jgrb51352-sup-0001-s01

advertisement
1
2
Journal of Geophysical Research – Solid Earth
3
Supporting Information for
4
Intraplate rotational deformation induced by faults
5
Neta Dembo1,2, Yariv Hamiel2, Roi Granot1
6
7
8
1Department
of Geological and Environmental Sciences, Ben-Gurion University of the Negev, Beer-Sheva, Israel,
2Geological Survey of Israel, Jerusalem, Israel.
9
10
11
12
13
14
Contents of this file
15
Introduction
16
17
18
The supporting information presented in Text S1 discusses the robustness of our elastic
modeling approach Text S2 provides a detailed description of the interseismic modeling
parameters and results.
19
20
Figure S1 provides a comparison between the rotational patterns predicted by an elastic halfspace solution to those of both elastic and elasto-plastic finite element models.
21
22
Figure S2 provides the same figures as in figure 2a-b in the main article, but for a right lateral
strike-slip fault.
23
24
25
Figure S3 provides 40Ar/39Ar step heating results used to determine basalt ages. Dating
measurements and analysis were performed by the New Mexico Geochronological Research
Laboratory at the New Mexico Institute of Mining and Technology.
Text S1 and S2
Figures S1 to S3
26
27
28
29
30
31
32
33
1
34
Text S1.
35
To assess the viability of our elastic modeling approach we compared the results of the long-
36
term rotational patterns predicted by an elastic half-space solution to the results of both elastic
37
and elasto-plastic finite element models.
38
We modeled a synthetic left-lateral strike slip fault extending to a length of 100 km using a slip
39
rate of 2.5 mm/yr. The elastic half-space solution was modeled using a Poisson ratio of 0.25
40
and given a fault extending from the surface downwards to a depth of 1000 km (in order to
41
mimic infinite fault width). The finite element modeling is done using COMSOL commercial
42
software. The computational domain that mimics the lithosphere used for the finite element
43
models has a size of 200 x 200 x 100 km3 with element length of ~2 km near fault and
44
increasing towards the block boundaries. Boundary conditions include tangential
45
displacements at the outermost lateral boundaries, which are parallel to the fault, and zero
46
displacement at the bottom of the block. A Poisson ratio of 0.25 and a Young’s modulus of 75
47
Mpa are used for the elastic finite element model. The same elastic moduli are used in the
48
elasto-plastic model until a yield stress of 150 MPa. After this yield stress an effective Young’s
49
modulus of 25 Mpa is defined.
50
Comparison between the deformation fields predicted by the three models resulted in similar
51
rotation patterns (Figure S1). Therefore, the results from our synthetic modeling suggest that
52
our approach using an elastic half-space dislocation model can be used to simulate patterns of
53
permanent deformation near faults.
54
55
2
56
57
58
Text S2.
59
rate of 4.9 mm/yr (in the range of 3.8-4.9 mm/yr) and locking depth of 15.9 km (in the range of
60
14.1-15.9 km) along the DSF segment found south of the junction with the CGFS (Figure 8) to a
61
slip rate of 3.8 mm/yr (in the range of 3.8-4.9 mm/yr) and locking depth of 14.1 km (in the range
62
of 14.1-15.9 km) along the DSF segment north of this junction. Due to small number of GPS
63
stations in the eastern side of this junction, we cannot point to an exact location where this
64
change in slip occurs, also we cannot rule out the idea that this transition is gradual.
Our best fitting interseismic model for the DSF in northern Israel shows a decrease from a slip
65
66
Our study confirms Sadeh et al. [2012] slip rate of 1.1±0.4 mm/yr along the Carmel Fault, which
67
is also consistent with the decrease in slip along the DSF (slip direction is in azimuth of ~6˚). We
68
infer a best fitted locking depth of 5 km (in the range of 5-15 km) for this fault. The model also
69
indicates a locking depth of 5 km (in the range of 5-15 km) along the Gilboa Fault together with
70
a slip rate of 0.75±0.75 mm/yr (slip direction in azimuth of ~31˚), which is in agreement with
71
geological slip rates reported for this fault by Hatzor and Reches [1990].
72
73
Based on our modeling, interseismic rotations are relatively insensitive to locking depths
74
confined between 5 and 16 km and to the range of examined slip rates along the DSF (3.8-4.9
75
mm/yr) and CGFS (up to 1.5 mm/yr). However, the RMS value of the best fitting model of the
76
DSF fault alone is still higher than the RMS of the model including both DSF and CFS faults.
77
This indicates that by adding the CFS faults to the model the fit to the GPS data improves,
78
which also strengthens the theory of slip portioning from the DSF to CGFS.
79
80
81
82
3
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
Figure S1. Rotation rates predicted by: (a) elastic half space model, (b) elastic finite element
model, (c) elasto-plastic finite element model. A detailed description of the models setup is
found in Text S1 of the auxiliary material.
4
117
118
119
120
121
122
123
124
125
126
127
128
129
130
Figure S2. Surface vertical axis rotation rates caused in response to motion of synthetic faults.
(a) An interseismic model of a vertical right-lateral strike slip fault locked to 15 km depth. (b) A
total deformation slip model of a vertical right-lateral strike slip fault that slips up to the
surface. Both faults were arbitrary set to a length of 200 km using a slip rate of 5 mm/yr. The
parameters were chosen to best reflect and emphasize the differences between the different
styles of deformation.
131
5
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
Figure S3. 40Ar/39Ar step heating of Neogenic basalts. Arrows show the steps included by the
plateau age calculation. Dating measurements and analysis were performed by the New
Mexico Geochronological Research Laboratory at the New Mexico Institute of Mining and
Technology. MSWD stands for the mean sum weighted deviates.
6
Download