Skip to main content
Log in

Fault Wear by Damage Evolution During Steady-State Slip

  • Published:
Pure and Applied Geophysics Aims and scope Submit manuscript

Abstract

Slip along faults generates wear products such as gouge layers and cataclasite zones that range in thickness from sub-millimeter to tens of meters. The properties of these zones apparently control fault strength and slip stability. Here we present a new model of wear in a three-body configuration that utilizes the damage rheology approach and considers the process as a microfracturing or damage front propagating from the gouge zone into the solid rock. The derivations for steady-state conditions lead to a scaling relation for the damage front velocity considered as the wear-rate. The model predicts that the wear-rate is a function of the shear-stress and may vanish when the shear-stress drops below the microfracturing strength of the fault host rock. The simulated results successfully fit the measured friction and wear during shear experiments along faults made of carbonate and tonalite. The model is also valid for relatively large confining pressures, small damage-induced change of the bulk modulus and significant degradation of the shear modulus, which are assumed for seismogenic zones of earthquake faults. The presented formulation indicates that wear dynamics in brittle materials in general and in natural faults in particular can be understood by the concept of a “propagating damage front” and the evolution of a third-body layer.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7

Similar content being viewed by others

References

  • Allix, O. and Hild, F. (2002) Continuum damage mechanics of materials and structures, Elsevier, 396 pp.

  • Archard, J.F. (1953) Contact and rubbing of flat surfaces, J. Appl. Phys., 24, 981–988.

    Google Scholar 

  • Archard, J.F., Hirch, W. (1956) Wear of metals under unlubricated conditions. Proc. R. Soc. Lond. A Math. Phys. Sci. 236, 397–410.

  • Bazant, Z.P. (1991) Why continuum damage is nonlocal: Micromechanics arguments, J. Eng. Mech., 1070–1087.

  • Bazant, Z.P. (2005) Scaling of structural strength. Elsevier, 327 pp.

  • Bazant, Z.P. and Jirasek, M. (2002) Nonlocal integral formulations of plasticity and damage: Survey of progress, J. Eng. Mech., 128, 1119–1149.

    Google Scholar 

  • Ben-Zion, Y., Lyakhovsky, V., (2002), Accelerating seismic release and related aspects of seismicity patterns on earthquake faults. Pure Appl. Geophys. 159, 2385–2412.

    Google Scholar 

  • Bercovici, D. and Ricard, Y. (2003) Energetics of a two-phase model of lithospheric damage, shear localization and plate boundary formation, Geophys. J. Int., 152, 581–596.

    Google Scholar 

  • Bercovici, D., Ricard, Y., and Schubert, G. (2001) A two-phase model for compaction and damage, part 1: general theory, J. Geophys. Res., 106, 8887–8906.

    Google Scholar 

  • Boneh, Y, Chang, J.J., Lockner, D.A., and Reches, Z. (2014) Fault evolution by transient processes of wear and friction, Pure. Appl. Geoph. (this volume).

  • Boneh, Y. (2012) Wear and gouge along faults: Experimental and mechanical analysis, Thesis, Univ. Oklahoma.

  • Boneh, Y., Sagy, A., and Reches, Z. (2013) Frictional strength and wear-rate of carbonate faults during high-velocity, steady-state sliding, Earth Planet. Sci. Lett., 381, 127–137.

  • Brodsky E.E., Gilchrist, J.J Sagy, A., and Collettini, C. (2011) Faults smooth gradually as a function of slip, Earth Planet. Sci. Lett., 302, 185–193.

  • Byerelee, J.D. (1967) Frictional characteristics of granite under high confining pressure, J. Geophys. Res., 72, 3639-3648.

    Google Scholar 

  • Byerlee, J.D., (1978), Friction of rocks. Pure Appl. Geophys., 116, 615–626.

  • Chang, J.C., Lockner, D.A., Reches, Z. (2012) Rapid acceleration leads to rapid weakening in earthquake-like laboratory experiments, Science, 338, 101, doi:10.1126/science.1221195.

  • Chester, F.M., and Chester, J.S. (1998) Ultracataclasite structure and friction processes of the San Andreas fault, Tectonophysics, 295, 199–221.

