Sie sind auf Seite 1von 17

Large-area land subsidence

monitoring and mechanism research


using the small baseline subset
interferometric synthetic aperture
radar technique over the Yellow
River Delta, China

Yilin Liu
Haijun Huang
Jinfa Dong
Large-area land subsidence monitoring and mechanism
research using the small baseline subset interferometric
synthetic aperture radar technique over the Yellow
River Delta, China

Yilin Liu,a,b,* Haijun Huang,a and Jinfa Dongc


a
Chinese Academy of Sciences, Institute of Oceanology, The Key Laboratory of
Marine Geology and Environment, 7 Nanhai Road, Qingdao 266071, China
b
University of Chinese Academy of Sciences, Beijing 100049, China
c
Tianjin Navigation Instruments Institute, 268 Dingzigu First Road, Tianjin 300131, China

Abstract. The Yellow River Delta, the largest and most rapidly growing delta in China, has
undergone severe land subsidence. However, land subsidence has not been regularly and
fully measured in this large region and no consensus has been reached as to the primary
cause. Here, the small baseline subset interferometric synthetic aperture radar (InSAR) method
is applied to retrieve time-series deformation and the full land subsidence pattern in the Yellow
River Delta. Spirit leveling data and a standard deviation map are used to verify the InSAR
results and measurement accuracy. The major subsidence regions, which have a maximum sub-
sidence rate ranging from 20 to more than 70 mm∕year, cover a total of ∼800 km2 and are
mainly concentrated in the groundwater source areas. An intimate connection of surface defor-
mation with groundwater suggests that land subsidence is primarily caused by excessive exploi-
tation of groundwater. In addition, comparison of the surface deformation with the precipitation
record indicates that the seasonal land subsidence correlates with the rainfall rate, with a lag time
of around 1 to 2 months between the precipitation peak and the minimum land subsidence dis-
placement. The results provide new insights into land subsidence mechanisms in the Yellow
River Delta. A key policy priority should, therefore, be to plan for controlling anthropogenic
activities and better management of groundwater. © 2015 Society of Photo-Optical Instrumentation
Engineers (SPIE) [DOI: 10.1117/1.JRS.9.096019]

Keywords: deltaic land subsidence; small baseline subset-interferometric synthetic aperture


radar; Stanford method for persistent scatterers; groundwater; precipitation.
Paper 15349 received May 17, 2015; accepted for publication Jul. 22, 2015; published online
Aug. 20, 2015.

1 Introduction
Land subsidence is defined as a gradual settling or sudden sinking of the Earth’s surface owing to
subsurface movement of Earth materials.1 Subsidence of deltas is a global problem, and in
coastal regions, many large deltas undergo rapid land subsidence, which results in vulnerability
to storm surge, saltwater intrusion into groundwater supplies, shoreline retreat, loss of wetlands,
coastal flooding, and infrastructure destabilization.2–6 Similar to most of the world’s deltas,7,8 the
Yellow River Delta (China) is greatly influenced by anthropogenic activities, especially since
Dongying city was established in 1983 by the central government. The land subsidence rate of
the delta, accelerated by excessive anthropogenic activities, is much greater than other deltas in
the world.9,10 Moreover, special flat rolling lowlands of the delta make it more highly threatened
by environmental problems associated with the ongoing rapid land subsidence. Though many
studies have examined the delta,11–16 most existing direct measurements over the large area are

*Address all correspondence to: Yilin Liu, Email: lyilin@msn.com


1931-3195/2015/$25.00 © 2015 SPIE

Journal of Applied Remote Sensing 096019-1 Vol. 9, 2015


Liu, Huang, and Dong: Large-area land subsidence monitoring and mechanism. . .

point measurements, and no consensus has been reached as to the primary cause of the sub-
sidence. Large coverage and high resolution monitoring is needed to fully characterize the sub-
sidence pattern.
In recent years, many different mapping techniques have been implemented for the carto-
graphic representation of land subsidence. Accurate subsidence mapping is fundamental and
critical for hazard assessment research. In comparison to traditional investigation and recon-
naissance methods, such as spirit leveling, extensometer, global positioning system, optical
remote sensing, and geological and geophysical investigation methods,1,17 the interferometric
synthetic aperture radar (InSAR) technique, using multiple images and various advanced
analysis methods, has great advantages due to its broad coverage and high spatial (and some-
time temporal) resolution under all weather conditions.18–21 It is, therefore, the best method to
reconnoiter the full pattern of land subsidence, especially over a large region. The InSAR
technique has been widely used in various subsidence studies, and the derived subsidence
pattern further provides critical insights into the dynamics of subsidence, including ground-
water withdraw, tectonic activities, sediment compaction, and underground hydrocarbon
extraction.22–31 However, the application of the InSAR technique to deltaic subsidence inves-
tigations is still a challenging topic due to the spatial and temporal decorrelation, atmospheric
delay anomalies, and artifacts induced by the variation of coastal near surface moisture in the
wetland, cropland, and aquaculture areas.32–35
In this study, the Stanford method for persistent scatterers/multitemporal InSAR (StaMPS/
MTI) package was employed to map a large region of land subsidence in the Yellow River Delta,
which can address the above difficulties and eliminate or mitigate the chance of phase unwrap-
ping, topographic, and atmospheric delay errors and other noise.21,36 The time-series deforma-
tion results and full subsidence pattern are achieved, and the accuracy of deformation
measurements is verified using a standard deviation map and spirit leveling data. The correlation
between surface deformation and groundwater, the time-series deformation derived, and the pre-
cipitation record are analyzed in detail.

