dam engineering grv butt

74
9 1 . I N T R O D U C T I O N 1.1. THE NEED FOR DAMS Dams are built to impound water and discharge it for a variety of purposes into channels, tunnels and conduits, to create head of water needed for hydropower generation and to protect the land against the ravages of floods. Above all, however they provide active storage for the management of water, which is for striking a balance between natural flow regimes on the one hand and patterns of demand, which are normally quite different on the other. Depending on the size of the reservoir, this equalizing function may occupy a daily, weekly or annual cycle, or even tide over multiannual periods. All the dams, with only very few exceptions, offer security against two extremes: (1) Against a lack of water bringing draught, power failures, dried out river beds and falling ground water levels and (2) Against too much water, especially too much too quickly, in the form of raging floods causing devastating inundation to farmland and people's homes. The primary function of dams worldwide is to provide reliable water supplies to a major proportion of the approximately 2 700 000 km 2 presently under irrigation in the world. And for the future the United Nations Development Programme (UNDP) requires a 3% compound rate of annual growth to meet the needs of an extra one billion people in the next ten years. The control of floods, which still cause 40% of fatalities from natural catastrophes worldwide, has always been particularly significant motive for dam construction everywhere, and sometimes its primary purpose. In many other cases flood control is a valuable spin-off from dams built primarily to serve other purposes, which also represent the subsequent source of revenue, in particular power production. With a total annual generation of 2.1 million GWh, hydropower today account for 20% of electricity consumption worldwide. Even at a conservative estimate, the total exploitable potential is at least six times larger - an incentive for the further development of this source of clean, renewable energy. Storage reservoirs form the backbone of innumerable water supply systems, and many more of them will still be needed for this purpose. Efforts to supply water have to keep up with the unprecedented growth in population that already face a serious scarcity of safe water. Augmentation of low flows in rivers, the improvement of water quality, the enhancement of tourism, recreation areas and fishery are all additional assets rightly claimed for many dams and their reservoirs.

Upload: tiberiu-cotoi

Post on 11-Dec-2015

60 views

Category:

Documents


18 download

DESCRIPTION

sd

TRANSCRIPT

Page 1: Dam Engineering GRV BUTT

9

1. INTRODUCTION

1.1. THE NEED FOR DAMS

Dams are built to impound water and discharge it for a variety of purposes into

channels, tunnels and conduits, to create head of water needed for hydropower

generation and to protect the land against the ravages of floods. Above all, however

they provide active storage for the management of water, which is for striking a

balance between natural flow regimes on the one hand and patterns of demand, which

are normally quite different on the other. Depending on the size of the reservoir, this

equalizing function may occupy a daily, weekly or annual cycle, or even tide over

multiannual periods.

All the dams, with only very few exceptions, offer security against two extremes: (1)

Against a lack of water bringing draught, power failures, dried out river beds and

falling ground water levels and (2) Against too much water, especially too much too

quickly, in the form of raging floods causing devastating inundation to farmland and

people's homes.

The primary function of dams worldwide is to provide reliable water supplies to a

major proportion of the approximately 2 700 000 km2 presently under irrigation in the

world. And for the future the United Nations Development Programme (UNDP)

requires a 3% compound rate of annual growth to meet the needs of an extra one

billion people in the next ten years.

The control of floods, which still cause 40% of fatalities from natural catastrophes

worldwide, has always been particularly significant motive for dam construction

everywhere, and sometimes its primary purpose. In many other cases flood control is

a valuable spin-off from dams built primarily to serve other purposes, which also

represent the subsequent source of revenue, in particular power production.

With a total annual generation of 2.1 million GWh, hydropower today account for

20% of electricity consumption worldwide. Even at a conservative estimate, the total

exploitable potential is at least six times larger - an incentive for the further

development of this source of clean, renewable energy.

Storage reservoirs form the backbone of innumerable water supply systems, and many

more of them will still be needed for this purpose. Efforts to supply water have to

keep up with the unprecedented growth in population that already face a serious

scarcity of safe water.

Augmentation of low flows in rivers, the improvement of water quality, the

enhancement of tourism, recreation areas and fishery are all additional assets rightly

claimed for many dams and their reservoirs.

Page 2: Dam Engineering GRV BUTT

10

At present reservoirs behind the dams store some 5500 km3 of water, of which about

3600 km3 is available for regular use. This figure is of the same magnitude as the

worldwide water withdrawals for consumptive use, estimated at present at about 3700

km3/year, of which 68% is used for the irrigation of croplands, 24% for industry and

8% for domestic and industrial purposes. The sheer magnitude of these figures shows

how much dams and appurtenant reservoirs have become an integral part of our large-

scale engineered infrastructure, of our man-made basis of survival. At the same time

the rate of growth in the world population and the pressures that growth is creating in

terms of water, food and energy supplies require the continued creation and

maintenance of further storage volumes.

Today about 45 000 large dams (higher than 15 m) and an estimated 800 000 smaller

ones improve the living conditions of many of the world's six billion people. Yet

about 1.5 billion of these people have no reliable access to clean water and hygienic

conditions, a fundamental human right. Moreover the world population continues to

grow, in some regions at a rapid rate and in concentrated areas.

Dams provide all these benefits, but there is usually also a cost. People, other life

forms and the natural environment are affected when dams and reservoirs are built,

and the impact may require resettlement and relocation. The challenge is to balance

the benefits and costs in the short and long term. When the needs are water supply,

food production, flood control and electricity production, what realistic options and

alternatives does society have, and what are the short and long-term consequences?

Along with the growth in size of the dams and their reservoirs their environmental

impact and the awareness thereof increased as well. In the late 1960's these aspects

became the object of public debates, which, at times degenerated into bitter

controversy.

Besides the change of landscapes - by no means always negative - dam constructions

had drastic consequences whenever large numbers of people had to be resettled from

the reservoir area. Unfortunately, several of these operations were poorly planned

and/or badly managed.

As is the case in other fields of technology, is not the dam construction that should be

blamed if anything goes wrong. To blame is the abuse that occurs for often short-lived

and/or particular advantages. Through a long history, dam engineers have acquired the

ability to serve both goals – need of water storage and environmental and social

compatibility-efficiently, securely and in an environmentally safe way.

Looking ahead, people will continue to build dams for many purposes, especially in

the developing world. The reason is that dams in very many cases will be judged to

represent the best option among the feasible alternatives to satisfy critical human and

societal needs. As the world's population continues to increase in water-scarce

regions, additional water resources must be developed. The only practical way to

achieve this on the scale required is to increase storage capacity and that means to

build more dams.

Page 3: Dam Engineering GRV BUTT

11

1.2. BRIEF HISTORY OF CONCRETE DAM EVOLUTION

The predecessors of modern dams are masonry dams dating back more than 2000

years, and some of these still remain in excellent conditions. In such dams the

upstream and downstream faces have been formed by dressed masonry and the main

body comprises rubble masonry bedded in lime mortar. With the advent in the last

century of Portland cement concrete and the increased labor costs of stone quarrying,

the concrete dam has largely replaced the masonry dam, especially in industrialized

countries.

Along the years the concrete dam has developed into a number of specific types such

as the double curvature arch and the buttress. More recently still, the increased use of

earth moving plant has firmly established roller compaction as the frequently

preferred method for mass concrete construction (Figure 1.1).

Fig.1.1. Development of concrete dam types

In their simplest forms both masonry and concrete dams are gravity structures with

near-vertical upstream faces and downstream slopes in the order of 0.8:1. The water

thrust from upstream is resisted by assumed sliding planes within the dam body and

ultimately through sliding resistance at the base. Friction is the principal

consideration, although in many cases some acknowledgement of effective cohesion

is also used. The science of designing such dams really developed in the last century

with the increased awareness of internal and foundation uplift pressures and their

fundamental importance to the stability of such dams.

RCC dam

FSHD

Page 4: Dam Engineering GRV BUTT

12

Gravity dams are relatively understressed structures, and attempts at economy have

been made by forming voids. Such hollow gravity dams are not common nowadays.

The logical progression from these has been the buttress dam, where a series of

"counterforts" or upstream/downstream walls are used to support a continuous facing.

The facing, or buttress heads, can be profiled to ensure that the head stays in overall

compression, thus minimizing reinforcement requirements.

In buttress and some hollow gravity dams, the foundations between the buttresses are

open, thus eliminating uplift pressure. The upstream faces are also typically sloped to

encourage additional hydraulic load to improve overall stability. In the cases of both

buttress and hollow gravity dams, stresses are higher than in conventional gravity

dams, and therefore the qualities of concrete and foundations needed are also higher.

The use of both types has waned, as a result of the increased costs of labor and

shuttering compared with simple gravity construction.

Arch dams, in masonry form, have been constructed for some several hundred years.

However, the last 50 years has seen the use of advanced mathematical techniques

such as finite element analyses, now enhanced by increased computing power and

availability, to optimize double curvature design. Such dams can of course only be

constructed where the topography and geology are appropriate. The requirement for

good quality concrete is generally high and complications such as joints grouting and

pre-and/or post-cooling of the concrete inevitably increase the cost.

The arch-gravity dam is sometimes seen as a type of dam in its own right; however,

simple analyses generally demonstrate that, in normal working conditions, the arch

action tends to be minimal and, indeed, the dam functions normally as a slender

gravity dam.

Perhaps the biggest advantage which concrete dams have over earth dams is their

ability to tolerate diversion floods with minimal damage and their ability to

accommodate, or act as, spillways, eliminating the need for special separate

structures.

In the case of concrete dams, the most recent revolution has been re-establishing the

concrete gravity dam as an economic alternative, given the introduction of roller

compacted concrete techniques. Initially, three schools emerged. At one extreme was

the use of a relatively lean and porous mix at dams. At the other extreme was the use

of a high paste mix. In the latter case, however, the tendency has been to increase

specification requirements, as problems have developed, and to lose some of the

benefits of cheap roller compaction.

In this sense the evolution of the concrete dam has in fact regressed back to the

largely gravity section by the use of a different construction method, although

examples also exist of both arch and arch-gravity RCC structures.

An intermediate form of gravity dam is the FSHD or "hardfill" dam. This form of

dam, in effect, represents a very weak form of RCC, but with cohesion and more self-

supporting than pure rockfill. The dam would slope both upstream and downstream.

Such a dam can be expected to be permeable and, therefore, some form of upstream

Page 5: Dam Engineering GRV BUTT

13

facing is required such as those currently used on concrete faced rockfill dams. Such a

facing can be incorporated in the dam construction, while a number of concrete dams

have now been waterproofed upstream with geomembranes.

Another intriguing option could be the fully drained gravity dam, which is a simple

gravity dam with an impermeable face founded on a confined layer of rockfill. A

secure, impermeable barrier, with drainage provision, would protect the main body

from water penetration, while a free-draining layer of rockfill under the dam would

mean that foundation seepage pressure could be relieved, eliminating upthrust.

Eliminating uplift can, in theory, reduce the downstream slope of a simple gravity

dam from 0.8:1 to 0.5:1.

1.3. BRIEF HISTORY OF EMBANKMENT DAM EVOLUTION

The basic concept of the embankment dam can be traced back 4500 years to the

earliest well known example of the Sadd-el-Kafara dam on the River Nile. This was a

zoned dam, although generally the earliest dams, and even today the simplest dams,

are homogeneous earth embankments: that is, the material used for the construction

remains essentially constant throughout. The material therefore has to provide both

sufficient impermeability to give the dam its purpose, while at the same time it must

have sufficient strength to retain its shape and resist water loads. Internal water

pressures will be based on a phreatic seepage line. Even when attempts are made at

zoning by using the same material compacted to different densities, the result, in

practice, is frequently homogeneity.

Figure 1.2. Simplified development of embankment dam types

The inevitable development from the homogeneous embankment was the zoned dam,

and there are good historic examples of these (Figure 1.2). In the case of the zoned

Bro

ad

ev

olu

tio

na

ry d

irec

tion

Homogenous

Embankment

Zoned

Embankment

Embankment with

Waterproof

Facing

Page 6: Dam Engineering GRV BUTT

14

earthfill or rockfill dam, materials are selected and used for their specific properties

and in proportions guided by local site availability. Typically, the central core of the

dam will be made of a relatively impermeable material such as clay, silt or moraine.

This will be supported upstream and downstream by shoulders, often of random

material. Cores have also been constructed of bitumen, concrete and, less commonly,

steel and plastic.

