Use of Mohr Diagrams to Predict Fracturing in a Potential Geothermal Reservoir
Abstract
:1. Introduction
- Which lithologies are most likely to fracture?
- What stresses and fluid pressures are needed for the reactivation of pre-existing fractures or the development of new fractures?
- Which orientations and types of fractures are most likely to be reactivated?
- Will reactivated or new fractures show shear or extension?
- What effects do heterogeneities (veins, joints, cleavage, bedding planes) have, and what are the different mechanical significances of veins vs. joints?
- What are the effects of Late Cretaceous and Tertiary exhumation and what amount of exhumation is needed to create joints?
2. Predictions about the Pre-Permian Geology beneath Göttingen
Factor | Meaning | Significance | Bunter Sandstone | Variscan rocks |
---|---|---|---|---|
Lithology | Rock types, their porosities and mechanical behaviour | Controls the thermo-mechanical behaviour of the rock. Mechanical behaviour can change significantly through time, especially as deformation occurs | Triassic sandstone (see Table 2 for mechanical properties) | Devonian and Carboniferous greywackes and slates (see Table 2 for mechanical properties) |
Fluid type | The chemistry and phase (liquid or gas) of the palaeo- and present-day fluid(s) | Controls the fluid pressure gradient and mineralisation events | Present-day: water (possibly brine) | Present-day: water, probably saline because of the overlying Zechstein. No information on gas content. During the Variscan: mineralising fluids and fluidised sediments |
Stress | Magnitudes and orientations of the applied stresses, including the vertical stress (overburden) and horizontal stresses. Horizontal stresses are related to the geostatic stress ratio, applied tectonic stresses and to internal stresses (e.g., related to temperature changes) | Along with fluid pressure, controls the effective stresses, which control the deformation | The vertical stress can be calculated using the mean density of the overburden, but the magnitudes and orientations of the horizontal stresses are uncertain | The vertical stress can be calculated using the mean density of the overburden, but the magnitudes and orientations of the horizontal stresses are uncertain |
Fluid pressure | Palaeo- and present-day fluid pressures | Along with the stresses, controls the effective stresses, which control the deformation | Presently probably hydrostatic | The Zechstein evaporites may allow present-day overpressure. Veins, breccias and possible remobilised sediments indicate phases of overpressure during the Variscan Orogeny |
Temperature | Palaeo- and present-day temperatures | Influences the style of deformation, with present-day temperature controlling commercial viability | Depends on the geothermal gradient | Depends on the geothermal gradient, but likely to be reduced because of the overlying salt. Possibly elevated by Tertiary igneous activity |
Strain | The amount of strain and the existing structures | Influences fluid flow in the sub-surface and present-day mechanical behaviour of the rocks | Controlled by Tertiary rifting. Likely to be influenced by salt tectonics and possibly by Tertiary igneous activity. Steeply-dipping joints and some normal faults are likely to occur | Dominated by: (1) Variscan Orogeny, with folds, thrusts and veins; (2) Late Cretaceous and Tertiary rifting and/or uplift, with normal faults and joints developing |
History | The relative and absolute timing of deformation (including mineralisation) events and structures | Controls the types of fractures (faults, veins, joints, etc.) and therefore their effects on fluid flow in the sub-surface | (1) Triassic sedimentation during Mesozoic basin development. (2) Cretaceous and Tertiary regional uplift. (3) Tertiary rifting (Leinetal Graben) and volcanism | (1) Sedimentation during the Devonian and Carboniferous. (2) Variscan Orogeny. (3) Permian and Mesozoic sedimentation and basin development. (4) Cretaceous and Tertiary regional uplift. (5) Tertiary rifting and volcanism |
3. Model Set-Up
3.1. Mohr Diagrams, Stresses and Failure Envelopes
- The rock properties used to define the failure envelope are the tensile strength (T), uniaxial compressive strength (UCS), cohesion (S0), and coefficient of internal friction (μ) [75];
- The stress state (σ), which is defined in terms of principal stresses, mean stress and differential stress [17]. Stresses are in turn controlled by factors, such as depth of burial (overburden), tectonic (horizontal) stresses, and other changes in the physical state of the material, such as expansion or contraction caused by temperature and volume change (e.g., [76]). Changes in stresses that lead to fracturing can either be by increasing [77] or reducing the [78] the applied compressive stresses;
- In the upper crust, fluid pressure in the pores and cracks combines with the applied stresses to produce an effective stress, where σ′ = σ − PF (e.g., [79,80,81]). In the absence of specific information, we use a Biot coefficient (B) of 1, where σ′ = σ − B.PF [82]. Changes in fluid pressure that can lead to fracturing can either be an increase in fluid pressure (e.g., [83]) or a reduction in fluid pressure, which can cause pore collapse (e.g., [84,85]). Pore collapse is not considered further in this paper.
3.2. Input Data, Assumptions, and Uncertainties
- An Andersonian stress system is assumed, i.e., with one of the principal axes of stress being vertical and the other two being horizontal [64];
- The analysis is carried out in two-dimensions, considering just vertical stress and horizontal stress. This simplifies the analysis and is, we argue, justified at the pre-drilling stage of analysis because of the magnitudes and orientations of the horizontal stresses are currently unknown. Hydrofracture data from three wells in the region suggest a thrust regime with a maximum horizontal stress orientated ~WNW-ESE [93];
- The vertical stress is produced by the weight of overburden;
- The fluids are hydrostatically pressured;
- The failure parameters used in the modelling (Table 3) are assumed to be representative of the rock properties in the sub-surface.
