33
Lithofacies and depth dependency of thermo- and petrophysical rock parameters of the Upper Jurassic geothermal carbonate reservoirs of the Molasse Basin Lithofazies- und Teufenabhängigkeit thermo- und petrophysikalischer Gesteinskennwerte der oberjurassischen geothermischen Karbonatreservoire im Molassebecken Dipl.-Ing. & M.Sc. Sebastian Homuth 1 , Prof. Dr. Annette E. Götz 1,2 , Prof. Dr. Ingo Sass 1 1 Technische Universität Darmstadt, Institute of Applied Geosciences, Geothermal Science and Technology, Schnittspahnstr. 9, 64287 Darmstadt, [email protected], [email protected] 2 co Rhodes University, Geology Department, Sedimentology, P.O. Box 94 Grahamstown 6140, South Africa, [email protected] Keywords: Reservoir characterization, outcrop analogue study, Upper Jurassic, Molasse Basin, thermo-physical parameters, thermofacies, geothermal energy Schlüsselwörter: Reservoircharakterisierung, Aufschlussanalogstudie, Oberer Jura, Molassebecken, thermophysikalische Kennwerte, Thermofazies, Geothermie Abstract In the early stages of hydrothermal reservoir exploration, the thermo-physical characterization of the reservoir is accomplished by evaluating drilling data and seismic surveys. Especially in carbonate reservoirs the distinction of different facies zones and heterogeneities in general is very complex. For economic reasons a sufficiently high flow rate toward the production well and an according high fluid temperature is necessary. For reservoir predictions and modelling, geothermal parameters such as permeability, thermal conductivity/diffusivity, and specific heat capacity have to be quantified. The thermo-physical parameters are facies related. The application of a thermofacies classification to Upper Jurassic limestones serves to understand the heterogeneities and to identify production zones. Outcrop analogue studies enable the determination and correlation of facies related thermo-physical parameters and structural geology data and thus the geothermal exploration concept becomes more precise and quantitative. The analogue outcrops of the Swabian and Franconian Alb represent the target formations of Upper Jurassic carbonate reservoirs in the adjacent Molasse Basin. These limestone formations contain the main flow paths through fractures, faults, and characteristic of

Lithofacies and depth dependency of thermo- and ......facies and thermo - and petrophysical investigations on a low cost basis. The integrated analysis of lithology, facies, and corresponding

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

  • Lithofacies and depth dependency of thermo- and petrophysical rock

    parameters of the Upper Jurassic geothermal carbonate reservoirs of the

    Molasse Basin

    Lithofazies- und Teufenabhängigkeit thermo- und petrophysikalischer

    Gesteinskennwerte der oberjurassischen geothermischen

    Karbonatreservoire im Molassebecken

    Dipl.-Ing. & M.Sc. Sebastian Homuth1, Prof. Dr. Annette E. Götz1,2, Prof. Dr. Ingo Sass1

    1Technische Universität Darmstadt, Institute of Applied Geosciences, Geothermal Science and Technology,

    Schnittspahnstr. 9, 64287 Darmstadt, [email protected], [email protected]

    2co Rhodes University, Geology Department, Sedimentology, P.O. Box 94 Grahamstown 6140, South Africa,

    [email protected]

    Keywords: Reservoir characterization, outcrop analogue study, Upper Jurassic, Molasse Basin,

    thermo-physical parameters, thermofacies, geothermal energy

    Schlüsselwörter: Reservoircharakterisierung, Aufschlussanalogstudie, Oberer Jura, Molassebecken,

    thermophysikalische Kennwerte, Thermofazies, Geothermie

    Abstract

    In the early stages of hydrothermal reservoir exploration, the thermo-physical characterization of the

    reservoir is accomplished by evaluating drilling data and seismic surveys. Especially in carbonate

    reservoirs the distinction of different facies zones and heterogeneities in general is very complex. For

    economic reasons a sufficiently high flow rate toward the production well and an according high fluid

    temperature is necessary. For reservoir predictions and modelling, geothermal parameters such as

    permeability, thermal conductivity/diffusivity, and specific heat capacity have to be quantified. The

    thermo-physical parameters are facies related. The application of a thermofacies classification to

    Upper Jurassic limestones serves to understand the heterogeneities and to identify production zones.

    Outcrop analogue studies enable the determination and correlation of facies related thermo-physical

    parameters and structural geology data and thus the geothermal exploration concept becomes more

    precise and quantitative. The analogue outcrops of the Swabian and Franconian Alb represent the

    target formations of Upper Jurassic carbonate reservoirs in the adjacent Molasse Basin. These

    limestone formations contain the main flow paths through fractures, faults, and characteristic of

  • limestone formations also through karstification. The type and grade of karstification is also facies

    related. In general the matrix permeability has only a minor effect on the reservoir’s sustainability

    except for some grainstones and dolomitized zones with higher porosities and permeabilities.

    Permeabilities range from 10-18 to 10-13 m² (0.001 mD to 100 mD). The permeability range of mud-

    and wackestones is about the same. A high variation of thermo-physical parameters is recognized

    within individual facies zones or stratigraphic units. Mud- and wackestones show thermal

    conductivities around 2 W/(mK), while mudstones have lower thermal conductivities than

    wackestones. The thermal conductivities of massive reefal limestones show values of 1.8 to 3.9

    W/(mK). Secondarily silicified reefal limestones and dolomites show the highest values of thermal

    conductivity. These parameters determined on oven dried samples have to be corrected for water

    saturated rocks under the according temperature and pressure conditions using transfer models. A

    comparison of calculated reservoir properties with measurements from deep drill cores confirms a

    good correlation. Based on the investigation of the matrix parameters in combination with reservoir

    transfer models, the reservoir prognosis and numerical simulation can be improved. The facies related

    characterization and prediction of reservoir formations is a powerful tool for the exploration,

    operation, extension and quality management of geothermal reservoirs in the Molasse Basin.

    Kurzfassung

    In der Planungsphase einer hydrothermalen Reservoirerkundung erfolgt die thermophysikalische

    Charakterisierung des Reservoirs durch die Auswertung von Bohrungsdaten und seismischen

    Erkundungen. Im Falle von Karbonatreservoiren ist die Differenzierung von verschiedenen

    Faziesbereichen bzw. Heterogenitäten im Allgemeinen sehr komplex. Aus wirtschaftlichen Gründen

    ist eine hohe Förderrate der Produktionsbohrung und eine entsprechend hohe Fluidtemperatur

    notwendig. Für eine Reservoirprognose und -modellierung müssen die geothermischen Kennwerte wie

    Permeabilität, Wärmeleitfähigkeit, Temperaturleitfähigkeit und spezifische Wärmekapazität im

    Reservoir quantifiziert werden. Diese thermo- und petrophysikalischen Kennwerte sind

    faziesabhängig. Die Anwendung einer thermofaziellen Klassifikation auf die Karbonate des Oberen

    Jura dient hierbei zum besseren Verständnis der Heterogenitäten und zur Identifikation von

    Produktionszonen. Aufschlussanalogstudien ermöglichen hierbei die Bestimmung faziesabhängiger

    thermophysikalischer Kennwerte und deren Korrelation mit strukturgeologischen Daten, wodurch die

    geothermische Erkundung präziser und quantitativer wird. Die analogen Aufschlüsse der

    Schwäbischen und Fränkischen Alb repräsentieren die Zielformation der oberjurassischen

    Karbonatreservoire des angrenzenden Molassebeckens. Die Hauptzuflüsse in diesen

    Kalksteinformationen erfolgen durch Klüfte, Störungen und charakteristisch für Kalksteine durch

    Karststrukturen. Der Typ und der Grad der Verkarstung ist ebenfalls faziesabhängig. Prinzipiell hat

    die Gesteinspermeabilität sämtlicher untersuchter Karbonate aber nur einen geringen Anteil an der

    hydraulischen Reservoirnachhaltigkeit mit Ausnahme einiger Grainstones und dolomitisierter

  • Bereiche, die höhere Porositäten und Permeabilitäten zeigen. Die Permeabilitäten reichen von 10-18 bis

    10-13 m² (0.001 mD bis 100 mD). Es wurde eine hohe Variabilität der thermophysikalischen

    Kennwerte innerhalb verschiedener Faziesbereiche bzw. stratigraphischer Einheiten festgestellt. Mud-

    und Wackestones zeigen Wärmeleitfähigkeiten um die 2 W/(mK), wobei Mudstones eine geringere

