Gahalaut Et Al Reservoir Triggered Seismicity

8
7/23/2019 Gahalaut Et Al Reservoir Triggered Seismicity http://slidepdf.com/reader/full/gahalaut-et-al-reservoir-triggered-seismicity 1/8 A new case of reservoir triggered seismicity: Govind Ballav Pant reservoir (Rihand dam), central India Kalpna Gahalaut  a , V.K. Gahalaut  a, , M.R. Pandey  b  National Geophysical Research Institute, Uppal Road, Hyderabad 500 007, India  b  Department of Mines and Geology, Lainchur, Kathmandu, Nepal Received 21 November 2006; received in revised form 27 March 2007; accepted 9 April 2007 Available online 14 April 2007 Abstract We report here that seismicity near Govind Ballav Pant reservoir is strongly influenced by the reservoir operations. It is the second largest reservoir in India, which is built on Rihand river in the failed rift region of central India. Most of the earthquakes occurred during the high water stand in the reservoir with a time lag of about 1 month. We use the concept of coulomb stress change and use Green's function based approach to estimate stresses and pore pressure due to the reservoir load. We find that the reservoir increases coulomb stress on the nearby faults of the region that are favourably oriented for failure in predominantly reverse slip manner under the NNESSW compression and thus promotes failure. The above two factors make it an obvious, yet so far unreported case of reservoir triggered seismicity. © 2007 Elsevier B.V. All rights reserved.  Keywords:  Reserior triggered seismicity; Coulomb stress; Rihand reservoir; Peninsular India 1. Introduction Increases in the frequency of occurrence of earthquakes due to man's engineering activities have resulted from the reservoir impoundment, quarrying and mining, and fluid injection and extraction (McGarr and Simpson, 1997 ). However, earthquakes caused by reservoir impoundment are stronger than by any other engineering activities. So far globally about hundred sites of Reservoir Triggered Seismicity (RTS) have been reported which include at least eight sites from India, namely, Koyna, Warna, Bhatsa, Dhamni, Gandipet, Idukki, Mula and Sriramsagar (Gupta, 2002). According to the mechanism of RTS, reservoir load and/or induced porepressure due to reservoir operationsis the cause of triggering of earthquakes on critically stressed  pre-existing faults in the vicinity of the reservoirs (Simpson, 1986; Gupta, 1992; McGarr and Simpson, 1997). In this article, we report a new case of RTS associated with the Govind Ballav Pant reservoir, central India (Fig. 1). The reservoir is located on the Rihand river, a tributary of Son river, India. The 92 m high Rihand dam was built in 1962 and the reservoir is the second largest reservoir in India occupying an area of about 45×15 km 2 having a maximum storage capacity of 10.6 km 3 , and a capacity of electric power generation of 300 MW. In this article, we analyse the correlation between the times of high water levels in the reservoir and occurrence of the maximum number of earthquakes, and simulate the effect of the reservoir on the nearby earthquake causative faults to verify that these earthquakes were triggered by the reservoir. Tectonophysics 439 (2007) 171 178 www.elsevier.com/locate/tecto  Corresponding author.  E-mail address: [email protected] (V.K. Gahalaut). 0040-1951/$ - see front matter © 2007 Elsevier B.V. All rights reserved. doi:10.1016/j.tecto.2007.04.003

Transcript of Gahalaut Et Al Reservoir Triggered Seismicity

Page 1: Gahalaut Et Al Reservoir Triggered Seismicity

7/23/2019 Gahalaut Et Al Reservoir Triggered Seismicity

http://slidepdf.com/reader/full/gahalaut-et-al-reservoir-triggered-seismicity 1/8

A new case of reservoir triggered seismicity: Govind Ballav Pant 

reservoir (Rihand dam), central India

Kalpna Gahalaut  a , V.K. Gahalaut  a,⁎, M.R. Pandey  b

a  National Geophysical Research Institute, Uppal Road, Hyderabad 500 007, India

 b  Department of Mines and Geology, Lainchur, Kathmandu, Nepal 

Received 21 November 2006; received in revised form 27 March 2007; accepted 9 April 2007

Available online 14 April 2007

Abstract

We report here that seismicity near Govind Ballav Pant reservoir is strongly influenced by the reservoir operations. It is the

second largest reservoir in India, which is built on Rihand river in the failed rift region of central India. Most of the earthquakes

occurred during the high water stand in the reservoir with a time lag of about 1 month. We use the concept of coulomb stress

change and use Green's function based approach to estimate stresses and pore pressure due to the reservoir load. We find that the

reservoir increases coulomb stress on the nearby faults of the region that are favourably oriented for failure in predominantly

reverse slip manner under the NNE–SSW compression and thus promotes failure. The above two factors make it an obvious, yet so

far unreported case of reservoir triggered seismicity.