  • Chester, F.M., Evans, J.P., Biegel, R.L. (1993) Internal structure and weakening mechanisms of the San Andreas fault, J. Geophys. Res., 98, 771–786.

    Google Scholar 

  • Chester, F.M., Logan, J.M. (1987) Composite planar fabric of gouge from the Punchbowl fault, California, J. Struct. Geol., 9, 621–634.

  • Di Toro, G., Han, R., Hirose, T., De Paola, N., Nielsen, S., Mizoguchi, K., Ferri, F., Cocco, M., Shimamoto, T. (2011) Fault lubrication during earthquakes. Nature, 471, 494–498.

  • Dieterich, J.H. and Kilgore, B.D., (1996), Imaging surface contacts; power law contact distributions and contact stresses in quarts, calcite, glass, and acrylic plastic, Tectonophysics, 256, 219–239.

  • Dieterich, J.H., (1972) Time-dependent friction in rocks. J. Geophys. Res., 77, 3690–3697.

    Google Scholar 

  • Dieterich, J.H., (1979) Modeling of rock friction 1. Experimental results and constitutive equations, J. Geophys. Res., 84, 2161–2168.

    Google Scholar 

  • Engelder, J. T. (1974), Cataclasis and the generation of fault gouge. Geol. Soc. Am. Bull., 85, 1515–1522.

  • Eringen, A.C. (1966) A unified theory of thermomechanical materials, Int. J. Energ. Sci., 4, 179–202.

  • Fillot, N., Iordanoff, I., and Berthier, Y. (2007) Wear modeling and the third body concept, Wear, 262, 949–957.

  • Fisher, R. (1937), The wave of advance of advantageous genes, Ann. Eugenics, 7, 355–369.

  • Godet, M. (1984) The third-body approach: A mechanical view of wear. Wear, 100, 437–452.

  • Grindrod, P. (1996) The theory and applications of reaction-diffusion equations: Patterns and waves, Oxford Applied Mathematics and Computing Science Series, The Clarendon Press, Oxford Univ. Press, New York, 2nd ed., 275 p.

  • Hamiel, Y., Liu, Y., Lyakhovsky, V., Ben-Zion. Y. and Lockner, D., (2004), A visco-elastic damage model with applications to stable and unstable fracturing. J. Geophys. Int. 159, 1155–1165.

    Google Scholar 

  • Hamiel, Y., Katz, O., Lyakhovsky, V. Reches, R. and Fialko, Y. (2006) Stable and unstable damage evolution in rocks with implications to fracturing of granite. Geophys. J. Int. 167, 1005–1016, doi:10.1111/j.1365-246X.2006.03126.x.

  • Han, R., Hirose, T., Shimamoto, T. (2010) Strong velocity weakening and powder lubrication of simulated carbonate faults at seismic slip rates. J. Geophys. Res., 115, B03412, doi:10.1029/2008JB006136.

  • Hansen, N.R. and Schreyer, H.L. (1994) A thermodynamically consistent framework for theories of elasticity coupled with damage, Int. J. Solids Struct., 31, 359–389.

    Google Scholar 

  • Heesakkers, V., Murphy, S., and Reches, Z. (2011a) Earthquake rupture at focal depth, Part I: Structure and rupture of the Pretorius fault, TauTona mine, South Africa, Pure Appl. Geophys., doi:10.1007/s00024-011-0354-7.

  • Heesakkers V., Murphy. S., Locker, D.A., and Reches, Z. (2011b) Earthquake rupture at focal depth, Part II: Mechanics of the 2004 M2.2 earthquake along the Pretorius fault, TauTona mine, South Africa, Pure Appl. Geophys., doi:10.1007/s00024-011-0355-6.

  • Hirose, T., Mizoguchi, K., Shimamoto, T. (2012) Wear processes in rocks at slow to high slip rates. J. Struc. Geol., doi:10.1016/j.jsg.2011.12.007.

  • Hoff, N.J. (1953) The necking and rupture of rods subjected to constant tensile loads, J. Apl. Mech., 20, 105–108.

    Google Scholar 

  • Johnson, P.A. and Jia, X., (2005), Non-linear dynamics, granular media and dynamic earthquake triggering, Nature, 437, 871–874.

