Skip to main content

Science – Society – Technology

Inferring the in situ stress regime in deep sediments: an example from the Bruchsal geothermal site

Abstract

Background

Knowledge of the ambient state of stress is of crucial importance for understanding tectonic processes and an important parameter in reservoir engineering. In the framework of the 2,500-m deep geothermal project of Bruchsal in the central part of the Upper Rhine Graben, new evidence is presented for the stress field in deep-seated sedimentary rocks.

Methods

With a sophisticated data analysis based on the concept of critical stress ratios, we evaluate the quality and uncertainty range of earlier stress field models in the Bruchsal area. New data from borehole logging and leak- off tests in deep sediments are used to propose an alternative stress profile for this part of the Upper Rhine Graben.

Results

The revised stress field model for the Bruchsal area predicts a normal with transition to strike-slip faulting regime. Stress field perturbations and potential decoupling process within specific clay-, salt-, and anhydrite-bearing units of the Keuper can be observed.

Conclusion

By comparison with other models, we can show a regional consistency of our stress field model that is reliable throughout the central Upper Rhine Graben extending from Bruchsal in the East to the Soultz-sous-Forêts EGS site in the West.

Background

In a regional context, the stress field is typically used for investigation of neotectonic and recent geodynamic processes. The world stress map (Heidbach et al. [2008]) provides a sound database with respect to determination of fault reactivation, tectonic deformation, and related earthquake hazard (e.g., Hergert and Heidbach [2011]). Moreover, stress is a key parameter in unconventional reservoir engineering. Faults and fractures that are favorably oriented and critically stressed for frictional failure often dominate fluid flow (Barton et al. [1995]; Townend and Zoback [2000]). In this respect, a higher resolution of the stress field is required and linear stress-depth profiles should be used with caution, as principal stress magnitudes can vary locally by topography, geological unconformities, stratifications, lithology, or geological structures like faults or fractures (Heidbach et al. [2010]; Zang and Stephansson [2010]). In sedimentary rocks, stress field orientation and principal stress magnitudes show significant variations depending on their rheological characteristics (Anderson et al. [1973]; Cornet and Burlet [1992]). Interstratification of stiff clastic sediments and clay-, salt-, and anhydrite-bearing formations causes significant deviations from linear stress-depth profiles in deep sedimentary basins such as the North German basin, the Paris Basin, and in continental rift systems such as the Upper Rhine Graben (URG) (Cornet and Röckel [2012]; Wileveau et al. [2007]). Thus, stress measurements in sediments (orientation and magnitude) need to be evaluated with respect to the lithological characteristics of the corresponding formation, and extrapolation of measured stress values to depth should be conducted with care, especially when only few measurements are available. With this in mind and although the world stress map provides a large amount of data, determination of local stress appears often to be insufficient.

In this study, we present a methodology for stress field estimation in areas where a detailed knowledge of the local stress conditions is unavailable. The approach is applied on the example of the Bruchsal geothermal site, where a number of earlier studies have been carried out with a similar aim. We will present a comparative review of the existing data and add new unpublished data from two leak-off tests (LOT) and our resulting approach applied to the Bruchsal area.

Geological setting

Bruchsal is located in the central segment of the URG close to its Eastern boundary fault (Figure 1). The Bruchsal geothermal doublet system operates through a 1,932-m-deep injection (GB1) and a 2,542-m-deep production well (GB2). The highly fractured geothermal reservoir, located at a depth ranging between 1.8 and 2.5 km, mainly consists of fine- to coarse-grained sandstones of the Lower Triassic (Buntsandstein) and gravelly sandstones and breccia conglomerates of the Upper Permian (Rotliegend and Zechstein). The overlying Triassic units are characterized by clay-rich formations with carbonate and dolomitic layers (Muschelkalk) and gypsum- and anhydrite-bearing layers of the Keuper.

Figure 1
figure 1

Compilation of data relevant to the stress field in the URG. Shown are stress field indicators derived from seismological and well test data compiled in the world stress map (Heidbach et al. [2008]). The underlying map shows a digital elevation model based on SRTM data and the major tectonic fault systems (gray lines; Illies and Greiner [1979]). Position of the boundary faults and major shear zones are displayed as red lines. WMBF, western main boundary fault; EMBF, eastern main boundary fault; LL/BB-SZ, Lalaye-Lubine/Baden-Baden shear zone; HTSBF, Hunsrück-Taunus southern boundary fault. For further information on the data, see text.

The 300-km-long URG represents the central part of the European Cenozoic rift system (Schumacher [2002]), extending over a distance of more than 1,000 km across central Europe. It is subdivided into a NNE-striking southern, a NE-striking central, and a NNW to N-striking northern segment (Figure 1). The evolution of the Cenozoic URG was controlled by polyphase reactivations of a complex set of crustal discontinuities of Late Paleozoic structures (Ziegler [1990]). The main extensional rifting and crustal NW-SE extension started in Late Eocene (Sissingh [1998]) during which Late Variscan and Permo-Carboniferous crustal discontinuities were transtensionally reactivated (Schumacher [2002]). The opening of the graben was controlled by a paleostress field with a SHmax orientation of NNE-SSW (Ahorner [1975]; Illies [1975]) resulting in the development of a NE-SW to NNE-SSW striking graben. A major reorientation of the regional stress field during early Miocene times established a NE-SW-extensional to transtensional stress field with reactivated fault segments showing sinistral and dextral oblique displacements but also local inversion and contraction (e.g., Illies and Greiner [1979]). The change of the regional stress field initiated a new tectonosedimentary regime. The synrift deposits and older strata in the central and southern segments were uplifted and partly eroded due to transpressional reactivation of these graben segments (e.g., Rotstein et al. [2005]; Rotstein and Schaming [2011]). A number of thermal anomalies in the western part of the central and northern part of the URG are linked to different structural features such as zones of uplift (Baillieux et al. [2013]). Subsidence and sedimentation were restricted to the northern graben segment with a maximum Cenozoic graben fill of up to 3.0 km (Bartz [1974]; Pflug [1982]).

