The Strength of Active Faults


George Hilley, Department of Geology, Arizona State University, Tempe, Arizona 85281

Abstract

The strength of active faults in the lithosphere influences the orientation of the stresses acting on the fault. For faults with low shear strength, stresses applied remotely must reorient themselves such that the shear traction applied to the fault does not exceed the strength of the fault. In this study, we take the formuations of Zoback et. al (1987) relating the stress rotation around a weak fault as a function of fault strength, and rederive the equation to relate the stress rotation to fault friction. We show that changes in strike of a strike slip fault can be used to calculate the driving stresses responsible for movement along the fault as well as fault friction.

We use these formulations relating fault friction and stress rotation to evaluate which processes may be responsible for frictional weakening of major strike-slip faults and which cannot. In particular, we investigate the hypothesis that frictional weakening along the fault occurs as the result of hyperbrecciation along the fault surface. We compare the stress rotation around two strike-slip faults with similar total displacment to determine the possible frictions acting on these faults. In the case that frictional weakening results from hyperbrecciation, two strike-slip faults with similar total displacement should have similar fricitonal strengths. We investigate the validity of this hypothesis by comparing the stress rotation around the Altyn Tagh fault in Central Asia and the San Andreas Fault in California. The San Andreas Fault shows significant stress rotation in the vicinity of the fault, indicating a low frictional strength of the fault. In contrast, the Altyn Tagh fault shows no such rotation, indicating that it is a well-oriented, strong fault. Both faults have accumulated hundreds of kilometers of total displacement. These results suggest that hyperbrecciation does not play a role in reducing the faults' frictional strength. We are currently investigating a total of 12 strike-slip faults in order to strengthen this suggestion and evaluate the correlation of low fault strength with strike-slip faults that are plate boundaries.


I. Introduction

Faults in the brittle upper crust behave as frictional sliding surfaces. In these types of faults, the displacement occurs after some critical combination of shear stress and normal stress is applied to the fault surface. This critical value scales linearly with the ratio of the normal stress to shear stress. Laboratory samples and most active faults have critical values at ratios of normal stress to shear stress around 0.6-0.8. However, a certain class of faults fail at anomalously low ratios. Three mechanisms have been invoked to explain this behavior. First, high pore pressure (near hydrostatic stresses) in the fault zone may act to weaken the fault. Second, accumulated slip along the fault may lead to hyperbrecciation which weakens the fault. Third, minerology of the fault zone may act to lubricate the fault surface, reducing the fault friction. Laboratory experiments show that only pure clay is sufficient to minerologically lubricate the fault zone. The stability of clays above 100C precludes the presence of the minerals' presence at seismogenic depths. Therefore, high pore pressures and hyperbrecciation may be the realistic processes that produce a weak fault.

Faults with low frictional strength will cause a rotation of the remote stresses near to the fault. The strength of the fault cannot be exceded without failure; therefore, the principle stress orientations must rotate into an orientation such that the shear traction resolved from the local stress tensor onto the fault surface equals the strength of the fault (Zoback et. al, 1989). This relationship can be derived as (from Zoback et. al, 1989):

equation1

Where alpha is the orientation of the greatest compressive stress relative to the fault plane in the area near the fault, SHmax is the maximum horizontal compressive stress, Shmin is the minimum horizontal compressive stress, Co is the fault strength (in MPa), and beta is the angle of the greatest compressive stress relative to the fault plane in the far-field. Hence, the driving stresses responsible for the formation of the fault and subsequent movement are oriented at an angle, beta, to the fault surface. As one approaches the fault, the driving stresses rotate to an orientation where alpha is the angle between the fault surface and the direction of the maximum horizontal compressive stress (SHmax).

We can write the strength of the fault (Co) in terms of the normal traction (sigma(n)) and the friction(mu) as:

equation2

These two sets of equations can be solved itteratively to determine the induced rotation due to the presence of a weak fault. Figure 1 shows the relationship between beta and alpha for faults of various strengths (dashed lines, using Zoback et. al (1989) formulation) and frictions (solid lines, using itterative technique described above).

figure 1
Figure 1: Relationship between alpha (angle between fault plane and SHmax near the weak fault) and beta (angle between far-field driving stresses and fault plane) for different strengths (in MPa, dashed lines) and frictions (solid lines).


II. Frictional Strength of the San Andreas Fault

Heat anonmalies, stress drops during earthquakes, and stress rotations adjacent to the San Andreas Fault suggest that the fault has low frictional strength. In the vicinity of the San Andreas Fault, the orientation of SHmax rotates nearly orthogonal to the fault (Figure 2).

california map
Figure 2: Stress orientations in California (taken from Zoback et. al, 1989; scan courtesey of Zack Washburn). Note the near-orthogonality of the orientation of SHmax to the San Andreas Fault.

