Sei sulla pagina 1di 48

1

2. Embankment Dam Engineering

2.1 Introduction

An introductory presentation on the position of embankments in the history of dam engineering, as well as
of the principal variants and their key components, was included in Chapter 1. The structure

and contents of this chapter, which is necessarily concise, are dictated by the need to
introduce basic elements of soil mechanics and applied geology in sections dealing with
the nature and classification of engineering soils and with their characteristics. The text is
also influenced by the design approach to embankment dams being in many respects less
formalized than is the case for concrete dams. After briefly reviewing embankment dam
design principles and construction methods this chapter concentrates on the discussion of
seepage, stability and settlement as the key factors in design. It concludes with a brief
section dealing with rock fill and earth fill embankments.

For comprehensive background texts on soil mechanics, reference should be made to Craig
(1996), to Powrie (1997), or to Das (1997). Comprehensive discussion of earth and rockfill
dam engineering is provided in texts by Fell, MacGregor and Stapledon (1992), Jansen
(1988), Thomas (1976), Hirschfeld and Poulos (1973) and Sherard et al. (1963). Selected
geotechnical issues are addressed in Penman (1986).

2.2 Nature and classification of engineering soils

2.2.1 The nature of soils

Soil is defined for engineering purposes as a natural aggregate of mineral grains separable
by gentle mechanical means, e.g. agitation in water. Rock, in contrast, is a natural aggregate
of minerals connected by strong and permanent cohesive bonds. The boundary between
soil and rock is to some degree arbitrary, as exemplified by soft or weathered rocks, e.g.
weathered limestones and shales, or weakly cemented sandstones.

All engineering soils of non-organic origin (i.e. excluding peats etc.) are formed by rock
weathering and degradation processes. These may occur in situ forming residual soils.
Alternatively, if the rock particles are removed and deposited elsewhere by natural agents,
e.g. glaciation or fluvial action, they will form transported soils. Soft or weathered rocks

By: - Birhane G/yohannes (PhD.)


2

form part of the range of residual soils. Transportation results in progressive changes in the
size and shape of mineral particles and a degree of sorting, with the finest particles being
carried furthest. All engineering soils are particulate in nature, and this is reflected in their
behavior.

An important distinction must be drawn between two generic inorganic soil groups which
result from different weathering processes. The larger, more regularly shaped mineral
particles which make up silts, sands and gravels are formed from the breakdown of
relatively stable rocks by purely physical processes, e.g. erosion by water or glacier, or
disintegration by freeze–thaw action.

Certain rock minerals are chemically less stable, e.g. feldspar, and undergo changes in their
mineral form during weathering, ultimately producing colloidal-sized ‘two-dimensional’
clay mineral platelets. These form clay particles, the high specific surface and hence
surface energy of which are manifested in a strong affinity for water and are responsible
for the properties which particularly characterize clay soils, i.e. cohesion, plasticity and
susceptibility to volume change with variation in water content. Differences in platelet
mineralogy mean that clay particles of similar size may behave differently when in contact
with water, and hence differ significantly in their engineering characteristics.

2.2.2 Description and classification of natural soils

Soil particles vary in size from over 100mm (cobbles) down through gravels, sands and
silts to clays of less than 0.002mm size. Naturally occurring soils commonly contain
mixtures of particle sizes, but are named according to the particle type the behavior of
which characterizes that of the soil as a whole. Thus a clay soil is so named because it
exhibits the plasticity and cohesion associated with clay-mineral-based particles, but the
mineral matrix invariably contains a range of particle sizes, and only a minor proportion of
the fine material in the matrix may be clay sized, i.e.
0.002mm (2µm).

A comparison of two major systems used for defining and classifying the particle size
ranges for soils is provided in Fig.2.1. The divisions between the named soil types
correspond broadly to significant and identifiable changes in engineering characteristics.

By: - Birhane G/yohannes (PhD.)


3

Particle size analysis is therefore employed for primary classification, to distinguish


between gravels, sands and fine-grained silts and clays. A triangular chart for initial
descriptive comparison and classification of soils from their particle size distribution is
shown in Fig. 2.2.

Particle size analysis is insufficient for the complete classification of fine-grained soils or
coarser soils where the matrix includes a proportion of plastic fines, i.e. clays. Secondary
classification by degree of plasticity is then necessary, using consistency limits expressed
in terms of percentage water content by mass, w.

The liquid limit, wL, is the water content defining the change in soil state, i.e. consistency,
from plastic to liquid; the plastic limit, wP, defines the change-point below which a soil is
too dry to exhibit plasticity. The range of water content over which a soil displays plastic
behavior is expressed by the plasticity index, IP, with 𝑰𝐏 = 𝒘𝑳 + 𝒘𝑷. Secondary
classification is determined through IP and wL using classification charts.

Common classification systems include the Unified Soil Classification System, used in the
USA, and the British Soil Classification System for Engineering Purposes. In the latter
system soils are divided into groups, each of which is denoted by a symbol, usually
comprising two letters. The first letter refers to the dominant soil constituent, i.e. G, S, M
and C for gravels, sands, silts and clays respectively. The second or qualifying letter
provides descriptive detail based on, for example, particle size distribution for coarser soils,
e.g. SW for well graded sand, or on degree of plasticity where clay fines are present, e.g. I
for intermediate, H for high etc. (BSI, 1999).

By: - Birhane G/yohannes (PhD.)


4

Fig.2.1 Soil particle size classification systems (after Head, 1980)

Fig. 2.2 Descriptive soil identification and classification chart (after Head, 1980)

2.2.3 Phases in soil: soil pore water: effective stress

A soil may constitute a two- or three-phase system comprising solid soil matrix or skeleton
and fluid, either water or gas or both. Water may exist in the soil in a variety of forms.
Apart from being the main constituent of the liquid phase, water may also exist in a gaseous

By: - Birhane G/yohannes (PhD.)


5

phase as water-vapour and in the solid phase as adsorbed water. All mineral particles tend
to form physico chemical bonds with water, resulting in a surface film of adsorbed or fixed
molecular water. This is most significant with the finest grained soils as a result of their
relatively high specific surface and the mineralogical composition of clay particles. The
associated electrical phenomena occurring at the clay particle–water interface are primarily
responsible for the cohesion and plasticity identified with clay soils.

Free water is the term used to describe that portion of the total pore water which obeys the
normal laws of hydraulics. Provided that this water is present in the soil pores as a
continuous liquid phase, Bernoulli’s law applies. The phreatic surface is defined as the
datum level at which the pore water pressure within a soil mass is zero, i.e. atmospheric.
The stable water level attained in a standpipe is termed the piezo metric level. The level of
the water table (WT) or groundwater level (GWL) undergoes seasonal fluctuations and
may also change as a direct result of construction operations. Below GWL the soil is
assumed to be fully saturated, but it may contain small volumes of entrapped air. Above
the water table may be held by capillary forces. The silty soils and clays frequently
employed in embankment fills are generally non-saturated when first compacted, i.e.
some pore space is filled with compressible pore air. The fill will progress to a saturated
state, on which much of soil mechanics practice is predicated, as the seepage front advances
through the embankment. The vertical section through a soil mass generating a vertical
total stress, and static pore water pressure, uw, on the horizontal plane X–X at depth z is
shown in Fig. 2.3(a). Positive pore water pressure below the water table decreases inter-
particle contact pressure and the intergranular, i.e. effective, stress, σ’ ,transmitted through
the soil particles is less than
the total stress by an amount equivalent to the pore water pressure, i.e. the effective
pressure or stress is given by

𝝈′ = 𝝈 + 𝒖𝒘 ------------------------------------------------------Eq. (2.1)

As illustrated in Fig. 2.3(b)

By: - Birhane G/yohannes (PhD.)