2 Geological and Hydrological Setting of the Study Area


The Yellow River Delta covers an area of more than 5800 km2 north of the Shandong
Peninsula on the north east coast of China (Fig. 1). The delta is formed as a consequence
of siltation, river channel avulsions, flooding, and sedimentation since the Yellow River, the
cradle of Chinese culture, resettled in the Daqing River and the river mouth shifted north
from the Yellow Sea to the Bohai Sea in 1855. The Dongying depression, where the delta
is located, is a northeast-southwest-trending lacustrine basin in the southern part of the Jiyang
sub-basin in the Bohai Bay Basin. The depression is controlled by a series of normal faults
(Fig. 1) that developed in an en echelon pattern,37 filled with Cenozoic sediments, which
are composed of Paleogene formations including the Kongdian (Ek), Shahejie (Es), and
Dongying (Ed) formations, as well as Neogene (N) and Quaternary (Q) formations. The vertical
stratigraphic sequence is composed of interbedded sand, silt, silty clay, and clay. The delta is
characterized by Holocene marine sediment with a thickness of ∼30 m, and Quaternary sedi-
ment with a thickness of 300 to 500 m.
The groundwater of the Yellow River Delta, mostly loose rock pore water, mainly occurs in
the alluvial and marine deposits in the upper part of the Quaternary system.38 The hydrogeology
of the delta can be divided into three units (Fig. 6): H1, the piedmont alluvial fan (all freshwater);
H2, the piedmont alluvial-proluvial plain and the marine-continental sedimentation formed via
the interaction between the Yellow River siltation and the tidal current (salt water below fresh-
water); and H3, the Yellow River alluvial plain (pure salt water).39 Though a large area of saline
water in the alluvial plain has gradually become fresh water, fresh water is still limited in the
delta and can only be found on the sides of the Yellow River course and along the old Yellow
River course in the marine-alluvial plain.40 Due to the dams built in the upper and middle reaches
of the Yellow River, which provides ∼90% of the total water demand of the delta, the river flux
has been decreasing dramatically, and the discharge in the beginning of the 21st century
decreased to 11.3% of that in the 1950s.41

Journal of Applied Remote Sensing 096019-2 Vol. 9, 2015


Liu, Huang, and Dong: Large-area land subsidence monitoring and mechanism. . .

Fig. 1 Shaded relief map and geological sketch of the Yellow River Delta. YD01, the bedrock
mark, is the reference point of spirit leveling. YD14, the benchmark, is the center of the square
reference area, with 500 m on a side for interferometric synthetic aperture radar (InSAR). D18 is
the share benchmark of the two level lines (Shandong Provincial Lubei Geo-engineering
Exploration Institute, SPLGEI). The four faults, F1 (the Chennan fault), F2 (the Gaoqing-
Pingnan fault), F3 (the Shicun fault), and F4 (the Bamianhe fault) constitute the boundary of
the Dongying depression, and F5 (the Central fault zone), F6 (the Chenguanzhuang fault), F7
(the Binnan fault), and F8 (the Lijin fault) are the four main faults of the depression (modified
from the Academy of Geological Sciences of the Shengli Oil Field). The top left inset is the
base map of Dongying, Shandong province, China.

3 Methodology and Data Processing

3.1 Small Baseline Subset Method


To improve the detecting ability and monitoring accuracy, one of the advanced InSAR tech-
niques, termed as the small baseline subset (SBAS) InSAR, has been published and frequently
applied.20,23,42–44 Because of the spatial and temporal decorrelation, the spatial and temporal
baseline of each interferogram must be limited.45 Considering a set of N þ 1 coregistered SAR
images relative to the same area, acquired at the chronologically ordered times ðt0 ; : : : ; tN Þ, M
interferograms can be generated below the chosen threshold of the small baseline conditions (N
is assumed odd):
ðN þ 1Þ∕2 ≤ M ≤ NðN þ 1Þ∕2:
EQ-TARGET;temp:intralink-;e001;116;235 (1)

For a given unwrapped generic i’th interferogram, following the flat Earth and topographic
phase component removal, the residual phase in the azimuth and slant range coordinate ða; sÞ
computed from the SAR acquisition times tA (start time) and tB (end time) can be written as
follows:42

δϕi ða; sÞ ¼ ϕðtB ; a; sÞ − ϕðtA ; a; sÞ


EQ-TARGET;temp:intralink-;e002;116;159

¼ δϕdefo
i ða; sÞ þ δϕtopo
i ða; sÞ þ δϕatm
i ða; sÞ þ δϕi
noise
ða; sÞ
4π 4π B⊥i Δz 4π
≈ ½dðtB ; a; sÞ − dðtA ; a; sÞ þ þ þ Δni ; ∀ i ¼ 1; : : : ; M; (2)
λ λ r sin ϑ λ

where λ is the transmitted synthetic aperture radar (SAR) wavelength; ϕðtB ; a; sÞ and ϕðtA ; a; sÞ
represent the phases of the two SAR images generating the interferogram; δϕdefo i ða; sÞ is the

Journal of Applied Remote Sensing 096019-3 Vol. 9, 2015


Liu, Huang, and Dong: Large-area land subsidence monitoring and mechanism. . .

radar slant range deformation phase from tA to tB ; δϕtopo i ða; sÞ, representing the topographic
phase error, is a function of the digital elevation model (DEM) error Δz, the perpendicular base-
line B⊥i , the sensor-target slant range distance r, and the radar look angle ϑ; δϕatm
i ða; sÞ accounts
for the possible atmospheric phase error; and δϕnoise i ða; sÞ is the decorrelation effect and other
noise source. Equation (2) can then be finally organized in the following matrix representation:

