grl53163-sup-0001-supinfo

advertisement
1
2
[Geophysical Research Letter]
3
4
[Interseismic deformation of the Shahroud fault system (NE Iran) from Space-
5
borne Radar Interferometry measurements]
6
[Z. Mousavi1, 2, E. Pathier1, R. T. Walker3, A.Walpersdorf1, F. Tavakoli4, H. Nankali4, M. Sedighi4,
7
M-P Doin1]
8
[1) ISTerre, Université Joseph Fourier, CNRS UMR 5275, Grenoble, France (zahra.mousavi@ujf-
9
grenoble.fr)
Supporting Information for
10
2) Department of Earth Sciences, Institute for Advanced Studies in Basic Sciences, Zanjan, Iran
11
3) Department of Earth Sciences, University of Oxford, South Parks Road, Oxford, OX1 3AN, UK.
12
4) National Cartographic Center, Geodetic Department, Tehran, Iran]
13
14
15
16
17
18
19
20
Contents of this file
21
Text S1. Removing the tropospheric delay, orbit error and residual DEM
22
The tropospheric delays can be split up into turbulent and stratified components. The
23
effect of the turbulent component on SAR acquisitions can be considered as random in
24
space and time and is mostly topography independent. Therefore it can be reduced by
25
stacking or time-series analysis [Schmidt and Burgmann, 2003]. The vertically stratified
26
component, correlated with topography, is not random in space [Doin et al., 2009] and it
27
should be removed before tectonic interpretation. Here we use the global atmospheric
Text S1. Removing the tropospheric delay, orbit error and residual DEM
Figures S1 to S5
Tables S1
1
28
model ERA-Interim provided by the European Center for Medium‐Range Weather
29
Forecast (ECMWF) to estimate the tropospheric delay map. ERA‐I is a global
30
atmospheric model on a ∼75 km grid, generated four times per day and on 37 pressure
31
levels. The used global atmospheric model like ERA-I have been shown reducing biases
32
in INSAR multi-years strain rate estimate [Jolivet et al., 2011]. We did a spatial bilinear
33
interpolation for the delay function of ERA data and then a spline interpolation for
34
altitude to predict the delay map for each single image [Jolivet et al., 2011]. Differential
35
delay maps corresponding to each interferogram can be estimated by comparing these
36
delay maps. The residual orbital components due to orbital errors can still be significant,
37
particularly when studying low rates of deformation. Such errors are reduced from the
38
wrapped interferometric phase by searching for a best fitting ramp in range, followed by
39
a best fitting ramp in azimuth. We do not try to correct for more complex orbital error at
40
that stage to avoid removing the tectonic signal. Indeed the interseismic signal related to
41
the SFS that we want to estimate should be mainly localized within a few tens of
42
kilometers from the fault zone as shown on other strike-slip fault studies (e.g. Wright et
43
al. 2001, Wang et al 2009, Jolivet et al 2014). In this case, using a simple phase ramp
44
orbital error model estimated over 300km-long track minimize the risk of removing
45
tectonics signal. Comparison with independent GPS data will be done to check a-
46
posteriori the validity of this assumption. The interferometric phase component due to
47
DEM errors varies with the perpendicular baseline between the two acquisitions and is
48
estimated from wrapped interferograms following the methodology from Ducret et al.,
49
[2013]. The main interest of this correction is to reduce the phase variance for helping the
50
unwrapping process. Some residual DEM errors may remain after this step; they are re-
2
51
estimated during the time-series analyses step described hereafter. Then, interferograms
52
are filtered using Goldstein's filter [Goldstein and Werner, 1998] and unwrapped using a
53
branch-cut algorithm [Goldstein et al., 1988]. A careful visual inspection of the corrected
54
interferograms indicates that there is no evidence for any sharp creep deformation signal
55
located along the faults over the 2003 to 2010 period. To investigate the long wavelength
56
tectonic signal due to interseismic strain accumulation, a time series analysis of the
57
selected images has been performed on a pixel basis in order to enhance the signal to
58
noise ratio.