© 2007 Elsevier B.V. All rights reserved.

 Keywords:  Reserior triggered seismicity; Coulomb stress; Rihand reservoir; Peninsular India

1. Introduction

Increases in the frequency of occurrence of earthquakes

due to man's engineering activities have resulted from the

reservoir impoundment, quarrying and mining, and fluid

injection and extraction (McGarr and Simpson, 1997).

However, earthquakes caused by reservoir impoundment are stronger than by any other engineering activities. So far 

globally about hundred sites of Reservoir Triggered

Seismicity (RTS) have been reported which include at 

least eight sites from India, namely, Koyna, Warna, Bhatsa,

Dhamni, Gandipet, Idukki, Mula and Sriramsagar (Gupta,

2002). According to the mechanism of RTS, reservoir load

and/or induced pore pressure due to reservoir operations is

the cause of triggering of earthquakes on critically stressed

 pre-existing faults in the vicinity of the reservoirs

(Simpson, 1986; Gupta, 1992; McGarr and Simpson,

1997). In this article, we report a new case of RTS

associated with the Govind Ballav Pant reservoir, central

India (Fig. 1). The reservoir is located on the Rihand river,

a tributary of Son river, India. The 92 m high Rihand damwas built in 1962 and the reservoir is the second largest 

reservoir in India occupying an area of about 45×15 km2

having a maximum storage capacity of 10.6 km3, and a

capacity of electric power generation of 300 MW.

In this article, we analyse the correlation between the

times of high water levels in the reservoir and occurrence

of the maximum number of earthquakes, and simulate

the effect of the reservoir on the nearby earthquake

causative faults to verify that these earthquakes were

triggered by the reservoir.

Tectonophysics 439 (2007) 171–178

www.elsevier.com/locate/tecto

⁎  Corresponding author.

 E-mail address: [email protected]  (V.K. Gahalaut).

0040-1951/$ - see front matter © 2007 Elsevier B.V. All rights reserved.doi:10.1016/j.tecto.2007.04.003

Page 2: Gahalaut Et Al Reservoir Triggered Seismicity

7/23/2019 Gahalaut Et Al Reservoir Triggered Seismicity

http://slidepdf.com/reader/full/gahalaut-et-al-reservoir-triggered-seismicity 2/8

2. Tectonics and earthquakes of the region

The reservoir is located in a failed rift zone, which is

known as the Narmada–Son–Tapti failed rift zone. The

ENE–WSW trending failed rift zone transects the Indian

 peninsular shield area into northern and southern blocks(Fig. 1). It is seismically the most active region of the

 peninsular India, which evolved during the Archean and

Proterozoic period (Mahadevan and Subbarao, 1999).

Many prominent faults have been mapped in the region

 but the faults bounding the failed rift zone to the south

are suggested to be active and have witnessed reac-

tivation till quaternary period. Most of the earthquake

activity is also suggested to be associated with the

southern faults (Rao et al., 2002). The reservoir lies

close to and south of Son– Narmada South fault. Nu-

merous NW–

SE and NE–

SW trending small neotec-tonic faults have also been mapped in the vicinity of the

reservoir (GSI, 2000), however, sense of motion on

these faults is not known (Fig. 1).

