Em - 1110-2-1901 Seepage Analysis and PDF
Em - 1110-2-1901 Seepage Analysis and PDF
Em - 1110-2-1901 Seepage Analysis and PDF
Engineer Manual
No. 1110-2-1901 30 April 1993
1. This change replaces Appendix D, "Filter and Drain Design and Construction" of EM 1110-2-1901,
dated 30 September 1986.
2. File this change sheet in front of the publication for reference purposes.
WILUAM D. BROWN
Colonel, Corps of Engineers
Chief of Staff
EM 1110-2-1901
ENGINEER MANUAL 30 September 1986
3. Discussion. All earth and rock-fill dams are subject to seepage through
the embankment, foundation, and abutments. Concrete gravity and arch dams are
subject to seepage through the foundation and abutments. Seepage control is
necessary to prevent excessive uplift pressures, sloughing of the downstream
slope, piping through the embankment and foundation, and erosion of material
by loss into open joints in the foundation and abutments. The purpose of the
project, i.e., long-term storage, flood control, etc., may impose limitations
on the allowable quantity of seepage.
Chief of Staff
Table of Contents
CHAPTER 1. INTRODUCTION
Purpose------------------------------------ 1-1 1-1
Applicability ------------------------------ 1-2 1-1
References--------------------------------- 1-3 1-1
Objective and Scope------------------------ 1-4 1-1
General Considerations--------------------- 1-5 1-1
i
EM 1110-2-1901
30 Sep 86
ii
EM 1110-2-1901
30 Sep 86
iii
EM 1110-2-1901
30 Sep 86
APPENDIX A. REFERENCES
Government Publications--------------------- A-1
Non-Government Publications----------------- A-11
iv
EM 1110-2-1901
30 Sep 86
v
EM 1110-2-1901
30 Sep 86
CHAPTER 1
INTRODUCTION
1-1. Purpose. This manual provides guidance and information concerning seep-
age analysis and control for dams.
1-4. Objective and Scope. The objective of this manual is to provide a guide
for seepage analysis and control for dams.
1-5. General Considerations. All earth and rock-fill dams are subject to
seepage through the embankment, foundation, and abutments. Concrete gravity
and arch dams are subject to seepage through the foundation and abutments.
Seepage control is necessary to prevent excessive uplift pressures, sloughing
of the downstream slope, piping through the embankment and foundation, and
erosion of material by loss into open joints in the foundation and abutments.
The purpose of the project, i.e., long-term storage, flood control, etc., may
impose limitations on the allowable quantity of seepage (Sowers 1977).
1-1
EM 1110-2-1901
30 Sep 86
CHAPTER 2
DETERMINATION OF PERMEABILITY OF SOIL AND
CHEMICAL COMPOSITION OF WATER
v = ki (2-1)
or since Q = VA
Q = kiAt (2-2)
or using q =
q = kiA (2-3)
\where
v = discharge velocity
Q = quantity of discharge
t = time of flow
q = rate of discharge
2-1
EM 1110-2-1901
30 Sep 86
head(l) and not of pressure gradients (Harr 1962 and Bear 1972). As shown in
figure 2-1, flow is directed from point A to point B even though the pressure
at point B is greater than that at point A.
Figure 2-1. Darcy's law for flow through inclined soil column
(prepared by WES)
(1) The elevation head at any point is the distance from some arbitrary datum.
The pressure head is the water pressure divided by the unit weight of the
water. The total head is the sum of the elevation head and the pressure
head.
2-2
EM 1110-2-1901
30 Sep 86
movement of the water, as measured with dye tracers for instance, is the seep-
age velocity (Harr 1962 and Casagrande 1937) which exceeds the discharge
velocity.
(2-4)
2-3
EM 1110-2-1901
30 Sep 86
where
seepage velocity
It follows that
(2-6)
where
Reynolds number
2-4
EM 1110-2-1901
30 Sep 86
= density of fluid
The critical value-of Reynolds number at which the flow in soils changes from
laminar to turbulent has been determined experimentally by various investiga-
tors to range from 1 to 12 (Harr 1962 and Chugaev 1971). Assuming a water
-3
temperature of 20º C, substituting = 998.2 kg/m3 and µ = 1.002 x 10 kg/m
sec into equation 2-6, and assuming values of D and solving for v with
= 1 and = 12 gives the relationship shown in figure 2-3 which defines
the upper bound of the validity of Darcy's law. Depending on the discharge
velocity, Darcy's law is generally applicable for silts through medium sands.
c. Turbulent Flow.
(2-7)
where
i = hydraulic gradient
2-5
EM 1110-2-1901
30 Sep 86
2-6
EM 1110-2-1901
30 Sep 86
(2-8)
or since Q = VA
(2-9)
or using q =
(2-10)
(2-11)
where
= intrinsic permeability
2-7
EM 1110-2-1901
30 Sep 86
2-8
10-5 w-6
~u~~~~u~~?~/_fo~,r-~~a-te_r~f·~t,
1ol 2 10
2
~,?u~~~~)~·~--ul~l!~•~·~·~·~~~~~~~u·~·~·~•u_u_
10 1
__~~~~~~_u~~-L-L--~~-LL-L-~ULLL~~--~~~~~--liLuuU-L-L-~
10-
• For water at 2o•c To convert from em/sec to: m/min JJ/sec ft/sec ft/min ft/yr Darcy*
Multiply by: 0.600 10 4 0.0328 1.968 1.034 )( 10 6 1.045 )( 103
201
Figure 2-4. Permeability conversion chart (courtesy of John Wiley & Sons, Inc. )
EM 1110-2-1901
30 Sep 86
(2-12)
Equation 2-12 may be used when dealing with more than one fluid or with temper-
ature variations. This is widely used in the petroleum industry where the
presence of gas, oil, and water occur in multiphase flow systems (Freeze and
Cherry 1979, and Bureau of Reclamation 1977). In seepage analysis for earth
dams where we are primarily interested in the flow of water subject to small
changes in temperature, this refinement is seldom required.
T = kt (2-13)
where
T = transmissivity factor
k = average permeability
t = aquifer thickness
2-9
EM 1110-2-1901
30 Sep 86
2-10
EM 1110-2-1901
30 Sep 86
a
Prepared by WES.
2-11
EM 1110-2-1901
30 Sep 86
(2) The viscosity varies up to 63.5 percent over the range of tempera-
tures ordinarily encountered in seepage analysis of dams. As indicated in
equation 2-11, the permeability is inversely proportional to the viscosity of
the water. As given in table 2-1, the viscosity of water decreases as tempera-
ture increases. Therefore, the coefficient of permeability of the soil
increases as the temperature of the water increases. Permeability tests are
run at the most convenient temperature and reported at 20º C.
(3) The total dissolved salts (TDS) present in the seepage water may
influence the permeability of the soil, particularly for cohesive soils (Quirk
and Schofield 1955 and Cary, Walter, and Harstad 1943). Available data indi-
cate that cohesive soils may be two to three orders of magnitude more permeable
to seepage water containing moderate amounts of dissolved salts (less than
300 parts per million by weight) than the distilled water (Carry, Walter, and
Harstad 1943).
(2-14)
where
S = degree of saturation
= volume of water
Vv = volume of voids
(2-15)
where
k = unsaturated permeability
us
m = constant with values between 2 (uniform grain size) and
4 (well-graded materials)
2-12
EM 1110-2-1901
30 Sep 86
(2-16)
where
H = head loss
2-13
EM 1110-2-1901
30 Sep 86
(2-17)
(2-18)
while deviations from the straight line at high gradients indicate turbulent
flow. Darcy's law for a fine sand, as shown in figure 2-8, is valid only for
the hydraulic gradient less than 2 for the loose state and 4.5 for the dense
state. For soils larger than a fine sand, Darcy's law is valid for progres-
sively smaller hydraulic gradients (Burmister 1948 and Burmister 1955).
(2-19)
where
(1) As shown in table 2-2, the exchangeable cation present influences the
permeability of clay minerals at constant void ratio (Scott 1963). The
permeabilities are much smaller when the exchangeable cation is sodium
which is one of the reasons why sodium montmorillonite is used to seal
reservoirs.
2-14
EM 1110-2-1901
30 Sep 86
Hazen's experiments were made on sands for which 0.1 mm < D10 < 0.3 mm and the
uniformity coefficient, C u < 5 , where
(2-20)
where
C u = uniformity coefficient
The coefficient 100 is an average of many values which ranged from 41 to 146,
but most of the values were from 81 to 117. Equation 2-19 makes no allowance
for variations in shape of the soil particles or void ratio.
2-15
EM 1110-2-1901
30 Sep 86
2-16
EM 1110-2-1901
30 Sep 86
For compacted soils it is also observed that the permeability is much lower
(x 10-l to 10-2) in soils compacted slightly wet of optimum than in soils
compacted dry of optimum; it is thought that this occurs because of the
parallel arrangement of clay platelets in the wetter material after
compaction.
a 251
Courtesy of Addison-Wesley Publishing Company, Inc. .
2-17
EM 1110-2-1901
30 Sep 86
(2-21)
where
e = void ratio
V s = volume of solids
There are considerable laboratory test data, shown in figure 2-11, which
indicate that a plot of void ratio versus log of coefficient of permeability
is frequently a straight line (Lambe and Whitman 1969).
2-18
Table 2-3. Influence of Particle Shape and Surface Roughness on Permeability of Sand(l)
Coefficient of Permeability
Relative
Void Density em/sec
Surface c<2) cO>
Tl~e of Material Rouahness S~hericity Roundness _ u_ _c_ Ratio ~ercent Measured Com~uted-Hazen( 4 )
Crushed Sudeten granite Very rough 0.68 0.10 3.09 I. 54 0.543 79 5.00 X 10- 5 1.21 X 10- 2
Nysd Klodzkd River sand Very rough 0.75 o. 15 5.47 2.20 0.451 79 5.19 X 10- 6 5.63 X 10- 3
Odra River sand Rough to smooth 0.84 0.55 2.58 0.87 0.595 58 9.41 X 10- 4 1.44 X 10- 2
Odra River sand Rough to smooth 0.87 0.65 2.42 0.83 0.608 63 5.44 X 10- 3 1.44 X 10- 2
Baltic Beach sand Smooth 0.87 0.85 I. 91 0.89 0.720 38 1.83 X 10- 3 1.32 X 10- 2
Glass spheres Very smooth 1.00 1.00 I. 38 0.92 0.691 67 6.45 X 10-J 6.40 X
10-3
2-19
146
(!)Courtesy of British Geotechnical Society
{2)
(3)
EM 1110-2-1901
cc 0 10 0 60
X
30 Sep 86
( 4) k • 100 D •
10
EN 1110-2-1901
30 Sep 86
(2-22)
2-20
MEDIUM
0
i=
<(
~ 2.0 r-------+-------~~----~------~r-------+-------~-------+------~~------+-----~
0
0
>
2-21
7 5 1 3
10 10 10 10 6 10 10 10 10 I
EM 1110-2-1901
7. VICKSBURG BUCKSHOT CLAY 16. SILT -BOSTON 25. SODIUM MONTMORILLONITE
8. SANDY CLAY 17. SILT -BOSTON 26.-30. SAND DAM FIL TEA
9. SILT -BOSTON 18. LOESS
30 Sep 86
Figure 2-11. 201
Relationship between void ratio and log of permeability (courtesy of Wiley )
EM 1110-2-1901
30 Sep 86
2-22
EM 1110-2-1901
30 Sep 86
b. Masch and Denny Method. For uniform or nonuniform dense clean sands,
classified SP or SW in the Unified Soil Classification System (U. S. Army
Engineer Waterways Experiment Station 1960), the permeability may be estimated
from an empirical method developed by Masch and Denny 1966, and Denny 1965.
The gradation curve is plotted in Krumbein 4 units (Krumbein and Pettijohn
1938) (using the chart in figure 2-13) as shown in figure 2-14 where
(2-23)
where
(2-24)
where
2-23
10 I I I I I I I I I I I I I I I I I I I I I
30 Sep 86
EM 1110-2-1901
8
6 ~
4
~ ~
~ ~'<:,_,.,
zd
~
2-24
0 ~
-2 ~ .........._
~
-4 1 I I I I I I I I I I I I I I I I I I I I
0.01 0.1 10
GRAIN SIZE, d, mm
Figure 2-13. Conversion chart for ~ and grain size in mm, for the range 0.01 to 10 mm
(prepared by WES)
EM 1110-2-1901
30 Sep 86
b. Coefficient of permeability
versus median grain size
2-25
EM 1110-2-1901
30 Sep 86
The median grain size, d50 in units, is determined from the gradation
curve as shown in figure 2-14a. Then knowing and d50 , the coefficient
of permeability in cm per minute can be obtained from figure 2-14b (Freeze and
Cherry 1979).
(2-25)
where
k = coefficient of permeability
e = void ratio
For sands and silt-sized (finer than 0.074 mm and coarser than 0.005 mm)
particles CsTo2 = 5 is a good approximation (Perloff and Baron 1976). The
specific surface may be obtained from (Loudon 1952)
(2-26)
where
s s = specific surface
A = angularity factor
The angularity factor, A , which varies from 1.0 for glass spheres to 1.8 for
crushed glass, may be determined by microscopic examination of the soil or
estimated from table 2-4 (Loudon 1952). The specific surface of spheres,
Si , between the mesh sizes dx and dy is (Loudon 1952)
(2-27)
Subrounded 1.2
Subangular 1.3
Angular 1.4
2-27
EM 1110-2-1901
30 Sep 86
4 to 6 382
6 to 8 538
8 to 10 696
10 to 16 985
16 to 20 1524
20 to 30 2178
30 to 40 3080
40 to 50 4318
50 to 70 6089
70 to 100 8574
100 to 140 12199
140 to 200 17400
(2-28)
where
c V = coefficient of consolidation
a v = coefficient of compressibility
for permeability testing may vary depending upon whether the sample is fine-
grained or coarse-grained, undisturbed, remolded, or compacted, and saturated
or unsaturated. The permeability of remolded coarse-grained soils is deter-
mined in permeameter cylinders, while the permeability of undisturbed coarse-
grained soils in a vertical direction can be determined using the sampling
tube as a permeameter. Samples which have become segregated or contaminated
with drilling mud during sampling operations will not give reliable results.
The permeability of remolded coarse-grained soils is generally used to
approximate the permeability of undisturbed coarse-grained soils in a hori-
zontal direction. Usually the laboratory permeability of remolded coarse-
grained soils is considerably less than the horizontal permeability of the
coarse-grained soil in the field, so the approximation may not be conserva-
tive. Pressure cylinders and consolidometers are used for fine-grained soils
2-29
EM 1110-2-1901
30 Sep 86
(1) Use of samples that are not representative of actual field condi-
tions. This can be minimized by thorough field investigation, attention to
details (take undisturbed samples from test fills for determination of perme-
ability of embankment materials, sampling along faults, fissures, clay seams,
and sand partings for determination of permeability of the dam foundation),
and by the use of large samples.
(4) Air dissolved in the water. As water enters the specimen, small
quantities of air dissolved in the water will tend to collect as fine bubbles
at the soil-water interface and reduce the permeability with increasing time.
Permeability tests on saturated specimens should show no significant decrease
in permeability with time if properly deaired distilled water is used. How-
ever, if such a decrease in permeability occurs, then a prefilter, consisting
of a layer of the same material as the test specimen, should be used between
the deaired distilled water reservoir and the test specimen to remove the air
remaining in solution.
2-30
EM 1110-2-1901
30 Sep 86
snow, sleet, or hail which runs into lakes and streams or seeps into the soil
and thence into the underlying rock formations. The percolating water moves
through the saturated subsurface materials and may reappear at the surface, at
a lower elevation than the level where it entered the ground, in the form of
springs and seeps which maintain the flow of streams in dry periods (TM 5-545;
Bureau of Reclamation 1977; and Johnson Division, Universal Oil Products
1972).
b. Water Table. The surface below which the soil or rock is saturated
is the water table, as shown in figure 2-17. The water table is not a level
surface but varies in shape and slope depending upon the variations in perme-
ability and areas of recharge and discharge. In general, the water table
reflects the surface topography but with less relief. Ground water is said to
be perched if it is separated from the main water table by unsaturated mate-
rials, as shown in figure 2-17. An aquifer is a saturated permeable geologic
unit that can transmit significant quantities of water under ordinary hydrau-
lic gradients. An unconfined aquifer is one that does not have a confining
layer overlying it as shown in figure 2-17. The water table, or upper surface
of the saturated ground water is in direct contact with the atmosphere through
the open pores of the overlying material and movement of the ground water is in
direct response to gravity. The aquifer may be a layer of gravel or sand,
permeable sedimentary rocks such as sandstones or limestones, a rubbly zone
between lava flows, or even a large body of massive rock, such as fractured
2-31
EM 1110-2-1901
30 Sep 86
..• o;
c
0
•
!:'
;;;
~
,,_ ___ _
- - - _ _ _cPiezometric sur foce
--- II
II II
II II
II I I
II II
2-32
1'1 I I
II II
II II
Wate~ II II
level II II
II
Unconfined oquifer
II
I I
I I
II II
Perforated
II Well cosin9 II
II
Aquiclude
Figure 2-17. .
Groundwater relationsh1ps (courtesy of Soil Conservation Service 68 )
EM 1110-2-1901
30 Sep 86
b. Test Pits and Bore Hole Tests.. In sands and gravels above the
ground-water level, field tests are normally carried out by measuring the down-
ward seepage from test pits or shallow boreholes (Cedergren 1977). Below the
ground-water table information about the order of magnitude and variability of
the coefficient of permeability may be obtained by conducting falling head
permeability tests in the exploratory boring as drilling proceeds. The hole
is cased from the ground surface to the top of the zone to be tested and
extends without support for a suitable depth below the casing. If the per-
vious stratum is not too thick, the uncased hole is extended throughout the
full thickness, otherwise the uncased hole penetrates only a part of the
pervious stratum. Water is added to raise the water level in the casing and
then the water level descends toward its equilibrium position. The elevation
of the water level is measured as a function of time and the coefficient of
permeability is calculated (Terzaghi and Peck 1967).
(2-29)
where
k = coefficient of permeability
2-33
EM 1110-2-1901
30 Sep 86
time interval
The falling head field permeability test often gives an observed permeability
that is too low because silt particles which are suspended in the water may
form a filter skin over the walls and bottom of the hole in the pervious mate-
rial. The results of such tests are little more than an indication of the
order of magnitude of the in situ permeability. More reliable data are
obtained from field pumping tests.
2-9. Chemical Composition of Ground Water and River (or Reservoir) Water.
2-34
EM 1110-2-1901
30 Sep 86
(b)
L ..
For if; <20"-, vse L••L
5 m ~ M ~ ~ ~ ~ ~
Ratio~
ro
(a)
4000 I IT II I I I II -r If T1
2000
~
-- --
/
/
/
1000
700
=
~
/ = -
- / -
1..> ~00
'
-~
-"
,;::
2~0
-
v
v -
-...: I/
!00
"'
(3
70
~ / =
-
t-
1-- I/ -
40
1--
l/
v -
20
v
JO
7 v
v I I I II I I I l11 I I I II ~
I 2 ~ 7 /0 20 40 70 /00 200 ~ 700 1000 2000 4000
Rt~tio ~
ro
(c)
2-35
EM 1110-2-1901
30 Sep 86
2-36
EM 1110-2-1901
30 Sep 86
CHAPTER 3
DETERMINATION OF PERMEABILITY OF ROCK
3-1
EM 1110-2-1901
30 Sep 86
(3-1)
or
(3-2)
3-2
EM 1110-2-1901
30 Sep 86
where
Since the two equations are identical, resolving them indicates that m =
and The use of the above laws, however, is usually restricted
to a homogeneous, isotropic, porous continuum. Since in soil and rock masses
there are complex systems of interconnected void spaces, an equivalent rather
than an absolute permeability should be determined. The coefficient of equiv-
alent permeability from the continuum approach assumes that flow occurs uni-
formly throughout the mass rather than within individual passageways. There-
fore, for equivalent permeability, Darcy's law and Missbach's law are written,
respectively,
(3-3)
and
(3-4)
where
The continuum approach, in some cases, is not applicable and therefore, the
discontinuum method of analysis for evaluating an equivalent permeability
should be used. Formulae for the discontinuum analysis for equivalent
permeability are presented later with various other methods of analysis.
3-3. Methods for Determining Rock Mass Permeability. Numerous methods have
been developed for determining or estimating rock mass permeabilities. All of
the available testing, as well as analytical techniques should be considered
and evaluated for each individual study, in order to optimize the advantages
and minimize the disadvantages inherent within each method for determining the
permeability of a rock mass.
3-3
EM 1110-2-1901
30 Sep 86
(3-5)
where
(3-6)
Comparison of the flow velocity equation with Darcy's law indicates that the
parallel plate permeability, kp , can be expressed by
(3-7)
where
3-4
EM 1110-2-1901
30 Sep 86
(3-8)
where Thus, it follows that the flow rate per unit width
becomes
(3-9)
(3-10)
(3-11)
where
3-5
EM 1110-2-1901
30 Sep 86
Tests of individual fissures can also be analyzed for turbulent flow according
to the Missbach law
(3-12)
where the volume flow rate per unit width can be expressed as
(3-13)
(3-14)
(3-15)
3-6
EM 1110-2-1901
30 Sep 86
(3-16)
where
For S < 0.033 equations for laminar equivalent permeability have been devel-
oped which yield
(3-17)
(3-18)
3-7
EM 1110-2-1901
30 Sep 86
(c) Tracer Tests. Tracer tests involve the injection of an inert solu-
tion, or tracer, into an existing flow field via a borehole or well. Tracer
tests are often desirable because they are passive-type tests and do not place
unnatural stress conditions on the flow system. The dilution rate of the
tracer at the injection well or its time of travel to another well can be used
to calculate the ground-water velocity and ultimately the permeability. Detec-
tion of the tracer, or concentration measurements, can be made by either
3-8
EM 1110-2-1901
30 Sep 86
(3-19)
or
(3-20)
With the hydraulic gradient determined from observation wells in the area, the
permeability can be computed directly. For tracer tests the seepage velocity
is determined according to the equation
(3-21)
where
(3-22)
(3-23)
(3-24)
3-9
EM 1110-2-1901
30 Sep 86
(3-25)
where
w d = well diameter
3-10
EM 1110-2-1901
30 Sep 86
unsteady-state conditions and for using either air or water for injection.
Methods of analysis of pressure injection tests are presented in Appendix C.
(a) Water Pressure Tests. Water pressure tests, also known as packer
tests (in Europe they are called Lugeon tests) are normally conducted by
pumping water at a constant pressure into a test section of a borehole and
measuring the flow rate. Borehole test sections are commonly sealed off by
one to four packers, with the use of one or two packers being the most widely
used technique. In comparison with a pumping test, a water pressure test
affects a relatively small volume of the surrounding medium, because frictional
losses in the immediate vicinity of the test section are normally extremely
large. The test, however, is rapid and simple to conduct, and by performing
tests within intervals along the entire length of a borehole, a permeability
profile can be obtained. Additionally, the water pressure test is normally
conducted in NX boreholes, and has the advantage of being conducted above or
below the ground-water table.
(b) Air Pressure Tests. Air pressure tests are similar to water pres-
sure tests except that air rather than water is used for the testing fluid.
The air pressure test was developed for testing above the ground-water table
and has predominantly been used for testing areas of high permeability such as
those characteristic of rubblelike, fallback material adjacent to explosively
excavated craters in rock. In such areas, water pressure tests have been
inadequate due to an inability to provide water at a flow rate high enough to
pressurize the surrounding media. Air pressure tests have an unlimited supply
of testing fluid, as well as the advantage of a wide variety of high capacity
air compressors. The disadvantage of such tests is that permeability equations
must be modified for application to a compressible fluid and a conversion from
the air permeability to a water permeability must be made to obtain usable
results.
(c) Pressure Holding Tests. Pressure holding or pressure drop tests are
usually conducted in conjunction with water pressure tests. The test is
analogous to the falling head test used in soil mechanics; however, in rock
the test section is normally pressurized to a value above that of the static
head of water between the test zone and the ground surface. The pressures are
normally measured in the test section with a transducer, and pressures versus
time are recorded. The pressure holding test offers the advantage of being
quick and simple to perform, as well as requiring significantly less water
than that used in conventional constant pressure tests.
(2) Flow Rates. Given the permeability and gradient, the flow rate
across a given area can be computed. With the effective porosity, the rate of
advance of a seepage front can be determined. Ground-water flow rates are
required for determining construction dewatering requirements for both surface
and subsurface excavations, as well as the seepage losses through dam founda-
tions and abutments.
(1) Index Tests. The determination of rock mass permeability for dam
foundations and abutments has historically been used more as an index test
than as an absolute test. As the state of the art advances this trend is
gradually changing. While index tests are a valuable tool to the experienced
foundation engineer, more complex methods of analysis are becoming available
3-12
EM 1110-2-1901
30 Sep 86
3-13
EM 1110-2-1901
30 Sep 86
CHAPTER 4
SEEPAGE PRINCIPLES
c. Entrances and Exits. The lines defining the area where water enters
or leaves the pervious soil mass are known as entrances or exits, respec-
tively. Along these lines (O-l and 8-G in figure 4-1(a) and AD and BE in
figure 4-1(b)) are lines of equal potential; that is, the piezometric level is
the same all along the line regardless of its orientation or shape. Flow is
perpendicular to an entrance or exit. Entrances and exits are also called
reservoir boundaries (Harr 1962).
4-1
EM 1110-2-1901
30 Sep 86
4-3. Confined and Unconfined Flow Problems. Two general cases of seepage are
considered in this manual: confined and unconfined flow. Confined flow
exists in a saturated pervious soil mass which does not have a line of seepage
boundary. Figure 4-1(a) is an example of confined flow. Unconfined flow,
figure 4-1(b), exists when the pervious soil mass has a line of seepage. Thus
confined flow has all boundaries defined while for unconfined flow the surface
of seepage and line of seepage must be defined in the analysis.
4-2
EM 1110-2-1901
30 Sep 86
(1) Heads h and h are constant and thus flow is steady state.
1 2
(2) Water is incompressible.
(5) The element has a dimension, dy , into the plane of the figure
which gives an element volume but no flow takes place perpendicular to the
plane of the figure, i.e., the flow is two-dimensional.
(where Setting
equal to gives:
4-3
EM 1110-2-1901
30 Sep 86
(4-1)
Using Darcy's law, v = ki and assuming the same permeability in the x and
z directions:
and
(4-2)
4-4
EM 1110-2-1901
30 Sep 86
4-5
EM 1110-2-1901
30 Sep 86
I SEEPAGE ANALYSIS I
I J J 1
FLOW NETS MODELS ANALYTICAL METHODS NUMERICAL AND
COMPUTER METHODS
FRAGMENTS 1,2,3-DIMENSIONAL
(2) DRY 2-D
b. SAND c. CLOSED FORM
2-D OR 3-D
c. VISCOUS FLOW
2-0
66
Figure 4-3. Seepage analysis methods (from Radhakrishnan )
EM 1110-2-1901
30 Sep 86
(2) Sand models which may use prototype materials can provide informa-
tion about flow paths and head at particular points in the aquifer. The sand
or porous material may be placed underwater to provide a homogeneous condi-
tion, or layers of different sand sizes may be used to study effects of inter-
nal boundaries or layers. If the flow is unconfined and the same material is
used for model and prototype, the capillary rise will not be scaled and must
be compensated for in the model. Flow can be traced by dye injection and
heads determined by small piezometers. Disadvantages include effects of
layering when the porous material is placed, difficulty in modeling prototype
permeability and boundary effects. Prickett (1975) provides examples of sand
tank models and discusses applications, advantages, and disadvantages.
(3) Viscous flow models have been used to study transient flow (e.g.,
sudden drawdown) and effects of drains. This method depends on the flow of a
viscous fluid such as oil or glycerin between two parallel plates and is nor-
mally used to study two-dimensional flow. As with sand models, dye can be
used to trace flow lines.' Construction is normally complicated and operation
requires care since temperature and capillary forces affect the flow. Flow
must be laminar, which can be difficult to achieve at the boundaries or at
sharp changes in boundary geometry.
b. Analytical Methods.
4-7
EM 1110-2-1901
30 Sep 86
(3) Closed form solutions exist for simpler seepage conditions such as
flow to a fully penetrating well with a radial source (Muskat 1946). Seepage
problems associated with dams typically require approximate solutions because
of complicated flow conditions.
4-8
EM 1110-2-1901
30 Sep 86
4-6. Graphical Method for Flow Net Construction. Flow nets are one of the
most useful and accepted methods for solution of Laplace's equation (Casagrande
1937). If boundary conditions and geometry of a flow region are known and can
be displayed two dimensionally, a flow net can provide a strong visual sense
of what is happening (pressures and flow quantities) in the flow region.
Equation 4-2, paragraph 4.4, is an elliptical partial differential equation
whose solution can be represented by sets of orthogonal (intersecting at right
angles) curves. One set of curves represents flow paths of water through the
porous media while curves at right angles to the flow paths show the location
of points within the porous media that have the same piezometric head. The
former are called flow lines, the latter equipotential lines. The flow net is
a singular solution to a specific seepage condition, i.e., there is only one
family of curves that will solve the given geometry and boundary conditions.
This does not mean that a given problem will have only one flow net--we may
choose from the family of curves different sets of curves to define the
problem, figure 4-5. The relationship between the number of equipotential
4-9
EM 1110-2-1901
30 Sep 86
a. Net drawn for four flow b. Net drawn for five flow
channels. channels.
4-10
EM 1110-2-1901
30 Sep 86
b. Guidelines for Flow Net Drawing. Once the section of porous media
and boundary conditions are determined, the flow net can be drawn following
general guidelines:
(a) Flow will be along and parallel to impermeable boundaries lines BCD
and FG, figure 4-5.
(b) Entrances and exits are equipotential lines, lines AB and DE,
figure 4-5, with flow perpendicular to them.
(d) Entrance and exit conditions for a line of seepage are shown in
figure 4-7 under "Conditions for Point of Discharge."
(2) Equipotential and flow lines must meet at right angles and make
curvilinear squares. Usually, it is best to make either the number of flow
channels a whole number (if the number of flow channels is a whole number, the
number of equipotential drops will likely be fractional).
(3) Generally, a crude flow net should first be completed and adjust-
ments applied throughout the net rather than defining one portion since
refinement of a small portion tends to shift the whole net.
(5) If, in the finished flow net, either equipotential drops or flow
channels end up as a whole number plus a fractional line of squares (equi-
potential drop or flow channel), this should not be a problem but must be used
in any calculations based on the flow net. It is convenient to locate a par-
tial equipotential drop in an area of uniform squares since this will make
accurate estimation of the fraction easier.
4-11
EM 1110-7-1901
30 Sep 86
(6) Use only enough flow lines and equipotential lines to bring out
flow net definition. If more information is needed in particular areas, the
squares may be subdivided into smaller squares for more detail of flow and
pressure distribution.
(8) Figures 4-7 and 4-8 provide some guidelines for entrances and exits
and particular areas within the flow region.
4-12
EM 1110-2-1901
30 Sep 86
(9) It is helpful to lay out the boundaries which will contain the flow
net in ink and use a soft pencil and eraser to develop the flow net to final
form.
