Thermally induced brittle deformation in oceanic lithosphere and the spacing of fracture zones

https://doi.org/10.1016/j.epsl.2008.02.025Get rights and content

Abstract

Brittle deformation of oceanic lithosphere due to thermal stress is explored with a numerical model, with an emphasis on the spacing of fracture zones. Brittle deformation is represented by localized plastic strain within a material having an elasto-visco-plastic rheology with strain softening. We show that crustal thickness, creep strength, and the rule governing plastic flow control the formation of cracks. The spacing of primary crack decreases with crustal thickness as long as it is smaller than a threshold value. Creep strength shifts the threshold such that crust with strong creep strength develops primary cracks regardless of crustal thicknesses, while only a thin crust can have primary cracks if its creep strength is low. For a thin crust, the spacing of primary cracks is inversely proportional to the creep strength, suggesting that creep strength might independently contribute to the degree of brittle deformation. Through finite versus zero dilatation in plastic strain, associated and non-associated flow rule results in nearly vertical and V-shaped cracks, respectively. Changes in the tectonic environment of a ridge system can be reflected in variation in crustal thickness, and thus related to brittle deformation. The fracture zone-free Reykjanes ridge is known to have a uniformly thick crust. The Australian-Antarctic Discordance has multiple fracture zones and thin crust. These syntheses are consistent with enhanced brittle deformation of oceanic lithosphere when the crust is thin and vice versa.

Introduction

The length of the mid-ocean ridge segments varies substantially among spreading centers and is correlated with several tectonic factors, including a positive correlation with spreading rate. The first order segments, bounded by transform faults (first order discontinuities), have an average length of 600 ± 300 km along fast (> 6 cm/yr) spreading ridges and 400 ± 200 km for slow (< 6 cm/yr) spreading ridges (MacDonald et al., 1991). Sandwell (1986) suggested that the length of the first order segments varies linearly with spreading rates. Although valid at the first order scale, such linear correlations are not supported at every level of the hierarchy. For example, the magmatic segments of slow to intermediate spreading ridges were shown to be 52.5 km long on average, independent of spreading rates (Briais and Rabinowicz, 2002).

Segment lengths are also apparently associated with regional variations in crustal thickness, creep strength, and mantle temperature. For instance, the Reykjanes ridge above the Iceland hot spot is known to have a uniform and much thicker crust for its spreading rate (Bunch and Kennett, 1980, Murton and Parson, 1993, Smallwood and White, 1998). This hot spot-affected ridge has been shown to exhibit signatures of wet mantle source for basaltic melt (Nichols et al., 2002). The water in crustal and mantle minerals has a strong weakening effect on creep strength although the preferential partitioning of water into melt phases complicates this straightforward relation (e.g., Karato, 1986, Hirth and Kohlstedt, 1996). In contrast, the Australian-Antarctic Discordance (AAD) has a highly rugged seafloor indicating increased fracturing (Hayes and Conolly, 1972, Weissel and Hayes, 1974) as well as anomalously thinner crust in comparison with other parts of the Southeast Indian Ridge (SEIR) (Tolstoy et al., 1995, Okino et al., 2004). Such regional features were attributed to colder mantle beneath the AAD (Weissel and Hayes, 1974), a hypothesis that was later supported by the systematics of major elements of basalt along the SEIR (Klein et al., 1991). These two regions, the Reykjanes ridge and the AAD, respectively exhibit reduced and enhanced segmentations at both the 1st and 2nd order compared with the other parts of the respective ridge systems. The degree of fracturing in those regions is substantially different in the profiles of free-air gravity anomaly (Fig. 1). The profile for the Reykjanes ridge (A–A') is smooth over the segment closer to Iceland and becomes rugged towards the southern end. The profile B–B' along the SEIR shows strong high frequency changes in depth and free-air gravity associated with the fracture zones over the AAD and with abrupt transition to a smooth segment east of the AAD.