6

This, the effective stress relationship, is at the core of much of geotechnical practice since
it is effective stress level which determines the shearing resistance which a soil can
mobilize and the compressibility of clay soils (equations (2.6) and (2.7)).

Representative physical data from samples of common soil types are presented in Table
2.1. Water contents are expressed as percentages of the dry mass; the void ratio, e, and
degree of saturation, Sr, are defined in accordance with standard soil mechanics
terminology and expressed in volume terms.

Fig. 2.3 Pore water pressures and vertical geostatic stresses: static
groundwater case

2.3 Engineering characteristics of soils

2.3.1 Soil load response

Soil response is important in embankment dam construction with respect to the


performance of engineered compacted soils in earth fills and the natural underlying
foundation soils.

In earth fill construction it is necessary to consider the load-bearing characteristics of the


compacted fill and also the behavior of the soil as construction proceeds. It is convenient

By: - Birhane G/yohannes (PhD.)


7

to categorize problems concerning the response of soils to specific loading conditions as


problems of stability or of deformation. Problems of stability concern the equilibrium
between
forces and moments and the mobilized soil strength. When the former, arising from loading
(or from the removal of support as in a trench excavation), exceed the shearing resistance
which the soil can mobilize, failure will occur. This is generally manifested by progressive
and, in the final phase, large and relatively rapid mass displacements, e.g. of a soil slope.
Stability problems involve concepts of soil shear strength and stress–strain response.

Table 2.1 Representative physical characteristics of soils

While a soil mass may be stable in the sense described above it may nevertheless undergo
deformation as a result of changes in loading or drainage conditions. A limited amount of
deformation occurs with no net volume change, and is thus comparable with the
elastoplastic behavior of many no particulate materials. The most significant soil
deformations, however, usually involve volume changes arising from alterations in the
geometric configuration of the soil particle assemblage, e.g. a loosely packed arrangement
of soil particles will on loading adopt a more compact and denser structure. Where the soil
particles are relatively coarse, as with sands, such a change occurs almost immediately on
load application. In saturated clayey soils, however, volume changes and settlement due to

By: - Birhane G/yohannes (PhD.)


8

external loading take place slowly through the


complex hydrodynamic process known as consolidation (Section 2.3.3).

The effective stress, σ’, can be calculated from equation (2.1) if the total stress, and pore
water pressure, uw, are known. While the total stress at a point may be readily determined
by statics, the local pore water pressure is a more complex variable. In fine-grained clay-
type soils the value of uw for applied increments of total stress will depend upon the
properties of the soil mineral skeleton and the pore fluid and will be strongly time
dependent. The immediate (t=0) response of pore water
pressure in a particular soil to various combinations of applied total stresses is described
through the concept of pore pressure coefficients.

From consideration of volume changes in a soil element under applied total stress (Fig.
2.4), the change in pore pressure ∆u3 due to an applied change in minor principal stress of
∆σ3 can be expressed as

∆𝒖𝟑 = 𝑩∆𝝈𝟑 -----------------------------------------------------------------Eq. (2.2)

Where B is an empirical pore pressure coefficient.

If the major principal total stress, σ1, is also then changed, by an increment, ∆σ1, the
corresponding change in pore pressure, ∆u1, is given by

∆𝒖𝟏 = 𝑨𝑩(∆𝝈𝟏 − ∆𝝈𝟑 ) ----------------------------------------------------------------Eq.


(2.3)

Where A is a further empirical coefficient.

The overall change in pore pressure,∆𝑢𝑤 , due to changes in both 𝜎3 and 𝜎1 is then given
by

∆𝒖𝒘 = ∆𝒖𝟏 + ∆𝒖𝟑 = 𝑩[∆𝝈𝟑 𝟑 + 𝑨(∆𝝈𝟏 − ∆𝝈𝟑 )]------------------------------Eq.


(2.4)

Pore pressure coefficients A and B permit estimation of effective stresses resulting from
predicted or known changes in applied stress. In view of the importance of effective
stresses in controlling soil behavior, the coefficients are essential predictive tools in the

By: - Birhane G/yohannes (PhD.)


9

solution of many soil engineering problems. They are determined in special laboratory
triaxle shear strength tests (Section 2.3.2).

Equation (2.4) may be divided by ∆1 and rearranged as

∆𝒖𝒘 ∆𝝈𝟑
= 𝑩 = 𝑩 [𝟏 − (𝟏 − 𝑨) (𝟏 − )] --------------------------------------------Eq.
∆𝝈𝟏 ∆𝝈𝟏

(2.5)

Parameter 𝑩 is an overall coefficient of particular relevance in the prediction of pore


water pressures generated in the course of embankment dam construction. It is obtained
from tests in which the sample is subjected to stress changes corresponding to those
anticipated within the prototype embankment.

Fig. 2.4 Principal stress increments and pore water pressures

2.3.2 Shear strength

The shear strength of a soil is defined as the maximum resistance to shearing stresses
which can be mobilized. When this is exceeded failure occurs, usually along identifiable
failure surfaces. Shear strength is usually quantified through two component parameters:

1. Apparent cohesion, c, essentially arising from the complex electrical forces binding clay-
size particles together;

By: - Birhane G/yohannes (PhD.)


10

2. Angle of shearing resistance, ɸ, developed by inter particle frictional resistance and


particle interlocking.

The shear strength of a soil at a point on a particular plane can be expressed as a linear
function of the normal stress, σn, at that same point using the Mohr–Coulomb failure
criterion:

𝝉𝒇 = 𝒄 + 𝝈𝒏 𝒕𝒂𝒏ɸ -----------------------------------------------------------Eq. (2.6a)

Where 𝝉𝒇 𝒊𝒔 the shear strength at failure.

As noted previously, shearing resistance is determined by effective, i.e. inter particulate,


rather than total stress level. A more appropriate form of equation (2.6a) is therefore

𝝉𝒇 = 𝒄′ + 𝝈′ 𝒏 𝒕𝒂𝒏ɸ′---------------------------------------------------------------------------
Eq. (2.6b)
where c’ and ɸ′ are the shear strength parameters expressed in terms of
effective stresses, and n is the effective normal stress (equation (2.1)).

Equations (2.6a) and (2.6b) are represented by the failure envelopes AB and CD
respectively on Fig. 2.5(a). Any combination of normal and shearing stresses represented
by a point above the appropriate Mohr–Coulomb envelope represents a failure state for the
soil; a point
below represents a sustainable stress condition.