4-7. Flow Net for Anisotropic Soil. Most naturally occurring soils and many
man-placed soils have greater horizontal permeabilities than vertical. This
affects the shape of a flow net since the flow net provides a solution to
Laplace's equation which is based on the assumption of an isotropic porous
media (paragraph 4.4b). To compensate for anisotropy, the dimensions of the
porous media are changed by the square root of the ratio of the two perme-
abilities. If kh is the horizontal permeability and kv is the vertical
permeability, then the horizontal dimensions of the porous media cross section
are changed by a ratio of , e.g., if the base of a dam is 300 feet,
then it would be changed by a factor of , or would be 300 feet times
The same ratio would be applied to all other horizontal dimensions
to produce a transformed section. Next the flow net is drawn on the trans-
formed section, as described in paragraph 4-6. Then the section, including the
flow net, is returned to the original (true section) which produces a nonsquare
flow net. Computations are made using the nonsquare flow net just as a square
flow net is used for isotropic conditions. This procedure is illustrated in
figures 4-9 and 4-10. In the same manner, dimensions in the vertical direction
could be changed by the factor , square or normal flow net drawn
on the transformed section, then returned to true section. Pore pressure
distribution and hydrostatic uplift may be taken from either section while
gradient and magnitude of seepage forces must be determined from the true
section.
4-8. Flow Net for Composite Sections. Commonly, projects requiring seepage
analysis involve different soils with different permeabilities, e.g., strati-
fied foundation materials and zoned dams. Certain rules apply to flow lines,
equipotential lines, and lines of seepage crossing internal boundaries between
soils of different permeabilities. Figure 4-11 illustrates the deflection of
flow lines and equipotential lines at interfaces. The essential principle is
that the more permeable soil allows the same amount of water to flow with less
restriction, thus drops in potential within the higher permeability soil will
be farther apart (i.e., less energy loss in the higher permeability soil for
the same length of flow as in the low permeability soil). It should be noted
in figure 4-11 that when flow goes from lower permeability soil to higher
permeability soil, the distance between flow lines decreases (flow channel gets
smaller) and the distance between equipotential drops increases, Figures 4-12
through 4-14 are examples of flow net construction for seepage through soils of
4-14
EM 1110-2-1901
30 Sep 86
Figure 4-9. Flow nets constructed on transformed section and redrawn on true
155
section (courtesy of John Wiley and Sons )
4-15
--------r---------1--------------
30 Sep 86
EM 1110-2-1901
40
,., ~
io u Stole (fl!ct
H ~
.s=S
flow t honnel
Notes• I. llorl,onlol Ira dormolion
futtur -v'li~ik "VIiiu() '0.1
1
2. k = v'(ky)(lch
3. Nrtu,, · 6.6!j 10 =o.GGS
4-16
4. q ,tt'h(t~r!N~,J)=k'(36X0665)
=7j 9k
__ ______ -----
:}/ Noles• I. Toe or~o mu I be enlor!Jed
to ol.lhJin vul e or l\ L.
2.
Figure 4-10. Example of technique for developing a flow net for an anisotropic foundation (from
123
U. S. Department of Agriculture )
[~ = b I
k2
d k.J tan
c = k2 tan
----Flow lines
---Equipotential lines
EM 1110-2-1901
30 Sep 86
EM 1110-2-1901
30 Sep 86
a. Identify boundaries
4-18
EM 1110-2-1901
30 Sep 86
4-20
EM 1110-2-1901
30 Sep 86
4-21
EM 1110-2-1901
30 Sep 86
Then:
which gives the quantity of seepage flow for each unit of thickness of porous
media perpendicular to the plane of the flow net. Figure 4-10(b) gives an
example of this calculation for anisotropic seepage conditions in a dam
foundation. The permeability, k' , used for anisotropic conditions,
k' = is derived by Casagrande (1937).
4-22
ttcad Water Pnuure. 1lonu 111•1•P.r liniN IJ,.unriJrv
lbollout ol w"h I
Ground Su~·-~..__ _ _ _ _ _ , _ __
'N.
CP . Ql
GJ (\)
-.
Ol
w
-- It
I ul ar
4-23
!TTTTr
·------·.-'
• Fu!l'l
-- ------
l~u
ol dropa
----
fhwd lou • Upllll preasur t
I desivnu!lon to point l~h • flo. drops) ( IB.B'- hd.lon) Y. E1cap1 gredlent 11 tocwoll end lram~·· ''" sill!
\!lj ___ (p_ s rI "
5,5 5,1 85 5
-----
5.6
---67 ---6.5 824
768
- ---
EM 1110-2-1901
7,6 7.2 72~L ___
---·--·-- --- -
__ ---
-F---e
E 0.0
.. fl. 2 661
30 Sep 86
------·~---
9.5 0.8 62 4
Figure 4-15. Base uplift and escape gradient for concrete drop spillway
. 1 ture 123)
(f rom U• S . Department o f A gr1cu
EM 1110-2-1901
30 Sep 86
unit weight of the soil and is the gradient at which upward drag forces on the
soil particles equal the submerged weight of the soil particles, figure 4-16.
The critical gradient is dependent on the specific gravity and density of the
soil particles and can be defined in terms of specific gravity of solids, Gs ,
void ratio, e , and porosity, n :
ranges for factor of safety for escape gradient, FSG from 1.5 and 15,
4-24
EM 1110-2-1901
30 Sep 86
may be developed by placing very pervious material on the exit face, which will
allow free passage of water but add weight to the exit face and thus add
downward force. This very pervious material must meet filter criteria to
prevent loss of the underlying soil through the weighting material.
4-25
EM 1110-2-1901
30 Sep 86
seepage force acts in the same direction as flow, i.e. along flow lines. Con-
sider the seepage force on plane A-A in figure 4-16. Since flow is vertically
upward, the direction of seepage force is up, the difference in piezometric
head is h , and the area perpendicular to flow is A . The seepage force,
Fs , is the part of the upward forces due to differential head, h , or:
Two methods of applying this force to use in stability analysis are described
by Cedergren (1977) and termed the gradient method and boundary pressure
method. EM 1110-2-1902 gives examples of embankment stability analyses
considering seepage forces. Additionally, the effect of buoyant forces on soil
mass stability must also be considered. The upward or buoyant force, Fb ,
causing reduction in effective stress on plane A-A, figure 4-15, is the
remainder of the upward forces on plane A-A:
4-26
EM 1110-2-1901
30 Sep 86
CHAPTER 5
CONFINED FLOW PROBLEMS
5-2. Gravity Dam on Pervious Foundation of Finite Depth. Figure 5-1, a copy
of figure 4-15, provides an example of an impervious structure on a pervious
foundation of finite depth with calculation of uplift and escape gradient.
Figure 4-12 illustrates a gravity dam on a composite pervious foundation of
finite depth and figure 4-10 shows an example of an impervious dam (though not
a gravity dam) on a finite anisotropic foundation. A cross section of a
classical gravity dam/weir, figure 5-2 indicates the effect of a partially
penetrating cutoff placed beneath the upstream portion of the dam. Comparing
figure 4-10(b) with figures 5-1 and 5-2 will show the reduction in gradient at
the downstream toe caused by embedment of the structure in the pervious
foundation. Embedment provides a longer upper flow line for a given structure
width and reduces the gradient at the downstream toe of the structure. Other
measures to reduce uplift and/or high gradients at the downstream toe include
cutoffs beneath the downstream portion of the dam and placement of drains
beneath the downstream portion of the dam.
5-3. Gravity Dam on Infinitely Deep Pervious Foundations. This case, illus-
trated by figure 5-3, is symmetrical (if there are no asymmetrical cutoffs or
drains beneath the dam) and at large distances becomes a series of half circle
arcs for flow lines and radial lines for equipotential lines. Since the up-
stream (entrance) and downstream (exit) boundaries extend to infinity and the
foundation depth is infinite , the flow quantity is infinite. Obviously the
analyst must decide the limits of the problem and calculate flow quantity
based on those limits. The same manipulations and effects of embedment, cut-
off, and drain described for the previous case will apply to this case.
5-1
lleod Woler
EM 1110-2-1901
30 Sep 86
_ _ _V..___ _ _ _ _ _ _ _~--- - -
Ground
Toil Wuler
lrurrervious
Uountlor y \
77.11111 I rrr I 11 TTirrTT~
5-2
13.8;
----
Porn! No. of drops Ueod loss : Uplift pressure
desiqnotio11 to point 11\h a No drops) ( 18. a'- hd loss) Y. Escape yr adient at torwall and tramverse sill:
0
(fl.) (ps.U .~
A "5,5 5.1 855
--·----
B 6 5.6 824
1---·-- ---·----
c 7 6.5 768
f----··
0
·------
l.8
---·-
l.2 724
-- --
E 8.8 8.2 G61
-·---- -------- ---- -- --
F 9.5 8.8 62 'l
··-
Figure 5-l. Base uplift and escape gradient for concrete drop spillway (from
123
U. S. Department of Agriculture )
EM 1110-2-1901
30 Sep 86
5-3
EM 1110-2-1901
30 Sep 86
CHAPTER 6
UNCONFINED FLOW PROBLEMS
6-1. Introduction. This chapter will consider unconfined flow problems for
cases involving earth dams. Because of their ability to give a strong visual
sense of flow and pressure distribution, flow nets will be used to define seep-
age. Other methods, such as transformations (Harr 1962), electrical analogy
models, and numerical methods, can provide pressures and flows and be used to
develop flow nets. Unconfined flow problems require the solution of flow and
pressure distribution within the porous media and definition of the line of
seepage boundary (phreatic surface within the dam).
6-2. Homogeneous Earth Dam on Impervious Foundation. The simplest earth dam
configuration consists of a homogeneous, pervious embankment on an impervious
foundation. Though rarely encountered in engineered embankments, this case
will introduce general methods of defining flow in embankments.
Figure 6-1. Line of seepage, BC, and seepage exit face, CD, for a
homogeneous earth dam on an impermeable foundation (prepared by
WES)
6-1
EM 1110-2-1901
30 Sep 86
I
/I
h
II
I I
r I
- ........
........ ('
.
........................
I I .......
tl
METHOD . EQUATIONS
I SCHAFFERNAK -
! a = _d- - j d2 - h2
I cos 0: cos2 0: ~
VAN ITERSON I
1 q • k .
a sm o: tan 0:
; L. CASAGRANDE
a= S -
o
J S 2 - -h2 -
o sin2 0:
i q = k a sin2 0:
~ KOZENY i
I a
Yo 1
=- =- [ VI d 2 + h
2
- d
·
l
I 0 2 2
I
I
I q = 2ka 0 = ky 0
6-2
EM 1110-2-1901
30 Sep 86
b. After this is done one of the four methods shown in figure 6-2 and
explained below can be used to determine the location of the exit face CD and
the line of seepage BC.
(1) < 30º Schaffernak-Van Iterson. The two formulas for this method
given in figure 6-2 assume gradient equals dy/dx and allow direct determi-
nation of a and q . Construction of basic parabola shown in figure 6-3 is
the first step in determining the upper line of seepage (Casagrande 1937).
From embankment geometry and headwater height, point A is located. d and y
are determined by scribing an arc, with radius DA through point E. Then the
point of vertical tangency of the basic parabola, F, is determined. Line AG,
parallel to the embankment base and horizontal axis of the parabola, is drawn
and divided into an equal number of segments (6 in the case in figure 6-3).
Line GF, the vertical tangent to the parabola, located at yo/2 from the
downstream toe of the embankment is divided into the same number of equal
segments as line AG. The points dividing line AG into segments are connected
with point F. The intersection of these lines with their counterpart lines
drawn from the points on line GF define the parabola. Thus the basic parabola,
dashed line A-F, is defined. The upstream portion of the line of seepage,
dotted line BH, is drawn by starting at point B perpendicular to the upstream
slope (since the upstream slope is an equipotential line and the line of
seepage is a flow line) and continuing downstream to make line BH tangent to
the basic parabola at point H which is selected based on judgment. This is an
entrance condition as shown in figure 4-7. The central portion of the line of
seepage is along the basic parabola while the downstream portion is a smooth
transition from the basic parabola to tangency with the downstream slope at
point C. Point C is located a distance a. from the downstream toe as
determined by the equation for Schaffernak-Van Iterson shown in figure 6-2.
With all seepage boundaries known and using the rules of Chapter 4, a flow net
may be constructed within the boundaries as shown in figure 6-4. This figure
points out the important basic flow net requirement that all equipotential
lines intercept the line of seepage and exit face at points with equal verti-
cal separation (in this case H/10 apart).
6-3
EM 1110-2-1901
30 Sep 86
(2) < 90° L. Casagrande. The gradient assumption for this method is
i = dy/ds where s is the distance along the line of seepage, and allows
greater accuracy than Schaffernak-Van Iterson method for steeper downstream
slopes. Use of the equations in figure 6-2 and the same general procedures
used for the Schaffernak-Van Iterson method apply for up to 60º. For
60° < < 90°, since a and s are interdependent, the location of point C
(or distance a) must be estimated to determine the value of so
then distance a calculated. This procedure is repeated until there is
satisfactory agreement between the portion of the distance so as measured
and a as calculated. Thus the seepage boundaries are established allowing
flow net construction.
(3) = 180° Kozeny. For this special case Kozeny described a solution
adapted by Casagrande (1937). Figure 6-5 illustrates the nomenclature and
construction method for this case. Embankment geometry, h , and drain loca-
tion control construction of the basic parabola. For this case the seepage
face is the distance a0 and the correction is not used. Again with
boundary definition, the flow net can be drawn.
(4) 30° < < 18.0° A. Casagrande. After study of model experiments
and construction of flow nets for various , A. Casagrande (1937)
developed a curve, figure 6-6, which relates a to the ratio,
Construction of the basic parabola is the first step in this procedure. The
point, Co , as shown in figure 6-2, where the basic parabola intercepts the
downstream slope is determined and distance a + is measured. Knowing
C can be found in figure 6-6 and calculated. Information is then suf-
ficient to draw the line of seepage and discharge face, determine q , and con-
struct the flow net. Casagrande (1937) provides a procedure for the condition
of tailwater on the downstream slope:
6-4
EM 1110-2-1901
30 Sep 86
6-5
EM 1110-2-1901
30 Sep 86
the intersection of the crest and upstream slope (y axis), Zone II between the
y axis and a vertical line at the intersection of the line of seepage with the
downstream slope, and Zone III which is composed of the remainder of the
downstream toe. Pavlovsky assumed horizontal flow in each zone and wrote the
basic equation q for each zone using the nomenclature of figure 6-7.
The equations for each zone are:
Zone I
(6-1)
6-6
EM 1110-2-1901
30 Sep 86
Zone II
(6-2)
Zone III
for ho > 0
(6-3)
for ho = 0
(6-4)
(6-5)
(6-6)
then a plot of ao versus h1 may be made of equations (6-5) and (6-6). The
intersection of the two curves representing (6-5) and (6-6) is the value of a0
and h 1 for solution. Equation (6-4) will then provide q . An example from
Harr (1962) is provided in figure 6-8.
6-7
EM 1110-2-1901
30 Sep 86
A. Pavlovsky's solution
50 19.6 50 15.9
52.5 18.7 52.5 17.7
55 17.4 55 19.7
60 13.9 60 24.1
65 8.4 65 29.3
6-8
EM 1110-2-1901
30 Sep 86
a = 3ao = 54.9 ft
h W = 70 ft
a = 73 ft
C. L. Casagrande solution:
a = 76 ft
2
q = (0.002 ft/min)(76 ft sin 18º26')
3
q = 0.015 ft /min per ft of embankment length
6-9
EM 1110-2-1901
30 Sep 86
6-3. Earth Dam with Horizontal Drain on Impervious Foundation. Figure 6-5
presents this case for a homogeneous embankment. However, since most earth
dams are built in horizontal layers, they very likely have a stratified
structure which may allow considerable flow to bypass the horizontal drain,
figure 6-9. The difference in vertical and horizontal structure also causes
differences in horizontal and vertical permeabilities. This can strongly
affect the location of the upper line of seepage, figure 6-10, which affects
stability considerations and methods of controlling seepage.
6-4. Earth Dam with Toe Drain on Impervious Foundation. Toe drains are
another method of controlling the line of seepage, figure 6-11. Again the
effects of anisotropy must be considered, figure 6-12. The geometry of the
embankment and the toe drain, height of reservoir, and the degree of anisotropy
will control the location of the line of seepage.
6-5. Earth Dam with Vertical or near Vertical Horizontal Drains on Impervious
Foundation. One very effective method of intercepting horizontal flows due to
stratification of the embankment, figure 6-9, is the incorporation of an
inclined or vertical drain into the central portion of the embankment, fig-
ure 6-13. This seepage analysis of a zoned, anisotropic embankment assumed
the rockfill to have infinite permeability with respect to the core materials
and used the method recommended by A. Casagrande for drawing the parabola to
determine the upper line of seepage. The interface of the core and inclined
drain is used as the downstream slope for the seepage face since the drain has
a much higher permeability than the core material. Provision must be made in
sizing the drain to pass all the water coming out of the core without building
up a tailwater on the downstream slope of the core. It can be noted that the
designers of this example used m/3 instead of 0.3m to determine the intercep-
tion point of the parabola and headwater elevation and the formula
6-10
EM 1110-2-1901
30 Sep 86
6-11
30 Sep 86
EM 1110-2-1901
£DAM
I
6-12
I
~CURTAIN GROUTING
I
TRANSFORMATION
EMBANKMENT SEEPAGE
KH ~ 4Kv, TRANSFORMATION r ACTOR--~ ='f'[ = _!.""' 0.5
v~ v; 2
TRUE SECTION
FOR NON-ISOTROPIC SOIL THE EFFECTIVE COEFFICIENT OF PERMEABILITY
TRANFORMEO SECTION HORIZONTAL DIMENSION= 0.5 TRUE SECTION HORIZONTAL DIMENSION
K' = KvKH, SINCE THE EMBANKMENT IMPERVIOUS Fill MATERIAL IS
Figure 6-13. Seepage analysis for Kettle Creek earth embankment, Kettle Creek, Pa. (from
78
U. S. Army Engineer District, Baltimore )
£DAM
F/3•36.~ I'
MAX WATER SURFACE EL 962.7 "
I
:I:
:TRANSFORMED SECTION
6-13
FLOW NET
LEFT ABUTMENT-STATION 17+00
TRANSFORMATION EMBANKMENT SEEPAGE
EM 1110-2-1901
yo
a+Jl.a = -COS a , TAN a • 2, a=2, a=63' -26', COS a • 0.447
1
47.48 Jl.a
30 Sep 86
a+Jl.a • _ = 85.86; a+Jl.a • 0.32 (AFTER A. CASAGRANDE)
0 553
drain are considered infinitely permeable with respect to the central earth
portion of the embankment. In some cases dams may contain adjoining zones
which, have relatively small but marked differences in permeability and it may
be desired to accurately analyze the flow through these zones. Equipotential
lines and flow lines, with the exception of the upper line of seepage, will
cross the interface of the zones in the same manner as given for confined
composite sections in Chapter 4, but the location of the upper line of seepage
(phreatic line) must first be determined. Once its location is determined the
upper line of seepage transfers between regions of different permeabilities in
the manner shown in figure 6-14. In determining the location of the upper
seepage line, the essential principle to remember is that the flow rate must
be the same through each zone. That is, for a unit depth of embankment per-
pendicular to the plane of the flow net, Q must be the same for each zone.
More permeable zones require less gradient and/or cross-sectional area to pass
the flow transmitted to them from less permeable zones. This idea can be seen
in the example and instructions taken from Cedergren 1977.
a. Locate the reservoir level and the tailwater level, noting the
difference in head as h and dividing h into a convenient number of equal
parts of increments Draw a series of light horizontal guide lines (head
lines) at intervals of across the downstream part of the section
(figure 6-15a).
6-14
EM 1110-2-1901
30 Sep 86
b. Guess a trial position for the phreatic line in both zones and draw a
preliminary flow net as shown in figure 6-15a, making squares in zone 1 and
rectangles in zone 2. Make the length-to-width ratios of all of the rectangles
in zone 2 approximately equal by adjusting the shape of the saturation line,
using an engineer's scale to measure the lengths and widths of the figures.
When this step is completed, the trial flow net should be reasonably well
drawn. It should satisfy the basic shape requirements of a flow net, but the
length-to-width ratio of the shapes in zone 2 probably will not satisfy
c/d = k2/k1 . Although the flow net has been drawn for a composite section,
the ratio of k2/k1 probably does not equal the k2/k1 ratio originally
assumed for the section.
c. Calculate the actual ratio of k2/k1 for the trial flow net just
constructed. To make this important check proceed as follows:
(1) Count the number of full flow channels between any two adjacent
equipotential lines in zone 1 and call this number nf-1. In the trial flow
net in figure 6-15a , n f - 1 = 4.0 .
(2) Count the number of full flow channels between any two adjacent
equipotential lines in zone 2 and call this number nf-2 . In figure 6-15a,
nf-2 is equal to the width-to-length ratio of the figures in zone 2, d/c ,
and equal to 0.5.
(3) The actual value of k2/k1 for the trial flow net in figure 6-15a
can now be determined from the equation
6-15
EM 1110-2-1901
30 Sep 86
or
(4) If the calculated k2/k1 ratio is too high, the saturation line in
zone 2 is too low and must be raised. If the calculated k2/k1 ratio is too
low, the saturation line in zone 2 is too high and must be lowered. Raise or
lower the general level of the saturation line in zone 2 as indicated and
construct another trial flow net.
(5) Repeat steps (1) through (4) until a flow net of the desired
accuracy is obtained. (Usually a few trials will be sufficient.) By applying
the above equation to the first trial flow net in this example (figure 6-15a)
k2/k1 = 4.0/0.5 = 8.0. Because the ratio of k2/k1 for this example was
assumed to be 5, the k2k1 ratio of the trial flow net is too high; hence the
general level of the saturation line in zone 2 is too low and must be raised.
For the second trial flow net (figure 6-15b) nf-1 = 3.5 and nf-2 = 0.7 .
The calculated ratio of k2/k1 = 3.5/0.7 = 5.0 , the value originally assumed.
The above equation may be derived by recalling that the quantity of seepage in
zones 1 and 2 (figure 6-15) must be equal. Using
6-16
EM 1110-2-1901
30 Sep 86
(1) The flow net for the foundation was drawn independently of the
embankment, i.e. considering the embankment to be impermeable, and assuming the
foundation to be isotropic.
(2) The embankment flow net reflects the influence of foundation flow
net in the location of equipotential and flow lines and is drawn for aniso-
tropic conditions.
6-17
EM 1110-2-1901
30 Sep 86
6-18
EL 602
6-19
k-->oo
BUCHANAN DAM
H. R. CEDERGREN
FLOW NET STUOIES
3-18-75 FULL RESERVOIR - STA 9+50
DIS TRANSITION FILL- kh = 16kv
EM 1110-2-1901
Figure 6-18. Buchanan Dam flow net studies full reservoir - Sta 9+50 D/S
25
30 Sep 86
transition fill - ~ = 16 kv true section (from Cedergren and U. S.
Army Engineer District, Sacramento 1977)·
30 Sep 86
EM 1110-2-1901
SEEPAGE PER FOOT OF EMBANKMENT LENGTH
0,. 1< 1 h S,. 7.5 X 10-8 CM/SEC X 192FT X 15/8 X 60 SEC/MIN X 1/30.48 FT/CM X 7.48 GAL/FT 3
0 a 0.00040 GAL/MIN PER FOOT OF EMBANKMENT
FOR 600FT EMBANKMENT
QTOTAL,. 600FT X 0.00040 GAL/MIN/FT E 0.24 GAL/MIN
DISTANCE. FT FROM CENTER LINE OF DAM
900 BOO 700 600 500 400 300 200 100 100 200 300 400 500 600 700
650 650
~
( K : 150 x 10-B CM/SEC TO 'THE CORE FLOW NET,
"
,..
6-20
ROCK
~ GROUT CURTAIN~A~O AND GRAV~L DRAIN ROCK TAILWATER
::
a. IMPERVIOUS FOUNDATIOI~
Figure 6-19. Seepage analysis, Blakely Mountain Dam, Arkansas (from U. S. Army Engineer
121
Waterways Experiment Station )
SEEPAGE PER FOOT OF EMBANKMENT LENGTH
CleMBANKMENT "''R'1 h , ... 7.5 X 10-B CM/SEC X 192FT X 13.5/8.6 X 60 SEC/MIN X 1/30.48 FT/CM
X 7,48 GAL/FT3"' 0.00033 GAL/MIN
OFOUNDATION • K 5 h #•
5 X 10- 4 CM/SEC X 192FT X 418.6 X 60 SEC/MIN X 1/30.48 FT/CM X 7.48
GAL/FT3 "' 0.66 GAL/MIN PER FOOT OF EMBANKMENT
FOR 600 FT EMBANKMENT
660 660
600 600
COEFFICIENTS OF PERMEABILITY
K 1 , K 2, K 3 , AND K 4 SAME AS ABOVE
K 5 = Kv • Kh= 5 x 1o-4 CM/SEC
500 500
400 400
... ...
; "'>-
6-21
100 100
TO THE C07
OF FLOW PATHS WITH REFERENCE
FLOW NET.
THE CORE AND RANDOM SECTION WOULD NOT
AFFECT THE FOUNDATION FLOW NET.
-100
b. PERVIOUS FOUNDATION
EM 1110-2-1901
30 Sep 86
EM 1110-2-1901
30 Sep 86
(4) For calculation of Q in figure 6-19(b) the flow nets were con-
sidered to be separate but to have the same number of equipotential drops.
6-22
EM 1110-2-1901
30 Sep 86
CHAPTER 7
SEEPAGE TOWARD WELLS
7-1. Use of Wells. Wells are used in a variety of ways to control seepage.
They may be placed-on the landward side of water retention structures to
reduce pressure at the lower boundary of impervious strata. Wells are also
used to maintain dry conditions in excavations during construction. In
addition to seepage control, well pumping tests serve as an accurate means of
field determination of permeability (see Chapter 2).
7-2. Analysis of Well Problems. The graphical flow net technique described
in Chapter 4 or the approximate methods described in Appendix B can be used in
the analysis of well problems. However, formulas obtained from analytical
solutions to well problems are the most common methods of analysis.
(7-1)
where
k = permeability (L/T)
As for a plan flow net, must be the same for all elements within
the net. Thus is a constant. When drawn in plan view (figure 7-1b)
the flow net consists of square elements as in the plane case described in
Chapter 4. When drawn in profile (figure 7-1c) the elements' aspect ratios
(b/R) are proportional to the radial distance r and are therefore not
squares. Thus, graphical construction of flow nets for radial flow problems
is generally not practical except for cases where the water bearing has a
constant thickness and only the plan view of the net is required.
7-1
EM 1110-2-1901
30 Sep 86
7-3. Basic Well Equations for Steady State Flow. Steady flow conditions
exist when the well flow rate and piezometric surface do not change with time.
If the regional piezometric surface does not fluctuate, steady state condi-
tions are achieved by pumping from a well at a constant rate for a long time
period. Design of wells for seepage control are often based on computations
assuming steady state conditions.
7-2
EM 1110-2-1901
30 Sep 86
(7-2)
where
(7-3)
(7-4)
Inserting equation 7-4 into equation 7-3, the constant term is found to be:
7-3
EM 1110-2-1901
30 Sep 86
7-4
EM 1110-2-1901
30 Sep 86
By substituting the constant term into equation 7-3 and combining logarithmic
terms, the well equation for confined flow is obtained.
(7-5)
The distance re is often defined as the radius beyond which the well has no
influence or radius of influence.
(7-6)
(7-7)
The constant term can be evaluated from the condition at the radius of influ-
ence r e as was done in equations 7-4 and 7-5. The constant term is given
by:
which when substituted back into equation 7-7 gives the well equation for
gravity flow
7-5
EM 1110-2-1901
30 Sep 86
(7-8)
where
= hi (confined flow)
qi = intensity factor
( confined flow)
Ci = constant
(7-9)
7-6
EM 1110-2-1901
30 Sep 86
7-7
EM 1110-7-1901
30 Sep 86
Substituting the above equation into equation 7-9, the well formula for two
wells becomes
(2) A rock bluff line at the edge of an alluvial fill valley which can
be idealized as an impervious boundary.
The superposition of solutions (equation 7-9) can be used to analyze the flow
near a boundary by introducing an artificial device called an image well. An
image well is identical to the actual well and located symmetrically on the
opposite side of the boundary. The superimposed effect of the real and image
well for an infinite well is identical to the influence of the real well and
boundary. If the real well is a pumping well then a recharging image well is
used to represent boundaries such as rivers and a pumping image well is used
to represent an impervious barrier. For either case, the absolute value of
the flow Q for the image well is equal to that of the real well. For
7-8
EM 1110-2-1901
30 Sep 86
(7-10a)
(7-10b)
(7-11)
head at the river. Substituting the constant term into equation 7-11, the
formula for a single well near a recharge boundary is
To describe the head distribution for confined flow near an impervious boundary
an image discharge well is used (figure 7-4b), By the procedure used above, h
would be obtained as
(7-12)
7-9
EM 1110-2-1901
30 Sep 86
7-10
EM 1110-2-1901
30 Sep 86
(7-13)
The image well method can also be applied to problems involving multiple
boundaries. For example, a common geologic situation involving multiple
boundaries would be a discharge well pumping from an alluvial terrace located
between a river and rock bluff (figure 7-5). In this case, the image well for
the river would have a second image well with respect to the rock bluff, which
in turn would have an image with respect to the river and so on. A similar
progression of image wells would be needed for the impermeable barrier.
Eventually, the location of each added-image well extends beyond its radius of
influence re from the pumping well and has no practical influence in the
solution.
7-4. Special Conditions. Although the simple well formula (equation 7-8) is
often used to analyze flow problems, it describes a relatively idealized con-
dition that is found rarely in practice. It is generally desirable to con-
sider the effects of partial penetration of wells, sloping aquifer, and
stratification of water bearing units in the analysis.
(7-14)
7-11
EM 1110-2-1901
30 Sep 86
b. SECTION
a. PLAN
c. PRIMARY IMAGE
WELLS TO ACCOUNT FOR
INFLUENCE OF BOUNDARIES
ON REAL WELL
7-12
EM 1110-2-1901
30 Sep 86
124
Figure 7-6. Beta function curve (from Warriner and Banks )
7-13
EM 1110-2-1901
30 Sep 86
from the boundary conditions at the well and at the radius of influence as
(7-15)
where
(7-16)
with
where
(7-17)
with
7-14
EM 1110-2-1901
30 Sep 86
where
a = constant (L)
(7-18)
where H is the head at the source and ris is the distance between well i
and the source. The drawdown at each well is computed from combining
equations 7-17 and 7-18
(7-19)
Equation 7-19 gives the drawdown for each well within a group of fully
penetrating wells. The values of Qi required to cause the drawdown
can be determined by solving the system of N equations (7-19) for the N
unknowns Qi . As for the single well, a shape factor can be defined as:
(7-20)
7-15
EM 1110-2-1901
30 Sep 86
where is the shape factor for each well within a group. This shape
factor can be corrected to account for partial penetration by
(7-21)
By replacing in equation 7-16 with the flow from the well group is
given as
(7-22)
The computations required to evaluate equations 7-19 through 7-22 are straight-
forward though they are time consuming for large well groups. Warriner and
Banks (1977) provide a FORTRAN code to compute discharge and drawdowns for
partially penetrating well groups within an arbitrarily shaped source boundary.
(7-23)
where
k = permeability (L/T)
7-16
EM 1110-2-1901
30 Sep 86
7-17
EM 1110-2-1901
30 Sep 86
(7-24)
where
(7-25)
(7-26)
or
(7-27)
(7-28)
7-18
EM 1110-2-1901
30 Sep 86
(7-29)
(7-30)
which when combined with the well equations for unconfined flow gives
(7-31)
(7-32)
where
d m = thickness of layer m
7-19
EM 1110-2-1901
30 Sep 86
Note that the permeability determined from a field pumping test is an average
of all units penetrated by the pumping well. A case where vertical flow can be
important is shown in figure 7-8. The discharging well is pumping from a
Figure 7-8. Flow to well with significant vertical flow through confining
164
layer (courtesy of John Wiley and Sons )
(7-33)
where
B = thickness of aquifer
C = B'/k' (L)
7-20
EM 1110-2-1901
30 Sep 86
7-5. Nonsteady State Flow. Nonsteady state flow may arise in several ways.