A magma supply model has been proposed to explain the fundamentally different characteristics between slow- and fast-spreading centers, as well as axial morphology of a single ridge segment. According to this model, the variable amount of available magma at spreading centers and its along-ridge transport are responsible for along-ridge variations in axial bathymetry and associated geophysical and geochemical observations (MacDonald et al., 1991, MacDonald, 1998). Relating mantle dynamics to the conceptual magma supply model, calculations of mantle flow beneath slow spreading centers exhibit 3-D patterns that are segmented along the axis (Parmentier and Phipps Morgan, 1990, Lin and Phipps Morgan, 1992, Barnouin-Jha et al., 1997, Madge and Sparks, 1997). A problem with such flow models is that the wavelengths of the segmented mantle upwelling are larger (150 km) than the observed average second order segment length (~ 50 km) (Barnouin-Jha et al., 1997). However, when the effect of melt extraction on the viscosity of the magma residual was taken into account, a much shorter wavelength of segmented flow (as short as 70 km) was achieved (Choblet and Parmentier, 2001). Related to the magma supply model, ridge migration with respect to a hot spot reference frame was suggested to cause asymmetric mantle upwelling and melt production (Carbotte et al., 2004). This model provides an explanation for the observation that the majority of “leading” segments (that is, those that step in the same direction as ridge migration direction) are magmatically more robust.

Although the magma supply model is consistent with a range of observations, the model has yet to be linked with the brittle manifestation of mid-ocean ridge segmentation. Thermal stress due to the cooling of oceanic lithosphere is one possible driving force responsible for brittle ridge segmentation among many others (cf. Kastens, 1987). Using an order of magnitude argument, Collette (1974) suggested that thermal stress associated with the cooling of oceanic lithosphere should exceed its strength. By computing the bending moment of a semi-infinite thin elastic plate experiencing top-down cooling, it was suggested that segment length should be determined such that a plate can release thermal stress by bending (Turcotte, 1974). Expanding on this theory, Sandwell (1986) showed that ridge-bounding first order discontinuities can release thermal stress effectively when their spacing is proportional to spreading rate. Decomposing thermal stress into contraction and bending components, Haxby and Parmentier (1988) speculated that thermal bending stress, not contraction, would govern the spacing of transform faults because the magnitude of thermal contraction stress was independent of the ridge segment length. These studies, however, provide only an upper bound or indirect estimate of the fracture zone spacing. Sandwell and Fialko (2004) focused on the optimal spacing between thermal cracks, which minimizes stored elastic energy in a bending plate. It is notable that the spacing is not given a priori but is determined by the principle of minimum elastic energy.

A theory of thermal cracks provides useful insight into the spacing of ridge discontinuities if we assume that ridge segmentation occurs due to thermal stress. The stress distribution as a function of distance from a two-dimensional crack has been analyzed by Lachenbruch (1962). In a thermally contracting elastic half space, stresses are assumed to be released on the wall of a vertical crack. At greater distances from the wall, stress will increase to an ambient level so that each crack has a finite zone of stress relief. Since the strength of the material is limited, another crack will form at a distance where the stress exceeds the strength. In this fashion, the spacing between cracks is related to material strength and the size of the stress relief zones, which is determined by the material's elastic properties. Although this model could successfully explain crack spacing in permafrost, it leads to an apparent paradox for mid-ocean ridge segmentation. The model predicts a shorter spacing of cracks when the ambient level of stress is higher or the depth extent of cracks is shallower (Lachenbruch, 1962). Consequently, for fast and hot spreading centers, the Lachenbruch model implies a smaller fracture zone spacing because the amount of thermal stress is larger and brittle layers thinner compared to slow cooler spreading centers. In fact, the opposite trend is observed.

In this study, we investigate the role of thermal stress on the formation of mid-ocean ridge segmentation in terms of brittle deformation in young oceanic lithosphere. We address the influence of crustal thickness and rheology, factors that reflect a ridge system's tectonic setting, on fracture zone formation. A numerical method is used in which brittle deformation is allowed within the framework of continuum mechanics. This is an exploratory attempt towards a better understanding of ridge segmentation processes: relatively simple numerical models are used to draw implications relevant to actual ridge systems.