    Wärmeleitfähigkeit als Wackestones bei gleichem Permeabilitätsbereich besitzen. Die

    Wärmeleitfähigkeit der Massenkalke liegt zwischen 1,8 und 3,9 W/(mK). Sekundär silifizierte

    Riffzonen und Dolomite weisen die höchste Wärmeleitfähigkeit auf. Die Kennwerte werden an

    ofengetrockneten Gesteinsproben bestimmt und müssen für entsprechende wassergesättigte

    Reservoirbedingungen korrigiert werden. Dies erfolgt mittels verschiedener Transfermodelle. Ein

    Vergleich von prognostizierten Reservoirkennwerten mit gemessenen Kennwerten an Kernen aus

    Tiefbohrungen zeigt eine gute Übereinstimmung der Prognosewerte mit den tatsächlichen

    Bedingungen im Reservoir. Basierend auf der Untersuchung der Matrixkennwerte in Verbindung mit

    Reservoirtransfermodellen kann sowohl die Reservoirprognose als auch die Grundlage für numerische

    Simulationen verbessert werden. Die faziesbezogene Charakterisierung und Prognose einzelner

    Reservoirhorizonte ist somit ein wirkungsvolles Instrument für die Erkundung, den Betrieb, die

    Erweiterung und Qualitätsüberwachung geothermischer Reservoire im Molassebecken.

    1. Introduction

    To assess the potential and productivity of a hydrothermal and petrothermal reservoir a comprehensive

    and detailed knowledge of the thermo-physical, the hydraulic and mechanical rock and formation

    properties of the system is mandatory. In general, the determination of these reservoir properties is

    limited to time and cost-consuming exploration drillings, which reflect only a small portion of the

    whole reservoir. Especially in terms of carbonates it is difficult to evaluate the heterogeneity of

    different facies zones in seismic sections (Chilingarian et al. 1992), which applies to the Jurassic

    (Malm) target formation of multiple planned geothermal power plant projects in the southern German

    Molasse Basin. These carbonates, characterized by a karst-fractured aquifer system (Schulz et al.

    2012), are located in 3,500 to 5,500 m below surface in the southern part of the Molasse Basin and

    have, like in the Southern German Scarplands, a typical dip of 5° in south-east direction which results

    in an outcrop area north of the Basin (Clauser et al. 2002, Fig. 2). The outcrop analogues of the

    Swabian and Franconian Alb (Fig. 1) represent the reservoir formations and can be used for detailed

    facies and thermo- and petrophysical investigations on a low cost basis. The integrated analysis of

    lithology, facies, and corresponding thermo- and petrophysical rock properties as well as the

    application of relevant reservoir transfer models, leads to an improved prognosis of the reservoir

    properties and conditions. The planning and the geothermal exploration phase of multiple planned

    combined heat and power (CHP) projects in the Molasse Basin can therefore be more cost effective

    and less time consuming. An outcrop analogue study of the target formation Malm (Upper Jurassic)

  • which is the most promising formation for deep geothermal projects in the German Molasse Basin has

    to include facies studies following a thermofacies concept (Sass & Götz 2012).

    Outcrop analogue studies enable the determination and correlation of thermo- and petrophysical

    parameters as well as structural geology data with regional facies patterns. An outcrop analogue

    investigation examines the same rock formations (stratigraphy, lithology, facies) as the potential

    reservoir formations from which fluids at according depth are discharged (Philipp et al. 2007). For the

    economic utilization of deep geothermal reservoirs, a sufficiently high flow rate of thermal waters

    throughout the reservoir to the production well is necessary. This flow rate is mainly controlled by the

    reservoir permeability (Nelson 1985). In the Molasse Basin, the limestone formations of the Upper

    Jurassic contain the main flow paths through tectonic elements such as faults, joints and fractures, and

    characteristic of limestone formations, through karst structures.

    To characterize such fractured reservoirs the orientation of fractures, fracture width, surface roughness

    of fractures as well as the connectivity of fractures and possible secondary mineralization in the

    fracture system is important. These analyses enable the assessment of natural or due to stimulation

    artificially generated reservoir permeability prior to drilling operations. To forecast geometries and

    further properties of existing fracture systems in three dimensions for geothermal appropriate depths,

    the resolution of seismic analyses is not high enough (Philipp et al. 2007). However, in most cases the

    main faults or structural trends, which are generally detectable with seismic surveys, control the

    direction of the fractures. Furthermore the density of fractures increasing towards faults. Therefore, if

    the data quality is sufficient enough this can be indirectly used to model the fracture direction and

    estimate the density of fractures. Also drill cores from exploration drillings are often not available or

    give only a limited set of information. Therefore, it is necessary to obtain thermo-physical and

    structural geological data from outcrops, a well-established exploration tool in oil and gas industry

    (Jahn et al. 2008). Finally, the correlation and the spatio-temporal development of distinct sedimentary

    facies and their thermo-physical parameters may contribute to establish integrated structural 3D

    reservoir models. The thermofacies concept was developed for low and mid enthalpy geothermal

    systems linked to graben settings and deep sedimentary basins located in Central Europe (Sass & Götz

    2012).

    To consider all facies types of the Upper Jurassic in the subsurface of the Molasse Basin, eighteen

    outcrops located in the Swabian and Franconian Alb were investigated to assess the whole accessible

    stratigraphic section of the Upper Jurassic (Malm) spanning Malm α to Malm ζ3 strata (Fig. 1, Fig. 4).

    The upper Malm ζ4-6 is almost completely eroded and only known from a few deep drillings in the

    Molasse Basin. Drill cores of two shallow and one deep drilling (Moosburg SC4) were investigated to

    cover the uppermost strata of the Malm formation. The aim is to identify lithofacies-related trends of

  • thermo- and petrophysical reservoir parameters. The results and inferred predictions from the

    laboratory testing can be validated with drilling and pump test data from recent drill sites.

    This paper gives a brief summary of the regional geological setting of the study area, including the

    main facies zones and their diagenesis. Subsequently the outcrop studies and laboratory methods for

    the determination of rock properties are introduced. Next, the thermofacies concept and facies-related

    rock properties are presented. Then, the results and their transferability to reservoir conditions are

    discussed by evaluating data from deep drill cores. In addition facies-controlled karstification

    structures and the tectonic setting is presented and discussed. Finally, the application and access of

    data is summarized.

    2. Geological Setting

    During the Mesozoic the study area was part of a broad epicontinental sea north of the Tethys Ocean

    (Fig. 3). To the north the shelf area was confined by an island archipelago of changing dimensions. In

    the southern, deeper part of this epicontinental sea, an extensive siliceous sponge-microbial reef belt

    developed. With the burial of the Vindelician Ridge a direct connection of the south German Jurassic

    Sea with the Tethys Ocean was established (Meyer & Schmidt-Kaler 1989). During the entire Upper

    Jurassic a high carbonate production on the shallow shelf resulted in thick limestone series (Selg &

    Wagenplast 1990). In the southern, deeper part of this epicontinental sea, a reefal facies, established in

    the Middle Oxfordian, and was part of a facies belt characterized by silicious sponge reefs spanning

    the northern Tethys shelf (Pieńkowski et al. 2008).

    The main reef building organisms are microbial crusts, however, coral reefs are also present and

    becoming increasingly important towards the end of the Upper Jurassic, mirroring the overall

    shallowing sea level (Pieńkowski et al. 2008). In addition, clay-rich sediments from the Mid German

    Swell were shed into the shelf area. During times of low carbonate production, the clay content of the

    sediments increased which resulted in the sedimentation of marl (Meyer & Schmidt-Kaler 1990).

    According to Meyer and Schmidt-Kaler (1989, 1990), the Swabian facies as the central part of the reef

    belt formed a deeper-water area between the shallower Franconian-Southern Bavarian platform in the

    East and the Swiss platform in the West. In the Southwest, the Swabian shelf facies deepens gradually

    towards the pelagic facies of the Helvetian Basin. The Helvetic facies is characterized by dense,

    bituminous limestones with in places interbedded oolith layers. This facies describes the transition of

    the Germanic facies into the Helvetic facies, which is considered as sediments of a deeper shelf area of

    bedded limestones with very low permeabilties. Also karstification is not observed, so the northern

    boundary of the Helvetic facies is considered as the southern boundary of the Malm aquifer of the

    Molasse Basin (Villinger 1988).

