</strong> research team.<br />

Thomas Aigner studied geology at the universities<br />

<strong>of</strong> Stuttgart, Tübingen, Reading <strong>and</strong> Miami.<br />

After working for 6 years with Shell Research, he<br />

has been a pr<strong>of</strong>essor <strong>and</strong> head <strong>of</strong> the sedimentary<br />

geology group at the University <strong>of</strong> Tübingen since<br />

1991. His research group focuses on reservoir<br />

geology, <strong>carbonate</strong> reservoir characterization <strong>and</strong><br />

3D <strong>modelling</strong>. He was an AAPG European Distinguished<br />

Lecturer in 1996. In 2007–2008 he spent<br />

a sabbatical with Petroleum Development Oman<br />

<strong>and</strong> Qatar Shell.<br />

Björn Seyfang studied geology at the University<br />

<strong>of</strong> Tübingen <strong>and</strong> finished his M.Sc. on outcrop analogue<br />

studies for <strong>carbonate</strong> shoal reservoirs in 2006.<br />

In 2009, he completed his Ph.D. on facies <strong>and</strong> reservoir<br />

<strong>modelling</strong> at the Sedimentary Geology Group<br />

<strong>of</strong> Tübingen University, initiated by GDF-SUEZ<br />

<strong>and</strong> conducted in cooperation with ExxonMobil,<br />

Wintershall <strong>and</strong> RWE-Dea. In September 2009, he<br />

joined Total as an operations geologist.<br />

Sergio Nardon graduated in geology at the University<br />

<strong>of</strong> Trieste, Italy, in 1982 <strong>and</strong> joined Eni<br />

(former Agip) in 1984 as clastic sedimentologist.<br />

His experience spans from fractured reservoir<br />

characterization to <strong>carbonate</strong> sedimentology <strong>and</strong><br />

he has worked on several projects in Middle East,<br />

Europe <strong>and</strong> North Africa. His main interests are<br />

the integration <strong>of</strong> <strong>carbonate</strong> architecture, outcrop<br />

data <strong>and</strong> geological concepts into the 3D environment.<br />

He is currently technical leader <strong>of</strong> the research<br />

<strong>and</strong> development team at the geology department <strong>of</strong><br />

Eni E&P.<br />

This work is part <strong>of</strong> an integrated research <strong>and</strong> development<br />

project funded by Eni E&P. We want to express our<br />

thanks to Eni Management for their support <strong>and</strong> the permission<br />

to publish. We are grateful for discussions with<br />

many colleagues working on the German Muschelkalk,<br />

notably H. Hagdorn, T. Simon, M. Urlichs <strong>and</strong><br />

R. Borkhataria. Assistance was provided by members<br />

<strong>of</strong> the Sedimentary Geology group at the University <strong>of</strong><br />

Tübingen (A. Allgöwer, E. Dmitrieva, M. Looser, A.<br />

Schmid-Röhl, C. Schneider, M. Zeller) <strong>and</strong> A. Satterley,<br />

B. White <strong>and</strong> K. Mair from Eni E&P. P. Jeiseke produced<br />

the thin sections. Several quarry companies, for<br />

example Schön & Hippelein, Hohenloher Schotterwerke,<br />

Schotterwerk Schuhmann, J. Heumann <strong>and</strong> SHB Schotterwerke<br />

Hohenlohe generously allowed access to their quarries.<br />

The Geological Survey <strong>of</strong> Baden-Württemberg <strong>and</strong><br />

Bayern supplied us with borehole cores <strong>and</strong> well data.<br />

We wish to thank A. Glocke <strong>and</strong> Schlumberger for<br />

the access to Petrel (TM <strong>of</strong> Schlumberger) <strong>and</strong> to<br />