The far-field driving stresses responsible for movement along the San Andreas Fault are complicated by the scale of the fault and the presence of other stress provinces in the area. In order to determine the orientation of the far-field driving stresses, we study stress orientations in the region of the Big Bend. If the San Andreas is being fueled by a relatively constant far-field stress orientation, then the change in the strike of the fault results in a change in beta in Figure 1. Therefore, the angular relationship between the fault strike and the strike of SHmax near the fault should change through the bend. We use this fact to calculate the best-fit stress orientations driving the San Andreas Fault and the corresponding fault friction required to rotate the stresses to the angle that they make with the strike of the fault. We take two populations of stress measurements, one north of the Big Bend and one south of the Bend. Since beta in Figure 1 must be 33 degrees greater than Beta in Figure 1 south of the bend, we can perform a least squares inversion minimizing the misfit between the data and the predicted rotation from Figure 1 and backsolve for the orientation of SHmax in the far-field and the friction of the fault. The results of this inversion are shown in Figures 3 and 4. Figure 3 lets all variables optimize (free variables are differential stress, friction, and orientation of the far-field driving stresses). Figure 4 shows the inversion fixing the differential stress estimated by Zoback et. al (1989). The differential stress determined from the full optimization was 150.0800 MPa heuristically agrees well with the Zoback et. al (1989) estimate of a differential stress of 136 MPa. Computed orientations of SHmax is almost exactly N-S and computed frictions were low (mu = 0.0282 for full optimization and mu = 0.0090 for constrained differential stress optimization).


optimized solution
Figure 3: Constrained optimization of stress data around San Andreas Fault. Friction, differential stress, and remote stress orientation were optimized in this calculation. Note the N-S orientation of SHmax (strike of SAF for points in left half of figure is 315 degrees) and the low fault friction (mu = 0.0090).

Constrained Optimization
Figure 4: Constrained optimization of stress data around San Andreas Fault. Friction and remote stress orientation were optimized in this calculation with a constant differential stress of 136 MPa. Again, the fault friction and SHmax orientation are consistent with the results of the fully optimized model and the heat flow data collected along the San Andreas Fault.


III. Processes responsible for the formation of a low-friction fault

From the above investigation, we conclude that the San Andreas Fault is a weak fault. If we assume that this weakening is the result of hyperbrecciation resulting from accumulated displacement along the fault, then faults with similar displacement should have similar frictions. These low frictions should be apparent in the stress data as rotations of the remote stress near the fault. In order to test this hypothesis, we consider the Altyn Tagh fault in Central Asia. Slip along the Altyn Tagh is on the order of the San Andreas fault; however, the fault is not accomodating direct plate motion as a transform fault (as is the San Andreas) but is likely accomodating the escape of Asia from the Indo-Eurasian collision to the south. The location and some of the stress data from the region is shown in Figure 5.

china image
Figure 5: Location and stress orientations of the Alytn Tagh Fault, Central Asia.


Data from the World Stress Map Project was used to determine the degree of rotation in the vicinity of the Altyn Tagh fault. Table 1 shows the mean and standard deviation of the stress measurements for the different data quality classes defined by the World Stress Map Project and the mean and standard deviation for the entire population of stress data. In short, there is strong evidence of no stress rotation around the Altyn Tagh fault. From this, we infer that the fault is strong. Based on a comparison with the San Andreas Fault which has similar amounts of total displacement, we find it difficult to accept hyperbrecciation as a process of fault weakening.

Table 1: Results from Stress Analysis of Altyn Tagh Fault
Quality RankingMeanStd. Deviation
A35.510.6
B51.314.9
C40.216.7
D41.610.5
E18.94.6
All4016.7



IV. Conclusions and Future Work

We conclude that 1) the San Andreas Fault is a weak fault as evidenced from the stress rotation surrounding the fault, 2) by comparison of the Altyn Tagh fault with the San Andreas Fault, we feel that hyperbrecciation may not be responsible for decreases in fault friction. It is extrememly tenuous to base such grandiose conclusions on two data points. Therefore, we are currently investigating other active strike-slip faults in order to determine the relationship between total accumulated offset, tectonic setting (plate-bounding faults and non-plate-bounding faults) and fault friction.


V. References:

Zoback, M. D., and 12 others, New Evidence on the State of Stress of the San Andreas Fault System, Science, November 1987, pp. 1105-1111