3.3. Base Case Models
3.4. A Range of Stress States for Fracturing
4. Effects of Key Parameters
5. Potential for Reactivating and Generating Fractures
5.1. Bunter Sandstone
Bunter Sandstone K0 = 0.19 | Bunter Sandstone K0 = 0.54 | Greywacke K0 = 0.125 | Greywacke K0 = ratio = 0.41 | Slate K0 = 0.283 | Slate K0 = 0.41 | Units | Notes | |
---|---|---|---|---|---|---|---|---|
Fluid pressure, cohesionless | 9.81 | 24 | 19.62 44.15 | 50 115 | 51 115 | 51 115 | MPa | Shear on favourably-orientated cohesionless fractures |
Fluid pressure, gently-dipping cohesional | 29.5 | 30 | 72 138 | 72 138 | 68 134 | 68 134 | MPa | Gently-dipping extension fracture develop |
Fluid pressure, all orientations of extension fractures | 54 | 34.4 | 210 278 | 100 168 | 105 172 | 90 155 | MPa | Steep extension fractures develop |
a. Base case | Unstable | Stable | Unstable Unstable | Stable | Stable | Stable | MPa | Stable stress state for higher K0 |
b. Decreased tectonic stress, reactivation starts | 0 | −3.7 | 0 0 | −7.5 −17 | −2 −5 | −6.5 −14 | MPa | Shear on favourably-orientated cohensionless fractures |
c. Decreased tectonic stress, many fractures reactivated | −5 | −8 | −10 −20 | −20 −30 | −10 −22 | −14 −31 | MPa | Shear on a cohesionless fractures with a wide range of orientations |
d. Reduced tectonic stress, extension fractures develop | −8.2 | −13.5 | −24 −29 | −33 −50 | −24 −33 * | −28.5 −43 * | MPa | Extension fractures develop perpendicular to least compressive stress |
e. Reduced tectonic stress, extension fractures in all orientations | N/A | N/A | N/A N/A | N/A N/A | N/A N/A | N/A N/A | MPa | Requires increase in fluid pressure |
f. Increased tectonic stress, some reactivation of cohesionless | 48 | 43 | 0 (normal) 180 (thrusts) 0 (normal) 400 (thrusts) | 170 370 | 145 320 | 140 315 | MPa | Reactivation of cohesionless fractures in shear |
g. Increased tectonic stress, new shear fractures can develop | 95 | 90 | 370 590 | 360 580 | 275 450 | 270 440 | MPa | Creation of new shear fractures |
5.2. Devonian and Carboniferous Greywackes and Slates
- Very low geostatic stress ratios (e.g., 0.125) in the greywackes and slates are required to reactivate cohesionless fractures without fluid overpressure or applied tectonic stresses (base-case model; e.g., Figure 3a);
- A fluid pressure of about 50 MPa would be needed to reactivate cohesionless fractures in both the greywackes and slates at a depth of 2 km if the geostatic stress ratio is high (e.g., 0.41; e.g., Figure 3b), with fluid pressures reaching lithostatic pressures;
- Gently-dipping extension fractures will start to develop in the greywackes at a depth of 2 km if the fluid pressure is about 72 MPa, but may develop in the slates at slightly lower pressures (about 68 MPa). The models predict that, in the absence of cohesionless fractures, increasing fluid pressure will initially create gently-dipping extension fractures. This is because of the assumption of uniaxial strain (i.e., that the rocks are laterally confined). Higher fluid pressures will be required to generate steeply-dipping extension fractures, in the absence of horizontal tensile stresses, such as those related to tectonic forces or to cooling (Section 6);
- Steeply-dipping extension fractures are predicted to develop in the greywackes at 2 km depth if there is a tensile tectonic stress between about −24 and −40 MPa, but are likely to develop in the slates at tensile tectonic stress between about −24 to −33 MPa (e.g., Figure 3d);
- Extension fractures in all directions may develop in the greywackes at 2 km depth if the fluid pressure is between about 100 and 210 MPa, but would develop in the slates if the fluid pressure is between about 90 and 105 MPa (e.g., Figure 3e);
- Shear (normal faulting) may begin on favourably-orientated cohesionless fractures in the greywackes at a depth of 2 km if there is a tensile tectonic stress between about 0 and −7.5 MPa, and in the slates if there is a tensile tectonic stress between about −2 and −6.5 MPa (e.g., Figure 3b);
- Favourably-orientated cohesionless fractures may be reactivated as thrusts in the greywackes at 2 km depth if there is a compressive tectonic stress of about 170 MPa, and in the slates at about 140 MPa (e.g., Figure 3f);
- New thrusts will begin to develop in the greywackes at a depth of 2 km if there is a compressive tectonic stress of about 360 MPa, but will develop in the slates if there is a compressive tectonic stress of about 270 MPa (e.g., Figure 3g).
5.3. Possible Effects of Exhumation on the Bunter Sandstone
Parameter | Value | Unit |
---|---|---|
Density | 2.68 | g/cm3 |
Porosity | 10 | % |
Tensile strength | 6 | MPa |
Cohesion | 12 | MPa |
Poisson ratio | 0.25 | |
Geostatic stress ratio | 0.333 | |
Young’s modulus | 22 | GPa |
Coefficient of thermal expansion | 11.25 | 10−6 °C−1 |
Geothermal gradient | 30 | °C per km |
Fluid pressure | Hydrostatic | |
Tectonic stresses | Zero | |
Initial top of unit | 1.4 | km |
Initial base of unit | 2 | km |
6. Discussion
6.1. Possible Effects of Stimulation on Devonian and Carboniferous Rocks at 2 km Depth
- Shear can occur along favourably-orientated cohesionless fractures (e.g., joints) in the slates at fluid pressures of about 50 MPa, which is an overpressure of about 30 MPa (Figure 7b);
- Gently-dipping extension fractures can be generated in the slates a fluid pressure of about 68 MPa, and gently-dipping cohesionless fractures in the greywackes may be reactivated as extension fractures (Figure 7c);
- Gently-dipping extension fractures can be generated in the greywackes at a fluid pressure of about 72 MPa (Figure 7d);
- Extension fractures with any orientation may develop in the slates at fluid pressure between about 90 and 105 MPa (Figure 7e);
- Extension fractures with any orientation may develop in the greywackes if the fluid pressure is between about 100 and 210 MPa (Figure 6f).
