Hostname: page-component-76fb5796d-skm99 Total loading time: 0 Render date: 2024-04-29T10:54:38.463Z Has data issue: false hasContentIssue false

Fracture mechanics model of a fault termination zone

Published online by Cambridge University Press:  25 March 2011

A. RABINOVITCH*
Affiliation:
Department of Physics, The Ben Gurion University of the Negev, Beer Sheva, Israel
M. FRIEDMAN
Affiliation:
Department of Physics, Nuclear Research Center Negev, Beer Sheva, P. O. Box 9001, Israel
D. BAHAT
Affiliation:
Department of Geology and Mineralogy, The Ben Gurion University of the Negev, Beer Sheva, Israel
*
Author for correspondence: avinoam@bgu.ac.il

Abstract

A fault termination zone from Middle Eocene chalks near Beer Sheva, Israel, is analysed via a fracture mechanical technique. The zone consists of the end of the primary fault, three secondary faults and joints associated with these structures. We demonstrate that the shape of the first secondary fault can be obtained from theoretical fracture mechanical calculations. This shape also enables us to obtain the set of conditions which induced the observed structures. The technique reveals the relative importance of the different variables that appear in the theory. The most significant parameters in determining the shape of the first secondary fault are the vertical dimension of the primary fault, the ratio of the two horizontal differential stresses and the initiation angle of the secondary fault. Results indicate that the fault termination zone was created under an almost pure shear load. The true lateral displacement of the primary fault is unknown; hence, exact calculation of the stresses leading to the secondary fault is hardly possible. However, an estimation, based on a partial conversion of this displacement to an uplift, yields σ1 and σ3 values at initiation of between 3.8 and 7.6 MPa and 3.65 and 7.3 MPa, respectively.