Key elements in the design of a zoned dam also include the transitions or filters

between the core and the adjacent shoulder material. These can be particularly critical

on the downstream side, where they will act both as filters to retain the migration of

the fine core material and as drainage layers to avoid pressures building up in the

downstream shoulder. Vertical chimney drains may be supplemented, where appro-

priate, by horizontal drains through the shoulders. It should be noted that in the case

of such a zoned embankment with a narrow core, the hydraulic thrust is resisted

almost entirely by the downstream shoulder. It will do this by internal friction and

cohesion, both through the material itself and at and through the foundations. Many

thick cores transfer a major portion of thrust directly into the foundation via shear.

The construction of zoned embankment dams has advanced greatly in the last century

as a result of the efforts of engineers such as Terzaghi who, through rigorous research

and testing, enhanced the art of dam design with sound engineering science. It is

gratifying to the designer, however, to note that design still retains an element of both

art and science. One change which has accompanied the science of earth dams has

been the increased use of optimal compaction densities, a trend which is being relaxed

in circumstances such as wet core construction in high rainfall areas.

Rockfill as a principal construction medium for embankment dams has also advanced

considerably since the 1950s. Early construction often used quarry-run rockfill,

dumped relatively dry.

The next great change came in the 1960s with the increased compaction effected by

vibratory rollers. This permitted increased heights for all types of rock-fill

embankment but, in particular, resulted in a considerable increase of confidence in the

construction of faced rockfill dams. In such dams, the principal body is composed

entirely of rockfill, and the waterproof membrane is placed on the sloping upstream

face. Modern facings tend to be of reinforced concrete or asphalt, although early

examples also included steel and timber. Good design guidelines and precedents exist

now for the confident design of such dams to heights well in excess of 100 m. Careful

attention is needed to aspects such as the peripheral joint and to local seepage

gradients at the lower plinth.

In the case of such dams it should be noted that the entire rockfill body remains

protected from the reservoir and also acts to resist hydraulic water loads. Indeed, the

upstream sloping face will attract a stabilizing vertical load from the water upstream.

In very recent years rockfill embankments have been built which can withstand

overtopping. The design of embankments to overtop is, however, still rare.

Page 7: Dam Engineering GRV BUTT

15

1.4. DAM WORLD SITUATION

Figure 1.3 shows the evolution in the total number of large dams in operation

(according to International Commission of Large Dams a large dam is the one that is

higher than 15 m). At the beginning of the 20-century the total number of large dams

was only 427. In 1950 the dam number reached 5268. Since the World War II the

number of dams has increased considerably, especially in Asia. The last accurate

inventory of large dams was in 1992, when a total number of 36,235 large dams in

operation were registered.

Fig.1.3. Evolution of the number of large dams in the world

The significant increase of in dams since 1992 does not result only from new

constructions but is largely because of the correction of statistics made by China.

In the 21st century the dam construction is expected to maintain its rate of

development at least in Asia and Latin America. It should be noted that in

industrialized countries most of the potential dam construction sites have already been

exploited.

In 2001, a total number of 361 large dams with heights of more than 60 m, and 35

with heights of more than 150 m, were under construction in the world. China,

Turkey, India, Japan and Iran hold the first five places on the list.

Selecting the highest concrete dams, the list of dams under construction includes:

Gravity dams – Kishau (India), H = 236 m; Three Gorges (China), H = 175 m;

Shafaroud (Iran), H = 159 m.

Arch dams – Ertan (China), H = 240 m; Karun IV- Monj (Iran), H = 230 m;

Nukui (Japan), H = 156 m.

The multiple arch dam – Berke (Turkey), H = 201 m.

Selecting the highest embankment dams, the list of dams under construction includes:

Europe

Asia

America

Africa

Australia

TOTAL

Year

Page 8: Dam Engineering GRV BUTT

16

Earthfill dams – Yusufeli (Turkey), H= 270 m; Tehri (India), H= 261 m;

Gotrand (Iran), H=180 m; Xianolangdi (China), H= 154 m.

Impervious core rockfill dams – San Roque (Philippines), H= 200 m; Sykia

(Greece), H= 175 m; Yacambu (Venezuela), H= 158 m.

Concrete face rockfill dams – La Miel (Colombia), H=188 m; Tianshengqiao

(China), H= 178 m; Ita (Uruguay), H= 154 m.

On the other hand in Europe, once the location of all the dam records, only four

concrete dams with heights of more than 100 m were under construction (all of them

gravity dams), three in Spain and one in Italy.

Figure 1.4. shows the evolution of maximum dam heights around the world. The two

main types of dams are presented: concrete dams and embankment dams. For concrete

dams the world record is still held by the Grande Dixance gravity dam built in 1962 in

Switzerland at 285 m. For embankment dams the world record is held by Nurek earth

core rockfill dam built in 1980 in Russia at 308 m.

Fig. 1.4. Evolution of maximum dam heights

The highest arch dam in the world is Vaiont dam (Italy), with a total height of 262 m.

Since the reservoir behind the Vaiont dam is empty due to an unfortunate accident of

the reservoir slopes, the height record among the arch dams belongs to Mauvoisin

(Switzerland) with 237 m.

The records in the field of buttress dams are differentiated by dam types. Among the

massive head buttress dams the highest is Hatanagi (Japan), H = 125m. The highest

hollow gravity dam was completed in 1986 for the Itaipu waterpower development,

with a total height of 186 m. The highest multiple arch dam is Daniel Johnson (built

Page 9: Dam Engineering GRV BUTT

17

in 1970Canada), H + 214 m. As far as flat slab and buttress dams are concerned they

are of a smaller size, the record being realized in 1949 in Argentina, with a height of

88 m.

Although more recently developed, RCC dams are also very well represented in the

list of the highest dams. La Miel I dam in Colombia, commissioned in 2003, has a

height of 188 m. Two other high dams were recently built in Japan: Urayama (H =

156 m, in 1999) and Miyagase (H = 155 m, in 2001).

After Nurek, the second clay core rockfill dam in terms of height is Chicoasen dam in

Mexic, with a total height of 261 m.

The record in the field of earthfill dams is held by Mica dam built in 1973 in Canada,

as a dam with an incined central core. Oroville dam in USA with a height of 230 m is

the second of the same concept.

The highest concrete face rockfill dam in operation is Tianshengqiao, in China, with a

total height of 178 m. Storglomvatn dam in Norway, with a maximum height of 125

m is the highest rockfill dam with asphaltic core in the world.

Looking at history, it can be observed that the welfare of human beings has always

related to the development of hydraulic schemes, among which the most important

civil engineering works are the dams. In the future, hydraulic schemes and dams will

retain this fundamental importance for humanity.

BIBLIOGRAPHY

Hoeg, K. Lecture presented at the World Water Forum, The Hague, in March 2000.

Hydropower and dams. World Atlas 2000.

Hydropower and dams. World Atlas 2000.

Mason, P., J. The evolving dam. Hydropower and dams, Issue five, 1997.

Schnitter, N. A history of dams – the useful pyramids. A.A.Balkema, 1994.

Schleiss,A.,J. The importance of hydraulic schemes for sustainable development in

the 21st century. Hydropower and dams, Issue one, 2000.

Page 10: Dam Engineering GRV BUTT

18

2. GRAVITY DAMS

2.1. INTRODUCTION

Gravity dams are dams that by their own weight resist the forces imposed on them

with a desired factor of safety. Gravity dams are designed so that every unit of length

is stable, independent of every other unit of length; hence, the dams can be straight,

curved or arched, or even crooked. The typical gravity dam is straight, although many

designers curve the axis upstream to gain added stability through arch action.

One hundred years ago, the typical gravity dam was constructed of masonry. Large

stone blocks were shaped and fitted into the structure, and the joints filled with

cement and sand mortar. The materials for concrete gravity dams evolved gradually

into mass concrete. When dams constructed in this way became too expensive, in

comparison with earthfill and rockfill dams, a new construction method evolved for

concrete gravity dams, which placed and compacted concrete materials using earthfill

construction techniques, resulting in the term roller compacted concrete.

2.2. STRUCTURAL FEATURES

2.2.1. LAYOUT

The general layout of a gravity dam is shown in figure 2.1. The dam body is divided

by transverse contraction joints in to blocks each of which carries its load to the

foundation without any transfer of the load from or to adjacent vertical elements. The

block length between two adjacent contraction joints is usually in the range of 15...18

m. The joints are designed to avoid extensive random cracking of mass concrete that

can be induced by shrinkage and thermal contraction. The large amount of heat that is

generated during the concrete hardening leads to a significant change from the

maximum temperature to the final stable temperature inducing large volume changes

with consequent random cracking.

Properly designed contraction joints constitute only one provision for the control of

cracks. Other provisions are as follows:

- The construction of blocks by pouring the concrete in shallow vertical lifts,

preferably not more than 2 m high in large dams;

- The use of low cement content, with a low rate of heat generation;

- Precooling of aggregates and of the mixing water;

- Allowance of sufficient time between the constructions of vertical lifts to

permit a large loss of heat from the surface;

The horizontal line that is located in the middle of the dam crest and connects the two

dam abutments is the dam axis. The dam axis is straight or only slightly curved in

plane.

Page 11: Dam Engineering GRV BUTT

19

The maximum cross section (upstream-downstream) through the dam body is called

the dam profile. The shape of the dam profile is determined by the prescribed loading

conditions, the shear resistance of the foundation rock and the height of the maximum

section.

Fig. 2.1. General layout of a gravity dam:

a. plan view; b. longitudinal and transverse cross sections:

1-dam axis; 2-nonoverflow block; 3-overflow block; 4-lifts; 5-contraction joint; 6-

dam crest; 7-spillway; 8-bottom outlet; 9-stilling basin; 10-ground line; 11-

foundation contour; 12-upstream face; 13-downstream face; 14-dam heel; 15- dam

toe; 16-foundation surface; 17-inspection gallery; 18-grouting and drainage

gallery; 19-grout curtain; 20-drainage drillings.

Several dam profiles, corresponding to representative gravity dams, are shown in

figure 2.2. The upstream face of a gravity dam is usually made vertical or slightly

inclined towards upstream to concentrate the concrete weight at the upstream face

where it acts to overcome the effects of the reservoir water load. The downstream face

usually have a uniform slope that is determined by both stress and stability

requirements at the base. The ratio between the dam basis B and the dam height H

(k=B/H) is in the range of 0.75 ... 0.85.

A batter may be used on the lower part of the upstream face to increase the thickness

at the base to improve the sliding safety of the blocks. If a batter is used, stresses and

stability should be checked where the batter intersects the vertical upstream face.

The crest thickness is dictated by roadway or other access requirements. Its value is

usually in the range of 5.00 to 10.00 m. When additional crest thickness is used the

downstream face should be vertical from the downstream edge of the crest to an

intersection with the sloping downstream face.

In most of the cases several blocks of a gravity dam are designed as overflow or

spillway section. The overflow profile is inscribed in the nonoverflow section. The

curves describing the spillway crest and the junction of the slope with the energy

dissipater are designed to meet the hydraulic requirements.

Page 12: Dam Engineering GRV BUTT

20

Fig. 2.2. Gravity dam profiles

The entire area to be occupied by the base of the gravity dam is excavated to firm

material capable of withstanding the loads imposed by the dam. Generally, a

foundation surface appears as horizontal in the transverse (upstream-downstream)

direction. However, where an increased resistance to sliding is desired, the surface can

be sloped, upward from heel to toe of the dam.

If the canyon profile for a dam site is relatively narrow with steep sloping walls, each

vertical section of the dam, from the center towards the abutments is shorter in height

than the preceding one. Consequently section closer to the abutments will be deflected

less by the reservoir load. A sharp break in the excavated profile of the canyon will

result in an abrupt in the height of the dam. The effect of the irregularity of the

foundation rock causes a marked change in stresses in both the dam and foundation

and in stability conditions. For this reason, the foundation is shaped so that a

uniformly varying profile is obtained along the longitudinal contour, free of sharp

offsets or breaks.

In order to establish an effective barrier to seepage under the dam and to consolidate

the foundation rock the foundation treatment requires rock grouting. A grout mix,

basically consisting in water and cement, is injected in holes drilled into the rock

foundation.

