Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Water Management Proceedings of the Institution of Civil Engineers

Volume 166 Issue WM8 Water Management 166 September 2013 Issue WM8
Pages 452–462 http://dx.doi.org/10.1680/wama.11.00084
Uplift force, seepage, and exit gradient Paper 1100084
under diversion dams Received 12/09/2011 Accepted 19/03/2012
Published online 03/11/2012
Tokaldany and Shayan Keywords: hydraulics & hydrodynamics/models (physical)/river engineering

ICE Publishing: All rights reserved

Uplift force, seepage, and exit


gradient under diversion dams
j
1 E. Amiri Tokaldany MSc, PhD j
2 H. Khalili Shayan MSc
Associate Professor, Irrigation and Reclamation Engineering MSc Graduate, Irrigation and Reclamation Engineering Department,
Department, University of Tehran, Karaj, Iran University of Tehran, Karaj, Iran

j
1 j
2

Correct estimation of uplift force, seepage discharge and exit gradient is very important in stability analysis of
hydraulic structures. In this research, by carrying out a set of experiments on a laboratory model, the
application of various methods for estimating uplift pressure and seepage discharge under hydraulic structures,
and the exit gradient have been investigated. The results show that for the soil type used in the experiments,
the method of Khosla et al. gives a better estimation of seepage effects than creep theories. Moreover,
compared with Lane’s method, it is found that Bligh’s theory has better agreement with the observed data of
uplift pressure, seepage discharge and piping. By using the finite-element method, the magnitude of pressure
head upon a dam foundation in different conditions is calculated, and a relationship between the anisotropic
conductivity ratio and the ratio of coefficients in Lane’s creep theory is introduced. It is found that while the
accuracy of Lane’s theory is reduced by increasing the anisotropic conductivity ratio, the new relation estimates
the amount of uplift force very well. Based on the finite-element method, a set of graphs is presented to
estimate the exit gradient in different conditions with the presence of a cutoff wall at the downstream end or
without any cutoff wall.

Notation P uplift pressure (N)


b length of foundation (m) q seepage discharge per unit width (m2 /s)
C allowable exit gradient in creep theories RSQ correlation coefficient
D depth of impervious layer X any point along creep length
d length of cutoff wall (m) x distance of cutoff wall from upstream of floor (m)
FP uplift force (N) ªW specific weight of water (N/m3 )
FP0 uplift force in the absence of cutoff wall ˜h difference between upstream and downstream
H upstream head (m) heads (m)
P
h total head in any point (x, y) (m) L total horizontal percolation length (m)
P H
hX uplift pressure at any desired point (X) (m) LV total vertical percolation length (m)
IE exit gradient at the toe when L , b
IEmax maximum exit gradient at the toe of dam
i available gradient 1. Introduction
iexit exit gradient Due to the difference in water levels between the upstream and
i0 exit gradient in the presence of cutoff wall downstream sides of water structures constructed on permeable
kx horizontal hydraulic conductivity (m/s) foundations, seepage occurs under the foundation of these
ky vertically hydraulic conductivity (m/s) structures. The effects of seepage on the foundation of hydraulic
ky /kx anisotropic conductivity ratio structures can be classified into three parts: uplift force, seepage
L downstream permeable length discharge and exit gradient. The uplift force, by reducing shear
Leq equivalent length (m) resistance between the dam and its foundation, results in gener-
LX creep length for any point (X) (m) ating strain tension, and finally decreasing the safety factor
MAE mean absolute error against sliding or overturning of the dam structure. The exit
m,n weighting factors gradient is, on the other hand, the main design criterion in

452
Water Management Uplift force, seepage, and exit gradient
Volume 166 Issue WM8 under diversion dams
Tokaldany and Shayan

determining the safety of hydraulic structures against the piping 1X X


2:
Leq ¼ LH þ LV
phenomenon. 3

Bligh (1910) introduced the creep length theory for the flow
P
passing under hydraulic structures. He defined the creep length as in which Leq is total equivalent length, LH is total horizontal
P
the route of the first line of seepage which is in contact with the percolation length (walls with slope less than 458) and LV is
structure’s foundation. Bligh stated that hydraulic gradient is total vertical percolation length (walls with slope more than 458).
constant along the creep line and energy loss along this path As with Bligh’s method, the uplift pressure distribution under the
varies linearly with creep length – that is, uplift pressure structure foundation is assumed to be linear.
distribution is linear under the structure foundation. So according
to Bligh’s theory, uplift pressure (h X ) at any desired point (X) of Since undermining begins from the downstream toe of the
a creep path is (Figure 1) structure, in order to prevent this phenomenon, the available exit
gradient (i x ) has to be less than the allowable exit gradient, C.
So, to prevent undermining
LX
hX ¼ H  ˜h
2d þ b
˜h Leq 1
8 ix ¼ <C! > ) Leq > C˜h
<X 0<X <x 3: Leq ˜h C
LX ¼
: X þ 2d X .x
1:
It should be noted that the allowable exit gradient is defined for
various types of soil in both methods, but with different values
(e.g. see Leliavsky, 1965).

