You are on page 1of 8

GEOPHYSICAL RESEARCH LETTERS, VOL. ???, XXXX, DOI:10.

1029/,

In-Situ Measurement of the Hydraulic Diffusivity of the Active


Chelungpu Fault, Taiwan
M.L. Doan,1,3 E.E. Brodsky,1 Y. Kano,2 and K.F. Ma3

Hydraulic diffusivity controls fluid pressure and hence af- 1. Introduction


fects normal effective stress during rupture. Models sug-
The hydraulic diffusivity of the damage zone is a funda-
gest a spectacular example of fluid pressurization during the
mental property of earthquake rupture because it controls
Mw =7.6 1999 Chichi earthquake when pressurization may the fluid overpressure that can be maintained on a fault.
have reduced high-frequency shaking in the regions of large In between earthquakes, this fluid pressure can control fail-
slip if the fault was sufficiently sealed. We investigate in ure [Sibson et al., 1975]. During earthquakes, Rice [2006]
situ hydraulic diffusivity through a cross-hole experiment. and Andrews [2002] both calculated that the length scale
We find a diffusivity of D=(7 1) 105 m2 /s, a low value over which slip weakening occurs depends critically on the
compatible with pressurization of the Chelungpu fault dur- hydraulic diffusivity.
ing the earthquake. In most poroelastic media, the storativ- Permeability is the most variable part of hydraulic diffu-
sivity (see description of variables in appendix A). Labora-
ity S lies between 107 and 105 , so that the permeability tory measurements of core samples give a lower bound on the
k along the fault zone is probably between 1018 m2 and diffusivity [Lockner et al., 2005; Tanikawa et al., 2005; Chen
1016 m2 . This permeability is at most one hundred times et al., 2005], but permeability is notoriously scale dependent
larger than the value obtained on core samples from the host [Manning and Ingebritsen, 1999]. The most important frac-
rock. The fault zone is overpresssurized by 0.06 to 6 MPa, tures may be on scales much larger than a core sample and
which is between 0.2% and 20% of the lithostatic pressure. therefore it is necessary to measure the properties in situ.
Despite the fundamental importance of hydraulic diffu-
sivity, it has never been successfully measured in situ on an
active large-scale fault. Hydraulic tests in deep boreholes
intersecting the Nojima fault were attempted by Kitagawa
et al. [1999, 2002] but they were tapping flow in the hanging
1 Earth Science Department, University of Santa Cruz, wall at least 50 m away from the fault core and therefore at
the diffuse end of the damaged zone. In the present study,
1156, Santa Cruz, California, USA
2 Disaster Prevention Research Institute, Kyoto we study a pair of boreholes intersecting the Chelungpu fault
University, Uji, Japan
that are perforated closer to the fault core.
3 National Central University, Chung-li, 320-54, Taiwan Water was pumped out of hole A (Fig. 1). The resulting
head change propagated along the fault zone to produce a
hydraulic anomaly in hole B. We report the results of the
Copyright 2006 by the American Geophysical Union. experiment. After effectively modeling leaks in the casing,
0094-8276/06/$5.00 we extract a value of diffusivity for the fault damage zone
of 7 105 m2 /s, which is sufficiently low to confine pres-
surized fluid during an earthquake.
Hole A Hole B
pumping hole monitoring hole
2. Hydraulic tests on the Chelungpu fault
  

 
 

 
  
 
 1108.0m


  2.1. Boreholes of the Taiwan Chelungpu-fault Drilling
1111m  1109.2m


  Project

 
  1138m The pair of boreholes of this experiment are separated by
  

  40 m. Both holes are fully cased and cemented in their an-
  1138.0m

  nuli. Both are perforated only near their intersection with
  1139.2m

  ~1m
 40m
the Chelungpu fault. Hole A is perforated above the fault;
hole B directly below (Fig. 1).
Hole A intersects the fault at a depth of 1111 m within
the silty shales of the Chinshui formation. The slip is con-