Grouting operations may be performed from the surface of the excavated foundation,

from the top of concrete placements for the dam, from galleries within the dam and

from tunnels driven into the abutments. The preliminary low-pressure shallow

consolidation grouting is followed by high-pressure, deep curtain grouting.

LONGYANGXIA

Page 13: Dam Engineering GRV BUTT

21

Although a well-executed grouting program may materially reduce the amount of

seepage, some means are provided to intercept the water that percolates through and

around the grout curtain. Drainage is usually accomplished by drilling one or more

lines of holes downstream from the high - pressure grout curtain. As a general rule,

hole depths vary from 20% to 40% of the reservoir depth and 35% to 75% of the deep

curtain grouting depth. Drain holes are drilled from foundation and drainage galleries

within the dam, or from the downstream face of the dam if no gallery is provided.

Several examples of gravity dams are shown on the following figures. At the time of

its completion in 1942, Friant Dam (figure 2.3) was the fourth largest concrete dam

in the world. The dam is a concrete gravity structure with a structural height of 97 m

and a crest 1060 m long. 128 mm diameter drainage holes were drilled into the

foundation rock. In addition, the foundation was grouted by a cement grout curtain

under the upstream toe.

Fig. 2.3. Friant dam

Shasta Dam (figure 2.4) is a curved concrete gravity structure. The dam has a crest

length of 1054 m, a maximum height of 184 m, and a maximum base thickness of 269

m. The downstream face of the dam has a slope of 0.8 horizontal to 1 vertical, and the

Page 14: Dam Engineering GRV BUTT

22

upstream face is vertical except for the portion falling below El. 720.0, which slopes

upstream at o.5 horizontal to 1 vertical.

Fig. 2.4. Shasta dam

The dam, while curved in plan to accommodate the configuration of the site, was

designed as a gravity structure throughout, resisting the pressure of the reservoir by its

sheer mass. Necessity of passing a major portion of the flood flow over the

downstream face of the dam led to the design of the world's highest man-made

waterfall.

Page 15: Dam Engineering GRV BUTT

23

Izvorul Muntelui dam (Figure 2.5) holds the height record for gravity dams in

Romania. The dam has a crest length of 453 m, a maximum height of 127 m, and a

maximum base thickness of 119 m. The upstream face of the dam is inclined towards

upstream presenting two slopes while the downstream face has a more uniform slope

of 0.681 horizontal to 1 vertical, on the average. The dam body is divided into 30

blocks of 15 m wide, out of which four are overflowing blocks located in the central

area.

Two types of concrete were used, according to a concrete zoning concept: (1) face and

foundation concrete, with a cement content of 270 kg/m3 and (2) inner concrete with a

cement content of 175 kg/m3. The drainage system of the dam comprises the drains in

the dam body that collects the infiltrated water through 30 cm diameter porous pipes

and 15 m deep drilling holes in foundation, arranged on three successive rows from

downstream of the grout curtain to the dam toe.

Fig.2.5. Izvorul Muntelui Dam:

1- concrete block; 2- spillway; 3- bottom outlet; 4- gate chamber; 5- access shaft to

the gate chamber; 6- stilling basin; 7- road at the dam crest; 8- inspection galleries; 9-

grout curtain; 10- drainage galleries; 11- lateral wall.

MAIN CROSS SECTION

Page 16: Dam Engineering GRV BUTT

24

2.2.2 FOUNDATION TREATMENT

Excavation – As it was mentioned before, the entire area to be occupied by the base

of the gravity dam should be excavated to firm material capable of withstanding the

loads imposed to the dam, reservoir, and appurtenant structures. Very often the

exploratory drilling or final excavation uncovers faults, seams, or shattered or inferior

rock extending to such depths that it is impracticable to attempt to clear such areas out

entirely. These conditions require special treatment in the form of removing the weak

material and backfilling the resulting excavations with concrete. This procedure of

reinforcing and stabilizing such weak zones is frequently called "dental treatment"

Over excavation to achieve symmetry on a nonsymmetrical site is not recommended.

Sharp breaks in the excavation profile should be avoided. Thrust blocks may be used

to establish an artificial abutment where a natural one does not exist, but not to

provide symmetry.

Foundation Grouting - The principal objectives of foundation grouting continue to

be consolidation (under low pressures) and establishment of an effective barrier to

seepage under and around the dam (under high pressures). Spacing, length and

orientation of grout holes, and the procedure to be followed in grouting a foundation

are dependent on the height of the structure and the geologic characteristics of the

foundation.

Consolidation grouting - Low pressure grouting to fill voids, fracture zones, and

cracks at and below the surface of the excavated foundation is accomplished by

drilling and grouting relatively shallow holes. For structures over 30 m in height, a

preliminary consolidation grout program will call for lines of holes parallel to the axis

of the dam extending from heel to toe, and spaced about 3 to 6 m apart. Holes with a

minimum diameter of 45 mm are staggered on alternate lines to provide better

coverage of the area. The depth varies from 10 to 15 m depending on local conditions

and to some extent on the height of the dam. Such a program is carried out using the

principal of split spacing. Drilling is usually accomplished from the excavated

surface, although in some cases drilling and grouting to consolidate steep abutments

has been accomplished from the tops of concrete placements in the dam to prevent

"slabbing" of the rock. Water-cement ratios for grout mixes may vary widely

depending on the permeability of the foundation rock. Starting water-cement ratios

usually range from 8:1 to 5:1 by volume. An admixture such as sand or clay may be

added if large voids are encountered.

Curtain Grouting - Construction of a deep grout curtain near the heel on the dam to

control seepage is accomplished by drilling deep holes and grouting them using

higher pressure. To permit application of high pressure without causing displacement

in the rock or loss of grout through surface cracks, this grouting procedure is carried

out subsequent to consolidation grouting and after some of the concrete has been

placed. Usually, grouting will be accomplished from galleries within the dam from

tunnels driven into the abutments especially for this purpose.

The alignment of holes should be such the base of the grout curtain will be located the

vertical projection of the hell of the dam. If drilled from a gallery that is some distance

Page 17: Dam Engineering GRV BUTT

25

from the upstream face, the holes may incline as much as 150 upstream from the plane

of the axis. If the gallery is near the upstream face, the holes will be nearly vertical.

The usual procedure will be first to drill and grout holes approximately 10 to 12 m

apart, or as far apart as necessary to prevent grout from one hole leaking into another

drilled but ungrouted hole. Drilling and grouting of additional intermediate holes,

splitting the spaces between completed holes, will continue until the desired spacing is

reached or until the amount of grout accepted by the last group of intermediate holes

indicates no further grouting is necessary. During the progress of the grouting, local

conditions may determine the actual or final depth of grouting.

The depth of holes depends on the characteristics of the foundation rock. For hard,

dense, relatively tight rock, the depth may be only 30 to 40 percent of the head, but

for rock with more permeable zones, the depth may reach 70 to 100 percent of the

head. The spacing of holes may be only 3 m or less. Foundations usually increase in

density and tightness of seams at greater depth.

Two general methods of grouting are used that permit higher pressures in lower

zones: descending stage grouting and ascending stage grouting. Descending stage

grouting consists of drilling a hole to a limited depth or to its intersection with an

open seam, grouting to that depth, cleaning out the hole after the grout has taken its

initial set, and then drilling and grouting next stage. Ascending stage grouting consists

of drilling a hole to its final depth and grouting the deepest high-pressure stage first by

use of a packer that is seated the top of this stage. After grouting this stage, the grout

pipe is raised so that the packer is at the top of the next stage, which is subsequently

grouted using somewhat lower pressure. This staged process is repeated, working

upward until the hole is completely grouted. Ascending stage grouting is becoming

more generally used, as it reduces the chances for displacement of the foundation

rock, gives better control as to the zones of injection, and expedites drilling.

Foundation Drainage - The purpose of drainage is to collect water that passes

through and around the grout curtain and to prevent the build up of high hydrostatic

pressures at the base of the dam and in the abutments. Usually one or more lines of 75

to 100 mm holes are drilled downstream of the grout curtain. Spacing, depth and

orientation are all influenced by foundation conditions. The holes are usually spaced 3

to 5 m on center. As a general rule, the depth varies from 20 to 40 percent of the

reservoir depth and 35 to 75 percent of the depth of the high-pressure grout curtain.

Drain holes should be drilled after all foundation grouting has been completed in the

area. They can be drilled from foundation and drainage galleries within the dam, or

from the downstream face of the dam if no gallery is provided.

In some instances where the stability of a rock foundation may be beneficiated by

reducing the hydrostatic pressure along planes of potentially unstable rock masses,

drainage holes have been introduced to alleviate this condition. A collection system

for drainages should be designed so that flows can be gathered and removed from the

area.

Page 18: Dam Engineering GRV BUTT

26

2.2.3. JOINTS

Cracking in concrete dams is undesirable because cracking in random locations can

destroy the monolithic nature of the structure, thereby impairing its serviceability and

loading to an early deterioration of the concrete. Cracking will occur when tensile

stresses are developed which exceed the tensile strength of the concrete. These tensile

stresses may occur because of imposed loads on the structure, but more often occur

because of restraint against volumetric change. The largest volumetric change in mass

concrete results from change in temperature and the cracking tendencies occur as a

result.

Joints placed in mass concrete are designed cracks, located where they can be

controlled and treated to minimize any undesirable effects. Two principal types of

joints are used in concrete dams: contraction and construction joints.

Contraction joints - In order to control the formation of cracks in mass concrete

dams, the current practice is to construct the dam in blocks separated by transverse

contraction joints. These contraction joints are vertical and extend from the

foundation to the top of the dam and are continuous from the upstream face to the

downstream face.

The location and spacing of transverse contraction joints should be governed by the

physical features of the dam site, results of temperature studies, placement methods

and the probable concrete mixing plant capacity. For average conditions a spacing of

15 m has proved to be satisfactory. In dams where pozzolan and retarders are used,

spacing up to 18 m have been acceptable.

A typical disposition of contraction joints is shown in Figure 2.6.

Fig. 2.6. Contraction joints:

a. Metal water stop; b. Plastic water stop:

1-copper strip ; 2-copper strip Z; 3- drainage shaft; 4-concrete; 5-PVC strip; 6-

bituminous mastic; 7-pipe for regenerating steam; 8- inspection shaft

Page 19: Dam Engineering GRV BUTT

27

The water stop provided by the metal or plastic seal is located at 1 to 1.5 m from the

upstream face. Stainless steel or copper strips were extensively used in the past. The

metal strip thickness is in the range of 2 to 3 mm, the width is between 0.75 and 1.5 m

and the length is some 3… 4 m. The strips are welded on site during the dam erection.

The wave placed within the joint surface (sometimes in a bituminous mastic box) may

have the shape of ,V, M or Z (figure 2.7).

Fig. 2.7. Metal water stops:

a. simple; b. shape; c. polygonal shape; d. Z shape

The plastic strips, usually PVC, are more common in the present. Such seals are 3 to 6

mm thick by 30 to 40 cm long, with a 1.5 cm radius bulb at each end and a cored 2 cm

radius bulb at the intersection of the joint (Figure 2.8). The plastic strips are provided

at much longer length, over 10 m, and are easily welded during the construction.

Fig. 2.8. Plastic PVC water stops:

a. strip types; b. capability of strips to withstand relative movements; c. joining

elements:

1- adjacent concrete blocks; 2- concrete blocks separated by bituminous mastic

Contraction joints should be constructed so that no bond exists between the concrete

blocks separated by the joint. However, in the American practice vertical keys are

sometimes provided in transverse joints. The keys are used primarily to provide

increased shearing resistance between locks. When the joints and keys are grouted a

monolithic structure is created which has greater rigidity and stability because of the

transfer of load from one block to another through the keys.

Page 20: Dam Engineering GRV BUTT

28

Drainage shafts are placed downstream of the water stops. The shaft section may have

circular or square shape with a radius / side length of 0.8 to 1.2 m. From the drainage

shaft towards the downstream face the joint is free draining.

Construction joints - Mass concrete is usually placed in 1.5 or 2 m lifts separated by

construction joints. The integrity of a concrete structure is dependent to a large extent

on the proper preparation of construction joints before placing fresh concrete upon the

construction joint surfaces. Bond is desired between the old and new concretes. Each

successive layer must be placed while the next lower layer is still plastic. The

vibrators must penetrate through each layer and revibrate the concrete in the upper

portion of the underlying layer to obtain a dense monolithic concrete throughout the

lift. Such a procedure will also prevent cold joints within the placement lift.