When pumping is started, time is required to establish a virtually steady state
condition. Flow during this period must be assumed to be nonsteady state. If
pumping occurs intermittently, a steady state condition may not be established.
Also, if large fluctuations occur at the source, potential steady state flow
conditions are not maintained. The steady state condition can be viewed as the
end condition that is reached after pumping for a long time period. In the
design of a well system for seepage control, it is generally adequate to
consider only the steady state condition. However, the determination of
coefficient of permeability from test data often requires analysis based on
nonsteady state condition. The duration of many well tests is too short to
(a) Prepared from more extensive tables presented by Kruseman and De Ridder
(1970).
7-21
EM 1110-2-1901
30 Sep 86
(7-34)
where
k = permeability (L/T)
(7-35)
where
The storage coefficient S represents the amount of water removed from stor-
age as a result of consolidation of the aquifer and expansion of water in
response to the decline in head. Physically S is given by
(7-36)
7-22
EM 1110-2-1901
30 Sep 86
1 1 W(u)
f u f = 1.0 f = 2.0 f = 8.0
7-29
EM 1110-2-1901
30 Sep 86
where
B = thickness of aquifer
n = porosity (dimensionless)
(1) A plot is made of W(u) (log scale) versus u (log scale). This
plot is referred to as the type curve.
(3) Superimpose the test data over the type curve in such a way that
the drawdown data best fit the type curve (figure 7-9). The coordinate axes
of the two curves should be kept parallel.
7-24
EM 1110-2-1901
30 Sep 86
Figure 7-9. Use of type curve for analysis of nonsteady state flow
164
(courtesy of John Wiley and Sons )
(5) Compute the value of kB from equation 7-34 using the matching
point value of H - h and W(u) . Compute the value of S from equa-
2
tion 7-35 using the matching point values of u and r /t combined with the
previously computed value of kB . The above procedure is carried out for
each observation well. Ideally, the computed values of kB and S should be
the same for all observation wells. Differences in the computed values may be
caused by geologic variations in the aquifer and hydrologic boundaries not
accounted for in the analysis.
7-25
EM 1110-2-1901
30 Sep 86
(7-37)
(7-38)
(7-39)
Note that ro represents the radius of influence for the well at time equals
t . Thus the radius of influence for the steady state condition re is equal
to ro as t tends to infinity. This implies that the radius of influence
7-26
EM 1110-2-1901
30 Sep 86
b. Simultaneous observations
7-27
EM 1110-2-1901
30 Sep 86
(7-40)
where
(7-41)
where
Sy = specific yield
(1) Type curves for several values of r/B should be plotted. The
curve giving the best fit to the initial time-drawdown data is used to
estimate r/B .
7-28
EM 1110-2-1901
30 Sep 86
a FAMILY OF BOULTON TYPE CURVES: W (UA, r/B) VERSUS 1/uA AND W (Uy, r/b) VERSUS
1/Uy FOR DIFFERENT VALUES OF r/B.
7-29
EM 1110-2-1901
30 Sep 86
(2) The time-drawdown data overlay may be moved to obtain the best fit
for the latter time-drawdown data. Both initial time and latter time fits
should give the same value of r/B and kB .
(7-42)
where
r 2S
u=4kBt
r = radius from well (L)
k = permeability (L/T)
7-30
NONEQUILIBRIUM
TYPE CURVE
0.0
.1 0.075 0 .05
0 3 0.2 0.15
·0 5°.4 .
o.s-·
1.0 0.8
1 14.6 OW!u, r/Bl
~-------- 1.5 5"'
T
----------2.0 u=
2100 ,-2s
Tt
7-31
~---------r/8=2.5
0.1~----~~~----------------~----------~~~~r~~-------1------~
r /8= /IK'/b') vT
0.01~~~~~-~~~~-~~~~~~~WUUL~~~~~~~~~ti-~-L~~~-L-L~~
10- 1 1.0 10 lol 103 10 4 10 5 106 107
1/u
Figure 7-12. Type curve for nonsteady flow in a semiconfined aquifer (courtesy of
EM 1110-2-1901
164
John Wiley and Sons )
30 Sep 86
EM 1110-2-1901
30 Sep 86
The application of the type curve method for the leaky aquifer problem is simi-
lar to the application to the delayed yield problem. The time-drawdown data
are matched to the standard type curve with the curve giving the best fit being
used to estimate r/B .
(7-43)
(7-44)
where
7-32
EM 1110-2-1901
30 Sep 86
Note, then, when the function W(u) can be replaced with a logarithmic
approximation, as in the Jacob's method (equation 7-37), equation 7-44 can be
approximated as
(7-45)
From equation 7-45 it is seen that as u becomes small, flow becomes virtually
steady state (compare equation 7-45 with the steady state case, equation 7-11).
Thus the presence of a recharge boundary in an aquifer tends to shorten the
time needed to reach steady state (Davis and Dewiest 1966).
7-33
EM 1110-2-1901
30 Sep 86
CHAPTER 8
SEEPAGE CONTROL IN EMBANKMENTS
8-1. General. All earth and rock-fill dams are subject to seepage through
the embankment, foundation, and abutments. Seepage control is necessary to
prevent excessive uplift pressures, instability of the downstream slope, pip-
ing through the embankment and/or foundation, and erosion of material by
migration into open joints in the foundation and abutments. The purpose of
the project, i.e., long-term storage, flood control, etc., may impose limita-
tions on the allowable quantity of seepage.
8-2. Methods for Seepage Control. The three methods for seepage control in
embankments are flat slopes without drains, embankment zonation, and vertical
(or inclined) and horizontal drains.
8-3. Flat Slopes Without Drains. For some dams constructed with impervious
soils having flat embankment slopes and infrequent, short duration, high
reservoir levels, the phreatic surface may be contained well within the down-
stream slope and escape gradients may be sufficiently low to prevent piping
failure. For these dams, when it can be assured that variability in the
characteristics of borrow materials will not result in adverse stratification
in the embankment; no vertical or horizontal drains are required to control
seepage through the embankment. A horizontal drain may still be required for
control of underseepage (see Chapter 9). Examples of dams constructed with
flat slopes without vertical or horizontal drains are Aquilla Dam, Aubrey Dam,
and Lakeview Dam (U. S. Army Engineer District, Fort Worth 1976a, 1976b, 1980).
Figure 8-1 shows the analysis of through embankment seepage for Aubrey Dam,
Texas (now called Ray Roberts Dam). As shown in figure 8-1a, this is a zoned
embankment with relatively flat slopes due to a weak stratum in the founda-
tion. The slopes could be steepened from IV:10.6H to lV:8H if the weak foun-
dation gains shear strength due to consolidation during construction (the dam
is scheduled for completion in 1988). A 3-ft-thick horizontal drainage blanket
and collector system will be provided under the downstream embankment from
sta 136+00 to sta 142+60 to control any seepage through the foundation. For
the analysis of through embankment seepage, shown in figure 8-1b, the steady
state phreatic surface was developed graphically for the conservation pool by
considering a homogeneous nonisotropic embankment and an impervious foundation.
Since the escape seepage gradients were computed to be less than 0.3 to 0.4
(see paragraph 4.9b), it was concluded that no vertical or horizontal drains
were required.
8-1
30 Sep 86
EM 1110-2-1901
ct_ EMBANKMENT
SELECT
IMPERVIOUS I
44'
FILL
UPSTREAM 18" RIPRAP ON 6 DOWNSTREAM
.--
EL 511.0
FLOODPLAIN EMBANKM'NTSECTION
STATION 100+00 TO STATION 142+25
...
SCALES
HORIZONTAL
30 0 30 60FT
MW ~ERT!CAL
a. FLOODPLAIN EMBANKMENT SECTION
8-2
£EMBANKMENT
UPSTREAM DOWNSTREAM
FLOODPLAIN EMBANKMENT
LEGEND
...
b. EMBANKMENT SEEPAGE ANALYSIS
50 100FT
KH =HORIZONTAL PERMEABILITY
Kv ~VERTICAL PERMEABILITY
Figure 8-1. Analysis of through embankment seepage for Aubrey Dam, Texas
81
(from U. S. Army Engineer District, Fort. Worth )
EM 1110-2-1901
30 Sep 86
8-3
EM 1110-2-1901
30 Sep 86
but below the bottom of the frost zone, A vertical core located near the
center of the dam is performed over an inclined upstream core because the
former provides higher contact pressure between the core and foundation to
prevent leakage, greater stability under earthquake loading (Sherad 1966,
1967), and better access for remedial seepage control. An inclined upstream
core allows the downstream portion of the embankment to be placed first and the
core later and reduces the possibility of hydraulic fracturing (Nobari, Lee,
and Duncan 1973).
(1) Laboratory filter tests are not a routine laboratory test. Standard test-
ing procedures have not been developed. The conduct of laboratory filter
tests should be under the direction of a specialist and should be carried
out in a research laboratory.
8-4
EM 1110-2-1901
30 Sep 86
8-5
EM 1110-2-1901
30 Sep 86
8-6
EM 1110-2-1901
30 Sep 86
8-7
EM 1110-2-1901
30 Sep 86
(8-1)
where
8-8
EM 1110-2-1901
30 Sep 86
8-9
EM 1110-2-1901
30 Sep 86
KH = KV
KH = 9KV
KH = 9KV
8-10
EM 1110-2-1901
30 Sep 86
required. Well-graded materials are internally unstable and should not be used
as filters when Cu > 20.(l)
(8-2)
(1)
8-11
EM 1110-2-1901
30 Sep 86
where
k = coefficient of permeability
Q = quantity of discharge
i = hydraulic gradient
Figure 8-9. Example of design procedure for inclined and horizontal drains
155
to assure adequate drain capacity (courtesy of John Wiley and Sons )
8-12
EM 1110-2-1901
30 Sep 86
(8-3)
Every filter must be permeable enough to have a reasonable reserve for higher
than expected flows. The filter should have a minimum permeability after
placement and compaction of at least 20 times that calculated theoretically.
Therefore, the required permeability for the inclined drain is
(8-4)
(8-5)
8-13
EM 1110-2-1901
30 Sep 86
(8-6)
To design the horizontal drain, select a drain height and calculate the
required minimum permeability. Apply a factor of safety of 20 to the calcu-
lated permeability and select a drain material from available aggregates.
Select a drain height of 4 ft.
8-14
EM 1110-2-1901
30 Sep 86
(8-7)
From figure 8-10, for screened fine gravel (3/8-in. to 1/2-in. size) with
1 = 0.007 the reduction factor for permeability is 0.9.
(8-8)
The permeability of the screened fine gravel (3/8-in. to 1/2-in. size) reduced
for turbulence is greater than the required permeability for design:
k = 0.1 ft/day
8-15
EM 1110-2-1901
30 Sep 86
foundation to the toe of the dam. The thickness of the horizontal drain must
be sufficient to satisfy the discharge requirements. When filter or drain
material is not available locally and must be hauled to the site at substantial
cost, a thin (2 ft or greater) horizontal drain has been used (U. S. Army Engi-
neer District, Tulsa 1975 and U. S. Army Engineer District, Louisville 1974).
Stringers or finger drains may be used as an alternative to a continuous hori-
zontal drain, as shown in figure 8-11, when the drain material is costly. The
cross-sectional area of the stringer drains must be sufficient to satisfy the
discharge requirements. The stringer drain may be constructed either by
trenching into the embankment or foundation for narrow (6 ft) widths or by
placing the adjacent impervious fill and then the drain material for wider
(50 ft) widths (U. S. Army Engineer District, Kansas City 1974, 1978). In
either case, the side slopes of the stringer drains should be sloped instead of
8-17
EM 1110-2-1901
30 Sep 86
vertical (see figure 8-11) to avoid stress concentrations which could cause
vertical transverse cracks. The stringer drain material must be thoroughly
compacted (see paragraph 8-4c) to ensure that consolidation does not occur upon
saturation leaving an open seepage conduit in the top of the trench, bridged by
the overlying embankment, and susceptible to progressive erosion (Jansen 1980).
The downstream end of horizontal drains and stringer drains must be able to
discharge freely and must be protected against siltation and erosion. This may
be accomplished by providing a weighted filter (riprap overlying bedding mate-
rial) or a toe drain as shown in figure 8-12 (U. S. Army Engineer District,
Mobile 1965 and U. S. Army Engineer District, Tulsa 1975). The toe drain has
the advantage of lower maintenance requirements and preventing the development
of localized wet areas at the surface along the downstream toe of the
embankment.
8-6. Seepage Control Against Earthquake Effects. For dams located where
earthquake effects are likely, there are several considerations which can lead
to increased seepage control and safety.- The core material should have a high
resistance to erosion (Arulanandan and Perry 1983). Relatively wide transition
and filter zones adjacent to the core and extending for the full height of the
dam can be used. Additional screening and compaction of outer zones or shells
will increase permeability and shear strength, respectively. Geometric con-
siderations include using a vertical instead of inclined core, wider dam crest,
increased freeboard, and flatter embankment slopes, and flaring the embankment
at the abutments (Sherard 1966, 1967).
8-18
EM 1110-2-1901
30 Sep 86
8-19
EM 1110-2-1901
30 Sep 86
CHAPTER 9
SEEPAGE CONTROL IN EARTH FOUNDATIONS
9-2. Selection of Method for Seepage Control. The methods for control of
underseepage in dam foundations are horizontal drains, cutoffs (compacted back-
fill trenches, slurry walls, concrete walls, and steel sheetpiling (1)), up-
stream impervious blankets, downstream seepage berms, toe drains, and relief
wells. To select an underseepage control method for a particular dam and
foundation, the relative merits and efficiency of different methods should be
evaluated by means of flow nets or approximate methods as described in
Chapter 4 and Appendix B, respectively. As shown in table 9-1, the changes in
the quantity of underseepage, factor of safety against uplift, and uplift pres-
sures at various locations should be determined for each particular dam and
foundation. Since the anisotropy ratio of the foundation has a significant
influence on the results of the underseepage analysis, this parameter should be
9-4. Cutoffs.
9-1
Table 9-1. Selection of Underseepage Control Method
EM 1110-2-1901
30 Sep 86
Changes in i (a)
c
Quantity of F
i
Uplift Pressure
s
Method Under seepage e Under Dam Toe of Dam Downstream of Dam
Horizontal Drain
Complete Cutoff r---------------------,
1 Fill
in table for each particular 1
Partial Cutoff
I dam and foundation. Vary the I
Upstream Impervious Blanket
I permeability anisotropy of the :
Downstream Seepage Berm
Toe Drain ILI foundation
______
(K
1~_=_1,_10, 25, 100) as appropriate.~
_ _ _ _ _ _ _ _ _ _ _ _ ...J
Relief Wells
9-2
(a) Factor of safety against uplift at downstream toe of dam given by the ratio of the
critical hydraulic gradient to the vertical component of the exit gradient.
(prepared by WES)
EM 1110-2-1901
30 Sep 86
9-3
EM 1110-2-1901
30 Sep 86
(3) If the tailwater conditions are such that ponds of water exist
downstream of the dam so that underseepage would emerge underwater and could
not be observed, it is desirable to be more conservative in evaluating the
need for a complete seepage cutoff.
(4) The amount of silt and clay sized particles in suspension in the
river water which contributes to siltation of the reservoir with time and
tends to diminish underseepage.
Theory and model tests indicate that it is necessary for a cutoff to penetrate
a homogeneous isotropic foundation at least 95 percent of the full depth before
there is any appreciable reduction in seepage beneath an earth embankment as
shown in figure 9-2 (Telling, Menzies, and Coulthord 1978; and Mansur and
Perret 1949). The effectiveness of the partial cutoff in reducing the quan-
tity of underseepage decreases as the ratio of the width of the dam to the
depth of penetration of the cutoff increases (see figure 9-2). Partial cut-
offs are effective only when they extend down into an intermediate stratum of
lower permeability. This stratum must be continuous across the valley founda-
tion to ensure that three-dimensional seepage around a discontinuous stratum
does not negate the effectiveness of the partial cutoff.
(9-1)
where
(9-2)
9-4
EM 1110-2-1901
30 Sep 86
where
The head efficiency is more widely used because the field performance may be
established from piezometric data taken during construction, before and during
initial filling of the reservoir, and subsequently as frequently as necessary
to determine changes that are occurring and to assess their implications with
respect to safety of the dam, as described in Chapter 13. The flow efficiency
may only be approximated since the rate of underseepage without the cutoff
cannot be directly established and since it is difficult to measure the rate
of underseepage with the cutoff because, except for special cases, only part
of the underseepage discharges at the ground surface immediately downstream of
the dam (Telling, Menzies, and Simons 1978a and Marsal and Resendiz 1971). The
flow efficiency of a compacted backfill partial cutoff in a foundation of
permeable soils of moderate thickness overlying an impervious rock is shown in
figure 9-3. This figure also illustrates the high seepage gradients that occur
along the base of the cutoff and on its downstream face in both the foundation
and embankment zones. Suitable filters must be provided to prevent piping of
soil at faces A-B-C in figure 9-3a and 9-3b (Cedergren 1977 and Klohn 1979).
As shown in figure 9-4, a partial cutoff in a homogeneous isotropic foundation
will lower the line of seepage in the downstream embankment somewhat but exit
gradients at the downstream toe (as reflected by the distance between the
equipotential lines) are reduced only slightly (Cedergren 1973). When the
pervious foundation is cut off by a compacted backfill or slurry trench, (1) the
rate of underseepage may be estimated by (Ambraseys 1963 and Marsal and
Resendiz 1971)
(9-3)
where
3
Q o = rate of underseepage in m /set per running meter of dam
(1) This approach neglects the contribution of the filter cake that forms on
the trench walls to the overall slurry trench permeability. When the
permeability of the backfill placed in the trench is high, the overall
slurry trench permeability will be controlled by the filter cake
(D'Appolonia 1980).
9-5
EM 1110-2-1901
30 Sep 86
For a concrete wall or steel sheetpiling with defects (openings in the cutoff)
the rate of underseepage per unit length of cutoff is given by (Ambraseys 1963
and Marsal and Resendiz 1971)
(9-4)
where
9-6
0~----------~-----------T------------~----------~----------~
L
20
0
J:o D
1-x
n..o
w, 40 ~-----------+------------~------------+-------------+-------~~~
0"'0
lL..
0 0 ..
1-w
9-7
zl-
w~ 60~----------+-----------4------------r-----------+----~~~~
ucr
a:f-
ww
CLZ
w
Q..
20 40 60 80 100
EM 1110-2-1901
PERCENT OF TOTAL SEEPAGE
30 Sep 86
Figure 9-2. Effect of depth of penetration of partial cutoff on seepage reduction for a
homogeneous isotropic foundation (prepared by WES)
EM 1110-2-1901
30 Sep 86
a. Partial cutoff
b. Complete cutoff
9-8
EM 1110-2-1901
30 Sep 86
9-9
EM 1110-2-1901
30 Sep 86
Material and compaction requirements are the same as those for the impervious
section of the dam (EM 1110-2-1911). When constructing a complete cutoff (see
para 9-4a), the trench must fully penetrate the pervious foundation and be
carried a short distance into unweathered and relatively impermeable foundation
soil or rock. To ensure an adequate seepage cutoff, the base width should be
at least one-fourth the maximum difference between the reservoir and tailwater
elevations but not less than 20 ft, and should be wider if the foundation
material under the cutoff is considered marginal in respect to imperviousness.
As previously mentioned (see para 9-4b), high seepage gradients occur along the
base of the cutoff and on its downstream face in both the foundation and
embankment zones. Suitable filters must be provided (see Appendix D for design
of filters) to prevent piping of soil at these interfaces. The trench
9-10
EM 1110-2-1901
30 Sep 86
excavation must be kept dry to permit proper placement and compaction of the
impervious backfill. Dewatering systems of wellpoints or deep wells are
generally required during excavation and backfill operations when below
groundwater levels (TM 5-818-5). Because construction of an open cutoff trench
with dewatering is a costly procedure , the trend has been toward use of the
slurry trench method of construction (EM 1110-2-2300 and Cedergren 1977).
d. Slurry Trench.
(1) Introduction. When the cost of dewatering and/or the depth of the
pervious foundation render the compacted backfill trench too costly and/or
impractical, the slurry trench cutoff may be a viable method for control of
9-11
EM 1110-2-1901
30 Sep 86
(a) History of Use. The first use of the slurry trench method of con-
struction was by the U. S. Army Engineer District, Memphis, in September 1945,
to form a partial cutoff along the Mississippi River levee on the Arkansas
side of the river just below Memphis, Tennessee (Clay 1976 and Kramer 1946).
The idea for the project probably evolved from the use at that time of puddle
clay trench cutoffs combined with the use of drilling mud for advancing
borings. A paddle wheel mixing device was constructed for making slurry from
native clays. Trenches were dug to a 20-ft depth using a trenching machine and
to a 35-ft depth using a dragline with a 100-ft boom and 2-cu-yd bucket.
Backfill was mixed in windrows at the site from hauled-in clay gravel and
native materials and pushed into the trench by a bulldozer when the length of
the trench was equal to about twice the trench depth. BG Hans Kramer foresaw
the use of the slurry trench method for the construction of cutoffs for earth
dams. It is amazing that after 38 years, the technique is still about the same
as it was when first developed by the Memphis District. A soil-bentonite
cutoff was constructed under the Kennewick Levee adjacent to the Columbia River
as part of the McNary Dam Project in Washington by the Walla Walla District in
1952 (Jones 1961). The first application of a soil-bentonite slurry trench
cutoff for control of underseepage at a major earth dam was at Wanapum Dam on
the Columbia River in Washington in 1959 (La Russo 1963). Subsequently, soil-
bentonite cutoffs have been used for control of underseepage at a number of
dams as shown in table 9-2. The cement-bentonite slurry trench cutoff was
first used to tie into the abutment zones at the Razaza Dam on the Euphrates
River in Iraq in 1969 (Soletanche 1969). Subsequently, cement-bentonite cut-
offs were installed as remedial seepage control measures through the embank-
ment and foundation of four existing dams in Mexico from 1970 to 1972
(Soletanche 1970, 1971, 1971-1972, 1972). As shown in table 2, the first
cement-bentonite cutoff in the United States was constructed at the Tilden
Tailings Project to store tailings from the Tilden Mine in Michigan in 1976
(Meier and Rettberg 1978). The first cement-bentonite cutoff constructed at a
dam on a river retaining a reservoir in the United States was completed in
1978 at the Elgo Dam (formerly the San Carlos Dam) in Arizona (Anonymous 1978
and Miller and Salzman 1980).
9-12
EM 1110-2-1901
30 Sep 86
Table ?-2. Comparison of Slurry Trench Cutoffs
Max. Hal<.
Differential
!late Head Width De~th
Pro1ect Location Owner Constructed Foundation Uaterial ft __!!._ __!!._ Location ~ Reference
Soil-Bentonite Slurrv Trench
Kennewick Levee, Columbia River .. Corps of Engineers 1952 Sandy or silty gravel with 15 22 Center of 2. 5 Jones 1961
t-tcNa.ry Dam Wash. z:ones of open gravel dam
Wou>aj>ua n- Col.-hi.a Uve.r. Pllfllie UUlUy 1962 S<md, g1"3"ftls ...,d. g,;ave!.ly 88 10 80 Center o£ 8.8 L.a Russo 1963
wasn. District 2 .. Grant sands underlain by open da10
County, Wasb. wrk gravels
Wells Dam Columbia River, l'ublic Utility 1964 Pervious gravels 70 80 Center of 8.8 Jones 1967
Wash. District 1. dam
Douglass County,
Wasb.
Yards Creek Hew Jersey Public Service 1964 Sands~ gravels, cobbles, 55 8 40 Center of 6.9 Jones 1967
Lo""r Reservoir Electricity and and boulders daa
Gas
Comanche Dam - Moke1UII!IIIe East Bay 1966 Opper stratum of clayey 45 95 Upstream 5.6 Anton and
Dike 2 River, CaUf. )!unicipal District silts, silts, and clayey of toe of Dayton 1972
sands .. Lower stratum of da..
sand over zone of gravel
\lest Point Dat'l Chattabocchie Corp~ of Eng1neers 1966 Upper stratum of alluvial 61 60 Upstream 12.2 u.s. Army
River, Ca. soil~ alternating layers toe of dam Engr~ Dist,
Plum Creek Plum Creek., Soil Conservation 1973 Sands, gravelly sands. 54 40 Ce.ntet:' of 10.8 Knabach and
llaa Wis. Service silty gravels, silty sands dam Dingle 1914
Upper Big Upper Big Soil Cortservet ion 1978 Sands and gravels 62 40 Center of 15.5 Bloom, Dynes,
&lue Daa Blue R.iver, Service dam Glol!sett
Ind. 1979
Addicks lla,.(a) Buffalo Bayou, Co-rps of Engineers 1979 Silty aands 30 70 Varies 10.0 u.s. Anny
Tex~ Engr .. Dist,
Galveston
Barker Dall(a) Buffalo Bayou, Corps of Engineers 1979 Sandy clay, clayey sand, 24 55 Cenr:er of 8.0
ailty sand 1983 (sallie
Tex .. for Barker Da11)
(a) Slurry trench c.ucof[ ins called aa reaedial seepage control for exlatina da11 (aee Chapter 12).
(b) Planrted maximum differential head wben ta1.11ng;a dam is at final height (45 ft in February 1983}.
9-13
EM 1110-2-1901
30 Sep 86
Item Soil Bentonite Slurry Trench Cutoff Cement-Bentonite Slurry Trench Cutoff
1. Excavation Long slope of backfill requires trenching con- Construction sequence is more flexible to meet site
tinuously in one direction. constraints.
Can accommodate interruptions in construction. Excavation for each panel should progress uninter-
rupted so each panel is completed before the slurry
begins to set.
Difficult to work with coarse gravels, with Easier to remove coarse gravels with clam shell
backhoe and/or dragline, which settle out and mounted on a rigid Kelly bar, from trench trench
accumulate on the trench bottom reducing the bottom. Better suited for excavation in areas
certainty of a tight seal. prone to failure.
Wider trench required to prevent segregation dur- Narrower trench (2 - 3 ft) may be used.
ing backfilling (> 3 ft) and prevent piping
failure of the sl~rry trench backfill into the
adjacent soil downstream of the cutoff (1 ft
width for each 10 ft of differential head)
2. Slurry Mixture of bentonite and water used to support Mixture of cement, bentonite, and water used to
excavation during trenching. Slurry displaced by support ~xcavation foe panel and later sets to form
backfill is desanded and reused. Upon completion permanent backfill. No disposal or desanding of
of trench, unused slurry must be disposed of in slurry required.
an environmentally acceptable fashion.
9-15
' 3. Backfill Mixture of slurry plus excavated material from None (see 2). Therefore•not dependent on availa-
trench and/or imported select backfill materials bilityor quality of soil for backfill and more
.suitable for work in confined areas.
6. Relative cost Generally lower if cost of backfill is not Generally higher due to cost of cement. More
EM 1110-2-1901
prohibitive (see 3) competitive as depth of trench increases due to
narrower trench width required.
7. Patent applicable No (expired 1973) Yes (United States Patent No. 3,759,044 dated
30 Sep 86
September 18, 1973)
backfill material, the factor of safety against instability is (Nash and Jones
1963)
(9-5)
where
F = factor of safety
Cu = undrained cohesion
For a slurry trench excavated in dry cohesionless soil (Nash and Jones 1963)
(9-6)
(9-7)
where
For arbitrary levels of ground water and slurry in cohesionless soil, as shown
in figure 9-7a, a slightly conservative (neglects arching effect of short
9-16
EM 1110-2-1901
30 Sep 86
trenches and stabilization of the soil adjacent to the trench face due to
slurry penetration and gelation) estimate of the slurry density required to
ensure stability of the trench is (Morgenstern and Amir-Tahmasseb 1965)
(9-8)
where
Equation 9-6 may be solved by use of the nomograph shown in figure 9-7b
(Duguid et al. 1971).
(2) Slurry. The slurry has three basic functions in slurry trench
construction (Ryan 1977):
9-17
EM 1110-2-1901
30 Sep 86
9-18
EM 1110-2-1901
30 Sep 86
bentonite should be used for slurry trench construction. The pH of the water
used for mixing with the bentonite should equal 7.0 ± 1.0. Water hardness
should not exceed 50 ppm (parts per million). Total dissolved solids should
not exceed 500 ppm. The amount of oil, organics, or other deleterious sub-
stances should be limited to no more than 50 ppm each (Stanley Consultants,
Inc., and Woodward-Clyde Consultants 1977). If the use of poor quality water
cannot be avoided, it will require more bentonite and longer mixing times to
achieve the desired properties.
(b) Mixing. The methods used to prepare the hentonite slurry vary with
project size and layout. Always add the bentonite to the water, never the
water to the bentonite. For small jobs the batch system is used where
specific quantities of water and bentonite are placed in a tank and mixed at
high speeds with a circulationpump or paddle mixer and the slurry is dis-
charged into the trench. Mixing is usually complete in a matter of minutes
for the 2- to 5-cu-yd batch produced by this method. The most commonly used
method is the flash or Venturi mixer and circulation ponds which is well
adapted for bulk handling of large slurry volumes. A flash mixer introduces
dry bentonite into a turbulent water jet which discharges into a low speed
circulation pond. When Marsh Funnel viscosity readings stabilize, the slurry
is stored in a second pond prior to using the trench (Spooner et al. 1982 and
D'Appolonia 1980).
9-19
30 Sep 86
EM 1110-2-1901
Table 9-4. Summary of Quality Control Tests for Slurry Trench Cutoff
Bentonite Viscosity Each railroad car or truck load API RP 13A > 40 seconds at 65°F
Bentonite slurry (after Viscosity Twice per day API RP 13B > 40 seconds at 65°F
mixing) pH Twice per day API RP 13A 7 - 10
Bentonite content Twice per day None(b) 3 - 7 percent
3
Unit weight Twice per day API RP 13B 1.01 - 1.04 g/cm
3
Filtrate or water loss Twice per day API RP 13B > 20 cm at 100 psi in 30 min
Bentonite slurry (in trench) Viscosity Twice per day API RP 13B > 40 seconds at 65°F
Unit weight Twice per day API RP 13B minimum(c)_ 85 pcf
Cement-bentonite slurry Viscosity Twice per day API RP 13B 40-50 seconds at 65°F
(after mixing) pH Twice per day API RP 13A 10 - 13
Bentonite content Twice per day None 4 - 7 percent
1.03 - 1.04 g/c~
3
Unit weight Twice per day API RP 13B
3
< 165 cm at 100 psi in 30 min
Filtrate or water loss Twice per day API RP 13B
Cement/water ratio Twice per day None 0.17 - 0.20
mixing and in the trench to determine that it is dense enough to stabilize the
trench, but not so dense as to cause the backfill to settle too loosely, and
that it has sufficient viscosity to maintain cuttings in suspension (Stanley
Consultants, Inc., and Woodward-Clyde Consultants 1976).
(b) Bottom Treatment. The aquiclude used for the slurry wall founda-
tions should be continuous, and relatively free of fractures and other pervious
zones. The cutoff wall should extend a minimum of 2 ft into clay (or 1 ft into
rock) to prevent weathered zones, desiccation cracks, or other geological
features from permitting seepage under the cutoff (Spooner 1982). As the
trench is excavated, heavier soil particles such as sand and gravel fall to the
bottom of the trench. The amount of sand accumulation on the trench bottom
depends upon the coarseness of the strata being excavated as well as the
9-21
EM 1110-2-1901
30 Sep- 86
excavation technique used. Although this sand layer may not have a direct
effect on trench stability, it may adversely affect the permeability of the
slurry trench cutoff wall (Spooner 1982). An air lift pump should be used to
remove the sand and gravel particles from the trench bottom prior to backfill-
ing. When the slurry trench is keyed into a soil aquiclude after the trench
bottom has been cleaned thoroughly, a minimum of one split-spoon sample shall
be taken every 50 ft along the length of the trench to determine if additional
excavation is required (Winter 1978).