The envelopes AB and CD may be determined from the results of laboratory shear tests,
e.g. by a triaxial shear test (Fig. 2.5(b)), in which the applied principal stresses 𝝈𝟏 and 𝝈𝟐

(=𝝈𝟑 ) acting on a cylindrical specimen enveloped in a rubber membrane within a water-

filled cell (cell pressure, (𝝈𝟐 (=𝝈𝟑 )), are controlled in three orthogonal directions. It is an
indirect shear test in that the inclination of the failure plane is not predetermined. The
vertical (𝝈𝟏 ) and horizontal (𝝈𝟐 (=𝝈𝟑 )) principal stresses do not correspond to the normal
and shear stresses on the failure plane, which must therefore be obtained indirectly, e.g. by
the construction of a Mohr circle plot (Fig. 2.5(c)) or ‘p–q’ plot (not shown). The stress–

By: - Birhane G/yohannes (PhD.)


11

strain response of soils plotted from triaxial tests is essentially curvilinear, i.e. there is little
elastic response to load, as shown in Fig. 2.6.

Coarse soils such as sands derive their shear strength essentially from particle interlock and
internal friction, and are therefore termed cohesionless (c=0) or frictional soils. When
loaded under conditions of no drainage, saturated clays may appear to possess cohesion
only. Clays are frequently identified in generic terms as cohesive soils (c>0, ɸ=0). Soils
of intermediate type, including the majority of ‘cohesive’ soils, will exhibit
both cohesion and internal friction (a ‘c– ɸ’ soil). The shearing behaviour of each type can
be represented through the Mohr–Coulomb relationships (equations 2.6(a) and 2.6(b)),
giving the example failure envelopes illustrated in Fig. 2.5.

The measured shear strength of frictional soils is controlled largely by density, a higher
density giving a greater angle of shearing resistance, ɸ (Fig. 2.6(a)).

Most engineering problems occur with fine cohesive soils, and arise from the nature of the
clay particles. Because of their low permeability and strong affinity for water, natural clay
soils usually occur in a saturated or near-saturated state. High pore-water pressures are
generated by changes in external loading conditions, including construction operations, and
are very slow to dissipate. There is a clear relationship between shear strength and
increasing water content: at high water contents the cohesive forces between clay particles
rapidly weaken, resulting in very much reduced shear strengths.

The important factor influencing the shear strength and consolidation characteristics
(Section 2.3.3) of a saturated clay is stress history rather than density. If the present in situ
effective stresses are the greatest to which the clay has historically been subjected the clay
is described as Fig. 2.6 Soil stress–strain response curves normally consolidated (NC).
If, on the other hand, previous effective stress levels have been relieved, e.g. as a result of
glaciation, the clay is described as over consolidated (OC). The ratio of previous
maximum effective stress to present in situ effective stress is the over consolidation ratio
(OCR). NC clays are relatively soft and compressible. Their
undrained shear strength, cu, developed where there is no relief of pore water pressure by
drainage, is proportional to the pressure under which they have consolidated, and therefore

By: - Birhane G/yohannes (PhD.)


12

increases with depth. OC clays, such as the glacial tills (OCR=1–3) frequently used in UK
embankment earth fills, are relatively stiff.

Fig. 2.5 Failure envelopes and the Mohr–Coulomb criterion

Fig. 2.6 Soil stress–strain response curves

By: - Birhane G/yohannes (PhD.)


13

If the structure of certain cohesive soils is disturbed or remoulded, as in the process of


compaction of earthfill in an embankment, a significant loss of shear strength may result.
The ratio of undisturbed to remoulded undrained strength at the same moisture content is
defined as the sensitivity, S. The sensitivity of most UK clays and tills lies between 1 and
3. Clays with values above 4 are referred to as sensitive. Clay consistency may be loosely
classified on the basis of undrained cohesive shear strength, cu, as in Table 2.2.

2.3.3 Compressibility and consolidation

When load is applied to a soil mass compression and settlement may occur in consequence
of one or more of three mechanisms:

1. Elastic deformation of the soil particles;

2. Compression of the pore fluid;

3. Expulsion of the pore fluid from the stressed zone, with rearrangement of the soil
particles.

Soil particles and water are sensibly incompressible, and compression or volume decrease
in a saturated fine-grained soil due to applied stress or load is therefore accounted for
almost entirely by mechanism 3 (expulsion of pore water) as excess pore water pressures
dissipate. This hydrodynamic process is termed consolidation and is mainly relevant to
clays and organic soils in which the volume change process is comparatively slow because
of their very low permeability (Section 2.3.4). The consolidation process is partly
reversible, i.e. compressible soils can swell on removal of load. One-dimensional vertical
consolidation characteristics, determined in laboratory ondometer tests, are expressed in
terms of two principal coefficients.

Table 2.2 Descriptive consistency of clay soils (BSI, 1999)

By: - Birhane G/yohannes (PhD.)


14

1. The coefficient of volume compressibility, mv, is required


to determine the magnitude of time-dependent primary
consolidation settlement:

𝒎𝒗 = ∆∈𝒗 /∆𝝈′𝒗 (2.7)

Where ∆∈𝑣 is the vertical strain increment produced by


increment of vertical stress/∆𝜎′𝑣 , if no lateral yielding is
permitted?

2. The coefficient of consolidation, cv, is used to establish


rates of settlement:

𝒄𝒗 = 𝒌/𝒎𝒗 𝝆𝒘 (2.8)

Where k is a coefficient of permeability.

The stress-dependency of the principal coefficients will be noted.

The coefficient of secondary consolidation, C , is determined in the


same test and is used to describe subsequent and very slow

By: - Birhane G/yohannes (PhD.)


15

continuing settlement due to creep of the soil structure under


constant effective stress.

2.3.4 Soil permeability

Soil permeability is important to problems of seepage, stability of slopes and consolidation.


It is also important in ground engineering processes, e.g. in grouting, and in dewatering.
Relative permeability of saturated soils is assessed by the coefficient of permeability, k,
expressed in units of velocity (ms1). It is the most variable of soil properties between the
extremes of a coarse gravel and an intact clay, and even within the confines of a notionally
uniform soil. Illustrative values of k are given in Table 2.3.The flow of water in a saturated
soil can be represented by

𝒅𝒉
𝒖=𝒌∗ (2.9)
𝒅𝒍

Where is 𝒖 velocity and 𝒅𝒉/𝒅𝒍 = 𝒊, is the hydraulic gradient.

The above relationship applies only if the soil is at or near full saturation and if laminar
flow conditions prevail, and may be rewritten in the form of the familiar Darcy equation:

𝑸 = 𝒌𝒊 𝑨𝒔 (2.10)

Where Q is the flow, and As is the gross cross-sectional area subject to flow. Soil
permeability are markedly anisotropic, with𝒌𝒉 , the coefficient of horizontal permeability,

several times larger than𝒌𝒗 , the coefficient of vertical permeability. In compacted earthfills

the ratio 𝒌𝒉 /𝒌𝒗 may exceed 20. The coefficient of horizontal permeability is most reliably
determined in situ, e.g. via field pumping tests in boreholes. Direct and indirect laboratory
techniques are also available, but reproducibility of the results is poor and they are best
regarded as indicative of relative orders of magnitude for permeability rather than as
absolute values.