A. Henriette (ALT) for access to WellCAD.<br />

References<br />

Aigner, T. 1985. Storm Depositional Systems. Dynamic<br />

Stratigraphy in Modern <strong>and</strong> Ancient Shallow Marine<br />

Sequences. Lecture Notes in Earth Sciences, Vol. 3,<br />

Springer, Berlin.<br />

Aigner, T. 1999. Dynamische Stratigraphie des Oberen<br />

Muschelkalks am Beispiel Südwestdeutschl<strong>and</strong>s,<br />

115–128.<br />

Aigner, T., Braun, S., Palermo,D.&Blendinger,W.<br />

2007. 3D geological <strong>modelling</strong> <strong>of</strong> a <strong>carbonate</strong> shoal<br />

complex: reservoir analogue study using outcrop<br />

data. First Break, 25, 65–72.<br />

Alabert, F. 1987. The practice <strong>of</strong> fast conditional simulations<br />

through the LU decomposition <strong>of</strong> the covariance<br />

matrix. Mathematical Geology, 19, 369–386.<br />

Allgöwer, A. 2006. Stratigraphy, petrophysics <strong>and</strong><br />

facies analysis <strong>of</strong> epeiric <strong>carbonate</strong>s in the Upper Muschelkalk.<br />

An outcrop analogue study for skeletal/<br />

oolitic <strong>carbonate</strong> s<strong>and</strong> bodies in the Middle East<br />

(Upper Muschelkalk, South German Basin). Diploma<br />

thesis, University <strong>of</strong> Tübingen.<br />

Azerêdo, A. C. 1998. Geometry <strong>and</strong> facies dynamics<br />

<strong>of</strong> Middle Jurassic <strong>carbonate</strong> ramp s<strong>and</strong>bodies. In:<br />

Wright, V.P.&Burchette, T. P. (eds) Carbonate<br />

Ramps. Geological Society, London, Special Publications,<br />

149, 281–314.<br />

Bachmann, G. H . 1973. Die karbonatischen Best<strong>and</strong>teile<br />

des Oberen Muschelkalks (Mittlere Trias) in Südwest-<br />

Deutschl<strong>and</strong> und ihre Diagenese. Arbeiten des Instituts<br />

für Geologie und Paläontologie der Universität<br />

Stuttgart, 68, 1–99.<br />

Borgomano, J. R. F., Masse, J.-P. & Al Maskiry, S.<br />

2002. The lower Aptian Shuaiba <strong>carbonate</strong> outcrops<br />

in Jebel Akhdar, northern Oman: Impact on static<br />

<strong>modelling</strong> for Shuaiba petroleum reservoirs. AAPG<br />

Bulletin, 86, 1513–1529.<br />

Braun, S., 2003. Quantitative analysis <strong>of</strong> <strong>carbonate</strong> s<strong>and</strong>bodies:<br />

outcrop analogue study from an epicontinental<br />

basin (Triassic Germany). PhD thesis, University <strong>of</strong><br />

Tübingen.<br />

Burchette, T. P., Wright, V.P.&Faulkner, T.J.<br />

1990. Oolitic s<strong>and</strong>body depositional models <strong>and</strong><br />

geometries, Mississippian <strong>of</strong> southwest Britain: implications<br />

for petroleum exploration in <strong>carbonate</strong> ramp<br />

settings. Sedimentary Geology, 68, 87–115.<br />

Cavallo, L.J.&Smosna, R. 1997. Predicting porosity<br />

distribution within oolitic tidal bars. In: Kupecz,<br />

J. A., Gluyas, J. & Bloch, S. (eds) <strong>Reservoir</strong><br />

Quality Prediction in S<strong>and</strong>stones <strong>and</strong> Carbonates.