centrated within a 12 cm thin layer of fine-grained clayish




Perforation on casing Chelungpu fault damaged zone
Cement Chelungpu fault slip zone black material, which is distinct from the nearby 30 cm
Leakage Perforation depth thick layer of grayish gouge. Hole B intersects regions of
black material at 1137.5 m and 1138.0 m that are also inter-
preted as slip zones [Ma et al., 2005]. In both holes, the
Figure 1. Configuration of the cross-hole hydraulic test fault core is surrounded by breccia and fractured rocks that
on the Chelungpu boreholes. The two holes are separated form a 1 m wide damage zone.
by 40 m and perforated near the fault with a density of 4
shots per foot. Blue thick numbers indicate the top and 2.2. Pumping test
bottom depths of the perforations. Perforation location We pumped water out of hole A on Nov. 18, 2005 and
is accurate to within 0.5 m. This schematic is not true then continuously recorded the water level in both holes for
scale.
1
X-2 DOAN ET AL.: IN-SITU HYDRAULIC PROPERTIES OF THE CHELUNGPU FAULT

the following 3 months. Fig. 2 shows the evolution of water parameters, which are storativity S=106 and transmissiv-
level in hole A during and after the pumping. The water ity T =107 m2 /s. The model of Cooper et al. [1967] does
level in the pumping well recovers in one week and then not take into account the overpressure in the aquifer so that
stays at zero when it reaches the wellhead. the computed curve was shifted by the overpressure in the
Fig. 3 shows the evolution of water level in hole B. Both aquifer, which is 0.3 MPa.
before and after pumping, the water level continuously de- Even before the perforation, hole A was artesian with a
flow rate of 106 m3 /s due to a leaky casing. Prior to per-
creased due to leaks in the casing resulting in a loss of more foration, a pumping test was done on hole A (Chia-Shyun
than 18 m in 3 months. This large loss from leakage ob- Chen, pers. comm.). The results suggest that the leaks tap
scures the more subtle drop in water level that is created by an aquifer with an overpressurized hydraulic head of 30 -
the pumping in hole A. Therefore, we need to model and re- 50 m and a transmissivity close to 107 m2 /s. These values
move the effects of the leaks in hole B to detect the transient are similar to those retrieved in Fig. 2, indicating that the
induced by the pumping. recovery in hole A is dominated by the leakage rather than
by the perforated fault zone.
Although the model of Cooper et al. [1967] gives satis-
3. Analysis factory results, it does not take into account the constant
water level when water reaches the surface. The input signal
To recover the hydraulic properties of the fault zone, we in hole A was approximated by the exponential fit of Fig. 2.
analyze how the monitoring well (hole B) responds to the This exponential approximation will be used in section 3.3
pumping well (hole A). We: (1) model the recovery of wa- to fit the anomaly in hole B.
ter level in hole A, which is the cause of the anomaly in
3.2. Removal of the first-order trend induced by
hole B, (2) remove the effects of the leaks in hole B and (3)
compare the remaining anomaly in hole B with a prediction leaks in hole B
based on the variations in water level in hole A to recover Fig. 3 shows that the water level in hole B is dominated
the hydraulic properties of the fault zone. by a continuous decrease in water level. This decrease ex-
isted prior to the perforation in hole B and is due to leakage
3.1. Modeling the pumping hole in the casing. To extract the disturbance due to the pump-
ing in hole A, an accurate approximation of this leakage is
The sudden change in water level from the pumping in needed.
hole A disturbed the aquifers tapped by the well. For a Let us suppose a quasi-permanent radial flow inside the
single isotropic poroelastic aquifer with no lateral bound- aquifer tapped by the leakage. The flow Q through the leak
aries, Cooper et al. [1967] computed a now-standard solu- is given by the Thiem equation:
tion (Eq. (A5)). Fig. 2 displays the model with the best-fit
2T
Q = (hw hf ) (1)
ln (R /rb )

Hole A where R is the radius of influence beyond which the pres-


100 sure in the formation is undisturbed, rb is the borehole ra-

0
Hole B
100
10 data
exponential fit with h=69.7m, =270d
water level (m)

200 12 exponential fit with h=54.7m, =200d

14
300
water level (m)