9-22
EM 1110-2-1901
30 Sep 86
accordance with ASTM C-143 (1) (Winter 1978 and Ryan 1976). The backfill is
placed continuously from the beginning of the trench in the direction of the
excavation to the end of the trench. Free dropping of the backfill material
through the slurry would produce segregation and is not allowed. Depending on
the steepness of the excavated slope, it may be necessary to lower the initial
backfill to the bottom of the trench with a crane and clamshell bucket until a
slope at the angle of repose of the backfill has been formed from the bottom of
the trench to the top of the excavation. The toe of the backfill slope is kept
to within 50 to 150 ft of the leading edge of the excavation to minimize the
open length of the slurry-supported trench while allowing enough space behind
the excavation for cleaning the trench bottom. Additional backfill is placed
by a bulldozer in such a manner that the backfill enters the trench and slides
progressively down the slope of the previously placed backfill and produces a
slope ranging from 1V on 5H to 1V on 10H. The slope of the backfill shall be
measured with soundings starting at the toe of the backfill in the bottom of
the trench and progressing up the backfill slope at 25-ft horizontal intervals.
A set of soundings shall be made at least for every 25 ft horizontal advance-
ment of backfill placement. Once the natural slope of the backfill is estab-
lished during initial placement of the backfill, the slope should remain nearly
the same. If the slope, or a portion of the slope, suddenly gets steeper, it
could be an indication that sediment is being trapped or that the backfill has
a pocket of relatively clean material (slurry not mixed in properly or was
washed out). If the slope suddenly gets flatter, it could indicate that a
pocket of slurry was trapped in the backfill or that the backfill does not
contain sufficient sand or coarser material (Stanley Consultants, Inc., and
Woodward-Clyde Consultants 1976; Winter 1978; and Ryan 1976).
(1) American Society for Testing and Materials standard. If a desirable back-
filled slope (1V on 5H to 1V on 10H) cannot be maintained in the trench
with a 5 in. ± 1 in. slump, the slump may be altered to meet construction
conditions. Such was the case at the soil-bentonite slurry trench cutoff
constructed at W. G. Huxtable Pumping Plant, Marianna, Arkansas (U. S.
Army Engineer District, Memphis 1978).
9-23
EM 1110-2-1901
30 Sep 86
(b) Blowout Requirements. Once the slurry trench is installed, the dam
has been constructed, and the reservoir filled, there is a substantial differ-
ential head acting on the slurry trench (see table 9-2 for typical values).
Depending upon the characteristics of the backfill material and pervious foun-
dation, the hydraulic gradient acting across the slurry trench may be suffi-
cient to cause blowout or piping of backfill material into the surrounding per-
vious foundation. This is especially critical when the foundation contains
openwork gravel where the piping process could result in the formation of
channels and cavities that may breach the slurry wall. Based upon laboratory
tests conducted on widely graded gravel containing no sand, the blowout gradi-
ent ranges from 25 to 35, depending on the properties of the backfill mate-
rial (La Russo 1963 and Nash 1976). The factor of safety against blowout is
(9-9)
where
(9-10)
where
9-24
EM 1110-2-1901
30 Sep 86
(9-11)
(9-12)
where
tb = backfill thickness
k b = backfill permeability
9-25
EM 1110-2-1901
30 Sep 86
9-26
EM 1110-2-1901
30 Sep 86
9-27
EM 1110-2-1901
30 Sep 86
(f) Mix Design. The gradation of the backfill for the soil-bentonite
slurry trench is selected by conducting permeability, shear strength, and com-
pressibility tests on a range of materials including soil to be excavated from
the trench. Such a procedure was followed in the mix design for the backfill
of the soil-bentonite slurry trench installed for remedial seepage control at
Addicks Dam, Texas (U. S. Army Engineer District, Galveston 1977c). The allow-
able range set on the gradation of the backfill should produce a material which
contains enough fines to reduce the seepage through the slurry trench cutoff to
an acceptable level and sufficient coarse particles to approximate the strength
and compressibility of the surrounding ground. If sufficient fines are not
present in material excavated from the trench, borrow sources should be identi-
fied or alternatively a higher bentonite content specified for the backfill.
If sufficient coarse particles are not present in material excavated from the
trench, approved sources should be identified for obtaining natural sound,
hard, durable sand and gravel. Crushed limestone, dolomite, or other crushed
calcareous materials should not be used. The maximum particle size of the
gravel shall be 3 in. and the material should be well graded.
9-28
Table 9-5. Summary of Shear Strength Data from Post-Construction Testing of Backfill Material from Soil-Bentonite
Slurry Trench Cutoff at Saylorville Dam, Towa(a)
STU-7 87+80
EM 1110-2-1901
30 Sep 86
( ) 106
a From U. s. Army Engineer District, Rock Island
(b) Conducted on undisturbed soil specimens 5 in. in diameter by 11 in. high; unconsolidated-undrained (Q) and consolidated-undrained with
·pore pressure measurements (R) shear tests.
EM 1110-2-1901
30 Sep 86
e PLASTIC FINES
0 NON-PLASTIC FINES
•
/
~
/
e / 1-D COMPRESSION
/e
0 / STRESS INCREMENT
0.5 TO 2.0 kg /cm2
0 ~~--~--~--~--~~~~--~--~--~--~~
0 .02 .04 .06 .08 .10 .12
COMPRESSION RATIO, Cc/ I+ e 0
80
70
/.
.~
60 0
50
0
/
40
0./ • ISOTROPIC COMPRESSION
30
20
:;t .
10
0
/ STRESS INCREMENT
0.5 T02.0 kg/cm2
0
0 .02 .04 .06 .08 .10 .12
9-30
EM 1110-2-1901
30 Sep 86
(d) Set Time. The set time is important because of the construction
technique employed. After the cement-bentonite slurry in the first set of
panels has set, the areas between them can be excavated. A normal cement-
bentonite mixture begins to set after a few hours and has a consistency simi-
lar to lard after 12 hours. The second day the cement-bentonite slurry can be
walked on and final set is normally taken at 90 days (Ryan 1977).
(e) Resistance to Hydraulic Pressure. Once the slurry trench has been
completed, the embankment constructed, and the reservoir filled, there is a
substantial differential head acting on the slurry trench (see table 9-2 for
typical values). The time between completion of the slurry trench and reser-
voir filling is generally sufficiently long (> 90 days) to allow the cement-
bentonite slurry trench to develop its final set. The resistance of the
cement-bentonite material to withstand gradients comparable to those which
will exist in the field should be tested in the laboratory by subjecting
intact specimens which have developed full set to hydraulic pressure and
measuring the increase (if any) of permeability with time (Spooner et al. 1982
and Jefferis 1981).
9-31
EM 1110-2-1901
30 Sep 86
the surrounding ground will be higher than normal and in some instances as
great as 100 percent of the trench volume (Xanthakos 1979). For design
purposes, specimens should be prepared from the cement-bentonite with varying
percentages of sand and gravel, cured for 28 days under consolidation pressures
existing in the field, and laboratory permeability tests conducted
(EM 1110-2-1906).
9-32
Table 9-6. Summary of Shear Strength Data from Post-Construction Testing
(a)
of Cement-Bentonite Slurry Trench Cutoff at Tilden Tailings Project, Michigan
EM 1110-2-1901
30 Sep 86
EM 1110-2-1901
30 Sep 86
100
I I rl
80 f-
Bentonite Type I
e SPV 200
r.
• UG 180 I
60 - I
.,
I
I
-
40
• I
~
/
.... 20
0"
1:
41
...
l<
.,
\1'1
0
-..."'"'
>
41
0 O.l'J o. 20 0. 30
Cement-Water Ratio C/W
0.40 0.50 0.60
0.
E 140
u
0
...c: I I I I
.....
...."' ...
...... \ LEGEND
....., .0 t\.
X
..... 120 • Slurry prepared in the laboratory -
.
~:\
c: o Slurry sampled during constrUction
:::::1
r;,?
......,Cl
........E
.....
:::::1
100
•,\ ' '''
~
~
.,>.
Q
I
80
'~
.
....
a!
' ........
', ~ • .........
60
i' .•""'
'' ~
.......
J
--
00
0
.._
-
40
' ~ :~ 1
• t-0 •
20
' '-,.,
~"-,
.............
•nf~>.ZS"%
.....
0
0 5 10 15 20 25 30
9-34
EM 1110-2-1901
30 Sep 86
(c) Gaps (or Windows) in the Slurry Wall. Trench collapse or improper
placement of backfill can create gaps (or windows) which result in zones of
higher permeability as well as variations in wall thickness and strength
(Spooner et al. 1982). The continuity of the trench should be tested before
backfilling by passing the bucket or clamshell of the excavating tool verti-
cally and horizontally along each segment of the trench. As mentioned pre-
viously, for soil-bentonite slurry trenches irregularities in the backfill
slope are indications that pockets of clean material (slurry not mixed in
properly or was washed out) or slurry were trapped in the backfill or that the
backfill does not contain sufficient sand or coarse material.
e. Concrete Wall.
9-37
EM 1110-2-1901
30 Sep 86
(1)
The workability of the concrete, discussed under Mix Design, is of primary
importance with respect to tremie placement of the concrete.
9-38
EM 1110-2-1901
30 Sep 86
b. Interlocking of primary
and secondary elements
9-39
Table 9-7. Comparison of Corps of Engineers Concrete Cutoff Walls
30 Sep 86
EM 1110-2-1901
t.;utorr
Max. Max. (a)
Dif- Devi-
fer- at ion Max.
Date Embankment- ential Max. from Head
Con- Foundation Head Width Depth Ver- Cutoff
Project Location structed Material ft. ft ....!!__ Location tical Width Reference
Kinzua (formerly Allegheny River 1964 Silts, 140 2.5 185 Upstream 1:370 56.0 U. S. Army
Allegheny) Dam Pennsylvania sands and of toe Engineer District,
gravels, of dam Pittsburgh 1965
boulders
Wolf Creek Dam(b) Cumberland River 1978 Embankment 187 2.0 300 Center of 1:600 93.5 U. S. Army
Kentucky (200ft.); dam Engineer District,
cavernous Nashville 1975;
limestone Fetzer 1979
(100 ft.)
Mill Creek Dam Mill Creek 1981 Silt 51 2.0 170 Upstream 1:100 25.5 U. S. Army
Washington overlying toe of Engineer District,
cot)glomerate dam Walla Walla 1980a,b
Clemson Lowerb Hartwe 11 Lake 1982 Embankment 44 2.0 85 Center of 1:100 22.0 u. s. Army
9-40
(a)
(b) Deviation from vertical in any direction at any depth.
Concrete cutoff installed as remedial seepage control for existing dam (see Chapter 12).
Table 9-8. Comparison of Concrete Mix Proportions for Corps of Engineers Concrete Cutoff Walls
Wolf Creek Dam 564 123 275 0.40 1,619 1,369 6.5-7.5 3,000 4,685
Mill Creek Dam 299 l34(b) 260 0.60 1,628 1,587 8 1,500 1,970
Clemson Lower 507 110 334 0.54 1,540 1,310 60.9.0 1,300
Diversion Dam
9-41
EM 1110-2-1901
30 Sep 86
(a)
(b) Weights are saturated surface dry.
Fly ash percentage replaced by solid volume was raised from 35 to 45 before wall was completed.
EM 1110-2-1901
30 Sep 86
of 3/4 in. and a water cement ratio of 0.6, the permeability is usually lower
-10
than 10 cm/sec (Xanthakos 1979). The permeability of a concrete cutoff
wall is influenced by cracks in the finished structure and/or by void spaces
left in the concrete as a result of honeycombing or segregation (see Equa-
tion 9-4 and figure 9-5). The joints between panels are not completely
impermeable but the penetration of bentonite slurry into the soil in the
immediate vicinity of the joint usually keeps the flow of water very small
(Hanna 1978). Measured head efficiency for concrete cutoff walls from
piezometric data generally exceeds 90 percent (Telling, Menzies, and Simons
1978b). At Kinzua Dam (formerly Allegheny Dam), the measured head efficiency
was 100 percent, i.e., the head just downstream of the concrete cutoff wall was
of the magnitude established by vertical seepage through the upstream
connecting blanket (Fuquay 1968).
(9-13)
where
manufactured aggregate. Since the concrete is poured into the trench through
tremie pipes and displaces the bentonite slurry from the bottom of the exca-
vation upward, the concrete must have a consistency such that it will flow
under gravity and resist mixing with the bentonite slurry. Admixtures may be
used as required to develop the desired concrete mix characteristics. Fly ash
is often used to improve workability and to reduce heat generation. The unique
problems inherent at each project require studies to develop an adequate con-
crete mix (Holland and Turner 1980). Some typical concrete mixes used in Corps
of Engineers concrete cutoff walls are given in table 9-8. The placement
techniques used for the concrete are of equal importance in assuring a satis-
factory concrete cutoff wall.
9-43
EM 1110-2-1901
30 Sep 86
should be used. (1) The dry tremie is placed in the hole with a metal plate and
rubber gasket wired to the end of the tremie. The tremie pipe is lifted,
breaking the wires and allowing the concrete flow to begin. Concrete is added
to the hopper at a uniform rate to minimize free fall to the surface in the
pipe and obtain a continuous flow. The tremie apparatus is lifted during
placement at a rate that will maintain the bottom of the pipe submerged in
fresh concrete at all times and produce the flattest surface slope of concrete
that can practically be achieved. The flow rate (foot of height per hour) and
surface slope of the concrete shall be continuously measured during placement
with the use of a sounding line. A sufficient number of tremies should be
provided so that the concrete does not have to flow horizontally from a tremie
more than 10 ft. As soon as practical, core borings should be taken in
selected panels through the center of the cutoff wall to observe the quality of
the final project. Unacceptable zones of concrete such as honeycombed zones,
segregated zones, or uncemented zones found within the cored panels or elements
should be repaired or removed and replaced. One means of minimizing such
problems at the start of a job is to require a test section in a noncritical
area to allow changes in the construction procedure to be made early in the
project (Hallford 1983; Holland and Turner 1980; and Gerwick, Holland, and
Komendant 1981).
(1)
At Wolf Creek Dam concrete problems (areas of segregated sand or coarse
aggregate, voids, zones of trapped laitance, and honeycombed concrete)
occurred for tremie-placed 26-in. -diam cased primary elements. This must
be considered in future projects which involve tremie-placed elements of
small cross-sectional areas (Holland and Turner 1980).
9-44
EM 1110-2-1901
30 Sep 86
9-45
EM 1110-2-1901
30 Sep 86
likely. Concrete cutoff walls located under or near the toe of the dam are
subject to possible rupture from horizontal movements of the foundation soil
during embankment construction. This effect can be minimized by constructing
the dam embankment prior to the concrete cutoff wall. As mentioned previously,
concrete cutoff walls located under the center of the dam are subject to pos-
sible compressive failure due to negative skin friction as the foundation
settles under the weight of the embankment. The probability of this occurring
would depend upon the magnitude of the negative skin friction developed at the
interface between the concrete cutoff wall and the foundation soil and the
stress-strain characteristics of the concrete cutoff wall. Also, as previously
mentioned, a centrally located concrete cutoff wall may punch into and crack
the overlying core material unless an adequate connection is provided between
the concrete cutoff wall and the core of the dam.
f. Steel Sheetpiling.
9-46
EM 1110-2-1901
30 Sep 86
STRAIGHT ARCH Z
a. Sections
b. Interlocking of sections
(2) History of use. Steel sheetpiling was first used by the Corps of
Engineers to prevent underseepage at Fort Peck Dam, Montana (U. S. Army Engi-
neer District, Omaha 1982). The steel sheetpiling, driven to Bearpaw shale
bedrock with the aid of hydraulic spade jetting, reached a maximum depth of
163 ft in the valley section (see table 9-9). An original plan to force grout
into the interlocks of the steel sheetpiling was abandoned during construction
as impractical. Steel sheetpiling was used as an extra factor to prevent pip-
ing of foundation soils at Garrison Dam, North Dakota (U. S. Army Engineer
District, Omaha 1964). At Garrison Dam, underseepage control was provided for
by an upstream blanket and relief wells and the contribution of the steel
sheetpiling to reduction of underseepage was neglected in the design of the
relief wells. Steel sheetpiling and an upstream blanket were installed for
control underseepage at Oahe Dam, South Dakota. Relief wells were installed
for remedial seepage control to provide relief of excess hydrostatic pressures
developed by underseepage (U. S. Army Engineer District, Omaha 1961).
9-48
Table 9-9. Comparison of Corps of Engineers Steel Sheetpiling Cutoff Walla
(a) After 17 years (relief wells were installed in 1942 aa remedial underseepage control)
(b) After 10 years (upstream blanket and relief wells were constructed prior to impoundment)
(c) After 4 years (upstream blanket constructed prior to impoundment and relief wells installed in 1962 as remedial underseepage control)
9-49
EM 1110-2-1901
30 Sep 86
EM 1110-2-1901
30 Sep 86
9-50
EM 1110-2-1901
30 Sep 86
9-51
EM 1110-2-1901
30 Sep 86
9-52
EM 1110-2-1901
30 Sep 86
(5) Both the blanket and substratum have a constant thickness and are
horizontal.
(9-14)
where
(9-15)
where
9-53
EM 1110-2-1901
30 Sep 86
(9-16)
where
h o = pressure head under the blanket at the downstream toe of the dam
The critical pressure head under the blanket at the downstream toe of the dam
is
(9-17)
where
The factor of safety against uplift or heaving at the downstream toe of the dam
is
(9-18)
9-54
EM 1110-2-1901
30 Sep 86
(9-19)
where qf is the rate of discharge through the pervious foundation per unit
length of dam. The acceptable rate of discharge or underseepage depends upon
the value of the water or hydropower lost, availability of downstream right-
of-way, and facility for disposal of underseepage. The following procedure is
used to determine the length of an upstream blanket when there is a downstream
blanket present (see figure 9-19b):
(d) Determine the rate of discharge through the pervious foundation per
unit length of dam from equation 9-19. If the rate of discharge is excessive,
a reduction can be obtained by increasing the thickness of the upstream blanket
or reducing the permeability of the upstream blanket material by compaction.
When these methods are used, steps 1 to 4 are repeated before going to step 5.
(9-20)
(f) Enter figure 9-20 with c and L1 and obtain Lo , which is the
distance from the upstream toe of a homogeneous impervious dam or the imper-
vious core section of a zoned embankment to where a discontinuity in the up-
stream blanket will have no effect on the uplift at the downstream toe of the
dam or rate of discharge through the pervious foundation. This is the point
beyond which a natural blanket may be removed in a borrowing operation.
Also, Lo would represent the distance upstream from the toe of the dam to
which a streambed should be blanketed to ensure the continuity of a natural
upstream blanket. If there is no downstream blanket the pressure head under
the blanket at the downstream toe of the dam will be zero (see equation 9-16)
9-55
EM 1110-2-1901
30 Sep 86
and the following procedure is used to determine the length of the upstream
blanket:
9-56
EM 1110-2-1901
30 Sep 86
9-57
EM 1110-2-1901
30 Sep 86
9-58
EM 1110-2-1901
30 Sep 86
downstream natural blanket is present, the downstream seepage berm should have
a thickness so that the factor of safety against uplift or heaving at the down-
stream toe of the dam is at least 3 and width so that the factor of safety
against uplift at the downstream toe of the seepage berm is at least 1.5.
Formulas for the design of downstream seepage berms where a downstream natural
blanket is present are given in figure 9-22. If there is no downstream natural
blanket present, the need for a downstream seepage berm will be based upon
Bligh's creep ratio.
(9-21)
where
Minimum acceptable values of Bligh's creep ratio are given in table 9-10. If
the creep ratio is greater than the minimum value, a downstream seepage berm is
not required. (1) If the creep ratio is less than the minimum value, the width
of the downstream seepage berm should be made such that the creep ratio is
above the minimum value shown in table 9-10. The thickness of the downstream
seepage berm at the toe of the dam will be determined so that the factor of
safety against uplift or heaving at the downstream toe of the dam is at
least 3. The pressure head beneath the downstream seepage berm at the landside
toe of the levee is
(9-22)
where
h o = pressure head under the seepage berm at the downstream toe of the
dam
(1)
A downstream seepage berm may be required to correct other problems such
as excessive seepage gradients under the dam (could be detected by check-
ing the rate of underseepage).
9-59
EM 1110-2-1901
30 Sep 86
The rate of discharge through the pervious foundation per unit length of dam
is
(9-23)
where
9-60
X
PeRVIOUS SUBSTRATUM
NOTATIONS
BERM=~
X = REQUIRED BERM WIDTH
c = h0 = HEAD AT LANDS IDE TOE OF DAM WITHO\.)T
s + )(3 t = REQUIRED THICKNESS OF BERM AT TOE OF DAM
h~ = HEAD AT LANDSIDE TOE OF DAM WITH BERM
k 1 = VERTICAL PERMEABILITY OF BERM
SUBMERGED UNIT WEIGHT OF TOP STRATUM i 0 = ALLOWABLE UPWARD GRADIENT AT LANDSIDE TOE OF DAM
F = FACTOR OF SAFETY AGAINST UPLIFT AT TOE
SUBMERGED UNIT WEIGHT OF BERM
9-61
FORMULAS
IMPERVIOUS BERM (k, = 01 SEMIPERVIOUS BERM lk 1 - k0 ci SAND BERM PERVIOUS BERM WITH COLLECTOR
, Hx
3 h = h = -3 -
h , =H (:-x- +X
-) WHERE IN·
0 0 s + )(3
o S + x3 +X
A= 6 + 3 SC (r + 1)
1i
1+--
F 1w
h;,- 1 Z1
0 =~ (1 -~,;
EM 1110-2-1901
0
USE F;;. 1.5 t=---
1 + i0 B S + x3
30 Sep 86
Figure 9-22.
Design of downstream seepage berms where a downstream natural blanket is
120
present (from U.S. Army Engineer Waterways Experiment Station )
EM 1110-2-1901
30 Sep 86
Minimum Bligh's
Material Creep Ratio
Station120
become covered with sod and stabilized so that soil particles carried by sur-
face runoff and erosion will not clog the seepage berm. If it is necessary to
construct the downstream seepage berm at the time the earth dam is built or
before it has become covered with sod, an interceptor dike should be built at
the intersection of the downstream toe of the dam and the seepage berm to pre-
vent surface wash from clogging the seepage berm. A free-draining downstream
seepage berm is one composed or random fill overlying horizontal sand and
gravel drainage layers with a terminal perforated collector pipe system (U. S.
Army Engineer Waterways Experiment Station 1956a).
9-62
EM 1110-2-1901
30 Sep 86
V-lWTCH WEIR
METAL WELL GUARD -
VARIABLE- TO TOP OF
MINIMUM WATER TABLE
SAND BACKFILL
BLANK PIPE
THROUGH VERY
FINE SAND STRATA
o•.
0 ..
...
• 0
•• 0
...
0 ••
0 PERFORATED OR
••• SLOTTED SCREEN
MIN
9-63
EM 1110-2-1901
30 Sep 86
their tops are accessible for cleaning, sounding for sand, and pumping to
determine discharge capacity. Relief wells should discharge into open ditches
or into collector systems outside of the dam base which are independent of toe
drains or surface drainage systems. Experience with relief wells indicates
that with the passage of time the discharge of the wells will gradually
decrease due to clogging of the well screen and/or reservoir siltation. A
comprehensive study of the efficiency of relief wells along the Mississippi
River levee showed that the specific yield of 24 test wells decreased 33 per-
cent over a 15-year period. Incrustation on well screens and in gravel filters
was believed to be the major cause (Montgomery 1972). Therefore, the amount of
well screen area should be designed oversized and a piezometer system installed
between the wells to measure the seepage pressure, and if necessary additional
relief wells should be installed (EM 1110-2-2300, U. S. Army Engineer Waterways
Experiment Station 1956a, Singh and Sharma 1976).
9-64
EM 1110-2-1901
30 Sep 86
many dams to prevent excessive uplift pressures and piping through the
foundation.
c. Design Considerations.
9-65
EM 1110-2-1901
30 Sep 86
and the presence of any geologic features and/or man-made features which would
result in an open or blocked seepage exit. The procedure for computation of
the seepage exit distance, rate of discharge through the pervious foundation
per unit length of dam, and pressure head without relief wells is given in
figure 9-26. Generally relief wells have diameters of 6 to 18 in. and screen
lengths of 20 to 100 ft, depending on the requirements. Some types of screens
used for wells are slotted or perforated steel pipe, perforated steel pipe
wrapped with steel wire, slotted wood stave pipe, and slotted plastic pipe.
Riser pipe usually consist of the same material as the screen but does not
contain slots or perforations. The open area of a well screen should be suf-
ficiently large to maintain a low entrance velocity (< 0.1 ft per sec) at the
design flow in order to minimize head losses across the screen and reduce the
incrustation and corrosion rates. The entrance velocity is calculated by
dividing the expected or desired yield of the relief well by the area of open-
ings in the screen (Johnson Division, Universal Oil Products Co. 1972). Fil-
ter packs around relief wells are usually 6 to 8 in. thick and must meet the
criteria specified in Appendix D. Head losses within the relief well system
consist of entrance head loss, friction head loss in the screen and riser
pipes, and velocity head loss as shown in figure 9-27. The effective well
radius is that radius which would exist if there were no hydraulic head loss
into the well. For a well without a filter, the effective well radius is
one-half the outside diameter of the well screen. Where a filter has been
placed around the well, the effective well radius is the outside radius of the
well screen plus one-half of the thickness of the filter.
9-66
EM 1110-2-1901
30 Sep 86
(9-24)
where
9-67
EM 1110-2-1901
30 Sep 86
A factor •••••••••••••••••••••••••• c =
CASE l, 1
3
= ...
X ., -
1
3 c
h =h t-cx
X 0
(t = 2. 718)
1
x3 =c tanh(cL )
3
cosh· c(L
3
- x}
h
X = h
0 cosh(cL )
3
h
0
h
x=L3 = cosh (ct )
3
sinh c{L - x)
3
hx = ho sinh (cL )
3
hx=L =0
3
NOTE: x can be COI:If·t:.ted ~r-on. formulas for x 1 by
1
inserting the ~ength of upstrerum blanket L1 fO~ L
in the appropn.ute exnressi<"'r, when upstrenm condl.tiods
are similar to the above do~nstreP~ conditions.
Figure 9-26. Computation of rate of discharge and pressure heads for semi-
pervious downstream top stratum and no relief wells (from U. S. Army
120
Engineer Waterways Experiment Station )
9-68
TYPE OF PIPE c
STEEL \NEW) I ?5 O.fi7
STEEL IAVG CONDITION! ,, a· 0.83
WOOD STAVE RISER 130 0.71
fOTAL. HYDRAULIC HEAD LOSS IN A WELL (Hw) IS WOOD STAVE SCF<EEN 100 .00
CORRUGATED M£TAL 70 .92
Jr
~ r"
~"<HERE He ENTRANCE HEAD LOSS (SCREEN AND FILTER)
t- Jr
ESTIMATE FROM CURVE a "- 0
Ul •
H HEAD LOSS IN SCREENED SECTION OF WELL "'I
0_j
' fSTIMATE FROM CURVE b FOR A DISTANCE OF ONE- w
•
(b)
\0.0
3.0~-------.----.----.----.---.----,----,----,----,
s.o
w f-
u "- 3.0
z ..::
<t
Jr f- !
';; "- 1.0
LJj I 2,0 "'"'0 0.5
z • _j
WI
w-
Jr Ul ..
0
w
0.3
~0 I
.J
u }- 0.1
"7 {J t .0 DEVELOPED FROM----- 1-
;;: "w PUMPING TEST u 0.05
rr r DATA. 16-IN.-DIAM 0
w .J
WELL,\00-SQ-IN SCREEN w 0.03
EM 1110-2-1901
f-
_j
WITH 5/32-IN SLOTS, >
6-IN.·THICK FILTER
Co - - - J. - - - -- _j ____..L____j__ _..L__ _ i_ _.J....____j_ _....l
0 1C 20 30 40 so
30 Sep 86
Wl:LL DISCHARGE PER FT OF SCREEN, GPM DISCHARGE, GPM
(a) (c)
205
Figure 9-27. Hydraulic head loss in relief wells (after Leonards )
EM 1110-2-1901
30 Sep 86
(9-25)
(9-26)
(9-27)
(9-28)
where
The percent penetration of the well screen into the transformed pervious
foundation is
9-70
EM 1110-2-1901
30 Sep 86
(9-29)
(9-30)
where
The factor of safety against uplift or heaving at the downstream toe of the dam
(1)
provided by the relief well system should be at least 1.5 .
(4) Infinite and Finite Relief Well Systems. Formulas for the design
of relief wells are based on the assumption that the flow is laminar,
(1) Relief wells should be designed to reduce the excess head to zero to pre-
vent upward seepage from occurring beneath the downstream top stratum.
9-71
EM 1110-2-1901
30 Sep 86
artesian, and continuous and that a steady-state condition exists. Relief well
systems are considered to be infinite or finite in length. The term infinite
is applied to a system of wells that conforms approximately to the following
idealized conditions:
(d) The boundaries at the ends of the relief wells are impervious,
normal to the line of wells, and at a distance equal to one-half the well
spacing beyond the end wells of the system.
If these conditions exist, the flow to each well and the pressure distribution
around each well are uniform for all wells along the line. Therefore, there
is no flow across planes centered between wells and normal to the line, hence
no overall longitudinal component of flow exists anywhere in the system. The
term infinite is applied to such a system because it may be analyzed mathe-
matically by considering an infinite number of wells; the actual number of
wells in the system may be from one to infinity. Normally, a line of relief
wells below a dam extending entirely across a valley and terminating at rela-
tively impervious valley walls should be designed as an infinite line. A
finite system of wells in any system that does not approximate the idealized
condition for the infinite system. Whenever a major and abrupt change in the
character of the system such as penetration or well spacing might result in
an appreciable component of flow parallel to the line of wells, the use of
design procedures for finite systems will be used (see U. S. Army Corps of
Engineers 1963).
(9-31)
(1) Also applicable-when the top stratum is semipervious provided the well
system is located in a drainage ditch and the head is kept below the
ground surface on the downstream side of the dam to prevent any seepage
upward through the top stratum. Under these conditions, the downstream
top stratum acts as if it is impervious.
9-72
EM 1110-2-1901
30 Sep 86
where
a = well spacing
(9-32)
where
The total drawdown at the well, neglecting hydraulic head losses in the well,
is that at the slot plus that due to the well
(9-33)
(9-34)
9-73
EM 1110-2-1901
30 Sep 86
The head midway between wells will exceed the head at the well by
(9-35)
where is the head increase midway between wells. The drawdown midway
between wells is
(9-36)
At a distance downstream from the well system, the head will exceed that at
the well by
9-74
EM 1110-2-1901
30 Sep 86
(9-37)
(9-38)
(9-39)
where is the average uplift factor (obtained from figure 9-29). The
total drawdown at the partially penetrating well, neglecting hydraulic head
losses in the well, is that at the slot plus that due to the well
(9-40)
The head midway between partially penetrating wells will exceed the head at
the well by
(9-41)
where
is the midpoint uplift factor (obtained from figure 9-29). The draw-
down midway between partially penetrating wells is
(9-42)
9-75
EM 1110-2-1901
30 Sep 86
a.lrw
~i
.
-
~ '
t· ~
I - 0.5
o.l
- - - -
I
·- I
I
~ru
\()0 - - -
•.!!.
0
'!-0 It i
10
iS /!'l
l
I
- -- I
2' >---
,o ! -- E
Q I I CD
....
...."'"'
~!l !