Geological Society, London, Special Publications published online June 27, 2012 as doi: 10.1144/SP370.6<br />

CARBONATE SAND BODIES IN GERMANY<br />

American Association <strong>of</strong> Petroleum Geologists, Tulsa,<br />

OK, Memoirs, 69, 211–229.<br />

Choquette,P.W.&Pray, L. C . 1970. Geologic nomenclature<br />

<strong>and</strong> classification <strong>of</strong> porosity in <strong>carbonate</strong>s.<br />

Bulletin AAPG, 54, 107–250.<br />

Chou, Y. 1975. Statistical Analysis with Business <strong>and</strong><br />

Economic Applications. Elsevier, Amsterdam.<br />

Cressie, N. 1993. Statistics for Spatial Data. Wiley,<br />

New York.<br />

Dercourt, J., Ricou, L.E.&Vrielynck, B. (eds) 1993.<br />

Atlas Tethys Paleoenvironmental Maps. Gauthier-<br />

Villars, Paris.<br />

Deutsch,C.V.&Journel, A. G. 1992. GSLIB: Geostatistical<br />

S<strong>of</strong>tware Library <strong>and</strong> user’s Guide. Oxford<br />

University Press, New York.<br />

Dmitrieva, E. 2006. Facies distribution, sequence stratigraphy<br />

<strong>and</strong> poroperm analysis <strong>of</strong> <strong>carbonate</strong> shoal<br />

bodies in the Upper Muschelkalk (middle Jagst<br />

valley, South German Basin). An outcrop analogue<br />

study for <strong>carbonate</strong> reservoirs in the Middle East.<br />

Diploma thesis, University <strong>of</strong> Tübingen.<br />

Eisenberg, R. A., Harris, P. M., Grant, C. W., Goggin,<br />

D. J. & Conner, F. J. 1994. Modeling reservoir heterogeneity<br />

within outer ramp <strong>carbonate</strong> facies using an<br />

outcrop analog, San Andres formation <strong>of</strong> the Permian<br />

Basin. AAPG Bulletin, 78, 1337–1359.<br />

Flügel, E. 2004. Micr<strong>of</strong>acies <strong>of</strong> Carbonate Rocks. Analysis,<br />

Interpretation, Application. Springer, Heidelberg.<br />

Gawthorpe, R.L.&Gutteridge, P. 1990. Geometry<br />

<strong>and</strong> Evolution <strong>of</strong> Platform-Margin Bioclastic Shoals,<br />

Late Dinantian (Mississippian). International Association<br />

<strong>of</strong> Sedimentologists, Derbyshire, Special Publications,<br />

9, 39–54.<br />

Geyer, O.F.&Gwinner, M. P. 1991. Geologie von<br />

Baden-Württemberg. E. Schweizerbart’sche Verlagsbuchh<strong>and</strong>lung,<br />

Stuttgart<br />

Grammer, G. M., Harris, P.M.&Eberli, G. P. 2004.<br />

Integration <strong>of</strong> outcrop <strong>and</strong> modern analogs in reservoir<br />

modeling. American Association <strong>of</strong> Petroleum<br />

Geologists, Tulsa, OK, Memoirs, 80.<br />

Grant, C. W., Goggin, D.J.&Harris, P. M. 1994.<br />

Outcrop analog for cyclic-shelf reservoirs, San<br />

Andres formation <strong>of</strong> Permian Basin: stratigraphic framework,<br />

permeability distribution, geostatistics, <strong>and</strong><br />

fluid-flow modeling. AAPG Bulletin, 78, 23–54.<br />

Gringarten, E. & Deutsch, C. V. 2001. Variogram<br />

interpretation <strong>and</strong> modeling. Mathematical Geology,<br />

33, 507–534.<br />

Hagdorn, H. 1991. The Muschelkalk in Germany – An<br />

introduction. In: Hagdorn, H. (ed.) Muschelkalk, A<br />

Field Guide. Goldschneck, Korb.<br />

Hagdorn,H.&Simon, T. 1988. Geologie und L<strong>and</strong>schaft<br />

des Hohenloher L<strong>and</strong>es. Thorbecke, Sigmaringen.<br />

H<strong>and</strong>ford, R. C. 1988. Review <strong>of</strong> <strong>carbonate</strong> s<strong>and</strong>-belt<br />

deposition <strong>of</strong> ooid grainstones <strong>and</strong> application to<br />