  • In general 400 to 600 m of carbonate rocks were deposited during the Upper Jurassic and two major

    lithofacies-types can be distinguished (Geyer and Gwinner 1979, Pawellek & Aigner 2003):

    (1) a basin facies, consisting of well-bedded limestones and calcareous marls (mud-/wackestones), and

    (2) a reefal or massive facies, when bedding is either absent, indistinct or very irregular (rud-/float-

    /grainstones).

    The massive limestones are built by microbial crusts (stromatolites and thrombolites) and siliceous

    sponges that have been interpreted by various authors as relatively deep and quiet water ‘‘reefs’’,

    mounds or bioherms (Gwinner 1976, Leinfelder et al. 1994, 1996, Pawellek & Aigner 2003). The

    basin facies may either interfinger with the reefs or onlap onto the reefs (Gwinner 1976; Pawellek

    2001). In the upper parts of the Upper Jurassic, a coral facies developed locally upon the microbial

    crust-sponge reefs. The abundance of reef facies differs regularly through time. Reef expansion phases

    correlate with an increase in the carbonate content within the basin facies, while phases of reef retreat

    correlate with increasing abundance of marls within the basin facies (Meyer and Schmidt-Kaler 1989,

    1990, Pawellek 2001). A comparison of the different lithostratigraphic units in the Franconian and

    Swabian Alb is given in Figure 4.

    3. Materials and Methods

    The investigations are carried out on three different scales: (1) The macro scale including an outcrop

    mapping to detect the lithotypes, structural elements and facies patterns in the outcrop; (2) the meso

    scale, to determine thermo- and petrophysical properties of different lithotypes in the laboratory of

    representative rock samples; and (3) the micro scale, to analyze microstructures, cements, porosities,

    etc. in thin sections. This approach is based on the concept of Aigner (1985) and was applied to the

    Upper Jurassic by Schauer (1998), Aigner & Schauer (1998) and Pawellek & Aigner (2004).

    The rock samples were classified in different facies groups after the scheme developed by Pawellek

    (2001). The following simplifications have to be considered: Pawellek (2001) differentiates within the

    massive reefal limestones crust-dominated and sponge-dominated rocks. This detailed differentiation

    is for the thermo- and petrophysical property investigation not necessary, because no significant

    difference is observed in terms of thermo- and petrophysical properties of these rock types. In

    addition, a simplified facies classification (Tab. 1, Fig. 7) with respect to the dominant reservoir facies

    types is necessary, matching the resolution of reservoir formations by geophysical exploration

    methods. According to the Dunham (1962) and Embry & Klovan (1971) classification of carbonate

    rocks the following facies types are detected in the study area: mudstones, wackestones, grain-

    /packstones and float-/rudstones. The rock classification is based on previous studies of the Malm

    formations in Southern Germany (Schauer 1998, Pawellek 2001).

  • Reservoir characterization based on thermo-and petrophysical parameters including permeability and

    thermal conductivity data is rarely available measured at the same sample. For direct correlation in this

    study all parameters are determined at the same sample. More than 350 rock samples from 17 outcrop

    locations as well as shallow and deep core drillings in Baden-Wuerttemberg and Bavaria (Fig. 2) are

    collected and analyzed. For statistical purposes on each rock sample 3 to 10 single measurements of

    different rock properties were conducted, i.e. in total over 1150 measurement pairs or triples are

    collected.

    To determine the thermo-physical properties of the sampled formations and to generate reproducible

    results the samples were dried at 105°C to mass constancy and afterwards cooled down to 20°C in an

    exsiccator. A thermal conductivity scanner (optical scanning method after Popov et al. 1985) and a gas

    pressure permeameter (Jaritz 1999, Hornung & Aigner 2004) and porosimeter are used. With a

    thermal conductivity scanner the thermal conductivity and diffusivity are determined. The

    measurement is based on a contact free temperature measurement with infrared temperature sensors

    (Bär et al. 2011). The measurement accuracy is stated by the manufacturer of about 3 %. The

    determination of the grain and bulk density as well as the porosity was done by measuring the grain

    and bulk volume of the samples, using a helium pycnometer and a powder pycnometer. The specific

    heat capacity cp was calculated with the measured thermal conductivity λ, density ρ and thermal

    diffusivity α (converted Debye-Equation):

    𝑐𝑐𝑝𝑝 =𝜆𝜆

    𝜌𝜌 ∙ 𝛼𝛼 Eq. 1

    For the determination of the rock permeability a combined column and mini permeameter was used.

    The method offers either the measurement of the apparent gas permeability which afterwards is

    converted in permeability or the direct measurement of the intrinsic permeability. The basis for the gas

    driven permeameter is the Darcy law which is enhanced by the terms of compressibility and viscosity

    of gases. The maximal measurement error depending on the measurement setting is 10% while the

    minimal error is about 1%.

    4. Results and Discussion

    4.1 Thermofacies

    The matrix permeability of all measured carbonates is quite low except for some grainstones with

    higher permeabilities and porosities. Thick-bedded and platy limestones show thermal conductivities

    around 2 W/mK, characteristic of limestones. Permeabilities range from 10-18 to 10-13 m² (K in [m²] =

    K in [D] · 9.8692 10-13). Marly limestones have lower thermal conductivities than thick-bedded and

    platy limestones, showing the same range of permeabilities as the thick-bedded limestones. It seems

  • that the higher clay content of the marly limestones is decreasing the thermal conductivity by

    insulating the heat conduction and at the same time showing only minor effects on permeability which

    shows the same range for mud- and wackestones. The thermal conductivities of different reefal

    limestones have values of 1.8 to 3.9 W/(mK), related to the higher content of secondarily silicified reef

    bodies and due to dolomitization of reefal structures. The layers with increased silica content are

    identified as silicified sponge layers (Leinfelder et al. 1994, 1996). The dolomitized carbonates show

    the highest values of thermal conductivity of all investigated carbonates in this study. The herein

    presented thermal conductivity and porosity values were in good accordance to results reported from a

    limited number of samples in recent works (Koch et al. 2007). However, this study focuses on the

    facies relation of thermo- and petrophysical rock parameters of Malm carbonates with respect to

    geothermal reservoir characterization. Figure 5 shows that for some stratigraphic units trends are

    detectable (increasing thermal conductivity from Malm α to Malm ζ, due to decreasing clay content

    and increasing dolomitization (the maximum of the dolomitization is also found within the Malm δ by

    Schauer (1996, 1998)). The peak of thermal conductivity observed in the Malm δ also correlates with

    increased silica content supplied by silicified sponge layers). Thermo-physical correlations between

    different reservoir properties are controlled by lithofacies types. A thermofacies classification based on

    the rock parameters thermal conductivity and permeability and the related facies types is shown in

    Figure 6, including the nomenclature after Pawellek (2001).

    Diagenetic processes caused dolomitization and de-dolomitization of reef structures and their adjacent

    transition zones to the basin facies, resulting in an increase of inter-crystalline porosity and therefore

    increased matrix permeability. On the other hand, if de-dolomitization led to the formation of

    saccharoidal limestone, permeability is decreasing due to reduced crystalline porosity.

    With increasing dolomite content an increase of thermal conductivity is observed due to the higher

    thermal conductivity of the dolomite crystal structure. The dolomitized areas related to the geometry

    of the massive reefal limestone complexes can span over several stratigraphic units of the Malm

    predominantly in vertical direction (Schauer 1998, Koch 2011). In the study of Wolfgramm et al.

    (2011) an overview and estimation in percentage of the dolomitized areas of the Purbeck and Malm is

    given. Figure 7 includes the measurement ranges of different thermo- and petrophysical matrix

    properties to improve the reservoir assessment.

    Based on the Debye-Equation (eq. 1) and the results of measurements on deep drill cores (Fig. 11), it

    can be inferred that the thermophysical properties for tight carbonate rocks are density controlled (see

    also section 5.2 Transferability). Density itself is strongly dependent on the lithofacies of the carbonate

    rock, so the massive and basin facies has direct influence on the formations’ hydraulic conductivity. In

    the transition zone of basin facies to reef facies sub-vertical fractures caused through differential

    compaction between massive facies and adjacent basin facies can also be observed in the studied

    outcrops. Due to the increased fracture density in this zone, the karstification process is favored, which

  • results in dissolution of carbonate. The increased hydraulic conductivity results in the disintegration

    into dolomite sand or in the process of de-dolomitization (re-calcification) either. In this context it is

    important to consider that zones of dolomites due to their primary facies interfingering and genesis

    also on a small scale are laterally variable and developed across fractures and porous zones into

    adjacent facies (Koch 2011). Therefore, the identification and location of dolomitized areas is of

    special interest for the geothermal reservoir prognosis in terms of hydraulically prospective reservoir

    formations.