where H is upstream total head, ˜h is difference between Khosla et al. (1936) presented a method of estimating the
upstream and downstream heads, b is length of foundation, Lx is distribution of uplift pressure under foundations through solving
distance between the upstream end of the structure to any point complex potential functions. They investigated the flow network
(X) in the contact line under the foundation, x is distance of under a hydraulic structure which was constructed on a permeable
cutoff wall from upstream end of floor, and d is length of cutoff foundation. Khosla et al. assumed that flow and potential lines
wall. are concentric ellipses and hyperbolic, respectively. Considering
no cutoff wall, the proposed relationship by Khosla et al. to
Lane (1935) based his investigation on over 200 damaged estimate uplift pressure distribution along the floor is (Figure 2)
hydraulic structures and reported that there is a difference (Leliavsky, 1965)
between horizontal and vertical creep paths. Consequently, he
presented his weighted creep theory assigning coefficients of 0.33
Hªw 2x b b
and 1.0 for total horizontal and vertical percolation lengths, P¼ cos1 for  <x<
respectively. Therefore, according to Lane’s weighted creep
4:  b 2 2
theory, the equivalent creep length is defined as (Leliavsky, 1965)

Therefore, the amount of uplift force per unit width along the
floor by using the method of Khosla et al. is given by

b/2 b/2
Δh x
H
b
H/2
X L
d H Hγw 2x
P⫽ cos⫺1
D π b
y

Bligh’s theory

Figure 2. Comparison of uplift distribution according to the Bligh


Figure 1. Definition of parameters in creep theory and Khosla et al. theories

453
Water Management Uplift force, seepage, and exit gradient
Volume 166 Issue WM8 under diversion dams
Tokaldany and Shayan

ð x¼(b=2)
H 2x horizontal and vertical hydraulic conductivity, respectively. This
FP ¼ ªw cos1 dx
x¼(b=2)  b method presents an exact solution for solving the cases with
different boundary conditions, but the equation is difficult for
ð x¼(b=2)
ªw H 2x engineering applications and includes difficult integrals when a
¼ cos1 dx cutoff wall or change in extent of seepage flow is introduced
 x¼(b=2) b
(Harr, 1962). Abedi Koupaei (1991) in a case study calculated
ð x¼(b=2) the distribution of uplift pressure using the four above-mentioned
2x
As cos1 dx methods. He stated that the amount of uplift pressure estimated
x¼(b=2) b
by using Bligh and Lane’s theories is less than both Khosla et al.
2 sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 3 and the finite-difference methods.
 2
b 42x 1 2x 2x 5 b
¼ cos  1  ¼ Griffiths and Fenton (1997), by combination of the techniques of
2 b b b  x¼(b=2) 2
x¼(b=2) random fields generating with the finite-element (FE) method,
tried to model the 3-D steady seepage in which the permeability
1 was randomly distributed within the soil body. Opyrchal (2003)
5: (Then) FP ¼ Hªw b
2 presented the application of the fuzzy concept to identify the
seepage path within the body of a dam. Jie et al. (2004)
investigated the seepage beneath dams and embankments using
Although there is a difference between the forms of pressure the finite-difference method (FDM) based on a boundary-fitted
distribution, Equation 5 gives nearly the same results for uplift coordinate transformation. Ahmed and Bazaraa (2009) investi-
force as Bligh’s theory when there is no cutoff wall under the gated three-dimensional seepage below and around hydraulic
foundation. As can be seen in Figure 2, at the downstream end of structures by using an FE method based on a computer program.
the impervious floor, the slope of the tangent line on Khosla et They compared the results of 3-D analysis with 2-D analysis for
al.’s curve is infinite and upward resulting in a piping phenomen- the estimation of exit gradient, seepage losses and uplift force.
on. Hence, a cutoff wall at the downstream end is required. The Ahmed (2011) investigated different configurations of sheet pile
Khosla et al. method can be applied for compound structures on the variations of uplift force, seepage losses and exit gradient
when there is more than one cutoff wall, for sloping floors and for designing hydraulic structures considering flow through canal
for thick foundations, and also takes into account the effects of banks.
multiple cutoff walls over each other. The exit gradient in the
Khosla et al. method is given by (Leliavsky, 1965) Due to broad acceptance by engineers around the world in using
the four principal methods mentioned above to estimate uplift
H force, exit gradient, and seepage under hydraulic structures, the
iexit ¼ pffiffiffiffi ,
d Y research reported in this paper discusses the use of a physical
pffiffiffiffiffiffiffiffiffiffiffiffiffiffi model, the application of finite-element analysis, and the per-
1þ 1 þ Æ2 b formance and accuracy of these methods. In this regard, in this
6:
where Y ¼ and Ƽ
2 d research the authors carried out a large number of laboratory
experiments upon a physical model, gathered a large quantity of
data, and investigated the variation of the parameters due to
in which iexit is the exit gradient. installation of a cutoff wall with different heights and positions
under the dam foundation. Also, they investigated the application
Although the theory of Khosla et al. is generally more reliable range of the coefficients recommended by Lane for seepage lines
than the creep theories of Bligh and Lane, in cases dealing with a in vertical and horizontal directions and their dependency on the
compound foundation it is required to solve very complicated anisotropic conductivity ratio of the materials under the dam
equations and also has a low accuracy when applied to aniso- foundation.
tropic foundations.
2. Experimental specifications and
The fourth method of estimating uplift pressure under hydraulic procedure
structures is based on solving the Laplace equation. The Laplace In order to compare the methods and to estimate uplift pressure,
equation, for steady conditions, is given by seepage discharge and exit gradient based on laboratory data, an
experimental programme was performed in a flume with a length
@2 h @2 h of 1.70 m and width of 0.18 m located in the Hydraulic Labora-
kx þ k y ¼0 tory of the Department of Irrigation and Reclamation Engineer-
7: @x 2 @ y2
ing, University of Tehran. The experimental set consisted of an
upstream impervious bed, an impervious wall as a dam body,
where h is the total head at any point (x, y), and kx and ky are various cutoff walls, and a piezometric network, all of which were