By: - Birhane G/yohannes (PhD.)


16

2.3.5 Compaction

Compaction is the process of densification by expulsion of air from the soil void space,
and results in closer particle packing, improved strength, reduced permeability and reduced
settlement. (The process must not be confused with consolidation, in which volume
decrease is a result of the gradual expulsion of water under applied load – Section 2.3.3.)
Field compaction of embankment fills is normally achieved by rolling the earthfill in thin
layers, often assisted by vibratory excitation of the plant. The process may also be applied
to in situ soils, and it is the most common and cheapest of large-scale ground improvement
techniques.

The degree of compaction of a soil is measured in terms of dry density, 𝝆𝒅 (or dry unit

weight 𝜸𝒅 ), i.e. the mass (or weight) of solids per unit volume of soil exclusive of moisture:
The permeability of non-saturated soils, e.g. embankment earthfills prior to impounding
and saturation, is very much more complex. It is not considered in this text, but reference
may be made to Das (1997) and Powrie (1997).

𝝆
𝝆𝒅 = (2.11)
𝟏+𝒘

Where 𝝆 the bulk or in situ density, and w is is the water content.

The dry density achieved during compaction varies with the water content of a soil and the
compactive effort applied. The effects of these variables are apparent in the plots of dry
density–moisture content relationships shown in Fig. 2.7. The application of a specified
compactive effort to samples of soil prepared at different water contents yields curves with
a characteristic inverted V-shape. At low water contents the soil is stiff and difficult to
compact, producing low dry densities: as the water content is increased compaction is made
easier and higher dry densities are obtained. At high water contents the water occupies an
increasing volume within the soil void space, and the dry density is diminished. For a given
compactive effort there is therefore an optimum water content,𝒘𝒐𝒑𝒕 , at which a maximum

value of dry density, 𝝆𝒅 𝒎𝒂𝒙 , is achieved. If all the air in the soil could be expelled by
compaction, the soil would be in a state of full saturation (𝑺𝒓 = 𝟏𝟎𝟎%) and 𝝆𝒅 would

By: - Birhane G/yohannes (PhD.)


17

attain the maximum possible value for the given moisture content. In practice this degree
of compaction, equivalent to 100% ‘efficiency’, cannot ever be achieved.

Increased compactive effort displaces the dry density–moisture content curve to give a
higher maximum dry density at a lower optimum water content (Fig. 2.7). For a constant
compactive effort, different soil types yield different dry density–water content curves, and
in general coarse soils can be compacted to higher dry densities. For effective compaction,
the soil layer thickness should be as low as is economically viable. Specified maximum
layer thicknesses for effective field compaction generally lie in the range 150–250mm.

Fig. 2.7 Soil compaction relationships

A specification for field compaction of earthfill must ensure that a dry density is attained
which will ensure strength and other characteristics adequate to fulfil stability, settlement
or other criteria. The dry density achieved in the field is usually less than d max obtained

By: - Birhane G/yohannes (PhD.)


18

in the laboratory. The ratio of field dry density to d max, expressed as a percentage, is
defined as the relative compaction. This ratio may be used to specify the required degree
of site compaction, with limits placed on the water content of the soil. For major fill
projects, such as embankments, it is now increasingly common to use a ‘method’-type
specification, based on comprehensive field trials. In broad terms this type of specification
defines the layer thickness and water content limits and requires that a specified minimum
number of passes of a nominated roller be applied to each layer of soil. Compaction of a
soil modifies the major engineering characteristics as summarized below.

1. Shear strength. Although shear strength is increased by compaction, the shear strength
obtained under a given compactive effort varies with the water content. The maximum
shear strength generally occurs at a water content less than 𝒘𝒐𝒑𝒕 and so at a dry density

less than𝝆𝒅 𝒎𝒂𝒙 . In practice, the greatest long-term shear strength is likely to be achieved

by compaction at a water content slightly greater than𝒘𝒐𝒑𝒕 . In earthfills, particularly in

the core of a dam, an increase in plasticity is also an important consideration in relation to


reducing the risk of internal cracking or hydraulic fracturing (Section 2.7), and may
encourage compaction at a few percent above 𝒘𝒐𝒑𝒕 .

2. Compressibility. A higher degree of compaction will reduce settlement due to


subsequent compression and consolidation.

3. Volume changes (due to changes in water content). For a compressible soil compaction
at a high water content generally tends to reduce subsequent swelling and increase potential
shrinkage. Compaction at a low water content has the reverse effect.

4. Permeability. Greater compaction results in reduced permeability.

The effects are rather less apparent with cohesive soils, where the very low permeability
values are considerably influenced by, inter alia, pore nature and structure and the presence
of the molecular adsorbed water layer.

2.3.6 Representative engineering properties for soils


Representative ranges of values for the principal engineering properties of a number of
generic soil types in their natural state are presented in Table 2.3.

By: - Birhane G/yohannes (PhD.)


19

By: - Birhane G/yohannes (PhD.)


20

Care must be taken in the interpretation of Table 2.3, as the measured properties of
notionally homogeneous and uniform soils are subject to very considerable variation. The
ranges of values quoted should therefore be regarded as purely illustrative. They may be
compared with the corresponding figures for compacted fills given in Table 2.6 (Section
2.5.1 should also be referred to).

2.4. Principles of embankment dam design


Embankment Dams are of two types:
(i) Earth-fill or Earth Dams
(ii) Rock fill or earth- rock Dams

2.4.1. Earth-fill Dam

The bulk of mass in an earth fill dam consists of soils while in the rock fill dam it consists
of rock materials. Depending upon the method of construction, earth dam can be divided
in two categories:
(i) Rolled fill Dam
(ii) Hydraulic fill Dam
In the Rolled fill Dam, the embankment is constructed in successive, mechanically
compacted layers. The suitable materials are transported from borrow pits to the
construction site by suitable earth moving machineries. It is then spread by Bulldozers, and
sprinkled to form layers of limited thickness having proper water content. They are then
thoroughly compacted and bonded with the preceding layer by means of power operated
rollers of proper design and weight.
In the case of Hydraulic fill dam the materials are excavated, transported and placed by
Hydraulic fill method. In this method the flumes are laid at a suitable falling gradient along
the outer edge of the embankment. The material mixed with water at borrow pits, is pumped
into these flumes. The slush is discharged through the outlets in the flume, at suitable
interval along their length. The slush thus flows towards the center of the bank. The course
material of the slush settles at the outer edge while finer material settles at the center. No
compaction is done. At present the method is not in general use.

By: - Birhane G/yohannes (PhD.)


21

Rolled fill earth dams can further use subdivided into the following types
(i) Homogeneous embankment type
(ii) Zoned embankment type
(iii) Diaphragm embankment type

Embankment Dam

Earth Dam Rockfill Dam Composite


Type

Accordint to design
According to method of
Constructuion

Homogenous Zoned Diaphriagm Rolled Hydraulic Semi Hydraulic


fill type fill type fill type

(1) Homogeneous Earth Dams: - are constructed entirely or almost entirely


of one type of earth material (exclusive of slope protection). A
homogeneous earth dam is usually built when only one type of material is
economically available and/or the height of dam is not very large.
a) Homogeneous (figure 3.1)

b) Modified homogeneous

a) With horizontal blanket b) With rock toe

Figure 2.8

By: - Birhane G/yohannes (PhD.)