    The grain density of the outcrop samples ranges between 2.59 and 2.84 g/cm³ (Fig. 8), the bulk density

    is between 2.31 and 2.75 g/cm³. Dolomites showing typically higher densities than limestones, while

    dolomitized limestones also elevate the grain density of effected well bedded and reefal limestones.

    The porosity calculation based on these values is accordingly less than 15%. The massive limestones

    have porosities less than 8%, grainstones and dolomitized zones showing increased porosities up to

    15%. From these measurements (Fig. 9) it can also be inferred that the lower Malm α to γ shows very

    poor hydraulic conductivity due to average porosities around 5% or less and the upper Malm δ to ζ has

    poor to good hydraulic conductivity due to a considerable share of extreme values which are showing

    porosities up to 18%. These high porosity values are often correlated to increased permeability of

    grain- and dolostones in the upper Malm strata. This observation also correlates with pump test data

    from deep drill sites in the Molasse Basin (Schulz et al. 2012).

    The permeability measurements state in general very low matrix permeabilties. Only grainstones,

    reef/coral debris limestones and dolomitized zones show higher permeability ranges up to 10-14 m² (10

    mD). A comparison of permeability and porosity indicates that high porosities occur in grain- and

    dolostones and also cause higher permeability. But for all other lithofacies types no correlation

    between porosity and permeability in regard to interconnected porosity can be inferred (Fig. 10).

    Compared with other carbonate reservoirs mainly explored for hydrocarbons (Ehrenberg und Nadeau,

    2005), the rock permeability and porosity of the Malm carbonates of the Molasse Basin is rather low.

    Including fracture network, dolomitization and karstification, a positive shift of the permeability-

    porosity relationship across several magnitudes can be observed (Fig. 11). This shift identifies a high

    hydrothermal potential of the deep Malm aquifer system in the Molasse Basin.

    The assumption of a positive 2-3 magnitude reaching permeability correction is based on pump test

    data and comparisons of matrix and formation productivity from different deep drilling locations in the

    Molasse Basin (Böhm et al. 2011, Schulz et al. 2012).

  • 4.2 Transferability of Reservoir Characterization Parameters

    The presented data of outcrop analogue studies are based on rock measurements on oven dried cores,

    which are conducted under laboratory conditions with atmospheric pressure and room temperature of

    20°C. This approach guarantees a very good reproducibility of the results but also requires a correction

    of the measured data for reservoir conditions. It is assumed that the reservoir is completely saturated.

    For the following analyses the temperature and pressure conditions of a 5000 m deep and 150°C hot

    reservoir, which are realistic values for the Molasse Basin (Fig. 12), are estimated.

    The thermal conductivity of water saturated rocks can be calculated following the model of

    Lichtenecker and numerous other authors (Pribnow 1994, Clauser & Huenges 1995, Pribnow & Sass

    1995, Williams et al. 1995, Schoen 1996, Popov et al. 2003, Hartmann et al. 2005, u. a.). The water

    saturated geometric mean thermal conductivity (λgeo) is calculated with the thermal conductivity of the

    rock matrix (λrock), the known or assumed thermal conductivity of the pore space filling fluid (λfluid),

    and the porosity (Φ):

    𝜆𝜆𝑔𝑔𝑔𝑔𝑔𝑔 = 𝜆𝜆𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝛷𝛷 ∙ 𝜆𝜆𝑟𝑟𝑔𝑔𝑟𝑟𝑟𝑟1−𝛷𝛷 Eq. 2

    Respectively, the calculation of the specific heat capacity under water saturated conditions is done

    using eq. 1.

    For the temperature and pressure correction numerous transfer models for reservoir parameters exist.

    Temperature dependency models of thermo-physical properties of different rock types can be found in

    Tikhomirov (1968), Huenges et al. (1989), Somerton (1992), Pribnow (1994), Vosteen &

    Schellschmidt (2003), and Abdulagatova et al. (2009). All authors provide different correction

    functions based on comparisons of their own laboratory tests and published data. In general the

    thermal conductivity decreases with increasing temperature and increases with increasing pressure

    (Clauser & Huenges 1995). The fundamental effects are the reduction of pore space and the increasing

    temperature with increasing depth (Clauser et al. 2002). Both parameters control the fluid and matrix

    conditions, although in terms of carbonates the temperature depending porosity reduction is the

    dominant factor. Also for carbonates the lithostatic pressure has only minor to negligible influence on

    the porosity-permeability relation. The pressure depending increase of thermal conductivity is only

    relevant for unconsolidated rocks.

    For large-scale models of the upper earth crust and also for deep geothermal systems the function after

    Zoth & Hänel (1988) for the correction of the thermal conductivity for different rock types is

    commonly used, where λ(T) is the thermal conductivity at temperature T:

    𝜆𝜆(𝑇𝑇) = 𝐴𝐴350+𝑇𝑇

    + 𝐵𝐵 Eq. 3

  • In this equation specific petrophysical coefficients are applied. For the Malm carbonates the

    coefficients A = 0.13 and B = 1073 valid for a temperature range of 0 – 500°C are chosen. The

    correction function of Vosteen & Schellschmidt (2003) is also valid for a temperature range of 0 –

    500°C and the rock specific coefficients for carbonates are a = 0,0034 +/- 0,0006 and b = 0,0039 +/-

    0,0014.

    𝜆𝜆0 = 0.53 ∙ 𝜆𝜆25 +12�1.13 ∙ (𝜆𝜆25)2 − 0.42 ∙ 𝜆𝜆25 Eq. 4

    𝜆𝜆(𝑇𝑇) = 𝜆𝜆00.99+𝑇𝑇∙�𝑎𝑎−𝑏𝑏𝜆𝜆0

    � Eq. 5

    Here λ0 is the thermal conductivity at 0°C, λ25 is the thermal conductivity at 25°C (laboratory

    temperature) and T is temperature in °C.

    The model of Sass et al. (1971) is given in equation 6 and 7:

    𝜆𝜆(𝑇𝑇) = 𝜆𝜆01.007+𝑇𝑇 �0.0036−0.0072𝜆𝜆0

    � Eq. 6

    𝜆𝜆(0) = 𝜆𝜆25 ∙ �1.007 + 25 ∙ �0.0037−0.0074𝜆𝜆25

    �� Eq. 7

    With the parameters as defined above the heat capacity is inversely proportional (cf. eq. 1) to the

    thermal conductivity and increases with increasing temperature. The thermal diffusivity depends on

    the thermal conductivity, density and heat capacity and therefore decreases with increasing

    temperature and increasing pressure. After Vosteen & Schellschmidt (2003) the thermal conductivity

    of rocks decreases in a temperature interval of 1 to 300°C about 35 ± 10 %, while the thermal

    diffusivity in the same temperature range decreases about 48 ± 6 %, showing that this process is

    mainly temperature controlled. The same order of magnitude is stated by other authors as well

    (Somerton 1992, Clauser et al. 2002). Table 2 shows the range of measured values and calculated

    transfer values of different thermo-physical rock properties for different facies types of the Malm

    carbonates. For a comparison also the calculated reservoir rock properties with respect to the

    mentioned reservoir conditions and according correction functions are listed.

    In terms of matrix porosity and permeability it is concluded that the low rock porosity measured on the

    outcrop samples will not change significantly with increasing depth in regards to hydraulic

    conductivity. In terms of the mean reservoir porosity the temperature of the carbonate systems is the

    dominant factor with regard to the thermal expansion and carbonate chemistry and not the depth

    function of the reservoir (Schmoker 1984, Bjørkum & Nadeau 1998). A comparison with other

    carbonate reservoirs confirms this approach (Ehrenberg & Nadeau 2005).

    Measurements of thermal conductivity and permeability of drill cores from greater depths (up to 1600

    m, drilling: Moosburg SC4) in the Molasse Basin are shown in Figure 13. The measurements confirm

  • the above mentioned assumptions and correction functions applied on the near surface values. In terms

    of permeability no significant change with depth is inferable. The permeability values obtained from

    cores in depth of 1600 m show typical values of permeability comparable to the value range of outcrop

    samples. Due to the slight temperature difference of maximum 40°C the thermal conductivity shows

    only minor to negligible differences compared with the outcrop results, except for the depth range of

    1300 to 1500 m where a dolomitized area of massive facies in the Malm ζ1 and ζ2 is encountered. The

    dolomitization process results here in a significant increase (up to 2 W/(mK)) of thermal conductivity.