Numerous local studies have been carried out to determine the tectonic stress field in the URG. Most of them are based on the analysis and interpretation of earthquakes, tectonic studies, overcoring data, hydraulic fracturing data, and borehole breakouts (Greiner [1975]; Baumann [1981]; Larroque and Laurent [1988]; Plenefisch and Bonjer [1997]; Valley and Evans [2007]). Figure 1 presents an extensive data compilation for the present-day stress field in the URG and shows the abovementioned major structural units of the rift system. The seismological data are derived from fault plane solutions of earthquakes from 1971 to 1980 and incorporate 33 fault plane solutions selected by Larroque et al. ([1987]), based mostly on data from Bonjer et al. ([1984]). The fault traces from Illies and Greiner ([1979]) are based on the interpretation of 2D seismic sections. The information on the stress field orientation and the faulting regime are taken from the world stress map (Heidbach et al. [2008]).

The compilation of stress field indicators in Figure 1 highlights the generally uniform NW-SE orientation of SHmax demonstrated earlier by Müller et al. ([1992]). This general trend is confirmed by stress inversion of earthquake fault plane solutions (Delouis et al. [1993]; Plenefisch and Bonjer [1997]). The general trend of the stress field shows a local variation with SHmax orientation in the northern URG ranging N130°E to N135°E and in the southern URG/northern Switzerland ranging N145°E to N160°E. Interpretations of fault plane solutions also reveal a change in faulting regime in the URG (e.g., Plenefisch and Bonjer [1997]). The northern part of the URG is characterized by an extensional stress state and active normal faulting (σ1 = Sv, σ2 = SHmax). In the seismically more active southern part, strike-slip faulting (σ1 = SHmax, σ2 = Sv) with secondary normal faulting is the predominant mechanism. The transition of the stress orientation and the change of the faulting regime by permutation of σ1 and σ2 (Larroque et al. [1987]) occurs in the central segment of the URG, in the area of the site of investigation, probably causing a transitional stress state between normal and strike-slip faulting.

At the western central margin of the URG, an extensive set of in situ stress data is available for the geothermal site of Soultz-sous-Forêts (France). Measured and derived orientations of SHmax determined down to 5 km varies between N125°E and N185°E with a mean value of N175°E ± 10° (Cornet et al. [2007]) and indicate a transitional stress state down to 5 km with a change from normal to strike-slip faulting at depths below 3 km (Cuenot et al. [2006]).

Methods

SHmaxorientation

For the Bruchsal geothermal wells, Eisbacher et al. ([1989]) have derived the SHmax orientation from borehole breakouts in GB1 and GB2 using oriented caliper logging. The logs were acquired in the depth range of 1,632 to 1,900 m (GB1) and 2,023 to 2,525 m (GB2) in the Keuper, Muschelkalk, and Buntsandstein formations and have been azimuth-corrected for the deviated wells. The values were subdivided into zones of fairly homogeneous orientation (Table 1) ranging between 50 and 100 m in depth. SHmax orientation was determined by stacking caliper data in each zone, with uncertainties of up to 20°. In addition to the stress-relevant data, the classification of reservoir rocks is indicated in Table 1. The italicized table entries indicate SHmax orientations determined in the clay-, gypsum-, and anhydrite-bearing formations of the Muschelkalk and Keuper. These low-permeable units seal the reservoir which mainly consists of sandstones and conglomerates.

Table 1 Depth intervals of the analyzed borehole breakouts and determined S Hmax orientations in wells GB1 and GB2

Svand PPcalculation

The magnitude of the vertical stress, Sv, is generally equal to the weight of the overburden and can be calculated by integration of the rock densities from the surface to the depth of interest. Consequently, the stratigraphic units of the overburden were subdivided in two major groups. The first group includes the quaternary and tertiary formations of the graben fill, while the second one includes the occurring Mesozoic formations. We assume an average density of 2,400 kg m−3 for the Cenozoic sedimentary succession based on Rotstein et al. ([2006]). For the Triassic Muschelkalk and Buntsandstein, densities were determined from a litho-density log (LDL) acquired in GB1 between 1,650 and 1,900 m (Figure 2). A weighted mean rock density of about 2,500 kg m−3 is indicated in the reservoir formations. This value is close to the literature data (Mueller [1988]; Plaumann [1967]). With an average thickness of the Tertiary graben sediments in Bruchsal of about 1,350 m, a mean density of 2,430 kg m−3 for the overburden is calculated for a reservoir depth of 2,000 m. This leads to a vertical stress magnitude of Sv = 47.7 MPa and a gradient of 23.8 MPa km−1.

Figure 2
figure 2

Measured rock densities obtained from a litho-density log in GB1 between 1,650 and 1,900 m. Red lines show mean density values for the drilled stratigraphic units. Zones of very small rock densities (<2,000 kg m−3) and increased porosities in the Buntsandstein are associated with major fracture and fault zones in the geothermal reservoir.

The pore pressure, PP, was calculated similarly, assuming that it is close to hydrostatic. With an average depth of the free water table 60 m below ground level, the reservoir reveals tendency to slight under pressure condition. The fluid density of the geothermal brine is 1,070 kg m−3 (T. Kölbel 2013, pers. comm.). At mean reservoir depth of 2,000 m, a pore pressure of PP = 20.4 MPa and a gradient of 10.2 MPa km−1 was calculated. This leads to a ratio of pore pressure to vertical stress magnitude of PP = 0.43SV.