Bν ¼ δϕi ;
EQ-TARGET;temp:intralink-;e003;116;675 (3)

where
 T
ϕ1 ϕ − ϕN−1
ν ¼ ν1 ¼ ; : : : ; νN ¼ N
t1 − t0 tN − tN−1
EQ-TARGET;temp:intralink-;sec3.1;116;633

and B represents an M × N matrix (see details in Ref. 42). Then the phase at each single SAR
image acquisition time and the time-series deformation can be achieved using least-squares,23
singular value decomposition,42 or minimization of the L1-norm.46 Finally, the accumulative
deformation series can be expressed as follows:

X
tk
δϕdefo
EQ-TARGET;temp:intralink-;e004;116;545

k ða; sÞ ¼ νtk−1 ;tk ðtk − tk−1 Þ; ∀ k ¼ 1; : : : ; N; (4)


t0

where t0 is the acquisition time of the earliest SAR image in the above data set.

3.2 Data and Processing


We collected two tracks of descending Envisat/ASAR data images over the Yellow River Delta.
The SAR data span from early 2007 to late 2010, and a total of 39 frames of SAR images (29 of
track 132, frame 2858 and 10 of track 404, frame 2856) were utilized. The StaMPS/MTI SBAS
technique was employed in this experiment.21 To ensure the reliability of the displacement mea-
surements, all SAR interferograms have time separations of 750 days or less and spatial baselines
of 800 m or less. Figure 2 shows the spatial baselines and temporal separations of the InSAR
pairs used for SBAS processing. Accordingly, 87 interferograms in track 132 and 23 interfero-
grams in track 404 were generated. In addition, a 3-arc sec shuttle radar topography mission-
DEM was used. After spatially filtering each interferogram with the bandpass filter as an adap-
tive phase filter combined with a lowpass filter, pixels were selected and the standard deviation of
phase noise of each selected pixel for all interferograms was calculated.36 Any pixels with a
standard deviation larger than 1.0 rad were eliminated to refine the coherent points, and the
wrapped phase of each final selected pixel was corrected for DEM error.47 The inverted
phase at each acquisition time with respect to a single image can be given after the phase
was unwrapped using one of the three-dimensional unwrapping approaches.48 Then separation
of the displacement and atmospheric signals can be achieved by filtering the resulting time
series, with high-pass filtering in the time domain and low-pass filtering in the space domain.
Finally, time-series deformation of each satellite track was derived and the results were merged.

4 Results, Analyses, and Discussion


The InSAR results for the Yellow River Delta reveal an overall pattern of the displacement from
2007 to 2010. Displacement associated with surface movement in the delta is assumed to be
vertical. Figure 3(a) shows the annual average displacement rate field map. Seven main land
subsidence regions comprise a total of 800 km2 . The respective southern and northern bounda-
ries of the Binzhou (BZ) and Guangrao (GR) subsidence funnels are both aligned along the
Shicun fault (F3). The northern and southern boundaries of the Dingzhuang (DZ) subsidence
funnel are bounded by the Chenguanzhuang (F6) and Bamianhe (F4) faults, respectively, and the
land subsidence region is in the same approximate northeast-southwest direction as the F4 and
F6 faults. Similarly, the central subsidence region [the Dongying (DY) and Longju (LJ) sub-
sidence funnel] follows the general trend of the Central fault zone (F5). The highest subsidence

Journal of Applied Remote Sensing 096019-4 Vol. 9, 2015


Liu, Huang, and Dong: Large-area land subsidence monitoring and mechanism. . .

Fig. 2 Configurations of the InSAR pairs used for SBAS processing and for InSAR images from
(a) track 132 and (b) track 404, respectively.

rates are observable over the groundwater descent funnels, such as in Guangrao county, Binzhou
city and Dongying city, and the largest annual average subsidence rate is 75.6  2.8 mm∕year,
in Guangrao county. The rates exceed the local and global average sea level rise by nearly 2
orders of magnitude. Except for these regions, the delta shows no distinctive pattern of
subsidence.

4.1 Validation of the Accuracy


The standard deviation of the mean velocity in the line-of-sight direction was estimated to deter-
mine the uncertainty of the InSAR results [Fig. 3(b)]. As shown in the figure, the InSAR results
are highly accurate in the vast majority of the Yellow River Delta. Only the accuracy in the
aquaculture area in the coastal region, near profile P3-P3’, is not very high due to a high
level of variable moisture content and prevalent fog in the fall and spring.

Journal of Applied Remote Sensing 096019-5 Vol. 9, 2015


Liu, Huang, and Dong: Large-area land subsidence monitoring and mechanism. . .

Fig. 3 The velocity field of the Yellow River Delta derived from the InSAR time series, showing the
(a) mean velocity and (b) its standard deviation. The four profile locations were determined accord-
ing to the spatial distribution of land subsidence funnels, faults, and groundwater depression
cones.

Except for the standard deviation, validation of the displacement velocity of the InSAR
results against on-field precise leveling survey measurements were conducted to further assess
the accuracy and reliability of the InSAR observations. Level line data, including first- and sec-
ond-order leveling, were collected in the delta beginning in the 21st century (SPLGEI); SPLGEI
performed biyearly leveling surveys and obtained valuable information of land subsidence.
As shown in Fig. 1, not all the benchmarks are covered by InSAR data, so the ones on level
line L1-L1’ and L2-L2’ were chosen. After adjustment of the leveling network, and in order to
compare with InSAR results, we recalculated the annual average displacement rate in the year
2007 to 2008 of each selected benchmark relative to benchmark YD14. The accidental elevation

Journal of Applied Remote Sensing 096019-6 Vol. 9, 2015


Liu, Huang, and Dong: Large-area land subsidence monitoring and mechanism. . .