6.2. Possible Effects of Stimulation on Devonian and Carboniferous Rocks at 4.5 km Depth
6.3. Possibility of Open Fractures in the Devonian and Carboniferous Rocks below Göttingen
6.4. Potential Benefits, Problems, and Improvements
- The Mohr diagram models used give little direct information about potential fluid flow in the sub-surface. The approach could, however, be used in combination with other modelling approaches. For example, it would be useful to compare predictions of critically stressed fractures from Mohr diagrams with distinct element analysis of fracture networks (e.g., [106,107]);
- The values for rock properties used are based on triaxial tests, which are probably over-estimates because small, unfractured samples are generally used (e.g., [86]). More accurate methods for estimating the material properties of rock masses are available (e.g., [108]), and these methods could be used when more detailed information becomes available about the fracture patterns in the rock mass;
- Similarly, we have used rock mechanical properties from the literature and have made various simplifying assumptions (e.g., no applied tectonic stresses, Biot coefficient = 1). The modelling can be improved and the assumptions properly tested as the input parameters become better constrained, for example as borehole data become available;
- The anisotropy of the slates has not been modelled in a sophisticated way here, and this can be improved using more detailed information about the relationships between the angle between in situ stresses and cleavage (e.g., [109]);
- The Mohr diagram analysis used here is two-dimensional, mainly because the magnitudes and orientations of the horizontal stresses are unknown. The analysis could be expanded to three dimensions when such information becomes available, for example from well data. Although predictions can be made about the stresses involved in the Variscan Orogeny and the formation of the Leinetal Graben, those predictions do not help with making predictions about the present-day stresses.
7. Conclusions
Author Contributions
Funding
Data Availability Statement
Acknowledgments
Conflicts of Interest
References
- Leiss, B.; Tanner, D.; Vollbrecht, A.; Wemmer, K. Tiefengeothermisches Potential in der Region Göttingen—Geologische Rahmenbedingungen. In Neue Untersuchungen zur Geologie der Leinetalgrabenstruktur; Leiss, B., Tanner, D., Vollbrecht, A., Arp, G., Eds.; Universitätsverlag Göttingen: Göttingen, Germany, 2011; pp. 163–170. [Google Scholar]
- Leiss, B.; Romanov, D.; Wagner, B. Risks and challenges of the transition to an integrated geothermal concept for the Göttingen University Campus. In Proceedings of the European Association of Geoscientists and Engineers, Conference Proceedings, 1st Geoscience and Engineering in Energy Transition Conference, Strasbourg, France, 16–18 November 2020; pp. 1–5. [Google Scholar] [CrossRef]
- Leiss, B.; Wagner, B.; Heinrichs, T.; Romanov, D.; Tanner, D.C.; Vollbrecht, A.; Wemmer, K. Integrating deep, medium and shallow geothermal energy into district heating and cooling system as an energy transition approach for the Göttingen University Campus. In Proceedings of the World Geothermal Congress 2021, Reykjavik, Iceland, 24–27 October 2021; pp. 1–9. [Google Scholar]
- Arp, G.; Vollbrecht, A.; Tanner, D.C.; Leiss, B. Zur Geologie des Leintalgrabens—Ein kurzer Überblick. In Neue Untersuchungen zur Geologie der Leinetalgrabenstruktur; Leiss, B., Tanner, D., Vollbrecht, A., Arp, G., Eds.; Universitätsverlag Göttingen: Göttingen, Germany, 2011; pp. 1–7. [Google Scholar]
- Trullenque, G.; Genter, A.; Leiss, B.; Wagner, B.; Bouchet, R.; Léoutre, E.; Malnar, B.; Bär, K.; Rajšl, I. Upscaling of EGS in different geological conditions: a European perspective. In Proceedings of the 43rd Workshop on Geothermal Reservoir Engineering, Stanford University, Stanford, CA, USA, 12–14 February 2018. [Google Scholar]
- Dalmais, E.; Genter, A.; Trullenque, G.; Leoutre, E.; Leiss, B.; Wagner, B.; Mintsa, A.C.; Bär, K.; Rajsl, I. MEET Project: toward the spreading of EGS across Europe. In Proceedings of the European Geothermal Congress, Den Haag, The Netherlands, 11–14 June 2019; p. 8. [Google Scholar]
- Leiss, B.; Wagner, B.; MEET Konsortium. EU-Projekt MEET: Neue Ansätze »Enhanced Geothermal Systems (EGS)«—Göttinger Unicampus als Demoprojekt. Geotherm. Energ. 2020, 91, 26–28. [Google Scholar]
- Welsch, B. Technical, Environmental and Economic Assessment of Medium Deep Borehole Thermal Energy Storage Systems. Ph.D. Thesis, Technische Universität Darmstadt, Darmstadt, Germany, 2019; 192p. [Google Scholar]
- Krzywiec, P. Triassic-Jurassic evolution of the Pomeranian segment of the Mid-Polish Trough—Basement tectonics and subsidence patterns. Geol. Quart. 2006, 50, 139–150. [Google Scholar]
- Doornenbal, H.; Stevenson, A. (Eds.) Petroleum Geological Atlas of the Southern Permian Basin Area; EAGE: Houten, The Netherlands, 2010; 342p. [Google Scholar]
- Arp, G.; Hoffmann, V.E.; Seppelt, S.; Riegel, W. Trias und Jura von Göttingen und Umgebung. In Geobiologie 2—Jahrestagung der Paläontologischen Gesellschaft; Reitner, J., Reich, M., Schmidt, G., Eds.; Universitätsverlag Göttingen: Göttingen, Germany, 2004; Volume 74, pp. 147–192. [Google Scholar]
- Leiss, B.; Tanner, D.; Vollbrecht, A.; Arp, G. (Eds.) Neue Untersuchungen zur Geologie der Leinetalgrabenstruktur; Universitätsverlag Göttingen: Göttingen, Germany, 2011; 170p. [Google Scholar]
- Kirnbauer, T. (Ed.) Geologische Exkursionen in die Region um Göttingen; Jahresberichte und Mitteilungen des Oberrheinischen geologischen Vereins; E. Schweizerbart‘sche Verlagsbuchhandlung: Stuttgart, Germany, 2013; Volume 95, 319p. [Google Scholar]
- Sanderson, D.J. Field-based structural studies as analogues to sub-surface reservoirs. In The Value of Outcrop Studies in Reducing Subsurface Uncertainty and Risk in Hydrocarbon Exploration and Production; Bowman, M., Smyth, H.R., Good, T.R., Passey, S.R., Hirst, J.P.P., Jordan, C.J., Eds.; Geological Society London, Special Publications: London, UK, 2015; Volume 436. [Google Scholar]
- Martel, S.J. Progress in understanding sheeting joints over the past two centuries. J. Struct. Geol. 2017, 94, 68–86. [Google Scholar] [CrossRef] [Green Version]
- Fairhurst, C. Fractures and fracturing: hydraulic fracturing in jointed rock. In Effective and Sustainable Hydraulic Fracturing; Jeffrey, R., Ed.; IntechOpen Limited: London, UK, 2013; pp. 47–79. ISBN 978-953-51-6341-1. [Google Scholar]
- Ramsay, J.G. The Folding and Fracturing of Rocks; McGraw-Hill: New York, NY, USA, 1967; 568p. [Google Scholar]
- Fossen, H. Structural Geology, 2nd ed.; Cambridge University Press: Cambridge, UK, 2016; 524p. [Google Scholar]
- Mahmoodpour, S.; Singh, M.; Turan, A.; Bär, K.; Sass, I. Key parameters affecting the performance of fractured geothermal reservoirs: a sensitivity analysis by thermo-hydraulic-mechanical simulation. Geophysics. in review.