16
400
18
data
Pumping time

500 7 2 6 20
Cooper et al. T=10 m /s, S=10 , h=30m
exponential fit
600 22
0 5 10 15 20
Time since pumping (day) 24

26
Figure 2. Recovery of the water level in the pump-
ing hole through time. We lowered the water level of Dec Jan Feb
hole A by 400 m. (There was a small transient that
dropped the level to -500 m while the pump was de-
ployed). The curve fits the evolution predicted by Cooper Figure 3. Evolution of the water level in hole B rela-
et al. [1967] with the transmissivity T =107 m2 /s and tive to the wellhead. It is compared with the exponential
the storativity S=106 (blue dashed curve), provided we solutions computed with Eq. (2). We present here the
take into account the overpressure of the leaky aquifer two extreme sets of parameters =200 days, h =54.7 m
(about 0.3 MPa, equivalent to 30 m of water). This the- (top red dot-dashed line) and =270 days, h =69.7 m
oretical result does not take into account the fixed level (bottom green dashed line), that delineate a range of pos-
of head at the surface. The red dot-dashed curve depicts sible fitting exponentials (shaded area). The maximum
the exponential function used to compute analytically the departure is 70 cm, over a total change of 18 m. The error
expected response of hole B. is thus less than 3.5% over 3 months.
DOAN ET AL.: IN-SITU HYDRAULIC PROPERTIES OF THE CHELUNGPU FAULT X-3

dius (7.8 cm for both holes A and B), T is the transmissivity the perforations of one hole and the leak of the other hole,
of the formation, hw is the hydraulic head in the well and (3) a flow between the perforations of the two holes.
hf is the hydraulic head in the formation. Combined with The observed signal might come from the propagation of
the conservation of mass of fluid in the borehole, the water the pressure disturbance between the leaks of the two holes.
level in hole B is described by an exponential function: However, the leakage of hole A, where water goes into the
borehole, differs from the leakage in hole B, where water
hw = h + (hw (t0 ) h ) e(tt0 )/ (2) flows out of the borehole. Moreover, the value of diffusivity
found from the cross-hole experiment is much smaller than
where = rc2 / ln(R
2T
/rb )
. Notice that the value of hw does the one derived for the leakage in hole A. These two obser-
not depend on the choice of t0 as hw (t0 ) is adjusted appropri- vations suggest that the aquifer tapped by the leaks in hole
ately. The leakage is then characterized by two parameters A is not connected to the leak in hole B.
only: the decay time and the far-field hydraulic head h . The observed signal might also be due to a connection
This is as complex a model as is warranted by the available between the leak of one hole to the perforations of the other
data. hole. In hole A, a cement bonding log shows that the cement
We calibrate the model of Eq. (2) using different sub- in the annular interval between the casing and the borehole
sequences of the data of November 2005 in order to find a wall is missing at the fault zone depth but also reveals that
range of possible values. Using intervals of 6-8 days dura- cement plugs of good quality exist over 10 m in the annu-
tion starting at times between Nov. 15 and Nov. 26 yields lar interval at 1000 m and 1040 m. The perforations in hole
a range of fit parameters given by the series of curves in the A are therefore decoupled from the permeable Kueichulin
shaded portion of Fig. 3. Even though this simple, steady- formation. No such log exists for hole B. However, if the
state model should only be valid for short times, it fits the cement were very poor all along the casing on hole B, the
observed data quite well for 3 months after the pumping. perforations in hole B would have allowed water from the
Fig. 4 shows the remaining signal in hole B after the Cholan formation to enter the borehole with a rate similar
leakage modeled by Eq. (2) is removed. The first-order ex- to the one observed in hole A. This is not what have been
ponential trend seen in Fig. 3 explains the data until 12 observed.
days after the pumping in hole A. For the entire probable The most plausible hypothesis is a flow between the two
range of parameters, the anomaly in hole B is in the shaded perforations of the borehole located at the fault zone. Core
region of Fig. 4. In all cases, the anomaly begins suddenly studies [Lockner et al., 2005; Tanikawa et al., 2005] suggest
12 days after the pumping in hole A. that the fault zone has the permeability pattern described by
Caine et al. [1996]: a fault core less permeable than the host
3.3. Hydraulic diffusivity from the hydraulic anomaly
rock, bounded by a damaged zone of enhanced permeabil-
recorded in the observation well ity. Since the Chinshui shale is a relatively low permeability
We now use our fit of the water level in hole A to predict
the anomaly in the corrected water level in hole B.
The anomaly in hole B cannot be explained with models
involving a single aquifer that both refills hole A and empties 0
D=7x105 m2/s
hole B. For instance, the direct application of the standard 0.1 5 2
D=6x10 m /s
model of Cooper et al. [1967] gives a best fitting curve with a
Water level exponential fit (m)