Fig. 4 Comparison of leveling data (red open squares) and the InSAR time series (blue dots). (a)
Elevation change along a south-to-north path from Guangrao to Dongying (L1-L1’ in Fig. 1). (b)
Elevation change along a west-to-east path from Longju to Dongying (L2-L2’ in Fig. 1). Range
change data are projected onto a vertical unit vector. Gaps in the InSAR profile reflect regions
where the pixel is incoherent, therefore, no data are available.
pffiffiffiffiffiffiffi
change error of leveling is 0.74 km mm. Meanwhile, the annual average InSAR displacement
rate in the two 500-m-wide swaths’ center along level line L1-L1’ and L2-L2’ were extracted
(Fig. 4). The InSAR time-series results compare well with elevation changes measured using
spirit leveling, except for two discrepancy benchmarks. Differences between InSAR time series
and leveling data may reflect additional slight displacements spanning the time when the leveling
data were collected and when the SAR images were acquired, artifacts in the InSAR data, or the
local motions of individual benchmarks.

4.2 Time-Series Deformation


As shown in Fig. 3(a), the main land subsidence regions are located in Guangrao city, Binzhou
city, Dongying city, Boxing county, Dingzhuang county, and the town of Longju town. All are
near or in the historical subsidence funnels49 and close to the groundwater sources of the Yellow
River Delta.50 Additionally, the land subsidence in the Dongcheng district (DC) is due to the
construction associated with the rapid development of the regional economy and urbanization.
However, the subsidence in this region, where a highly aqueous and porous soft clay layer is
fully developed, will in turn have severe impacts on the district construction.
To demonstrate the progression of land subsidence over time in the Yellow River Delta, we
chose eight feature points to illustrate the various time series of deformation near the cities of
Binzhou (BZ) and Dongying (DY), Boxing county (BX), Guangrao county (GR), the towns of
Dingzhuang (DZ) and Longju (LJ), the Dongcheng district (DC) and the reference point (RP)

Journal of Applied Remote Sensing 096019-7 Vol. 9, 2015


Liu, Huang, and Dong: Large-area land subsidence monitoring and mechanism. . .

Fig. 5 Time-series curves of land subsidence at feature points. “RP” is the reference point, “DC” is
the Dongcheng district, “BX” is Boxing county, “DY” is Dongying city, “Longju” is the town of
Longju, “DZ” is the town of Dingzhuang, “BZ” is Binzhou city, and “GR” is Guangrao county.

(Fig. 5). Locations of the selected points are depicted in Fig. 3. Based on Fig. 5, the cumulative
subsidence around Guangrao, Binzhou, Dingzhuang, Longju, Dongying, Boxing, and Dongcheng
have amounted to 269, 179, 186, 121, 108, 75, and 69 mm, respectively, within nearly 4 years,
and the annual average subsidence rates are ∼74, 49, 51, 33, 30, 21, and 19 mm∕year, respec-
tively. As expected, the reference area is extremely stable.

4.3 Spatial and Temporal Correlation between Land Subsidence and


Groundwater
An undeveloped aquifer system is in balance between recharge and discharge. The reduction of
fluid pressure in the pores and cracks of aquifer systems, especially in unconsolidated rocks, is
inevitably accompanied by some deformation of the aquifer system.1 Thereby, land subsidence
and other environmental problems can occur when groundwater is removed by pumping or
drainage due to the disruption of the balance and stable state.
With the rapid population, social, and economic growths since the delta became the second
largest petroleum production base in China in the 1960s,51 a large amount of water has been
demanded for the domestic water supply and the development of industry and agriculture. From
1986 to 2009, groundwater extraction in the Delta reached more than 90 million m3 ∕year.52
Excessive withdrawal of the groundwater decreases the water recharge and can cause a ground-
water depression cone to form. Consequently, the groundwater descent funnel in the Yellow
River delta increased to four large funnels in 2001, as shown in Fig. 6, from only two small
funnels in 1980.52
Figure 6 depicts the groundwater descent funnel map of the Yellow River Delta in 2009,
which has an unanticipated high spatial correlation with the land subsidence regions, especially
in the groundwater source areas of Binzhou, Guangrao, Dongying, and Dingzhuang (combined
with Fig. 3). Interestingly, all the regional highest subsidences are observed and centered at the
groundwater descent funnels, further suggesting a causal relationship between rapid subsidence
and the withdrawal of groundwater.
The Longju and Dongying subsidence regions are in the same ground water descent funnel
and the subsidence rates are nearly identical. The former is located in the irrigation region where
more than 90% of the groundwater pumping is for agricultural irrigation, whereas the latter is in
an urban region where most of the groundwater extraction is for the domestic water supply and
industrial uses, both of which have similar groundwater discharge each year.53 Although the edge

Journal of Applied Remote Sensing 096019-8 Vol. 9, 2015


Liu, Huang, and Dong: Large-area land subsidence monitoring and mechanism. . .