In past 100 years only one major earthquake, namely

the June 2, 1927, Son Valley earthquake ( M   6 1/2;

Gutenberg and Richter, 1954, p.207), has occurred

within 100 km of the reservoir (Chandra, 1977). Threeestimates of epicentre are available (Fig. 1) and suggest 

variation in the estimate of up to 100 km. The earth-

quake is assumed to have occurred in the lower crust of 

failed rift regions of Narmada Son and Tapti, similar to

the March 14, 1938 Satpura ( M ∼6 1/4 Gutenberg and

Richter, 1954) and May 21, 1997 Jabalpur earthquakes

(Singh et al., 1999; Rao et al., 2002; Gahalaut et al.,

2004). The rupture of the recent 1997 Jabalpur 

earthquake (Mw 5.8), which occurred about 300 km

WSW of the reservoir, is suggested to lie in the downdip

 parts of the Son–

 Narmada South fault. Focal mechan-isms of this (Singh et al., 1999) and 1970 Broach

Fig. 1. Broad tectonic features of the part of the failed rift region of Narmada–Son and Tapti over a smooth topographic map. Three estimates of 

epicentre of 1927 Son Valley earthquake ( M s 6.5), by Gutenberg and Richter (1954), ISC and USGS, are also shown. Stars denote earthquakes

reported in the ISC and IMD catalogues during 1984–2004 ( M ≥3) while small circles denote epicentres by DMG during Jan. 1997–Dec. 1999

( M ≤3). Bold arrows show the direction of maximum compression due to plate movement (Gowd et al., 1992). Dark brown colour faults, marked as

 F 1, F 2 and F 3 and the Son Narmada South fault are the neotectonic faults (GSI, 2000). The 1970 Broach and 1997 Jabalpur earthquakes are shown inthe inset. Inset also shows the Indian and Nepalese national seismological networks.

172   K. Gahalaut et al. / Tectonophysics 439 (2007) 171 – 178

Page 3: Gahalaut Et Al Reservoir Triggered Seismicity

7/23/2019 Gahalaut Et Al Reservoir Triggered Seismicity

http://slidepdf.com/reader/full/gahalaut-et-al-reservoir-triggered-seismicity 3/8

earthquake, Mw 5.4, (Chung, 1993), which occurred at 

the western extreme of the failed rift zone (Fig. 1),

suggest predominance of the compressive regime. Since

1984, about sixty small magnitude ( M ≤3) earthquakes

have been reported in the vicinity of the reservoir by the

national network (Fig. 1) of India MeteorologicalDepartment (IMD) and International Seismological

Centre (ISC). Prior to 1984, there is no documentary

evidence of earthquake occurrence in the region. How-

ever, it does not imply that such earthquakes did not 

occur prior to 1984 and after the impoundment of the

reservoir in 1962. Probably they were missed due to

absence of nearby seismic stations. Incidentally, the year 

1984 approximately coincides with the period of world-

wide strengthening of seismological network. However,

from earthquake catalogues, it appears certain that no

strong or even moderate magnitude earthquake hasoccurred in the region since the reservoir impoundment.

Department of Mines and Geology, Kathmandu, Nepal

(DMG), which operates a 17 station network across Nepal

(Fig. 1), about 400 km north of the reservoir, has reported

numerous small magnitude ( M ≤3) earthquakes from the

study region. These earthquakes cluster close to the

reservoir (Fig. 1). The data from January 1997 to

December 1999 only are available to us. Epicentres of 

these earthquakes as reported by IMD, ISC and DMG,

coincide within 10–15 km in a tight cluster around the

reservoir, though their focal depths are unreliable. We

agree that the location of these earthquakes contains error 

of about 10–15 km, but it does not change our results and

interpretation in any significant way. Occurrence of earthquakes near the reservoir makes it imperative to

test the hypothesis of RTS.

3. Temporal variation of seismicity

Temporal variation of the earthquakes reported by

DMG that occurred within 30 km from the reservoir, and

the maximum water levels in the reservoir are shown in

Fig. 2. Frequency of earthquake occurrence is the lowest 

during June–July in each year and then it increases. The

water level in the reservoir is the minimum during lateMay and then it increases with the onset of monsoon

each year. We discretized the two time series at equal

interval of 1 month and calculated the cross-correlation.

The cross correlation coefficient between the two time

series is 0.71 at an average lag of about a month between

reservoir water level and frequency of earthquake

occurrence. The correlation decreases to 0.67 if we

consider earthquakes within 50 km from the centre of 

Fig. 2. Monthly frequency of earthquakesreported by DMG( M ≤3) that occurredwithin30 km fromthe reservoir, the reservoir water levels andthe Cross-

correlation function between the two time series. The correlation is the maximum (0.71) at a time lag of about 1 month. Graph with dashed line shows thetemporal variation of change in stress in wet state, ΔS w, at a point located at a depth of 8 km under the reservoir and the prominent earthquake cluster.