  • Kachanov, L.M. (1958) On the time to rupture under creep conditions, Izv. Acad. Nauk SSSR, OTN 8, 26-31 (in Russian).

  • Kachanov, L.M. (1986), Introduction to Continuum Damage Mechanics, Martinus Nijhoff Publishers, 135 p.

  • Kachanov, M. (1994) On the concept of damage in creep and in the brittle-elastic range, Int. J. Damage Mech., 3, 329–337.

    Google Scholar 

  • Karrech, A. Regenauer-Lieb, K., and Poulet, T. (2011) Continuum damage mechanics for the lithosphere. J. Geophys. Res., 116, B04205.

    Google Scholar 

  • Kato, K., Adachi, K. (2000) Wear mechanisms, In: Bhushan, B. (ed.), Modern Tribology Handbook. CRC Press, Boca Raton, Florida, 273–300.

  • Katz, O., Reches, Z. E. and Baer, G. (2003) Faults and their associated host rock deformation: Part I. Structure of small faults in a quartz–syenite body, southern Israel, J. Struct. Geol., 25, 1675–1689.

    Google Scholar 

  • Kolmogorov, A.N., Petrovskii, I.G., and Piscounov, N.S. (1937) A study of the diffusion equation with increase in the amount of substance, and its application to a biological problem, Bull. Moscow Univ., Math. Mech. 1, 1–25. Translated by V.M. Volosov in V.M. Tikhomirov, editor, Selected Works of A. N.

  • Krajcinovic, D. (1996) Damage Mechanics, Amsterdam, Elsevier. 774 p.

  • Kroner, E., (1968) Elasticity theory of materials with long-range cohesive force. Int. J. Solids Struct., 3, 731–742.

    Google Scholar 

  • Lemaitre, J. (1996) A Course on Damage Mechanics, Springer-Verlag, Berlin.

  • Levy, A. V., and Jee, N. (1988) Unlubricated sliding wear of ceramic materials, Wear, 121, 363–380.

  • Lifshitz, E.M. and Pitaevskii, L.P. (1981) Physical Kinetics, Course of Theoretical Physics, L.D. Landau and E.M. Lifshitz, Vol. 10, Elsevier, 452 p.

  • Lockner, D.A., Morrow, C., Moore, D. and Hickman, S. (2011) Low strength of deep San Andreas fault gouge from SAFOD core, Nature, 472, p. doi:10.1038/nature09927.

  • Lockner, D.A., Recher, Z., Moore, D.E. (1992) Microcrack interaction leading to shear fracture. In Proc. 33rd U.S. Symposium on Rock Mechanics (ed. W. Wawersik), A.A. Balkema Rotterdam.

  • Lu, K., Brodsky, E.E., and Kavehpour, H.P. (2007) Shear-weakening of the transitional regime for granular flow, J. Fluid Mech., 587, 347–372.

    Google Scholar 

  • Lyakhovsky, V., Ben-Zion, Y. and Agnon, A., (2005) A viscoelastic damage rheology and rate- and state-dependent friction. Geophys. J. Int. 161, 179–190.

    Google Scholar 

  • Lyakhovsky, V., and Ben-Zion Y., (2014a), Damage-Breakage rheology model and solid-granular transition near brittle instability, J. Mech. Phys. Solids, 64, 184–197.

  • Lyakhovsky V. and Ben-Zion Y. (2014b) Continuum damage-breakage model for faulting accounting for solid-granular transition, Pure Appl. Geophys. (this volume).

  • Lyakhovsky V., Hamiel, Y. and Ben-Zion, Y. (2011) A non-local visco-elastic damage model and dynamic fracturing. J. Mech. Phys. Solids, 59, 1752–1776. doi:10.1016/j.jmps.2011.05.016.

    Google Scholar 

  • Lyakhovsky, V., Ben-Zion, Y., and Agnon, A. (1997) Distributed damage, faulting, and friction, J. Geophys. Res., 102, 27635–27649.

    Google Scholar 

  • Ma, S. K. (2000), Modern theory of critical phenomena, Westview press, 561 p.

  • Marone, C., (1998), Laboratory-derived friction laws and their application to seismic faulting, Annu. Rev. Earth Planet. Sci., 26, 643–649.