Stress field profiles for Bruchsal and adjacent areas

Eisbacher et al. ([1989]) prepared two different stress profiles including the minimum (Shmin) and maximum (SHmax) horizontal stress components. The first stress field profile is based on the linear extrapolation of overcoring data from outcrops and shallow wells measured by Greiner ([1978]) hereafter referred to as the Greiner profile. In detail, the model is based on SHmax and Shmin magnitudes of 4.9 and 3.7 MPa, respectively, from the 140-m-deep Auerbach well, about 60 km north of Bruchsal (Figure 1). These data were interpolated with measurements from the Wössingen outcrop, 10 km SE of Bruchsal (Figure 1), with magnitudes of SHmax = 2.2 MPa and Shmin = 1.0 MPa. The obtained stress profile results in a normal faulting regime for the Bruchsal area of:

S Hmax = 2.2 + 0.019 · z MPa
S hmin = 2.2 + 0.019 · z MPa

The second profile is based on the stress field compilation of Rummel and Baumgärtner (1982, unpublished) for central Europe with data originating from 120 hydraulic fracturing measurements. In SW Germany, this compilation is based on wells at depths down to 500 m. The linear extrapolation (hereafter referred to as the Rummel and Baumgärtner profile) predicts a strike-slip regime with:

S Hmax = 0.8 + 0.034 · z MPa
S hmin = 0.9 + 0.021 · z MPa

At a reservoir depth of 2,000 m, the Rummel and Baumgärtner and Greiner profiles differ in Shmin by 3.9 MPa only, but in SHmax by 28.6 MPa. In the following, we compare these profiles to the well-defined stress field models prepared for the Soultz geothermal site. In Soultz, two profiles have been prepared recently: Cornet et al. ([2007]) show that the SHmax magnitude is close to SV (mean SV gradient of 24.5 MPa km−1) resulting in a range of 0.95SV < SHmax < 1.1SV between 2,800 and 3,600 m (Figure 3A) and Valley ([2007]) characterized the stress at Soultz between 1,500 and 5,000 m to be in the range of 0.90SV < SHmax < 1.05SV (with the same SV gradient, see Figure 3A). Both describe the transitional stress field between a normal faulting (NF) and strike-slip (SS) regime.

Figure 3
figure 3

Comparison of existing stress-depth relationships and stress ratios for the geothermal sites of Bruchsal and Soultz. Stress magnitudes of SHmax(A) and Shmin(B) are normalized by Sv. The stress profiles of Greiner and Rummel and Baumgärtner are based on the combination and linear extrapolation of stress magnitudes obtained from near surface hydro frac and overcoring measurements. The Soultz stress profiles from Cornet et al. ([2007]) and Valley and Evans ([2007]) are based on evaluations of large-scale injection experiments and interpretations of borehole breakouts in the deep crystalline reservoir of Soultz.

With exception of SHmax of the Greiner profile, the two Bruchsal stress profiles differ significantly from the Soultz profiles (Figure 3). With a rather similar Sv gradient of 23.8 MPa km-1, the Rummel and Baumgärtner and the Greiner stress profiles would yield SHmax/SV ratios at depths over 500 m of 1.4 to 1.5 and 0.8 to 1.0, respectively (Figure 3A). Hence, the Rummel and Baumgärtner profile predicts a strike-slip regime, while the Greiner profile indicates normal faulting at depths below 500 m. In contrast, the ratios of Shmin/SV in both Bruchsal profiles are similar (0.8 to 0.9, see Figure 3B) but differ from the Soultz profiles with ratios of approximately 0.5. Finally, it must be stated that the two Bruchsal profiles are debatable since they are derived from measurements at shallow depth only.

Critical stress concept

In situ measurements and detailed analyses to determine the magnitude of SHmax are rare in the URG. In such conditions, a critical stress concept may constrain the SHmax order of magnitude (Zoback et al. [2003]). According to frictional equilibrium, one can assume that the ratio between maximum and minimum effective stresses cannot exceed the one required to cause slip on pre-existing faults that are optimally oriented to the principal stress field (Jaeger et al. [2007]; Moos and Zoback [1990]). Hence, effective differential stresses are bounded by a critical ratio. Assuming that one of the principal stresses is vertical and that there is no cohesion, the critical stress ratio can be written after Jaeger et al. ([2007]) as:

σ 1 / σ 3 = S 1 P p / S 3 P p μ 2 + 1 + μ 2
(1)

where PP is the in situ pore pressure and μ the coefficient of frictional sliding. To predict limiting stress differences at depth, Anderson's faulting theory has to be applied to determine which of the principal stresses SHmax, Shmin, and Sv correspond to S1, S2, and S3. This will depend on whether it is a normal, strike-slip, or reverse faulting regime:

Normal faulting σ 1 / σ 3 = S v P p / S hmin P p μ 2 + 1 + μ 2
(2)
Strike slip faulting σ 1 / σ 3 = S Hmax P p / S hmin P p μ 2 + 1 + μ 2
(3)
Reverse faulting σ 1 / σ 3 = S Hmax P p / S v P p μ 2 + 1 + μ 2
(4)

Based on these equations, the critical ratio of the principal stress magnitudes depends on depth, pore pressure, and the coefficient of frictional sliding. A coefficient of friction of 0.85 has been shown to be applicable under shallow crustal conditions for normal stresses up to 200 MPa, i.e., depths of approximately 6 km (Byerlee [1978]). However, it is generally accepted that hydrothermal alteration and clay content in shear planes contribute to a significant reduction of the friction angle during fault reactivation (Krantz [1991]). But, since no data on the mineralogical composition of the faults and fractures in Bruchsal were acquired, we assume μ = 0.85, which results in an effective stress ratio of σ1/σ3 < 4.68. The possible mechanically stable stress conditions in different stress regimes have been comprised by Zoback et al. ([2003]) to the so-called stress polygon. For a given friction coefficient, the area inside the polygon represents the allowed stress, whereas values lying outside the borders are instable and yield failure.

