Auxiliary_Material_text

advertisement
1
1. Fault rotation model
2
In a previous publication [Olive and Behn, 2014], we proposed a conceptual
3
framework for understanding the rapid rotation of normal faults from a steep 60º-
4
initiation angle down to dip angles in the 30–45º range, over a few km of offset. We
5
proposed that fault evolution proceeded in a manner that systematically minimized the
6
mechanical work required to sustain fault slip, and showed that (1) this assumption
7
results in rotation rates that scale roughly as the inverse of the faulted layer thickness, and
8
(2) accounting for rotation is essential to correctly predict fault lifespan in a force balance
9
model. Here we adopt a complementary approach, which is significantly less
10
computationally challenging to implement, but produces very similar rotation rates (see
11
below). Specifically, instead of assuming work minimization, we model fault rotation as
12
the passive advection of the fault plane in the displacement field induced by flexural
13
relaxation of the footwall and hanging wall blocks.
14
If we denote
15
fault rotation rate will scale roughly as
the average rotation rate of the near-fault displacement field, then the
(S1)
16
To first order, the main source of rotation in the flexural displacement field is the lateral
17
gradient of vertical motion (Figure 1a), which is maximal at the fault and equal to V tanθ,
18
and becomes negligible a fraction of the flexural wavelength (γα, the lever arm, where γ
19
is a scaling factor and α is the flexural wavelength) away from the fault. We therefore
20
write
(S2)
21
Plugging (S2) into (S1) with the constraint h = 2Vt yields
¶q
tanq
=¶h
4ga
22
(S3)
Which can be integrated into
æ
æ 1
q = sin ç sinq 0 exp ç è 4g
è
-1
dh ö ö
ò0 a (h) ÷ø ÷ø
h
(S4)
23
In addition, we taper the rotation rate at heaves greater than 80% of the faulted layer
24
thickness so that rotation ceases at large offsets, when the fault has become curved and
25
the idealized fault geometry depicted in Figure 1a no longer applies.
26
Fault rotation enters the force balance model in several ways. As extension
27
proceeds, the elastic-plastic iterations predict the fault-induced topography [Buck, 1988;
28
1993]. At every extension step (h), we fit this topography with the elastic solution of
29
Weissel and Karner [1989] and determine an equivalent elastic thickness for the entire
30
plate. We then use it to calculate an equivalent flexural wavelength, α(h), which enters
31
into the calculation of the next fault dip at step (h+Δh) following Equation (S4). We find
32
that a scaling factor γ = 0.25 provides a good fit to the numerical models, regardless of
33
layer thickness and other parameters. In other words, the lever arm associated with fault
34
rotation corresponds to about a fourth of the effective flexural wavelength of the faulted
35
layer.
36
37
Once fault dip has been updated, it enters the calculation of the frictional force
FFRIC, following [Behn and Ito, 2008]
FFRIC =
38
mr gH / 2 + HC
m sin 2 q + sinq cosq
(S5)
For reference, the threshold for breaking a new fault is given by
FBREAK =
mr gH / 2 + HC0
m sin 2 q 0 + sinq 0 cosq0
(S6)
39
where θ0 corresponds to the Andersonian dip for a new fault breaking intact lithosphere.
40
(All notations are summarized in Figure 1a).
41
Fault dip also enters the calculation of the bending and topographic forces. A fault
42
that rotates as it accumulates offset generates less throw than if it were to retain its initial
43
angle. We therefore slightly modify the method of Weissel and Karner [1989] by using
44
an effective dip θEFF in the loading term of the flexure equation (w*(x) in Olive and Behn,
45
2014). For a given amount of heave (h) this effective dip yields the amount of throw that
46
would have been reached by a rotating fault, and can be written:
æ1h
ö
ò tanq (h) dh÷ø
èh
q EFF = tan -1 ç
0
(S7)
47
θEFF then enters the calculation of the fault-induced topography and the associated work
48
terms. Specifically, the bending work WBEND is obtained by integrating the bending stress
49
times the bending strain over the entire layer [Olive and Behn, 2014], while the
50
topographic work WTOPO corresponds to the change in gravitational potential energy
51
associated with offsetting the air-rock density contrast, Δρ, from its initial flat state, and is
52
written
+¥
WTOPO =
ò Dr g y ( x ) dx
2
-¥
(S8)
53
54
2. Comparison with the energy minimization model of Olive and Behn [2014]
55
In [Olive and Behn, 2014], we suggested that fault rotation proceeds in a manner
56
that minimizes the work needed to sustain fault slip, but did not explicitly consider the
57
physical mechanism that causes rotation. The present approach suggests that passive
58
advection of the fault plane in a flexural displacement field provides a good candidate to
59
explain the kinetics of rotation observed in numerical experiments. Indeed, this new
60
model predicts the rotation kinematics observed in our simulations (Figure 2b). Much
61
like the energy minimization model, it accounts for rapid rotation of normal faults down
62
to dip angles of 30–45º over fault heaves shorter than half of the faulted layer thickness.
63
Further, the inverse scaling of rotation rate with layer thickness is evident from Equation
64
(S3) and is derived from the fact that thicker layers bend on a larger wavelength, making
65
faults rotate about a longer lever arm.
66
Finally, we find that the present kinematic model for fault rotation predicts dips
67
that remain close to the lowest energy configuration throughout fault evolution, even
68
though energy minimization was not assumed. Supplementary Figure fs01 compares the
69
kinematic elastic-plastic solution for a 15 km thick layer (red line in Figure 2b) to an
70
energy minimization model where plasticity is accounted for by assuming an effective
71
elastic thickness of 7.5 km. Colors indicate the total amount of work required to sustain
72
slip on the fault. The kinematic rotation model predicts slightly slower rotation then the
73
energy minimization model for small offsets, which provides a better fit to numerical
74
experiments (Figure 2b).
75
76
Supplementary captions
77
78
Supplementary Figure fs01 – Total external work (WEXT = WFRIC + WBEND + WTOPO)
79
required to keep a normal fault active as a function of fault dip and heave for increasing
80
amounts of extension in a 15 km thick elastic pseudo-plastic layer with an effective
81
elastic thickness of 7.5 km. The white dashed line indicates the lowest energy path
82
computed following the model of Olive and Behn [2014] taking into account all energy
83
terms: WFRIC, WBEND, and WTOPO, the last of which was previously ignored. The thick
84
white line shows the kinematic rotation model used in this study.
85
86
Supplementary Table ts01 – Summary of parameters and results of our numerical
87
experiments.
88
Download