Section snippets

Numerical method

We use SNAC, an explicit finite difference code, to solve for the equations of momentum and heat energy conservation (Choi et al., in preparation). Although the code is fully three-dimensional, because of computational requirements, we only solve for 2-D problems here. The conservation equations are solved by the energy-based finite difference method (Bathe, 1996), which makes SNAC equivalent to a finite element code with linear tetrahedral elements except for the lack of explicit references to

Model setup

A series of 2-D models are constructed to represent a vertical cross-section along a straight ridge segment (Fig. 2). A 500 km long and 50 km deep domain is discretized into 1 × 1 km quadrilateral elements. To avoid complexities involving ridge axial processes, the domain is assumed to be initially at a small distance from the spreading center such that the initial temperature field is 0.3 My old lithosphere given by a half-space cooling model. Temperature is fixed at 0 °C on the top surface and

Models with weak crust

The temporal evolution of topography, temperature, viscosity, and plastic strain for the model with weak and normal thickness (7 km) crust (model 1 in Table 1) is shown in Fig. 3. The viscosity field is almost entirely determined by temperature while the dependence on stress is minimal. Lower viscosities are consistently found in the lower crust. Although brittle deformation occurred in the high viscosity uppermost part of the crust and mantle, the amount of plastic strain was only about 1%.

Effects of crustal thickness and creep strength

The models show that creep strength and crustal thickness strongly influence brittle deformation by the release of thermal stresses. Through the centers of surface troughs (grabens) that are connected to the primary cracks at depth, we measure the average spacing between the primary cracks (Table 3; Fig. 8).

Crustal thickness determines whether primary cracks are created. If the crust is thicker than a global mean (6–7 km, Chen, 1992, White et al., 1992) and thus has a weak lower part, then the

Conclusions

We show that crustal thickness, crustal creep strength, and the rule for plastic flow can substantially influence the brittle deformation of oceanic lithosphere. Crustal thickness determines whether brittle deformation would evolve into primary cracks or stay at secondary cracks without associated topographic features. Primary cracks only emerge when a crust is thinner than a certain threshold. Lower crustal creep strength has a net effect of shifting this threshold: When the creep strength is

Acknowledgements

We thank Joann Stock, Laetitia Le Pourhiet, and Paul Asimow for fruitful discussions and Dietmar Müller for suggestions on the manuscript. We would like to thank Claude Jaupart, Louis Geli, and two other anonymous reviewers for their useful and constructive reviews. This is contribution number 9174 of the Division of Geological and Planetary Sciences and 72 of the Tectonics Observatory. Development of SNAC was partially supported by the NSF ITR program under EAR-0205653. All calculations

References (67)

  • Barnouin-JhaK. et al.

    Buoyant mantle upwelling and crustal production at oceanic spreading centers: on-axis segmentation and off-axis melting

    J. Geophys. Res.

    (1997)
  • BatheK.-J.

    Finite Element Procedure

    (1996)
  • BellR. et al.

    Crustal control of ridge segmentation inferred from observations of the Reykjanes ridge

    Nature

    (1992)
  • BirdP.

    An updated digital model of plate boundaries

    Geochem. Geophys. Geosyst.

    (2003)
  • BoleyB.A. et al.

    Theory of Thermal Stresses

    (1960)
  • BriaisA. et al.

    Temporal variations of the segmentation of slow to intermediate spreading mid-ocean ridges 1. synoptic observations based on satellite altimetry data

    J. Geophys. Res.

    (2002)
  • BunchA.W.H. et al.

    The crustal structure of the Reykjanes ridge at 59° 30′N

    Geophys. J. R. Astron. Soc.

    (1980)
  • CarbotteS.M. et al.

    The influence of ridge migration on the magmatic segmentation of mid-ocean ridges

    Nature

    (2004)
  • ChenY.

    Oceanic crustal thickness versus spreading rate

    Geophys. Res. Lett.

    (1992)
  • ChenY. et al.

    Rift valley/no rift valley transition at mid-ocean ridges

    J. Geophys. Res.

    (1990)
  • ChenY. et al.

    A nonlinear rheology model for mid-ocean ridge axis topography

    J. Geophys. Res.

    (1990)
  • Choi, E., Thoutireddy, P., Lavier, L., Quenette, S., Tan, E., Gurnis, M., Aivazis, M., Appelbe, B., in preparation....
  • ChopraP.N. et al.

    The role of water in the deformation of dunite

    J. Geophys. Res.

    (1984)
  • ColletteB.J.

    Thermal contraction joints in a spreading seafloor asorigin of fracture zones

    Nature

    (1974)
  • CundallP.

    Numerical experiments on localization in frictional materials

    Ingenieur Arch.

    (1989)
  • de Sousza NetoE.A. et al.