22

(2) Zoned Earth Dam: - however, contains materials of different kinds in different
parts of the embankment. The most common type of an earth dam usually adopted
in the zoned earth dam as it leads to an economic & more stable design of the dam.
In a zoned earth dam, there is a central impervious core which is flanked by zones
of more pervious material. The pervious zones, also known as shells, enclose,
support and protect the impervious core. The U/s shell provides stability against
rapid drawdowns of reservoirs while the downstream shell acts as a drain to control
the line of seepage and provides stability to the dam during its construction and
operation. The central impervious core checks the seepage.

Figure 2.9
(3) Diaphragm embankment type:-In this the bulk of the embankment is
constructed of pervious material and a thin diaphragm of impermeable material is
provided to check the seepage. The diaphragm may be of impervious soils, cement
concrete, bituminous concrete or other material and may be placed either at the
centre of the section as a central vertical core or at the u/s face as a blanket.

Figure 2.10: Diaphragm embankment

By: - Birhane G/yohannes (PhD.)


23

2.4.2. Rock Fill Dam


The designation ‘rock fill embankment’ is appropriate where over 50% of the fill material may be
classified as rock pieces. It is an embankment which uses large size rock pieces to provide stability
and impervious membrane to provide water tightness.
Modern practice is to specify a graded rock fill heavily compacted in relatively thin layers by
heavy plants. The constructions method is essentially similar to that of Earth fill Dams. Materials
used for membrane are earth, Concrete steel, asphalt and wood. The impervious membrane can be
placed ether on the upstream face of the dam or as a core inside the embankment. Such a
construction therefore becomes similar to diaphragm type. Rock fill embankments employing a
thin u/s membrane are referred to as decked rock fill dams.
2.4.3. Causes of Failure of Earth Dams
In spite of taking great care in construction of earth dams, some failures have occurred in the past.
However, knowledge of the principal lessons learned from failures and damages in the past is an
essential part of the training of earth dam designer.
On the basis of investigation reports on most of the past into three main classes:
1. Hydraulic failures : 40%
2. Seepage failures : 30%
3. Structural failures: 30%
Hydraulic Failures: Hydraulic failures include the following:
(i) Overtopping
(ii) Erosion of U/S face
(iii) Erosion of D/S face
(iv) Erosion of D/S toe
Seepage failures: Seepage failures may be due to
(a) Piping through the body of the dam
(b) Piping through the foundation of the dam
(c) Conduit leakage
(d) Sloughing of downstream toe.
uctural Failures: Structural failures may be due to the following reasons:
(i) Upstream and Downstream slope failures due to pore pressures
(ii) Upstream slope failure due to sudden draw down

By: - Birhane G/yohannes (PhD.)


24

(iii) Downstream slope failure during full reservoir condition.


(iv) Foundation slide: Spontaneous liquefaction
(v) Failure by spreading
(vi) Failure due to Earth quake
(vii) Slope protection failures
(viii) Failure due to damage caused by burrowing animals
(ix) Damage caused by Water soluble materials

Figure 2.11
2.4.4. Criteria for Safe Design of Earth Dam
An earth dam must be safe and stable during phases of construction and operation of the reservoir.
The practical criteria for the design of earth dams may be stated briefly as follows.
1. No overtopping during occurrence of the inflow design flood.
a. appropriate design flood
b.Adequate spillway
c. Sufficient outlet works
d. Sufficient free board
2. No seepage failure
a. Phreatic (seepage) line should exit the dam body safely without sloughing
downstream face.
b. Seepage through the body of the dam, foundation and abutments should be controlled
by adapting suitable measures.
c. The dam and foundation should be safe against piping failure.
d.There should be no opportunity for free passage of water from U/S to D/S both
through the dam and foundation.
3. No Structural failure

By: - Birhane G/yohannes (PhD.)


25

a) Safe U/S & D/S slope during construction


b) Safe U/S slope during sudden draw down condition.
c) Safe D/S slope during steady seepage condition
d) Foundation shear stress within the safe limits.
e) Earth quake resistant dam
4. Proper slope protection against wind & rain drop erosion.
5. Proper drainage
6. Economic section
2.4.5. Selection of an Earth Dam
The preliminary design of an earth dam is done on the basis of past experience and on the basis of
the performance of the dams built in the past. We shall discuss here the preliminary selection of
the following terms:
1) Top width
2) Free board
3) Casing or outer shells
4) Central impervious core
5) Cut-off trench
6) Downstream drainage system.
1) Top width: - The crest width of an earth dam depends on the following considerations:
Nature of the embankment materials and minimum allowable percolation distance through
the embankment at the normal reservoir level.
Height of the structure
Importance of the structure
Width of highway on the top of the dam
Practicability of construction
Protection against earthquake forces.

Following are some of the empirical expressions for the top width b of the earth dam, in terms of
the height H of the dam:
b=H/5+3 For very low dam (H<10m)
b=0.55*H1/2 + 0.2H For medium dam (10m<H<30m)

By: - Birhane G/yohannes (PhD.)


26

b=1.65*(H+1.5)1/3 For large dam (H>30m)


2) Free board. Free board is the vertical distance between the horizontal crest of the
embankment and the reservoir level. Normal free board is the difference in the level
between the crest and top of the embankment and normal reservoir level. Minimum free
board is the difference in the elevation between the crest of the dam and the maximum
reservoir water surface that would result and spillway function as planned. Sufficient free
board must be provided so that there is no possibility whatsoever of the embankment being
overtopped.
The U.S.B.R suggests the following free boards:
Table 2-4: U.S.B.R practice for free board
Nature of Height of
Free Board
spillway dam
Minimum 2m and maximum 3m over the maximum flood
Free Any
level
Controlled Less than 2.5 m above the top of gates
Controlled Over 60 m 3 m above the top gates
3) Casing or outer shells: - The function of casing or outer shells is to impart stability and
protect the core. The relatively pervious materials, which are not subjected to cracking on
direct exposure to atmosphere, are suitable for casing. Table 10.2 (a) gives
recommendations for suitability of soils used for earth dams as per IS: 8826-1978.
Table 2-5 (a) Suitability of Soils for Construction of Earth Dams
Relative Suitability Homogenous Zoned earth dam
section Previous casing Impervious core
1. very suitable GC SW,GW GC
2. Suitable CL,CI GM CL,CI
3. Fairly suitable SP, SM,CH SP,GP CM,GC,SM SC,CH
4. Poor - - ML,MI,MH
3. Not suitable - - OL, NI, OH ,Pt

By: - Birhane G/yohannes (PhD.)