All laitance and inferior surface concrete must be removed from the old surface with

air and water jets and wet sandblasting as necessary. All surfaces should be washed

thoroughly prior to placing the new concrete, but should be surface dry at the time

they are covered with the fresh concrete. Generally a layer of concrete a few

centimetres thick with a finer particle size or a mortar layer is placed prior to

resuming mass concreting.

Rock surfaces to be covered with concrete must be sound and free of loose material

and should also be saturated, but surface dry, when covered with fresh concrete or

mortar. Mortar should be placed only on those rock surfaces that are highly porous or

are horizontal or nearly horizontal absorptive surfaces.

2.2.4. GALLERIES AND DRAINAGE SYSTEMS WITHIN THE DAM BODY

Drainage is provided primarily to relieve uplift pressure. Transverse joints are drained

by vertical shafts. In the body of the dam the inspection galleries provide additional

relief.

The galleries located within the dam body can be horizontal or inclined, following the

foundation contour along the longitudinal profile of the dam. The galleries are

provided for:

- Drainage collection of seepage water that percolates through the upstream

face of the dam or through the foundation;

- Accommodation of drilling and grouting works for the foundation treatment;

- Direct access in the dam body for regular inspections of the concrete

behavior, rate of cracking and joints performance (water tightness and drainage

control);

- Reading the monitoring parameters collected by the monitoring devices

embedded into the dam body;

- Access to the gate or valve chambers to operate the hydro mechanical and

electric equipment of the spillways and bottom outlets.

The galleries are usually provided close to the upstream face of the dam at distances

that vary from 5 to 10 m. A general rule establishes a minimum thickness between the

upstream face and the gallery of 1/10 of the water depth at the gallery elevation. An

interval of 20 ... 30 m is provided between the horizontal galleries along the dam

height foundation drainage.

Page 21: Dam Engineering GRV BUTT

29

The dimensions of the inspection galleries are limited to 1.20 m width and 2.00 m

height in order to avoid stress concentration. For the same reason their shapes are

rounded (fig. 2.9). The grouting and drainage galleries located in the vicinity of the

foundation have larger dimensions required to accommodate the drilling equipment.

The contours of galleries and shafts are reinforced in order to avoid concrete cracking.

Fig. 2.9. Typical cross sections of inspection and grouting galleries:

1- collecting trench; 2 – concrete fill

2.2.5. MASS CONCRETE FOR DAMS

a. General

The structural quality of concrete is closely related to the structural quality of

aggregate of which it is composed, to the cement type and its characteristics, to the

admixtures incorporated and to the water quality. Thorough and comprehensive series

of tests of all ingredients alone and also the proposed mixes is mandatory for large

dams.

b. Concrete ingredients

Aggregates for use in concrete should be of good quality and reasonably well graded.

When good quality natural sand and coarse aggregate is available, use of crushed sand

and/or coarse aggregate is generally limited to that needed to make up deficiencies in

the natural materials. In these instances, crushing is usually restricted to crushing of

oversize materials and/or the excess of any of the individual sizes of coarse aggregate.

Where little or no natural coarse aggregate is available in a deposit, it may be

necessary to use crushed coarse aggregate from a good quality quarry rock.

An ideal particle size curve is established, which enabled a mix to be produced with

the least possible cement content, while ensuring a good workability. Workability

requirements do not vary much, since the concrete mix is carried with buckets and

compacted in layers with vibrating needles.

Another design consideration is the type of cement to be used. Limitations on the heat

of hydration of this cement are specified when determined necessary to minimize

cracking in the concrete structure. Further limitation on the heat of hydration, if more

INSPECTION GALLERIES

GROUTING GALLERY

Page 22: Dam Engineering GRV BUTT

30

stringent control of heat is needed, can be obtained by providing cement of 70 calories

per gram at 7 days or 80 calories per gram at 28 days, or both. To reduce the heat of

hydration, special cement was selected in some cases.

Other methods of heat reduction include use of lower cement contents, inclusion of a

pozzolan as part of the cementitious material, use of a pipe cooling system, and use of

a specified maximum placing temperature of the concrete. An ideal particle size curve

can be established, which enables a mix to be produced with the least possible cement

content, while ensuring a good workability.

Admixtures are incorporated into the mix design. Admixtures have varying effects on

concretes, and should be employed only after a thorough evaluation of their effects.

Most commonly used admixtures are accelerators; air-entraining agents; water

reducing, set-controlling admixtures (WRA) and pozzolans.

Air-entraining agents should be used to increase the durability of the concrete,

especially if the structure will be exposed to cycles of freezing and thawing. Use of

such agents will expedited the placing of concrete under difficult conditions, such as

for large concrete placements in hot weather.

Good quality pozzolans can be used as a replacement for cement in the concrete

without sacrificing later-age strength. Pozzolan is generally less expensive than

cement and will, as previously indicated aid in reducing heat of hydration.

The water used in the concrete mix should be reasonably free of silt, organic matter,

alkali, salts, and other impurities.

c. Placing and curing

As it was shown in the previous section, mass concrete is placed in 1.5 or 2.0 m lifts

and each of these lifts is made up of 30- to 50 cm layers. Each successive layer must

be placed while the next lower layer is still plastic. The vibrators must penetrate

through each layer and revibrate the concrete in the upper portion of the underlying

layer to obtain a dense monolithic concrete throughout the lift. Bond is desired

between the old and new concretes. Generally a layer of concrete a few centimetres

thick with a finer particle size or a mortar layer is placed prior to resuming mass

concreting.

Thickness of lifts, time interval between lifts, height differentials between blocks, and

seasonal limitations on placing of concrete are main factors that control the concreting

of the dam.

Owing to hydration of the cement, a temperature rise will take place in the concrete

after placement. After the peak temperature is reached, the temperature will decline

depending upon the thickness of section, the exposure conditions, the rate and amount

of continued heat of hydration.

A considerable portion of the total heat of hydration in a placement lift can be lost

through the top exposed surface before the next lift is placed. Shallow lifts and longer

delays between placement lifts will result in the minimum temperature rise in the

Page 23: Dam Engineering GRV BUTT

31

concrete under these conditions. The minimum elapsed time between placing of

successive lifts in any one block is usually restricted to 72 hours, but temperature

studies should be made to relate heat loss or heat gain to the placement lifts.

From a temperature standpoint, an even temperature distribution throughout the

structure will be obtained when all blocks in the dam are placed in a uniform and

continuous manner. This even temperature distribution is desirable because of the

subsequent uniform pattern of contraction joint openings. Extreme temperature

gradients on the exposed sides of blocks will also be lessened when each lift is

exposed for a minimum length of time.

Minimizing the overall height differential between the highest and lowest blocks in

the dam will cause construction of the dam to progress uniformly up from the bottom

of the canyon. In practice, the maximum height differential between adjacent blocks is

usually 7.5 m when 1.5 m lifts are used or 8 m when 2 m lifts are used. The maximum

differential between the highest block in the dam and the lowest block is usually

limited to 12 ….16 m.

Curing of concrete is very important if high quality is to be obtained. The full

effectiveness of water curing requires that it be a continuous, not intermittent,

operation. Protection of the newly placed concrete against freezing is important.

When freezing temperatures may occur, enclosures and surface insulation should also

be required. One of the most important factors associated with protection of concrete

is advance preparation for the placement of concrete in cold weather.

d. Control testing during construction

Control testing is carried out at regular intervals as concreting progresses. They

comprise both testing during the fabrication process (with aggregates control and

fresh concrete testing), and testing on hardened concrete, based on samples and cores

taken from the placed concrete. Among the properties to be measured are the

compressive strength, the tensile strength, as well as various characteristics such as

density, and dynamic and static moduli of elasticity.

Specific tests are carried out every 300 to 500-600 m3

of placed concrete, or as a

function of the number of lifts to be concreted per day/per team, and thus as a

function of the dam's dimensions.

The test results (together with their variability) are compiled for each concrete type.

The variability of the results makes it possible to assess the safety margin with respect

to the effective stresses in the structure. If necessary, the cement content is adapted to

correct the mechanical strength. In this process, testing of the aggregates, the fresh

concrete and the seven-day compressive tests are of the utmost importance so that any

possible defect or decrease in production quality can be discovered as soon as

possible.

In dam construction, compressive strength has been the parameter most often monitored on

hardened concrete. Testing is usually; carried out in a site laboratory after 7, 28, 90 and

365 days. Some samples are tested after they have been kept for a longer period of time.

Page 24: Dam Engineering GRV BUTT

32

On average, strength increases by approximately 1 N/mm2

for every additional 10 kg/m3 of

cement content, which confirms the linear relationship between the water – cement (W/C)

ratio and strength, assuming constant water content (thus constant consistency):

,/6072)90( CWc in N/mm2

e. Cracking

Concrete cracking is a very complex process. On the one hand it is related to the

properties of the concrete itself, and on the other to the geometry of the structure or

of parts of the structure. It is necessary to distinguish the origin of the cracks and

their propagation.

It is known that the origin of the cracks is micro cracking caused by the prevented

shrinkage of the cement paste in the concrete mass. Temperature and moisture

gradients then induce microscopic cracks. During concrete placement, temperature

gradients develop in massive elements. These induce cracks if adequate preventive

measures are not taken. This type of cracking occurs within days, or sometimes

weeks, after the concrete has been placed. In general, these are surface cracks, which

often propagate in a vertical plane.

The volumetric changes of concern are those caused by the temperature drop from the

peak temperature, occurring shortly after placement, to the final stable temperature of

the structure. A degree of control over the peak temperature can be attained by

limiting the placing temperature of the fresh concrete and by minimizing the

temperature rise after placement. The placing temperature can be varied, within limits,

by precooling measures that lower the temperatures of one or more of the ingredients

of the mix before batching.

Precooling of aggregates and the use of low-heat cements, reduced cement, and

pozzolans are normally adopted as temperature control measures for gravity dams

containing no longitudinal joints.

Concrete loses some of the mixing water after removal of the framework. A water content

gradient then develops. If the resulting shrinkage is excessive, cracking can be

expected. In massive elements, desiccation can also create a state of stress that may lead to

internal cracks of random orientation.

Cracking during normal operation can have various causes: unfavourable geometry,

poorly executed construction joints, temperature distributions within the dam at the end of

construction, at closure time, or during impounding, and so on. Concrete cracking often

begins during the construction phase, and then develops.

f. Creep and shrinkage

As dams age, and in particular dams with annual operating cycles, one can often

observe year after year an increase in the maximum and minimum displacements,

both at full and low water levels. Such a trend in the annual maximum and minimum

deformations can reach several centimetres and is often still noticeable after several

years of operation.

Page 25: Dam Engineering GRV BUTT

33

Shrinkage, as one usually understands it, corresponds to a volume loss due to the

evaporation of the water not chemically bonded (drying shrinkage). It partly depends

on the concrete temperature, but is mostly related to the moisture gradient that exists

between the core of the structure and its surface.

Creep is a long-term deformation of concrete under constant loads. Concrete is a

visco-elastic material, which means that mechanical loading leads to both a quasi-

instantaneous deformation and a deferred deformation (creep). Creep develops quite

quickly at the beginning and then slows down. It keeps affecting the structure for

several tens of years after loading. Younger concrete undergoes a much stronger

specific creep than older concrete. After 15 or 20 years, creep-induced deformations

are in the order of two to four times larger than instantaneous deformations.

The respective influences of shrinkage and creep on a dam are not easy to distinguish.

Their combined effects are usually manifested by a contraction of the structure. Such

contraction has fewer consequences on gravity dams than on thinner structures like

arch or buttress dams. On such structures, the compressive stresses are higher and

shrinkage occurs on the particularly stressed superficial zones.

2.3. STRUCTURAL ANALYSIS

2.3.1. INTRODUCTION

Under reliably specific conditions the analysis methods available permit an accurate

assessment of the complete behavior of the dam and its foundation. The modern

computational tools can be involved to provide a reasonable answer for most

problems encountered. However some traditional methods of analysis of concrete

dams, which have been used with success in the post, remain in current use.

A simple two dimensional linear static analysis of a concrete gravity structure is based

intrinsically on the assumption that the shear resistance is uniformly distributed across

the potential sliding surface and that the normal stresses on that surface vary linearly

from upstream to downstream. An analysis of the same structure using linear static

finite element method will demonstrate significant stress concentrations at the heel

(tension) and at the toe (compression) of the dam. With the adoption of the "no

tension" criterion the finite element method will lead to a more conservative design.