Mississippian <strong>Reservoir</strong>, Damme Field, Southwestern<br />

Kansas. AAPG Bulletin, 72, 1184–1199.<br />

Harris, P.M.&Kowalik, W. S. 1994. Satellite images<br />

<strong>of</strong> <strong>carbonate</strong> depostional settings, examples <strong>of</strong><br />

reservoir- <strong>and</strong> exploration-scale geologic facies variation.<br />

AAPG Methods in Exploration Series, 11, 147.<br />

Jennings,W.J.Jr. 2000. Spatial Statistics <strong>of</strong> Permeability<br />

Data from Carbonate Outcrops <strong>of</strong> West Texas <strong>and</strong> New<br />

Mexico: implications for Improved <strong>Reservoir</strong><br />

Modelling. Bureau <strong>of</strong> Economic Geology. Report <strong>of</strong><br />

Investigations no. 258. University <strong>of</strong> Texas.<br />

Jennings, J. W., Ruppel, S.C.&Ward, W. B. 1998.<br />

Geostatistical analysis <strong>of</strong> <strong>petrophysical</strong> data <strong>and</strong><br />

<strong>modelling</strong> <strong>of</strong> fluid-flow effects in <strong>carbonate</strong> outcrops.<br />

Society <strong>of</strong> Petroleum Engineers Annual Technical<br />

Conference <strong>and</strong> Exhibition, New Orleans, 27–30<br />

September, SPE Paper 49025.<br />

Jensen, J. L. 1991. Use <strong>of</strong> the geometric average for effective<br />

permeability estimation. Mathematical Geology,<br />

23, 833–840.<br />

Kerans,C.&Tinker, S. W. 1997. Sequence Stratigraphy<br />

<strong>and</strong> Characterization <strong>of</strong> Carbonate <strong>Reservoir</strong>s.<br />

OSEPM Short Course Notes, 40.<br />

Kittridge, M. G., Lake, L. W., Lucia,F.J.&Fogg,G.E.<br />

1990. Outcrop/subsurface comparison <strong>of</strong> heterogeneity<br />

in the San Andres Formation. Society <strong>of</strong> Petroleum<br />

Engineers Formation Evaluation, 5, 233–240.<br />

Kostic, B. 2001. Sedimentäre Strukturen, Fazies und<br />

Poroperm-Eigenschaften in ausgewählten ‘Karbonats<strong>and</strong>en’:<br />

Quaderkalk, Oberer Muschelkalk. Diploma<br />

thesis, University <strong>of</strong> Tübingen.<br />

Kostic,B.&Aigner, T. 2004. Sedimentary <strong>and</strong> poroperm<br />

anatomy <strong>of</strong> shoal-water <strong>carbonate</strong>s (Muschelkalk,<br />

South German Basin): an outcrop-analogue study <strong>of</strong><br />

interwell spacing scale. Facies, 50, 113–131.<br />

Looser, M. 2006. Facies <strong>and</strong> poroperm parameters <strong>of</strong><br />

epeiric <strong>carbonate</strong>s in the Upper Muschelkalk <strong>of</strong> the<br />

middle Kocher valley. Diploma thesis, University <strong>of</strong><br />

Tübingen.<br />

Lucia, F. J. 1983. Petrophysical parameters estimated<br />

from visual description <strong>of</strong> <strong>carbonate</strong> rocks: a field<br />

classification <strong>of</strong> <strong>carbonate</strong> pore space. Journal <strong>of</strong><br />