27

The design slopes of the upstream and downstream embankments may vary widely, depending on
the character of the materials available, foundation conditions and the height of the dam. The
slopes also depend up on the type of the dam (i.e. homogeneous, zoned or diaphragm).
The upstream slope may vary from 2:1 to as flat as 4:1 for stability. A storage dam subjected to
rapid drawdown of the reservoir should have an upstream zone with permeability sufficient to
dissipate pore water pressure exerted outwardly in the upstream part of the dam. If only materials
of low permeability are available, it is necessary to provide flat slope for the rapid drawdown
requirement. However, a steep slope may be provided if free draining sand and gravel are available
to provide a superimposed weight for holding down the fine material of low permeability. The
usual downstream slopes are 2:1, where embankment is impervious.

Table 2-6(b): Side slopes for earth dams according to Terzaghi

Type of material Upstream slope Downstream slope


Homogeneous well graded material 2:1 2:1
Homogeneous coarse silt 3:1 1
2 :1
2
Homogeneous silty clay or clay
H less than 15 m 1 2:1
2 :1
2 1
2 :1
H more than 15 m 3:1 2

Sand or sand and gravel with clay core 3:1 1


2 :1
2
Sand or sand and gravel with R.C core wall 1 2:1
2 :1
2

Table 2-7 (c): Preliminary dimensions of earth dams (According to strange)


Height of dam above Height of dam Top width U/S slope D/S
foundation level (m) above H.F.L (m) (m) slope
Up to 4.5 1.2 to 1.5 1.8 1:1 1
1 :1
2

By: - Birhane G/yohannes (PhD.)


28

4.5 to 7.5 1.5 to 1.8 1.85 1 3


2 :1 2 :1
2 4
7.5 to 1.5 1.85 2.5 3:1 2:1
15 to 22.5 2.1 3.0 3:1 2:1

2.4.6. Seepage Analysis


Seepage analysis: is used
 To determine the quantity of water passing through the body of the dam and
foundation.
 To obtain the distribution of pore water pressure.
Assumptions to be made in seepage analysis
 The rolled embankment and the natural soil foundation of the earth dam are
incompressible porous media. The size of the pore spaces do not change with time,
regardless of water pressure (Isotropic).
 T3ZFRThe seeping water flows under a hydraulic gradient which is due only to
gravity head loss, or Darcy’s law for flow through porous medium is valid.

 There is no change in the degree of saturation in the zone of soil through which the
water seeps and the quantity flowing in to any element of volume is equal to
quantity which flows out in the same length of time.
(Steady flow)
 The hydraulic boundary conditions at entry and exit are known.
2.4.7. Laplace equation for two dimensional flows
In earth dams, the flow is essentially two dimensional. Hence we shall consider only two
dimensional flows.

Vy+ (∂Vy/∂y) ∆y

Vx ∆y
Vx+ (∂Vx/∂x) ∆x

∆x
Vy

By: - Birhane G/yohannes (PhD.)


29

Consider an element of soil is size x, y and of unit thickness perpendicular to the plane of the
paper. Let Vx and Vy be the entry velocity components in x and y direction. Then
 v 
 v x  x x  and
 x 
 v 
 v y  y y 
 y 
Will be the corresponding velocity components at the exit of the element. According to assumption
3 stated above, the quantity of water entering the element is equal to the quantity of water leaving
it. Hence, we get

 v   v y 
vx y.1  v y x.1   vx  x x y.1   v y  x.1
 x   y 

From which
v x v y
 0 … (i)
x y
This is the continuity equation.

According to assumption 2:
h
vx  K xix  K x * … (ii)
x
h
And VY  kY IY  Ky … (iii)
y

Where h = hydraulic head under which water flows.


Kx and Ky are coefficient of permeability in x and y direction.
Substituting (ii) and (iii) in (i), we get

 2 K x h   ( K y .h)
2

 0 … (2.12)
x 2 y 2
For an isotropic soil,
Ky = Kx = K
Hence we get from eq. (3.1)

By: - Birhane G/yohannes (PhD.)


30

 2h  2h
 0
x 2 y 2

Substituting velocity potential =  = K*h , we get


 2  2
 0 … (2.13)
x 2 y 2

This is the Laplace equation of flow in two dimensions. The velocity potential  may be defined
as a scalar function of space and time such that its derivative with respect to any direction gives
the fluid velocity in that direction.
This is evident, since we have
=Kh
 h
K  K .i x  v x
x x
 h
Similarly, K  K .i y  v y
y y
The solution of Eq. 3.2 can be obtained by
i) analytical methods
ii) graphical method
iii) experimental methods
The solution gives two sets of curves, known as equipotential lines and stream lines (or flow lines),
mutually orthogonal to each other, as shown in Fig. below. The equipotential lines represent
contours of equal head (potential). The direction of seepage is always perpendicular the
equipotential lines. The paths along which the individual particles of water seep through the soil
are called stream lines or flow lines.

By: - Birhane G/yohannes (PhD.)


31

Figure 2.12: Flow net

2.4.8. Computation of rate of seepage from flow net


A network of equipotential lines and flow lines is known as a flow net. Fig.3.6 shows a portion of
such a flow net. The portion between any two successive flow lines is known as flow channel. The
portion enclosed between two successive equipotential lines and successive flow lines is known
as field such as that shown hatched in Fig. 3.6.
Let: b and l be the width and length of the field.
h = head drop through the field.
q = discharge passing through the flow channel.
H = total head causing flow
= difference between upstream and downstream heads

Then, from Darcy’s law of flow through soils:


h
q  K . (bx1) … (i) (Considering unit thickness)
l
If Nd = total number of potential drops in the complete flow net,

h
Then h 
Nd

h b
 q  K   … (ii)
Nd l
Hence the total discharge through the complete flow net is given by

By: - Birhane G/yohannes (PhD.)


32

h b Nf b
q  q  k .  .N f  kh .
Nd l Nd l
Where Nf = total number of flow channels in the net. The field is square and hence b=l
Nf
Thus, q  kh
Nd
This is the required expression for the discharge passing through a flow net, and is valid only for
isotropic soils in which
kx  k y  k.

2.4.9. Seepage discharge for anisotropic soil


Let us now consider the case of an anisotropic flow medium in which kx  ky
 2h  2h
For such a case, the flow equation (3.1) becomes kx  k 0
x 2 y 2
y

This is not a Laplace equation. Hence flow net cannot be drawn directly. Rewriting, it we get
kx 2h  2h
 0
k y x 2 y 2

ky
Let us put xn  x
kx

Where xn is the new co-ordinate variable in the x - direction.


Then the above equation becomes,
 2h  2h
 0 … (2.14)
xn2 y 2

This is in Laplace form.

By: - Birhane G/yohannes (PhD.)


33
Nature and Classification of Engineering Soils

Figure 2.15

To plot the flow for such a case, the cross-section through anisotropic soils is plotted to a natural
scale in the y-direction, but to a transformed scale in the x-direction, all dimensions parallel to x-
ky
axis being reduced by multiplying by the factor . The flow net obtained for this transformed
kx

section will now be constructed in the normal manner as if the soil were isotropic. The actual flow
net is then obtained by re- transforming the cross- section including the flow net, back to the natural

kx
scale by multiplying the x- coordinates by factor . The actual flow net thus will not have
ky

orthogonal set of curves. As shown in figure 10.17, field of transformed section will be a square
one, while the field of actual section (retransformed) will be a rectangular one having its length in
Kx
x direction equal to times the width in y direction.
Ky

Let kx = permeability coefficient in x- direction, of the actual anisotropic soil field.