Consequently, any method of the analysis has to be directly related to the basic

criteria.

Stress analyses of gravity dams fall into two classifications-those analyses based on

trapezoidal law (or gravity method) and those based on the finite element method.

Both methods are based on the assumption that a straight gravity dam is comprised of

a number of vertical elements of unit length, each of which carries its load to the

foundation without any transfer of the load from or to adjacent vertical elements.

Specific assumptions are made for each method and will be presented in the

corresponding sections.

An adequate assessment of sliding stability must account for the basic structural

behavior, the mechanism of transmitting compressive and shearing loads to the

Page 26: Dam Engineering GRV BUTT

34

foundation, the reaction of the foundation to such loads, and the secondary effects of

the foundation behavior on the structure. Sliding stability of concrete gravity dams

can be adequately assessed by using a limit equilibrium approach.

2.3.2. LOADS

Dead Load

The magnitude of dead load is considered equal to the weight of concrete plus

appurtenances such as gates and bridges. For preliminary design the unit weight of

concrete ( b ) is assumed to be 24 kN/m3. For final design the unit weight of concrete

should be determined by laboratory tests.

Reservoir and Tailwater

The hydrostatic pressure at any point on the dam faces is equal to the hydraulic head

at that point (h) times the unit weight of water ( = 10 kN/m3):

hp (2.1)

For the reservoir water load the normal design elevation is the highest elevation that

water is normally stored. Maximum design reservoir elevation is the highest

anticipated water surface elevation that occurs in conjunction with the routing of the

inflow design flood through the reservoir. For preliminary design the maximum

reservoir level can be assumed to be at the crest level.

The tailwater elevation should be the minimum that can be expected to occur with a

particular reservoir elevation.

In the case of a triangular profile as in Figure 2.10,a the water load is divided in

vertical and horizontal components:

Fig. 2.10. Reservoir and tailwater loads:

a. water pressure on faces; b. actual pressure at the upstream heel:

1- linear distribution; 2- reduced pressure

Page 27: Dam Engineering GRV BUTT

35

- horizontal upstream (reservoir) load:

20

2

1HP (2.2)

- horizontal downstream (tailwater) load:

2'0

2

1hP (2.3)

- vertical upstream load:

21

2

1HPv (2.4)

- vertical downstream load:

2'

2

1hPv (2.5)

where 1 is the upstream face slope ( 1 - Horizontal: 1 vertical) and is the

downstream face slope ( 1 - Horizontal: 1 Vertical).

If the dam is deeply embedded into foundation rock (Fig. 2.10,b) the hydrostatic

pressure may decrease under the rock levels. The m coefficient (m<1) depends on the

rock characteristics and the concrete-rock contact. In the engineering practice the

pressure reduction is neglected in spite of existing field data that support the pressure

diagram correction.

Uplift

Hydrostatic pressure from reservoir water and tailwater act on the dam and occur

within the dam and foundation as internal pressures in the pores, cracks, joints and

seams. The component of internal pressure acting to reduce the vertical compressive

stresses in the concrete on a horizontal section through the dam or at its base is

referred to as uplift. Internal hydrostatic pressures and uplift should be used for

analyses of the foundation, the dam body and overall stability of the dam at its contact

with the foundation.

In the hypothetical case of a perfect homogeneous dam without joints and drains the

uplift distribution on a horizontal section of the dam is assumed to vary linearly from

full hydrostatic head at the upstream face to zero or tailwater pressure at the

downstream face. The same distribution of uplift pressures is assumed on dam

foundation contour (Fig. 2.11). When drains are provided in the dam body the internal

pressure should be modified in accordance with the size, location and spacing of the

drains.

The internal pressure distribution through the foundation is dependent on rock

porosity, jointing and faulting and to a large extent on the grout curtain effects, and

the drainage system. Determination of the internal pressures can be made from flow

nets computed by numerical methods, modeling the effects of drainage and grouting

curtains. The numerical model has to include the jointing, faulting, variable

permeability and other geologic features.

Page 28: Dam Engineering GRV BUTT

36

Fig. 2.11 Theoretical distribution of uplift on foundation area

For preliminary design purposes some conventional uplift distributions are assumed

as illustrated by Figure 2.12. The uplift diagram is made from a straight - line drop

from reservoir level at the heel of the dam to a fraction of the difference in head

between reservoir and tail water along the line of drains and then another straight-line

drop to tailwater at the downstream toe. The value of that fraction is in the range of

0.25 to 0.50.

Fig.2.12. Uplift assumptions – tail water

If there is no tailwater, the downstream end of the uplift diagram is zero at the

downstream toe. The pressure is assumed to act over the entire foundation area

(Figure 2.13).

Maximum tail water

Page 29: Dam Engineering GRV BUTT

37

Fig. 2.13. Uplift distribution – no tail water

The reduction is based on observations on existing dams. Figure 2.14 showing the

uplift pressure gradients measured at Hiwassee dam renders evident that the actual

pressures at the drains fall well below the present design assumptions.

Fig. 2.14. Uplift pressures at the base of Hiwassee dam

In some European countries the uplift distribution for preliminary design is assumed

to vary linearly from a fraction m (m<1) of the headwater pressure at the dam heel to

the tailwater pressure at the downstream toe (Figure 2.15).

Page 30: Dam Engineering GRV BUTT

38

Fig. 2.15. Uplift distribution – simplified linear diagram:

a. with tail water; b. without tail water

The uplift force on the dam foundation is:

with tailwater BhHmhU2

1 (2.6)

without tailwater BHmU2

1 (2.7)

The m coefficient is called the uplift reduction coefficient or uplift coefficient.

Conservative values for uplift coefficient are in the range of 0.70...0.80. The current

practice recommends different m values according to the dam height and the rock

foundation characteristics. For large dams provided with grout curtains and drainage

wells the m values are defined by the foundation rock types:

m = 0.75 - stratified and fissured rocks;

m = 0.67 - slightly fissured rocks;

m = 0.50 - sound, homogeneous rock

In the design criteria it is suggested that the safety of large dams should not depend on

the proper functioning of grout curtains and drains. It must be remembered that large

dams may be in existence for hundreds of years. Drains have been known to clog.

Dams should still have a large margin of safety if they do. A conservative assumption

regarding the uplift distribution provides this reserve.

Silt pressure

In some reservoir finely divided silt or clay is deposited in substantially horizontal

layers against the upstream face of the dam. These deposits are produced by flows of

muddy water along the bottom of the reservoir and are to be distinguished from the

deltas of silt and sand that form in the upstream end of the reservoir. It is usually

assumed that the material is under complete saturation and is completely fluid.

Page 31: Dam Engineering GRV BUTT

39

Horizontal silt pressure is equivalent to that of a fluid weighing 14 kN/m3. Vertical silt

pressure is determined as if silt were a soil having a wet unit weight of 19.2 kN/m3,

the magnitude of pressure varying directly with depth.

More precise calculations can be made by determining the horizontal component of

the silt load from the Rankine formula, neglecting cohesion:

sin1

sin1

2

2ss

silth

P (2.8)

where s is the unit weight of the silt;

sh is the silt depth;

is the internal angle of friction of silt material.

Ice load

Ice pressure is created by thermal expansion of the ice. The pressure values are

dependent on the temperature rise of the ice, the thickness of the ice sheet, the

coefficient of expansion, the elastic modulus and the strength of the ice.

A significant factor is the rate of temperature rise. If this rate is reduced the plastic

deformation of ice attenuates the pressures acting on the dam face. In several

European countries the ice pressure is evaluated by using Royen's relationship:

3 2119 tg

n

tgtghgPice kN/m (2.9)

where hg is the ice thickness (m);

tg is the maximum temperature rise of ice sheet during n hours.

The ice temperature rise tg is evaluated in terms of air temperature rise in the same

time interval at ( tg = 0.35 at ).

Not all dams are subjected to ice pressure and based on the above factors an

allowance for ice pressure may be decided. An acceptable estimate of ice load to be

expected on the face of a structure may be taken in terms of the ice depth:

Ice depth (m) 0.5 0.7 1.0 1.20 1.50

Ice load (kN per linear m) 70 100 150 200 280

Ice load may be also created by the wind rag of floating ice. For large dams this ice

load is negligible. However, for thin, flexible structural element and especially for

gates the dynamic effect of floating ice can be important.

Page 32: Dam Engineering GRV BUTT

40

Temperature

Seasonal temperature changes of air and reservoir water propagate within the dam

concrete only in the first 5...6 m from the faces. The daily air temperature variation

propagates in a very thin concrete layer (20...30 cm) from the dam face. Comparing

with the very large concrete mass of the dam the affected zones are to small to induce

significant volumetric changes in the dam. Even if some volumetric changes are

larger, the contraction joints allow free deformation and the horizontal thrust is

negligible.

However, secondary stresses can occur at the faces of the dam and around openings

(drainage galleries, valve chambers) due to temperature differentials. These

temperature differentials are caused by differences in the temperature of the concrete

surfaces due to ambient air and temperature variation, solar radiation and air

movement in openings. These secondary stresses may produce cracks that could lead

to progressive deterioration.

Wave pressure

The upper portions of the dam are subject to the wave action, that induces additional

water pressures. The wave pressure depends on the dimensions of the wave (heights

and lengths) that in their turn depend on the extent of the water surface and the

velocity of the wind, among other factors.

For a gravity dam with a vertical or slightly inclined upstream face (see Figure 2.16)

the wave pressure load may be calculated as:

Fig. 2.16. Wave pressure:

a. pressure distribution; b. wave characteristics:

1- dam; 2- normal water level; 3- increased water surface

Page 33: Dam Engineering GRV BUTT

41

20

2

12

2

1HaHhhHPwave (2.10)

where:

h2 is the wave height;

0h is the average heightening of water level during the wind action;

a is wave pressure at the dam heel.

The wavelength and the wave height 2h are usually evaluated on the basis of

empirical formulas in terms of wind velocity (v) and the "fetch" or straight length of

water subject to wind action (F). The increased height of water surface 0h is then

calculated in terms of wave characteristics:

2

2

2

)2( 2

0H

htch

h (2.11)

or in terms of primary parameters:

H

Fvh

460

10 (m) (2.12)

where H is the reservoir depth, is the wind velocity (km/h) and F is the fetch

(km).

The wave pressure at the dam heel is approximated as

2

2

2

Hhc

ha for shallow reservoirs

0a for deep reservoirs

The wave pressure is a significant load for small dams with dam height less than 20 m

where its value may reach up to 12...15% of the water load. For very high dams (dam

height larger than 100 m) the wave load is much lower but is still a 2.5...4% of the

water load.

Seismic Loads

The problem of the earthquake response analysis of concrete dams is the subject of a

distinct chapter of this book. Under this section the earthquake forces are treated

simply as static forces to be combined with the hydrostatic pressure and gravity loads.

In representing the effects of ground motion - transverse to the axis of the dam - by

static lateral forces the dynamic response of the dam-water-foundation rock system is

neglected. Two types of static forces are included i.e. forces associated with the mass

of the dam and water pressures in addition to the hydrostatic pressure.

Page 34: Dam Engineering GRV BUTT

42

The earthquake intensity is characterized by the earthquake acceleration or by seismic

coefficient that is the ratio between the acceleration (aE) and the gravity a=aE /g.

Most dams in seismically active regions have been designed for an acceleration of

one-tenth gravity, or a = 0.1. According to different national standards, the seismic

coefficient depends on the seismicity of the site and ranges between 0.05 and 0.14.

The intensity of the inertia force associated to the mass of the dam is found from the

equation:

Gaag

GC Eg (2.13)

where gC is the earthquake force and G is the weight of the dam body (Figure 2.17).

The inertia force gC may by assumed acting through the center of gravity of the dam

or block in a downstream direction for horizontal acceleration and in an upward

direction for vertical acceleration.

The increased water pressure due to the inertia of the water in the reservoir can be

represented by a diagram of the form presented in figure 2.17.