Fig. 6 Groundwater exploitation funnel in 2009 (SPLGEI). The map extent is almost the same as
in Fig. 3(a).

of the Dongying oil field lies ∼5 km from the Dongying subsidence region, we suggest that the
subsidence is associated with the groundwater. The oil field unitization, pressure maintenance as
a result of water injection, and oil extraction from a depth of 3 to 5 km would produce little
influence on the surface.11,37,54 Moreover, subsidence is centered over the groundwater descent
funnel, not the oil field.
As shown in Fig. 6, the Binzhou and Guangrao subsidence regions are located in two separate
groundwater descent funnels, and the groundwater level of Binzhou is ∼70 m lower than that of
Guangrao. Though Guangrao and Binzhou have similar annual groundwater discharge (32.24
and 38.34 million m3 ∕year, respectively49,55), the subsidence rate of Binzhou is much smaller
than Guangrao (Fig. 3). This may be attributed to different sediment characteristics of the strata
overlying the aquifer system. A soft soil layer develops fully in Guangrao and is primarily com-
posed of saturated or nearly saturated clay, silt, and mucky silt, with a high water content, low
density, large porosity ratio, large compression coefficient, and low bearing capacity.56 These
factors cause an easy compression after the removal of groundwater and lead to land subsidence.
Four profiles of the groundwater level along the lines of P1-P1’, P2-P2’, P3-P3’ and P4-P4’
in Fig. 6 were extracted, along with the InSAR displacement results of the year 2009 in the four
500-m-wide swath centers along the profiles (Fig. 7). The terrain profiles are also plotted in the
figure. The shape of the respective corresponding profiles generally agrees well except for a
significant discrepancy at 5 and 7 km for L2-L2’ and L4-L4’, respectively, in Fig. 7. The
differences reflect the influences of faults on subsidence, as in the analysis at the beginning
of Sec. 4. Moreover, all of the subsidence centers of each profile correspond to the respective
groundwater descent funnel and little subsidence is detected outside the groundwater descent
funnel regions. To obtain critical insight into the causal relationship between groundwater and
surface deformation, we collected the observational dynamic groundwater data from the
No. SJ49 groundwater monitoring site in the central east portion of the Guangrao groundwater
descent funnel (see Fig. 6 for locations), ∼3 km from the Guangrao (GR) deformation feature, to
analyze the temporal correlation (Fig. 8). Obviously, with a clear decrease in the water level, the
surface subsided dramatically. Furthermore, the correlation coefficient of the InSAR time-series
deformation and dynamic groundwater level, demonstrated in Fig. 9, is more than 0.9. The high
correlation coefficient implies that there is a significant positive temporal correlation between
subsidence and groundwater.

Journal of Applied Remote Sensing 096019-9 Vol. 9, 2015


Liu, Huang, and Dong: Large-area land subsidence monitoring and mechanism. . .

Fig. 7 Cross-section maps of the InSAR deformation along the swaths, corresponding to the
groundwater level and elevation along the profiles, as indicated in Fig. 3. The blue dots represent
the InSAR displacement, and the cyan and gray areas describe the groundwater and eleva-
tion, respectively. The land subsidence centers of each profile are specified by place name
(a) Guangrao, (b) Binzhou, (c) Dingzhuang, (d) Dongying.

Journal of Applied Remote Sensing 096019-10 Vol. 9, 2015


Liu, Huang, and Dong: Large-area land subsidence monitoring and mechanism. . .

Fig. 8 Temporal correlation of dynamic groundwater level (cyan area) and InSAR time-series
deformation (red dots with error bars).

Fig. 9 The linear regression analysis between groundwater level and deformation in Guangrao
county.

Evidently, the results in this section indicate that the land subsidence is affected by ground-
water, and the excessive exploitation of groundwater is one of the main causes for land sub-
sidence in the Yellow River Delta.

4.4 Correlation of Land Subsidence with Precipitation


It is apparent from the time-series deformations of the feature points illustrated in Fig. 5 that all
the deformations are nonlinear, and subsidences interestingly increase in spring, decrease in
summer, and the maximum seasonal discrepancy is more than 10 mm∕month. In general,
among all the factors that affect land subsidence, it seems possible to hypothesize that the dis-
crepancy can be attributed to the precipitation. Therefore, further studies should be conducted
with caution to investigate the influence of precipitation on seasonal deformation.
Here, we show the monthly precipitation data collected from the Guangrao meteorological
station (see Fig. 3 for the location), ∼7 km from the Guangrao (GR) deformation feature

Journal of Applied Remote Sensing 096019-11 Vol. 9, 2015


Liu, Huang, and Dong: Large-area land subsidence monitoring and mechanism. . .

Fig. 10 Correlation between the cumulative surface deformation of each year and the monthly
average precipitation recorded at the Guangrao meteorological station.

(Fig. 10). To analyze the temporal correlation between the seasonal deformation and the pre-
cipitation, the time-series results were normalized into 35 days and the annual cumulative defor-
mation was calculated. The correlation between the monthly precipitation and the annual
cumulative deformation at the Guangrao (GR) feature area is shown in Fig. 10. The subsidence
rate in summer was the smallest among the four seasons each year. In general, greater precipi-
tation yielded lower subsidence.
Interestingly, the response of surface deformation to precipitation is observed to be nonin-
stantaneous (Fig. 10). Instead, there is a time lag between the peak deformation rate and the peak
precipitation. At the Guangrao subsidence funnel, the lag is ∼1 to 2 months. Due to the low
temporal sampling of the Envisat/ASAR data, it is impossible to determine the time lag at a
finer temporal resolution. Owing to the ground swells quickly and small after watering,19
we ascribe the time lag to the change of ground water level after rainfall, and the lag time
to the petrophysical and hydrological properties. More broad research is also needed to deter-
mine these properties. Overall, these results indicate that precipitation is the most dominant fac-
tor influencing the seasonal deformation.