454
Water Management Uplift force, seepage, and exit gradient
Volume 166 Issue WM8 under diversion dams
Tokaldany and Shayan

Pervious region

Piezometric tubes
70 cm

Impervious
region Impervious

40 cm
region

Cutoff wall positions

Flow inlet Seepage


40 cm 90 cm 40 cm measurement

Figure 3. Dimensions of the physical model

constructed inside the flume (Figure 3). The length of cutoff walls the results obtained from Bligh, Khosla et al. and Lane’s method
varied from 2.5 cm to 30 cm and they were located in various are indicated. As shown in Figure 4(b), the accuracy of Bligh’s
positions beginning from 40 cm upstream to 115 cm downstream method is greater than Lane’s method while Khosla et al.’s
from the impervious wall. The upstream head water was taken method has excellent accuracy in estimating the exit gradient.
from 2.5 cm to 20 cm and the upstream water level was fixed by
using a floating body. The downstream water level was set to Also, Figure 4(c) shows that the accuracy of Lane’s method on
zero. Distribution of uplift pressure was measured by the piezo- estimating seepage discharge is lowest while the FE method
metric network, which consisted of 39 piezometers (13 rows with computed the seepage discharge exactly as it is. The accuracy of
each row including three piezometers, Figure 3). Bligh’s method is good.

To prevent uncontrolled piping, a gravel filter (D50 ¼ 2.0 mm) To make a quantitative evaluation of the accuracy of the methods
was dumped in both upstream and downstream beds. Beach sand, used to estimate the above parameters, two statistical parameters
classified as the most unsuitable type of soil from the point of were used: mean absolute error (MAE) and correlation coefficient
view of stability of hydraulic structures, was selected as pervious (RSQ) defined as
bed material. The hydraulic conductivity of the soil was measured
using the constant-head method and injected dye. The soil X
n

horizontal and vertical hydraulic conductivities were estimated as ð xi  ^xi Þ2


i¼1
0.00143 m/s and 0.001202 m/s, respectively. A total of 110 RSQ ¼ 1 
X
n
experiments were conducted in various conditions of upstream ð xi  xÞ2
heads, cutoff depths, and cutoff wall positions. In each experi- 8: i¼1
ment, the seepage discharge was measured based on a volumetric
method. Uplift pressure distribution was taken by reading water
levels in piezometeric tubes.
1X n
MAE ¼ j xi  ^xi j
9: n i¼1
3. Results and discussion
3.1 Experimental versus theoretical results
In Figure 4 the results of total uplift force, exit gradient and where xi is the observed data, ^xi is the calculated data, and xi is
seepage discharge obtained from experimental tests versus the the average of observed data. In Table 1 the amounts of MAE
results obtained from Bligh, Lane, Khosla et al. and FE methods and RSQ for different parameters are presented. Based on Figure
are indicated. As shown in Figure 4(a), the accuracy of Bligh, 4, despite the fact that Lane’s theory was presented as a modified
Khosla et al. and FE methods to estimate the amount of uplift method of Bligh’s creep theory, Bligh’s theory was found to show
pressure are nearly equal, while the accuracy of Lane’s method is greater accuracy than Lane’s in estimating uplift pressure,
lower than the other methods. seepage discharge and exit gradient in this study.