173 K. Gahalaut et al. / Tectonophysics 439 (2007) 171 – 178

Page 4: Gahalaut Et Al Reservoir Triggered Seismicity

7/23/2019 Gahalaut Et Al Reservoir Triggered Seismicity

http://slidepdf.com/reader/full/gahalaut-et-al-reservoir-triggered-seismicity 4/8

the reservoir. The diffused earthquake cluster north of 

the reservoir may not be correlated with the reservoir 

water level variations as the correlation decreases to

0.60 once we include it.

The only limitation here is that the catalogue is of 

shorter duration. The IMD and ISC catalogues cannot be

used for such quantitative correlation exercise, as there arenot many events in these catalogues and their complete-

ness may also be doubtful. However, we qualitatively

analysed the temporal variation in the occurrence of these

earthquakes reported by ISC and IMD since 1984. As

reported earlier, in the period from 1984–2004, 59

earthquakes of  M N3 occurred in the region. Maximum

number of earthquakes occurred in the months of 

December and January, followed by occurrences in the

months of February, March and April (Fig. 3). Earthquake

occurrences are the least in remaining months of the year.

Even this correlation, though qualitative, indicates that majority of the earthquakes occurred in the periods after 

high water level was attained in the reservoir. Thus the

good correlation between the water level and earthquakes

(Figs. 2 and 3) suggests that the frequency of earthquake

occurrence is influenced by the annual changes in the

reservoir water levels.

4. Effect of reservoir operation on the earthquake

causative faults

We consider stress changes due to reservoir water 

load and the pore pressure. Since most of the earth-

quakes appear to have occurred during the periods of 

high water level in the reservoir, first we considered the

effect of reservoir water load alone in earthquake

triggering. In the subsequent analysis we included the

effect of pore pressure also.

4.1. Effect of reservoir water load 

Following Chander and Kalpna (2000), we simulated

the load of Rihand reservoir through a series of 

rectangular loads. The load was uniform in each

rectangle but its magnitude decreased with distance of 

the rectangle upstream from the dam. The maximum

water depth was considered as 90 m at the dam. We used

formulas based on 3d Boussineq solutions (Jaeger and

Cook, 1969, p.281) to compute cumulative values of six

stress tensors. These stress tensors were used to

calculate change in normal (Δσ) and shear (Δτ ) stresson a given fault plane, which were then used to calculate

change in stability in dry state, i.e., without considering

 pore pressure, ΔS d, (= Δτ −μΔσ, where μ  is the friction

coefficient) due to reservoir load (Jaeger and Cook,

1969; Gough and Gough, 1970; Bock, 1980) on the

considered faults. Positive values of   ΔS d   suggest 

destabilization and promote failure on the considered

 planes and vice-versa.

Unfortunately, focal mechanisms of the earthquakes

that occurred close to the reservoir are not available. We

assumed that these earthquakes accompanied similar 

motion as was involved during 1997 Jabalpur and 1970

Fig. 3. Average monthly frequency of earthquakes ( M ≥3) in ISC and IMD catalogues during the period 1984–2004 along with the average reservoir 

water levels.

174   K. Gahalaut et al. / Tectonophysics 439 (2007) 171 – 178

Page 5: Gahalaut Et Al Reservoir Triggered Seismicity

7/23/2019 Gahalaut Et Al Reservoir Triggered Seismicity

http://slidepdf.com/reader/full/gahalaut-et-al-reservoir-triggered-seismicity 5/8

Broach earthquakes. It has been suggested that during

these earthquakes reverse motion occurred on the south

dipping plane (Singh et al., 1999; Gahalaut et al., 2004)

which is consistent with the NE to NNE directed

compressional regime of central and peninsular India

(Gowd et al., 1992). Thus we calculated   ΔS d   on thesouth dipping fault planes of 1997 Jabalpur (strike,  ϕ

61°, dip,   δ   64° and rake,   λ   74°) and 1970 Broach

earthquake (ϕ 35°, δ  49° and λ  44°) at a depth of 10 km

assuming that all the earthquakes near the reservoir 

occurred at shallow depth in the upper crust, as at deeper 

depth, reservoir effects will not be very significant. This

is in contrast with the focal depth of 36 km for the 1997

Jabalpur earthquake (Singh et al., 1999), but consistent 

with 11 km focal depth for 1970 Broach earthquake

(Chung, 1993). Focal depths of both the earthquakes

were determined from the waveform modeling (Chung,

1993; Singh et al., 1999) and hence appear to be reliable.