    In general the thermophysical matrix parameters seem to be density controlled. Density increases with

    depth typically and with dolomite content especially, therefore thermal conductivity increases but also

    decreases with increasing temperature, which is the dominant factor. Specific heat capacity increases

    with depth and temperature (Tab. 2).

    4.3 Diagenesis

    The Malm and Purbeck carbonates are diagenetically altered. In terms of the Molasse Basin three

    phases are known: (1) the marine-phreatic early diagenesis, (2) the shallow burial diagenesis with (2a)

    a hydrothermal phase and (3) the meteoric karstification diagenesis (Reinhold 1998) and deeper burial.

    (1) In the early-diagenesis phase calcitic cements and microbiological pyrite was built.

    Furthermore, the silicious yielding skeletons of sponges were dissolved (Reinhold 1998) and

    afterwards lithified by different carbonate-based cements.

    (2) During the shallow burial first fractures, grain margin dissolution and styloliths occur

    (Wolfgramm et al. 2011). Next, dolomitization of the matrix starts from fractures and

    styloliths (Schauer 1996, 1998). During the burial phase the carbonate partly dolomitized

    completely, so that diagenesis generated dolomites, which have the same petrophysical

    behaviour than grainstones (increased porosity and permeability). A selective early diagenetic

    dolomitization occurred which was related to primary facies with probably increasing

    alteration of the pore water chemistry (Liedmann & Koch 1990). Dolomite was formed in

    residual pore spaces which were not closed during early diagenetic cementation-phases of the

    sponge-algae reefs. The basinal facies dolomitization is associated with pre-formed pressure-

    solution seams. The dolomitization fluids percolated predominantly through relatively

    permeable packstones/floatstones and grainstones of the sponge-algal bioberms (Liedmann &

    Koch 1990). Within these massive dolomites secondarily formed inter-crystalline pore space

    is detected (Wolfgramm et al. 2011). Also during the burial the Ca-rich matrix dolomites can

    be altered into Ca-poor dolomites. After tectonic phases (2a) dissolution (karstification) along

    fractures occurred. Along the fractures Ca-rich dolomitic cements are attached. Due to the

  • supply of CO2-rich fluids these dolomitic cements were dissolved again. Afterwards blocky

    calcidic cements precipitated.

    (3) During emersion of the Malm formation related to the Alpine orogenesis, the effects of

    meteoric and hydrothermal waters became more dominant. De-dolomitization and

    karstification of the matrix dolomites started from the fracture network. Evidence for long-

    term karstification can be found in the increasing opening width of fractures caused by

    dissolution (Reinhold 1998) and karst structures in great depth observed in seismic sections

    (Lüschen et al. 2011). Especially in fault zones increased opening widths of fractures and karst

    structures, which were secondarily filled by bitumen, are recognized (Wolfgramm et al. 2011).

    Due to different diagenetic processes and depending on primary porosity conditions, the intensity of

    dolomitization varies in the different facies types and can result in substantially increased rock

    permeability (cf. Böhm et al. 2013).

    4.4 Tectonics and karstification

    The present study shows that the rock permeability of the investigated carbonates in general is very

    low. Thus, it is concluded that no significant matrix supported hydraulic influence on the short-term

    production capacity of the Malm aquifer can be inferred, except for grainstone or dolomitized facies

    zones, which are often associated with fault zones. But in terms of the long term sustainability of the

    total reservoir system the matrix potential has to be considered. Accordingly, numerical simulations

    under consideration of facies-related variations of the matrix potential should be applied (cf. Schulz et

    al. 2012), in which the compiled data of this study can be an important contribution. However, in

    places the Malm aquifer is very productive, producing discharge rates of over 100 l/s. So the main

    hydraulic flow paths must be related to dense hydraulically open fracture networks and associated

    karstification. By looking at karst structures two general facies-related karstification types can be

    observed (Fig. 14): (1) In the case of thick-bedded or platy limestone (basin-) facies (mudstone,

    wackestone) karstification occurs only along major fractures and faults; (2) in the case of reefal

    limestones and bioclastic limestones (grainstone, rud-/floatstone) the karstification occurs on a larger

    scale due to the primary and secondary porosity of these facies.

    In addition, the above mentioned processes of dolomitization and de-dolomitization can also cause

    considerably higher rock permeabilities. While the massive limestones are in general favorable for

    karstification, the basin facies or thick to platy limestones show a karstification resistance due to their

    significantly higher clay content (Wolfgramm et al. 2011), so that CO2-rich groundwater can only

    dissolve small amounts of carbonate.

  • According to Lüschen et al. (2011) karstification in the region north of Munich is observed in seismic

    sections. For the following statements it is assumed that karstification is at least in the north eastern

    part of the Molasse Basin widespread throughout the aquifer and its intensity and characteristics

    depend on facies zones. During the Purbeck a first karstification phase occurred, especially in the

    northern part of the Molasse Basin. This part is characterized by the absence of Purbeck sediments, so

    meteoric waters could enter the Malm carbonates easily and in depth. The waters migrated also to the

    south, where the karstification intensity decreases (Birner et al. 2011). During the orogenesis of the

    Alps the south German Massif was tilted. In the north the Jurassic rock formations are exposed and in

    the south pushed under the Alps. Large-scale stress analyses show that in the Molasse Basin

    transtensive to strike-slip structures are dominant. Due to the N-S aligned main stress axis and the

    shape of the Molasse Basin also sinistral NE-SW striking transtension and transpression faults are

    locally observed (Reinecker et al. 2010, Fig. 15).

    As an example the main fault structure in the 3D seismic exploration of the Unterhaching region is

    interpreted as a sinistral transtension regime with according sinistral transpression regime (Lüschen et

    al. 2011). Attached to the major faults often Riedel shears are observed. These tectonic structures

    cause a disaggregation of the formation in the E-W and NE-SW striking related major fault zones in

    the central Molasse Basin. Due to their development in a transtensive regime adjacent faults are

    typically formed at the deep block which can also be confirmed by seismic profiles and according

    structural analyses carried out in the Molasse Basin (Wolfgramm et al. 2011). These faults zones were

    also active for hydrothermal fluids during different diagenesis phases (Liedmann 1992). This caused

    an intensive dissolution, extension and dolomitization process in the vicinity of the fault zones and

    therefore leads to potentially higher hydraulic conductivities and permeabilities, respectively.

    In terms of geothermal exploration more than 50 well bores in the Molasse Basin confirm that the

    facies pattern is of comparable importance as tectonics in terms of the intensity of karstification and

    thus also for the formation permeability (Schulz et al. 2012). Inferred from the spatial distribution of

    total circulation losses of hydrocarbon and hydrothermal drillings in the karst and fracture aquifer of

    the Malm, the karstification along major fault zones with adjacent fracture networks is due to their

    improved hydraulic conductivity considerably increased (Böhm et al. 2013). The tectonical setting,

    karstification and adjacent dolomitization are the key factors for a successful geothermal exploration

    in the Molasse Basin, as they are controlling the flow rates of the Upper Jurassic carbonates.

    5. Conclusions

    The obtained data from the Upper Jurassic limestones of South Germany show that prognosis of

    reservoir properties by applying facies models to the deeper subsurface can be implemented as an

    additional exploration tool. The determination of geothermal reservoir properties serves in general to

  • distinguish between petrothermal and hydrothermal systems (Sass & Götz 2012) and can also be used

    for optimized drilling and stimulation design purposes. Outcrop analogue studies offer effective

    opportunities to gain data to be predicted via transfer models to greater depths and higher

    temperatures, which lead to a better understanding of production capacities of geothermal reservoirs.

    Furthermore, these studies provide a sufficient data base to determine thermo-physical reservoir

    characteristics of the rock matrix of geothermal reservoir formations. Facies concepts are applied as

    exploration tool producing conservative results. Adding information on secondary porosities,

    karstification, dolomitization and stress field into a reservoir model will lead to realistic reservoir

    capacities.