maximum amplitude of 4 m, which is five times larger than 0.2 5 2


D=8x10 m /s
observed. Also, the observations in section 3.1 indicate that 0.3
hole A is predominantly filled by an aquifer that is separate
from the perforated fault zone. 0.4
Instead we use a Green function solution that propa- 0.5
gates the anomaly from hole A to hole B through a separate
aquifer than the one responsible for the recovery of hole A 0.6
(appendix A). The model has a single parameter: the hy- 0.7 data corrected with h =54.7m, =200d
draulic diffusivity of the aquifer.
data corrected with h=61.1m, =230d
Fig. 4 shows that the modified model fits the observed 0.8
data corrected with h=69.7m, =270d
anomaly with the hydraulic diffusivity D=7 105 m2 /s. 0.9 theoretical model, modelled pumping recovery
The input signal used for the computation is the exponen- numerical simulation, observed pumping recovery
tial approximation of the recovery in hole A (Fig. 2). This 1
0 10 20 30 40 50 60 70 80 90
approximation has a slower recovery than the observed sig- Time since pumping (day)
nal and the calculated anomaly in hole B is expected to be
slightly larger than observed. Figure 4. Pumping from hole A observed in hole B.
The prediction is refined with numerical methods in order We removed the fit of Fig. 3 for the range of parame-
to use directly the time series of the water level data of hole ters described in section 3.2. The extreme parameters are
A of Fig. 2 instead of its exponential approximation. Our
=200 days, h =54.7 m and =270 days, h =69.7 m.
numerical solution was obtained by the direct implementa-
tion of pressure diffusion equation for a 2D radial flow with In all cases, a residual anomaly begins 12 days after
the finite elements method software Comsol3.2. With solu- pumping. Appendix A explains the observed anomaly as
tions using both modeled and observed recoveries in hole A, the propagation of the disturbance along the fault zone
the best fit is obtained with D=(7 1) 105 m2 /s. with a hydraulic diffusivity D=7105 m2 /s (bold lines),
using two different methods: (1) an analytical method
based on the exponential fit of the recovery data in hole
4. Discussion A of Fig. 2, and (2) a numerical simulation with the true
4.1. Identification of the tested aquifer data from hole A. Modeled curves for D=6 105 m2 /s
(upper thin lines) and D=8105 m2 /s (lower thin lines)
We interpreted the data as the propagation of the pres- provide estimates of the accuracy of our value of hy-
sure front induced by pumping within an aquifer. But which draulic diffusivity. A set of intermediate parameters gives
aquifer? Fig 1 suggests three main options: (1) a flow from
a candidate experimental curve that would be fitted by
the leak of hole A to the leak of hole B, (2) a flow between
the expected curves for D=7 105 m2 /s.
X-4 DOAN ET AL.: IN-SITU HYDRAULIC PROPERTIES OF THE CHELUNGPU FAULT