59
Reference:
60
Doin, M. P., C. Lasserre, G. Peltzer, O. Cavalié, and C. Doubre (2009), Corrections of stratified
61
tropospheric delays in SAR interferometry: Validation with global atmospheric models, J. Appl.
62
Geophys., 69, 35–50.
63
Ducret, G., M.-P. Doin, R. Grandin, C. Lasserre, S. Guillaso (2013), DEM Corrections Before Unwrapping
64
in a Small Baseline Strategy for InSAR Time Series Analysis, Geoscience and Remote Sensing
65
Letters, IEEE, Volume: PP , Issue: 99, 1 – 5.
66
67
68
69
70
71
Goldstein, M. R., A. H. Zebker, and C. L. Werner (1988), Satellite radar interferometry: Two-dimensional
phase unwrapping, Radio Sci., 23(4), 713–720.
Goldstein, M. R., and L. C. Werner (1998), Radar interferogram filtering for geophysical applications,
Geophys. Res. Lett., 25(21), 4035–4038.
Jolivet, R., R. Grandin, C. Lasserre, M.-P.Doin, and G. Peltzer (2011), Systematic InSAR tropospheric
phase delay corrections from globa l meteorological reanalysis data, Geophys. Res. Lett., 38, L17311.
72
Jolivet, R., P. S. Agram, N. Y. Lin, M. Simons, M. P. Doin, G. Peltzer, and Z. Li (2014), Improving InSAR
73
geodesy using global atmospheric models. Journal of Geophysical Research: Solid Earth, 119(3),
74
2324-2341.
3
75
Schmidt, D. A., and R. Bürgmann (2003), Time-dependent land uplift and subsidence in the Santa Clara
76
valley, California, from a large Interferometric synthetic aperture radar data set, J. Geophys. Res.,
77
108(B9), 2416
78
79
Wang, H., T. J. Wright, and J. Biggs (2009), Interseismic slip rate of the northwestern Xianshuihe fault
from InSAR data, Geophys. Res. Lett. 36, no. 3).
80
Wright, T., B. Parsons, and E. Fielding (2001), Measurement of interseismic strain accumulation across the
81
North Anatolian Fault by satellite radar interferometry, Geophys. Res. Lett. 28, no. 10, 2117-2120.
82
Figures:
83
84
85
86
87
88
89
Figure S1. Diagrams of all processed interferograms selected using a maximum
perpendicular baseline criterion (500m). Perpendicular baselines with respect to the July
2003 orbit for descending track D020 (left plot), and January 2004 orbit for ascending
track A156 (right plot), are plotted as function of acquisition dates. Track D020 uses 22
images that are combined into 63 interferograms and track A156 uses 23 images that
are combined into 51 interferograms.
4
90
91
92
93
Figure S2. Velocity versus altitude for the area covered by track D020 from which it is
clear that there is no correlation between altitude and deformation.
94
95
96
97
98
Figure S3. Solution space plot for modeling the location of fault and slip rate for the two
tracks A156 (left) and D020 (right). The red stars are the best-solution of modeling the
location and slip rate of fault based on fix locking depth at 12km and the green stars are
the location of fault that we choose for the 2d modeling of slip rate and locking depth.
5
99
100
101
Figure S4. One example of a perturbed dataset: a) the original interferogram, b) the
perturbation image c) perturbed data d) the RMS misfit for perturbed data.
102
6
103
104
105
106
107
Figure S5. Solution space plot for modeling the location of fault and slip rate for the two
tracks A156 (left) and D020 (right). The red stars are the best-solution of modeling the
location and slip rate of fault based on _x locking depth at 12km and the green stars are
the location of fault that we choose for the 2d modeling of slip rate and locking depth.
108
109
Tables:
110
111
112
113
Table S1. Correlation coefficients between LOS mean velocity map and topography,
estimated separately on the northern and southern part (with respect to the fault zone) of
each InSAR tracks. These values are all low, indicating that there is no significant
correlation between the mean velocity and topography.
114
Northern part
Southern part
0.1643
0.0733
-0.4874
-0.129
correlation coefficient track
D020
correlation coefficient track
A156
115
116
7
Download