    The studied rocks of the Upper Jurassic are not a homogenous formation of limestones. Even on a

    small scale different facies zones and their interfingering, which can be differentiated in geometry,

    structure, fabric and composition, can be identified. These differences affect the thermo-physical

    properties of the rocks and show facies related trends. The hydraulic parameters vary in the order of 4

    magnitudes within a stratigraphic unit or facies zone, but show in general a range of poor to very poor

    hydraulic conductivity according to the classification of the DIN 18130. It can be inferred from the

    outcrop studies that hydraulic active pathways are bound to fracture networks, faults and adjacent

    karstification and/or dolomitized zones. The secondary reservoir permeability is strongly related to the

    tectonic setting and history and additionally depending on the hydro chemical conditions of the

    carbonate reservoir, which also is facies controlled, to maintain open flow paths. Based on the

    investigation of the matrix parameters the sustainable heat transport into the utilized geothermal

    reservoir can be assessed. Thus, the long-term capacities for different utilization scenarios can be

    calculated more precisely. With the help of 3D seismic surveys the investigations on the lateral

    extension and facies heterogeneity will give insight on the transmissibility of different target horizons.

    The thermofacies characterization and prediction of geothermal reservoir parameters is therefore a

    powerful tool for the exploration, extension, operation and quality management of planned and

    existing geothermal reservoirs, respectively. The key to a reliable reservoir prognosis, reservoir

    stimulation, and sustainable reservoir utilization for the Malm in the Molasse Basin is to integrate

    statistically tested databases of tectonic, hydraulic and thermofacies models into 3D reservoir models.

    The presented geothermal rock parameters for the different lithofacies can be used to improve

    numerical simulations and 3D models of different project locations in the Molasse Basin. The database

    will be available via the Fachinformationssystem Geophysik (FIS-GP) of the Leibniz Institute of

    Applied Geophysics (LIAG).

  • 6. Acknowledgements

    We thank the Bayerisches Landesamt für Umwelt (LfU) for the support and access to drill cores from

    shallow and deep research drillings for the investigation of physical rock parameters. Furthermore, we

    acknowledge the contribution of numerous students, who performed measurements of geothermal rock

    parameters in the framework of their theses at the Institute of Applied Geosciences of the Technische

    Universität Darmstadt. We kindly acknowledge the helpful and constructive comments and remarks of

    Sonja Philipp, University of Göttingen, and two anonymous reviewers which greatly improved the

    manuscript.

    7. References

    Abdulagatova, Z., Abdulagatov, I.M., Emirov, V.N. (2009): Effect of temperature and pressure on the

    thermal conductivity of sandstone. - International Journal of Rock Mechanics & Mining Sciences,

    46: 1055-1071.

    Aigner, T. (1985): Storm depositional systems: dynamic stratigraphy in modern and ancient shallow-

    marine sequences. – Lecture Notes in Earth Sciences Series, Vol. 3, 174 p. Berlin (Springer).

    Aigner, T. & Schauer, M. (1998): Exploration for industrial minerals and rocks using "dynamic

    stratigraphy": example ultra-pure limestones. - Z. angew. Geol., 44: 159-163.

    Bär, K., Arndt, D., Fritsche, J.-G., Götz, A.E., Kracht, M., Hoppe, A. & Sass, I. (2011): 3D-

    Modellierung der tiefengeothermischen Potenziale von Hessen: Eingangsdaten und

    Potenzialausweisung. – Z. Dt. Ges. Geowiss., 162 (4): 371-388.

    Bertleff, B. W., Joachim, H., Koziorowski, G., Leiber, J., Ohmert, W., Prestel, R., Stober, J., Strayle,

    G., Villinger, E., Werner, W. (1988): Ergebnisse der Hydrogeothermiebohrungen in Baden-

    Württemberg. - Jb. Geol. L-Amt Baden-Württemberg, 30: 27-116.

    Birner, J., Mayr, C., Thomas, L., Schneider, M., Baumann, T. & Winkler, A. (2011): Hydrochemie

    und Genese der tiefen Grundwässer des Malmaquifers im bayerischen Teil des süddeutschen

    Molassebeckens. - Z. geol. Wiss., 39: 291-308.

    Bjørkum, P.A. & Nadeau, P.H. (1998): Temperature controlled porosity/permeability reduction, fluid

    migration, and petroleum exploration in sedimentary basins. - Australian Pet. Prod. & Expl. Assoc.

    Journal, 38, 453-464.

    Böhm, F., Birner, J., Steiner, U., Koch, R., Sobott, R., Schneider, M., Wang, A. (2011): Tafelbankiger

    Dolomit in der Kernbohrung Moosburg SC4, Ein Schlüssel zum Verständnis der Zuflussraten in

    Geothermiebohrungen. – Z. geol. Wiss., 39: 117-157.

    Böhm, F., Savvatis, A., Steiner, U., Schneider, M., Koch, R. (2013): Lithofazielle

    Reservoircharakterisierung zur geothermischen Nutzung des Malm im Großraum München.

    Grundwasser, 18 (1): 3-13.

    http://link.springer.com/journal/767/18/1/page/1

  • Chilingarian, G. V., Mazzullo, S. J., Rieke, H. H., (1992): Carbonate Reservoir Characterization: A

    Geologic-Engineering Analysis. - Elsevier Sci. Publs. B.V., 639 p, Amsterdam (Elsevier).

    Clauser, C. & Huenges, E. (1995): Thermal Conductivity of Rocks and Minerals. – Rock Physics and

    Phase Relations, A Handbook of Physical Constants. - AGU Reference Shelf, 3: 105-126.

    Clauser, C., Deetjen, H., Höhne, F., Rühaak, W., Hartmann, A., Schellschmidt, R., Rath, V. &

    Zschocke, A. (2002): Erkennen und Quantifizieren von Strömung: Eine geothermische

    Rasteranalyse zur Klassifizierung des tiefen Untergrundes in Deutschland hinsichtlich seiner

    Eignung zur Endlagerung radioaktiver Stoffe. – Endbericht zum Auftrag 9X0009-8390-0 des

    Bundesamtes für Strahlenschutz (BfS), Applied Geophysics and Geothermal Energy E.ON Energy

    Research Center, 159 S., Aachen (RWTH ).

    DIN 18130-1 (1998): Baugrund - Untersuchung von Bodenproben – Bestimmung des

    Wasserdurchlässigkeitsbeiwertes - Teil 1: Laborversuche, Beuth-Verlag.

    Dunham, R. J. (1962): Classification of carbonate rocks according to depositional texture. - In: Ham,

    W.E. (ed.): Classification of carbonate rocks. - AAPG Memoir, 1: 108-171.

    Ehrenberg, S. N. & Nadeau, P. H. (2005): Sandstone versus carbonate petroleum reservoirs: a global

    perspective on porosity-depth and porosity-permeability relationships. - AAPG Bulletin, 89: 435-

    445.

    Embry, A. F. & Klovan, J. E., (1971): A Late Devonian reef tract on northeastern Banks Island,

    N.W.T. - Canadian Petroleum Geology Bulletin, 19: 730-781.

    Geyer, O. F. & Gwinner, M. P. (1979): Die Schwäbische Alb und ihr Vorland. - Sammlung geol.

    Führer, 271 S., Stuttgart-Berlin (Gebrüder Borntraeger).

    Geyer, O. F.& Gwinner, M. P. (1991): Geologie von Baden-Württemberg. - 482 S., Stuttgart

    (Schweizerbart).

    Gwinner, M. P. (1976): Origin of the Upper Jurassic of the Swabian Alb.- Contrib. Sedimentol., 5: 1 –

    75.

    Hartmann, A., Rath, V. & Clauser, C. (2005): Thermal conductivity from core and well log data. -

    International Journal of Rock Mechanics & Mining Science, 42: 1042-1055.

    Hornung, J. & Aigner, T. (2004): Sedimentäre Architektur und Poroperm-Analyse fluviatiler

    Sandsteine: Fallbeispiel Coburger Sandstein, Franken. – Hallesches Jahrb. Geowiss., Reihe B (18):

    121–138.

    Huenges, E., Buntebarth, G., Zoth, G. (1989): Die Wärmestromdichte in der KTB-Oberpfalz. - VB

    KTB Report 89-3: B404.

    Jahn, F., Graham, M., Cook, M. (2008): Hydrocarbon Exploration & Production. - Vol. 55, Second

    Edition (Developments in Petroleum Science), 470 p., Philadelphia (Elsevier Science).