Results

In the following, we will present the results of the determination of Shmin by two leak-off tests from the well GB2 as well as the application of the critical stress concept to further constrain Shmin and SHmax.

Leak-off tests and Shminmagnitude

For the determination of the least principal stress magnitude, Shmin, we interpreted new available data sets of two LOT that were carried out in 1984 (Table 2).

Table 2 Key parameters of the two leak-off tests conducted in the geothermal well Bruchsal GB2 in 1984

The leak-off pressure (LOP) at the depth of the test is the sum of the well head pressure and the pressure in the wellbore due to the wellbore fluid. According to Zoback et al. ([2003]), we consider the clear LOP of each test (corresponding to a distinct break-in-slope of the linear pressure build-up) to be approximately equal to the Shmin magnitude. While LOT-2023 is conducted in the weak Keuper, LOT-2245 is located in stiff units of the Middle Buntsandstein. We would like to emphasize that LOT-2023 results in a higher LOP of 31.0 MPa compared to LOT-2245 with a LOP of 27.6 MPa. This finding is consistent with Shmin magnitudes obtained in comparable environments (Cornet and Röckel [2012]). The determined Shmin/Sv ratios show contradictory results to those from Soultz.

Critical stresses of Shmin

Based on the statements in Section 2, a normal faulting regime (NF) close to strike-slip transition is assumed for the Bruchsal area. In a first step, the SV gradient of 23.8 MPa km−1 is calculated for a mean rock density of 2,430 kg m−3. The application of the critical stress concept is illustrated in Figure 4A. Applying Equation 2, the Shmin magnitude of LOT-2023 (15.3 MPa km−1) results in a σ1/σ3 ratio that is smaller than the critical value of σ1/σ3 < 4.68, whereas the σ1/σ3 ratio of LOT-2245 (12.3 MPa km−1) is larger in depths over 500 m. The latter ratio violates the critical stress concept. This may be due to errors in the determination of the Sv (in the following called upper bound) or Shmin (in the following called lower bound) gradient. It should be mentioned here that a higher magnitude of μ also reveals higher critical stress ratios. The interpretation of LOT-2245 (Shmin/Sv = 0.52) leads to stress ratios between 6 and 7 at 3,000 m depth. In order to be consistent with the critical stress concept, a μ of 1.1 is needed. We consider this rather unlikely and discuss potential uncertainties in Sv and Shmin determination only.

Figure 4
figure 4

Application of the critical stress concept and uncertainty analysis of the determined S v gradient. (A) Comparison of the effective stress ratios σ1/σ3 for Bruchsal calculated for the Shmin magnitudes of the leak-off tests LOT-2023 (green line) and LOT-2245 (red line) in GB2. The vertical dashed line represents the critical σ1/σ3 ratio of 4.68, i.e., the maximum stress ratio for which the critical stress concept is valid. The blue line shows corrected stress ratios for LOT-2245 to be consistent with the critical stress concept (see text). (B) Overview of maximum possible rock density for derived Shmin magnitudes with respect to fulfilling the criteria for the critical stress concept.

Therefore, a sensitivity study on the rock density (or vertical stress gradient) was conducted to fit the critical stress concept. In this case, the lower bound of the stress ratio is modified. For the Shmin of LOT-2245, the assumed density of the overburden has to be decreased at least from 2,430 to 2,000 kg m−3 (Figure 4B), yielding a decrease of the SV gradient by 18%, from 23.8 to 19.6 MPa km−1. After comparing this value with the LDL measurements in GB2 (Subsection 3.2), we can conclude that it is unrealistically low. On the other hand, the lower bound of the ratio can be investigated, keeping the density distribution fixed. To comply with the critical stress concept, Shmin of LOT-2245 needs to be increased by 7%, from 12.3 to 13.2 MPa km−1. Figure 5 shows the corresponding stress profiles including the measured LOT data.

Figure 5
figure 5

Uncertainty analysis of the determined S hmin gradients derived from LOT-2023 and LOT-2245. According to the critical stress concept, the effective stress ratio σ1/σ3 should not exceed a factor of 4.68 (gray shaded area). Assuming a Sv gradient of 23.8 MPa km−1 and a NF regime, the σ1/σ3 ratio for LOT-2023 is within this range whereas LOT-2245 shows inconsistent ratios between 6 and 7. The slight increase of Shmin from 12.3 to 13.2 MPa km−1 (blue shaded area) indicates the minimum necessary variation to keep the σ1/σ3 ratio of LOT-2245 within the predicted threshold of σ1/σ3 ≤ 4.68. Uncertainty range of Shmin magnitudes is estimated following observations of Valley ([2007]) for deep well injection tests in Soultz.

There are no error estimates given on the Bruchsal LOT measurements. Based on the example from Soultz, we can assume a reasonable uncertainty range for Shmin magnitudes of ± 0.45∙z [km] + 1.82, as derived from several large and small volume injection tests (Valley [2007]). If we assume similar error margins for Bruchsal, a bandwidth of the Shmin profile can be determined. An increase of the Shmin/Sv ratio from 0.52 to 0.55 is consistent with the critical stress concept and remains consistent with the error margins determined for the deep geothermal wells in Soultz. This represents a minimum Shmin gradient of LOT-2245, referred to as ‘LOT-2245 (SL)’ in the following (Figure 4, blue lines).