Petroleum Technology, 3, 626–637.<br />

Lucia, F. J. 1999. Carbonate <strong>Reservoir</strong> Characterization.<br />

Springer, Berlin.<br />

Matheron, G. 1963. Principles <strong>of</strong> Geostatistics. Economic<br />

Geology, 58, 1266–1963.<br />

Ockert, W. 1988. Lithostratigraphie und Fossilführung<br />

des Trochitenkalks (Unterer Hauptmuschelkalk, mo1)<br />

im Raum Hohenlohe. In: Hagdorn, H. (ed.) Neue<br />

Forschungen zur Erdgeschichte von Crailsheim. Sonderbände<br />

der Gesellschaft für Naturkunde in Württemberg,<br />

Goldschneck, Korb, 1, 43–69.<br />

Palermo, D., Aigner, T., Geluk, M., Poppelreiter, M.<br />

& Pipping, K. 2008. <strong>Reservoir</strong> potential <strong>of</strong> a lacustrine<br />

mixed <strong>carbonate</strong>/siliciclastic gas reservoir: The Lower<br />

Triassic Rogenstein in the Netherl<strong>and</strong>s. Journal <strong>of</strong><br />

Petroleum Geology, 31, 61–96.<br />

Palermo, D., Aigner, T., Nardon,S.&Blendinger,W.<br />

2010. 3D facies <strong>modelling</strong> <strong>of</strong> <strong>carbonate</strong> s<strong>and</strong> bodies:<br />

outcrop analogue study in an epicontinental basin (Triassic,<br />

SW Germany). AAPG Bulletin, 94, 475–512.<br />

Pranter, M. J., Hirstius, C.B.&Budd, D. A. 2005.<br />

Scales <strong>of</strong> lateral <strong>petrophysical</strong> heterogeneity within<br />

dolomite lith<strong>of</strong>acies as determined from outcrop<br />

analogs: Implications for 3-D reservoir <strong>modelling</strong>.<br />

AAPG Bulletin, 89, 645–662.<br />

Pranter, M. J., Reza, Z. A. & Budd, D. A. 2006.<br />

<strong>Reservoir</strong>-scale characterization <strong>and</strong> multiphase fluidflow<br />

<strong>modelling</strong> <strong>of</strong> lateral <strong>petrophysical</strong> heterogeneity<br />

within dolomite facies <strong>of</strong> the Madison Formation,<br />

Sheep Canyon <strong>and</strong> Lysite Mountain, Wyoming,<br />

USA. Petroleum Geoscience, 12, 29–40.


Geological Society, London, Special Publications published online June 27, 2012 as doi: 10.1144/SP370.6<br />

D. PALERMO ET AL.<br />

Qi, L., Carr,T.R.&Goldstein, R. H. 2007. Geostatistical<br />

three-dimensional <strong>modelling</strong> <strong>of</strong> oolite shoals,<br />

St. Louis Limestone; southwest Kansas. AAPG Bulletin,<br />

91, 69–96.<br />

Rankey, E. C., Riegl, B.&Steffen, K. 2006. Form,<br />

function, <strong>and</strong> feedbacks in a tidally dominated ooid<br />

shoal, Bahamas. Sedimentology, 53, 1191–1210.<br />

Ruf, M. 2001. Facies distribution, petrophysics <strong>and</strong><br />

mapping <strong>of</strong> selected <strong>carbonate</strong> s<strong>and</strong> bodies in the<br />

Upper Muschelkalk, South German Basin: a reservoir<br />

analogue investigation. Diploma thesis, University <strong>of</strong><br />

Tübingen.<br />

Ruf,M.&Aigner, T. 2004. Facies <strong>and</strong> poroperm characteristics<br />

<strong>of</strong> a <strong>carbonate</strong> shoal (Muschelkalk, South<br />

German Basin): A reservoir analogue investigation.<br />

Journal <strong>of</strong> Petroleum Geology, 27, 215–239.<br />

Savary, B. & Ferry, S. 2004. Geometry <strong>and</strong> <strong>petrophysical</strong><br />

parameters <strong>of</strong> a calcarenitic turbidite lobe<br />

(Barremian-Aptian, Pas-de-la-Cluse, France). Sedimentary<br />

Geology, 168, 281–304.<br />

Senger, R. K., Lucia, F. J., Kerans, C.&Ferris, M.A.<br />

1991. Dominant control on reservoir-flow behavior in<br />

<strong>carbonate</strong> reservoirs as determined from outcrops<br />

studies. In: Burchfield, T. E. & Wesson, T. C.<br />

(eds) Third International <strong>Reservoir</strong> Characterization<br />

Technical Conference. National Institute for Petroleum<br />

<strong>and</strong> Energy Research <strong>and</strong> US Department <strong>of</strong> Energy,<br />