    Jaritz, R. (1999): Quantifizierung der Heterogenität einer Sandsteinmatrix am Beispiel des

    Stubensandstein (Mittlerer Keuper, Württemberg). - Tübinger Geol. Abhandlungen, Reihe C (48):

    104.

  • Koch, A., Hartmann, A., Jorand, R., Mottaghy, D., Pechnig, R., Rath, V. Wolf, A. & Clauser, C.

    (2007): Erstellung statistisch abgesicherter thermischer und hydraulischer Gesteinseigenschaften

    für den flachen und tiefen Untergrund in Deutschland (Phase 1 – Westliche Molasse und nördlich

    angrenzendes Süddeutsches Schichtstufenland). - Schlussbericht zum BMU-Projekt FKZ 0329985:

    220 p., Aachen (RWTH).

    Koch, A., Jorand, R., Arnold, J., Pechnig, R., Mottaghy, D., Vogt, C. & Clauser, C. (2009): Erstellung

    statistisch abgesicherter thermischer und hydraulischer Gesteinseigenschaften für den flachen und

    tiefen Untergrund in Deutschland (Phase 2 – Westliches Nordrhein-Westfalen und bayerisches

    Molassebecken). - Abschlussbericht zum BMU-Projekt FKZ 0329985: 174 p., Aachen (RWTH).

    Koch, R. (2011): Dolomit und Dolomit-Zerfall im Malm Süddeutschlands – Verbreitung,

    Bildungsmodelle, Dolomit-Karst. - Laichinger Höhlenfreund, 46: 75-92.

    Leinfelder, R. R., Krautter, M., Laternser, R., Nose, M., Schmid, D. U., Schweigert, G., Werner, W.,

    Keupp, H., Brugger, H., Hermmann, R., Rehfeld-Kiefer, U., Schroeder, J. H., Reinhold, C., Koch,

    R., Zeiss, A., Schweizer, V., Christmann, H., Menges, G., Luterbacher, H. (1994): The origin of

    Jurassic reefs: current research developments and results. - Facies 31, 1 –56.

    Leinfelder, R. R., Werner, W., Nose, M., Schmid, D. U., Krautter, M., Laternser, R., Takacs, M.,

    Hartmann, D. (1996): Paleoecology, growth parameters and dynamics of coral, sponge and

    microbolite reefs from the late Jurassic. - Gött. Arb. Geol. Paläontol., 2: 227– 248.

    Liedmann, W. (1992): Diageneseentwicklung süddeutscher Malmkarbonate. - Diss., 307 S.,

    Heidelberg (Uni. Heidelberg).

    Lüschen, E., Dussel, M., Thomas, R., Schulz, R. (2011): 3D-Seismik Unterhaching 2009 im Rahmen

    hydrothermaler Exploration und Modellierung. - Der Geothermiekongress 2011, Bochum (GtV

    Bv. e.V).

    Meyer, R. K. F. & Schmidt-Kaler, H. (1989): Paläogeographischer Atlas des süddeutschen Oberjura

    (Malm). - Geol. Jb., A 115: 3-77.

    Meyer, R. K. F., Schmidt-Kaler, H. (1990): Paläogeographie und Schwammriffentwicklung des

    süddeutschen Malm – ein Überblick. – Facies, 23 (1): 175-184.

    Nelson, R. A. (1985): Geological Analysis of Naturally Fractured Reservoirs. – Contr. In Petrol.

    Geology and Eng., 320 S., Housten, Texas (Gulf Publishing).

    Pawellek, T. (2001): Fazies-, Sequenz- und Gamma-Ray-Analyse im höheren Malm der Schwäbischen

    Alb (SW-Deutschland) mit Bemerkungen zur Rohstoffgeologie (Hochreine Kalke). - Tübinger

    geowissenschaftliche Arbeiten, Reihe A, Band 61, Tübingen (Universität Tübingen).

    Pawellek, T. & Aigner, T. (2003): Apparently homogenous „reef“-limestones built by high-frequency

    cycles Upper Jurassic, SW-Germany. - Sediment. Geol., 160: 259-284.

    Pawellek, T. & Aigner, T. (2004): Dynamic stratigraphy as a tool in economic mineral exploration:

    ultra-pure limestones (Upper Jurassic, SW Germany). - Marine and Petroleum Geology, 21: 499-

    516.

  • Philipp, S. L., Gudmundsson, A. & Oelrich A. R. I. (2007): How structural geology can contribute to

    make geothermal projects successful. - European Geothermal Congress 2007 Unterhaching, 10 p.

    Pieńkowski, G., Schudack, M. E., Bosák, P., Enay, R., Feldman-Olszewska, A., Golonka, J.,

    Gutowski, J., Herngreen, G. F. W., Jordan, P., Krobicki, M., Lathuilière, B., Leinfelder, R.R.,

    Michalík, J., Mönnig, E., Noe-Nygaard, N., Pálfy, J., Pint, A., Rasser, M.W., Reisdorf, A.G.,

    Schmid, D.U., Schweigert, G., Surlyk, F., Wetzel, A. & Wong, T.E. (2008): Jurassic. - In:

    McCann, T. (ed.): The Geology of Central Europe. - Volume 2: Mesozoic and Cenozoic: 823-922,

    London (The Geological Society).

    Popov, Y. A., Berezin, V.V., Semenov, V.G., Korostelev, V.M. (1985): Complex detailed

    investigations of the thermal properties of rocks on the basis of a moving point source. - Phys.

    Solid Earth, 21 (1), 64-70.

    Popov, Y. A., Tertychnyi, V., Romushkevich, R., Korobkov, D., Pohl, J. (2003): Interrelations

    between thermal conductivity and other physical properties of rocks: experimental data. - Pure app.

    geophys., 160, 1137-1161.

    Pribnow, D. & Sass, J. (1995): Determination of Thermal Conductivity from Deep Boreholes. - J.

    Geophys. Res., 100: 9981-9994.

    Pribnow, D. (1994): Ein Vergleich von Bestimmungsmethoden der Wärmeleitfähigkeit unter

    Berücksichtigung von Gesteinsgefügen und Anisotropie. - VDI Fortschrittsberichte, Reihe 19 (75):

    111 S.

    Reinecker, J., Tingay, M., Müller, B., Heidbach, O. (2010): Present-day stress orientation in the

    Molasse Basin. - Tectonophysics, 482 (1-4): 129-138.

    Reinhold, C. (1998): Multiple episodes of dolomitization and dolomite recrystallization during

    shallow burial in Upper Jurassic shelf carbonates: eastern Swabian Alb, southern Germany. -

    Sedimentary Geology, 121: 71-95.

    Sass, I. & Götz, A. E. (2012): Geothermal reservoir characterization: a thermofacies concept. - Terra

    Nova, 24: 142-147.

    Sass, J. H., Lachenbruch, A. H., Munroe, R. J., Greene, G. W., Moses, T. H., Jr. (1971): Heat Flow in

    the Western United States. - Journal of Geophysical Research, 76 (26): 6376-6413.

    Schauer, M. (1996): Untersuchungen an Dolomitsteinen und zuckerkörnigen Kalksteinen auf der

    Mittleren Schwäbischen Alb. - Laichinger Höhlenfreund, 31: 5-18.

    Schauer, M. (1998): Dynamische Stratigraphie, Diagenese und Rohstoffpotenzial des Oberjura

    (Kimmeridge 1-5 der mittleren Schwäbischen Alb.- Tübinger geowissenschaftliche Arbeiten, Reihe

    A, Band 36, Tübingen (Universität Tübingen).

    Schmoker, J. W. (1984): Empirical relation between carbonate porosity and thermal maturity: an

    approach to regional porosity prediction. - AAPG Bulletin, 68: 1697-1703.

  • Schoen, J. H. (1996): Physical Properties of Rocks: Fundamentals and Principles of Petrophysics. - In:

    Handbook of Geophysical Explorations. Section I, Seismic Exploration. – V.18, Trowbridge

    (Redwood Books).

    Schulz, R. (ed.), Thomas, R., Dussel, M., Lüschen, E., Wenderoth, F., Fritzer, T., Birner, J., Schneider,

    M., Wolfgramm, M., Bartels, J., Hiber, B., Megies, T., Wassermann, J. (2012): Geothermische

    Charakterisierung von karstig-klüftigen Aquiferen im Großraum München. - Final report,

    Förderkennzeichen 0325013A, Hannover (LIAG).