    Google Scholar 

  • Muhuri, S.K., Dewers, T.A., Scott, T.E., and Reches, Z. (2003) Interseismic fault strengthening and earthquake-slip stability: Friction or cohesion? Geology, 31(10), 881–884.

  • Murray, J. D. (2002) Mathematical Biology: I. An Introduction, 3rd ed., Springer, 551 p.

  • Newman W.I. and Phoenix, S.L. (2001) Time-dependent fiber bundles with local load sharing, Phys. Rev. E., 63(2), 021507, doi:10.1103/PhysRevE.63.021507.

  • Nielsen, S., Di Toro, G. and Griffith, W.A. (2010), Friction and roughness of a melting rock surface. Geophys. J. Int. 182, 299–310.

  • Nolen, J., Roquejoffre, J.M., Ryzhik, L. and Zlatoš, A. (2012) Existence and non-existence of Fisher-KPP transition fronts, Arch. Rational Mech. Anal., 203, 217–246, doi:10.1007/s00205-011-0449-4.

  • Onsager, L., (1931), Reciprocal relations in irreversible processes. Phys. Rev., 37, 405–416.

    Google Scholar 

  • Paterson, M.S. and Wong, T.-F. (2005) Experimental Rock Deformation—The Brittle Field. Berlin, Heidelberg, New York: Springer-Verlag, 348 pp.

  • Petit, J.P. (1987) Criteria for the sense of movement on fault surfaces in brittle rocks, J. Struct. Geol., 9, 597–608.

    Google Scholar 

  • Power, W.L., Tullis, T.E. and Weeks, J.D. (1988) Roughness and wear during brittle faulting, J. Geophys. Res. 93, 15268–15278.

    Google Scholar 

  • Queener, C.A., Smith, T.C., and Mitchell, W.L. (1965) Transient wear of machine parts, Wear, 8, 391–400.

  • Rabinowicz, E., Dunn, L.A., and Russell, P.G. (1961) A study of abrasive wear under three-body conditions, Wear, 4, 345–355.

  • Rabotnov, Y.N. (1959) A mechanism of a long time failure, in: Creep problems in structural members, 5–7, USSR Academy of Sci. Publ., Moscow.

  • Rabotnov, Y.N. (1988) Mechanics of Deformable Solids, Moscow, Science, 712 p.

  • Reches, Z. and Lockner, D.A., (1994) Nucleation and growth of faults in brittle rocks. J. Geophys. Res., 99, 18159–18173, doi:10.1029/94JB00115.

  • Reches, Z., and Lockner, D.A. (2010) Fault weakening and earthquake instability by powder lubrication, Nature, 467, 452–456, doi:10.1038/nature09348.

  • Ricard, Y. and Bercovici, D. (2009) A continuum theory of grain-size evolution and damage, J. Geophys. Res., 114. doi:10.1029/2007JB005491.

  • Robinson, E.L. (1952) Effect of temperature variation on the long-term rupture strength of steels, Trans. Am. Soc. Mech. Eng., 174, 777–781.

    Google Scholar 

  • Sagy, A. and Brodsky, E.E. (2009) Geometric and rheological asperities in an exposed fault zone, J. Geophys. Res. 114, B02301. doi:10.1029/2008JB005701.

  • Sammis, C., Lockner, D., Reches, Z., 2011. The role of adsorbed water on the friction of a layer of submicron particles. Pure Appl. Geophys. 168, 2325–2334.

  • Scholz, C.H. (1987) Wear and gouge formation in brittle faulting, Geology, 15, 493–495.

  • Scholz, C.H., (2002) The Mechanics of Earthquakes and Faulting, 2nd ed. Cambridge University Press, 471 p.

  • Shcherbakov, R. and Turcotte, D.L. (2003) Damage and self-similarity in fracture, Theor. Appl. Fracture Mech., 39, 245–258.

    Google Scholar 

  • Shipton, Z.K., Evans, J.P., Abercrombie, R.E., and Brodsky, E.E. (2006) The missing sinks: slip localization in faults, damage zones, and the seismic energy budget, In: Abercrombie, R. (eds.) Earthquakes: Radiated Energy and the Physics of Faulting, 217–222. Washington, DC.

  • Sibson, R.H. (1977) Fault rocks and fault mechanisms. J. Geol. Soc. 133, 191–213.