--
- --
Q_
.I I "
a>
_J
.J
cJ
~
t[
II
IIj ---- -f-
I
20
.... ·~ '1 j
I
2 EXAMPLE.
II
w
<)
a: ,.;: Where Q /rw :100, 9av~~ I 0, 9'6-= 1.0
w
"- Step 1.· Locate Sav: I 0 on ca/rw line a 100 and mark point
2>
- Step 2. Locate 0 /Q = 1.0 on Ola. hne for 9av
~; Step ·3 Place tr•onqle edc;~e on obova point$ located by steps I ond 2.
/
Step 4 Slide tnonqle parallel to intttol position to po.nt 0/a : 1.00 on Q/a line for 9m
Reod off value at 9m on a/rw l1ne = 100.
--
-1 - - -
-
3.0
Step 5.
Step 6.
em: 1.05
Slide tnongle parallel to 1nil!ol posttion to pole potnt on 0/a line of Oav
Step 7. Read off percent screen penetrotton = 46% on alrw ""1000
g g
N
qlr,.
Figure 9-29. Nomograph for obtaining uplift factors for design of partially penetrating
120
relief wells (from U.S. Army Engineer Waterways Experiment Station )
EM 1110-2-1901
30 Sep 86
(a) Compute the allowable pressure head under the top stratum at the
downstream toe of the dam from
(9-43)
where
(b) Assume that the net head in the plane of the wells equals the
allowable pressure head under the top stratum at the downstream toe of the dam
and compute the net seepage gradient toward the well line
(9-44)
where
9-77
EM 1110-2-1901
30 Sep 86
(9-45)
(c) Assume a well spacing and compute the flow from a single well
(9-46)
where
a = well spacing
(d) Estimate the total hydraulic head loss in the well from
figure 9-27.
(e) Compute the net average head in the plane of wells above the total
head loss in the well including elevation head loss (see figure 9-25) from
(9-47)
where
h a v g = net average head in the plane of wells above the total head loss
in the well including elevation head loss
9-78
EM 1110-2-1901
30 Sep 86
(9-48)
(h) The first trial well spacing is that of value a for which
from step (f) equals from step (g).
(i) Find from figure 9-29 for the first trial well spacing and the
corresponding values of a/rw and D/a .
(j) If repeat steps (c) to (i) using the first trial well
spacing in lieu of the spacing originally used in step (c), and determine the
second trial well spacing. This procedure should be repeated until relatively
consistent values of a are obtained on two successive trials. Usually the
second trial spacing is sufficiently accurate.
(m) Compute the net head beneath the top stratum midway between the
wells above the total head loss in the well including elevation head loss (see
figure 9-25) from
(9-49)
where
hm = net head beneath the top stratum midway between the wells above the
total head loss in the well including elevation head loss
9-79
EM 1110-2-1901
30 Sep 86
= net head beneath the top stratum midway between the wells
Hm
= total head loss in the well including elevation head loss
HW
(n) Using from steps (h) and (i), respectively, compute
ha v g from
(9-50)
(o) Using Hw and h a v g from steps (1) and (n), respectively, and
compute H a v g f r o m
(9-51)
(q) Using hm and from steps (m) and (p), respectively, compute
for various values of a from
(9-52)
(r) Find from figure 9-29 for the values of a used in step (q)
and the corresponding a/r w and D/a values.
(s) The second trial well spacing is that value of a which from
step (q) equals from step (r).
(t) Find from figure 9-29 for the second trial well spacing and
the corresponding values of a/rw and D/a .
(u) Determine the third trial well spacing by repeating steps (k) to
(t) using the second trial well spacing in lieu of the spacing originally
assumed in step (k), and in step (n) using the values of and from
steps (s) and (t), respectively, instead of those from steps (h) and (i). This
procedure should be repeated until relatively consistent values of a are
obtained on two successive trials. Normally, the third trial is sufficiently
accurate.
9-80
EM 1110-2-1901
30 Sep 86
9-81
4.0
Assumptions:
EM 1110-2-1901
30 Sep 86
~ .!. x3
(1) Penetration of pervious aquifer by ve11 screen = 100%
1 50 250 {2) Semipervious top stratum infinite in landvard extent
{3) Effective ve11 radius = 1.0 ft
3.~ 2 10() 500-
Note:
3 150 '500
(l)H = Head midvay betveer. vells in infinite vell line
4 200 500 m...
....
..{
IX'
(3) a= Well specing in feet
9-82
(4) x
3
= Exit length in feet
l.O r-----~~r\--------t---------T---~----T---------+---------+---------+---------J
Figure 9-31. Ratio of head midway between relief wells at center of a finite well system to
head midway between wells in an infinite system (from U.S. Army Engineer Waterways Experi-
.
ment s tat1on
120)
EM 1110-2-1901
30 Sep 86
(1) For partially penetrating relief wells, the bottom plug should be such
that future screen extension will be possible,
9-83
EM 1110-2-1901
30 Sep 86
conjunction with relief well systems to collect seepage in the upper pervious
foundation that the deeper relief wells do not drain. If the volume of seepage
is sufficiently large, the trench drain is provided with a perforated pipe. A
trench drain with a collector pipe also provides a means of measuring seepage
quantities and of detecting the location of any excessive seepage (U. S. Army
Engineer Waterways Experiment Station 1956a, EM 1110-2-1911, EM 1110-2-1913,
and Cedergren 1977).
9-84
EM 1110-2-1901
30 Sep 86
9-85
EM 1110-2-1901
30 Sep 86
b. Drain holes drilled in the rock foundation downstream from the grout
curtain can be discharged into the gallery and observations of the quantities
of seepage in these drain holes will indicate where foundation leaks are
occurring.
9-86
EM 1110-2-1901
30 Sep 86
The minimum size cross section recommended for galleries and access shafts is 8
by 8 ft to accommodate drilling and grouting equipment. A gutter located along
the upstream wall of the gallery along the line of grout holes will carry away
cuttings from the drilling operation and waste grout from the grouting opera-
tion. A gutter and system of weirs located along the downstream wall of the
gallery will allow for determination of separate flow rates for foundation
drains (EM 1110-2-3502, and Blind 1982).
9-87
EM 1110-2-1901
30 Sep 86
CHAPTER 10
SEEPAGE CONTROL THROUGH EARTH ABUTMENTS ADJACENT TO STRUCTURES
AND BENEATH SPILLWAYS AND STILLING BASINS
10-2. Adjacent to Outlet Conduits. When the dam foundation consists of com-
pressible soils, the outlet works tower and conduit should be founded upon or
in stronger abutment soils or rock. When conduits are laid in excavated
trenches in soil foundations, concrete seepage cutoff collars shall not be
provided solely for the purpose of increasing seepage resistance since their
presence often results in poorly compacted backfill around the conduit. Col-
lars, with a minimum projection from the conduit surface, will be used over
conduit joints to protect against joint displacements resulting from differen-
tial movement on yielding foundations. Excavations for outlet conduits in
soil foundations shall be wide enough to allow for backfill compaction
parallel to the conduit using heavy rolling compaction equipment. Equipment
used to compact along the conduit should be free of framing that prevents its
load transferring wheels or drum from working against the structure. Excavated
slopes in soil for conduits should be no steeper than 1V to 2H to facilitate
adequate compaction and bonding of backfill with the sides of the excavation.
Drainage layers should be provided around the conduit in the downstream zone of
embankments without pervious shells. A concrete plug shall be used as backfill
in rock cuts for cut-and-cover conduits within the core zone to ensure a water-
tight bond between the conduit and vertical rock surfaces. The plug, which can
be constructed of lean concrete, should be at least 50 ft long and extend up to
the original rock surface. In embankments having a random or an impervious
downstream shell, horizontal drainage layers should be placed along the sides
and over the top of conduits downstream of the impervious core. Where outlet
structures are to be located in active seismic areas, special attention must be
given to the possibility of movement along-existing or possibly new faults
(EM 1110-2-2300).
10-1
EM 1110-2-1901
30 Sep 86
EL. 1169 TOP Of DAM
------------------------------------
1150
[
--------c------ 1100
10-2
I
I
L SANDS &
GRAVELS\
ASSUMED SEEPAGE
CHANNEL LIMITS
END SEEPAGE
CUTOff
1050
----------t------ T1LL"\.
25+00 ' 15+00
Figure 10-1. Profile of left pervious abutment at North Branch of Kokosing Dam, Ohio
87
(from U. S. Army Engineer District, Huntington )
EM 1110-2-1901
30 Sep 86
10-3
EM 1110-2-1901
30 Sep 86
10-4
EM 1110-2-1901
30 Sep 86
CHAPTER 11
SEEPAGE CONTROL IN ROCK FOUNDATIONS AND ABUTMENTS
11-1
EM 1110-2-1901
30 Sep 86
c. Blankets may sometimes give adequate control of seepage water for low
head structures, but for high head structures it is usually necessary to incor-
porate a downstream drainage system as a part of the overall seepage-control
design. The benefits derived from abutment impervious blankets are due to the
dissipation of a part of the reservoir head through the blanket. The propor-
tion of head dissipated is dependent upon the thickness, length, and effective
permeability of the blanket in relation to the permeability of the adjacent
soil, or rock.
11-2
EM 1110-2-1901
30 Sep 86
c. Grouting of steep abutment slopes has the potential for causing dif-
ficulties and possibly can do more harm than good. Care must be exercised when
grouting in abutments to avoid displacements within the rock mass. Even rela-
tive low grouting pressures can cause joint opening and decrease the integrity
of the abutment.
11-4
EM 1110-2-1901
30 Sep 86
CHAPTER 12
REMEDIAL SEEPAGE CONTROL
(2) Risk.
(4) cost.
12-2
EM 1110-2-1901
30 Sep 86
b. 140-ft-High Earth Dam (Ley 1974). In the left abutment, gypsum had
apparently formed in the bedding planes and fractures of folded and faulted
shale and siltstone. After water was impounded, leakage, carrying dissolved
gypsum, occurred from the abutment. Settlement and gradual increase in seep-
age also indicated that gypsum was being removed from the formation. Built in
1915, the dam underwent a grouting program from 1930-1933, resulting in place-
ment of about 35,000 cu ft of grout in a series of holes along the dam crest,
the left abutment, and at the bottom of the hill forming the left abutment.
This program reduced seepage quantities by 75 percent. Approximately 30 years
later, over 32,000 cu ft of cement-bentonite grout (colored with iron oxide to
distinguish from previously placed grout) was placed in 137 holes to again
reduce seepage and replace material removed by solution. Cores indicated good
penetration with most seams from hairline to 1/8 in. thick. Seepage was
greatly reduced. Other geologic materials may also be dissolved when subjected
to seepage. In a similar manner, silt and clay in limestone cavities may also
be removed by seepage. Grouting may only be a temporary solution to a seepage
problem if solution of a soluble foundation continues after grouting.
12-3
EM 1110-2-1901
30 Sep 86
e. Hills Creek Dam (Jenkins and Bankofier 1972). Hills Creek Dam, con-
structed by the Portland District, has a maximum height of 338 ft and consists
of a central impervious core with gravel and rock shells. Minor seepage occur-
red near the left abutment during first filling, but decreased with time.
Seepage markedly increased in extent and volume after 6 years of normal opera-
tion. Vertical drains placed in the downstream shell as an initial remedial
measure were not effective in lowering water levels in the downstream shell and
seepage continued to increase. An investigation to determine the seepage
source continued during the remedial action. Initially it was thought that
leakage was through the upstream blanket into the foundation and abutment, but
further observations indicated flow was through the core or core-foundation
contact. Grouting , which injected 4,500 sacks of cement, most in a 1-1/2:1 mix
at zero psi, resulted in elimination of almost all seepage. Four 42-in. bucket
auger holes, as well as several smaller borings, were drilled to inspect
grouting of the core and foundation. The main source of seepage was at a point
along the core foundation contact where a haul road had crossed the abutment.
Twelve years later, seepage is still negligible. Frequently, the source of
seepage is not obvious. The engineer must consider all possibilities and,
after choosing and installing a remedial measure, try to understand what post-
remedial monitoring is indicating. The extent of the engineer's knowledge of
the foundation, embankment materials, and construction history will greatly
influence the accuracy of his analysis of the seepage problem. Often available
foundation and construction information will not be adequate and further
geotechnical investigation will be required.
12-4
EM 1110-2-1901
30 Sep 86
materials are usually placed on the reservoir bottom. If sloped areas such as
the reservoir sides of upstream embankment slope are to be sealed, considera-
tion must be given to protection against wave attack and erosion from runoff.
Additionally, fine-grained materials placed on the upstream embankment slope
may be removed during drawdown because of low saturated strength and high
saturated weight. If seepage can also go through the upstream portion of the
embankment and then into the foundation an upstream blanket will be less effec-
tive and another remedy may be necessary, e.g., cutoff beneath dam, fig-
ure 12-1. The nature of reservoir bottom materials must be considered. Any
large voids must be filled with a stable material such as compacted soil,
stabilized soil, concrete etc. High gradients will likely exist through the
blanket during high reservoir levels, particularly close to the embankment. It
may be necessary to place a filter material before placing the blanket to pre-
vent piping of the blanket material into the foundation. The extent of the
blanket is determined by analysis and will depend on several factors, including
extent of desired decrease in seepage quantities and pressures and blanket
material available (quantity and permeability) (EM 1110-2-1913 and Barron
1977). Man-made liners have provided a seal for reservoirs with pervious
foundations when fine-grained materials were not economically available. They
are usually rather expensive, require relatively smooth surface for placement,
and coverings (normally soil) to protect them from puncture in stressed areas
and deteriorating exposure to sunlight. Joining of sections is one of the most
critical and difficult aspects of man-made liners. Field seams, especially
under difficult field conditions and with other than highly experienced person-
nel, can be an appreciable source of leakage. Quality control of seaming
should be strict. One example of the use of an impervious upstream blanket was
given in paragraph 12-3c; another is provided below:
12-5
EM 1110-2-1901
30 Sep 86
the piping criteria, D15 (alluvium) < 5D85 (blanket), Appendix D. As the
reservoir emptied after first filling, several sinkholes and cracks were noted
in the blanket. Sinkholes ranged from 1 to 15 ft in diameter and 4 to 6 ft in
depth. It was felt that uneven settlement during the first reservoir filling
caused tension and compression cracks in the blanket which allowed considerable
seepage into the underlying sand-choked gravel. In areas where the sand was
less dense, the seepage moved the sand down to form a layer in the lower part
of the gravel. This created open work gravel just beneath the blanket, and
fines from the blanket moved into and through this open layer forming the sink-
holes. Sinkholes were filled with filter material and mounded over with
blanket material. Typically, the blanket mounds were approximately 15 ft high
and extended 30 to 35 ft beyond the sinkhole edge. After filling of the reser-
voir, sinkholes were located by side-scan sonar and filled with a mixture of
filter material and silt from self-propelled bottom dump barges. Each sinkhole
generally received 50 barge loads of material. Sinkholes continued to be dis-
covered and covered over another 3-4 years after the initial remedial action.
Siltation on the reservoir blanket and filling of sinkholes have reduced
seepage about one half.
12-6. Downstream Berm. Berms control seepage by increasing the weight of the
top stratum so that the weight of the berm plus top stratum is sufficient to
resist uplift pressure. If of low permeability, they will reduce seepage, but
increase uplift pressures beneath the downstream toe of the dam since they
force seepage to exit further downstream of the dam. If pervious, they must
be designed as a filter or with an underlying filter to prevent upward migra-
tion of line particles from the foundation materials beneath them. Again, a
seepage analysis must be made to determine the resisting load required of the
berm. Downstream slope stability of the embankment will normally increase
because of the resistance to sliding provided by the berm. Huntington District
has employed berms as remedial measures at several flood control dams in the
Muskingum River flood control system (Coffman and Franks 1982). Similarity of
the embankments and environments allowed a standard remedial action for several
of the dams at the downstream embankment toe. A 3- to 7-ft-thick pervious
blanket of appropriate length is placed over the soft seepage areas at the
downstream toe. This adds weight and provides a working platform for instal-
lation of relief wells at points of excessive seepage.. Another example of a
stability berm is given in the Addicks and Barker Dams example,
paragraph 12-7a.
12-7. Slurry Trench Cutoff. Two major technical considerations in the use of
slurry trenches as remedial seepage control measures are (a) the effect on
stability of the embankment due to excavation of the trench and the presence
of a vertical plane of relatively weak soil (in the case of a soil-bentonite
backfill) and (b) tying the slurry trench to other existing or proposed seep-
age control measures. If a competent upstream blanket exists, the trench may
be placed upstream of the embankment and tied to the blanket or may be placed
through the dam and any pervious substratum if stability requirements are met.
A cement-bentonite backfill may be placed in panels or a concrete wall may be
placed in separately excavated elements if an open trench and the relatively
weak soil-bentonite backfill are unacceptable because of stability risks. The
following experiences with slurry trenches provide general examples of this
cutoff type as a remedial measure.
12-6
EM 1110-2-1901
30 Sep 86
a. Addicks and Barker Dams, Houston, Tex. (U. S. Army Engineer District,
Galveston 1977a; U. S. Army Engineer District, Galveston 1977b; U. S. Army
Engineer District, Galveston 1983). Completed in the late 1940's, Addicks and
Barker Dams are rolled earth embankments providing flood control in the Hous-
ton, Texas, area, with respective maximum heights of 48.5 and 36.5 ft above
streambed. Neither normally impound water except in periods of rainfall. The
embankments contain some silts and sands, foundations have silt and sand
layers, and upstream borrow areas expose the foundation permeable layers. At
the time of construction, these conditions were not considered, significant
because of the large discharge capability and short detention time of the
reservoirs. Residential and commercial development of the downstream local
area caused several changes in operating conditions which increased detention
time and made the effect of seepage more critical. These included restriction
of discharge rates and construction of drainage channels on non-Federal land
within 200-300 ft downstream of the center line of the dams which expose the
pervious portion of the foundation. Erosion of the drainage channel slopes on
the side of the channel nearest the dam and boils in the channel bottom during
times of low reservoir impoundment indicated the potential for dangerous seep-
age conditions during high reservoir levels. Downstream piezometers also
indicated a quick response to changes in reservoir levels. This example
describes remedial actions at Addicks Dam; actions at Barker Dam were similar.
Several remedial measures were considered:
(2) Downstream drainage blanket, stability berm, and relief well system
and downstream slurry trench - (relief wells between embankment toe and slurry
trench) very positive control (blanket and berm control embankment seepage
while wells and slurry trench control underseepage), but very costly, long-
term well maintenance required, and all seepage forces would be directed at
the embankment toe.
(3) Same plan as (2) except slurry trench replaced with steel sheet pile
cutoff - same reasoning as (2) except sheet pile would greatly increase cost.
12-7
EM 1110-2-1901
30 Sep 86
!i
SLURRY
PORTION OF EMBANKMENT THAT MAY BE REMOVED TRENCH EXISTING EMBANKMENT
DURING CONSTRUCTION OF SLURRY TRENCH AND
REPLACED AFTER COMPLETION OF SLURRY TRENCH
UPSTREAM DOWNSTREAM
LIMIT OF DEGRADING ROWl
---
_;.r;:::-'
3 - ...._ ....__2_, 1 EXISTING GROUND
1
- --
EL 94.3--........
--~,_,
~EXISTING GROUND
_..- _,.- _.-
TYPICAL SECTION
STA 350+50 TO STA. 377+00
NO SCALE
12-8
Ci SLURRY TRENCH
-------------r'-orrr~~--------~~-
Figure 12-2. Addicks Dams remedial slurry trench for embankment and foundation seepage
83
control (from U.S. Army Engineer District, Galveston )
EM 1110-2-1901
30 Sep 86
plugged with cloth. Backfill gradation is shown in table 12-1. For some por-
tions of the project, percent passing for the No. 200 sieve were 15-30 per-
cent. Backfill mixing and transport to the trench were conducted in several
ways. Some backfill was batched dry, placed in concrete trucks with slurry
added, then mixed and transported to the trench. The higher fines content
backfill in some cases proved too sticky to mix in trucks. Mixing was con-
ducted on the ground next to the trench but occasionally excess fines were
picked up from the working surface. A concrete mixing pad was used as an
alternative though wear from the mixing equipment destroyed the concrete.
Excess or unsatisfactory material was deposited in old borrow areas upstream of
the embankment to reduce underseepage. In one area, a slurry trench located at
the upstream toe of the embankment provided underseepage control while a down-
stream berm provided embankment stabilization. The berm of sandy clay had
permeability characteristics similar to the embankment and provided a 1V on 8H
slope. Several of the discharge conduits which suffered from seepage and pip-
ing were resealed, after cleaning, with ethafoam backer rods and a polyurethane
sealant. Where the sealant would not adhere to the concrete, joints were
talked with oakum soaked with a grouting compound. Well screens were placed in
weep holes to prevent loss of soil, and relief wells with submersible pumps
were installed. For certain portions of Barker Dam, use of an upstream clay
blanket and a downstream stability berm (1V on 8H) was more cost effective than
a slurry trench. There was intermittent surface exposure of pervious founda-
tion materials and a source of CH materials for the blanket was available
within the reservoir. Prior to placement of the blanket, ponded water and soft
surface materials were removed. The blanket was placed in 8-in. layers and
compacted with tamping rollers at natural moisture content.
Table 12-1. Backfill Mix for Slurry Trench, Addicks Dam (a)
Though not yet severely tested, the control measures have performed satisfac-
torily based on the following observations:
12-9
EM 1110-2-1901
30 Sep 86
(b) Phreatic surface has been raised upstream of the slurry trench.
b. Wolf Creek Dam, Ky. (Fetzer 1979). Constructed in the 1940's, Wolf
Creek Dam is a 200-ft-high combination earthfill and concrete dam founded on
limestone containing shale and solution cavities. During excavation of a
10-ft-wide cutoff trench, several interconnected solution cavities were dis-
covered in the limestone. These were backfilled for a short distance with
impervious material, and a 50-ft-deep single-line grout curtain was placed
beneath the bottom of the cutoff trench. In 1967, muddy flow was observed in
the tailrace, a small sinkhole developed near the downstream toe, and wet areas
existed near the downstream toe. In 1968, a larger sinkhole developed (13 ft
wide, 10 ft deep) and drilling revealed solution features running perpendicular
and parallel to the dam axis. It was concluded that reservoir water was
passing beneath the cutoff trench. Grout lines were placed along the dam axis
near the embankment-concrete contact and downstream of this area. During
1971-1972, an overall assessment of the seepage problem was made since the
remedial grouting had only addressed about 200 ft of the 4,000-ft embankment
portion of the dam. A diaphragm concrete cutoff wall was considered the best
solution because it could be installed without draining the reservoir, a very
costly operation due to reservoir use. Explorations , which included borings
spaced on 3.1-ft centers along the axis of the wall (parallel to the dam axis),
defined the depth and length of the wall. Depth was 10 ft below the lowest
indication of solution activity (maximum depth 278 ft) and length was 2,239 ft.
In 1974, a request for technical proposals resulted in seven proposals with two
acceptable. In the second stage, a bid invitation was issued and an award was
made for a wall in the area of the switchyard and 989 ft of the wall along the
dam axis. The award in 1975 was followed by a second competition and an award
in 1977 for the remaining 1,250 ft of the axis wall. The wall consists of
alternate cylindrical primary elements and connecting secondary elements
installed using bentonite slurry, figure 9-14. Primary elements are
2.17-ft-diam steel casings filled with tremied concrete (see table 9-8 for mix
proportions). Weak cement grout fills the volume between excavation walls and
the casing. A 25-ft-deep core hole was drilled beyond the bottom of each pri-
mary element to explore for cavities and was pressure-tested and grouted prior
to the placement of a closed-end primary casing. The primary element was
required to set for a minimum of 20 days before excavation of the secondary
element which is also filled with tremied concrete. Frequent piezometer
readings (as often as every 4 hours) were made during construction to determine
the hydraulic condition of the embankment and foundation and warn of any
potentially critical seepage conditions. Excavation and drilling were closely
monitored to observe any drill rod drops or mud losses. Sealers and reserve
mud, constantly on hand, provided for emergencies. Grout takes around
12-10
EM 1110-2-1901
30 Sep 86
the primary casings and volume of concrete used in the secondary elements were
closely monitored as was the embankment in general. Efficient management of a
large number of observations was necessary to determine the current condition
of the dam. The lack of major losses of slurry, grout, or concrete during
construction was probably due to the densely spaced borings and grouting done
during the earlier exploration program. Wall construction, completed in 1979,
took approximately 4 years and two construction contracts. Subsequent
piezometric levels indicate the wall is a successful seepage barrier.
12-11
EM 1110-2-1901
30 Sep 86
Table 12-2. Backfill Mix for Slurry Trench, Camanche Dike 2 (a)
12-12
EM 1110-2-1901
30 Sep 86
CHAPTER 13
MONITORING PERFORMANCE OF SEEPAGE CONTROL MEASURES
13-1
EM 1110-2-1901
30 Sep 86
13-2
NOTE: ALTHOUGH PIEZOMETER
LEADS ARE SHOWN
LEGEND EXTENDING THROUGH
IMPERVIOUS ZONE, THIS
o EMBANKMENT PIEZOMETER IS NO LONGER PERMITTED.
• FOUNDATION PIEZOMETER
i PROJECT CONTAINS OTHER
TYPES OF INSTRUMENTATION
THAT ARE NOT SHOWN.
13-3
PERVIOUS FOUNDATION
SOIL·S
SCALE IN FEET
40 0 80
Figure 13-2. Piezometer installation for dams on pervious foundations, Deer Creek Dam (from
EM 1110-2-1908)
EM 1110-2-1901
30 Sep 86
EM 1110-2-1901
30 Sep 86
~ NOTE: ALL PIEZOMETERS ARt OPEN-SYSTEM
TYPE. PROJECT CONTAINS OTHER TYPES
0 U.AI!IANK'-IENT PIEZOMETER OF INSTRU'-IENTATION THAT ARE NOT
e FOUNDATION PIEZOMETER SHOWN.
RCLArtVELY PERVIOUS
UNCOMPACrCD MArERIAL
13-4
roc DRAIN
SCALE IN fEET
40 00
Figure 13-3. Piezometer installation with a thin impervious topstratum and relief well
systems, Enid Dam (from EM 1110-2-1908)
! SEE SHEETS 249 AND 251
FOil LOCATION)
;
8 & ~ I~ •
8
0
8
'
A-111
I
13-5
A·ZIID I
I
SA· S
LEGEND
EM 1110-2-1901
30 Sep 86
Figure 13-4. Piezometer installation for dam abutments having grout
curtains or cut-slope drains, Applegate Lake Dam (from U. S. Army
104
Engineer District, Portland )
EM 1110-2-1901
30 Sep 86
D\STANCl IN f'EET I="ROM Ct.NTI:._'<.LI~'E or OMvl
800 700 600 500 400 300 200 \00 0 \00 200 300 ~00 sor; 6C:i ?OO 800
,-.-----T - --~ --- T --- T T- r---·r-------r-- y-----T - ---~-- -r --r- - --,-~
SELECT IMPERVIOUS
FILL----
o?o
18 "RIPPAP ON 6 "8£0/)/A/r. -
S60
24 ~RIPRAP (}N 4 ,.8!00/NG
650
J €>40
"'
~ 630
ti- b20
w
....
;;e 610
~ GOO
13-6
j: 590
~
~ '580
w
570
se.o
550
3' HO/'!Z0/1/r/lt.
ORfil•'-46£ R!.•'lhfXT
[OUFrTOf' SYSTEM_/
Figure 13-5. Estimated seepage for Aubrey Dam, Texas (from U. S. Army Engineer
81
District, Fort Worth )
EM 1110-2-1901
30 Sep 86
13-7
EM 1110-2-1901
30 Sep 86
GROUT TREATMENT
DIAPHRAGM WALL I I
Ol 0 100 200 300 400 500 609 BOO 900m
~~----_.~~~~~-r~T-~-r~-r~~-r~~-r~~~~~~~~-r~-r-r-r-+--------~~
330 330
320 320
310 310
300 300
290 290
280 mO.D. m00. 280
340-
ICEY
610 620 630 640 650 660 670
13-8
Figure 13-6. Layout of Baldershead Dam piezometer installation in upstream shale fill
212
(courtesy of British Geotechnical Society )
TYPICAL SECTION- FLOOD PLAIN
13-9
• SUFFACE MONUMENT
bea:. Y •lo 100
I
+ POPE PRESSURE DEVICE IN SILT FOUNDATION
SUSQUEHANNA RIVER BASIN
..t. EMBANKMENT PORE PRESSURE DEVICE
TIOGA· H~MWOND LAKES
Figure 13-7. Piezometer installation for dams with through seepage and underseepage (from
79
U. S. Army Engineer District, Baltimore )
EM
1110-2-1901
30 Sep 86
EM 1110-2-1901
30 Sep 86
OOWNSTREAM
HP 04lPAOJI
20' MAINTENANCE BERN
00
..,
SCALE OF FEET
IDO
13-10
LEGEND
• SURFACE MONUMENT
Figure 13-8. Piezometer installation for dams with lateral drains (from
79
u. S. Army Engineer District, Baltimore )
•-••cr-t:•n
A TT"' I'IU:OifUIIIII'OUifDoUICNII
- TIUIU:II •110 IIOWIII c•kl •t.IIIIIIIIIT PLAN J:h
13-11
/
SUftfAt:l[ MONUMUT , "UOM[T[ft AND
SECTION@
EM 1110-2-1901
Figure 13-9. Drain and piezometer installation, Applegate Lake, Oregon (from
Portland 104 )
30 Sep 86
·
U• S • Army Eng1neer D1str1ct,
· ·
EM 1110-2-1901
30 Sep 86
e. Near Relief Wells. If a relief well system has been installed, pore
pressures should be checked in the vicinity to evaluate the efficiency of the
system. Piezometers should be located both upstream and downstream of the
relief wells and should intercept the stratum being drained, figure 13-3. If
there is a line of wells, piezometers should also be installed generally at
the midpoint between the wells or at the point expected to have the highest
pore pressures.
c. Relief Wells. Relief wells, as the name implies, are widely used to
relieve pressures and control seepage through pervious strata beneath earth
dams, spillways, and outlet works. A thorough knowledge of the geologic con-
ditions and characteristics of the soils at the dam must be available to
design a system as part of initial construction or as remedial work. To be
effective, the well must flow but must not allow the loss of foundation
13-12
... "!\"~·· ··~··· ...... ,
f
I
If
•''
<?-
+
•
r
13-13
..,.,...,,J __ •f_,., ..
llt•nuttol .. 1-1 p.rr- .-.#
OUTLl'T STII:UCTUillll[)
e
.......
, . . _ . , Nl:n•• "lo.cr . .._.._..
PLAN AND PROfiLE
• Af.ZOIIU."'tTR
•
"..,.._,""•"""'•n••'"u..t w IIMI.I.Y
·EO--~ I "' ,._,_~1--
a..I.T..O•
1" •
Cll'fiC.I. 01 TMI. ...........T, ••att»....l
1111.,1.111 C-IIIIIS-
:::.~c, w•aa ••~o
.~.-:-_!.~· ..,.
~--..
;.;::_. ---- ~--'-
EM 1110-2-1901
Figure 13-10. Piezometer installation for dams with outlet structures (adapted from
30 Sep 86
115
U. S. Army Engineer District, Vicksburg )
30 Sep 86
EM 1110-2-1901
S"'ILLWAT
COMPLETED PLAN
>fO I :IOOC:C.t K ....'( t•IO•JOfl
..... ~ I I < 1 • •t I ,. • .,.. H~>•
C"+Cf M Tfll l"tllil&ID(NT, '*'SSISS_.f'l ••v(tl (Qt,H,ooSltOO<
.,.,(oU~..... S · Nllll.t•c•
Su-ftlO ~••OvtO.