Region of increased  ΔS d  in both cases lie southeast of 

the reservoir, whereas the region of earthquake

occurrence lie west of the reservoir. Thus the two did

not correlate in any significant way. Computations of ΔS d   even on the north dipping planes did not produce

favourable results. Thus we suggest that the reservoir 

load does not destabilize the faults of 1997 Jabalpur or 

1970 Broach earthquake type. We even changed the

depth at which   ΔS d   is calculated but the results and

interpretation do not change in any significant way.

Unfavourable results of the above section prompted

us to search for the faults on which the reservoir effects

were favourable. We considered all the geologically

mapped faults in the vicinity of the reservoir for our 

Fig. 4. Stress changes ΔS  (in kPa) on the SSW dipping planes (ϕ 120°, δ  70° and  λ  120°, corresponding to the plane   ‘a’ of the derived fault plane

solution, shown at the top right) are shown. Geologically mapped neotectonic faults having similar orientation ( F 1, F 2 and  F 3) are also shown. Red

colour contours indicate the region of destabilization. (a)Δ

S d at 8 km depth with contour interval of 10 kPa. (b)Δ

S d in a vertical section along A–

A′,

contour interval is 5 kPa. (c) ΔS w  at 8 km depth (d) ΔS w  in a vertical section along B–B′.

175 K. Gahalaut et al. / Tectonophysics 439 (2007) 171 – 178

Page 6: Gahalaut Et Al Reservoir Triggered Seismicity

7/23/2019 Gahalaut Et Al Reservoir Triggered Seismicity

http://slidepdf.com/reader/full/gahalaut-et-al-reservoir-triggered-seismicity 6/8

analysis. Throughout our analysis, the sense of motion

on geologically mapped faults is chosen in such a way

so as to be consistent with NNE–SSW compression due

to plate tectonic forces. This is also a condition in the

analysis related to the RTS, as earthquakes will be

triggered only on those faults which are criticallystressed for failure under the ambient stresses (Simpson,

1986; McGarr and Simpson, 1997; Talwani, 1997;

Gupta, 2002). The reservoir acts only as trigger. Thus

under the compressive environment of central and

 peninsular India, all the faults present in the region will

either be reverse or strike slip faults, depending upon the

orientation of the faults with respect to the NNE to NE

directed maximum compression (Gowd et al., 1992). We

considered the orientation of the geologically mapped

faults and performed a grid search to estimate the dip

direction, dip amount and slip direction of these faults insuch a way that the reservoir causes destabilization on

these faults in the region of earthquake occurrences. The

computations of  ΔS d  were done at several depths but 

here we show  ΔS d  at 8 km depth (Fig. 4a), where the

effect of the reservoir load is most pronounced (Fig. 4 b).

We find that WNW–ESE oriented geologically mapped

neotectonic faults (GSI, 2000), shown as  F 1, F 2 and  F 3in Figs. 1 and 4, with δ  of 70°±10° and λ  of 120°±10°,

having strike of about N120° are destabilized by the

reservoir load in the region of prominent earthquake

cluster reported by DMG and ISC (Fig. 4). Specifically,

the east –west trending south dipping Son– NarmadaSouth fault, the largest fault in the region, did not 

 produce destabilization in the region of earthquake

occurrence. It caused destabilization north of it. This was

expected, as a reservoir located in the hanging wall will

stabilize a thrust fault, rather than destabilizing it 

(Roeloffs, 1988). The faults   F 1,   F 2  and   F 3   have been

mapped, before reservoir impoundment in 1962, by the

Geological Survey of India (GSI), in their district level

map series (e.g., see geological and tectonic maps of 

Sarguja, Koriya, Sidhi and Mirzapur districts published

 by GSI) and are shown in Figs. 1 and 4. The derived fault  plane solution corresponding to the above plane is also