- Friedel, C.H.; Huckriede, H.; Leiss, B.; Zweig, M. Large-scale Variscan shearing at the southeastern margin of the eastern Rhenohercynian belt: a reinterpretation of chaotic rock fabrics in the Harz Mountains, Germany. Int. J. Earth Sci. 2019, 108, 2295–2323. [Google Scholar] [CrossRef]
- Fielitz, W. Variscan transpressive inversion in the northwestern central Rhenohercynian belt of western Germany. J. Struct. Geol. 1992, 14, 547–563. [Google Scholar] [CrossRef]
- Franke, W. The mid-European segment of the Variscides: tectonostratigraphic units, terrane boundaries and plate tectonic evolution. In Orogenic Processes: Quantification and Modelling in the Variscan Belt; Franke, W., Haak, V., Oncken, O., Tanner, D., Eds.; Geological Society of London, Special Publications: London, UK, 2000; Volume 179, pp. 35–61. [Google Scholar]
- Franke, W. Rheno-Hercynian belt of central Europe: review of recent findings and comparisons with south-west England. Geosci. SW Eng. 2007, 11, 263–272. [Google Scholar]
- Friedel, C.H.; Hoffmann, C. Stopp 4: Schichtgebundene Deformationsstrukturen in der Kulmgrauwacke des Oberharzes (Bundesstraße 242). In Harzgeologie 2016. 5. Workshop Harzgeologie Kurzfassungen und Exkursionsführer; Friedel, C.H., Leiss, B., Eds.; Universitätsverlag Göttingen: Göttingen, Germany, 2016; pp. 85–90. [Google Scholar]
- Zeuner, M. Conceptual 3D-Structure Model of the Variscan “Culm Fold Zone” in the Vicinity of the Oktertal dam (Harz Mountains, Germany) in Terms of Geothermal Reservoir Development. Master’s Thesis, Georg-August-Universität Göttingen, Göttingen, Germany, 2018. [Google Scholar]
- Wagner, B.; Leiss, B.; Tanner, D.C. Stopp 2: Gefalteter unterkarbonischer Kieselschiefer am Bielstein nördlich Lautenthal (Innerstetal). In Harzgeologie 2016. 5. Workshop Harzgeologie Kurzfassungen und Exkursionsführer; Friedel, C.H., Leiss, B., Eds.; Universitätsverlag Göttingen: Göttingen, Germany, 2016; pp. 69–78. [Google Scholar]
- Franzke, H.J.; Hauschke, N.; Hellmund, M. Spätpleistozäne bis holozäne Tektonik an der Harznordrand-Störung bei Benzingerode (Sachsen-Anhalt). In Harzgeologie 2016. 5. Workshop Harzgeologie Kurzfassungen und Exkursionsführer; Friedel, C.H., Leiss, B., Eds.; Universitätsverlag Göttingen: Göttingen, Germany, 2016; pp. 13–17. [Google Scholar]
- De Graaf, S.; Lüders, V.; Banks, D.A.; Sośnicka, M.; Reijmer, J.J.G.; Kaden, H.; Vonhof, H.B. Fluid evolution and ore deposition in the Harz Mountains revisited: isotope and crush-leach analyses of fluid inclusions. Miner. Depos. 2020, 55, 47–62. [Google Scholar] [CrossRef] [Green Version]
- Ramsey, J.M.; Chester, F.M. Hybrid fracture and the transition from extension fracture to shear fracture. Nature 2004, 428, 63–66. [Google Scholar] [CrossRef]
- Wagner, B.; Günther, S.; Ford, K.; Sosa, G.; Leiss, B. The “Hexagon concept”: a fundamental approach for the geoscientific spatial data compilation and analysis at European scale to define the geothermal potential of Variscan and pre-Variscan low- to high-grade metamorphic and intrusive rocks. In Proceedings of the World Geothermal Congress 2021, Reykjavik, Iceland, 24–27 October 2021; pp. 1–10. [Google Scholar]
- Arp, G.; Tanner, D.C.; Leiss, B. Struktur der Leinetalgraben-Randstörung bei Reiffenhausen (Autobahn 38 Heidkopftunnel-Westportal). In Neue Untersuchungen zur Geologie der Leinetalgrabenstruktur; Leiss, B., Tanner, D., Vollbrecht, A., Arp, G., Eds.; Universitätsverlag Göttingen: Göttingen, Germany, 2011; pp. 17–21. [Google Scholar]
- Tanner, D.C. Exkursion nördliches Leinetal. In Neue Untersuchungen zur Geologie der Leinetalgrabenstruktur; Leiss, B., Tanner, D., Vollbrecht, A., Arp, G., Eds.; Universitätsverlag Göttingen: Göttingen, Germany, 2011; pp. 27–38. [Google Scholar]
- Arp, G.; Bielert, F.; Hoffmann, V.E.; Löffler, T. Palaeoenvironmental significance of lacustrine stromatolites of the Arnstadt Formation (“Steinmergelkeuper”, Upper Triassic, N-Germany). Facies 2005, 51, 419–441. [Google Scholar] [CrossRef]
- Ritter, M.; Vollbrecht, A.; van den Kerkhof, A.; Wemmer, K. Sedimentgänge im Bausandstein der Solling-Folge NW’ von Billingshausen. In Neue Untersuchungen zur Geologie der Leinetalgrabenstruktur; Leiss, B., Tanner, D., Vollbrecht, A., Arp, G., Eds.; Universitätsverlag Göttingen: Göttingen, Germany, 2011; pp. 89–106. [Google Scholar]
- Reyer, D.; Bauer, J.F.; Philipp, S.L. Fracture systems in normal fault zones crosscutting sedimentary rocks, Northwest German Basin. J. Struct. Geol. 2012, 45, 38–51. [Google Scholar] [CrossRef]
- Grupe, O. Uber die Zechsteinformation und ihr Salzlager im Untergrunde des hannoverschen Eichsfeldes und angrenzenden Leinegebietes nach den neueren Bohrergebnissen. Z. Fur Prakt. Geol. 1909, 17, 185–205. [Google Scholar]
- Tanner, D.C.; Musmann, P.; Wawerzinek, B.; Buness, H.; Krawczyk, C.M.; Thomas, R. Salt tectonics of the eastern border of the Leinetal Graben, Lower Saxony, Germany, as deduced from seismic reflection data. Interpretation 2015, 3, T169–T181. [Google Scholar] [CrossRef]