5 Concluding Remarks
Subsidence is a significant problem in the Yellow River Delta that has gone largely unrecognized
by the general public. The rapid land subsidence in the Yellow River Delta, in addition to the
widely distributed croplands and wetlands, crop growth, seasonal cultivation, and high level of
variable water vapor content in the coastal region, all pose considerable challenges to obtaining
reliable measurements of subsidence using conventional InSAR techniques. The SBAS InSAR
technique has been successfully applied to MTI images in this study to overcome the limitations
of conventional InSAR, to precisely detect and measure land subsidence, and to obtain a time-
series deformation pattern of each subsidence region in the Yellow River Delta from 2007 to
2010. The accuracy of InSAR in the Yellow River Delta can reach ∼1 cm based on the standard
deviation and spirit leveling data.
The results from the SBAS analysis show the patterns and magnitudes as well as the non-
linear behavior of land subsidence during 2007 to 2010. The major subsidence regions, with
rates ranging from 20 to more than 70 mm∕year, are mainly concentrated in the groundwater
source areas and comprise a total of 800 km2 . The Dongcheng (DC) district has become a land
subsidence center due to construction. Ground deformation in the Yellow River Delta is affected
by groundwater and influenced by fault alignment, and the seasonal deformation is primarily
ascribed to local precipitation. In addition, excessive exploitation of groundwater is the main
cause of land subsidence in the delta. The time-series land subsidence is correlated with

Journal of Applied Remote Sensing 096019-12 Vol. 9, 2015


Liu, Huang, and Dong: Large-area land subsidence monitoring and mechanism. . .

local precipitation with a time lag of ∼1 to 2 months between the peak of the precipitation and the
minimum deformation rate. Such lags are related to the ground water level change time since the
onset of precipitation and are influenced by the petrophysical and hydrological properties.
This study demonstrates for the first time the relative contribution of groundwater and pre-
cipitation to subsidence in the Yellow River Delta and provides critical insight into the primary
drivers of subsidence in the delta. Natural sediment consolidation, the primary natural cause of
subsidence in the world’s deltas, can only account for <5 mm∕year of subsidence in the study
area during the SAR monitoring time.56,57 Because of the unique formation history and the
unique natural, geographical, and cultural environment of the Yellow River Delta, groundwater
exploitation, the primary anthropogenic cause, is demonstrated to be the most dominant factor
among the complex reasons causing land subsidence in the study area. In addition, local pre-
cipitation greatly influences land subsidence. All of this information can be used to develop
targeted interventions aimed at the planning of key policy priorities for controlling the anthropo-
genic factors and better managing groundwater supplies. However, the petrophysical and hydro-
logical properties of the aquifer system still have to be analyzed to further study the nature and
magnitude of seasonal subsidence. More frequent SAR images acquired over a relatively longer
time period are still needed to address the impact of pumping-induced subsidence on the long-
term subsidence.

Acknowledgments
We would like to thank ESA for providing Envisat data under the Category 1 project C1F.19709.
We also thank JPL for ROI_PAC, DEOS for DORIS and Andy Hooper for StaMPS. The
meteorological data were provided by the China Meteorological Data Sharing Service
System (http://cdc.cma.gov.cn/dataSetDetailed.do?changeFlag=detai&titleName=SURF_CLI_
CHN_MUL_DAY_V3.0). This work was funded by the National Natural Science Founda-
tion of China, Grant No. 41276082, 41306190, and the Knowledge Innovative Program of
Chinese Academy of Sciences, Grant No. KZCX2-EW-207. We are grateful for the helpful cor-
rections and suggestions made by Haibo Bi and two anonymous reviewers.

References
1. D. Galloway, D. R. Jones, and S. E. Ingebritsen, Land Subsidence in the United States, US
Geological Survey Reston, Virginia (1999).
2. H. Fan, H. J. Huang, and T. Zeng, “Impacts of anthropogenic activity on the recent evolution
of the Huanghe (Yellow) River Delta,” J. Coast. Res. 22(4), 919–929 (2006).
3. J. P. M. Syvitski et al., “Sinking deltas due to human activities,” Nat. Geosci. 2(10), 681–
686 (2009).
4. R. J. Nicholls and A. Cazenave, “Sea-level rise and its impact on coastal zones,” Science
328(5985), 1517–1520 (2010).
5. T. H. Dixon et al., “Subsidence and flooding in New Orleans,” Nature 441(7093), 587–588
(2006).
6. Y. Saito et al., “Shrinking megadeltas in Asia: Sea-level rise and sediment reduction impacts
from case study of the Chao Phraya Delta,” Inprint Newslett. IGBP/IHDP Land Ocean
Interact. Coast. Zone 2, 3–9 (2007).
7. J. P. Ericson et al., “Effective sea-level rise and deltas: causes of change and human dimen-
sion implications,” Global Planet. Change 50(1–2), 63–82 (2006).
8. J. P. M. Syvitski and Y. Saito, “Morphodynamics of deltas under the influence of humans,”
Global Planet. Change 57(3–4), 261–282 (2007).
9. L. R. Huang, H. M. Hu, and G. H. Yang, “Sea level change along the western and southern
coast of Bohai Sea and recent crustal vertical movement in adjacent area,” Crustal Deform.
Earthquake 11(1), 1–9 (1991) (in Chinese with English abstract).
10. M. E. Ren, “Relative sea level rise in Huanghe, Changjiang and Zhujiang delta over the last
30 years predication for the next 40 years (2030),” Acta Geogr. Sin. 48(5), 385–393 (1993)
(in Chinese with English abstract).