    Google Scholar 

  • Turcotte, D.L., Newman, W.I., and Shcherbakov, R. (2003) Micro and macroscopic models of rock fracture, Geophys. J. Int., 152, 718–728.

  • Wang, W. and Scholz, C.H. (1994) Wear processes during frictional sliding of rock: a theoretical and experimental study. J. Geophys. Res., 99, 6789–6799.

    Google Scholar 

  • Wibberley, C.A., Yielding, G., and Di Toro, G. (2008) Recent advances in the understanding of fault zone internal structure: A review. Geol. Soc., London, Special Publ., 299, 5–33.

  • Wilson, B.T., Dewers, T., Reches, T. and Brune, J.N. (2005) Particle size and energetics of gouge from earthquake rupture zones. Nature, 434, 749–752.

  • Zang, A., Wagner, F., Stanchits, S., Janssen, C. and Dresen, G., 2000. Fracture process zone in granite, J. Geophys. Res., 105, 23651–23661.

    Google Scholar 

Download references

Acknowledgments

The manuscript benefitted from useful comments by Y. Ben-Zion (editor), W. Ashley Griffith (reviewer), and an anonymous reviewer. The study was supported by the NSF, Geosciences, Equipment and Facilities, Grant No. 0732715, with partial support of NSF, Geosciences, Geophysics, Grant No. 1045414, and ConocoPhillips Foundation grant. VL acknowledges support by the US–Israel Binational Science Foundation (Grant 2008248); AS acknowledges support by the Israel Science Foundation (Grant 929/10).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Vladimir Lyakhovsky.

Appendix

Appendix

1.1 Damage Rheology Model

Key derivations of the damage rheology model leading to the estimate of the speed of the propagating damage front are presented in this “Appendix”. Following thermodynamic balance relations and the Onsager (1931) principle, Lyakhovsky et al. (1997) developed a kinetic equation for the damage state variable, α (weakening and healing) which is a function of the progressive deformation. Non-linear elasticity that connects the effective elastic moduli to a damage variable and loading conditions allows accounting for the transition from damage accumulation to healing. This transition is controlled by the strain invariants ratio \( \xi = {{I_{1} } \mathord{\left/ {\vphantom {{I_{1} } {\sqrt {I_{2} } }}} \right. \kern-0pt} {\sqrt {I_{2} } }} \), where I 1 = ɛ kk and I 2 = ɛ ij ɛ ij are the invariants of the elastic strain tensor ɛ ij . The ξ value is a conjugate quantity to the ratio between shear and normal stress expressed in terms of strains, instead of stresses. The rate of damage/healing accumulation is given by Lyakhovsky et al. (1997):

$$ \frac{{{\text{d}}\alpha }}{{{\text{d}}t}} = \left\{ {\begin{array}{*{20}c} {C_{\text{d}} I_{2} \left( {\xi - \xi_{0} } \right)} & {{\text{for }}\xi > \xi_{0} } & {} \\ {C_{1} \exp \left( {\frac{\alpha }{{C_{2} }}} \right)I_{2} \left( {\xi - \xi_{0} } \right)} & {{\text{for }}\xi < \xi_{0} } & {} \\ \end{array} } \right.\!\!\!\!. $$
(4)

where the coefficient C d is the rate of positive damage evolution (material degradation) that is constrained by laboratory experiments (Lyakhovsky et al. 1997; Hamiel et al. 2004, 2006). The value ξ = ξ 0 controls the onset of damage accumulation or transition from material healing to weakening associated with microcrack nucleation and growth. The value of the critical strain invariants ratio also called “modified internal friction” (see Fig. 3 in Lyakhovsky et al. 1997) is explicitly related to the internal friction angle of Byerlee’s law (Byerlee 1978) and Poisson ratio of the intact rock.

The rate of damage recovery (healing) is assumed to depend exponentially on α and produces logarithmic healing with time in agreement with the behavior observed in laboratory experiments with rocks and other materials (e.g., Dieterich and Kilgore 1996; Scholz 2002; Muhuri et al. 2003; Johnson and Jia 2005). Lyakhovsky et al. (2005) showed that the local damage model reproduces the main phenomenological features of the rate- and state-dependent friction, and constrained the healing parameters C 1, C 2 by comparing the model calculations with empiric parameters of the slow-rate frictional sliding (e.g., Dieterich 1972, 1979; Marone 1998).