rock, it is probable that the easiest path between the per- By extrapolating storativity values of core samples, the
forations is the zone of highest permeability: the damaged permeability is constrained to be at most 1016 m2 , one hun-
zone of the Chelungpu fault. Because of the weak cement in dred times the value obtained for unfractured core sample
the annular interval of hole A, the impermeable fault core of the host rock. Core samples show extensive calcite crys-
is short-circuited and does not disturb the propagation of tallization in the breccia zone. A possible explanation of the
the head anomaly from the perforations of hole A (above small enhancement of permeability of the damaged zone is
the fault core) to the perforations of hole B (below the fault
core). that the recovery of the fault zone is well advanced only 6
years after the Chichi earthquake.
4.2. Estimation of the permeability of the fault Is the observed hydraulic diffusivity small enough to
damaged zone maintain the pore pressure inside the fault during an earth-
quake ? A naive interpretation suggests that fluid will leak
Most hydromechanical models studying the effect of fault in a time equal to L2 /D, where L is the thickness on the
zone deal with permeability. To extract this parameter zone of intensive shear during the earthquake, at most equal
from the hydraulic diffusivity, the specific storage Ss is to the thickness of the damaged zone. For our observed val-
needed. For the black material forming the fault core, Lock- ues of hydraulic diffusivity and a 1 m wide damaged zone,
ner et al. [2005] found a specific storage ranging from 1.3 to this leakage time is a couple hours long, much longer than
7 107 m1 . However, this highly deformed material dif- the duration of the earthquake. More careful formulation
fers from the silty shale forming the matrix of the damaged
zone. Chen et al. [2005] obtained a narrow range of specific by Andrews [2002] still allows pressurization to occur with
storages of 8 107 m1 to 3 106 m1 on core samples the observed diffusivity D. Thermal or hydrodynamic pres-
of host rock far from the fault. Large scale media should be surization during rupture is still a plausible mechanism.
more porous and less rigid than core samples as the specific Acknowledgments. We thank Prof. Ji-Hao Hung for his
storage of poroelastic media is given by Ss = ( Kf + K1s ) f g, information on the wells and Chang-Wei Tsao, Sonata Wu and
where is the porosity, f is the fluid density, g is the gravity Hsin-I Lin for their technical help. We are grateful to Andy Fisher
acceleration, Kf and Ks are the bulk modulus of water and for its thoughtful remarks and Joe Andrews for his constructive
solid matrix respectively. In the damaged zone, the poros- review. Doan and Brodsky were funded in part by the NSF. Kano
ity and the matrix compressibility are expected to be higher was partially funded by KAGI21.
than for the host rock. The specific storage might then range
from 107 m1 (laboratory values) to 105 m1 (large scale
value from the theoretical expression). As cores and logs
suggest, the densely damaged zone has a thickness b ' 1 m. References
The storativity lies in the range S=Ss b ' 107 105 .
The transmissivity T =D S is then deduced to be be- Andrews, D. J., A fluid constitutive relation accounting for ther-
tween 1011 m2 /s and 109 m2 /s. Transmissivity T is re- mal pressurization of pore fluid, J. Geophys. Res., 107 (B12),
lated to permeability (section A1). For water at 50 C and doi:10.1029/2002JB001,942, 2002.
b ' 1 m, the permeability is between 1018 m2 and 1016 m2 .
Caine, J. S., J. P. Evans, and C. B. Forster, Fault zone architec-
This permeability range is larger than the value 1019 m2 to
ture and permeability structure, Geology, 24 (1), 10251028,
1021 m2 obtained for core samples from the fault slip zone 1996.
[Lockner et al., 2005] but it is not much larger from that
of siltstone reference samples, ranging from 1019 m2 trans- Chen, N., W. Zhu, T. F. Wong, and S. Song, Hydromechanical
versely to bedding and 1018 m2 parallel to bedding [Chen behavior of country rock samples from the Taiwan Chelungpu
et al., 2005]. Drilling Project, Eos Trans. AGU, 86 (52), Fall Meet. Suppl.,
Abstract T51A1324, 2005.
4.3. Estimation of the overpressure of the fault
Cooper, H. H., J. D. Bredehoeft, and I. S. Papadopulos, Response
damaged zone of a finite-diameter well to an instantaneous charge of water,
Water Resour. Res., 3 (1), 263269, 1967.
The overpressure in the fault zone is estimated from the
change in flow rate in hole B. In hole B, the rate of water Kitagawa, Y., N. Koizumi, K. Notsu, and G. Igarashi, Water in-
level decrease changed after the perforation of the casing. jection experiments and discharge changes at the Nojima fault
The water level dropped by 24.5 cm/day prior to perforation in Awaji island, Japan, Geophys. Res. Lett, 26 (20), 31733176,
and by 22 cm/day after the perforation. This corresponds doi:10.1029/1999GL005,263, 1999.
to a change in flow rate Q=5 109 m3 /s. Assuming the
Kitagawa, Y., K. Fujimori, and N. Koizumi, Temporal change in
range in transmissivity inferred above and an influence ra-
permeability of the rock estimated from repeated water injec-
dius R =105 rb ' 8000 m in Eq. (1) (as it intervenes in- tion experiments near the Nojima fault in Awaji island, Japan,
side a logarithmic function, the result is poorly sensitive to Geophys. Res. Lett, 2002.
the choice of this parameter), the overpressure in the well
is computed to be between 0.06 MPa and 6 MPa (between Lockner, D. A., C. Morrow, S. Song, S. Tembe, and T. Wong,
0.2% and 20% of the lithostatic pressure). Permeability of whole core samples at Chelungpu fault, Tai-
wan TCDP scientific drillhole, Eos Trans. AGU, 86 (52), Fall
Meet. Suppl., Abstract T43D04, 2005.
5. Conclusion
Ma, K. F., C. Wang, J. H. Hung, S. Song, H. Tanaka, E. Yeh,
We found that the hydraulic diffusivity of the damaged and Y. Tsai, Dynamics of Chi-Chi earthquake rupture: Discov-
zone around the Chelungpu fault is close to 104 m2 /s. This ery from seismological modeling and Taiwan Chelungpu-fault
is to our knowledge the first in situ measurement of hy- Drilling Project (TCDP), Eos Trans. AGU, 86 (52), Fall Meet.
draulic diffusivity obtained for the damaged zone very near Suppl., Abstract T13A1353, 2005.
a major crustal fault, as similar experiments performed on
Manning, C. E., and S. E. Ingebritsen, Permeability of the con-
other fault drilling projects give rather properties of the
tinental crust: Implications of geothermal data and metamor-
aquifers surrounding the fault [Kitagawa et al., 1999, 2002].
phic systems, Rev. Geophys., 37 (1), 127150, 1999.
Our results suggest also that the damaged zone is mod-
erately overpressurized (at most 20% of the lithostatic pres- Rice, J. R., Heating and weakening of faults during earthquake
sure). A much smaller overpressure is also consistent with slip, J. Geophys. Res., 111 (B5), doi:10.1029/2005JB004006,
our data. 2006.
DOAN ET AL.: IN-SITU HYDRAULIC PROPERTIES OF THE CHELUNGPU FAULT X-5