Fig.2.17. Earthquake forces – static approach;

Increased water pressure represented by: 1- an ellipse; 2- a parabola

According to Westergaard, whose solutions are most frequently applied, the true

equation of the increased water pressure may be expressed by an ellipse or by a

parabola without appreciable error. The resulting equations are:

- for parabola equivalence:

HzCap pc (2.14)

Page 35: Dam Engineering GRV BUTT

43

3

2

6

/

10

75.71

8170mN

T

H

C p (2.15)

2

3

2HCaC pa (2.16)

- for ellipse equivalence:

zHzCap ec 2 (2.17)

)/(

10

75.71

6540 3

2

26

mN

T

H

Ce (2.18)

2

4HCaC ea (2.19)

where:

cp is the additional unit water pressure;

H = dam height or the depth of water acting on the dam;

z = current vertical coordinate;

pC and eC are factors depending on the height of the dam (H) and the

earthquake period (T);

aC = the additional total water pressure.

For medium dam heights the pC and eC factors may be approximated as pC =

0.830 and eC = 0.660.

Fig.2.18. Reduction coefficient K in terms of upstream slope 1

If the upstream face is inclined the additional water pressure normal to the inclined

face is lower than the one acting on a vertical face:

zHCaKp pc (2.20)

Page 36: Dam Engineering GRV BUTT

44

where K is a reduction coefficient that depends on the upstream face slope 1 (see

figure 2.18).

2.3.3. STRESS ANALYSIS - TRAPEZOIDAL LAW

Assumptions

As it was mentioned at the beginning of this section, the stress analysis based on

trapezoidal law (or gravity method) is based on the assumption that a straight gravity

dam is comprised of a number of vertical elements of unit length, each of which

carries its load to the foundation without any transfer of the load from or to adjacent

vertical elements. Specific assumptions are made for this method:

(1) The concrete in the dam is a homogeneous, isotropic, and uniformly elastic

material.

(2) There are no differential movements, which occur at the dam site due to

water loads on the reservoir walls and floors.

(3) All loads are carried by the gravity action of vertical, parallel-side

cantilevers, which receive no support from the adjacent elements on either side.

(4) Unit vertical pressures, or normal stresses on horizontal planes, vary

uniformly as a straight line from the upstream face to the downstream face.

(5) Horizontal shear stresses have a parabolic variation across horizontal

planes from the upstream face to the downstream face of the dam.

Vertical stresses - triangular profile

Only the more conventional analysis for a vertical section having the width of 1 m is

considered in this section. A simplified triangular profile is considered. The active

loads are the reservoir hydrostatic pressure, the dead load and the uplift (Figure 2.19).

The internal pressure (uplift) distribution, both within the dam body and through the

dam-foundation contour, is assumed to vary linearly from a fraction m of the head

water pressure to zero at the downstream end of the horizontal section.

Fig.2.19. Loading system - simplified triangular profile

Page 37: Dam Engineering GRV BUTT

45

For a horizontal section located at the depth z the vertical normal stresses at the

upstream and downstream faces are computed by the eccentric compressive formula:

W

M

A

Vupdw

(2.21)

where:

V is the algebraic summation of all active vertical forces;

M - the overturning moment taking into consideration all the

forces that act on the section;

A - area of the section;

W - modulus of strength.

The loads and their moment arms referred to the center of the gravity of the section

are (see figure 2.19):

- Dead load of the downstream prism:

2

2

1zG b arm:

6)3( 1

z

- Dead load of the upstream prism:

2

112

1zG b arm:

6)3( 1

z

- Horizontal reservoir load:

20

2

1zP arm:

3

z

- Vertical reservoir load:

21

2

1zPv arm:

6)3( 1

z

- Uplift (internal pressure) force:

21 )(

2

1zmU arm:

6)( 1

z

The resultant vertical force is:

2

1112

1zmUPGGV bv (2.22)

The resultant moment acting on the center of the section is:

Page 38: Dam Engineering GRV BUTT

46

32111

21

2 2312

1zmM b (2.23)

The section area is zA 1 and the corresponding area modulus is

221

6

1zW .

By replacing the above-evaluated terms in the Equation 2.21 the vertical normal

stresses up

at the upstream heel and dw at the downstream toe can be evaluated by

the formula:

zm

m

b

b

up

dw

}23

2

1

2

111

2

1

2

11

2

12

1 (2.24)

For the empty reservoir case 0 and the normal vertical stresses are:

zz bdwbup

1

1

1

(2.25)

Analyzing the normal stress formulas one can notice that:

- under both hypotheses - full and empty reservoir - the stresses depend

linearly on the depth z;

- maximum normal vertical stresses act on the upstream heel for empty

reservoir condition and on the downstream toe of the dam for full reservoir condition,

where z = H.

For a dam with a vertical upstream face the vertical normal stresses are:

- full reservoir:

zzm dwbup 22)( (2.26)

- empty reservoir:

zbup 0dw (2.27)

In the case of a real dam - with crest thickness, batter on the lower part, variable

downstream slope etc. the same general formula (2.21) has to be applied. The section

characteristics A, Wup and Wdw are properly evaluated. According to the specific

situation other significant loads can be included in the analysis. The uplift pressure

may have any conventional distribution (see figures 2.12 to 2.15).

In important structures, it may be desirable to determine the intensities and directions

of the principal and maximum shearing stresses at various points throughout the dam

section. The usual procedure in analyzing these stresses is to divide the section into

elementary prisms and consider each prism to be held in equilibrium by the stresses

Page 39: Dam Engineering GRV BUTT

47

acting on it. By starting with the vertical normal stresses acting on the upper and

lower surfaces of each elementary prism, it is possible to determine be successive

integration the vertical and horizontal shearing stresses intensities, the horizontal

normal stresses, and the first and second principal stress.

Principal and shearing stresses at the dam faces

Once again the conventional triangular profile is considered (figure 2.20). The

upstream and downstream face directions are directions of principal stress since the

shearing stress is zero along them. The directions normal to the faces are also

directions of principal stress.

Fig. 2.20. Principal stresses at the dam faces

To evaluate the principal stresses the first step is to separate from the dam face

triangular elements shown in figure 2.20. Each element is then assumed to be held in

equilibrium by the normal and shearing stresses:

upstream triangle: 12

112 sincospz (2.28)

downstream triangle: 22 sinz (2.29)

The principal stresses may be computed readily from the equilibrium equations as:

upstream: 21

211

2

121 1

sinpctgp z

z (2.30)

p2

downstream: 01

2

122 1

sinz

z (2.31)

The principal shearing stress:

21max2

1

Page 40: Dam Engineering GRV BUTT

48

has the values:

upstream: 2max 1

2

1pz (2.32)

downstream: 2max 1

2

1z (2.33)

Based on the above relationships it is important to note that:

- the principal stresses at the upstream face may become tensile stress when 21

211 pz ; in order to avoid the concrete cracking a more gentle slope of the

face is required in such cases;

- the maximum principal stress develop at the downstream face; at the

downstream toe, where z = H, the stress value is up to 2.5 times larger than the

maximum hydrostatic pressure;

- a change of the downstream face slope near the dam toe is recommendable in

order to reduce the maximum compressive stresses (figure 2.21);

Fig. 2.21. Local curvature of the downstream slope

- the local curvature of the downstream face near the dam toe is beneficial

since the principal shearing stresses (Equation 2.33) may reach quite large values.

Mention should be made that larger shearing stresses are allowable provided these are

not in the presence of tension.

2.3.4. STRESS ANALYSIS - FINITE ELEMENT METHOD

The structural behavior of high gravity dams, as determined by field measurements,

leads to the conclusion that the classic trapezoidal law produces stress patterns, which

do not even remotely resemble those obtained from field measurements.

Figure 2.22, for example, shows a comparison of measured and calculated stresses on

the foundation of Hiwassee dam under both low and high reservoir levels. Figure 2.23

shows foundation pressures under the Shasta dam measured when reservoir was

nearly full. The maximum pressure was measured at a point close to the center of the

base. The sharp reduction in pressures near the downstream face indicates that about

the downstream third of the concrete is taking a relatively minor part of the load.

Page 41: Dam Engineering GRV BUTT

49

Fig. 2.22. Hiwassee dam – stresses on the foundation

Fig. 2.23. Foundation pressure measured for Shasta dam

Page 42: Dam Engineering GRV BUTT

50

It will be noted that the upstream half of the base takes a greater percentage of the

load than that evaluated by the trapezoidal distribution. It will also be noted that the

maximum stress do not occur at the downstream toe and that there may be tension

instead of compression at the upstream heel.

The finite element method is an important advance in stress analysis of dams. By this

method the dam and its foundation are divided into elementary, contiguous finite

elements, and the elastic properties of each part of the dam – foundation system are

properly specified. It is not the object of this section to present into details the method.

Some general considerations are made in the context of gravity dam analysis.

Assumptions regarding loading conditions, initial stress, etc., have to be considered

before an analysis is carried out. In addition, after the results become available some

criteria concerning their interpretation will be needed.

The foundations of the dam are an integral part of the structure, and their

deformability can be included in the analysis without difficulty. As the stresses caused

by the dam within the foundations "dissipate" rapidly it is possible to use larger

elements where stress gradients are less significant. .

Initial stresses due to weight and various kinds of tectonic action exist within the

foundation rock prior to the construction of the dam. It is a general rule that these

stresses are estimated (or computed) before the interconnection with the dam. This

means that no gravity loads within the foundation should be applied to the analysis of

the dam-foundation complex. The initial gravity, tectonic, and pore-water-pressure

stresses are obviously simply superposed on the additional effects of dam stresses.

In the analysis it is important to know at least the approximate elasticity constants of

the rock material. With the finite-element process no difficulty arises in dealing with

nonhomogeneity or anisotropy, as will be.

In the dam itself it is often accurate enough to consider the gravity as if it were

applied externally to the completed, weightless structure. This assumption is

manifestly not true, and for special cases the step-by-step construction process can be

followed in the finite-element analysis. The approximation of applying gravity to the

whole structure gives results, which do not differ by a large amount from those of the

more exact process.

The finite element method may be used for evaluating the piezometric levels resulting

from the flow through the dam body and the underlying rock. The basis of these

calculations is the Darcy law. Anisotropic permeability resulting from the orientation

of the discontinuities can be taken into account. Variations in permeability induced by

the foundation treatment – grouting curtain and drainage system – can be easily

modeled.

Temperature effects on stresses are treated easily in the general formulation as simply

one of the load systems. The two types of temperature distribution which arise (1)

from external causes such as seasonal cycles and (2) from heat of hydration present

somewhat different problems. In the first case a reasonably elastic behavior of the

concrete can be assumed without serious error. The second acts on the concrete during

Page 43: Dam Engineering GRV BUTT

51

its early life when creep effects are most serious. While again the solution of such

creep problem is possible by the finite-element method, it is subject to certain

uncertainties and indeed.

It is usually no more expensive to obtain separate solutions for each load and obtain

totals by automatic addition. The separate effects can thus be studied individually or

in sum. All programs should be capable of computing the principal stress system

acting at various points of the structure. It is these stresses, which will finally decide

whether suitable safety is ensured, and decisions on modification of the design will be

taken on that basis.

One of the basic criteria usually introduced is that no tensile stress should be present

in the final structure. Certainly this must apply in the rock, which is usually fissured,

and probably it is desirable to extend this to the concrete mass, where cracking could

have occurred by thermal or shrinkage action. However, in elastic stress analysis it is

almost impossible to design a dam structure with no tension in any part. The criterion

must therefore be modified. A reasonable approach is to consider the criterion as

applicable after a limited amount of cracking has occurred

While the elastic stress analysis gives a good picture of working stresses and makes it

possible to estimate the least factor of safety against failure, the true failure load can

only be obtained by the introduction of plastic nonlinearities into the analysis. Also,

the introduction of cracks into finite-element analysis presents little difficulty.

In order to demonstrate the power of the finite element analysis in the field of dam

safety evaluation an example is presented in the followings. A complete stress and

seepage analysis has been performed for Eder dam in Germany to verify the

effectiveness of the rehabilitation measures.

The Eder dam is a gravity dam with a total height of 48 m, relatively thin (figure

2.24). Achieving stability was possible either by restricting the water level, or by

taking other steps to improve the dam condition. Because of economic considerations

a solution involving the anchoring of the dam to the rock by installing permanent

anchors was given preference.