Since it is not possible to obtain experimentally the amount of 3.2 Estimating exit gradient using the Khosla et al.
the exit gradient, the FE method employed in Geostudio2007 equation
software was used to determine the exit gradient. In Figure 4(b), For further investigation of the accuracy of the Khosla et al.
the results for exit gradient obtained from the FE method versus method (Equation 6) for estimating the exit gradient, it should be

455
Water Management Uplift force, seepage, and exit gradient
Volume 166 Issue WM8 under diversion dams
Tokaldany and Shayan

300 Bligh’s theory 0·40


Lane’s theory 0·35
250 Khosla’s method 0·30
Calculated uplift force: N

Finite element 0·25 y ⫽ 0·651x⫺0·58

IEd/H
0·20 R 2 ⫽ 0·997
200 0·15
MAEL ⫽ 98·816 0·10
150 MAEB ⫽ 5·807 0·05 Finite element Khosla’s method (Equation 6)
MAEFE ⫽ 1·719 0
100 MAEKh ⫽ 11·974 0 2 4 6 8 10 12 14 16 18 20
b/D
50
Figure 5. Estimated exit gradient by using the Khosla et al. and
0 finite-element methods
0 50 100 150 200 250 300
Observed uplift force: N
(a)
0·8 MAE RSQ
Bligh’s theory
0·7 Lane’s theory
Hydraulic gradient Bligh 0.13 0.810
Bligh/Lane/Khosla methods

Khosla’s method
0·6 Lane 0.05 0.756
0·5 MAEL ⫽ 0·13 Khosla et al. 0.043 0.966
MAEKh ⫽ 0·043 Seepage discharge Bligh 2.40 3 1006 0.985
0·4 MAEB ⫽ 0·05 Lane 1.23 3 1005 0.854
0·3 Finite element 7.04 3 1008 0.999
0·2 Uplift force Bligh 5.807 0.992
Lane 98.816 0.849
0·1
Finite element 1.719 0.997
0 Khosla et al. 11.974 0.942
0 0·1 0·2 0·3 0·4 0·5 0·6 0·7 0·8
Finite-element method
Table 1. Evaluation of accuracy of results from Bligh, Lane, Khosla
(b)
et al. and finite-element methods, based on experimental data
0·000015
Calculated discharge: cm

than those estimated by the FE method. Considering Figure 5, a


0·000010 new equation is introduced in the form of Equation 6 to estimate
MAEB ⫽ 2·40 ⫻10⫺6 exit gradient by using FE
MAEL ⫽ 1·23 ⫻10⫺5
MAEFE ⫽ 7·04 ⫻10⫺8
0·000005 iexit d 1 :
Finite element ¼ pffiffiffiffi ¼ 0:6513Æ0 589
Bligh’s theory
H  Y
Lane’s theory :
0 10: ! Y ffi 0:23886Æ1 178
0 0·000005 0·000010 0·000015
Observed discharge: cms
(c)

3.3 Estimating exit gradient with no cutoff wall


Figure 4. Evaluation of different methods to estimate (a) total
Since the Khosla et al. method cannot estimate the amount of
uplift force, (b) exit gradient, and (c) seepage discharge
exit gradient when there is no cutoff wall at the downstream end,
in this research the authors aimed to introduce a new equation for
this condition. In this regard, considering the geometry of an
noted that for the case when there is no cutoff wall at the experimental model, the authors supposed an upstream head (H)
downstream end – that is, when d ! 0 – the exit gradient is of 2.5, 5, 7.5, 10, 12.5, 15, 17.5 and 20 cm, impermeable dam
infinite. Therefore, to compare the results of the Khosla et al. foundation lengths (b) of 10, 20, 30, 40, 50, 60, 70, 80, 90 and
method with those obtained from the FE analysis, the current 100 cm, a downstream permeable length (L) of 5, 10, 20, 30, 40,
authors considered a cutoff wall at the downstream end with 50, 60, 70, 80, 90 and 100 cm, and a constant value for the depth
various heights of 10, 15, 20 and 30 cm under an upstream fixed of the impervious layer (D). Then, using the FE method, they
head of 2.5, 5, 7.5, 10, 12.5, 15, 17.5 and 20 cm. Based on the estimated the maximum exit gradient at the toe of the dam
results shown in Figure 5, it can be seen that the prediction of (IEmax ) for any combination of h, b and L. For the case of L ¼ b,
exit gradient by the Khosla et al. method is nearly 33% lower the estimated values of exit gradient are shown in Figure 6(a).