Figure 13-11. Piezometer installation for spillways (adapted from U. S. Army Engineer
115
District, Vicksburg )
EM 1110-2-1901
30 Sep 86
SLOPE VARIES
MAX 5 • 0.0100_
OliN. 5 • 0.0020
REGULATING OUTLET
CHANNEL
EL. 1762.:1
PIPELINE A A
,----~~=-=T~-=---~
~--~==~~--~ ~~==~~~~~~_oEL.1772.~
1
I TYPE m
r---••o'---1 - •I
RIQ Ct<AHNEL
~~ .....-..---_-:;_-_-_-_-__
PIPELINE C 1773.~
_)__,.,..._____.
~~~...._R2_~EYoSED _l:~oDOS
---- - - - -~--- -
~
ANO Sci)P£.!_
A~
--y
7111A1t01 : R£VrS[D PLAN PA()fll.£S
~--- -~---- ;.~~---
-
'J. S, £NQ!t-IE~ OISTRICT, PORTI.ANO
- EMBANKMENT DAM
APPLEGATE LAKE
EMBANKMENT
-
~
.fwt!.
DRAINAGE SYSTEM
PLAN AND PROFILE
~~, Dol'~"-:
::::t~~~..H'-.- --------
Tl hA-..OL
4~tl~••&.\\···ot(l0
~~If-. .
-•n >o AGE -1-1/23
13-15
30 Sep 86
EM 1110-2-1901
FURNISH AND INSTALL I" MARINE PLY
_ \ CMP
SELECT BACKFILL
__ ,,,__
APPL£G4.T£ Al'w'ER
} -SELECT APPLEGATE LAKE
f1U-t----'~r'-----"!l_:j BACK FILL ,. EMBANKMENT DAM [W8ANKW[frljf
I 1/.l•.x 1/4••,.••• st- Hit• • _,,
~151/Z •.,_..UA ,..,. DRAINAGE SYSTEM
MINIMUM EXCAVATION LINE ~~! ., •tt:lt . . . . .- •t
DETAILS n
(1/JIIS'flll . ., ,_~-Ill U.l
lttltlltll,..,,..,._,."'''.,
.
~'"" .,.. ., "•"'-'••'
-- ~_s
Figure 13-13. Typical weir installation (from U. S. Army Engineer District, Portland 104 )
EM 1110-2-1901
30 Sep 86
13-17
EM 1110-2-1901
30 Sep 86
observe the situation especially during high reservoir levels. All seepage
outlets should be monitored to determine long-term trends. Sand boils devel-
oping downstream require immediate attention. An estimate of the pore pres-
sure involved can be determined by sandbagging around the boil to measure the
height of water rise. If the seepage exiting at the toe or abutment is severe
and will require remedial work but the dam is not in imminent danger of fail-
ing, collection systems must be designed as a pattern develops. Temporary
control measures such as toe drains and surcharge berms might be installed
with weirs to establish the severity and the trends of the seepage for use in
design of remedial work. Any changes noted during monitoring, i.e., volume
change, sediment load change, etc., should be considered important. After the
remedial work has been designed it may include any number of the monitoring
systems discussed previously in this chapter. If the seepage is exiting from a
drain system that is not being monitored internally, visual monitoring should
continue at specified intervals and during heavy runoffs and high reservoir
levels.
13-19
EM 1110-2-1901
30 Sep 86
13-20
EM 1110-2-1901
30 Sep 86
--r-
Ne+K~----t-----~~~~~~+-----+---~a
·C.~----t-----+-----+-----+-----+---~HCOa
Ma~--~~--~-----+-----+----~----~so.
F•~----t-----+-----+-----+-----+---~co,
12-6
15-l
30 u 20 15 10 5 0 10 25
CATIONS. IN MILUEQUIVAU:NTS ANIONS, IN MILLIEQUIVALENT$
PER UTER PER LITER
13-21
EM 1110-2-1901
30 Sep 86
CHAPTER 14
INSPECTION, MAINTENANCE, AND REHABILITATION OF
SEEPAGE CONTROL MEASURES
14-2. Inspection. The procedure for periodic inspection and continuing evalu-
ation of dams is given in ER 1110-2-100. Procedures for reporting evidence of
distress in dams are given in ER 1110-2-101. Procedures in these two ER's are
often supplemented by Division Regulations as well. Details concerning the
monitoring performance of seepage control measures are given in Chapter 13.
The first general field inspection for new earth and rock-fill dams is carried
out immediately after topping out the embankment. The initial inspection of
concrete dams is accomplished prior to impoundment of reservoir water. The
second inspection for earth and rock-fill and concrete dams is made at a
reasonable stage of normal operating pool but no later than one year after
initial impoundment has begun. Subsequent inspections will be made at one-year
intervals for the next four years, at two-year intervals for the following
four years, and then may be extended to every five years if warranted. The
periodic inspections provide the opportunity for a group of specialists to
critically examine a project for existing and/or potential problems, to recom-
mend remedial action or changes in instrumentation , and to direct the attention
of the operating personnel toward the significant and critical features of a
project. However, the occasional inspection cannot take the place of daily
observations required to detect potentially dangerous problems at an early and
repairable stage. Table 14-1 outlines the inspection, instrumentation, mainte-
nance, and rehabilitation of seepage control facilities. Some seepage control
methods such as embankment zonation, cutoffs, and upstream impervious blankets
are not amenable to visual inspection. Other methods such as flat slopes
downstream of the dam and downstream seepage berms are most accessible and
should be inspected daily during periods of full reservoir pool to ascertain
that they are functioning properly in controlling seepage. Other seepage con-
trol facilities should be inspected on a regular schedule as shown in
table 14-1.
Plat elope• Vithout Dally Wet •pot•. •louch ... Monthly Surf•c:e IIODUMI'I.U Pertilid.nl, .ow- btond dop41 thtckDan
drain• tna, •roaton ina, flll1na and lonath
Yari.. (b) (d)
PIIIOM:ter
Variu(b) (d)
Vertical and horizontal' Weekly Turbidity and dia- PUia.eter
~~train• charge rate
Control of Under•••p•••
(d)
Hortaontal drain Wukly Turbidity and dio- V.eriel(b) PiiiOMter
chat·a• rate
Cutoff
__ (d)
c..poeud baekt111
treneh
Slurry trench
Toe trench dr.ain w.. kly Tnbldity ond dh- Chock partodically<•l
c:hAr&• rau
bltef velle Vat hi (b) Flow nte and sand Ph.aoeeter Chock par1od1cally(f) RohabUitate ... u.
infUtrat1on
Relief wlla Varh1 (b) Plow rate and sand Chock periodtcolly(f) Rehob!l1tate vella
tnfiltrat1on
Var'he (b) Flow rate and 1.and Varies (b) Piezo.e ur Chack. ptriodically(f) Rehabilitate vella en
lnt'iltraticm unwatertna of
structuTe
(a) Schedule to be folloved upon obtainaant of full reservoir pool. The schedule durina initial fillin& will ba deterained durin& th.e initial
periodic inapection.
(b) Obearvat tou should be aad~ one week after ...xiatl• reservoir level and at three aubsequent fall ina raa&rvoir ataaea.
(c)V.eries dttpencUn& upon potential aed:t. . ntation probleM, 1..,act on project perfo~nc:e,. and: t..pact on 1tre. . ayat. . (aae EM 11 t0-2-4000).
(d}Rialna or f'aUtn& bead test abould be conducted on pier.oaeteu annually before the atora.,e aeaaon (Me EK lil0-2-1908, Part l).
(e)C.tcb baelu, unholea, ditcbel, drainage pipe, and weir• ahculd be cle&ned, •• • ainiau. in thtl! fall tn preparation for the vlnter
••••on, and •aa1n in the apring.
(f} Annually. before atoraae aeaaon, c:h&ck valva a, aaak.ata, vall au.rda. cover pl1tea,. flap aataa em the outl•t• and other appurtanaa:c&s (see
EM liiQ-1-1908, Part 1).
(&)Should be axaatned fo-r atra•• cract.. hula••· •hUta of ali&ftiMnt. ••ce•atve lealc.a&•• &net dabrt• cleaned fr,._ &utter• and weirs.
14-2
EM 1110-2-1901
30 Sep 86
14-3
EM 1110-2-1901
30 Sep 86
APPENDIX A
REFERENCES
Government Publications
1. TM 5-818-5.
2. TM 5-545.
3. ER 1110-2-100.
4. ER 1110-2-101.
5. EM 1110-2-1901.
6. EM 1110-2-1902.
7. EM 1110-2-1905.
8. EM 1110-2-1906.
9. EM 1110-2-1908.
10. EM 1110-2-1911.
11. EM 1110-2-1913.
12. EM 1110-2-2300.
13. EM 1110-2-2400.
14. EM 1110-2-2501.
15. EM 1110-2-3501.
16. EM 1110-2-3506.
17. EM 1110-2-4000.
18. Albritton, J., Jackson, L., and Bangert, R., "Foundation Grouting Prac-
tices at Corps of Engineers Dams," TR GL-84-13, 1984, U. S. Army Engineer
Waterways Experiment Station, P. O. Box 631, Vicksburg, MS 39180-0631.
A-1
EM 1110-2-1901
30 Sep 86
22. Bedinger, M. S., "Relationship Between Median Grain Size and Permeability
in the Arkansas River Valley, Arkansas," Professional Paper 424-C, 1961,
U. S. Department of the Interior, Geological Survey, 1717 H Street N.W.,
Washington, DC 20277.
25. Cedergren, H. R., "Flow Net Studies for Several Dams," for U. S. Army
Engineer District, Sacramento, Mar 1975, 650 Capitol Mall, Sacramento, CA
95814.
29. Cooley, R. L., Harsh, J. F., and Lewis, D. C., "Principles of Ground-
Water Hydrology," Hydrologic Engineering Methods for Water Resources
Development, Vol 10, Apr 1972, U. S. Army Engineer Hydrologic Engineering
Center, 609 Second Street, Davis, CA 95616.
31. Cooper, S. S., Koester, J. P., and Franklin, A. G., "Geophysical Investi-
gation at Gathwright Dam," Miscellaneous Paper GL-82-2, Mar 1982, U. S.
Army Engineer Waterways Experiment Station, P. O. Box 631, Vicksburg, MS
39180-0631.
A-2
EM 1110-2-1901
30 Sep 86
35. Duncan, J. M., "Seepage Analysis for Columbia Lock and Dam," Miscellane-
ous Paper No. 3-503, Jun 1962, U. S. Army Engineer Waterways Experiment
Station, P. O. Box 631, Vicksburg, MS 39180-0631.
37. Edris, E. V., Jr., and Vanadit-Ellis, W., "Geotechnical Computer Program
Survey," Miscellaneous Paper GL-82-1, Mar 1982, U. S. Army Engineer
Waterways Experiment Station, P. O. Box 631, Vicksburg, MS 39180-0631.
39. Franke, P., "The Concrete-Pile Cutoff Wall According to the I.C.O.S. -
Vedner Patent," Translation No. 54-6, Sep 1954, U. S. Army Engineer
Waterways Experiment Station, P. O. Box 631, Vicksburg, MS 39180-0631.
40. Gerwick, B. C., Jr., Holland, T. C., and Komendant, G. J., "Tremie Con-
crete for Bridge Piers and Other Massive Underwater Placements," Report
No. FHWA-RD-81-153, Sep 1981, U.S. Department of Transportation, Federal
Highway Administration, Washington, D. C. Available from: Technical
Information Center, U. S. Army Engineer Waterways Experiment Station,
P. O. Box 631, Vicksburg, MS 39180-0631.
41. Guertin, J. D., and McTigue, W. H., "Groundwater Control Systems for
Urban Tunneling," Groundwater Control in Tunneling, Vol 1, Report
No. FHWA-RD-81-073, Apr 1982, U. S. Department of Transportation, Federal
Highway Administration, Washington, D. C. Available from: Technical
Information Center, U. S. Army Engineer Waterways Experiment Station,
P. O. Box 631, Vicksburg, MS 39180-0631.
A-3
EM 1110-2-1901
30 Sep 86
43. Hallford, C. R., "Fluid Trench Construction, Concrete Diaphragm Wall Con-
struction," Lecture for Construction of Earth and Rockfill Dams Course,
Mar 1983, U. S. Army Engineer Waterways Experiment Station, P. O.
Box 631, Vicksburg, MS 39180-0631.
47. Holland, T. C., and Turner, J. R., "Construction of Tremie Concrete Cut-
off Wall, Wolf Creek Dam, Kentucky," Miscellaneous Paper SL-80-10, Sep
1980, U. S. Army Engineer Waterways Experiment Station, P. O. Box 631,
Vicksburg, MS 39180-0631.
48. Jansen, R. B., Dams and Public Safety, 1980, U. S. Department of the
Interior, Bureau of Reclamation, Denver Federal Center, Denver, CO
80225.
51. Leach, R. E., "Waterbury Dam Seepage Study," draft report prepared for
U. S. Army Engineer District, New York, 1982, U. S. Army Engineer Water-
ways Experiment Station, P. O. Box 631, Vicksburg, MS 39180-0631.
A-4
EM 1110-2-1901
30 Sep 86
53. Mansur, C. I., and Perret, W. R., "Efficacy of Partial Cutoffs for Con-
trolling Underseepage Beneath Levees," Technical Memorandum No. 3-267,
Jan 1949, U. S. Army Engineer Waterways Experiment Station, P. O.
Box 631, Vicksburg, MS 39180-0631.
54. Mantei, C. L., and Cobb, J. E., "Seepage Analysis, Progress Report -
Stage 2, Specifications, Calamus Dam, Nebraska," Technical Memorandum
No. V-222-B-1-F, 1981, Bureau of Reclamation, P. O. Box 25007, Denver, CO
80225.
55. Mantei, C. L., Esmiol, E. E., and Cobb, J. E., "Seepage Analysis, Calamus
Dam - Stage I, Pick-Sloan Missouri River Basin Project, Nebraska,"
Technical Memorandum No. V-224-A, Mar 1980, Bureau of Reclamation, P. O.
Box 25007, Denver, CO 80225.
59. Moser, J. H., and Huibregtse, K. R., "Handbook for Sampling and Sample
Preservation of Water and Wastewater," EPA-600-4-76-049, Sep 1976,
Environmental Protection Agency, 26 W. St. Clair Street, Cincinnati, OH
45268.
61. Nobari, E. S., Lee, K. L., and Duncan, J. M., "Hydraulic Fracturing in
Zoned Earth and Rockfill Dams," Contract Report CR S-73-2, Jan 1973,
U. S. Army Engineer Waterways Experiment Station, P. O. Box 631,
Vicksburg, MS 39180-0631.
62. Pace, M. E., et al., "Seepage Analysis of Confined Flow Problems by the
Method of Fragments," Instruction Report K-84-8, Sep 1984, U. S. Army
A-5
EM 1110-2-1901
30 Sep 86
63. Perry, E. B., "Laboratory Tests on Granular Filters for Embankment Dams,"
Draft Technical Report, 1986, U. S. Army Engineer Waterways Experiment
Station, P. O. Box 631, Vicksburg, MS 39180-0631.
67. Sherard, J. L., "Filters for Glacial Tills and Other Coarse Impervious
Soils," Memorandum IG, May 1984, Soil Conservation Service Soil Mechanics
Laboratory, Lincoln, Nebraska. Available from: Technical Information
Center, U. S. Army Engineer Waterways Experiment Station, P. O. Box 631,
Vicksburg, MS 39180-0631.
72. Talbot, J. R., and Nelson, R. E., "The Mechanics of Seepage Analysis,"
Soil Mechanics Note No. 7, Oct 1979, U. S. Department of Agriculture,
Soil Conservation Service, P. O. Box 2890, Washington, DC 20013.
73. Tracy, F. T., "An Interactive Graphics Finite Element Method Grid Genera-
tor for Two-Dimensional Problems," Miscellaneous Paper K-77-5, Aug 1977a,
A-6
EM 1110-2-1901
30 Sep 86
74. Tracy, F. T., "An Interactive Graphics Postprocessor for Finite Element
Method Results," Miscellaneous Paper K-77-4, Aug 1977b, U. S. Army Engi-
neer Waterways Experiment Station, P. O. Box 631, Vicksburg, MS
39180-0631.
75. Tracy, F. T., "A Plane and Axisymmetric Finite Element Program for
Steady-State and Transient Seepage Problems," Miscellaneous Paper K-73-4,
May 1973a, U. S. Army Engineer Waterways Experiment Station, P. O.
Box 631, Vicksburg, MS 39180-0631.
76. Tracy, F. T., "A Three-Dimensional Finite Element Program for Steady-State
and Transient Seepage Problems," Miscellaneous Paper K-73-3, May 1973b,
U. S. Army Engineer Waterways Experiment Station, P. O. Box 631,
Vicksburg, MS 39180-0631.
80. U. S. Army Engineer District, Ft. Worth, "Brazos River Basin, Aquilla
Creek, Texas, Aquilla Lake, Embankment, Spillway, and Outlet Works,"
Design Memorandum No. 7, May 1976a, P. O. Box 17300, Ft. Worth, TX 76102.
81. U. S. Army Engineer District, Ft. Worth, "Trinity River Basin, Elm Fork,
Trinity River, Aubrey Lake, Texas, Embankment and Spillway (Revised),"
Design Memorandum No. 5, Jun 1976b, P. O. Box 17300, Ft. Worth, TX 76102.
82. U. S. Army Engineer District, Ft. Worth, "Trinity River Basin, Mountain
Creek, Lakeview Lake, Texas, Embankment and Spillway," Design Memorandum
No. 9, Apr 1980, P. O. Box 17300, Ft. Worth, TX 76102.
A-7
EM 1110-2-1901
30 Sep 86
90. U. S. Army Engineer District, Kansas City, "Hillsdale Lake, Osage River
Basin, Big Bull Creek, Kansas," Plans for Hillsdale Dam, Apr 1978,
601 E. 12th Street, Kansas, City, MO 64106.
91. U. S. Army Engineer District, Kansas City, "Long Branch Lake, Chariton
River Basin, East Fork Little Chariton River, Missouri," Jul 1974,
601 East 12th Street, Kansas City, MO 64106.
93. U. S. Army Engineer District, Los Angeles, "Painted Rock Reservoir, Gila
River, Arizona, Seepage Control Measures," Oct 1981, P. O. Box 2711, Los
Angeles, CA 90053.
A-8
EM 1110-2-1901
30 Sep 86
105. U. S. Army Engineer District, Rock Island, "Local Flood Protection Proj-
ect, Rockford - Stage I, Kent Creek, Winnebago County, Illinois,"
Specifications for Page Park Dam, Jun 1978a, Clock Tower Building, Rock
Island, IL 61201.
106. U. S. Army Engineer District, Rock Island, "Slurry Trench Cutoff, Dam-
Foundation Report, Saylorville Lake, Des Moines River, Iowa," Part VI,
Binder 3 of 5, Jan 1978b, Clock Tower Building, Rock Island, IL 61201.
A-9
EM 1110-2-1901
30 Sep 86
111. U. S. Army Engineer District, St. Louis, "Clarence Cannon Dam and
Reservoir, Upper Mississippi River Basin, Salt River, Missouri, Embank-
ment Design, Main Dam," Design Memorandum No. 12, Nov 1969, 210 Tucker
Boulevard, St. Louis, MO 63101.
112. U. S. Army Engineer District, St. Louis, "John H. Overton Lock and Dam,
Red River Waterway," Design Memorandum No. 17, Detail Design, Vol 2,
Appendices, Design Computations, Nov 1978, St. Louis, MO. (prepared for
U. S. Army Engineer District, New Orleans).
113. U. S. Army Engineer District, Tulsa, "Big Hill Lake, Big Hill Creek,
Kansas, Embankment, Outlet Works, and Spillway," Design Memorandum
No. 6, Oct 1975, P. O. Box 61, Tulsa, OK 74121-0061.
114. U. S. Army Engineer District, Tulsa, "El Dorado Lake, Walnut River,
Kansas, Embankment and Spillway," Design Memorandum No. 6, Jul 1974,
P. O. Box 61, Tulsa, OK 74121-0061.
A-10
EM 1110-2-1901
30 Sep 86
Non-Government Publications
132. Anonymous, "Fast Value Engineering Saves Over Budget Dam," Engineering
News Record, Vol 201, No. 19, 9 Nov 1978, pp 24-25. Available from:
Technical Information Center, U. S. Army Engineer Waterways Experiment
Station, P. O. Box 631, Vicksburg, MS 39180-0631.
A-11
EM 1110-2-1901
30 Sep 86
133. Anonymous, "Ground Water: Our Priceless Resource," Ground Water Age,
Vol 14, No. 8, Apr 1980, pp 33-36, 64-66. Available from: Technical
Information Center, U. S. Army Engineer Waterways Experiment Station,
P. O. Box 631, Vicksburg, MS 39180-0631.
134. Anonymous, "World's Deepest Cutoff Wall Reaches 430 Ft," Engineering
News Record, Vol 188, No. 1, 6 Jan 1972, pp 26-28. Available from:
Technical Information Center, U. S. Army Engineer Waterways Experiment
Station, P. O. Box 631, Vicksburg, MS 39180-0631.
135. Anton, W. F., and Dayton, D. J., "Comanche Dike-2 Slurry Trench Seepage
Cutoff," Proceedings-of the Conference on Performance of Earth and
Earth-Supported Structures, Purdue University, Vol 1, Part 1, 1972,
pp 735-749. Available from: Technical Information Center, U. S. Army
Engineer Waterways Experiment Station, P. O. Box 631, Vicksburg, MS
39180-0631.
136. Arulanandan, K., and Perry, E. B., "Erosion in Relation to Filter Design
Criteria for Earth Dams," Journal of the Geotechnical Engineering
Division, American Society of Civil Engineers, Vol 109, No. 5, May 1983,
pp 682-698. Available from: Technical Information Center, U. S. Army
Engineer Waterways Experiment Station, P. O. Box 631, Vicksburg, MS
39180-0631.
138. Barron, R. A., "The Design of Earth Dams," Handbook of Dam Engineering,
A. R. Golze, ed., Van Nostrand, New York, 1977, pp 291-318. Available
from: Technical-Information Center, U. S. Army Engineer Waterways
Experiment Station, P. O. Box 631, Vicksburg, MS 39180-0631.
139. Basak, P., and Madhav, M. R., "Upper and Lower Limit of Darcy's Law,"
Indian Geotechnical Journal, Vol 9, No. 2, Apr 1979, pp 134-153.
Available from: Technical Information Center, U. S. Army Engineer
Waterways Experiment Station, P. O. Box 631, Vicksburg, MS 39180-0631.
140. Bear, J., Dynamics of Fluids in Porous Media, Elsevier, New York, 1972.
Available from: Technical Information Center, U. S. Army Engineer
Waterways Experiment Station, P. O. Box 631, Vicksburg, MS 39180-0631.
141. Blind, H., "Inspection Galleries in Earth and Rockfill Dams," Water
Power and Dam Construction, Vol 34, No. 4, Apr 1982, pp 25-31. Avail-
able from: Technical Information Center, U. S. Army Engineer Waterways
Experiment Station, P. O. Box 631, Vicksburg, MS 39180-0631.
142. Bloom, E., Dynes, S., and Glossett, C., "Soil Bentonite Slurry Trench in
a PL-566 Rolled Earth Dam in Indiana," Paper No. 79-2579 presented at
American Society of Agricultural Engineers Meeting, New Orleans, La.,
A-12
EM 1110-2-1901
30 Sep 86
143. Boer, S. A. De, and Molen, W. H. Van Der, "Electrical Models: Conduc-
tive Sheet Analogies," Drainage Principles and Applications, Interna-
tional Institute for Land Reclamation and Improvement, Wageningen,
Netherlands, Publication 16, Vol 1, 1972, pp 201-221. Available from:
Technical Information Center, U. S. Army Engineer Waterways Experiment
Station, P. O. Box 631, Vicksburg, MS 39180-0631.
145. Brand, E. W., and Armstrong, R., "Sand Model Studies of Seepage Through
Earth Dams," Proceedings, Symposium on Earth and Rockfill Dams, Indiana,
Vol 1, 1968, pp 188-196. Available from: Technical Information Center,
U. S. Army Engineer Waterways Experiment Station, P. O. Box 631,
Vicksburg, MS 39180-0631.
146. Bros, B., and Orzeszyna, H., "The Influence of Particle Angularity and
Surface Roughness on Engineering Properties of Sand," Design Parameters
in Geotechnical Engineering, British Geotechnical Society, London, 1979,
pp 11-14. Available from: Technical Information Center, U. S. Army
Engineer Waterways Experiment Station, P. O. Box 631, Vicksburg, MS
39180-0631.
148. Burmister, D. M., "The Importance and Practical Use of Relative Density
in Soil Mechanics," Proceedings, American Society for Testing and
Materials, Vol 48, 1948, pp 1249-1268. Available from: Technical
Information Center, U. S. Army Engineer Waterways Experiment Station,
P. O. Box 631, Vicksburg, MS 39180-0631.
149. Cary, A. S., Walter, B. H., and Horstad, H. T., "Permeability of Mud
Mountain Dam Core Material," Transactions, American Society of Civil
Engineers, Vol 108, 1943, pp 719-737. Available from: Technical
Information Center, U. S. Army Engineer Waterways Experiment Station,
P. O. Box 631, Vicksburg, MS 39180-0631.
A-13
EM 1110-2-1901
30 Sep 86
151. Casagrande, A., "Seepage Through Dams," New England Waterworks Associa-
tion, Vol LI, No. 2, Jun 1937, Dedham, Mass. Available from: Technical
Information Center, U. S. Army Engineer Waterways Experiment Station,
P. O. Box 631, Vicksburg, MS 39180-0631.
152. Case International Company, "Case Slurry Wall Notebook," 1982, Roselle,
Ill. Available from: Case International Company, P. O. Box 40,
Roselle, IL 60172.
155. Cedergren, H. R., Seepage, Drainage and Flow Nets, 2nd ed., Wiley, New
York, 1977. Available from: Technical Information Center, U. S. Army
Engineer Waterways Experiment Station, P. O. Box 631, Vicksburg,
MS 39180-0631.
156. Christian, J. T., "Flow Nets by the Finite Element Method," Ground
Water, Vol 18, No. 2, Mar-Apr 1980a, pp 178-181. Available from:
Technical Information Center, U. S. Army Engineer Waterways Experiment
Station, P. O. Box 631, Vicksburg, MS 39180-0631.
157. Christian, J. T., "Flow Nets from Finite Element Data," International
Journal for Numerical and Analytical Methods in Geomechanics, Vol 4,
Apr-Jun 1980b, pp 191-196. Available from: Technical Information
Center, U. S. Army Engineer Waterways Experiment Station, P. O. Box 631,
Vicksburg, MS 39180-0631.
159. Coffman, J. A., Jr., and Franks, L. W., "Rehabilitating the Muskingum
River System," Transactions of the Fourteenth International Congress on
Dams, Q52 R.50, Rio de Janeiro, Brazil, 1982, pp 827-845. Available
from: Technical Information Center, U. S. Army Engineer Waterways
Experiment Station, P. O. Box 631, Vicksburg, MS 39180-0631.
160. Couch, F. B., and Ressi deCervia, A. L., "Seepage Cutoff Wall Installed
Through Dam is Construction First," Civil Engineering, American Society
of Civil Engineers, Vol 49, No. 1, Jan 1979, pp 62-66. Available from:
A-14
EM 1110-2-1901
30 Sep 86
161. Coxon, R. E., and Crook, D. E., "Some Simple Approaches to Leakage
Detection in Dams," Transactions of the Twelfth International Congress
on Large Dams, Vol II, 1976, Mexico City, pp 527-540. Available from:
Technical Information Center, U. S. Army Engineer Waterways Experiment
Station, P. O. Box 631, Vicksburg, MS 39180-0631.
164. Davis, S. N., and Dewiest, R. J. M., Hydrogeology, Wiley, New York,
1966. Available from: Technical Information Center, U. S. Army Engi-
neer Waterways Experiment Station, P. O. Box 631, Vicksburg,
MS 39180-0631.
165. Denny, K. J., Grain Size Distribution and Its Effect on the Permeability
of Unconsolidated Sands, M.S. Thesis, 1965, University of Texas, Austin,
Tex. Available from: Technical Information Center, U. S. Army Engineer
Waterways Experiment Station, P. O. Box 631, Vicksburg, MS 39180-0631.
167. Desai, C. S., and Abel, J. F., Introduction to the Finite Element
Method, Van Nostrand Reinhold, New York, 1972. Available from:
Technical Information Center, U. S. Army Engineer Waterways Experiment
Station, P. O. Box 631, Vicksburg, MS 39180-0631.
168. Duguid, D. R., et al., "The Slurry Trench Cut-Off for the Duncan Dam,"
Canadian Geotechnical Journal, Vol 8, No. 1, Feb 1971, pp 94-108.
Available from: Technical Information Center, U. S. Army Engineer
Waterways Experiment Station, P. O. Box 631, Vicksburg, MS 39180-0631.
A-15
EM 1110-2-1901
30 Sep 86
173. Fetzer, C. A., "Wolf Creek Dam - Remedial Work, Engineering Concepts
Actions, and Results," Transactions of the Thirteenth International
Congress on Large Dams, Vol 2, 1979, New Delhi, India, pp 57-82.
Available from: Technical Information Center, U. S. Army Engineer
Waterways Experiment Station, P. O. Box 631, Vicksburg, MS 39180-0631.
175. Freeze, R. A., and Cherry, J. A., Groundwater, Prentice-Hall, New York,
1979. Available from: Technical Information Center, U. S. Army
Engineer Waterways Experiment Station, P. O. Box 631, Vicksburg, MS
39180-0631.
176. Fuquay, G. A,, "Foundation Cutoff Wall for Allegheny Reservoir Dam,"
Closure to discussion, Journal of the Soil Mechanics and Foundations
Division, pp 1363-1366, American Society of Civil Engineers, New York,
Vol 94, No. SM6, Nov 1968. Available from: Technical Information
Center, U. S. Army Engineer Waterways Experiment Station, P. O. Box 631,
Vicksburg, MS 39180-0631.
178. Greer, D. M., Moorhouse, D. C., and Millet, R. A., "Sheet Pile Interlock
Seepage Study," Woodward-Moorhouse & Associates, Inc., New York, 1969.
Available from: Woodward-Clyde & Associates, Inc., Two Pennsylvania
Plaza, New York, NY 10001.
A-16
EM 1110-2-1901
30 Sep 86
180. Harr, M. E., Groundwater and Seepage, McGraw-Hill Book Company, New
York, 1962. Available from: Technical Information Center, U. S. Army
Engineer Waterways Experiment Station, P. O. Box 631, Vicksburg, MS
39180-0631.
182. ICOS, "The ICOS Company in Underground Works," Vol 3, 1968, Milano,
Italy. Available from: ICOS Corporation of America, Four West 58th
Street, New York, NY 10019.
188. Jenkins, J. D., and Bankofier, D. E., "Hills Creek Dam Seepage Correc-
tion," Performance of Earth and Earth-Supported Structures, American
Soceity of Civil Engineers, New York, 1972, pp 723-733. Available from:
Technical Information Center, U. S. Army Engineer Waterways Experiment
Station, P. O. Box 631, Vicksburg, MS 39180-0631.
A-17
EM 1110-2-1901
30 Sep 86
189. Johnson Division, Universal Oil Products Co., Ground Water and Wells,
St. Paul, Minn., 1972. Available from: Technical Information Center, U.
S. Army Engineer Waterways Experiment Station, P. O. Box 631, Vicksburg,
MS 39180-0631.
191. Jones, L. B., "Reservoir Seepage and Ground Water Control, McNary
Reservoirs," paper presented at the American Society of Civil Engineer
Convention, Phoenix, Ariz., 1961. Available from: Shannon & Wilson,
Inc., 1105 North 38th Street, Seattle, WA 98103.
193. Klohn, E. J., "Seepage Control for Tailings Dams," Mine Drainage,
G. O. Argall, Jr., ed., Miller Freeman Publications, San Francisco, 1979,
pp 671-725. Available from: Technical Information Center, U. S. Army
Engineer Waterways Experiment Station, P. O. Box 631, Vicksburg, MS
39180-0631.