The hypothesis on the NF tectonic regime also influences the consistency with the critical stress concept. Following the same procedure, it is obvious that a strike-slip regime with S1 = SHmax > SV would result in an effective stress ratio of σ1/σ3 > 4.68. This would rather lead to a larger increase of the Shmin gradient for LOT-2245 to comply with the critical stress concept. Accordingly, this option is not further considered in this study.

Critical stresses of SHmax

As described by Zoback et al. ([2003]), the applied critical stress concept is illustrated in Figure 6 by the stress polygon. We assume a pore pressure of PP = 0.43Sv and a frictional coefficient of μ = 0.85. Points at the periphery of the polygon correspond to a state of stress at the frictional limit (I, radial extension; II, normal faulting; III, transitional normal/strike-slip faulting; IV, strike-slip faulting; V, transitional strike-slip/reverse faulting; VI, reverse faulting; VII, radial compression). According to Equations 2 and 4, the lower bound on SHmax is given by SHmax = 0.55Sv (see previous discussion, point I in Figure 6A) and the upper bound by SHmax = 3.10Sv (line between points V and VII in Figure 6A).

Figure 6
figure 6

Stress polygon concept in Bruchsal to constrain allowable stress ratios and related faulting regimes. (A) Possible range for horizontal stress components SHmax and Shmin (each normalized by Sv). (B) SHmax magnitudes accounting for Shmin from the leak-off tests LOT-2023 and LOT-2245 (SL). Gray shaded area indicates the range of possible stress states between a strike-slip and a radial extensional faulting regime.

The range of possible values from Shmin (i.e., measured LOT-2023 and corrected LOT-2245 (SL)) is constrained in Figure 6B. The lower bound of SHmax as defined from LOT-2245 (SL) would correspond to a radial extensional faulting regime. According to our assumptions on a NF regime, SHmax may take any value between Shmin and Sv (segments I to III in Figure 6B) and reaches the limit to a strike-slip regime. For LOT-2023 a stress ratio of SHmax = Shmin = 0.64Sv marks the lower bound for SHmax. With a relation of SHmax = 1.42Sv, a strike-slip regime seems possible, too.

Discussion

The following discussion includes an assessment of the obtained stress profile in a regional context and with respect to the well-defined Soultz stress field. It should be mentioned that the sedimentary cover at Soultz reaches a maximum depth of about 1,400 m. Consequently, at the depth of the Bruchsal reservoir at Soultz, we are already in the upper granitic reservoir (Genter et al. [2010]).

Stress field variations

The orientation of SHmax in Bruchsal between 125° and 145° agrees with the general NW-SE trend of the regional stress field in SW Germany (Heidbach et al. [2008]; Müller et al. [1992]) and thus differs from the values in Soultz by about 30° to 40°. This variation may be attributed to decoupling at the interface between the crystalline basement and the overlying sedimentary rock. Such a behavior is also observed in the Otterbach-2 well in Basel. Breakouts and drilling-induced fractures identified above and below the sediment-basement interface suggest a clockwise 30° rotation of SHmax (Valley and Evans [2009]). In Bruchsal, internal variation of orientation seems to be linked to lithological changes. A comparison of the SHmax orientations for different depth sections including their variation with the litho-log of the two wells GB1 and GB2 is presented in Figure 7.

Figure 7
figure 7

Comparison of the stratigraphic and lithological well profiles of the Bruchsal geothermal wells. Stress field orientations based on breakout analysis of oriented caliper logging data (Eisbacher et al. [1989]). Mean SHmax azimuths were determined by stacking breakout data in zones of fairly homogeneous orientations to take lithological characteristics into account.

Generally, SHmax orientation exhibits a high variability in the upper part of the investigated section of the Mesozoic sediments. Orientations stabilize at about 140° below the Muschelkalk in GB1. With the exception of the Middle Buntsandstein to Upper Permian in GB2, the orientations stabilize below the Keuper at about 130°. Thus, highest variation is observed for the sections with a high clay, gypsum, or anhydrite content. LOT-2023 is located in such lithology. Measured SHmax orientation is by about 30° to 40° off the general GB2 trend in this depth range. The determined Shmin magnitude of LOT-2023 is 3.4 MPa higher than that of LOT-2245. This observation is consistent with results summarized by Cornet and Röckel ([2012]) and the increase in Shmin by about 4 MPa in the Callovo-Oxfordian argillite of the Paris basin (Wileveau et al. [2007]). It suggests that the leak-off tests in rocks with a low friction coefficient behave generally in this manner, which is also indicated by laboratory experiments (e.g., Krantz [1991]). The stronger rocks of the Middle Buntsandstein can support higher differential stresses without failing than the weaker rocks of the Middle Keuper. It is rather noteworthy that for the same stratigraphic unit, internal deformation processes have also been described (Wirth [1962]). Litho-log interpretations of numerous oil wells of the former oil field Forst, 3 km North of Bruchsal, suggest massive internal shearing and sub-horizontal deformation within the Keuper and especially in the evaporitic layers of the Lower Keuper. With regard to our results, both orientation and magnitude of LOT-2023 show similar deviations and, thus, are not considered to be representative for the reservoir zone of Bruchsal. In contrast, LOT-2245 is located within the section of stable SHmax orientation.