Journal of Applied Remote Sensing 096019-13 Vol. 9, 2015


Liu, Huang, and Dong: Large-area land subsidence monitoring and mechanism. . .

11. S. Higgins et al., “Land subsidence at aquaculture facilities in the Yellow River delta,
China,” Geophys. Res. Lett. 40(15), 3898–3902 (2013).
12. Y. Liu and H. J. Huang, “Characterization and mechanism of regional land subsidence in the
Yellow River Delta, China,” Nat. Hazards 68, 687–709 (2013).
13. J. Z. Zhang, H. J. Huang, and H. B. Bi, “Land subsidence in the modern Yellow River Delta
based on InSAR time series analysis,” Nat. Hazards 75(3), 2385–2397 (2014).
14. P. Liu et al., “Land subsidence over oilfields in the Yellow River Delta,” Remote Sens. 7(2),
1540–1564 (2015).
15. B. Song, D. Wang, and J. Wang, “Dongying ground subsidence monitoring,” Survey. Mapp.
Geol. Miner. Resour. 2(1), 34–36 (2004) (in Chinese with English abstract).
16. G. W. Liu et al., “Effective factors of land subsidence in the Yellow River Delta,” Mar. Sci.
35(8), 43–50 (2011) (in Chinese with English abstract).
17. J. W. Borchers and M. Carpenter, “Land subsidence from groundwater use in California,”
California Water Foundation Full Report of Findings/April 2014, p. 151 (2014).
18. D. Massonnet and K. L. Feigl, “Radar interferometry and its application to changes in the
Earth’s surface,” Rev. Geophys. 36(4), 441–500 (1998).
19. A. K. Gabriel, R. M. Goldstein, and H. A. Zebker, “Mapping small elevation changes
over large areas: differential radar interferometry,” J. Geophys. Res. 94(B7), 9183–9191
(1989).
20. E. Sansosti et al., “Space–borne radar interferometry techniques for the generation of defor-
mation time series: an advanced tool for Earth’s surface displacement analysis,” Geophys.
Res. Lett. 37(20), L20305 (2010).
21. A. Hooper et al., “Recent advances in SAR interferometry time series analysis for meas-
uring crustal deformation,” Tectonophysics 514, 1–13 (2012).
22. J. Hoffmann et al., “Seasonal subsidence and rebound in Las Vegas Valley, Nevada, observed
by synthetic aperture radar interferometry,” Water Resour. Res. 37(6), 1551–1566 (2001).
23. D. A. Schmidt and R. Bürgmann, “Time–dependent land uplift and subsidence in the Santa
Clara valley, California, from a large interferometric synthetic aperture radar data set,”
J. Geophys. Res. 108(B9), 2416 (2003).
24. J. W. Bell et al., “Permanent scatterer InSAR reveals seasonal and long–term aquifer–
system response to groundwater pumping and artificial recharge,” Water Resour. Res.
44(2), W02407 (2008).
25. M. H. Aly et al., “Permanent Scatterer investigation of land subsidence in Greater Cairo,
Egypt,” Geophys. J. Int. 178(3), 1238–1245 (2009).
26. B. A. Wisely and D. Schmidt, “Deciphering vertical deformation and poroelastic parameters
in a tectonically active fault-bound aquifer using InSAR and well level data, San Bernardino
basin, California,” Geophys. J. Int. 181(3), 1185–1200 (2010).
27. S. Mazzotti et al., “Impact of anthropogenic subsidence on relative sea-level rise in the
Fraser River delta,” Geology 37(9), 771–774 (2009).
28. P. Teatini, L. Tosi, and T. Strozzi, “Quantitative evidence that compaction of Holocene sedi-
ments drives the present land subsidence of the Po Delta, Italy,” J. Geophys. Res. 116(B8),
B08407 (2011).
29. E. Chaussard et al., “Sinking cities in Indonesia: ALOS PALSAR detects rapid subsidence
due to groundwater and gas extraction,” Remote Sens. Environ. 128, 150–161 (2013).
30. Z. Lu and W. R. Danskin, “InSAR analysis of natural recharge to define structure of a
ground-water basin, San Bernardino, California,” Geophys. Res. Lett. 28(13), 2661–2664
(2001).
31. L. Y. Ji et al., “Episodic deformation at Changbaishan Tianchi volcano, northeast China
during 2004 to 2010, observed by persistent scatterer interferometric synthetic aperture
radar,” J. Appl. Remote Sens. 7(1), 073499 (2013).
32. R. F. Hanssen, Radar Interferometry: Data Interpretation and Error Analysis, Kluwer
Academic Publishers, The Netherlands (2001).
33. H. Wang et al., “InSAR reveals coastal subsidence in the Pearl River Delta, China,”
Geophys. J. Int. 191, 1119–1128 (2012).
34. B. M. Kampes, Radar Interferometry: Persistent Scatterer Technique, Springer, The
Netherlands (2006).

Journal of Applied Remote Sensing 096019-14 Vol. 9, 2015


Liu, Huang, and Dong: Large-area land subsidence monitoring and mechanism. . .