Sibson, R. H., J. M. Moore, and A. H. Rankin, Seis- 95064, Santa Cruz, USA (mdoan@es.ucsc.edu)
mic pumping - a hydrothermal fluid transport mechanism, E.E. Brodsky, Earth Science Department, Earth & Marine Sci-
J.Geol.Soc.London, 1975. ences Building, University of Santa Cruz, 1156 High Street, CA
Tanikawa, W., T. Shimamoto, H. Noda, and H. Sone, Hydraulic
properties of Chelungpu, Shuangtung and Shuilikeng fault 95064, Santa Cruz, USA (ebrodsky@es.ucsc.edu)
zones and their implications for fault motion during 1999 Chi- Y. Kano, Disaster Prevention Research Institute, Kyoto Uni-
Chi earthquake, Eos Trans. AGU, 86 (52), Fall Meet. Suppl., versity, Gokasho, Uji, Kyoto 611-0011, Japan. (kano@eqh.dpri.kyoto-
Abstract T43D08, 2005. u.ac.jp)
M.L. Doan, Earth Science Department, Earth & Marine Sci- K.F. Ma, National Central University, Chung-li, 320-54, Tai-
ences Building, University of Santa Cruz, 1156 High Street, CA wan. (fong@rupture.gep.ncu.edu.tw)
In-Situ Measurement of the Hydraulic Diffusivity
of the Active Chelungpu Fault, Taiwan
APPENDIX
Doan, M.L., Brodsky, E.E., Kano, Y. and Ma, K.-F.
July 11, 2006