456
Water Management Uplift force, seepage, and exit gradient
Volume 166 Issue WM8 under diversion dams
Tokaldany and Shayan

1·8
1·6 3.5 Dependence of Lane’s theory on the anisotropy of
1·4 the soil
1·2 In this research it was found, by measuring the hydraulic
IEmax D/H

1·0 conductivities in the x- and y-directions, that the anisotropic


0·8
0·6 conductivity ratio (ky /kx ) was equal to 0.84. It was found that the
0·4 great accuracy of Bligh’s method in this study was due to the soil
0·2 type, in which the hydraulic conductivity in both directions was
0 nearly the same. Therefore it can be expected that there is a
1 1 1 1 1 2
b/D relation between anisotropic conductivity ratio with the weighted
(a) ratio in Lane’s creep theory. For this purpose, a model was
2·0 considered according to Figure 7.
1·9
1·8
1·7 The authors assumed an upstream constant head of 20 m, a cutoff
1·6
IE / IEmax

1·5 wall with length of 15 m and thickness of 50 cm, with different


1·4 positions at the distance of 0.15, 30, 45 and 60 m from the
1·3 beginning of the floor. For each case, the authors considered
1·2
1·1 different anisotropic conductivity ratios and then calculated uplift
1·0 pressure distribution at specified points using Lane’s and the FE
0 0·2 0·4 0·6 0·8 1·0
L /b methods. The results are shown in Figure 8, from which the
(b) following can be seen.

Figure 6. Determination of exit gradient under no cutoff wall (a) The amount of uplift pressure calculated from Lane’s theory
condition was found to be sometimes more and sometimes less than
that calculated from the FE method when a cutoff wall has
been installed upstream or downstream, accordingly.
(b) When the cutoff wall is installed in the middle, the amount of
For cases when L , b, since the exit gradient is distributed in a uplift pressure calculated from Lane’s theory is less than that
smaller area, the value of the exit gradient is obviously increased calculated from the FE method from the beginning to the
at the downstream toe of the dam. To determine exit gradient middle, but it is more than that from the middle to the end of
when L , b, Figure 6(b) was produced by which the relative the floor.
increase in the amount of exit gradient versus the different values
of L/b can be calculated. The estimated value from Figure 6(a) It can be seen that in all cases the difference between the results
for the case of L ¼ b can be multiplied by the value obtained using Lane’s and the FE method is decreased by reducing the
from the vertical axis of Figure 6(b) and corresponds to the ratio anisotropic ratios (ky /kx ). This point suggests that the weighting
of L/b. factors in Lane’s theory may be dependent on the anisotropic
conductivity ratio of the soil. Therefore, in order to achieve the
3.4 Evaluation of hydraulic gradient distribution corrections in Lane’s relationship, Lane’s equivalent length can be
uniformity in creep theories written as
X X
To evaluate the accuracy of creep theories in estimating the Leq ¼ m LH þ n LV
11:
distribution of hydraulic gradient along the bed foundation of
hydraulic structures, a cutoff wall was installed with a height of
15 cm in different positions: at the upstream and downstream
ends and in the middle. A constant head of 20 cm was provided To obtain the value of coefficients m and n, first, the authors
at the upstream end of the physical model. The method of FE
in Geostudio2007 software was used to determine the amount 45 m 45 m
of the hydraulic gradient along the floor. The amount of 5m
hydraulic gradient was also calculated at the same points by
45 m

using Bligh and Lane’s theories. It was found that Bligh’s and
Lane’s theories assumed that the hydraulic gradient decreases
uniformly along the floor, while the gradient changes signifi-
cantly in the location of the cutoff wall in the FE method. In
150 m
the other words, the creep theories only consider the height of
the cutoff walls, while their results are not affected by the Figure 7. Dimensions of model for investigation of dependence of
location of the cutoff wall along the floor in corresponding creep weight factors on anisotropic ratio
conditions.