Stress profile

The discussion on the suitability of earlier stress field concepts as well as the new leak-off test data lead us to propose an alternative stress profile for the Bruchsal area that is presented in Figure 8. Our resulting stress field model is based on the LOT-2245 (SL) applying the critical stress concept. It should be mentioned that the critical stress concept restricts the ratio σ1/σ3 depending on the coefficient of frictional sliding. An increase of this ratio leads to shear failure on the most favorable oriented fault or fracture plane. We are aware that, if this fracture trend does not exist in the subsurface, larger stress ratios may be applicable. However, with no detailed information on the orientation of the fracture network, such as in Bruchsal, the application of the critical stress concept is a rather conservative approach.

Figure 8
figure 8

Comparison of the estimated stress-depth relations and stress ratios for the Bruchsal geothermal site. (A) Predicted magnitudes for SHmax based on the stated Greiner and Rummel and Baumgärtner profiles and the newly derived Bruchsal profile, respectively. The range between the assumed upper and lower bound of SHmax covers a potential normal (NF), a radial extensional (RE), and a transitional (NF/SS) faulting regime for the Bruchsal model (see Figure 4). (B) Overview of the minimum horizontal stress component Shmin. (C, D) Comparison of the potential stress magnitudes of SHmax and Shmin normalized by Sv.

Theoretically, our stress model has a SHmax-depth relation constrained by LOT-2245 (SL) with radial extension as a lower bound faulting regime and NF/SS transition as an upper bound (Figure 8A). Geologically, the validity of the radial extension under the tectonic settings of Bruchsal is, however, questionable. As discussed in Section 2, a transitional to normal faulting stress regime in the central part of the URG in general and in Bruchsal in particular is rather likely. This is consistent with the Soultz stress profiles or determined earthquake fault plane solutions (e.g., Ritter et al. [2009]). Thus, we suggest our model to be valid for NF and NF/SS faulting regimes. It should be mentioned that the Greiner profile is within our range of SHmax for depths larger than 500 m.

The proposed Shmin-depth relation (Figure 8B) presents the steepest gradient (except the results for the uncorrected LOT-2245) and clearly deviates from the earlier Greiner and Rummel and Baumgärtner profiles.

Although located on the opposite side of the URG at the Eastern boundary fault and despite the different lithologies, the determined stress ratios SHmax/Sv of 1.0 and Shmin/Sv of 0.55 for Bruchsal are close to those obtained at the same depth at Soultz (Figure 8C,D).

According to these considerations, a stress range can at best be estimated. To better constrain this, we therefore strongly recommend detailed stress measurements, such as those proposed by Cornet and Burlet ([1992]) in geothermal wells, in particular with respect to the option of hydraulic stimulation.

Conclusions

World-wide, geothermal projects are often located in active tectonic terrains with highly fractured subsurface reservoir rocks. The evaluation of the stress field in these projects is of main importance. It improves the understanding of the structural setting of the reservoir and supports the definition of an optimum exploration strategy by targeting zones of high transmissivity. The geothermal sites in the URG in central Europe are typically characterized in this context. The Rhine Graben tectonics yields large differential stress levels and favors therefore the existence of large fault and fracture systems. Due to a rather complex history, local evidence of the stress field needs to be evaluated. However, this importance of the stress field has only been pointed out gradually, and many earlier projects provide only little relevant data. In this perspective, new concepts and methodologies need to be considered. Our application of the critical stress concept to measured data may lead to better constrained results and this within their uncertainty of the measurements. The critical stress concept provides a number of solutions for the stress field model depending on possible stress regimes. Using indications such as the local geology and structural setting, the number of solutions has been reduced to normal faulting and transition of normal faulting to strike-slip regimes. Finally, based on new leak-off test data, we have constrained a new stress field model for the Bruchsal area.

Our results show that in accordance with earlier studies, sedimentary layers may have a large influence on both the orientation of the mean principal stress direction and the magnitudes. This applies in particular to clay- and anhydrite-rich layers that occur in this case in the Keuper formations and partly in the Muschelkalk. Stress decoupling and aseismic deformation within this formation cannot be excluded as also indicated by the results from the nearby Forst oil reservoir. We could show that stress magnitudes obtained in sedimentary formations vary significantly due to different rheological characteristics, and a linear stress-depth relation has to be treated only as a first approach. The result from a leak-off test in the Middle Buntsandstein in Bruchsal is, however, consistent with the Shmin/Sv ratio of Soultz. We may attribute this consistency across the entire E-W extension of the URG and different lithologies to small differences in the mechanical properties of the Soultz granitic basement and the sedimentary rocks of the Middle Buntsandstein in Bruchsal.

References

  • Ahorner L: Present-day stress field and seismotectonic block movements along major fault zones in Central Europe. Tectonophysics 1975, 29(1–4):233–249. 10.1016/0040-1951(75)90148-1

    Article  Google Scholar 

  • Anderson R, Ingram D, Zanier A: Determining fracture pressure-gradients from well logs. Journal of Petroleum Technology 1973, 25(NOV):1259–1268. 10.2118/4135-PA

    Article  Google Scholar 

  • Baillieux P, Schill E, Edel J-B, Mauri G: Localization of temperature anomalies in the Upper Rhine Graben: insights from geophysics and neotectonic activity. International Geology Review 2013, ᅟ: 1–19.

    Google Scholar 

  • Barton C, Zoback M, Moos D: Fluid flow along potentially active faults in crystalline rock. Geology 1995, 23(8):683–686. 10.1130/0091-7613(1995)023<0683:FFAPAF>2.3.CO;2

    Article  Google Scholar 

  • Bartz J: Die Mächtigkeit des Quartärs im Oberrheingraben. Approaches to Taphrogenesis. Inter-Union Commission on Geodynamics, Scientific report No. 8., Schweizerbart, Stuttgart, Germany. 1974.