194. Knabach, M. L., and Dingle, S. O., "Slurry Trench Cutoff in Dam
Construction," Paper No. 74-2014 presented at American Society of
Agricultural Engineers Meeting, Stillwater, Okla., 1974. Available
from: American Society of Agricultural Engineers, P. O. Box 229, St.
Joseph, MI 49085.
195. Koerner, R. M., Lord, A. E., Jr., and McCabe, W. M., "Acoustic Emission
(Microseismic) Monitoring of Earth Dams," The Evaluation of Dam Safety,
American Society of Civil Engineers, New York, 1977, pp 274-291.
Available from: Technical Information Center, U. S. Army Engineer
Waterways Experiment Station, P. O. Box 631, Vicksburg, MS 39180-0631.
196. Kramer, H., "Deep Cutoff Trench of Puddled Clay for Earth Dam and Levee
Protection," Engineering News Record, Vol 136, No. 26, 27 Jun 1946,
pp 76-80. Available from: Technical Information Center, U. S. Army
Engineer Waterways Experiment Station, P. O. Box 631, Vicksburg, MS
39180-0631.
A-18
EM 1110-2-1901
30 Sep 86
199. Kruseman, G. P., and DeRidder, N. A., "Analysis and Evaluation of Pump-
ing Test Data," Bulletin 11, 1970, International Institute for Land
Reclamation and Improvement, Wageningen, Netherlands. Available from:
Technical Information Center, U. S. Army Engineer Waterways Experiment
Station, P. O. Box 631, Vicksburg, MS 39180-0631.
200. Lambe, T. W., Soil Testing for Engineers, Wiley, New York, 1951. Avail-
able from: Technical Information Center, U. S. Army Engineer Waterways
Experiment Station, P. O. Box 631, Vicksburg, MS 39180-0631.
201. Lambe, T. W., and Whitman, R. V., Soil Mechanics, Wiley, New York, 1969.
Available from: Technical Information Center, U. S. Army Engineer
Waterways Experiment Station, P. O. Box 631, Vicksburg, MS 39180-0631.
202. Lane, K. S., and Wohlt, P. E., "Performance of Sheet Piling and Blankets
for Sealing Missouri River Reservoirs," Transactions of the Seventh
International Congress on Large Dams, Vol 4, 1961, pp 25, Rome, Italy.
Available from: Technical Information Center, U. S. Army Engineer
Waterways Experiment Station, P. O. Box 631, Vicksburg, MS 39180-0631.
203. La Russo, R. S., "Wanapum Development - Slurry Trench and Grouted Cut-
Off," Grouts and Drilling Muds in Engineering Practice, 1963,
pp 196-201, Butterworths, London. Available from: Technical Informa-
tion Center, U. S. Army Engineer Waterways Experiment Station, P. O.
Box 631, Vicksburg, MS 39180-0631.
207. Lewis, D. C., Kriz, G. J., and Burgy, R. H., "Tracer Dilution Sampling
Techniques to Determine Hydraulic Conductivity of Fractured Rock," Water
Resource Research, Vol 2, No. 3, 1966. Available from: Technical
A-19
EM 1110-2-1901
30 Sep 86
209. Logani, K. L., and Lhez, M. H. H., "Dispersive Soils Used in the
Construction of the Ullum Dam in Argentina," Proceedings of the Sixth
Panamerican Conference on Soil Mechanics and Foundation Engineering,
Vol 111, Dec 1979, Lima, Peru, pp 394-411. Available from: Technical
Information Center, U. S. Army Engineer Waterways Experiment Station,
P. O. Box 631, Vicksburg, MS 39180-0631.
210. London, A. G., "The Computation of Permeability from Simple Soil Tests,"
Geotechnique, Vol 3, No. 4, 1952, pp 165-183. Available from: Techni-
cal Information Center, U. S. Army Engineer Waterways Experiment
Station, P. O. Box 631, Vicksburg, MS 39180-0631.
211. Louis, c., "A Study of Groundwater Flow in Jointed Rock and Its Influence
on the Stability of Rock Masses," Rock Mechanics Research Report No. 10,
1969, Imperial College, London, England. Available from: Technical
Information Center, U. S. Army Engineer Waterways Experiment Station,
P. O. Box 631, Vicksburg, MS 39180-0631.
213. Lowe, J., III. "Foundation Design - Tarbela Dam," The Fourth Nabor
Carrillo Lecture, Mexican Society for Soil Mechanics, Mexico, 1978.
Available from: Technical Information Center, U. S. Army Engineer
Waterways Experiment Station, P. O. Box 631, Vicksburg, MS 39180-0631.
214. Lynch, E. J., Formation Evaluation, Harper and Row, New York, 1962.
Available from: Technical Information Center, U. S. Army Engineer
Waterways Experiment Station, P. O. Box 631, Vicksburg, MS 39180-0631.
A-20
EM 1110-2-1901
30 Sep 86
217. Mantei, C. L., and Harris, D. W., "Finite Element Seepage Analysis on
Reclamation Dams," Preprint 3691, paper presented at American Society of
Civil Engineers Convention and Exposition, Atlanta, Ga., Oct 1979.
Available from:. Technical Information Center, U. S. Army Engineer
Waterways Experiment Station, P. O. Box 631, Vicksburg, MS 39180-0631.
219. Masch, F. D., and Denny, K. J., "Grain Size Distribution and Its Effect
on the Permeability of Unconsolidated Sands," Water Resources Research,
Vol 2, No. 4, 1966, pp 665-677. Available from: Technical Information
Center, U. S. Army Engineer Waterways Experiment Station, P. O. Box 631,
Vicksburg, MS 39180-0631.
220. McDaniel, T. N., and Decker, R. S., "Dispersive Soil Problem at Los
Esteros Dam," Journal of the Geotechnical Engineering Division, Vol 105,
No. GT9, Sep 1979, pp 1017-1030, American Society of Civil Engineers,
New York. Available from: Technical Information Center, U. S. Army
Engineer Waterways Experiment Station, P. O. Box 631, Vicksburg,
MS 39180-0631.
222. Miller, E. A., and Salzman, G. S., "Value Engineering Saves Dam Proj-
ect," Civil Engineering, American Society of Civil Engineers, Vol 50,
No. 8, Aug 1980, pp 51-55. Available from: Technical Information
Center, U. S. Army Engineer Waterways Experiment Station, P. O. Box 631,
Vicksburg, MS 39180-0631.
223. Millet, R. A., and Perez, J.-Y., "Current USA Practice: Slurry Wall
Specifications," Journal of the Geotechnical Engineering Division,
American Society of Civil Engineers, Vol 107, No. GT8, Aug 1980,
pp 1041-1056. Available from: Technical Information Center, U. S. Army
Engineer Waterways Experiment Station, P. O. Box 631, Vicksburg, MS
39180-0631.
A-21
EM 1110-2-1901
30 Sep 86
225. Missbach, A., "Listy Cukrovar," Vol 55, 1937, p 293, Prague,
Czechoslovakia. Available from: Technical Information Center, U. S.
Army Engineer Waterways Experiment Station, P. O. Box 631, Vicksburg, MS
39180-0631.
226. Mitchell, J. K., Fundamentals of Soil Behavior, Wiley, New York, 1976.
Available from: Technical Information Center, U. S. Army Engineer
Waterways Experiment Station, P. O. Box 631, Vicksburg, MS 39180-0631.
227. Mitchell, J. K., Guzikowski, F., and Villet, W. C. B., "The Measurement
of Soil Properties In-Situ," Report No. LBL-6363, Mar 1978, Department
of Civil Engineering, University of California, Berkeley, Calif. Avail-
able from: Technical Information Center, U. S. Army Engineer Waterways
Experiment Station, P. O. Box 631, Vicksburg, MS 39180-0631.
229. Muskat, M., The Flow of Homogeneous Fluids Through Porous Media,
J. W. Edwards, Ann Arbor, Mich., 1946. Available from: Technical Infor-
mation Center, U. S. Army Engineer Waterways Experiment Station, P. O.
Box 631, Vicksburg, MS 39180-0631.
230. Nash, J. K. T. L., "Slurry Trench Walls, Pile Walls, Trench Bracing,"
Proceedings, Sixth European Conference on Soil Mechanics and Foundation
Engineering, Vienna, Austria, Vol 2.1, 1976, pp 27-32. Available from:
Technical Information Center, U. S. Army Engineer Waterways Experiment
Station, P. O. Box 631, Vicksburg, MS 39180-0631.
231. Nash, J. K. T. L., and Jones, G. K., "The Support of Trenches Using
Fluid Mud," Grouts and Drilling Muds in Engineering Practice, 1963,
pp 177-180, Butterworths, London. Available from: Technical Informa-
tion Center, U. S. Army Engineer Waterways Experiment Station, P. O.
Box 631, Vicksburg, MS 39180-0631.
233. Olson, R. E., and Daniel, D. E., "Field and Laboratory Measurement of
the Permeability of Saturated and Partially Saturated Fine-Grained
Soils," paper presented at the American Society for Testing and
Materials Symposium on Permeability and Groundwater Contaminant
Transport, Philadelphia, Pa., Jun 1979. Available from: Technical
A-22
EM 1110-2-1901
30 Sep 86
234. Parker, D. G., and Thornton, S. I., "Permeability of Fly Ash and Fly Ash
Stabilized Soils," Dec 1976, Department of Civil Engineering, University
of Arkansas, Fayetteville, Ark. Available from: Technical Information
Center, U. S. Army Engineer Waterways Experiment Station, P. O. Box 631,
Vicksburg, MS 39180-0631.
235. Pavlovsky, N. M., Collected Works, 1956, Akad. Nauk USSR, Leningrad
(cited in Harr 1977). Available from: Technical Information Center,
U. S. Army Engineer Waterways Experiment Station, P. O. Box 631,
Vicksburg, MS 39180-0631.
237. Perloff, W. H., and Baron, W., Soil Mechanics, Ronald Press, New York,
1976. Available from: Technical Information Center, U. S. Army Engi-
neer Waterways Experiment Station, P. O. Box 631, Vicksburg,
MS 39180-0631.
238. Pettyjohn, W. A,., et al., "A Ground-Water Quality Atlas of the United
States," May 1979, National Demonstration Water Project, Washington,
D. C. Available from: Technical Information Center, U. S. Army Engi-
neer Waterways Experiment Station, P. O. Box 631, Vicksburg,
MS 39180-0631.
A-23
EM 1110-2-1901
30 Sep 86
245. Rouse, H., and Ince, S., History of Hydraulics, State University of
Iowa, Iowa City, 1957. Available from: Technical Information Center,
U. S. Army Engineer Waterways Experiment Station, P. O. Box 631,
Vicksburg, MS 39180-0631.
247. Rushton, K. R., and Redshaw, S. C., Seepage and Groundwater Flow, Wiley,
New York, 1979. Available from: Technical Information Center, U. S.
Army Engineer Waterways Experiment Station, P. O. Box 631, Vicksburg,
MS 39180-0631.
248. Ryan, C. R., "Slurry Cut-Off Walls, Design and Construction," paper
presented at Slurry Wall Construction, Design, Techniques, and
Procedures Course, Chicago, Ill., Apr 1976. Available from: GO-CON,
Inc., P. O. Box 17380, Pittsburgh, PA 15235.
249. Ryan, C. R., "Slurry Cut-Off Walls, Design Parameters and Final
Properties," paper presented at Slurry Wall Construction, Design,
Techniques, and Procedures Course, Miami, Fla., Feb 1977. Available
from: GO-CON, Inc., P. O. Box 17380, Pittsburgh, PA 15235.
A-24
EM 1110-2-1901
30 Sep 86
252. Sherard, J. L., A Study of the Influence of the Earthquake Hazard on the
Design of Embankment Dams, Jul 1966, Woodward, Clyde, Sherard & Asso-
ciates, Oakland, Calif. Available from: Technical Information Center,
U. S. Army Engineer Waterways Experiment Station, P. O. Box 631,
Vicksburg, MS 39180-0631.
256. Sherard, J. L., et al., Earth and Earth-Rock Dams, Wiley, New York,
1963. Available from: Technical Information Center, U. S. Army Engi-
neer Waterways Experiment Station, P. O. Box 631, Vicksburg,
MS 39180-0631.
257. Singh, B., and Sharma, H. D., Earth and Rockfill Dams, Sarita Prakashan,
India, 1976. Available from: Technical Information Center, U. S. Army
Engineer Waterways Experiment Station, P. O. Box 631, Vicksburg,
MS 39180-0631.
259. Soletanche, "Laguna Dam, Mexico, 1970," Paris, France. Available from:
RECOSOL, Rosslyn Center, 1700 North Moore Street, Suite 2200, Arlington,
VA 22209.
260. Soletanche, "Los Reyes Dam, Mexico, 1972," Paris, France. Available
from: RECOSOL, Rosslyn Center, 1700 North Moore Street, Suite 2200,
Arlington, VA 22209.
A-25
EM 1110-2-1901
30 Sep 86
262. Soletanche, "Tenango Dam, Mexico, 1971," Paris, France. Available from:
RECOSOL, Rosslyn Center, 1700 North Moore Street, Suite 2200, Arlington,
VA 22209.
263. Soletanche, "The Razaza Dam, Iraq, 1969," Paris, France. Available from:
RECOSOL, Rosslyn Center, 1700 North Moore Street, Suite 2200, Arlington,
VA 22209.
264. Sowers, G. F., Earth and Rockfill Dam Engineering, Asia Publishing
House, New York, 1962. Available from: Technical Information Center,
U. S. Army Engineer Waterways Experiment Station, P. O. Box 631,
Vicksburg, MS 39180-0631.
265. Sowers, G. F., "Earth Dam Failures," Lectures of the Seminar, Failures
of Large Dams, Reasons and Remedial Measures, W. Wittke, ed., Publica-
tion No. 4, 1977, Institute for Foundation Engineering, Soil Mechanics,
Rock Mechanics, and Waterways Construction, Aachen, Germany, pp 178-226.
Available from: Technical Information Center, U. S. Army Engineer
Waterways Experiment Station, P. O. Box 631, Vicksburg, MS 39180-0631.
267. Strohm, W. E., Nettles, E. H., and Calhoun, C. C., "Study of Drainage
Characteristics of Base Course Materials," Symposium on Subsurface
Drainage, Highway Research Record 203, 1967, pp 8-28, Highway Research
Board. Washington. D. C. Available from: Technical Information Center,
U. S. Army Engineer Waterways Experiment Station, P. O. Box 631,
Vicksburg, MS 39180-0631.
268. Taylor, D. W., Fundamentals of Soil Mechanics, Wiley, New York, 1948.
Available from: Technical Information Center, U. S. Army Engineer Water-
ways Experiment Station, P. O. Box 631, Vicksburg, MS 39180-0631.
269. Taylor, H., and Chow, Y. M., "Design Monitoring, and Maintaining
Drainage System of a High Earthfill Dam," Transactions, Twelfth
International Congress on Large Dams, Mexico, Vol 2, 1976, pp 147-167.
Available from: Technical Information Center, U. S. Army Engineer
Waterways Experiment Station, P. O. Box 631, Vicksburg, MS 39180-0631.
271. Telling, R. M., Menzies, B. K., and Coulthard, J. M., "A Design Method
for Assessing the Effectiveness of Partially Penetrating Cut-Off Walls,"
Ground Engineering, Vol 11, No. 8, Nov 1978, pp 48-51. Available from:
Technical Information Center, U. S. Army Engineer Waterways Experiment
Station, P. O. Box 631, Vicksburg, MS 39180-0631.
A-26
EM 1110-2-1901
30 Sep 86
272. Telling, R. M., Menzies, B. K., and Simons, N. E., "Cut-Off Efficiency,
Performance and Design," Ground Engineering, Vol 11, No. 1, Jan 1978a,
pp 30-43. Available from: Technical Information Center, U. S. Army
Engineer Waterways Experiment Station, P. O. Box 631, Vicksburg,
MS 39180-0631.
273. Telling, R. M., Menzies, B. K., and Simons, N. E., "The Effectiveness of
Jointed Cut-Off Walls Beneath Dams on Pervious Soil Foundations," Ground
Engineering, Vol 11, No. 4, May 1978b, pp 27-37. Available from:
Technical Information Center, U. S. Army Engineer Waterways Experiment
Station, P. O. Box 631, Vicksburg, MS 39180-0631.
274. Terzaghi, K., Theoretical Soil Mechanics, Wiley, New York, 1943. Avail-
able from: Technical Information Center, U. S. Army Engineer Waterways
Experiment Station, P. O. Box 631, Vicksburg, MS 39180-0631.
275. Terzaghi, K., and Peck, R. B., Soil Mechanics in Engineering Practice,
2d ed., Wiley, New York, 1967. Available from: Technical Information
Center, U. S. Army Engineer Waterways Experiment Station, P. O. Box 631,
Vicksburg, MS 39180-0631.
277. Thomas, H. H., The Engineering of Large Dams, Vol II, Wiley, 1976.
Available from: Technical Information Center, U. S. Army Engineer
Waterways Experiment Station, P. O. Box 631, Vicksburg, MS 39180-0631.
278; Todd, D. D., and Bear, J., "River Seepage Investigation," Water Resources
Center Contribution No. 20, Sep 1959, Hydraulic Laboratory, University of
California, Berkeley. Available from: Technical Information Center,
U. S. Army Engineer Waterways Experiment Station, P. O. Box 631,
Vicksburg, MS 39180-0631.
279. Todd, D. K., Groundwater Hydrology, 2nd ed., Wiley, New York, 1980.
Available from: Technical Information Center, U. S. Army Engineer
Waterways Experiment Station, P. O. Box 631, Vicksburg, MS 39180-0631.
A-27
EM 1110-2-1901
30 Sep 86
282. Vaughan, P. R., "Design of Filters for the Protection of Cracked Dam
Cores Against Internal Erosion," Preprint 3420, Oct 1978, American
Society of Civil Engineers, New York. Available from: Technical Infor-
mation Center, U. S. Army Engineer Waterways Experiment Station, P. O.
Box 631, Vicksburg, MS 39180-0631.
283. Vaughan, P. R., et al., "Cracking and Erosion of the Rolled Clay Core of
Balderhead Dam and the Remedial Works Adopted for Its Repair," Transac-
tions of the Tenth International Congress on Large Dams, Vol 1, 1970,
Montreal, Canada, pp 73-93. Available from: Technical Information
Center, U. S. Army Engineer Waterways Experiment Station, P. O. Box 631,
Vicksburg, MS 39180-0631.
284. Veder, C., "Closing Address," Diaphragm Walls and Anchorages, 1975,
pp 221-225, Institution of Civil Engineers, London. Available from:
Technical Information Center, U. S. Army Engineer Waterways Experiment
Station, P. O. Box 631, Vicksburg, MS 39180-0631.
286. Vennard, J. K., Elementary Fluid Mechanics, 4th ed., Wiley, New York,
1965. Available from: Technical Information Center, U. S. Army Engi-
neer Waterways Experiment Station, P. O. Box 631, Vicksburg,
MS 39180-0631.
287. Walton, W. C., "Selected Analytical Methods for Well and Aquifer Evalua-
tion," Bulletin 49, 1962, Illinois State Water Survey, Urbana, Ill.
Available from: Technical Information Center, U. S. Army Engineer
Waterways Experiment Station, P. O. Box 631, Vicksburg, MS 39180-0631.
288. Wei, C.-Y., and Shieh, W. Y. J., "Transient Seepage Analysis of Guri
Dam," Journal, Technical Councils, American Society of Civil Engineers,
Vol 105, No. TCl, Apr 1979, pp 135-147. Available from: Technical
Information Center, U. S. Army Engineer Waterways Experiment Station,
P. O. Box 631, Vicksburg, MS 39180-0631.
289. Wilkins, J. K., "Flow of Water Through Rockfill and Its Application to
the Design of Dams," Proceedings, Second Australia-New Zealand
Conference on Soil Mechanics and Foundation Engineering, New Zealand,
1956, pp 141-149. Available from: Technical Information Center, U. S.
Army Engineer Waterways Experiment Station, P. O. Box 631, Vicksburg,
MS 39180-0631.
A-28
EM 1110-2-1901
30 Sep 86
290. Wilson, C. R., and Witherspoon, P. A., "An Investigation of Laminar Flow
in Fractured Porous Rocks," 1970, Department of Civil Engineering,
University of California, Berkeley, Calif. Available from: Technical
Information Center, U. S. Army Engineer Waterways Experiment Station,
P. O. Box 631, Vicksburg, MS 39180-0631.
291. Wilson, S. D., and Marsal, R. J., "Current Trends in Design and Con-
struction of Embankment Dams," 1979, American Society of Civil Engi-
neers. Available from: Technical Information Center, U. S. Army Engi-
neer Waterways Experiment Station, P. O. Box 631, Vicksburg,
MS 39180-0631.
292. Winter, G., and Nilson, A. H., Design of Concrete Structures, 9th ed.,
McGraw-Hill, New York, 1979. Available from: Technical Information
Center, U. S. Army Engineer Waterways Experiment Station, P. O. Box 631,
Vicksburg, MS 39180-0631.
293. Xanthakos, P. P., Slurry Walls, McGraw-Hill, New York, 1979. Available
from: Technical Information Center, U. S. Army Engineer Waterways
Experiment Station, P. O. Box 631, Vicksburg, MS 39180-0631.
295. Younger, J. S., and Lim, C. I., "An Investigation into the Flow Behavior
Through Compacted Saturated Fine-Grained Soils with Regard to Fines
Content and over a Range of Applied Hydraulic Gradients," Fundamentals
of Transport Phenomena in Porous Media, International Association for
Hydraulic Research, New York, 1972, pp 312-326. Available from:
Technical Information Center, U. S. Army Engineer Waterways Experiment
Station, P. O. Box 631, Vicksburg, MS 39180-0631.
A-29
EM 1110-2-1901
30 Sep 86
APPENDIX B
APPROXIMATE METHODS FOR ANALYSIS OF FLOW PROBLEMS
Sensitivity studies may be run to establish the effect of parameters not known
accurately.
B-1
EM 1110-2-1901
30 Sep 86
Table B-1. Analogy Between Darcy's Law and Ohm's Law (a)
anywhere in the fluid. Extensive use of the WES model has been made to:
(1) Determine uplift values and seepage quantities for use in the
design of Columbia Lock and Dam, Louisiana (Duncan 1962).
(2) Determine the uplift values and seepage quantities for fully and
partially penetrating well arrays from line and circular sources (Duncan 1963,
Banks 1963, and Banks 1965).
B-3. Sand Tank Model. The sand tank model (hydraulic model), as shown in
figure B-2, consists of a rigid, watertight container with a transparent front,
filled with sand, deaired water,(1) and measuring devices. The geometry of the
sand tank corresponds to that of the prototype. The sand may be placed under
water to provide a homogeneous condition, or layers of different sand sizes may
be used to study anisotropy. If the flow is unconfined and the same material
is used for model and prototype, the capillary rise must be compensated for in
the model. When a steady-state flow is reached, dye can be introduced at
various points along the upstream boundary close to the front wall to form
traces of the streamlines. Piezometers are used to measure the pressure heads
at various locations (Bear 1972 and Harr 1962). A sand tank model was employed
to investigate the effect of length of horizontal drain on the through seepage
flow nets and quantities for a homogeneous and isotropic sand embankment
(1)
For prolonged tests, disinfectants such as Formol should be added to the
water to prevent bacterial growth that causes clogging (Bear 1972).
B-2
B-3
EM 1110-2-1901
30 Sep 86
36
Figure B-1. Three-dimensional electrical analogy apparatus (after Duncan )
EM 1110-2-1901
30 Sep 86
TRACER DYE
SEEPAGE IN EMBANKMENTS
Figure B-2. Hydraulic or sand tank model
(prepared by WES)
B-4
EM 1110-2-1901
30 Sep 86
(Brand and Armstrong 1968). Sand tank models are also used extensively in
petroleum engineering, ground-water quality, and pollution research (Bear 1972
and Prickett 1979).
B-4. Viscous Flow Models. The viscous flow model, also called the Hele-Shaw
or parallel plate model, is based on the similarity between the differential
equations governing saturated flow in a porous medium and those describing the
flow of a viscous liquid in the narrow space between two parallel plates. The
viscous flow model contains the shape of the structure to be Investigated and
once a steady-state flow is obtained, colored dyes can be injected along the
upstream edge and patterns of streamlines can be observed. A camera (movie or
still) is normally used to record the results of experiments. Inhomogeneous
hydraulic conductivity, such as would exist in a zoned earth dam, can be simu-
lated by varying the width of the interspace between the parallel plates, as
shown in figure B-3. The viscous flow model experiments should be conducted in
a temperature-controlled room because viscosity plays an important role in
analog scaling. If this is not feasible, the temperature should be measured at
all inflow and outflow points during the test and scales must be recomputed
according to the varying average temperature of the liquid in the model (Bear
1972 and Harr 1962). A viscous flow model was constructed at WES to simulate
seepage conditions induced in streambanks by sudden drawdown of the river level.
The results from the model study compared favorably with field observations,
finite difference, and finite element methods (Desai 1970 and Desai 1973).
(B-1)
where
k - coefficient of permeability
B-5
EM 1110-2-1901
30 Sep 86
140
Figure B-3. Viscous flow model (courtesy of Bear )
B-6
EM 1110-2-1901
30 Sep 86
(B-2)
Since summation of the head loss in each fragment is equal to the total head
loss, the total quantity of flow can be expressed as
(B-3)
where h is the total head loss through the section . Along the same line, the
head loss in each fragment can be calculated from
(B-4)
(B-5)
B-7
EM 1110-2-1901
30 Sep 86
(B-6)
(B-7)
This elemental section will be used to derive the form factors for fragment
types IV, V, and VI.
Q/kh or
(B-8)
The form factor could also be expressed as the ratio of the elliptic integral
of the first kind with modulus m over the elliptic integral of the complemen-
tary modulus, m' . For this fragment type, the modulus value is a function of
the ratio S/T . The graph in figure. B-7 was obtained by solving the elliptic
integrals for various combinations of S/T . For the type II fragments, the
ratio of b/T equals 0 .
(4) Type IV. This type is an internal fragment with boundary length
b , embedment length S , in a pervious layer of thickness T . Figure B-9a
illustrates the two possible configurations. Pavlovsky divided the flow region
into active and passive parts based on the results of electrical analogue tests
as shown in figure B-9b by line AB . An angle of 45 deg was assumed for the
line dividing the two parts of the fragment. This resulted in two cases,
depending on the relation between b and S . For the case where b < S , the
B-8
Fragmcnl Form f<Klor, 11• t h is head Fragment Form fm.:lor. 11t ( h is head
Illustration llluslruliun
1rpc lns'i lhrou~h fru~mcnl) luss lhrou~h fntKmenll
·~"~- -- - - -- - -----·---
L s 2s:
ct• ... ,_ ~ = 2 In (I + :~ I
II v L2!2s:
~=21n (I+ ~)+L_~ 21
L 2:s' + s":
II ~= t<th),Fig.5-13 ~ = In I (I + .!:a
> (1 +~.:I I
II
I. - ( s' + s"l
+-------
T
VI
L ::s s' + s":
~ = In I (I + a~:)(I + II~:) I
where L+(s'-s"l
Ill ~ = I (~h), Fig. 5-13 ,'1' = - - - - 2 -
h" =!_,=(J'_-__!:'1
B-9
VII
b::ss: b y
«<•=ln(l+;;-)
IV VIII
1l>=ln(i+~)
+ b-
- ---'
T
IX
EM 1110-2-1901
30 Sep 86
Figure B-4. Summary of fragment types and form factors (courtesy of McGraw-Hill
181
Book Company )
EM 1110-2-1901
30 Sep 86
B-10
EM 1110-2-1901
30 Sep 86
1.5
1.4 l
j
1.3
1.2 \ I u
=-------_:~~ "- - - ~
·...Oo
i
1.1 \ I I - - - - - oo·a··
I -- - - - - - - - - 0 P.O. ..wa·O;t:E""
o
1.0
'r\ I I
I
I ,;;,,;gy:;..J§_2"fH··£C
~/,1'/ffiW/~$&##/m~r
0.9
\I I
\t I
-1~ 0.8
II
0.7
r--... ~ ~~ I
ol~
0.6
~~ I I
.-.!!so ~.
I
...... I
0.5
0. 7;-~~ '-
1 I
I
0.4
0.3
l.OO-
1.25
r---r---~
I ~ I
:::::::::::±-::~ ~
..:,~
I
1.50
0.2 -...~
~
I
........
~:
0.1
~
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
_s
T
h2 TTb
(f m = cos Ti
17"$
tan 2 t + tan 2 2T
TT S
Q 1 4
for m s 0.3, kh = -;r In m
09...Q_ _ _ ,.
for m 2 ~ . . kh - 1 2
21n(--...':!!.)
16
B-11
EM 1110-2-1901
30 Sep 86
(a) (b)
Figure B-8. Type III fragment (courtesy of McGraw-Hill
Book Company 181 )
(B-9)
If b > S, then the fragment can be divided into two fragments as shown in
figure B-9e. The first is a type IV with b > S and the second is a type I
fragment with L equal to b - S . Thus the form factor is the sum of the
form factors which would be
(B-10)
(5) Type V. This fragment type has two vertical boundaries of equal
embedment S in a pervious layer of thickness T . As shown in figure B-10,
the form factor for this fragment is twice that for the type IV fragment.
B-12
EM 1110-2-1901
30 Sep 86
B-13
EM 1110-2-1901
30 Sep 86
Since there were two cases of type IV fragment, there are two cases for the
type V fragment. The two cases are for L < 2s and L > 2S . For the first
case, the form factor is
(B-11)
For the second case which consists of a type I fragment within two type IV
fragments, the form factor is
(B-12)
(6) Type VI. This fragment type, illustrated in figure B-11, is the
same as the type V fragment except that the embedment lengths are different.
Using the same approximations as in fragment type IV, there are two cases for
the form factor. For the first case where L > (S' + S"), the form factor is
(B-13)
B-14
EM 1110-2-1901
30 Sep 86
For the second case where L < (S' + S"), the form factor is
(B-14)
where
(B-15)
B-15
EM 1110-2-1901
30 Sep 86
(B-16)
(B-17)
B-16
EM 1110-2-1901
30 Sep 86
(B-18)
(9) Type IX. This fragment type, shown in figure B-14, represents the
exit condition where the surface of seepage exists. The surface of seepage
(DE) is not an equipotential line or a streamline. Pavlovsky assumed that the
flow is horizontal. For the portion of the slope between D and E , the
flow is the coefficient of permeability times the integral of dy over
cot The flow for E to F is the permeability times the integral of
a2dy over the cot (a2 + h2 - y) . When the integration is performed, the
expression for the flow is
(B-19)
B-17
EM 1110-2-1901
30 Sep 86
(B-20)
where
As defined before, the modulus m is a function of both b/T and S/T and
is defined as
(B-21)
B-18
EM 1110-2-1901
30 Sep 86
Instead of calculating the various values, for type II fragments figure B-15
can be used with S/T to obtain the fraction for IES/hm . By substituting
the appropriate values, the exit gradient is calculated.
Region hi
1 0.641 1.51
2 6.514 15.33
3 0.495 1.16
= 7.650 = 18.00
The head along the bottom decreases from 16.49 ft at the upstream end to 1.16
at the downstream end. Using the assumption of a linear distribution of the
head loss within a fragment, the head at any point along the bottom of the
lock could be calculated as
B-19
EM 1110-2-1901
30 Sep 86
B-20
EM 1110-2-1901
30 Sep 86
type VIII fragment while region 2 is a type VII fragment and region 3 is a
type IX fragment. To calculate the flow through region 1, equation B-18 is
used with h = 32 ft; h d = 37 ft, and = 19.9 deg (cot = 2.76).
i
Substituting into the equation produces
(B-23)
(B-24)
(B-25)
B-22
EM 1110-2-1901
30 Sep 86
(B-26)
By substituting into equation B-26, there are four equations with four
unknowns, h , a2 , Q/k , and L . There are several methods to solve these
four equations (Harr 1977). For the case where h2 = 0 , a reduction of
two equations and two unknowns occurs. For this example, equation B-23 will
be combined with equation B-24 and equation B-26 will be substituted for L .