Type
Original Articles
Copyright
Copyright © Cambridge University Press 2011

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Aki, K. 1989. Geometric features of a fault zone related to the nucleation and termination of an earthquake rupture. In Proceedings of Conference XLV. Fault Segmentation Controls of Rupture Initiation and Termination (eds Schwartz, D. P. & Sibson, R. H.), pp. 1–9. United States Geological Survey Open-File Report 89-315.Google Scholar
Alik, I. Z, Mueller, B. & Schubert, G. 2005. Three-dimensional numerical modeling of contemporary mantle flow and tectonic stress beneath the earthquake-prone southeastern Carpathians based on integrated analysis of seismic, heat flow, and gravity data. Physics of the Earth and Planetary Interiors 149, 8198.Google Scholar
Anderson, E. M. 1951. The Dynamics of Faulting, 2nd ed. Edinburgh: Oliver and Boyd.Google Scholar
Bahat, D. 1987. Jointing and fracture interactions in Middle Eocene chalks near Beer-Sheva, Israel. Tectonophysics 136, 299321.CrossRefGoogle Scholar
Bahat, D. 1991. Tectonofractography. Heidelberg: Springer-Verlag, 354 pp.CrossRefGoogle Scholar
Bahat, D. 1999. Single-layer burial joints vs. single-layer uplift joints in Eocene chalk from the Beer Sheva syncline in Israel. Journal of Structural Geology 21, 293303.CrossRefGoogle Scholar
Bahat, D., Rabinovitch, A. & Frid, V. 2001. Fracture characterization of chalk in uniaxial and triaxial tests by rock mechanics, fractographic and electromagnetic methods. Journal of Structural Geology 23, 1531–47.CrossRefGoogle Scholar
Bahat, D., Rabinovitch, A. & Frid, V. 2005. Tensile Fracturing in Rocks. Springer.Google Scholar
Bartlett, W. L, Friedman, M. & Logan, M. J. 1981. Experimental folding and faulting of rocks under confining pressure, Part IX. Wrench faults in limestone layers. Tectonophysics 79, 255–77.CrossRefGoogle Scholar
Bayasgalan, A., Jackson, J., Ritz, J. F. & Carretier, S. 1999. Field examples of strike-slip fault terminations in Mongolia and their tectonic significance. Tectonics 18 (3), 394411.CrossRefGoogle Scholar
Bombolakis, E. G. 1976. Some constraints and aids for interpretation of fracture and fault development. In Proceedings of the Second International Conference on Basement Tectonics (ed. Podwysocki, M. H. & Earle, J. L.), pp. 289305. International Basement Tectonics Association, Inc.Google Scholar
Brace, W. F. & Bombolakis, E. G. 1963. A note on brittle crack growth in compression. Journal of Geophysical Research 68, 3709–13.CrossRefGoogle Scholar
Buchbinder, B., Benjamini, C., Mimran, Y. & Gvirtzman, G. 1988. Mass transport in Eocene pelagic chalk on the northwestern edge of the Arabian platform, Shefela area, Israel. Sedimentology 35, 257–74.CrossRefGoogle Scholar
Chinnery, M. A. 1963. The stress changes that accompany strike-slip faulting. Bulletin of the Seismological Society of America 53, 921–32.CrossRefGoogle Scholar
Chinnery, M. A. 1966 a. Secondary faulting I. Theoretical aspects. Canadian Journal of Earth Sciences 3, 163–74.CrossRefGoogle Scholar
Chinnery, M. A. 1966 b. Secondary faulting II. Geological aspects. Canadian Journal of Earth Sciences 3, 175–90.CrossRefGoogle Scholar
Chinnery, M. A. & Petrak, J. A. 1968. Dislocation fault model with a variable discontinuity. Tectonophysics 5, 513–29.CrossRefGoogle Scholar
Couzens, B. A. & Dunne, W. M. 1994. Displacement transfer at thrust terminations – the Saltville thrust and sinking creek anticline, Virginia, USA. Journal of Structural Geology 16 (6), 781–93.CrossRefGoogle Scholar
Cruikshank, K. M., Shao, G. & Johnson, A. M. 1991. Analysis of minor fractures associated with joints and faulted joints. Journal of Structural Geology 13, 865–86.CrossRefGoogle Scholar
Das, S. & Scholtz, C. 1982. Off-fault aftershock clusters caused by shear stress increase. Bulletin of the Seismological Society of America 71, 1669–75.CrossRefGoogle Scholar
Dhont, D., Chorowicz, J., Yurur, T., Froger, J. L., Kose, O. & Gundogdu, N. 1998. Emplacement of volcanic vents and geodynamics of central Anatolia, Turkey. Journal of Volcanology and Geothermal Research 85 (1–4), 3354.CrossRefGoogle Scholar
Du, Y. & Aydin, A. 1995. Shear fracture patterns and connectivity at geometric complexities along strike-slip faults. Journal of Geophysical Research 100, 18093–102.CrossRefGoogle Scholar
Kim, Y. S., Peacock, D. C. P. & Sanderson, D. J. 2003. Mesoscale strike-slip faults and damage zones at Marsalforn, Gozo Island, Malta. Journal of Structural Geology 25 (5), 793812.CrossRefGoogle Scholar
King, G. C. P. 1986. Speculation on the geometry of the initiation and termination process of earthquake rupture and its relation to morphology and geological structure. Pure and Applied Geophysics 124, 567–85.CrossRefGoogle Scholar
Klerks, J., Theunissen, K. & Delvaux, D. 1998. Persistent fault controlled basin information since the Proterozoic along the Western Branch of the East African Rift. Journal of African Earth Sciences 26 (3), 347–61.CrossRefGoogle Scholar
Knott, J. F. 1973. Fundamentals of Fracture Mechanics. New York: Wiley, 273 pp.Google Scholar
Koifman, L. & Flexer, A. 1975. Possibilities of Under-Ground Storage of Crude Oil in Israel. The Israel Institute of Petroleum and Energy, 217 pp.Google Scholar
Lawn, B. R. & Wilshaw, T. R. 1975. Fracture of Brittle Solids. London: Cambridge University Press, 204 pp.Google Scholar
Lensen, G. J. 1958. A method of graben and horst formation. Journal of Geology 66, 579–87.CrossRefGoogle Scholar
Little, T. A. & Roberts, A. P. 1997. Distribution and mechanism of Neogene to present-day vertical axis rotations, Pacific Australian plate boundary zone, South Island, New Zealand. Journal of Geophysical Research – Solid Earth 102 (B9), 20447–68.CrossRefGoogle Scholar
Muraoka, H. & Kamata, H. 1983. Displacement distribution along minor fault traces. Journal of Structural Geology 5, 483–95.CrossRefGoogle Scholar
Oskin, M. & Iriondo, A. 2004. Large-magnitude transient strain accumulation on the Blackwater fault, Eastern California shear zone. Geology 32 (4), 313–16.CrossRefGoogle Scholar
Petit, J-P. & Barquins, M. 1987. Formation de fissures en milier confine dans le PMMA (polymetachrylate de methyle): modele analogique de formation de structures tectoniques cassantes. Comptes Rendus de l'Académie des Sciences Paris 305, 5560.Google Scholar
Petit, J-P. & Barquins, M. 1988. Can natural faults propagate under mode II conditions? Tectonophysics 7, 1243–56.Google Scholar
Price, N. J. & Cosgrove, J. W. 1990. Analysis of Geological Structures. Cambridge: Cambridge University Press.Google Scholar
Rayleigh, B. & Marone, C. 1986. Dilatancy of quartz gouge in pure shear. In Mineral and Rock Deformation: Laboratories Studies (eds Hobbs, B. E. & Heard, H. C.), pp. 110. American Geophysical Union, Geophysical Monograph vol. 36, Washington, DC.CrossRefGoogle Scholar
Ringenbach, J. C., Pinet, N., Stephan, J. F. & Delteil, J. 1993. Structural variety and tectonic evolution of strike-slip basins related to the Philippine fault system, Northern Luzon, Philippines. Tectonics 12 (1), 187203.Google Scholar
Salvini, F., Billi, A. & Wise, D. U. 1999. Strike-slip fault-propagation cleavage in carbonate rocks: the Mattinata Fault Zone, Southern Apennines, Italy. Journal of Structural Geology 21 (12), 1731–49.CrossRefGoogle Scholar
Simancas, J. F., Azor, A., Martinez Poyatos, D., Gonzalez Lodeiro, F. & Caro, L. 1997. Accommodation of fault displacements by granite deformation: an example from the Southwestern Iberian Massif. Comptes Rendus De L'Academie des Sciences de la Terre et des Planetes 324 (3), 245–52.Google Scholar
Steketee, J. A. 1958. Some geophysical applications of the elasticity theory of dislocations. Canadian Journal of Physics 36, 1168–98.CrossRefGoogle Scholar
Trudgill, B. & Cartwright, J. 1994. Relay-ramp forms and normal-fault linkages, Canyonlands National Park, Utah.Bulletin of the Geological Society of America 106: 1143–57.2.3.CO;2>CrossRefGoogle Scholar