    Google Scholar 

  • Baumann H: Regional stress field and rifting in Western-Europe. Tectonophysics 1981, 73(1–3):105–111. 10.1016/0040-1951(81)90177-3

    Article  Google Scholar 

  • Bonjer K, Gelbke C, Gilg B, Rouland D, Mayer-Rosa D, Massinon B: Seismicity and dynamics of the Upper Rhinegraben. J Geophys-Z Geophys 1984, 55(1):1–12.

    Google Scholar 

  • Byerlee J: Friction of rocks. Pure and Applied Geophysics 1978, 116(4):615–626. 10.1007/BF00876528

    Article  Google Scholar 

  • Cornet F, Burlet D: Stress field determinations in France by hydraulic tests in boreholes. J Geophys Res-Sol Ea 1992, 97(B8):11829–11849. 10.1029/90JB02638

    Article  Google Scholar 

  • Cornet F, Röckel T: Vertical stress profiles and the significance of "stress decoupling". Tectonophysics 2012, 581: 193–205. 10.1016/j.tecto.2012.01.020

    Article  Google Scholar 

  • Cornet F, Berard T, Bourouis S: How close to failure is a ganite rock mass at a 5 km depth? International Journal of Rock Mechanics and Mining Sciences 2007, 44(1):47–66. 10.1016/j.ijrmms.2006.04.008

    Article  Google Scholar 

  • Cuenot N, Charlety J, Dorbath L, Haessler H: Faulting mechanisms and stress regime at the European HDR site of Soultz-sous-Forets, France. Geothermics 2006, 35(5–6):561–575. 10.1016/j.geothermics.2006.11.007

    Article  Google Scholar 

  • Delouis B, Haessler H, Cisternas A, Rivera L: Stress tensor determination in France and neighboring regions. Tectonophysics 1993, 221(3–4):413–438. 10.1016/0040-1951(93)90171-F

    Article  Google Scholar 

  • Eisbacher GH, Becker A, Clauß B: Geothermisches Projekt Bruchsal, Phase 5 und 6 : Tektonische Untersuchungen. Geologisches Institut und Geophysikalisches Institut, Universität Karlsruhe. 1989.

    Google Scholar 

  • Genter A, Evans K, Cuenot N, Fritsch D, Sanjuan B: Contribution of the exploration of deep crystalline fractured reservoir of Soultz to the knowledge of enhanced geothermal systems (EGS). Cr Geosci 2010, 342(7–8):502–516. 10.1016/j.crte.2010.01.006

    Article  Google Scholar 

  • Greiner G: In situ stress measurements in Southwest Germany. Tectonophysics 1975, 29(1–4):265–274. 10.1016/0040-1951(75)90150-X

    Article  Google Scholar 

  • Greiner G: Spannungen in der Erdkruste: Bestimmung und Interpretation am Beispiel von In-situ-Messungen im Süddeutschen Raum. Karlsruhe. 1978.

    Google Scholar 

  • Heidbach O, Tingay M, Barth A, Reinecker J, Kurfeß D, Müller B: The World Stress Map database release. 2008, ᅟ: ᅟ.

    Google Scholar 

  • Heidbach O, Tingay M, Barth A, Reinecker J, Kurfess D, Muller B: Global crustal stress pattern based on the World Stress Map database release 2008. Tectonophysics 2010, 482(1–4):3–15. 10.1016/j.tecto.2009.07.023

    Article  Google Scholar 

  • Hergert T, Heidbach O: Geomechanical model of the Marmara Sea region-II. 3-D contemporary background stress field. Geophys J Int 2011, 185(3):1090–1102. 10.1111/j.1365-246X.2011.04992.x

    Article  Google Scholar 

  • Illies J: Recent and paleo-intraplate tectonics in stable Europe and the Rhinegraben rift system. Tectonophysics 1975, 29(1–4):251–264. 10.1016/0040-1951(75)90149-3

    Article  Google Scholar 

  • Illies J, Greiner G: Holocene movements and state of stress in the rhinegraben rift system. Tectonophysics 1979, 52(1–4):349–359. 10.1016/0040-1951(79)90245-2

    Article  Google Scholar 

  • Jaeger JC, Cook NGW, Zimmermann RW: Fundamentals of Rock Mechanics. Blackwell Publishing, Malden MA, USA; 2007.

    Google Scholar 

  • Krantz R: Measurements of friction coefficients and cohesion for faulting and fault reactivation in laboratory models using sand and sand mixtures. Tectonophysics 1991, 188(1–2):203–207. 10.1016/0040-1951(91)90323-K

    Article  Google Scholar 

  • Larroque J, Laurent P: Evolution of the stress field pattern in the south of the Rhine Graben from the Eocene to the present. Tectonophysics 1988, 148(1–2):41–58. 10.1016/0040-1951(88)90159-X

    Article  Google Scholar 

  • Larroque J, Etchecopar A, Philip H: Evidence for the permutation of stresses s1 and s2 in the Alpine foreland - the example of the Rhine Graben. Tectonophysics 1987, 144(4):315–322. 10.1016/0040-1951(87)90299-X

    Article  Google Scholar 

  • Moos D, Zoback M: Utilization of observations of well bore failure to constrain the orientation and magnitude of crustal stresses: application to continental, deep-sea drilling project, and ocean drilling program boreholes. Journal of Geophysical Research-Solid Earth and Planets 1990, 95(B6):9305–9325. 10.1029/JB095iB06p09305

    Article  Google Scholar 

  • Mueller K: Anwendung moderner gravimetrischen Interpretationstechniken zur Erkundung der Kruste von Suddeutschland. Ph.D., Technical Univ, Clausthal, Germany; 1988.