- Tanner, D.C.; Leiss, B.; Vollbrecht, A. Strukturgeologie des Leinetalgrabens (Exkursionen G1 und G2 am 4. und 5. April 2013); Jahresberichte und Mitteilungen des Oberrheinischen Geologischen Vereins Band; E. Schweizerbart’sche Verlagsbuchhandlung: Stuttgart, Germany, 2013; Volume 95, pp. 131–168. [Google Scholar]
- Brink, H. The Variscan Deformation Front (VDF) in northwest Germany and its relation to a network of geological features including the ore-rich Harz Mountains and the European Alpine Belt. Int. J. Geosci. 2021, 12, 447–486. [Google Scholar] [CrossRef]
- Vincent, C.J. Porosity of the Bunter Sandstone in the Southern North Sea Basin Based on Selected Borehole Neutron Logs; IR/05/074; British Geological Survey Internal Report: Keyworth, UK, 2005; 20p. [Google Scholar]
- McNamara, D.D.; Faulkner, D.; McCarney, E. Rock properties of greywacke basement hosting geothermal reservoirs, New Zealand: preliminary results. In Proceedings of the Thirty-Ninth Workshop on Geothermal Reservoir Engineering Stanford University, Stanford, CA, USA, 24–26 February 2014. [Google Scholar]
- Gholami, R.; Rasouli, V. Mechanical and elastic properties of transversely isotropic slate. Rock Mech. Rock Eng. 2013, 47. [Google Scholar] [CrossRef]
- Menezes, F.F.; Lempp, C.; Svensson, K.; Neumann, A.; Pöllmann, H. Geomechanical behavior changes of Bunter Sandstone and borehole cement due to scCO2 injection effects. In Proceedings of the IAEG/AEG Annual Meeting, San Francisco, CA, USA, 17–21 September 2018; Shakoor, A., Cato, K., Eds.; Springer Nature: Cham, Switzerland, 2019; Volume 1, pp. 111–118. [Google Scholar]
- Agustawijaya, D.S. The uniaxial compressive strength of soft rock. Civil Eng. Dimen. 2007, 9, 9–14. [Google Scholar]
- Menezes, F.F.; Lempp, C. On the structural anisotropy of physical and mechanical properties of a Bunter Sandstone. J. Struct. Geol. 2018, 114, 196–205. [Google Scholar] [CrossRef]
- Havlíčková, D.; Závacký, M.; Krmíček, L. Anisotropy of mechanical properties of greywacke. GeoSci. Eng. 2019, 65, 46–52. [Google Scholar] [CrossRef]
- Rodríguez-Sastre, M.A.; Gutiérrez-Claverol, M.; Torres-Alonso, M. Relationship between cleavage orientation, uniaxial compressive strength and Young’s modulus for slates in NW Spain. Bull. Eng. Geol. Environ. 2008, 67, 181–186. [Google Scholar] [CrossRef]
- Price, N.J. Fault and Joint Development in Brittle and Semi-Brittle Rock; Pergamon Press: Oxford, UK, 1966; 176p. [Google Scholar]
- Martel, S.J. Effects of cohesive zones on small faults and implications for secondary fracturing and fault trace geometry. J. Struct. Geol. 1997, 19, 835–847. [Google Scholar] [CrossRef]
- Alneasan, M.; Behnia, M.; Bagherpour, H. The effect of Poisson’s ratio on the creation of tensile branches around dynamic faults. J. Struct. Geol. 2020, 131, 103950. [Google Scholar] [CrossRef]
- Dobbs, M.R.; Cuss, R.J.; Ougier-Simonin, A.; Parkes, D.; Graham, C.C. Yield envelope assessment as a preliminary screening tool to determine carbon capture and storage viability in depleted southern north-sea hydrocarbon reservoirs. Int. J. Rock Mech. Min. Sci. 2018, 102, 15–27. [Google Scholar] [CrossRef]
- Jefferies, G.M.; Crooks, J.H.A.; Becker, D.E.; Hill, P.R. Independence of geostatic stress from overconsolidation in some Beaufort Sea clays. Can. Geotech. J. 1987, 24, 342–356. [Google Scholar] [CrossRef]
- Cai, J.; Du, G.; Ye, H.; Lei, T.; Xia, H.; Pan, H. A slate tunnel stability analysis considering the influence of anisotropic bedding properties. Adv. Mater. Sci. Eng. 2019, 2019, 4653401. [Google Scholar] [CrossRef] [Green Version]
- Griffiths, L.; Heap, M.J.; Xu, T.; Chen, C.; Baud, P. The influence of pore geometry and orientation on the strength and stiffness of porous rock. J. Struct. Geol. 2017, 96, 149–160. [Google Scholar] [CrossRef]
- Wang, H.; Mang, H.; Yuan, Y.; Pichler, B.L.A. Multiscale thermoelastic analysis of the thermal expansion coefficient and of microscopic thermal stresses of mature concrete. Materials 2019, 12, 2689. [Google Scholar] [CrossRef] [Green Version]
- Pogacnik, J.; Elsworth, D.; O’Sullivan, M.; O’Sullivan, J. A damage mechanics approach to the simulation of hydraulic fracturing/shearing around a geothermal injection well. Comput. Geotech. 2016, 71, 338–351. [Google Scholar] [CrossRef]
- Lee, C.; Park, J.W.; Park, C.; Park, E.S. Current status of research on thermal and mechanical properties of rock under high-temperature condition. Tunn. Undergr. Space 2015, 25, 1–23. [Google Scholar] [CrossRef] [Green Version]
- Engelder, T.; Fischer, M.P. Influence of poroelastic behavior on the magnitude of minimum horizontal stress, Sh, in overpressured parts of sedimentary basins. Geology 1994, 22, 949–952. [Google Scholar] [CrossRef]
- du Rouchet, J. Stress fields, a key to oil migration. AAPG Bull. 1981, 65, 74–85. [Google Scholar]
- Mesri, G.; Hayat, T.M. The coefficient of earth pressure at rest. Can. Geotech. J. 1993, 30, 647–666. [Google Scholar] [CrossRef]