shown in  Fig. 4. In an effort to further strengthen our 

analysis, we performed the grid search to identify the

strike along with the dip and rake of the fault and found

that the above estimated planes (ϕ   120°,   δ   70° and

λ 120°) are the only fault planes on which the effect of 

reservoir is favourable in the region of intense

earthquake activity. We also show an east –west vertical

cross section of  ΔS d  across the reservoir, in which it is

seen that at all depths, the region west of the reservoir 

experiences destabilization due to reservoir load. A

nominal value of 0.65 for   μ   is assumed in these

calculations. Change in   μ   by ±0.2 does not alter the

 pattern of destabilization in any significant way.

4.2. Effect of pore pressure

We calculate stress changes in wet state, i.e., consid-ering the pore pressure, ΔS w (=Δτ −μ(Δσ−Δ P ), where

Δ P   is the pore pressure due to reservoir operations

since its impoundment started in 1961. Change in

normal (Δσ) and shear (Δτ ) stresses due the reservoir 

are calculated in the same manner as in the previous

section.   Δ P   is calculated by solving the following

inhomogeneous diffusion equation,

C j2 P ¼A

AT   P 

 B

3 h

;

where c is the hydraulic diffusivity, B is the Skempton's

coefficient and  θ /3 is the mean stress. The value of  c,

hydraulic diffusivity constant was considered to be

10 m2/s, by trial and error method, so as to give a time

lag of about 1 month at a point beneath the maximum

seismicity at an assumed depth of 8 km. The adopted

value of  c is consistent also with Talwani and Acree's

(1984) estimates derived from seismicity observations

near various reservoirs worldwide. Higher value of   c

decreases the time lag and increases the magnitude of 

 pore pressure, and vice-versa. B varies from 0 to 1, here

it is considered as 0.7 (Talwani et al., 1999). Thesolution of above equation is given by   Kalpna and

Chander (2000). Pore pressure due to cyclic loading and

unloading develops gradually in the initial periods (till

1970), but later on, pore pressure, and hence   ΔS w,

mimics the pattern of annual cycle of water level

changes in the reservoir with a time lag (Fig. 2).

Considering the lag of 1 month between the water 

level and monthly frequency of earthquakes, we

calculated   ΔS w   on a plane at depth of 8 km, after 

1 month of maximum water level in the reservoir. We

repeated the exercise of the previous section to computeΔS w and found that the pattern in the zone of stability did

not change significantly from the previous corres-

 ponding cases. Thus even after considering the pore

 pressure the region of increased  ΔS w  corresponding to

the 1997 Jabalpur and 1970 Broach earthquake faults

does not coincide with the region of earthquake

occurrence. However, the results of grid search method

suggested that once again faults with   ϕ   120°±10°,

δ   70°±10° and   λ   120° ± 10° are destabilized by the

reservoir load and induced pore in the region of prom-

inent earthquake cluster reported by DMG and ISC

(Fig. 4).

176   K. Gahalaut et al. / Tectonophysics 439 (2007) 171 – 178

Page 7: Gahalaut Et Al Reservoir Triggered Seismicity

7/23/2019 Gahalaut Et Al Reservoir Triggered Seismicity

http://slidepdf.com/reader/full/gahalaut-et-al-reservoir-triggered-seismicity 7/8

The east –west vertical cross section of  ΔS w  across

the reservoir (Fig. 4d) suggests that the region west of 

the reservoir experiences destabilization due to reservoir 

operations at all depths. However, the effect is most 

 pronounced at about 8 km. The analysis suggests that 

consideration of pore pressure not only leads to increasein the magnitude of stress changes (ΔS wNΔS d) on the

considered fault, it also leads to an increase in the region

of destabilization and now the region of destabilization

encompasses the entire region of earthquake activity

near the reservoir.