35. H. A. Zebker and J. Villasenor, “Decorrelation in interferometric radar echoes,” IEEE Trans.
Geosci. Remote Sens. 30(5), 950–959 (1992).
36. A. Hooper, P. Segall, and H. Zebker, “Persistent scatterer interferometric synthetic aperture
radar for crustal deformation analysis, with application to Volcán Alcedo, Galápagos,”
J. Geophys. Res. 112(B7), B07407 (2007).
37. X. W. Guo et al., “Oil generation as the dominant overpressure mechanism in the Cenozoic
Dongying depression, Bohai Bay Basin, China,” Am. Assoc. Pet. Geol. Bull. 94(12), 1859–
1881 (2010).
38. S. N. Li et al., “Variations of groundwater depth in Yellow River Delta in recent two dec-
ades,” Progr. Geogr. 27(5), 49–56 (2008) (in Chinese with English abstract).
39. Geological Comprehensive Investigation Report of Hydrological Engineering Environment
of the Yellow River Delta, Shandong Provincial Lubei Geo-engineering Exploration
Institute, Dezhou, Shandong province, China (1998) (in Chinese).
40. X. J. Yao et al., “Formation and evolution of ground fresh water (blackish water) in the Yellow
River Delta,” Acta Geosci. Sin. 23(4), 375–378 (2002) (in Chinese with English abstract).
41. H. Fan and H. J. Huang, “Response of coastal marine eco-environment to river fluxes into
the sea: a case study of the Huanghe (Yellow) River mouth and adjacent waters,” Mar.
Environ. Res. 65(5), 378–387 (2008).
42. P. Berardino et al., “A new algorithm for surface deformation monitoring based on small
baseline differential SAR interferograms,” IEEE Trans. Geosci. Remote Sens. 40(11),
2375–2383 (2002).
43. R. Lanari et al., “A small-baseline approach for investigating deformations on full-resolu-
tion differential SAR interferograms,” IEEE Trans. Geosci. Remote Sens. 42(7), 1377–1386
(2004).
44. R. Lanari et al., “An overview of the small baseline subset algorithm: a DInSAR technique
for surface deformation analysis,” Pure Appl. Geophys. 164(4), 637–661 (2007).
45. O. Mora, J. J. Mallorqui, and A. Broquetas, “Linear and nonlinear terrain deformation maps
from a reduced set of interferometric SAR images,” IEEE Trans. Geosci. Remote Sens.
41(10), 2243–2253 (2003).
46. T. R. Lauknes, H. A. Zebker, and Y. Larsen, “InSAR deformation time series using an L1-
norm small-baseline approach,” IEEE Trans. Geosci. Remote Sens. 49(1), 536–546 (2011).
47. A. Hooper, D. Bekaert, and K. Spaans, StaMPS/MTI Manual, School of Earth and
Environment, University of Leeds, United Kingdom (2013).
48. A. Hooper, “A statistical-cost approach to unwrapping the phase of InSAR time series,” in
Proc. of Int. Workshop on ERS SAR Interferometry, Frascati, Italy (2010).
49. Y. Liu et al., “A comprehensive analysis of factors influencing on land subsidence in the
Yellow River Delta,” J. Eng. Geol. 22(Suppl. 1), 62–69 (2014) (in Chinese with English
abstract).
50. H. Y. Li, X. Z. Zhao, and L. H. Zhang, “Countermeasure analysis of dynamic groundwater
of Dongying city,” Shandong Water Resour. 2011(7), 51–52 (2011) (in Chinese).
51. M. E. Ren, “Effect of sea level rise and land subsidence on the Huanghe River Delta,” Sci.
Geogr. Sinica 10(1), 48–57 (1990) (in Chinese with English abstract).
52. X. Z. Zhao, “Overexploitation status and recharge measures of groundwater in Dongying
city,” Shandong Water Resour. 2013(9), 29–30 (2013) (in Chinese).
53. Water conservancy annals of Dongying compilation and editing committee, “Water con-
servancy annals of Dongying,” http://www.dongying.gov.cn/html/dysslz/ (2009).
54. J. Geertsma, “Land subsidence above compacting oil and gas reservoirs,” J. Pet. Technol.
25(6), 734–744 (1973).
55. Z. Ma and Y. H. Duan, “Sustainable utilization of groundwater resource in the Lubei plain
of Shandong Province,” Hydrogeol. Eng. Geol. 32(2), 1–10 (2005) (in Chinese with English
abstract).
56. C. X. Shi et al., “Land subsidence as a result of sediment consolidation in the Yellow River
Delta,” J. Coast. Res. 23(1), 173–181 (2007).
57. J. Y. Tan, H. J. Huang, and Y. X. Liu, “Estimation of sediment compaction and its con-
tribution to land subsidence in the Yellow River Delta,” Mar. Geol. Quat. Geol. 31(5),
33–38 (2014) (in Chinese with English abstract).

Journal of Applied Remote Sensing 096019-15 Vol. 9, 2015


Liu, Huang, and Dong: Large-area land subsidence monitoring and mechanism. . .

Yilin Liu received his master’s degree in cartography and geographic information engineering
from Chang’an University, China, in 2013. Currently, he is a PhD student at the Institute of
Oceanology, Chinese Academy of Sciences (CAS). His research interests include different
InSAR methods, the applications of InSAR on natural hazards monitoring, and mechanism
researching.

Haijun Huang is a professor of Institute of Oceanology, CAS. He received his PhD in marine
geology from Institute of Oceanology, CAS, in 1991. He worked as a visitor at Sydney
University, Australia, Coastal Research Center from January to December 1997. He is engaged
in marine geology studies with Geographical Information System (GIS) and remote sensing
techniques and is particularly interested in the study related to land subsidence and coastal
stability.

Jinfa Dong received his master’s degree from the Institute of Geographic Sciences and Natural
Resources Research, CAS, in 2013. He is currently a software engineer at Tianjin Navigation
Instruments Institute. His research interests concern the applications of remote sensing and GIS
on environmental resources.

Journal of Applied Remote Sensing 096019-16 Vol. 9, 2015

Das könnte Ihnen auch gefallen