To verify the stability calculations several two-dimensional FE analyses were carried

out. The finite element discretization includes a 300 m-long and 159 m-high section

from the foundation, as well as the 48 m-high dam body (figure 2.24,b). Isotropic

behaviour in the elastic range was assumed for the dam body (E=7500 MPa) and for

the rock foundation (E=5000 MPa). The finite element mesh has 2757 isoparametric

elements, each with eight nodes, and 2881 nodal points. The fine meshed modelling

of the dam on the upstream and downstream sides can be seen. The object of this is to

calculate as accurately as possible the stresses in the area with large stress gradients.

The same discretization was used for the seepage analysis as well as for the stress

analysis. According to the results of the field tests the foundation rock was divided

into three zones with respect to permeability. The permeability of the grout curtain

zone was assumed as one order of magnitude lower than the average rock

permeability.

Page 44: Dam Engineering GRV BUTT

52

Fig. 2.24. Eder dam finite element analysis:

a. dam cross section; b. finite element mesh

The equipotential lines resulting from the seepage flow analysis are shown in figure

2.25. The draining effect of the inspection gallery and the effectiveness of the grout

curtain are rendered evident. The uplift pressure distribution along the dam foundation

a. MAIN CROSS SECTION

b. DISCRETIZATION

Page 45: Dam Engineering GRV BUTT

53

contact show that downstream of the grout curtain the pressures are 40% of the

hydrostatic pressure, thus confirming the assumptions used in stability analysis.

Fig. 2.25. Seepage analysis results

The seepage forces from the above-presented calculation, together with the loads from

the self-weight and anchors produced the principal stress distribution shown in figure

2.26. Because of the vertical tensile stresses that cannot coop with the actual strength

of the bedding rock, horizontal cracks occur at the heel of the dam on the upstream

side. Mention should be made that the stresses calculated by the classic trapezoidal

law led to the conclusion that this zone experiences compression. However, the area

covered by the horizontal cracks is small, and because they do not reach the injection

zone the watertightness of the grout curtain is not affected.

UPLIFT PRESSURES ON FUNDATION

EQUIPOTENTIALS

Analysis results

Measured values

Grout curtain

Page 46: Dam Engineering GRV BUTT

54

Fig. 2.26. Principal stresses under full reservoir hypothesis.

2.3.5. SLIDING STABILITY

Assumptions

An adequate assessment of sliding stability must account for the basic structural

behavior, the mechanism of transmitting compressive and shearing loads to the

foundation, the reaction of the foundation to such loads, and the secondary effects of

the foundation behavior on the structure.

Sliding stability of concrete gravity dams can be adequately assessed by using a limit

equilibrium approach. Assumptions and simplifications are listed below:

- A two dimensional analysis is accepted involving a unit length of the block;

- Only force equilibrium is satisfied. Moment equilibrium is not used;

- Considerations regarding displacements are excluded from the limit

equilibrium approach;

- Analyses are based on assumed plane failure surface defined by the dam-

foundation contour. The calculated factor of safety will be realistic only if the

assumed failure mechanism is kinematically possible;

- A linear relationship is assumed between the resisting shearing force and the

normal force acting along the failure surface.

- Shear strength is a combination of internal friction that varies with the

normal compressive stress and cohesive strength.

Sliding stability evaluation based on friction coefficient

The sliding of a gravity dam along its foundation contour takes place when the total

horizontal force H overpass the resisting force that can be mobilized at the contact.

Page 47: Dam Engineering GRV BUTT

55

The maximum resisting force can be assumed that is given by the product between the

total vertical load V acting normal to the foundation and a so called "concrete-rock

friction coefficient" f. Sliding stability condition is:

H < f V (2.34)

This condition can be expressed in terms of the Factor of Safety (FS) as:

FS H = f V (2.35)

The concrete-rock friction coefficient includes both the cohesion and friction effects.

The in situ tests performed in order to evaluate the shear strength at the concrete-rock

contact define the dependence between the shear stress along the block-rock

foundation contour and the normal vertical stress acting on the same contour. The

same field tests allow for relating the shear strength to the shear displacement along

the sliding contour. The test results can be processed assuming the Mohr-Coulomb

failure criterion:

tannc (2.36)

and thus the mobilized shear strength parameters c ( ) and tan ( ) are defined. That

is, for an allowable shear displacement one can evaluate the available cohesion c

and the available internal friction tan .

The relationship 2.34 corresponding to the limit equilibrium ( H = f V ) divided by

the foundation area becomes nf . Comparing this expression to the Equation

2.36 one can relate the concrete-rock friction coefficient to the available shear

strength parameters:

tann

cf (2.37)

For large dams, where the normal vertical stress n acting on the foundation contour

is rather large, the cohesion term nc / is neglected. For preliminary calculations the

concrete-rock friction coefficient is assumed to be:

f = 0.65 ... 0.70 for sound eruptive rocks

f = 0.50 ... 0.65 for sedimentary rocks

Corresponding to this simplifying assumption the required factor of safety is

moderate: FS = 1.15 ... 1.30 for normal static loading conditions and FS = 1.00 ... 1.05

for seismic loading conditions. FS is selected in terms of the class of importance of

the dam.

Sliding stability evaluation based on cohesion and friction

Although, generally, the foundation surface appears as horizontal in the transverse

(upstream-downstream) direction the actual irregularities of the foundation rock cause

a marked apparent cohesion. The cohesion term of the shearing strength is no longer

negligible. The sliding stability condition is:

Page 48: Dam Engineering GRV BUTT

56

AcVHFS tan (2.38)

where A is the foundation area.

Following this approach, the minimum required factor of safety is FS = 2.0 for normal

static loading conditions and FS = 1.3 for seismic loading condition. Mention should

be made that in the past the minimum required factor of safety for static loading

conditions was 4. The primary reasons for use of this conservative factor of safety

were the uncertainty in determining rock shear strength parameters and the peak shear

strength from in situ tests. In the last decades the methods of testing have substantially

improved and much better definition of mobilized cohesion and friction are now

possible.

Multiple - plane failure surface

When the dam structure is significantly embedded and some weak planes are

encountered in the foundation rock an additional sliding stability analysis is required.

The potential failure surface has to be defined based on the stratification, location and

orientation of the discontinuities and the configuration of the dam foundation.

Most frequently the failure surface is defined by two weak planes (Figure 2.27). The

assumed slide mass is divided in two wedges. The factor of safety FS is obtained from

the equations:

2

1

0

2

1

tan

tancos1

iii

iiiiii

i

i

VP

UVAcn

FS (2.39)

where

i

iiFSn

i 2tan1

tantan1

1

(2.40)

ic - cohesion on the sliding contour (segment) i;

iU - uplift force acting under the wedge i ;

iA - area (length) of the segment i;

iV - all applied vertical forces (body - dam and rock - and surcharge)

acting on the wedge i;

i - angle between the inclined plane (segment) i and the horizontal

( 0 for upslope sliding and 0 for downslope sliding);

i - angle of internal friction along the segment i.

Page 49: Dam Engineering GRV BUTT

57

Fig. 2.27. Multiple planes sliding mechanism

Equation 2.39 is implicit in FS since n is a function of FS. Therefore, the

mathematical solution of Equation 2.39 requires an iteration procedure. An initial

estimate of FS is inserted into the n term and a first FS is calculated. The calculated

FS is then inserted into the n term and the process is repeated until the calculated FS

converges with the inserted FS.

Mention should be made that Equation 2.39 is similar to the generalized method of

slices but in order to develop a simple analytic technique the vertical side forces due

to impending motion of the wedges between slices were assumed to be zero.

Therefore, although the equation satisfies complete horizontal static equilibrium,

complete vertical equilibrium is in general not satisfied.

Constructive measures

In order to increase the sliding stability of the dam some additional constructive

measures are provided that cannot be quantified. Besides the consolidation grouting

and the excavation to the firm rock the foundation contour itself may significantly

contribute to an improved stability (figure 2.28).

Fig. 2.28. Foundation contours:

a. upstream heel groin; b. irregularities along the foundation; c. inclined steps

Page 50: Dam Engineering GRV BUTT

58

Figure 2.28,a shows the effect induced by the upstream heel groin on the shear stress

distribution. Several irregularities along the foundation contour (Figure 2.28,b) create

a more even shear stress distribution and contribute to stability. If the foundation

contour is excavated following a pattern of inclined steps (Figure 2.28,c) the resultant

force under full reservoir condition is almost perpendicular to the steps thus reducing

the tangential component that induces sliding.

The beneficial effect of the upward sloped foundation, from heel to toe of the dam,

may be noticed in Figure 2.29.

Fig. 2.29. Forces acting on an inclined foundation

If is the angle of sloping, then the sliding stability condition may be expressed as:

NfTFS (2.41)

that is similar to the Equation 2.35. Since:

sincos VHT (2.42)

sincos HVN (2.43)

the actual factor of safety:

sin1

cos

sincos

H

H

FSf

fFSFS (2.44)

is increased in comparison to the factor of safety HFS of a dam with horizontal

foundation contour.

2.4. DESIGN OF GRAVITY DAM PROFILES

Design criteria

Assuming that the common profile has a triangular shape the design has to determine

the upstream and downstream slopes of the dam faces. Two main criteria are used:

Page 51: Dam Engineering GRV BUTT

59

First criterion. Under full reservoir condition the vertical normal stress at the dam

heel has to be a compressive stress. It is assumed that the dam-foundation joint is

incapable of resisting tensile stresses that may cause a slight opening of the joint.

Such an opening at the upstream heel is objectionable as it may admit full reservoir

pressure over the entire area not in compression with a further opening of the joint.

The progressive extension of the joint may be sufficient to cause failure. According to

the simplified trapezoidal law the criterion is written as:

0W

M

A

Vup (2.45)

where the notations are the same as for Equation 2.21.

Second criterion. The total frictional resistance to sliding on the foundation surface

must exceed by a safe margin the total horizontal force at the foundation level in order

to withstand the tendency to slide:

HVf (2.46)

where the notations are the same as for Equation 2.34.

The condition 2.46 may become the Equation HVf if the concrete-rock

friction coefficient f is properly corrected such as to include the cohesion and friction

effects and also the reduction corresponding to the required factor of safety.

Additional criteria are used in the design process. Among them are:

- Tension shall not exist in any joint of the dam under any condition of

loading;

- The ultimate shearing strength of any joint must exceed the total horizontal

force above the joint for all condition of loading;

- The principal compressive stresses in the dam and the foundation shall not

exceed certain allowable values.

Once the shape of the dam profile has been determined in accordance with the two

main criteria careful attention must be given to the additional criteria that can affect

the details of the design.

Dams with inclined upstream face

Assuming the load combination shown in figure 2.30: dead load, water pressure and

uplift, the two main criteria (Equations 2.45 and 2.46) may be written as:

Solving the system an equation on 1 is obtained:

(2.47)

(2.48)

Page 52: Dam Engineering GRV BUTT

60

Once the 1 value is determined the downstream slope can be determined from

Equation 2.48.

Fig. 2.30. Design of dams with inclined upstream face:

a. load combination; b. slopes variation in terms of coefficients f and m

Based on a parametric study the variation of the upstream and downstream slopes in

terms of concrete- rock friction coefficient f and uplift coefficient m is presented in

Figure 2.30. The upstream slope is strongly dependent on the uplift value, while the

downstream slope is more significantly defined by the friction coefficient.

If the dam is located in a seismic area the load combination includes the seismic

forces Ca and Cg as it is shown in figure 2.31:

Fig.2.31. Load condition and slope variation – seismic loads included:

a. loads; b. slope variation in terms of earthquake intensity

(2.49)

Page 53: Dam Engineering GRV BUTT

61

The design criteria (Equations 2.45 and 2.46) are written as:

where mb

i

b .

The equation in 1 becomes:

The 1 value is determined from the above Equation and then the downstream slope

is determined from Equation 2.51. The variation of the dam slopes in terms of the

seismic coefficient a (that depends on earthquake intensity) is shown in figure 2.31.

One can notice that the upstream slope becomes gentler as the earthquake intensity

increases while the downstream slope is practically constant.

Dams with vertical upstream face

The dam profile is defined by only one parameter, the downstream face slope . The

loading conditions for the two combinations: normal loads and extreme loads are

presented in figure 2.32.

Fig. 2.32. Load conditions for designing the profile with the vertical upstream face:

a. normal loads; b. extreme loads

(2.50)

(2.51)

(2.52)

Page 54: Dam Engineering GRV BUTT

62

Assuming the normal load combination, without earthquake loads, the first design

criterion – upstream compressive stress- leads to the Equation:

mb

(2.53)

The second criterion – sliding stability- is expressed by:

)( mf b

(2.54)

The actual slope value is selected as the maximum value given by the two equations.