    Google Scholar 

  • Müller B, Zoback M, Fuchs K, Mastin L, Gregersen S, Pavoni N, Stephansson O, Ljunggren C: Regional patterns of tectonic stress in Europe. J Geophys Res-Sol Ea 1992, 97(B8):11783–11803. 10.1029/91JB01096

    Article  Google Scholar 

  • Pflug R: Bau und Entwicklung des Oberrheingrabens. Erträge der Forschung 1982, 184: 145. Darmstadt (Wiss. Buchges.)

    Google Scholar 

  • Plaumann S: Rock density in the upper Rhine Valley. The Rhinegraben Progress Report 1967. Mem. Ser. Cart. Ceol. Alsace Lorraine 1967, 26: 114–115.

    Google Scholar 

  • Plenefisch T, Bonjer K: The stress field in the Rhine Graben area inferred from earthquake focal mechanisms and estimation of frictional parameters. Tectonophysics 1997, 275(1–3):71–97. 10.1016/S0040-1951(97)00016-4

    Article  Google Scholar 

  • Ritter JRR, Wagner M, Bonjer KP, Schmidt B: The 2005 Heidelberg and Speyer earthquakes and their relationship to active tectonics in the central Upper Rhine Graben. Int J Earth Sci 2009, 98(3):697–705. 10.1007/s00531-007-0284-x

    Article  Google Scholar 

  • Rotstein Y, Schaming M: The Upper Rhine Graben (URG) revisited: Miocene transtension and transpression account for the observed first-order structures. Tectonics 2011, 30:. 10.1029/2010TC002767

    Google Scholar 

  • Rotstein Y, Behrmann JH, Lutz M, Wirsing G, Luz A: Tectonic implications of transpression and transtension: Upper Rhine Graben. Tectonics 2005, 24(6):ᅟ. 10.1029/2005TC001797

    Article  Google Scholar 

  • Rotstein Y, Edel JB, Gabriel G, Boulanger D, Schaming M, Munschy M: Insight into the structure of the Upper Rhine Graben and its basement from a new compilation of Bouguer Gravity. Tectonophysics 2006, 425(1–4):55–70. 10.1016/j.tecto.2006.07.002

    Article  Google Scholar 

  • Schumacher ME: Upper Rhine Graben: role of preexisting structures during rift evolution. Tectonics 2002, 21(1):1006. 10.1029/2001TC900022

    Article  Google Scholar 

  • Sissingh W: Comparative tertiary stratigraphy of the Rhine Graben, Bresse Graben and Molasse Basin: correlation of Alpine foreland events. Tectonophysics 1998, 300(1–4):249–284. 10.1016/S0040-1951(98)00243-1

    Article  Google Scholar 

  • Townend J, Zoback M: How faulting keeps the crust strong. Geology 2000, 28(5):399–402. 10.1130/0091-7613(2000)28<399:HFKTCS>2.0.CO;2

    Article  Google Scholar 

  • Valley BC: The relation between natural fracturing and stress heterogeneities in deep-seated crystalline rocks at Soultz-sous-Forêts (France). 2007.

    Google Scholar 

  • Valley B, Evans K: Stress state at Soultz-sous-Forêts to 5 km depth from wellbore failure and hydraulic observations. In Proceedings of the 32nd Workshop on Geothermal Reservoir Engineering. Stanford University, Stanford, CA, USA; 2007.

    Google Scholar 

  • Valley B, Evans K: Stress orientation to 5 km depth in the basement below Basel (Switzerland) from borehole failure analysis. Swiss J Geosci 2009, 102(3):467–480. 10.1007/s00015-009-1335-z

    Article  Google Scholar 

  • Wileveau Y, Cornet F, Desroches J, Blumling P: Complete in situ stress determination in an argillite sedimentary formation. Physics and Chemistry of the Earth 2007, 32(8–14):866–878. 10.1016/j.pce.2006.03.018

    Article  Google Scholar 

  • Wirth E: Die Erdöllagerstätten Badens. Abh. geol. 1962, 4–5:. Landesamtes Baden-Württemberg, Freiburg i.Br

    Google Scholar 

  • Zang A, Stephansson O: Stress Fields of the Earth's Crust. Springer, Berlin; 2010.

    Book  Google Scholar 

  • Ziegler P: Collision related intra-plate compression deformations in Western and Central Europe. Journal of Geodynamics 1990, 11(4):357–388. 10.1016/0264-3707(90)90017-O

    Article  Google Scholar 

  • Zoback M, Barton C, Brudy M, Castillo D, Finkbeiner T, Grollimund B, Moos D, Peska P, Ward C, Wiprut D: Determination of stress orientation and magnitude in deep wells. International Journal of Rock Mechanics and Mining Sciences 2003, 40(7–8):1049–1076. 10.1016/j.ijrmms.2003.07.001

    Article  Google Scholar 

Download references

Acknowledgements

We would like to thank EnBW Energie Baden-Württemberg AG for providing the well logs and leak-off test data. Financial support has been kindly granted by the Federal Ministry for the Environment, Nature Conversation, Building and Nuclear Safety (BMU), and EnBW Energie Baden-Württemberg AG within the framework of the Logro project (FKZ 0325111 A). Further supports are from the Helmholtz Portfolio GEOENERGIE and the geothermal research center of Baden Württemberg (LFZG). We would like to thank I. Moeck, F. Cornet, B. Müller, and T. Röckel for fruitful discussions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jörg Meixner.

Authors’ original submitted files for images

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 2.0 International License (https://creativecommons.org/licenses/by/2.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Meixner, J., Schill, E., Gaucher, E. et al. Inferring the in situ stress regime in deep sediments: an example from the Bruchsal geothermal site. Geotherm Energy 2, 7 (2014). https://doi.org/10.1186/s40517-014-0007-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40517-014-0007-z

Keywords