The damage accumulation under constant stress in a simplified 1-D model with effective elastic moduli degrading proportionally to (1 − α) follows a power law solution (e.g., Ben-Zion and Lyakhovsky 2002; Turcotte et al. 2003):

$$ \alpha (t) = 1 - \left( {1 - 3\frac{{C_{\text{d}} \sigma^{2} }}{{G_{0}^{2} }}t} \right)^{{{\raise0.7ex\hbox{$1$} \!\mathord{\left/ {\vphantom {1 3}}\right.\kern-0pt} \!\lower0.7ex\hbox{$3$}}}} . $$
(5)

where G 0 is the elastic moduli of the intact rock, σ, the applied stress, and C d the damage rate coefficient. This solution allows us to introduce a time scale, t f, which is the time-to-failure when α = 1 (total damage) (e.g., Paterson and Wong 2005) that becomes

$$ t_{f} = \frac{{G_{0}^{2} }}{{3C_{d} \sigma^{2} }}. $$
(6)

This parameter controls the time scale of all processes associated with accumulated damage, including the growth rate of narrow fracture zones.

Recently, Lyakhovsky et al. (2011) developed a gradient-type damage rheology formulation that incorporates non-local behavior by enriching the local constitutive relations with a gradient of the damage state variable. This addition modifies the kinetic equation for the damage evolution. In addition to the source term controlling the damage growth in Eq. (4), the damage accumulation for ξ > ξ 0 in non-local formulation includes damage diffusion term with a coefficient D:

$$ \frac{\partial \alpha }{\partial t} = C_{\text{d}} I_{2} \left( {\xi - \xi_{0} } \right) + D\nabla^{2} \alpha . $$
(7)

The non-local gradient-type damage kinetic Eq. (7) has the form of a Fisher-KPP reaction–diffusion type equation for which the general one-dimensional form is

$$ \frac{\partial u}{\partial t} = \frac{{\partial^{2} u}}{{\partial x^{2} }} + f(u) $$
(8)

The solution of the Fisher-KPP Eq. (8) exhibits a traveling wave or fronts switching between equilibrium states f(u) = 0 (see text). Travelling fronts propagating with the speed c exist in homogeneous media when \( c \ge c_{*} \equiv 2\sqrt {f^{\prime}(0)} \). We calculate the source term f(α) = C d I 2(ξ − ξ 0) for α = 0 under loading condition that mimics the experimental set-up with the normal stress σ n and shear stress τ, and then estimate the speed of the propagating damage front.

Adopting damage model with (1 − α) reduction of the elastic moduli, the relations between stress and strain components (axial, ε a, and transversal, ε τ ) are (E 0, G 0, Young and shear moduli of the intact rock):

$$ \varepsilon_{\text{a}} = \frac{{\sigma_{\text{n}} }}{{E_{0} \left( {1 - \alpha } \right)}};\,\varepsilon_{\tau } = \frac{\tau }{{2G_{0} \left( {1 - \alpha } \right)}} $$
(9)

Using relations (9), the second strain invariant, I 2, and strain invariant ratio, ξ, are,

$$ I_{2} = \varepsilon_{\text{a}}^{2} + 2\varepsilon_{\tau }^{2} = \frac{1}{{\left( {1 - \alpha } \right)^{2} }}\left[ {\frac{1}{{E_{0}^{2} }}\sigma_{\text{n}}^{2} + \frac{1}{{2G_{0}^{2} }}\tau^{2} } \right] $$
(10)
$$ \xi = \frac{{\varepsilon_{\text{a}} }}{{\sqrt {\varepsilon_{\text{a}}^{2} + 2\varepsilon_{\tau }^{2} } }} = \frac{ - 1}{{\sqrt {1 + 2\left( {\frac{{E_{0} \tau }}{{2G_{0} \sigma_{n} }}} \right)^{2} } }} $$
(11)