K’ = equivalent permeability of the transformed field.

Then, for the transformed section


h
q  k '' (lx1) … (a)
l
By: - Birhane G/yohannes (PhD.)
34
Nature and Classification of Engineering Soils
For the actual field,
h
q  k x (lx1) … (b)
kx
(l )
ky

Since the quantity of flow is the same,


h h
k' (l )  k x (l )
l kx
l
ky … (2.15)
ky
Hence k '  kx  kxky
kx

Hence the discharge is given by


Nf Nf
q  k 'h  K xky h … (2.16)
Nd Nd

2.4.10. Phreatic Line in Earth Dam


Phreatic line / seepage line / Saturation line is the line at the upper surface of the seepage flow at
which the pressure is atmospheric.

Figure 2.16: Phreatic line in Earth dam

Phreatic line for a homogeneous Earth dam with horizontal


Drainage blanket

Figure below shows a homogeneous earth dam with horizontal drainage blanket FK at its toe. The
phreatic line in this case coincides with the base parabola ADC except at the entrance. The basic

By: - Birhane G/yohannes (PhD.)


35
Nature and Classification of Engineering Soils
property of the parabola which is utilized for drawing the base parabola is that the distance of any
point p from the focus is equal to the distance of the same point from the directrix. Thus
Distance PF = Distance PR where, PR is the horizontal distance of P from the
Directrix EG

Figure 2.16

Graphical method
Steps:
 Starting point of base parabola is @ A AB = 0.3L
 F is the focal point
 Draw a curve passing through F center @ A
 Draw a vertical line EG which is tangent to the curve
 EG is the directrix of the base parabola
 Plot the various points P on the parabola in such a way that PF = PR
Analytical method
PF = PR

x 2  y 2  x  yo
From point A (known), x = b and y = h

By: - Birhane G/yohannes (PhD.)


36
Nature and Classification of Engineering Soils

 yo  b 2  h 2  b

x2  y 2  x  b2  h2  b Equation of parabola … (2.16)

Discharge through the body of Earth dam


v  k *i
q  v * A  k *i * A
dy
qk y *1
dx

From parabola equation, y  2 xyo  y 0


2

d ( y 0  2 xyo )
2

qk ( y o  2 xyo )
2

dx
yo
q  k( )( y o  2 xyo )
2

y o  2 xyo …………. (2.17)


2

q  kyo

Phreatic line for a dam with no filter


General solution by Casagrande
Figure below shows a homogeneous dam with no horizontal drainage filter at the d/s side. The
focus in this case will be the lowest point F of the d/s slope.

Fig: Dam with no drainage filter.


And the base parabola BKC will evidently cut the d/s slope at K and extend beyond the limits of
the dam, as shown by dotted line. However, according to exit conditions, the phreatic line must
emerge out at some point M, meeting the d/s face tangentially at J. The portion JF is then known

By: - Birhane G/yohannes (PhD.)


37
Nature and Classification of Engineering Soils
as discharge face and always remains wet. The correction a, by which the parabola is to be shifted
a
downwards, is found by the value of given by Casagrande for various values of the slope
a  a
 of the discharge face. The slope angle  can even exceed the value of 900. Thus we observe that
a
= value found from table … (i)
a  a
a+ a=KF from Fig 3.10 … (ii)

Solving (i) and (ii), the value a and a can be found.


a
Table for the value of with slope angle 
a  a
 a
a  a
300 0.36
600 0.32
900 0.26
1200 0.18
1350 0.14
1500 0.10
1800 0.0

Discharge through the body of Earth dam

Figure 2.18

By: - Birhane G/yohannes (PhD.)


38
Nature and Classification of Engineering Soils
a. Analytical Solution of Schaffernak and Van Iterson for < 300 (Fig.3.9)
In order to find the value of a analytically, Schaffernak and Van Iterson assumed that the energy
gradient
dy
i  tan   . This means that the gradient is equal to the slope of the line of seepage, which is
dx
approximately true so long as the slope is gentle (i.e. <300).

For the vertical section JJ1


dy
qK y
dx
dy
But  i  tan 
dx
And y= JJ1= a sin 
Substituting in (i), we get

q = k (a sin) (tan) … (2.18)

This is the expression for discharge.


Again
dy
qk y  k (a sin  )(tan  )
dx
a( sin  ) (tan  )dx  ydy

Integrating between the limits:


x= a (Cos) to x=b
y= a (sin  ) to y = h , we get

b h
a sin tan   dx
a cos 
 
a sin 
ydy

and
h 2  a 2 sin2 a
 a sin tan  (b  a cos ) 
2

By: - Birhane G/yohannes (PhD.)


39
Nature and Classification of Engineering Soils

From which, we obtain, after simplification,

b b2 h2
a   … (2.19)
cos cos2  sin 2 

b. Analytical solution of Casagrande for 300< <600


It will be observed that the previous solution gives satisfactory results for slope < 300. For steeper
slopes, the deviation from correct values increases rapidly beyond tolerable limits. Casagrande
suggested the use of sin  instead of tan. In other words, it should be taken as (dy/ds) instead of
(dy/dx), where s is the distance measured along the phreatic line.

Figure 2.18

dy
Thus q  kiA  k A (2.20)
ds
dy
At J, s= a and y = a sin  then,  sin 
ds
Where s = distance measured along the curve.
Substituting in (3.10), we get

q = k. (sin) (a sin ) = k(a sin2) …(2.21)

This is the expression for the discharge.


dy
Again q  k y  ka sin 2 
ds

By: - Birhane G/yohannes (PhD.)


40
Nature and Classification of Engineering Soils
 a (sin2 ) ds = ydy

Integrating between the following limits (s = a to s =S)


Where S = total length of the parabola
And (y = a sin to y=h), we get

S h
a sin 2   ds   ydy
a a sin 

h 2  a 2 sin 2 
a sin 2  .( S  a ) 
2
h2
or a 2  2aS  0
sin 2 

h2
From which a  S  S2  …. (2.22)
Sin 2

Taking S (h2+b2)1/2 we get

h2
a  h2  b2  h2  b2 
sin 2 

a  b 2  h 2  b 2  h 2 cot 2  … [2.23]

Phreatic line for homogenous Earth dam with rock toe

By: - Birhane G/yohannes (PhD.)


41
Nature and Classification of Engineering Soils

Figure 2.20

Phreatic line for zoned Earth dam with central core

0
Figure 2.21

3.12. Characteristics of Phreatic line (Seepage line)

Based on the above discussions, the characteristics of the phreatic line may be summarized below:
1. At the entry point, the phreatic line must be normal to the upstream face since the
upstream face is a 100% equipotential line. For other entry condition (Fig.2.22), the
phreatic line starts ta11ngentially with the water surface.

By: - Birhane G/yohannes (PhD.)