We computed temporal variation of ΔS w (Fig. 2) at a

 point that lies at 8 km depth under the reservoir and

where the increased  ΔS w   is the maximum (Fig. 4c). It 

can be seen that earthquakes do not appear to have

occurred during the low water stand, even though the

effect of water load was to destabilize the fault evenduring that period (Fig. 2). Thus the analysis implies that 

the faults in the region are critically stressed for failure

under compressive regime and a small increase in stress

change due to reservoir operation, by about 25 kPa only,

corresponding to an annual increase of water load by

10–15 m (Fig. 2), triggers earthquakes on these faults.

Presence of critically stressed faults is also supported by

the observation that this region is the seismically most 

active region of peninsular India.

Thus the above analyses pertaining to the quantita-

tive effect of reservoir operation on the seismogenic

faults suggest that the reservoir operations destabilizethe nearby neotectonic faults.

5. Concluding remarks

Several factors control earthquake triggering by the

reservoir, important of them are, reservoir dimensions;

annual changes in the reservoir water level; ambient 

stresses and presence of faults and their orientation with

respect to the ambient stresses, etc. (Simpson, 1986). In

this case the Govind Ballav Pant reservoir on Rihand

river is very large, in fact the second largest in India,having annual water level changes of the order of about 

12±2 m, it is situated in a failed rift region of Narmada–

Son–Tapti, which is seismically the most active region in

the peninsular India (Gahalaut et al., 2004) and where

several faults have been mapped (Fig. 1). These factors

make it an ideal site where reservoir may trigger 

earthquakes on the nearby faults. A good correlation

 between temporal variation of seismicity and reservoir 

water levels (Fig. 2) suggest a possible case of reservoir 

triggering. Our simulation using a Green's function-

 based approach related to stability change on the

earthquake causative neotectonic faults due to reservoir 

operation (Fig. 4) supports the view that water level

changes in the reservoir have caused pore pressure

changes at hypocentral depths on the pre-existing faults,

which are favourably oriented, to trigger the earth-

quakes. This makes it a persuasive, yet so far unreported,

case of triggered seismicity due to the reservoir loading.In the region of interest, we have no knowledge about 

the earthquake occurrence immediately after the reser-

voir impoundment in 1962 and before 1984, the later 

year marks the year of beginning of reliable earthquake

catalogues. But good correlation between earthquakes

during 1984–2004 (for   M N3 from ISC and IMD

catalogues) and during 1997–99 (for  M b3 from DMG

catalogue) with reservoir water level possibly suggests

that the reservoir triggered earthquakes occurred in this

region after the impoundment and the reservoir 

continued to trigger earthquakes even after 40 years of impoundment. Cases of continued seismicity near a

reservoir have been reported from elsewhere as well,

e.g., Koyna–Warna in India (Gupta, 2002); Lake Mead

in USA (Talwani, 1997), Aswan in Egypt (Mekkawi

et al., 2004), Açu in Brazil (do Nascimento et al., 2004)

etc.

Acknowledgements

We are thankful to the Central Electricity Authority

for providing Rihand reservoir water levels, DMG,

 Nepal for providing the microseismicity data (acquiredunder DASE, France and DMG collaboration), and R.S.

Dattatrayam and G. Suresh of IMD, New Delhi for 

 providing the earthquake data. We benefited from the

comments of A. McGarr, Pradeep Talwani, Harsh Gupta,

Kusala Rajendran and two anonymous reviewers. We

thank Director, NGRI, R.K. Chadha and M. Ravi Kumar 

for their support.

References

Bock, Y., 1980. Load induced stresses and their relation to initial stress

field. J. Geophys. 48, 94–100.

Chander, R., Kalpna, 2000. On categorising induced and natural

tectonic earthquakes near new reservoirs. Eng. Geol. 46, 81–92.

Chandra, U., 1977. Earthquakes of Peninsular India  —   a seismotec-

tonic study. Bull. Seismol. Soc. Am. 67, 1387–1413.

Chung, W.Y., 1993. Source parameters of the rift-associated intraplate

earthquakes in peninsular India: the Bhadrachalam earthquake of 

April 13, 1969 and the Broach earthquake of March 23, 1970.

Tectonophysics 225, 219–230.

do Nascimento, A.F., Cowie, P.A., Lunn, R.J., Pearce, R.G., 2004.

Spatio-temporal evolution of induced seismicity at Açu reservoir,

 NE Brazil. Geophys. J. Int. 158, 1041–1052.