    Schulz, R., Agemar, T., Alten, A.-J., Kühne, K., Maul, A.-A., Pester, S. & Wirth, W. (2007): Aufbau

    eines geothermischen Informationssystems für Deutschland. - Erdöl Erdgas Kohle 123 (2): 76-81.

    Selg, M. & Wagenplast, P. (1990): Beckenarchitektur im süddeutschen Weißen Jura und die Bildung

    der Schwammriffe. - Jh. Geol. Landesamt Baden-Württemberg 32: 171-206.

    Somerton, W. H. (1992): Thermal properties and temperature-related behavior of rock-fluid systems. -

    Developments in Petroleum Science, 37 (8): 257 S.

    Tikhomirov, V. M. (1968): The thermal conductivity of rocks and its relationship to density, moisture

    content, and temperature (Russian text). - Neftianoe Khoziaistvo, 46: 36-40.

    Villinger, E. (1988): Hydrogeologische Ergebnisse. - In: Bertleff, B. et al.: Ergebnisse der

    Hydrogeothermiebohrungen in Baden-Württemberg. - Jh. d. Geol. Landesamtes Baden-

    Württemberg, 30: 27 – 116.

    Villinger, E. & Fleck, W. (1995): Symbolschlüssel Geologie (Teil 1) und Bodenkunde Baden-

    Württemberg. – Informationen Geologisches Landesamt Baden-Württemberg, 5: 1–69.

    Vosteen, H.-D. & Schellschmidt, R. (2003): Influence of temperature on thermal conductivity, thermal

    capacity and thermal diffusivity for different types of rock. - Physics and Chemistry of the Earth,

    28 (9-11): 499-509.

    Williams, C. F., Galanis, P. S., Grubb, A. D. & Moses, T. H. (1995): The Thermal Regime of Santa

    Maria Province (California). - In: Keller., M. A. (ed.): U.S. Geological Survey Bulletin - Evolution

    of Sedimentary Basins/onshore Oil and Gas Investigations, Santa Maria Province, USA.

    Wolfgramm, M., Dussel, M., Koch, R., Lüschen, E., Schulz, R., Thomas, R., (2011): Identifikation

    und Charakterisierung der Zuflusszonen im Malm des Molassebeckens nach petrographisch-

    faziellen und geophysikalischen Daten. - Der Geothermiekongress 2011, Bochum (GtV Bv. e.V.).

    Zoth, G. & Hänel, R. (1988): Thermal conductivity. - In: Hänel, R., Rybach, L., & Stegena, L. (eds)

    Handbook of Terrestrial Heat Flow Density Determination. - 449–453, (Kluwer Academic

    Publishers).

  • Figure captions

    Fig. 1: Study area with location of studied outcrops and drillings.

    Fig. 2: Cross section of the western Molasse Basin (modified, Clauser et al. 2002).

    Fig. 3: Palaeogeography of the Upper Jurassic in central Europe with study area Swabian and

    Franconian Alb (modified from Pawellek 2001, based on Meyer and Schmidt-Kaler, 1989, 1990).

    Fig. 4: Comparison of the different lithostratigraphic units in the Franconian and Swabian Alb based

    on reference sections after Geyer & Gwinner (1979) and Villinger & Fleck (1995).

    Fig. 5: A stratigraphic trend of increasing thermal conductivity is detected within the Malm from

    Malm α to Malm ζ.

    Fig. 6: Thermofacies classification of Upper Jurassic limestones of southern Germany based on the

    international Dunham classification and including the nomenclature introduced by Pawellek (2001);

    arrows indicate increasing values, arrow in both directions means approximately same range of values.

    Fig. 7: Schematic sketch of the depositional area of carbonates of the Purbeck and Malm and

    corresponding thermo- and petrophysical matrix properties in correlation with intensity of

    dolomitization; estimation of dolomitization intensity is based on Wolfgramm et al. (2011); values

    marked with * of the Purbeck are from Koch et al. (2007, 2009); TC: Thermal Conductivity, cp:

    specific heat capacity, n.d.: no data available.

    Fig. 8: Grain density for different lithotypes and dominant rock types (Dunham classification).

    Fig. 9: Porosities of Upper Jurassic stratigraphic units and hydraulic conductivity interpretation.

    Fig. 10: Porosity/Permeability relation of different lithotypes (only mean values are displayed).

    Fig. 11: Correlation of permeability and porosity. Overview of permeability and porosity relations of

    global carbonate reservoirs (including statistical trends: P90 (90% of reservoirs have porosity greater

    than this value), P50 (median), and P10) in comparison with matrix values of carbonates of the study

    area (modified, based on Ehrenberg und Nadeau, 2005); blue arrow indicates permeability shift of

    approximately two magnitudes considering fracture network, dolomitization and karstification.

  • Fig. 12: Temperature distribution at the top of the Malm formation in the south eastern part of the

    Molasse Basin (modified, based on Schulz et al. 2007 (www.geotis.de))

    Fig. 13: Near surface (outcrop) data in comparison with measurements from shallow (Solnhofen-

    Maxberg and Oberdolling) and deep core drillings (Moosburg SC4).

    Fig. 14: left: karstification along fault in thick-bedded and platy limestones (Solnhofen); right:

    karstification on larger scale (karst cavities) in reefal limestone (Hülen).

    Fig. 15: Present-day orientation of maximum horizontal stress (SH) determined from borehole

    breakouts and drilling-induced fractures. Symbol size is proportional to quality of the data. A (good) –

    D (poor) quality stress orientations (long axes of symbols) in 96 wells across the German Molasse

    Basin demonstrate that the present-day SH in the German Molasse Basin is predominantly oriented N–

    S (modified from Reinecker et al. 2010).

    Tab. 1: Number of measurements and mean values for different lithofacies types.

    Tab. 2: Facies types and associated thermo-physical properties at laboratory (measured values) and

    reservoir temperatures (calculated values) according to different transfer models; matrix dominated:

    Mud-/wackestone, grain-/component dominated: Rud-/floatstone, grain- and dolostone.

  • Figure 1

    Figure 2

  • Figure 3

  • Figure 4

  • Figure 5

  • Figure 6

  • Figure 7

    Figure 8

  • Figure 9

    Figure 10

  • Figure 11

    Figure 12

  • Figure 13

    Figure 14

  • Figure 15

  • Table 1

    Parameter Total number of

    measurements

    Number of measurments per facies- (lithotype)

    [mean value]

    Basin facies

    (mud-/wackestone)

    Massive reefal facies

    (rud-/float-/grainstone)

    Thermal

    conductivity

    862 426 [1,9 W/(m∙K)] 436 [2,6 W/(m∙K)]

    Thermal

    diffusivity

    801 407 [1,05 ∙ 10-6 mm/s] 394 [1,31 ∙ 10-6 mm/s]

    Specific heat

    capacity

    796 (calculated) 410 [775 J/(kg∙K)] 386 [733 J/(kg∙K)]

    Density 236 138 [2,66 g/cm³] 98 [2,73 g/cm³]

    Porosity 1126 552 [3.5%] 574 [6%]

    Permeability 1014 569 [5 ∙ 10-17 m²] 445 [1 ∙ 10-16 m²]

    Table 2

    Parameter Facies type Value range at

    20°C, dry

    Value range for

    saturated

    conditions at

    (20°C)

    Value range at

    (150°C),

    saturated

    Transfer model

    Specific heat

    capacity

    [J/kgK]

    Matrix dominated

    613 - 1045 645 - 1081 806 - 1227 Vosteen &

    Schellschmidt

    (2003) Grain/component

    dominated

    547 - 1167 565 - 1184 711 - 1330

    Thermal

    conductivity

    [W/(mK)]

    Matrix dominated 1.35 - 2.62 1.60 - 2.79 1.80 - 2.46 Zoth & Hänel

    (1988) Grain/component

    dominated

    1.72 - 4.87 1.85 - 4.94 1.92 - 3.53

    Matrix dominated 1.35 - 2.62 1.60 - 2.79 1.83 - 2.51 Sass et al.

    (1971)

    Grain/component

    dominated

    1.72 - 4.87 1.85 - 4.94 1.92 - 3.92

    Matrix dominated 1.35 - 2.62 1.60 - 2.79 1.41 - 2.25 Vosteen &

    Schellschmidt

    (2003), Clauser

    et al. (2002)

    Grain/component

    dominated

    1.72 - 4.87 1.85 - 4.94 1.56 - 3.71