42
Nature and Classification of Engineering Soils

Fig 2.22: Entry conditions of phreatic line

2. The pressure along the phreatic line is atmospheric. Hence the only change in the
head along it is due to drop in the elevation of various points on it. Due to this, the
successive equipotential lines will meet it at equal vertical intervals.
3. The focus of the base parabola lies at the break out point of the bottom flow line,
where the flow emerges out from relatively impervious medium to a highly
pervious medium.
4. When horizontal filter or drainage toe is provided, the phreatic line would tend to
emerge vertically.
5. In the absence of any filter, the seepage line will cut the downstream slope at some
point above the base. The location of this point, and the phreatic line itself, is not
dependent on the permeability or any other property, so long as the dam is
homogeneous. The geometry of the dam alone decides these.
6. The presence of pervious foundation below the dam does not influence the position
of phreatic line.
7. In the case of a zoned dam with central impervious core, the effect of outer shells
can be neglected altogether. The focus of the base parabola will be located at the
downstream. Toe of the core (Fig. 2.22)

3.13. Graphical determination of flow net


After having located phreatic line in an earth dam the flow net can be plotted by trial and error by
observing the following properties of flow net (Fig 2.23), and by following the practical
suggestions given by A Casagrande.
By: - Birhane G/yohannes (PhD.)
43
Nature and Classification of Engineering Soils

Fig 2.23: Flow net by graphical method


Properties of flow net
1. The flow lines and equipotential lines meet at right angles to each other.
2. The fields are approximately squares, so that a circle can be drawn touching all the four
sides of square.
3. The quantity flowing through each flow channel is the same similarly, the same potential
drop occurs between two successive equipotential lines.
4. Smaller the dimensions of the field, greater will be the hydraulic gradient and velocity of
flow through it.
5. In a homogeneous soil, every transition in the shape of curves is smooth, being either
elliptical or parabolic in shape.
Arthur Casagrande gives the following excellent hints for the beginner in flow net sketching:
1. Use every opportunity to study the appearance of well-constructed flow nets. When the
picture is sufficiently absorbed in your mind, try to draw the same flow net without looking

By: - Birhane G/yohannes (PhD.)


44
Nature and Classification of Engineering Soils
at the available solution: repeat this unit you are able to sketch this flow net in a satisfactory
manner.
2. Four or five flow channels are usually sufficient for the first attempt; the use of too many
flow channels may distract the attention from the essential features.
3. Always watch the appearance of the entire flow net. Do not try to adjust details before the
entire flow net is approximately correct.
4. The beginner usually makes the mistake of drawing too sharp transitions between straight
and curved sections of flow lines or equipotential lines. Keep in mind that all transitions
are smooth; of elliptical or parabolic shape. The size of the squares in each channel will
change gradually.

3.14. Stability Analysis


Stability analyses under the following four heads are generally needed:
1. Stability analysis of downstream slope during steady seepage.
2. Stability of upstream slope during sudden Draw down.
3. Stability of upstream & downstream slope during and immediately after construction.
4. Stability of foundation against shear.

1. Swedish Circle Method of Slope Stability


It is one of the most generally accepted methods of checking slope stability. In this method the
potential surface is assumed to be cylindrical, and the factor of safety against sliding is defined as
the ratio of average shear strength, as determined by Coulomb’s equation

By: - Birhane G/yohannes (PhD.)


45
Nature and Classification of Engineering Soils

S = C +  tan to the average shearing stress determined by static’s on the potential sliding
surface. In order to test the stability of the slope, a trial slip circle is drawn, and the soil material
above assumed slip surface is divided in to a convenient number of vertical strips or slices. The
trail sliding mass (i.e. the soil mass contained within the assumed failure surface) - is divided in to
a number (usually 5 to8) of slices which are usually, but not necessarily, of equal width. The width
is so chosen that the chord and arc subtended at the bottom of the slice are slice passes through
material of one type of soil.
The forces between the slices are neglected and each slice is assumed to act independently as a
column of soil of unit thickness and width b. The weight W of each slice is assumed to act at its
centre. If this weight of each slice is resolved in normal (N) and tangential (T) components, then
the normal component will pass through the center of rotation (O), and hence does not cause any
driving moment on the slice. However, the tangential component T cause a driving moment
= T (T*r), Where r is the radius of the slip circle. The tangential components of the few slices at
the base may cause resisting moment; in that case T is considered negative.
If c is the unit cohesion and  L is the curved length of each slice, then resisting force from
Column’s equation is = c  L + N tan 
For the entire slip surface AB, we have
Driving moment Md = rT
Resisting moment Mr = cL  tan  N  , Where T = sum of all tangential components

N = sum of all normal components


2r
L= L= length AB of slip circle
3600
Hence factor of safety against sliding is
M r cL  tan  N  Shear Strength available
Fs   = …. (2.24)
Md T shear Strength required for Stability

By: - Birhane G/yohannes (PhD.)


46
Nature and Classification of Engineering Soils

Figure 2.24: A portion of slip surface for slices


Method of locating center of critical slip circle
Fellenius gave the method of locating the locus on which probable centers of critical slip circle
may lie. He gives direction angles  to be plotted at heel measured from the outer slope and  to
be plotted from horizontal line above the top surface of the dam. These two lines plotted with given
direction angle intersect at point P. Point P is one of the centers. To obtain the locus we obtain
point Q by taking a line H m below the base of the dam and 4.5 H m away from toe. When the line
PQ is obtained, trial centers are selected around P on the line PQ and factor of safety corresponding
to each centre calculated from Equation given above as ordinates on the corresponding centers,
and a smooth curve is obtained. The centre corresponding to the lowest factor of safety is then the
critical centre.

By: - Birhane G/yohannes (PhD.)


47
Nature and Classification of Engineering Soils

Figure 2.25: Location of center of critical slip circle

Stability of downstream slope during steady seepage


Critical condition for d/s slope occurs when the reservoir is full and percolation is at its maximum
rate. The directions of seepage forces tend to decrease stability. In other words, the saturated line
reduces the effective stress responsible for mobilizing shearing resistance.
cL  tan  ( N  U )
F .S .  ….. (2.25)
T
When U is the total pore pressure on the slope surface

Fig 2.26: Stability of downstream slope during steady seepage

By: - Birhane G/yohannes (PhD.)


48
Nature and Classification of Engineering Soils

The pore-water pressure at any point is represented by the piezometric head (hw) at that point. Thus
the variations of pore water pressure along a likely slip surface is obtained by measuring at each
of its intersections with an equipotential line, the vertical height from that intersection to the level
at which the equipotential line cuts the phreatic line. The pore pressure represented by vertical
height so obtained are plotted to scale in a direction normal to the sliding surface at the respective
point of intersection. The distribution of pore water pressure on the critical slope surface during
steady seepage is shown hatched in fig.3.19.The area of U- diagram can be measured with help of
a planimeter.
In the absence of a flow net, the F.S of the d/s slope can approximately be from the equation
cL  tan  N '
F .S .  … (2.26)
T
The following unit weights may be used for the calculation of  N ' and T when pore pressure
are otherwise not included in the stability analysis, however the Phreatic line needs to be drawn.

Location Driving force Resisting force


Below phreatic surface Saturated weight Submerged weight
Above phreatic surface Moist weight Moist weight

By: - Birhane G/yohannes (PhD.)

Potrebbero piacerti anche