- Gercek, H. Poisson’s ratio values for rocks. Int. J. Rock Mech. Min. Sci. 2007, 44, 1–13. [Google Scholar] [CrossRef]
- Jaeger, J.C. Mohr diagram. In Structural Geology and Tectonics; Encyclopaedia of Earth Science; Springer: Berlin, Germany, 1987. [Google Scholar]
- Hoek, E.; Martin, C.D. Fracture initiation and propagation in intact rock—A review. J. Rock Mech. Geotech. Eng. 2014, 6, 287–300. [Google Scholar] [CrossRef] [Green Version]
- Anderson, E.M. The Dynamics of Faulting; Oliver & Boyd: Edinburgh, UK, 1951. [Google Scholar]
- Cook, N.G.W. The failure of rock. Int. J. Rock Mech. Min. Sci. Geomech. Abs. 1965, 2, 389–403. [Google Scholar] [CrossRef]
- Ucar, R. Determination of shear failure envelope in rock masses. J. Geot. Eng. 1986, 112, 303–315. [Google Scholar] [CrossRef]
- Patel, S.; Martin, C.D. Effect of stress path on the failure envelope of intact crystalline rock at low confining stress. Minerals 2020, 10, 1119. [Google Scholar] [CrossRef]
- Sibson, R.H. Brittle failure mode plots for compressional and extensional tectonic regimes. J. Struct. Geol. 1998, 20, 655–660. [Google Scholar] [CrossRef] [Green Version]
- Brace, W.F. An extension of the Griffith theory of fracture to rocks. J. Geophys. Res. 1960, 65, 3477–3480. [Google Scholar] [CrossRef]
- McClintock, F.A.; Walsh, J.B. Friction on Griffith cracks in rocks under pressure. In Proceedings of the fourth U.S. National Congress of Applied Mechanics, Berkeley, CA, USA, 18–21 June 1962; Volume 2, pp. 1015–1021. [Google Scholar]
- Jaeger, J.C.; Cook, N.G.W. Fundamentals of Rock Mechanics; Methuen: London, UK, 1969; 515p. [Google Scholar]
- Chang, C.; Zoback, M.D.; Khaksar, A. Empirical relations between rock strength and physical properties in sedimentary rocks. J. Pet. Sci. Eng. 2006, 51, 223–237. [Google Scholar] [CrossRef]
- Zoback, M.D. Reservoir Geomechanics; Cambridge University Press: Cambridge, UK, 2007; 449p. [Google Scholar]
- Sibson, R.H. A note on fault reactivation. J. Struct. Geol. 1985, 7, 751–754. [Google Scholar] [CrossRef]
- Scholz, C.H. The Mechanics of Earthquakes and Faulting, 2nd ed.; Cambridge University Press: Cambridge, UK, 2002; 504p. [Google Scholar]
- Meissner, R.; Strehlau, J. Limits of stresses in continental crusts and their relation to the depth-frequency distribution of shallow earthquakes. Tectonics 1982, 1, 73–89. [Google Scholar] [CrossRef]
- Malama, B.; Kulatilake, P.H.S.W. Models for normal fracture deformation under compressive loading. Int. J. Rock Mech. Min. Sci. 2003, 40, 893–901. [Google Scholar] [CrossRef]
- Corcoran, D.V.; Doré, A.G. Depressurization of hydrocarbon-bearing reservoirs in exhumed basin settings: evidence from Atlantic margin and borderland basins. In Exhumation of the North Atlantic Margin: Timing, Mechanisms and Implications for Hydrocarbon Exploration; Doré, A.G., Cartwright, J.A., Stoker, M.S., Turner, J.P., White, N., Eds.; Geological Society of London, Special Publications: London, UK, 2002; Volume 196, pp. 457–483. [Google Scholar]
- Terzaghi, K. Erdbaumechanik auf Bodenphysikalischer Grundlage; Franz Deuticke: Vienna, Austria, 1925. [Google Scholar]
- Terzaghi, K. Theoretical Soil Mechanics; Wiley: New York, NY, USA, 1943. [Google Scholar]
- Bredehoeft, J.D.; Wesley, J.B.; Fouch, T.D. Simulations of the origin of fluid pressure, fracture generation, and the movement of fluids in the Uinta Basin, Utah. AAPG Bull. 1994, 78, 1729–1747. [Google Scholar]
- Biot, M.A.; Willis, D.G. The elastic coefficients of the theory of consolidation. J. Appl. Mech. 1957, 24, 594–601. [Google Scholar] [CrossRef]
- Barnhoorn, A.; Cox, S.F.; Robinson, D.J.; Senden, T. Stress- and fluid-driven failure during fracture array growth: implications for coupled deformation and fluid flow in the crust. Geology 2010, 38, 779–782. [Google Scholar] [CrossRef]
- Crawford, B.R.; Webb, D.W.; Searles, K.H. Plastic compaction and anisotropic permeability development in unconsolidated sands with implications for horizontal well performance. In Proceedings of the 42nd US Rock Mechanics Symposium and 2nd U.S.-Canada Rock Mechanics Symposium, San Francisco, CA, USA, 29 June–2 July 2008. [Google Scholar]
- Ferrill, D.A.; Smart, K.J.; Cawood, A.J.; Morris, A.P. The fold-thrust belt stress cycle: superposition of normal, strike-slip, and thrust faulting deformation regimes. J. Struct. Geol. 2021, 148, 104362. [Google Scholar] [CrossRef]
- Becker, D.E. Testing in geotechnical design. Geotech. Eng. J. SEAGS AGSSEA 2010, 41, 10. [Google Scholar]
- Bär, K.; Arbarim, R.; Turan, A.; Schulz, K.; Mahmoodpour, S.; Leiss, B.; Wagner, B.; Sosa, G.; Ford, K.; Trullenque, G.; et al. Database of Petrophysical and Fluid Properties and Recommendations for Model Parametrization of the Four Variscan Reservoir Types; MEET Report, Deliverable D5.5; MEET Consortium European, June 2020; 110p, Available online: https://www.meet-h2020.com/project-results/deliverables/ (accessed on 23 September 2021).