    F-bar-based linear triangles and tetrahedral for finite strain analysis of nearly incompressible solids. Part I: formulation and benchmarking

    Int. J. Numer. Methods Eng.

    (2005)
  • DixonJ.E. et al.

    An experimental study of water and carbon dioxide solubilities in mid-ocean ridge basaltic liquids. Part I: Calibration and solubility models

    J. Petrology

    (1995)
  • HaxbyW.F. et al.

    Thermal contraction and the state of stress in the oceanic lithosphere

    J. Geophys. Res.

    (1988)
  • HayesD.E. et al.

    Morphology of the southeast Indian Ocean, in Antarctic Oceanography II: The Australian–New Zealand Sector

  • HirthG.

    Laboratory constraints on the rheology of the upper mantle

  • HirthG. et al.

    The rheology of the lower oceanic crust: implications for lithospheric deformation at mid-ocean ridges

  • JaegerJ. et al.

    Fundamentals of rock mechanics

    (1976)
  • KaratoS.

    Does partial melting reduce the creep strength of the upper mantle?

    Nature

    (1986)
  • Cited by (18)

    • Oblique continental rifting and long transform fault formation based on 3D thermomechanical numerical modeling

      2018, Tectonophysics
      Citation Excerpt :

      One possible reason that the long transform faults have not been reproduced in numerical models is that (due to computational limitations) they employed relatively small ( <200 km) computational domains (e.g., Allken et al., 2012; Gerya, 2013a,b, 2010; Liao and Gerya, 2015), which naturally limited the maximal size of the forming structures. Besides, most of the existing models investigating transforms nucleation commonly explored two perturbations (either thermal or compositional) with a certain relatively small offset ( <200 km, Allken et al., 2012; Choi and Gurnis, 2008; Gerya, 2013a,b; Püthe and Gerya, 2014), which to some extent pre-defined the length of the resulting transform structures. In this study, we aim at producing long transform faults from oblique continental rifting using 3D numerical thermomechanical modeling in a large computational domain.

    • Making Coulomb angle-oriented shear bands in numerical tectonic models

      2015, Tectonophysics
      Citation Excerpt :

      Finally, our approach is discussed in the light of the theory of strain localization and compared with the behaviors of natural faults. Numerical tectonic models often describe the plastic behavior of rocks with the Mohr–Coulomb (MC) or the Drucker–Prager (DP) model (e.g. Braun et al., 2008; Choi and Gurnis, 2008; Gerya and Yuen, 2007; Moresi et al., 2007; Popov and Sobolev, 2008). In spite of differences in details, time-independent stress projection onto a yield surface and isotropic hardening/softening are commonly associated with these plasticity models.

    • Accelerating DynEarthSol3D on tightly coupled CPU-GPU heterogeneous processors

      2015, Computers and Geosciences
      Citation Excerpt :

      The combination of an explicit finite element method, the Lagrangian description of motion, and the elasto-visco-plastic material model has been implemented in a family of codes following the Fast Lagrangian Analysis of Continua (FLAC) algorithm (Cundall and Board, 1988). These specific implementations of the generic FLAC algorithm have a track record of applications that demonstrate the method's aptitude for Long-term Tectonic Modeling (LTM) (e.g., Behn and Ito, 2008; Buck et al., 2005; Choi and Gurnis, 2008; Huet et al., 2011; Ito and Behn, 2008; Lyakhovsky et al., 2012; Poliakov and Buck, 1998; Poliakov et al., 1993). The original FLAC requires a structured quadrilateral mesh which severely limits the meshing flexibility, one of the major advantages of finite element method.

    • Early Cretaceous fracture zones in the Bay of Bengal and their tectonic implications: Constraints from multi-channel seismic reflection and potential field data

      2012, Tectonophysics
      Citation Excerpt :

      Radhakrishna et al. (2010) inferred a thin crust below the thick sedimentary regions of Bengal Fan, including the deep offshore Mahanadi basin. For this thin crust, Choi and Gurnis (2008) suggested the development of closely spaced fracture zones due to enhanced brittle deformation at the spreading ridge. The spacing of fracture zones (FZ10–FZ16) and the inferred thin crust in the deep offshore Mahanadi basin might provide together some clues with regard to the dynamics of seafloor spreading during the early breakup history of the margin.

    View all citing articles on Scopus
    View full text