457
Water Management Uplift force, seepage, and exit gradient
Volume 166 Issue WM8 under diversion dams
Tokaldany and Shayan

30 Ky ⫽ Kx Kx ⫽ 2Ky
amount of uplift pressure in the same locations using Lane’s
Ky ⫽ 2Kx Kx ⫽ 3Ky
25 Ky ⫽ 3Kx Kx ⫽ 4Ky weighted creep theory. To compare the results in terms of
Ky ⫽ 4Kx Kx ⫽ 5Ky quantity, mean absolute error (MAE) was used. The amount of
Uplift pressure: m

20 Ky ⫽ 5Kx Lane’s theory MAE for different values of the n/m ratio in sample cases is
calculated and shown in Figure 9. Using an interpolation method,
15
the minimum value of MAE for each case is calculated and the
10 results are shown in Figure 10, which shows that increasing the
anisotropic conductivity ratio results in decreasing the n/m ratio,
5 provided that the resistance against water flow in the vertical
direction is decreasing.
0
0 10 20 30 40 50 60
Distance along floor: m From Figure 10, it can be seen that the assumption of an n/m
(a) ratio equal to 3 (as in Lane’s weighted creep theory) is correct
25 when there is no cutoff wall and when the soil is isotropic.
Therefore, application of Lane’s theory results in considerable
20 error when a cutoff wall is installed under the structure floor and
Uplift pressure: m

the soil is anisotropic.


15
Ky ⫽ Kx
Ky ⫽ 2Kx To derive a relationship between the anisotropic conductivity of
10 Ky ⫽ 3Kx the soil (X = Ky /Kx ) and n/m ratios in the presence of a cutoff
Ky ⫽ 4Kx wall, the authors used values of the n/m ratio provided from the
Ky ⫽ 5Kx Kx⫽ 4Ky corresponding soil anisotropic conductivity ratio (Figure 11). The
5
Kx ⫽ 2Ky Kx ⫽ 5Ky authors found the following equation between the two coefficients
Kx ⫽ 3Ky Lane’s theory
0 with a coefficient correlation of 0.92.
0 10 20 30 40 50 60
Distance along floor: m  
: : Ky
(b) 4 564 þ 8 631
25 n Kx
¼"  2 #
m K Ky
y
20 1 þ 10:862 þ 2:795
12: Kx Kx
Uplift pressure: m

15
Ky ⫽ Kx
Ky ⫽ 2Kx To compare the accuracy of Equation 11 with Lane’s weighted
10 Ky ⫽ 3Kx creep theory, the authors assumed a constant upstream head of
Ky ⫽ 4Kx
Kx ⫽ 4Ky
20 cm and a cutoff wall with a depth of 20 cm in three positions:
5 Ky ⫽ 5Kx
Kx ⫽ 2Ky Kx ⫽ 5Ky upstream end, middle, and downstream end. The uplift pressure
Kx ⫽ 3Ky Lane’s theory at corresponding points was estimated from both Lane’s theory
0 and Equation 11, and the accuracy of the results was compared
0 10 20 30 40 50 60
Distance along floor: m by using the results obtained from the FE method (Figure 12).
(c) Figure 12 shows that as the anisotropic ratio of the soil increases,
the accuracy of Lane’s method decreases while the accuracy of
Figure 8. Distribution of uplift pressure along floor estimated
Equation 11 increases. Figure 12 also highlights the point that
from Lane’s and finite-element methods in soils with different
Equation 11 can be used to predict the distribution of uplift
anisotropic ratios: cutoff wall at (a) 45 m; (b) 75 m; and (c) 105 m
pressure along an impervious floor with good accuracy.
upstream of the floor
3.6 The effect of depth and situation of cutoff wall on
considered 135 samples including cases with no cutoff walls, as seepage discharge
well as cases in which the cutoff wall is installed at distances of In order to study the effect of depth of cutoff wall and its position
0, 15, 30, 45 and 60 m from the beginning of the floor. The on seepage discharge, the results of the physical model were used
depths of the cutoff wall were also considered, as 5, 15 and 25 m. to produce a series of dimensionless diagrams (Figure 13). On
For each case, we considered nine different anisotropic ratios as the basis of Figure 13 the following was concluded.
0.2, 0.25, 0.33, 0.5, 1.0, 2.0, 3.0, 4.0 and 5.0. At the second step,
using the FE method the authors determined the amount of uplift (a) In a certain depth by moving a cutoff wall along the floor
pressure in specific locations along the floor. Then, by assuming from upstream to the middle of the floor, seepage discharge
different values for the ratio of n/m, the authors calculated the increases so that it reached a maximum value in the middle

458
Water Management Uplift force, seepage, and exit gradient
Volume 166 Issue WM8 under diversion dams
Tokaldany and Shayan