- Tingay, M.R.P.; Hillis, R.R.; Morley, C.K.; Swarbrick, R.E.; Okpere, E.C. Variation in vertical stress in the Baram Basin, Brunei: tectonic and geomechanical implications. Mar. Pet. Geol. 2003, 20, 1201–1212. [Google Scholar] [CrossRef] [Green Version]
- Tan, C.P.; Willoughby, D.R.; Zhou, S.; Hillis, R.R. An analytical method for determining horizontal stress bounds from wellbore data. Int. J. Rock Mech. Min. Sci. Geomech. Abs. 1993, 30, 1103–1109. [Google Scholar] [CrossRef]
- Lund, B.; Townend, J. Calculating horizontal stress orientations with full or partial knowledge of the tectonic stress tensor. Geophys. J. Int. 2007, 170, 1328–1335. [Google Scholar] [CrossRef] [Green Version]
- Zhang, J.; Hüpers, A.; Kreiter, S.; Kopf, A.J. Pore pressure regime and fluid flow processes in the shallow Nankai Trough Subduction Zone based on experimental and modeling results from IODP Site C0023. J. Geophys. Res. 2021, 126, e20248. [Google Scholar] [CrossRef]
- Gao, B.; Flemings, P.B.; Nikolinakou, M.A.; Saffer, D.M.; Heidari, M. Mechanics of fold-and-thrust belts based on geomechanical modelling. J. Geophys. Res. 2018, 123, 4454–4474. [Google Scholar] [CrossRef]
- Heidbach, O.; Rajabi, M.; Reiter, K.; Ziegler, M.; WSM Team. World Stress Map Database Release 2016. V. 1.1. GFZ Data Serv. 2016. [Google Scholar] [CrossRef]
- Hobbs, B.E.; Means, W.D.; Williams, P.F. An Outline of Structural Geology; Wiley: New York, NY, USA, 1976; 671p. [Google Scholar]
- Peacock, D.C.P.; Tavarnelli, E.; Anderson, M.W. Interplay between stress permutations and overpressure to cause strike-slip faulting during tectonic inversion. Terra Nova 2017, 29, 61–70. [Google Scholar] [CrossRef] [Green Version]
- Laubach, S.E.; Olson, J.E.; Gross, M.R. Mechanical and fracture stratigraphy. AAPG Bull. 2009, 93, 1413–1426. [Google Scholar] [CrossRef]
- Japsen, P.; Green, P.F.; Nielsen, L.H.; Rasmussen, E.S.; Bidstrup, T. Mesozoic-Cenozoic exhumation events in the eastern North Sea Basin: a multi-disciplinary study based on palaeothermal, palaeoburial, stratigraphic and seismic data. Basin Res. 2007, 19, 451–490. [Google Scholar] [CrossRef]
- Nadan, B.J.; Engelder, T. Microcracks in New England granitoids: a record of thermoelastic relaxation during exhumation of intracontinental crust. Geol. Soc. Am. Bull. 2009, 121, 80–99. [Google Scholar] [CrossRef]
- Albero, F.; Tanner, D.C. Modellierung des Temperaturfeldes um eine tiefe Förderbohrung am östlichen Rand des Leinetalgrabens—Abschätzung des geothermischen Potentials. In Neue Untersuchungen zur Geologie der Leinetalgrabenstruktur; Leiss, B., Tanner, D.C., Vollbrecht, A., Arp, G., Eds.; Universitätsverlag Göttingen: Göttingen, Germany, 2011; pp. 115–123. [Google Scholar]
- Wilkins, S.J.; Gross, M.R.; Wacker, M.; Eyal, Y.; Engelder, T. Faulted joints: kinematics, displacement–length scaling relations and criteria for their identification. J. Struct. Geol. 2001, 23, 315–327. [Google Scholar] [CrossRef]
- Karnin, W.D.; Idiz, E.; Merkel, D.; Ruprecht, E. The Zechstein Stassfurt Carbonate hydrocarbon system of the Thuringian Basin, Germany. Pet. Geosci. 1996, 2, 53–58. [Google Scholar] [CrossRef]
- Bucher, K.; Stober, I.; Seelig, U. Water deep inside the mountains: unique water samples from the Gotthard rail base tunnel, Switzerland. Chem. Geol. 2012, 334, 240–253. [Google Scholar] [CrossRef]
- Van Der Heever, P. The Influence of Geological Structure on Seismicity and Rockbursts in the Klerksdorp Goldfield. Master’s Thesis, Rand Afrikaans University, Johannesburg, South Africa, 1982; 156p. [Google Scholar]
- Gumede, H.; Stacey, T.R. Measurement of typical joint characteristics in South African gold mines and the use of these characteristics in the prediction of rock falls. J. S. Afr. Inst. Min. Metall. 2007, 107, 335–344. [Google Scholar]
- Areshev, E.G.; Dong, T.L.; San, N.T.; Shnip, O.A. Reservoirs in fractured basement on the continental shelf of southern Vietnam. J. Pet. Geol. 1992, 15, 451–464. [Google Scholar] [CrossRef]
- Zhang, X.; Sanderson, D.J. Numerical study of critical behaviour of deformation and permeability of fractured rock masses. Mar. Pet. Geol. 1998, 15, 535–548. [Google Scholar] [CrossRef]
- Camac, B.A.; Hunt, S.P. Predicting the regional distribution of fracture networks using the distinct element numerical method. AAPG Bull. 2009, 93, 1571–1583. [Google Scholar] [CrossRef]
- Cundall, P.A.; Pierce, M.E.; Mas Ivars, D. Quantifying the size effect of rock mass strength. In Proceedings of the First Southern Hemisphere International Rock Mechanics Symposium, Perth, Australia, 16–19 September 2008; Potvin, Y., Carter, J., Dyskin., A., Jeffrey, R., Eds.; Australian Centre for Geomechanics: Perth, Australia, 2008; pp. 3–15. [Google Scholar]
- Donath, F.A. Experimental study of shear failure in anisotropic rocks. Geol. Soc. Am. Bull. 1961, 72, 985–989. [Google Scholar] [CrossRef]