A Hydraulic propagation of an imposed draw-


down from a well
A.1 General solution
Let us solve the change in hydraulic head induced at a distance r of a borehole
by a change in its water level (hydraulic head is proportional to the fluid pressure
minus the hydrostatic pressure [Domenico and Schwartz, 1990]). The following
set of equations controls the evolution of the hydraulic head h in an isotropic
and homogeneous aquifer surrounding the well whose water level hw is imposed:
 2 
h 1 h h
D + = (A1)
r2 r r t
h(r ) = 0 (A2)
h(r = rw ) = hw (t) (A3)

The first equation is Darcys law for a 2D radial flow. The hydraulic diffu-
sivity D = T /S is the ratio of transmissivity T over storativity S. The trans-
gb
missivity T = f includes the permeability , the thickness of the formation
b and the dynamic viscosity of water , f is the density of the fluid and g the
gravity acceleration. The two other equations are the boundary conditions at
infinity and around the borehole wall.
By applying the Laplace Transform, the first equation is simplified into a
ordinary differential equation whose solution is CK K0 (q r)p+ CI I0 (q r) where
K0 and I0 are modified Bessel functions of order 0 and q = p/D in which p is
the Laplace parameter. The boundary conditions constrain the coefficients CK
and CI so that:
K0 (q r)
h(r, p) = hw (p) (A4)
K0 (q rb )

1
This equation shows that we can get no more information than the hydraulic
diffusivity D if the water level evolution in the pumping well is not controlled
by the studied aquifer.
The inverse Laplace transform of equation is computed in time space with
the Stehfest algorithm [Stehfest, 1970].

A.2 Application to the Taiwan boreholes


The above general solution has been applied to model the response of hole
B to two analytical expressions of the recovery in hole A: the one given by
the solution of Cooper et al. [1967], which is a special case of Eq. (A4), and a
simplified exponential recovery to take into account the leveling of water level
when it reaches the surface.

A.2.1 Modeling with the model of Cooper and al. (1967)


Cooper et al. [1967] predict the drawdown due to a sudden change in water level
in a pumping well of borehole radius rb and casing radius rc :
rb S H0 K0 (q r) K0 (q r)
hCBP (r, p) = 2 2
= hCBP (rb , p)
T q [rb q K0 (q rb ) + 2 S rb K1 (q rb )/rc ] K0 (q rb )
(A5)
Eq. (A5) depends on both the storativity S and the diffusivity D separately.
The solution of Cooper et al. [1967] is a special case of our general model, where
both the recovery in hole A and the propagation of the hydraulic anomaly are
controlled by the same parameters.
Eq. (A5) applied at r = rb is used to model the recovery of water level
in hole A with the hydraulic properties of the leakage of hole A (blue dashed
line of Fig. 2). These parameters are different from the ones controlling the
propagation of the anomaly along the fault as used in Eq. (A4).

A.2.2 Modeling with an exponential recovery of the water level in


hole A
Once the water level reaches the surface, it stays at zero. Fig. 2 shows that
the recovery can be still approximated by hw (t) = P0 et/ , whose Laplace
transform is
P0
hw (p) = . (A6)
p + 1/
The blue dashed line in Fig. 4 is Eq. (A4) with Eq. (A6) substituted in the
right-hand side.

References
Cooper, H. H., J. D. Bredehoeft, and I. S. Papadopulos, Response of a finite-
diameter well to an instantaneous charge of water, Water Resour. Res., 3 (1),
263269, 1967.

2
Domenico, P. A., and F. W. Schwartz, Physical and Chemical Hydrogeology,
John Wiley & Sons, New York, 1990.

Stehfest, H., Algorithm 368: Numerical inversion of Laplace transform, Comm.


of ACM, 13 (1), 4749, 1970.

You might also like