3·0 No cutoff Cutoff 5 m deep, 15 m from upstream


Cutoff 5 m deep, 30 m from upstream Cutoff 5 m deep, 45 m from upstream
2·5 Cutoff 5 m deep, 60 m from upstream

2·0
MAE

1·5

1·0

0·5

0
0 1 2 3 4 5 6
n/m
Figure 9. Variations of MAE plotted against n/m ratios

6 No cutoff wall Cutoff wall in different positions

4
n/m

0
0 1 2 3 4 5
X ⫽ Ky /Kx

Figure 10. Variation of weighting factors ratio with anisotropic


conductivity ratio

of the floor. By moving the cutoff wall from the middle to 3.7 Effect of depth and position of cutoff wall on exit
downstream end, seepage discharge gradually decreases and gradient
the minimum seepage discharge will be obtained when the In Figure 14, the variation of the exit gradient against the
cutoff wall is at the downstream end. Thus, the best position variation of cutoff depth for different positions was found using
of a cutoff wall for reducing seepage discharge is at the the FE method. It can be seen that by moving the cutoff wall
downstream end. position from the upstream to the downstream end, first the exit
(b) In a certain position of cutoff wall, increasing the cutoff gradient increases and then decreases to its minimum value at the
depth results in a reduction in the seepage discharge. Based downstream end. In Figure 14, i is the exit gradient in the
on Darcy’s theory (Harr, 1962) this phenomenon can be presence of a cutoff wall, i0 is the exit gradient without cutoff
demonstrated by increasing creepage length and decreasing wall.
average hydraulic gradient.
3.8 Effect of depth and position of cutoff wall on
In Figure 13, q is seepage discharge per unit width, h is the uplift force
difference in water levels between the upstream and the down- In order to find the effects of the depth of the cutoff wall and its
stream ends, k is the hydraulic conductivity of soil, x is the position on the uplift force, the authors used the magnitude of
cutoff distance from the upstream floor, d is the cutoff wall depth, uplift pressure measured by a piezometric network. It was found
and D is the thickness of foundation. that uplift force varies by moving the position of the cutoff wall

459
Water Management Uplift force, seepage, and exit gradient
Volume 166 Issue WM8 under diversion dams
Tokaldany and Shayan

Cutoff 5 m deep in upstream Cutoff 15 m deep in upstream


Cutoff 25 m deep in upstream Cutoff 5 m deep, 15 m from upstream
Cutoff 15 m deep, 15 m from upstream Cutoff 25 m deep, 15 m from upstream
Cutoff 15 m deep, 30 m from upstream Cutoff 25 m deep, 30 m from upstream
Cutoff 15 m deep, 45 m from upstream Cutoff 25 m deep, 45 m from upstream
Cutoff 15 m deep, 60 m from upstream Cutoff 25 m deep, 60 m from upstream
2·5 Trendline

2·0

1·5
n/m

1·0

0·5

0
0 1 2 3 4 5
Ky /Kx

Figure 11. Variation of n/m ratio plotted against anisotropic


conductivity ratio

Upstream cutoff (Eq. 11) Upstream cutoff – Lane’s theory


Median cutoff (Eq. 11) Median cutoff – Lane’s theory
Downstream cutoff (Eq. 11) Downstream cutoff – Lane’s theory
1·00

0·98

0·96

0·94

0·92
RSQ

0·90

0·88

0·86

0·84

0·82

0·80
0 0·5 1·0 1·5 2·0 2·5 3·0 3·5 4·0 4·5 5·0
Ky /Kx

Figure 12. Comparison between Lane’s theory and Equation 12

460
Water Management Uplift force, seepage, and exit gradient
Volume 166 Issue WM8 under diversion dams
Tokaldany and Shayan

0·31

0·29

0·27

0·25
d/D ⫽ 3/4
q/(kh)

d/D ⫽ 5/8
0·23
d/D ⫽ 1/2
0·21 d/D ⫽ 3/8
d/D ⫽ 1/4
0·19 d/D ⫽ 1/8

0·17

0·15
0 0·1 0·2 0·3 0·4 0·5 0·6 0·7 0·8 0·9 1·0
x/b

Figure 13. Effect of depth and position of cutoff wall on seepage


discharge

2·5 x/b ⫽ 0 1·4


x/b ⫽ 1/6 1·2
2·0 x/b ⫽ 1/3
x/b ⫽ 1/2 1·0
x/b ⫽ 5/6
FP /FP0

0·8
1·5 x/b ⫽ 1 0·6
i / i0

1·0 0·4
0·2
0·5 0
0 0·1 0·2 0·3 0·4 0·5 0·6 0·7 0·8 0·9 1·0
0 x/b
0 0·1 0·2 0·3 0·4 0·5 0·6 0·7 0·8
d/D Figure 15. Changes of uplift force by position of cutoff wall with
Figure 14. Effect of depth and position of cutoff wall on exit constant depth along the floor
gradient