Gahalaut, V.K., Rao, V.K., Tewari, H.C., 2004. On the mechanism andsource parameters of the deep crustal Jabalpur earthquake, India, of 

177 K. Gahalaut et al. / Tectonophysics 439 (2007) 171 – 178

Page 8: Gahalaut Et Al Reservoir Triggered Seismicity

7/23/2019 Gahalaut Et Al Reservoir Triggered Seismicity

http://slidepdf.com/reader/full/gahalaut-et-al-reservoir-triggered-seismicity 8/8

May 21, 1997: constraints from aftershocks and change in static

stress. Geophys. J. Int. 156, 345–351.

Gough, D.I., Gough, W.I., 1970. Stress and deflection in the

lithosphere near Lake Kariba  — I. Geophys. J. 21, 65–78.

Gowd, T.N., Rao, S.V.S., Gaur, V.K., 1992. Tectonic stress field in the

Indian subcontinent. J. Geophys. Res. 97, 11879–11888.

GSI, 2000. Seismotectonics atlas of India. Geological Survey of India,Calcutta, India, p. 87.

Gupta, H.K., 1992. Reservoir Induced Earthquakes. Elsevier, Amsterdam.

355 pp.

Gupta, H.K., 2002. A review of recent studies of triggered earthquakes

 by artificial water reservoirs with special emphasis on earthquakes

in Koyna, India. Earth-Sci. Rev. 58, 279–310.

Gutenberg, B., Richter, C.F., 1954. Seismicity of the Earth, 2nd

edition. Princeton Univ. Press, Princeton, New Jersey.

Jaeger, J.C., Cook, N.G.W., 1969. Fundamentals of Rock Mechanics.

Methuen, London, p. 515.

Kalpna, Chander, R., 2000. Green's function based stress diffusion

solutions in the porous elastic half space for time varying finite

reservoir loads. Phys. Earth Planet. Inter. 120, 93–101.

Mahadevan, T.M., Subbarao, K.V., 1999. Seismicity of the Deccanvolcanic province—an evaluation of some endogenous factors. In:

Subbarao, K.V. (Ed.), Deccan Volcanic Province. Geol. Soc. India

Mem., vol. 43, pp. 453–484.

McGarr, A., Simpson, D., 1997. Keynote lecture: a broad look at 

induced seismicity. Rockbursts and Seismicity in Mines. Balkema,

Rotterdam, pp. 385–396.

Mekkawi, M., Grasso, J.R., Schnegg, P.A., 2004. A long-lasting

relaxation of seismicity at Aswan reservoir, Egypt, 1982–2001.

Bull. Seismol. Soc. Am. 94, 479–492.

Rao, N.P., Tsukuda, T., Koruga, M., Bhatia, S.C., Suresh, G., 2002.

Deep lower crustal earthquakes in central India: inferences from

analysis of regional broadband data of the 1997 May 21 Jabalpur 

earthquake. Geophys. J. Int. 148, 132–138.Roeloffs, E.A., 1988. Fault stability changes induced beneath a

reservoir with cyclic variations in water level. J. Geophys. Res. 93,

2107–2124.

Simpson, D.W., 1986. Triggered earthquakes. Annu. Rev. Earth

Planet. Sci. 14, 21–42.

Singh, S.K., Dattatrayam, R.S., Shapiro, N.M., Mandal, P., Pacheco,

J.F., Midha, R.K., 1999. Crustal and upper mantle structure of 

 peninsular India and source parameters of the 21 May 1997

Jabalpur earthquake (Mw= 5.8): Results from a new regional

 broadband network. Bull. Seismol. Soc. Am. 89, 1631–1641.

Talwani, P., 1997. On the nature of reservoir-induced seismicity. Pure

Appl. Geophys. 150, 473–492.

Talwani, P., Acree, S., 1984. Pore pressure diffusion and the

mechanism of reservoir-induced seismicity. Pure Appl. Geophys.122, 947–965.

Talwani, P., Cobb, J.S., Schaeffer, M.F., 1999. In situ measurements of 

hydraulic properties of a shear zone in northwestern South

Carolina. J. Geophys. Res. 104, 14993–15003.

178   K. Gahalaut et al. / Tectonophysics 439 (2007) 171 – 178