- Yao, C.; Shao, J.F.; Jiang, Q.H.; Zhou, C. A new discrete method for modeling hydraulic fracturing in cohesive porous materials. J. Pet. Sci. Eng. 2019, 180, 257–267. [Google Scholar] [CrossRef]
Definition and significance | Bunter sandstone | Greywacke | Slate | Unit | |
---|---|---|---|---|---|
Density | The mass per unit volume of the rock and/or the fluids in the rock. Mean density controls the vertical (overburden) stress | 2.68 [40] | 2.42 to 2.74 [41] | 2.7 to 2.9 [42] | g/cm3 |
Tensile strength | The stress needed to cause extension fracturing. Controls where the failure envelope intersects the zero shear stress axis of the Mohr diagram, and the magnitudes of the effective tensile stresses needed to create extension fractures | 6 [43] | 20.3 to 35.7 [41] | 4.4 normal to cleavage,14.4 along cleavage [42] | MPa |
Uniaxial compressive strength | The strength of a rock derived from a uniaxial compression test (e.g., [44]) | 70 to 134 [45] | Average ≈ 200 (range 41–209) [46] | 2.33 to 151.6 [47] | MPa |
Cohesion | The shear strength of a material when the stress normal to a shear surface is zero (e.g., [48,49]). Controls where the failure envelope intersects the zero normal stress axis of the Mohr diagram, and the magnitudes of the effective differential stresses needed to create shear fractures | 12 [45] | 49 to 51 [41] | 64 normal to cleavage, 11 when σ1 30° to cleavage [42] | MPa |
Poisson ratio | The relationship between the tendency to shorten in one direction and the tendency to expand in another direction (e.g., [50]) | 0.16 to 0.35 [51] | 0.11 to 0.29 [41] | 0.22 to 0.29 [42] | |
Geostatic stress ratio * | The ratio of the horizontal effective stress (σ′H) to the vertical effective stress (σ′V) (e.g., [52]). It gives the effect the overburden has on horizontal stresses. Influences the diameter of the Mohr circle. Values calculated using Equation (2) | 0.19 to 0.54 | 0.125 to 0.41 [41] | 0.28 to 0.41 | |
Angle of internal friction (φ) | The angle of the fracture to σ1 is ± (45° − φ/2) (e.g., [48]) | 27.6 to 37.9 [51] | 43 to 44 [41] | 30 to 50 [53] | Degrees |
Coefficient of internal friction (μ) ** | Controls the slope of the failure envelope in the compressional field of the Mohr diagram (e.g., [48]). μ = tan φ | 0.52 | 0.97 | 0.84 | |
Young’s modulus | The stiffness of a solid material (e.g., [54]) | 22 to 37 [51] | 2.3 to 7 [41] | 12 to 56 [42] | GPa |
Coefficient of thermal expansion | The extent to which a material expands when heated or contracts when cooled. Can influence the development of tensile stresses during exhumation and cooling of rocks | 11.25 [55] | 17 [56] | 8 [57] | 10−6 K−1 |
Bunter | Greywacke | Slate | Unit | ||
---|---|---|---|---|---|
Input parameters | Depth | 1000 | 2000 4500 | 2000 4500 | m |
Average rock density | 2.41 | 2.68 | 2.68 | g/cm3 | |
Fluid density | 1 | 1 | 1 | g/cm3 | |
Overpressure | 0 | 0 | 0 | MPa | |
Poisson’s ratio (ν) | 0.16 | 0.11, 0.29 | 0.22, 0.29 | 0.16 | |
Applied tectonic stress | 0 | 0 | 0 | MPa | |
Cohesion | 12 | 40 | 30 | MPa | |
Coefficient of internal friction (μ) | 0.7 | 0.97 | 0.84 | ||
Friction angle | 33 | 44 | 40 | Degrees | |
Tensile strength (T) | 6 | 20 | 15 | MPa | |
Bed dip | 5 | SE limb = 45° NW limb = 70° | SE limb = 45° NW limb = 70° | Degrees | |
Vein dip | 85 | SE limb = 45° NW limb = 25° | S limb = 45° NW limb = 25° | Degrees | |
Cleavage | N/A | 65 | 65 | Degrees | |
Joint dip | 90 | 85 | 85 | Degrees | |
Derived parameters | Fluid pressure | 9.81 | 19.6 44.1 | 19.6 44.1 | MPa |
Geostatic pressure ratio (k0) | 0.19 | 0.125, 0.41 | 0.283, 0.41 | ||
Vertical effective stress ( σ′V) | 13.85 | 33 74.2 | 33.4 75.1 | MPa | |
Horizontal effective stress (σ′H =σ´V k0) | 2.63 | 4.1, 13.5 9.3, 30.4 | 9.4, 13.7 21.2, 30.8 | MPa |
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations. |
© 2021 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (https://creativecommons.org/licenses/by/4.0/).
Share and Cite
Peacock, D.C.P.; Sanderson, D.J.; Leiss, B. Use of Mohr Diagrams to Predict Fracturing in a Potential Geothermal Reservoir. Geosciences 2021, 11, 501. https://doi.org/10.3390/geosciences11120501
Peacock DCP, Sanderson DJ, Leiss B. Use of Mohr Diagrams to Predict Fracturing in a Potential Geothermal Reservoir. Geosciences. 2021; 11(12):501. https://doi.org/10.3390/geosciences11120501
Chicago/Turabian StylePeacock, D.C.P., David J. Sanderson, and Bernd Leiss. 2021. "Use of Mohr Diagrams to Predict Fracturing in a Potential Geothermal Reservoir" Geosciences 11, no. 12: 501. https://doi.org/10.3390/geosciences11120501
APA StylePeacock, D. C. P., Sanderson, D. J., & Leiss, B. (2021). Use of Mohr Diagrams to Predict Fracturing in a Potential Geothermal Reservoir. Geosciences, 11(12), 501. https://doi.org/10.3390/geosciences11120501