available methods. For the soil sample used in the experiment, it


from the upstream end to the downstream end as is shown in was found that the method of Khosla et al. gives a better
Figure 15. It should be noted that in all experiments, the length estimation of seepage effects than the creep theories of Bligh and
of the cutoff wall and the upstream head were constant and Lane. It is also found that Bligh’s theory results in better
equivalent to 20 cm. Uplift force was calculated by the integra- agreement with the observed data of uplift pressure, seepage
tion of uplift pressure distribution (by using Simpson’s 3/8 rule). discharge, and piping than Lane’s method. Using the FE method,
It can be seen that the uplift force has its maximum value at the the authors found a relationship between the anisotropic conduc-
upstream end while it reaches its lowest value at the downstream tivity ratio and the ratio of coefficients in Lane’s creep theory.
end. They found that the accuracy of Lane’s theory is reduced by
increasing the anisotropic conductivity ratio, but the new relation
4. Conclusion estimates the amount of uplift force very well. It was also found
Although there are methods for estimating the amount of seepage that the best position to control seepage and exit gradient is at the
flow, exit gradient, and distribution of uplift pressure under the downstream end of the floor and the maximum and the minimum
foundation of hydraulic structures, they present some difficulties uplift forces for the corresponding cutoff wall are located at the
when applied to engineering purposes. In this research the authors downstream and upstream ends, respectively.
carried out a series of tests on an experimental model in a
laboratory composed of a soil with anisotropic conductivity ratio Acknowledgement
nearly equal to 1. They compared the results with those obtained This paper was financially supported under contract No.
from the tests and the software with those obtained using the 7102017/1/04 by the Applied Researches Office of the Faculty

461
Water Management Uplift force, seepage, and exit gradient
Volume 166 Issue WM8 under diversion dams
Tokaldany and Shayan

College of Agriculture and Natural Resources, University of through spatially random soil. Journal of Geotechnical and
Tehran. The authors would like to express their sincere thanks for Geoenvironmental Engineering 123: 153–160.
this support. Harr ME (1962) Groundwater and Seepage. McGraw-Hill, New
York, USA.
REFERENCES Jie Y, Jie G, Mao Z and Li G (2004) Seepage analysis based on
Abedi Koupaei J (1991) Investigation of the Effective Elements boundary-fitted coordinate transformation method. Computers
on Uplift Pressure upon Diversion Dams by Using Finite and Geotechnics 31(4): 279–283.
Difference. MSc thesis, University of Tarbiat Modarres, Khosla AN, Bose NK and McKenzie ET (1936) Design of Weirs on
Tehran, Iran (in Persian). Pervious Foundations. Publication No. 12 of the Central
Ahmed AA (2011) Design of hydraulic structures considering Board of Irrigation, Simla, India.
different sheet pile configurations and flow through canal Lane EW (1935) Security from underseepage: masonry dams
banks. Computers and Geotechnics 38(4): 559–565. on earth foundations. Transactions ASCE, 100(1): 1235–
Ahmed AA and Bazaraa AS (2009) Three-dimensional analysis of 1272.
seepage below and around hydraulic structures. Journal of Leliavsky S (1965) Design of Dams for Percolation and Erosion.
Hydrologic Engineering, ASCE 14(3): 243–247. Chapman and Hall, London, UK.
Bligh WG (1910) Dams, barrages, and weirs on porous Opyrchal L (2003) Application of fuzzy sets to identify seepage
foundations. Engineering News 64 (Dec): 708. path through dams. Journal of Hydraulic Engineering, ASCE
Griffiths DV and Fenton GA (1997) Three-dimensional seepage 129(7): 546–548.

WHAT DO YOU THINK?


To discuss this paper, please email up to 500 words to the
editor at journals@ice.org.uk. Your contribution will be
forwarded to the author(s) for a reply and, if considered
appropriate by the editorial panel, will be published as a
discussion in a future issue of the journal.
Proceedings journals rely entirely on contributions sent in
by civil engineering professionals, academics and students.
Papers should be 2000–5000 words long (briefing papers
should be 1000–2000 words long), with adequate illustra-
tions and references. You can submit your paper online via
www.icevirtuallibrary.com/content/journals, where you
will also find detailed author guidelines.

462

You might also like