This produces
(B-27)
(B-28)
Equations B-27 and B-28 have reduced the equations and unknowns by two. Thus
with two equations and two unknowns, a trial and error graphical process can
be used. The results of this process are shown in figure B-18 and indicate
that h = 28.9 ft and a2 = 0.9 ft., Substituting into equations B-23 and B-25
generates a Q/k value of 1.71 which results in an estimated flow of
242.4 ft3/day/ft of dam. Knowing h and a2 , the location of the phreatic
surface can be estimated.
(B-29)
where
B-23
EM 1110-2-1901
30 Sep 86
Equation B-27
h a2
28.8 -0.4
29.0 2.1
30.0 8.6
31.0 13.6
Equation B-28
h a
2
29.9 0.6
27.8 1.2
25.6 1.7
The ratio of permeabilities can vary from 0 to infinity. Over this range,
ranges from 0 to 1/2. The basis of this method is to determine the flow and
head losses for three certain special cases of and then interpolate
between these values. The three special cases are as follows:
B-25
EM 1110-2-1901
30 Sep 86
Point A Point B
Head Loss Total Head Head Loss Total Head
ft ft ft ft
0 5.2 52.8 12.8 45.2
1/4 6.2 51.9 12.0 46.0
For the case where = 1/2 , the head anywhere along the bottom of the struc-
ture is equal to half the total head loss, or for this case 9 ft. Thus the
total head on points A and B is equal to 49 ft. Figure B-20b is the plot of
the total head versus and shows, for an of 0.48, the total head at
point A is 49.4 ft while the total head at point B is 48.5 ft. The exit gra-
dient for each case is calculated by the procedure described in example 1.
For the case of 0 , the fraction IES/hm is 0.55 which with S = 24 ft
B-26
EM 1110-2-1901
30 Sep 86
04
0.3
0
'~-0.2
0.1
0.5
E
IAI
€
1- 51 IC)
LL.
ci'
<(
50
~
...J 49
<(
1-
~ 48
47
46
45
0.5
€
18)
Figure B-20. value plots for John H. Overton Dam and stilling
basin (from U. S. Army Engineer District, St. Louis 112 )
B-27
EM 1110-2-1901
30 Sep 86
and hm = 3.8 ft produces IE = 0.087 . For the = 1/4 case, the fraction
I E S / h m is 0.615 which with S = 24 ft and hm = 5.0 ft produces
IE = 0.128 . For the case where = 1/2 , the head loss is the total head and
the distance is the thickness of the top layer. Using the equation
(B-30)
the exit gradient is 0.265. The exit gradient versus plot is shown in
figure B-20c. For an value of 0.48, the exit gradient is 0.245.
B-28
EM 1110-2-1901
30 Sep 86
B-29
EM 1110-2-1901
31 Jul 86
1977b). A listing of finite element seepage computer programs used within the
Corps is available (Edris and Vanadit-Ellis 1982).
(1) WES studies. As discussed previously, the finite element method was
used at WES to simulate seepage conditions in streambanks induced by sudden
drawdown of the river level. This study included a viscous flow model, field
observations, and application of the finite difference and finite element
methods. The results of the study indicated that the finite element method
provided satisfactory solutions for transient unconfined fluid flow in porous,
anisotropic, and nonhomogeneous media (Desai 1970 and Desai 1973).
(2) Location of phreatic surface. The finite element method has been
used to determine the location of the phreatic surface in earth dams (Isaacs
1979; Isaacs 1980; Wei and Shieh 1979; and Desai and Kuppusamy 1980). The
finite element method was used to locate the phreatic surface within tailings
pond embankments and to define the subsurface flow of water from the pond.
Results predicted using the finite element model were confirmed with
measurements in a laboratory model and in the field (Kealy and Busch 1971).
(3) W.A.C. Bennett Dam. The finite element method was used to assess
the potential seepage flows and uplift pressures in the foundation rock for
W.A.C. Bennett Dam in British Columbia, Canada (see figure B-21). The finite
element analysis (see figure B-22) was carried out assuming the following
conditions:
The results of the finite element analysis, shown in figure B-23, indicate the
greatest reduction in seepage flow and hydrostatic pressure could be accom-
plished by an effective grout curtain and downstream-drainage system (Taylor
and Chow 1976).
B-30
EM 1110-2-1901
30 Sep 86
(4) Corps of Engineers levees. The finite element method was used by
the U. S. Army Engineer District, Rock Island, to study hydraulic sand fill
levees along the Mississippi River (Schwartz 1976). Finite element and
gradient plane (1) analyses were used in conjunction with data from a full scale
test levee to establish the material properties of the sand levees and to
determine the exit point of the free seepage surface, the quantity of through
seepage, and the exit gradients along the free discharge face. A parameter
study was performed and dimensionless design charts were developed.
(1) The gradient plane method is a graphical solution by means of the hodo-
graph (see description by Casagrande 1937).
B-31
EM 1110-2-1901
30 Sep 86
B-32
EM 111O-2-1901
30 Sep 86
foundation (Mantei and Harris 1979). Narrows Dam, Colorado, on the South
Platte River, was analyzed for seepage at the feasibility stage. Because of a
pervious foundation, the planners called for a positive vertical cutoff by
constructing a slurry wall down to the underlying shale. However, near the
right abutment the shale drops away to depths too great for economical slurry
wall construction. A three-dimensional finite element model (see figure B-24)
was used to determine the vertical exit gradients at the downstream toe of the
dam. The finite element method was used to study the effect of a toe drain,
partial depth slurry trench, partially and fully penetrating relief wells (see
figure B-25). Calamus Dam, Nebraska, on the Calamus River, was also analyzed
by the Bureau of Reclamation for seepage at the feasibility stage (Mantei and
Harris 1979; Mantei, Esmiol, and Cobb 1980; Mantei and Cobb 1981; and Cobb
1984). Calamus Dam has a setting very similar to Narrows Dam in the sense that
it is an earth dam on a pervious foundation. However, the underlying shale at
Calamus Dam is at such a great depth that it cannot be used as the base for a
cutoff wall as it was for Narrows Dam. Early thinking on the project involved
the use of a slurry trench cutoff down to a pervious sandstone fromation. A
three-dimensional finite element model (see figure B-26) was used to determine
the effects of an embankment toe drain, slurry trench under upstream blanket,
and/or relief wells at the downstream toe of the dam on the seepage rates and
hydraulic gradients in the dam foundation. Time and expense in operating the
large three-dimensional finite element models made it necessary that priority
be given to studying the various design alternatives using the best estimate
of permeability for each foundation material rather than conducting sensitiv-
ity studies to establish the effect of varying the permeability (see para-
graph B-l). The three-dimensional finite element models were five elements
deep, with the bottom layer of elements representing the sandstone, the next
layer sand and gravel, recent alluvium, interbedded fine sand, and dune sand.
A detailed three-dimensional finite element model was made for the outlet
works area that defined more of the design details, such as the filter blanket
under the stilling basin and channel and water table elevation controls, to
study the effectiveness of relief wells around the stilling basin.
(7) Corps of Engineer dams. The finite element method was used by the
U. S. Army Engineer District, Huntington, (1) in a reanalysis of the underseep-
age at Bolivar Dam, Ohio, completed in 1938 on Sandy Creek (U. S. Army Engineer
District, Huntington 1977a). A two-dimensional finite element model (see fig-
ure B-27) was used to determine the effects of an embankment toe drain, up-
stream impervious blanket, and proposed relief wells on seepage quantities,
exit gradients, and uplift pressures. A sensitivity study was conducted using
the two-dimensional finite element model to determine the influence of various
pool and rock surface levels, the permeability ratio of foundation soils, the
existence of a downstream gravel layer, and the effective source of seepage
entry upon underseepage. Typical test results for one set of boundary condi-
tions and permeability values are given in figure B-28. Additional applica-
tions of the two-dimensional finite element method to conduct sensitivity
analysis to assess the effect of permeability anisotropy and various seepage
control measures was given by Lefebvre and coworkers (Lefebvre, Part, and
(1) Work was performed by Soil Testing Services, Inc., Northbrook, Illinois.
B-33
EM 1110-2-1901
30 Sep 86
~xis of Dam
I
f§j ~J~I~1~~~~~~~ff-J *I
SECTION
I I I
: ; .: I
; I:
! l I :1 0
0
I ; j il N
(
0
:
~!
v.,
I
~l ~st\.
I
I I'
G
... I I ~;
-;;
i0 '
I
' I
I
.............
....End o kev rend
and cuto f
\ I\ 1\ g ~
Cl
\ \I ~oiiiiiiiF
-
u \
0
...0 ': ~~
e
•
~
', 0
...
I»
c: ...... I Ui
0 .... ,.. I 0.
:l
0 -o0
• I ~
0
v ..c:
I
·o...
Toe of dcim-........ 0
0
... - ./
I
C\1
.,...
Cl >
~
i I II ...
~
.,.
E I Q)
~
I .,.,E
:l
I
I
II 11
~
r
PLAN
B-34
EM 1110-2-1901
30 Sep 86
~~----~------~------~------~----~
~~----~-------+-------r------~----~
t-
z ~~----~-------+-------r------~----~
LLJ
0
<l: 0
o--- ~ .... -. -~--.
a::
(!)
t:
X
LLJ
_J <.0
~ 0~1----tr--------+-------~----~------~
(..)
t- ~
a:: 6~~--~-------+------~------~----~
~ ~---L- -
0
6~----~------~------~------~----~
-300·0 0·0 300·0 600·0 900-<J 1200-Q
LEGEND
D 3-0 MODEL WITHOUT REUEF WELLS
0 2-o MODEL W\THOUT RELIEF WELLS
~ QUASI 3-0 MODEL WtTH RELIEF WELLS
AT 100' SPACING
B-35
EM 1110-2-1901
30 Sep 86
B-36
llf. . .TION
I
•• ,
I
tJ
~
~::_.-;,_
I ~
- - c=: -- ----- - - -- -- - ----
_100
' I !
-- e---t-=1 . I
I
I
I-
-- ...
_
+=-+-
L ____
i
I
--t------
--
_•oo
B-37
SCAL£ tFtET)
Figure B-27. Grid for two-dimensional finite element model ·study of Bolivar Dam, Ohio
EM 1110-2-1901
88
(from U. S. Army Engineer District, Huntington )
30 Sep 86
30 Sep 86
EM 1110-2-1901
MODEl PREDICTIONS
-----l
I f (;:rilrotfTRlt }..~'-LL$ !'l 1
1
1,/JI\Nlll'i (f SC(f-'AG(. Am-t~'r..A..
[Uit-i.I ~"a~:.r:AUr'l.'j 1...{,- C!UJ'C"'
1 9
~ ~-~:-:t~,t.fl":(~2~~:r"'t:l! ~ I~ r.. '' (.4~ !It AI~~ I 1 !til f{'J lt.-o I
i~-.:._--: --:;-i-:-~-~J::Ji~---=i_:Ol_~fl
t 1 _ _ _ ~ _ _L _
-==-t;:::-t~
J _ _ t __ ~ :~ _~~ ~ ~:-
_ _jt-- __1:: _J
J
1 1fCLIJOl, fii[L.f .. ll~ rL(N • . , nr~-
~- .
,~ ~
~
!""
~
.!'0
B-38
.!!!!!.
'!"____ 'I""
NOIIIIOIIITAL DltT4f'Ct
~~
UI'AT. . . "'Ill
to . . . . . . UVIL DllfUM
-
....,.·-flle
__ __.... ,.... ,...
"'--AIUT"I' •m ••• , •• a.a, . . .,..
-.uwtD ---~
. .. fill. . . . ..,..
. . . . . . IITOTOlii'MI
Figure B-28. Results from two-dimensional finite element model study of Bolivar Dam, Ohio
88
(from U. s. Army Engineer District, Huntington )
EM 1110-2-1901
30 Sep 86
Tournier 1981). The finite element method was used by the U. S. Army Engineer
District, Huntington, (1) in a reanalysis of the underseepage at Mohawk Dam,
Ohio, completed in 1937 on the Walhonding River (U. S. Army Engineer District,
Huntington 1979b). A three-dimensional finite element model (see figure B-29)
was used to study the cause of unusually high relief well flows. Typical test
results for one set of boundary conditions are given in figure B-30.
(1) Work was performed by Soil Testing Services, Inc., Northbrook, Illinois.
B-39
EM 1110-2-1901
30 Sep 86
- IL ...
1\. •••
l '/
...'"
Ito
,'
}-LI,--~+------~---~~~=~~:::_~,~~~---~-~------~---~---
I
I ' ' ' •
.'
.
/1'
I '
I I ' ..,....... j
l
,/ ' . t
. ; 3i 0 g 8 8
! !
0
~ ~ .• .: .• ! 3
. .. ;
:,; . ! .
0 0
~ ~.; .• ;
g 0
•• !
0
i : ; 0 !l.l'l~ ...,.. . .ll't!.
• •
..
~ ~ ~ ~ ~ •I ,; ,;
.
0
"
I
- ·".
I I I __ ___, CDOC . . . Afl
"'I
I I I I I I I
'
I
t
B-40
-·-
-.· 1\f.
'
).:.' I
.
(i) ..............~
. ....
(i)
w ......
JCU'CUittaOUh...l"
~-~ tcl'I.UI
•• '""'
.... .___•••u•·SCAU
..., .. ..c-·•••
.
.
..
.......:, :J) •:..tU •'-·
""'
. ·-
. _ , ,_ _ ~k~
... -.a.u
,..
...""
f) ~,.._
(fl .......,~, ......
""'"'" 0 ....,. .......r "
_,.VICIUS ••
....
@ ..._.. OCDOI 0-
JfiUIIIOI''i(loOUI~
Figure B-29. Grid for three-dimensional finite element model study of Mohawk Dam, Ohio
89
(from U. S. Army Engineer District, Huntington )
OIIT~UO lt&ll
I l l fiiiiT
I
IliUM
- I
I 1;
I
1/ ~
~
~
I
'
' ~
I
i
!II
- .
Cf ..
I
I I I I I
I
i
I I
SU,...GI1\
(Ni'Hi
I
:= ~
i
~ •.,..c,.. j
;- i
:
i ·I
111
:
I
I
I
I
: I: I I: I . ~~I I=
I•
I• i! ~' 1- • i• .•• " ..
!j =~ jl if j if ~ -
al f-
! I!
I
i
i' i I
;
' '
i~l ~('
j~ ~~~
! I I I i
I I ; r~· :, I II I I
I
I I
i
i ·~ ~·
~- ! : I
!
I'
i
!
I
!
I
i
B-41
1T \ \
I
I
i
I t: ~I
I It~~~ II : •
-
ta4 trw
- .
.,. c, ..
.
I
I
I
.
I
I
I
i
I
I
I
I
I
I
·~·I
MI ~
. I""'-~··
I
I
i
'
Il
.=
i I
.
d L! <~ H J
:- r i" r- g 'i
1 2
\
~~:1i\
1'\ - II
i
I! I
I.
!.. ". : \
~--
1. . . " .
~--
I
. 1
j
Ir- ..
m
: I
!~ l I 'l/ I ~~~~/:
:
';·~ ~
·~·I
I I
-
t •
I l l " .. I
! I i I
I i
•·•o
1"·20
·-·
•·t
TOTIL WE.L.l DISC:>IMG( •
llV C:I'"M
7i"i"Cni"
EM 1110-2-1901
110 5
1100
30 Sep 86
Figure B-30. Results from three-dimensional finite element model study of Mohawk Dam, Ohio
89
(from U. S. Army Engineer District, Huntington )
EM 1110-2-1901
30 Sep 86
APPENDIX C
ANALYSIS OF PRESSURE INJECTION TESTS (Ziegler 1976)
(3) Radial flow from a cylindrical and vertical borehole test section
length, Radial flow implies that the equipotential surfaces form
cylinders symmetrical about the axis of the borehole test section.
(a) At r = ro . H = Ho .
Where
H o = excess pressure head at the center of the test section (L) (either
measured within the test section, or computed allowing for fric-
tional head losses within the drill pipe)
C-l
EM 1110-2-1901
30 Sep 86
{WATER FLO*)
corresponding to a 100 percent loss in excess head, Ho (L)
GROUND SURFACE
~ z
---
lLJ
...J
~j:: PACKER
lJ
lLJ
""I-
V,
lLJ
1-
/
"-- \ "'""
t
\.:)'
z
'"-I---PACKER ) ..
lLJ
...J
z0
i:
\
lJ
lLJ
""
1-
""lLJ1-
CLOSED
/ ) ~
C-2
EM 1110-2-1901
30 Sep 86
where
where
C-3
EM 1110-2-1901
30 Sep 86
2
A = area of an equipotential surface (L )
Thus
and integrated
(C-1)
The negative sign can be dropped by choosing flow away from the borehole as
positive, thus
C-4
EM 1110-2-1901
30 Sep 86
(C-2)
Equivalent Permeability,
Radius of Influence,
k e , ft/sec
R , ft
1
ke
10 2.1 k e
100 3.2 ke
1,000 4.3 k e
10,000 5.4 k e
100,000 6.5 ke
1,000,000 7.5 k e
Integration yields
C-5
EM 1110-2-1901
30 Sep 86
(C-3)
(3) Radial flow from cylindrical and vertical borehole test section of
length,
(a) At r = ro , H = Ho .
where
C-6
EM 1110-2-1901
30 Sep 86
Drop the negative sign by choosing flow away from the borehole as positive,
yielding
(C-4)
(C-5)
C-7
EM 1110-2-1901
30 Sep 86
where
The logarithms of terms in equation C-5 yield the equation of a straight line:
(3) Radial flow occurs within each fissure and is governed by Darcy's
law. No flow occurs in material between fissures.
(4) Each fissure has the same equivalent parallel plate aperture, e .
(a) At r = ro , H=Ho .
(b) At r = R , H=0.
j. The derivation of equations C-6 and C-7 proceeds in the same fashion
as that given above for laminar flow through a homogeneous isotropic medium
(equations C-2 and C-3). For the fissured medium the test section length,
is replaced by the quantity (ne) where n is the number of fissures
intersecting the test section and e is the equivalent parallel plate
aperture of each fissure. The resulting expression for the flow rate Q is
C-8
EM 1110-2-1901
30 Sep 86
(C-6)
(C-7)
C-9
EM 1110-2-1901
30 Sep 86
(C-8)
where
Equation C-8 is substituted into equation C-6 to solve for the equivalent
parallel plate aperture, e :
(3) Radial flow occurs within each fissure and is governed by the
Missbach law (vm = k'i). No flows occur in material between fissures.
j
(4) Each fissure has the same equivalent parallel plate aperture, e .
(a) At r = ro , H = Ho .
(C-9)
(C-10)
where
v = ki
C-11
EM 1110-2-1901
30 Sep 86
where
(C-11)
where
C-12
EM 1110-2-1901
30 Sep 86
where
The term [C] is dependent on assumed flow and boundary conditions. The follow-
ing equation is commonly used in analyzing water pressure test results.
(C-12)
where
(C-13)
(3) The flow pattern is ellipsoidal and symmetrical about the axis of
the borehole test section.
C-13
EM 1110-2-1901
30 Sep 86
(C-14)
(C-15)
(C-16)
where
(C-17)
C-14
EM 1110-2-1901
30 Sep 86
(C-18)
where
(C-19)
i. The weight flow rate from the test section, QWF , is constant along
the flow path and can be determined at the manifold by applying equation C-17:
C-15
EM 1110-2-1901
30 Sep 86
where
where
= unit weight of air at the manifold (lbF/ft3)
C-16
EM 1110-2-1901
30 Sep 86
where
C-17
EM 1110-2-1901
30 Sep 86
(C-20)
where
(1) Radial flow occurs from a vertical test section and is governed by
Darcy's law.
(3) Test zone is saturated. Maini (1971) suggests that before conduct-
ing tests in zones above the ground-water table, water be pumped into the bore-
hole test section to saturate the fissure system in the immediate vicinity of
the borehole.
(C-21)
(C-22)
C-18
EM 1110-2-1901
30 Sep 86
(C-23)
C-19
EM 1110-2-1901
30 Apr 93
Change 1
APPENDIX D
FILTER DESIGN
D-1. General. The objective of filters and drains used as seepage control
measures for embankments is to efficiently control the movement of water
within and about the embankment. In order to meet this objective, filters and
drains must, for the project life and with minimum maintenance, retain the
protected materials, allow relatively free movement of water, and have suffi-
cient discharge capacity. EoLdesign,. these. three. necessLties are termed,
respectively, piping or stability requirement, permeability requirement, and
discharge capacity. This appendix will explain how these requirements are met
for cohesionless and cohesive materials, and provide general construction
guidance for installation of filters and drains. The terms filters and drains
are sometimes used interchangeably. Some definitions classify filters and
drains by function. In this case, filters must retain the protected soil and
have a permeability greater than the protected soil but do not need to have a
particular flow or drainage capacity since flow will be perpendicular to the
interface between the protected soil and filter. Drains, however, while meet-
ing the requirements of filters, must have an adequate discharge capacity
since drains collect seepage and conduct it to a discharge point or area. In
practice, the critical element is not definition, but recognition, by the
designer, when a drain must collect and conduct water. In this case the drain
must be properly designed for the expected flows. Where ic is not possible to
meet the criteria of this appendix, che design must be cauciously done and
based on carefully controlled laboratory filter tests (Perry 1985).
D-2. Stability. Filters and drains< 1 l allow seepage to move out of a pro-
tected soil more quickly than the seepage moves within the protected soil.
Thus, the filter material must be more open and have a larger grain size than
the protected soil. Seepage from the finer soil to the filter can cause move-
ment of the finer soil particles from the protected soil into and through the
filter. This movement will endanger the embankment. <2 > Destruction of the
protected soil structure may occur due to the loss of material. Also, clog-
ging of the filter may occur causing loss of the filter's ability to remove
(1)
In paragraphs D-2 and D-3 the criteria apply to drains and filters; for
brevity, only the word filter will be used.
(2)
In practice, it is normal for a small amount of protected soil to move
into the filter upon initiation of seepage. This action should quickly
stop and may not be observed when seepage first occurs. This is one
reason that initial operation of embankment seepage control measures
should be closely observed by qualified personnel.
D-1
EM 1110-2-1901
30 Apr 93
Change 1
water from the protected soil. Criteria developed by many years of experience
are used to design filters and drains which will prevent the movement of pro-
tected soil into the filter. This criterion, called piping or stability cri-
terion, is based on the grain·size relationship between the protected soil and
the filter. In the following, the small character "d" is used to represent
the grain size for the protected (or base) material and the large character
"D" the grain size for the filter material.
3. Prepare adjusted gradation curves for base soils with particles larger
than the No. 4 (4.75 mm) sieve.
b. Multiply the percentage passing each sieve size of the base soil
smaller than No. 4 (4.75 mm) by the correction factor from step 3a.
d. Use the adjusted curve to determine the percent passing the No. 200
(0.075 mm) sieve in step 4.
4. Place the base soil in a category based on the percent passing the No.
200 (0.075 mm) sieve in accordance with Table D-1.
5. Determine the maximum D15 size for the filter in accordance with
Table D-2. Note that the maximum D15 is not required to be smaller than
0.20 nun.
( 3)
Guide for Determining the Gradation of Sand and Gravel Filters, Soil
Mechanics Note No. 1, U.S. Department of Agriculture Soil Conservation
Services, Engineering Division, Jan 1986.
D-2
EM 1110-2-1901
30 Apr 93
Change 1
1 >85
2 40-85
3 15-39
4 <15
7. Set the maximum particle size at 3 in. (75 mm) and the maximum passing the
No. 200 (0.075 mm) sieve at 5 percent. The portion of the filter material
passing the No. 40 (0.425 mm) sieve must have a plasticity index (PI) of zero
when tested in accordance with EM 1110-2-1906, "Laboratory Soils Testing."
8. Design the filter limits within the maximum and minimum values determined
in steps 5, 6, and 7. Standard gradations may be used if desired. Plot the
limit values, and connect all the minimum and maximum points with straight
lines. To minimize segregation and related effects, filters should have rela-
tively uniform grain-size distribution curves. without "gap grading"--sharp
breaks in curvature indicating absence of certain particle sizes. This may
require setting limits that reduce the broadness of filters within the maximum
and minimum values determined. Sand filters with D90 less than about 20 mm
generally do not need limitations on filter broadness to prevent segregation.
For coarser filters and gravel zones that serve both as filters and drains,
the ratio D90 /D 10 should decrease rapidly with increasing D10 size. The limits
in Table D-3 are suggested for preventing segregation during construction of
these coarser filters.
D-3
I I
'EM 1110-2-1901
30 Apr 93
Change 1
(a) Category designation for soil containing particles larger than 4.75 mm is
determined from a gradation curve of the base soil which has been adjusted
to 100% passing the No. 4 (4.75 mm) sieve,
(b) Filters are to have a maximum particle size of 3 in. (75 mm) and a maximum
of 5% passing the No. 200 (0.075 mm) sieve with the plasticity index (PI)
of the fines equal to zero. PI is determined on the material passing the
No. 40 (0.425 mm) sieve in accordance with EM 1110-2-1906, "Laboratory
Soils Testing." To ensure sufficient permeability, filters are to have a
D15 size equal to or greater than 4 X d 15 but no smaller than 0.1 mm.
(d) A - percent passing the No. 200 (0.075 mm) sieve after any regrading.
(f) In category 4, the D15 :,; 4 X d 85 criterion should be used in the case of
filters beneath riprap subject to wave action and drains which may be
subject to violent surging and/or vibration.
D-4
EM 1110-2-1901
30 Apr 93
Change 1
<0.5 20
0.5 1.0 25
1.0 2.0 30
2.0 5.0 40
5.0 10 50
10 50 60
D-3. PermeabilitY. The requirement that seepage move more quickly through
the filter than through the protected soil (called the permeability criterion)
is again met by a grain-size relationship criterion based on experience:
Permeability
D-4. Applicability. The previously given filter criteria in Table D-2 and
Equation D-1 are applicable for all soils (cohesionless or cohesive soils)
including dispersive soils.( 4 l However, laboratory filter tests for disper-
sive soils, very fine silt, and very fine cohesive soils with high plastic
limits are recommended.
( 4)
Sherard, J. L., L. P. Dunnigan, "Filters and Leakage Control in Embank-
ment Dams," Proceeding of the Symposium on Seepage and Leakage from Dams
and Impoundments, ASCE National Convention, Denver, Colorado, 1985.
D-5
EM 1110-2-1901
30 Apr 93
Change 1
D-5. Perforated Pipe.C 5 l The following criteria are applicable for prevent-
ing infiltration of filter material into perforated pipe, screens, etc.:
In many instances a filter material meeting the criteria given by Table D-2
and Equation D-1 relative to the material being drained is too fine to meet
the criteria given by Equation D-2. In these instances, multilayered or
"graded" filters are required. In a graded filter each layer meets the
requirements given by Table D-2 and Equation D-1 with respect to the pervious
layer with the final layer in which a collector pipe is bedded also meeting
the requirements given by Equation D-2. Graded filter systems may also be
needed when transitioning from fine to coarse materials in a zoned embankment
or where coarse material is required for improving the water-carrying capacity
of the system.
(5) EM 1110-2-2300 states, "Collector pipe should not be placed within the
embankment, except at the downstream toe, because of the danger of
either breakage or separation of joints, resulting from fill placement
and compaction operations, or settlement, which might result in either
clogging and/or piping."
D-6
u .4 , _ ...I ,.._
I I
r\,
'"'f"'
. •
1
... .... .. . -." ,. ..
I I
\l t. fi_,.PAM
••
•t~ NWI:!US
1• UIO
•
_,..
-- -~--- ·+ -- -
..
f- ., -- 1- -.
I - ..... -- ···- -
.
. -- - 1- ~
\.
- \- - - - 1-
-I ~
I' ~· .. ~
-- - .. \ I
I - ·-· - -- - ··- -I- 1- -- '' - -- - -- e
fi
.
-" ~-
. -- - -·- 1 - . ---- - . - -- I -
-
.. ·- -- ~-
1-r-1- -- ~-
. , __ •
-I _,. I
..
-
-
.......I
.
...
•
I - I
UWM ••
san
•!""
..
.._LIN(fla
~
I ''!"
-. -
. ...
I
1- --·
...
• • Dill Q.Af
- 1- -·
....
-- ...
D-9. Example of Graded Filter Design for Drain. Seldom, if ever, is a single
gradation curve representative of a given material. A material is generally
represented by a gradation band which encompasses all the individual gradation
curves. Likewise, the required gradation for the filter material is also
given as a band. The design of a graded filter which shows the application of
the filter criteria where the gradations are represented by bands is illus-
trated in Figure D-2. A typical two-layer filter for protecting an impervious
core of a dam is illustrated. The impervious core is a fat clay (CH) with a
trace of sand which falls in Category 1 soil in Table D-2. The criterion
D15 :5 9 x d 85 is applied and a "point a" is established in Figure D-2. Filter
material graded within a band such as that shown for Filter A in Figure D-2 is .-
acceptable based on the stability criteria. The fine limit of the band was (
arbitrarily drawn, and in this example, is intended to represent the gradation
of a readily available material. A check is then made to ensure that the
15-percent size of the fine limit of the filter material band (point b) is
equal to or greater than 3 to 5 times the 15-percent size of the coarse limit
of the drained material band (point c). Filter A has a minimum D10 size and a
maximum D90 size such that, based on Table D-3, segregation during placement
can be prevented. Filter A meets both the stability and permeability require-
ments and is a suitable filter material for protecting the impervious core
material. The second filter, Filter B, usually is needed to transition from a
fine filter (Filter A) to coarse materials in a zoned embankment dam.
Filter B must meet the criteria given by Table D-2 with respect to Filter A.
For stability, the 15-percent size of the coarse limit of the gradation band
for the second filter (point d) cannot be greater than 4 to 5 times the
85-percent size of the fine limit of the gradation band for Filter·A
(point e). For permeability, the 15-percent size of the fine limit (point f)
must be at least 3 to 5 times greater than the 15-percent size of the coarse
limit for Filter A (point a). With points d and f established, the fine and
coarse limits for Filter B may be established by drawing curves through the
points approximately parallel to the respective limits for Filter A. A check
is then made to see that the ratio of maximum D90 /minimum D10 size of Filter B
is approximately in the range as indicated in Table D-3. A well-graded filter
which usually would not meet the requirements in Table D-3 may be used if
segregation can be controlled during placement. Figure D-2 is intended to
show only the principles of filter design. The design of thickness of a
D-8
U.S. STANDARD SIEVE OPENING IN INCHES U.S. STANDARD SIEVE NUMBERS HYDROMETER
111512 10 I I 4 3 2 1¥1 1 ¥. ~ ~ 3 • I I 10 U 11 20 3D •o 50 70 100 140 2:00
100
I' \ ' .' ' '' . D
.. ' \e
"-..
D. (das) min
10
10 11 1\ 1\ 1\ 20
\ ' 1\ \
70 '\ \
~
.,.
\ "" ~~ 30
\
_1 ~
\ \ ':::<
?
..
'\ ~0.~
~ 1\ '\.~c
t:1
; 50
fP
\ '-o,..,
' 1\ \
"' "' \ \ ''-
\ a.;; 9 (das) min '
30
\ \ ,'\ ~ 70
\ I ;;.3 to 5a l\ ~
,.
' I
al '\
\
D
d.;; to 5e 1\ ~
--"'- 90
• 100 .. I I I
10
\
0.5
"" 0.1 0.05 0.01 0.005
100
0.001
D-10