Tulsa, OK, l, 107–150.<br />

Seyfang, B. 2006. Sedimentary <strong>and</strong> Poroperm Heterogeneities<br />

<strong>of</strong> Carbonate Shoals from Centimetre- to<br />

Kilometre-Scales. An Outcrop Analogue Study For<br />

Skeletal/Oolitic Carbonate S<strong>and</strong> <strong>Reservoir</strong>s in the<br />

Middle East. Diploma thesis, University <strong>of</strong> Tübingen.<br />

Skupin, K. 1969. Lithostratigraphische Pr<strong>of</strong>ile aus dem<br />

Trochitenkalk des Neckar-Jagst-Kocher-Gebietes.<br />

Jahresberichte und Mitteilungen des Oberrheinischen<br />

Geologischen Vereins, 63, 1–173.<br />

Ulrichs, M.&Mundlos, R. 1987. Revision der Gattung<br />

Ceratites de Haan 1825 (Ammonoidea, Mitteltrias),<br />

I.Stuttgarter Beiträge zur Naturkunde, 128, 36.<br />

Ulrichs, M. & Mundlos, R. 1990. Zur Ceratiten-<br />

Stratigraphie im Oberen Muschelkalk (Mitteltrias)<br />

Nordwürttembergs. Jahreshefte der Gesellschaft für<br />

Naturkunde in Württemberg, 145, 59–74.<br />

Vollrath, A. 1938. Zur Stratigraphie und Bildung des<br />

Oberen Hauptmuschelkalks in Mittel- und Westwürttemberg.<br />

Mitteilungen aus dem Mineralogisch-<br />

Geologischen Institut der TH Stuttgart, 33, 69–80.<br />

Vollrath, A. 1955. Zur Stratigraphie des Hauptmuschelkalks<br />

in Württemberg. Jahreshefte des<br />

Geologischen L<strong>and</strong>esamtes Baden-Württemberg, 1,<br />

79–168.<br />

Vollrath, A. 1957. Zur Entwicklung des Trochitenkalks<br />

zwischen Rheintal und Hohenloher Ebene. Jahreshefte<br />

des Geologischen L<strong>and</strong>esamtes Baden-Württemberg,<br />

2, 119–134.<br />

Vollrath, A. 1958. Beiträge zur Paläogeographie des<br />

Trochitenkalks in Baden-Württemberg. Jahreshefte<br />

des Geologischen L<strong>and</strong>esamtes Baden-Württemberg,<br />

3, 181–194.<br />

Vollrath, A. 1970. Ein vollständiges Pr<strong>of</strong>il des oberen<br />

Muschelkalks und ein neues Mineralwasser bei<br />

Ummenh<strong>of</strong>en, Gemeinde Untersontheim, L<strong>and</strong>kreis<br />

Schwäbisch Hall. Jahresberichte und Mitteilungen des<br />

Oberrheinischen Geologischen Vereins, 52, 133–148.<br />

Wagner, G. 1913. Beiträge zur Stratigraphie und Bildungsgeschichte<br />

des Oberen Hauptmuschelkalks und<br />

der Unteren Lettenkohle in Franken. Geologische<br />

und Paläontologische Abh<strong>and</strong>lungen, 2, 31–180.<br />

Warren,J.E.&Price, H. S. 1961. Flow in heterogeneous<br />

porous media. SPE Journal, 1, 153–169.<br />

Ziegler, P. A. 1990. Geological Atlas <strong>of</strong> Western <strong>and</strong><br />

Central Europe, Shell International Petroleum<br />

Maatschappij. 2nd edn. Elsevier, Amsterdam.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!