The minus sign in (11) implies that the compaction strains are negative. Note that without any shear loading (τ = 0), the strain invariants ratio is ξ = −1, which is slightly below typical ξ 0 range (Lyakhovsky et al. 1997). This implies that damage is not accumulated at loading by normal stress alone without shear loading. Substituting (10, 11) into equation for the damage accumulation (7), leads to the following relation for the source term:

$$ f(\alpha ) = C_{\text{d}} \frac{1}{{\left( {1 - \alpha } \right)^{2} }}\left[ {\frac{1}{{E_{0}^{2} }}\sigma_{\text{n}}^{2} + \frac{1}{{2G_{0}^{2} }}\tau^{2} } \right]\left[ {\frac{ - 1}{{\sqrt {1 + 2\left( {\frac{{E_{0} \tau }}{{2G_{0} \sigma_{n} }}} \right)^{2} } }} - \xi_{0} } \right] $$
(12)

The speed of the travelling damage front, which in the present work is defined as the wear-rate (text), is controlled by the value \( c_{*} = 2\sqrt {D \cdot f^{\prime}(0)} \) equal to

$$ c_{*} = 2\sqrt {2C_{\text{d}} D} \frac{{\sigma_{\text{n}} }}{{E_{0} }}\sqrt { - \xi_{0} \left[ {1 + 2\left( {\frac{{E_{0} \tau }}{{2G_{0} \sigma_{n} }}} \right)^{2} } \right] - \sqrt {1 + 2\left( {\frac{{E_{0} \tau }}{{2G_{0} \sigma_{n} }}} \right)^{2} } } $$
(13)

Taking μ S = τ/σ n for the steady-state friction coefficient, Eq. (13) is rearranged to:

$$ c_{*} = 2\sqrt {2C_{\text{d}} D} \frac{{\sigma_{\text{n}} }}{{E_{0} }}\sqrt { - \xi_{0} \left( {1 + A\mu_{\text{S}}^{2} } \right) - \sqrt {1 + A\mu_{\text{S}}^{2} } } $$
(14)

where A = 2(E 0/2G 0)2 = 2(1 + ν)2, and ν is the Poisson ratio. Note that typical values of the strain invariant ratio, ξ 0, controlling the onset of damage accumulation in (4) vary between ξ 0 = −1.2 and ξ 0 = −0.6. These values are taken from Fig. 3 of Lyakhovsky et al. (1997) for friction angle 30°, ν = 0.2 and for friction angle 40°, ν = 0.3. The steady-state friction coefficient should be above certain critical or threshold values (μ S ≥ μ cr) related to the material strength to give positive expression under the radical in (14). A negative value for μ S < μ cr implies that the applied shear stress is below the level needed for the onset of damage accumulation and the source term in damage growth Eq. (7) is negative; no any damage is accumulated, and no wear is anticipated.

The Poisson ratio for most rocks varies within the small range ν = 0.2–0.3. Variation of the steady-state friction, μ S, is also limited. It decreases from static friction values about 0.6–0.8 to dynamic values about 0.2–0.4. Accounting for this limited range of the material properties, the speed of the travelling damage front or wear-rate calculated using (14) only slightly depends on the Poisson ratio (black lines in Fig. 7). Hence, the wear-rate may be approximated by much simpler equation (red line in Fig. 7):

$$ {\text{wear\_rate}} \propto \sigma_{\text{n}} \sqrt {\mu_{\text{S}}^{2} - \mu_{\text{cr}}^{2} } $$
(15)

The units of the speed of the travelling damage front (13, 14) are length per time, while in the laboratory experiments wear-rate measured the thinning of the sample per unit slip. As the area of the surface is constant (Boneh et al. 2013), this thinning is proportional to the wear volume generation. Such unit conversion could be done under steady-state conditions with constant slip rate.

We note here that Eq. (14) was derived for the loading conditions of experimental set-up. More appropriate conditions for deep seismogenic zones should account for relatively large confining pressure, small damage-induced change of the bulk modulus and significant degradation of the shear modulus. Similar derivations lead directly to the more simple form (15) instead of (14) for the natural conditions.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Lyakhovsky, V., Sagy, A., Boneh, Y. et al. Fault Wear by Damage Evolution During Steady-State Slip. Pure Appl. Geophys. 171, 3143–3157 (2014). https://doi.org/10.1007/s00024-014-0787-x

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00024-014-0787-x

Keywords

Navigation