If the seismic loads are included the two equations are:

The slope selection is made by the same procedure.

Dams with upstream batter

In order to define the profile of a dam with an upstream batter one has to determine

three parameters (Figure 2.33): downstream slope , the height of the batter and the

slope of the batter o.

Fig.2.33. Profile definition and notations:

a. profile parameters and loads; b. additional design criterion

(2.55)

(2.56)

Page 55: Dam Engineering GRV BUTT

63

Two conditions are given by the main design criteria:

- compressive stress at the upstream heel:

The two equations are solved for o that in its turn depends on :

In order to define the profile parameters an additional criterion is introduced, that are

the limiting the tensile stresses in the section x-x from figure 2.33,b. If allowable

tensile stress is ta then the condition is written as:

ta

GR

H

MM

22

6

1

0 (2.60)

where:

MR is the moment corresponding to the foundation reactions along the batter

extension;

MGo is the moment due to the dead load of the batter zone.

The final form of the condition 2.60 is:

ta

b HH

)33()(

0

2

03

00

22

0 (2.61)

Since an explicit form of the equations is not available a multiple step design is

required. In the first step a certain value for is assumed. From Equation 2.59 results

the o slope, and then from Equation 2.58 results the downstream slope . In the

second step the profile parameters are introduced in inequality 2.60. If that is satisfied

with a rational margin of safety then the profile is defined. If it is not, a new value for

is proposed. By a trial and error process the final shape of the profile is established.

BIBLIOGRAPHY

Davis, C.,V., Sorensen, E.,K. Handbook of applied hydraulics. McGraw-Hill, 1969.

Popovici, A. Water storage dams. (in Romanian), Bucharest, 1992.

Priscu, R. Hydraulic structures. (in Romanian) , Bucharest, 1974.

- sliding stability:

(2.57)

(2.58)

(2.59)

Page 56: Dam Engineering GRV BUTT

64

Raphael, J.,M. Concrete gravity dams. Development of Dam Engineering in the

United States. Kolgaard, E.,B., Chadwick, W.,L. editors, Pergamon Press, 1988.

Romanian National Committee on Large Dams. Dams in Romania, Bucharest 2000.

Swiss Committee on Dams. Concrete of Swiss dams: Experiences and Synthesis,

September 2000.

Stematiu, D. Finite element analysis of hydraulic structures. (in Romanian).

Bucharest, 1988.

USBR. Design of gravity dams, Denver, 1976.

US Corps of Engineers. Sliding stability for concrete structures. ETL 1110-2-256,

June, 1981.

Page 57: Dam Engineering GRV BUTT

65

3. BUTTRESS DAMS

3.1. INTRODUCTION

The buttress dams are a derivation from the massive gravity type with the introduction

of intermediate spaces (Figure 3.1). These spaces allow water seepage through

foundation and dam body to discharge not only downstream, but also side and

upwards into them, thus greatly reducing the uplift pressures. Theoretically, the total

elimination of the uplift forces would permit a reduction of a gravity dam's mass by

40% without reducing its stability. On the other hand the intermediate spaces facilitate

the dissipation of the heat produced during the hardening of the concrete, so the

elaborate cooling measures are seldom needed during the construction of a buttress

dam.

However, the actual reduction of the uplift pressures proved considerably smaller

because the intermediate spaces had to be closed upstream with slabs, arches or a

thickening of the buttress heads to make them contiguous. The saving in costs is even

more modest due to the fact that a buttress dam requires more formwork for the more

complicated shape and higher cement content for a slimmer structure.

Fig.3.1. Derivation of the buttress dam types from the massive gravity dam

Given the absence of uplift, more substantial savings are possible by inclining the

upstream face and thus mobilizing the vertical water load on it for sliding stability.

Also the eccentricity of the resultant force is greatly improved, since it moves from

the downstream third of the base or horizontal sections in the case of vertical

Page 58: Dam Engineering GRV BUTT

66

upstream face to almost their center in the case of a 1H: 1V upstream face inclination.

Correspondingly, the vertical stresses are almost uniform over the section, although

several times larger. Except for high dams, this is welcome because it permits a better

utilization of the strength of the construction material.

3.2. BUTTRESS DAM TYPES

Buttress dams represent that class of structures having a component that retains the

reservoir and is supported by a series of buttresses. Three major types of buttress

dams exist:

- flat slab and buttress;

- multiple arch;

- massive head buttress.

The flat slab and buttress dam is a concrete structure comprising a flat reinforced

concrete slab inclined in the downstream direction that retains water and transmits

loads from the reservoir and slab to buttresses and thence to the foundation (Figure

3.2). The slab can be simply supported at the buttress (Ambursen type), made

continuous with the buttress or suspended from corbels cantilevered from the

upstream side of the buttresses. The buttresses usually are reinforced members

although in some cases where the member is relatively massive, reinforcement has

either been minimized or omitted entirely.

Fig.3.2. A typical flat slab and buttress dam:

a. cross sections; b. slab connection with the upstream end of the buttresses:

1- construction joint; 2- struts between buttresses; 3- plinth; 4- bituminous mastic

Page 59: Dam Engineering GRV BUTT

67

The multiple arch dam is a concrete structure that comprises a series of concrete

arches, semicircular or segmental in plan, whose thrust from the reservoir and dead

load is carried out by buttresses and thence to the foundation (Figure 3.3). The arches

are usually inclined in the downstream direction and are designed with reinforcement.

They are usually supported by, and constructed integrally with, equally spaced,

triangularly shaped buttresses.

Fig.3.3. A typical multiple arch dam

The massive head, buttress dam is a structure comprised of heavy counterforts having

their upstream ends enlarged into a bulb or mushroom to full span width (Figure 3.4).

Fig. 3.4. A typical massive head, buttress dam:

a. cross sections; b. rounded heads: buttress enlargement; 2- contraction joint; 3-

water stop

Page 60: Dam Engineering GRV BUTT

68

F.A. Noetzli proposed the original design in 1925. The rounded head transmitted the

water load to the buttress in compression. Each buttress is an independent entity and is

separated from its neighbor by a contraction joint. Inserting singular or multiple

waterstops at the upstream side of the joint ensures watertightness. The expanded

upstream end of the counterfort may have a variety of shapes that approximate the

mushroom head. Usually this approximation is made with a series of straight lines that

results in a faceted configuration or diamond head.

An alternative to the massive head buttress dam proposed by Noetzli is the hollow

gravity dam, promoted by Italian engineers and standardized by C. Marcello (Figure

3.5). Contraction joints are provided for at only every second intermediate space,

resulting in a shaped buttresses.

Fig. 3.5. Typical buttress, hollow gravity – type dam:

1- drainage pipe

Marcello’s double-buttress alternative, as well as the Noetzli’s single buttress design

became less important for economic reasons, the rise of labor cost in relation to the

prices of materials and the ensuing mechanization working strongly against them.

Nevertheless, the largest hydropower plant ever built so far, namely the 12,600 MW

Itaipu scheme, has a double-buttress main dam with an unprecedented height of 196

m.

3.3. FLAT SLAB AND BUTTRESS DAM – STRUCTURAL FEATURES

The structural features of a flat slab and buttress dam are presented on the basis of an

outstanding example that is La Prele Dam designed and constructed in USA during

the period 1906 to 1908 (Figure 3.6).

Page 61: Dam Engineering GRV BUTT

69

The dam, which is 42 m high and 99 m long at its crest, consists of a reinforced

concrete slab supported by 17 reinforced concrete buttresses. Two 100 mm diameter

steel pipes serve to regulate irrigation releases. A low-flow outlet consisting of one

steel pipes is used for releases to maintain minimum streamflow. The new spillway,

which replaced an overflow crest spillway, consists of a side channel located in

natural rock which discharges through a left abutment tunnel about 75 m downstream.

The flab slab is inclined approximately 400 from the horizontal and is supported by 17

buttresses at center-to center spacing of 5.50 m. The face slab varies from 1.50 m

thick at the base to 35 cm thick at the top.

Fig. 3.6. La Prele dam: cross section and plan view

PLAN VIEW

TRASHRACK

JOINTS IN DECK SLAB

CREST

NWL

DRAINAGE

SLIDE GATE

EXISTING SLAB

SUPLEMENTAL SLAB

CROSS SECTION

Page 62: Dam Engineering GRV BUTT

70

The slab deck was made up of independent slabs between each pair of buttresses.

Each buttress was provided with a double-grooved tongue in which the slab fits. This

groove was then painted with waterproofing consisting of successive coats of asphalt

primer, waterproof paint, felt and asphalt, after which the concrete slab was placed. A

small V-groove was left where the top of the slab meets the top of the tongue, and this

was finally filled with pitch as a last precaution against water entering the joint.

Struts are provided between buttresses. Cutoff walls are provided at the upstream and

downstream ends of the base slab.

In 1971 inspection showed that about 20 percent of the original face slab thickness

had eroded away. Rehabilitation included: (1) construction of a new face slab

over the existing slab, (2) grouting the abutments and foundation underneath the dam

in the river channel area, (3) installation of rock bolts to tie the structure to the

foundation, and (4) construction of a new side channel spillway in the left abutment.

The early popularity of the flat and buttress dam was enhanced by the large savings in

materials in conjunction with cheap skilled labor available at the time. Analysis of the

concrete slab was straightforward and followed the principles of reinforced concrete

design, treating the slab as a simply supported component free to expand or contract

in the direction parallel to the dam axis (Figure 3.7). Loads were imposed and tested

for shear and moment, a slab thickness determined, and reinforcement computed

accordingly.

Fig.3.7. Types of interfaces between slab and buttresses

Early designers considered the buttress component of flat slab and buttress dams as

cantilever beams of variable cross section. The dimensions of the "beam" had to be

sufficient to avoid tension at the upstream face of the member. For higher

structures, buckling became a concern. This concern was mitigated by treating the

buttress as a bearing wall with the minimum thickness governed by column criteria.

Additional strength against buckling was gained by including struts between

Page 63: Dam Engineering GRV BUTT

71

buttresses, by pilasters or flanges constructed along the width of the buttress, or by

diaphragm walls between buttresses.

As buttress design evolved, the cantilever beam approach was supplanted by

designing these members as columns. Those features were included in the buttresses

to facilitate these design assumptions (Figure 3.8). The buttresses were divided by

contraction joints into independent "columns". Keys were included along the

contraction joint to interlock each column segment to its neighbor. Although not

completely acting as independent columns, the approach more closely approximates

theory. Other advantages of the contraction joints include better control of cracks in

the buttress due to temperature and shrinkage, reduction in reinforcement steel, and

facilitation of construction.

Fig. 3.8. Layout of contraction joints in buttress

Flat slab and buttress dams also had their disadvantages. These disadvantages

included (1) a height limitation due to increased thickness of slab and the amount of

required reinforcement when compared to other types of concrete buttress dams, (2) a

need for exceptional quality control during construction, because the relatively thin

structural components are subject to deterioration, and (3) the rapidly escalating cost

of labour and materials from 1955 onwards. Moreover, early structures have required

extensive maintenance and repairs after being in service a relatively short time. In

many instances, the durability of buttress dams in extreme climatic conditions was

being questioned.

3.4. MULTIPLE ARCH DAM – STRUCTURAL FEATURES

The structural features of a multiple-arch dam are presented on the basis of two

examples. The first one is a common thin structure. The second one is a multiple

domes structure supported by massive buttresses.

Gem Lake Dam is a thin reinforced concrete dam with a height of 34 m (Figure 3.9),

with concrete arches varying in normal thickness from 30 cm at the top to 1.10 m at

the deepest point. The buttresses vary in thickness from 0,60 m at the top to 1.40 m at

Page 64: Dam Engineering GRV BUTT

72

the deepest point. The buttress counterforts are 1.50 m wide at their tops and 3.60 m

wide at the deepest point. Double struts between buttresses have a cross section of 30

cm by 45 cm.

The upstream faces of all of the arches are covered with one more layers of gunite of

varying thickness and extent where the arches have been extensively modified or

repaired. A so-called concrete "gravity section" was placed against the intrados of

each arch up to an elevation of 15 m above the foundation ground. This massive

support of the arches was judged necessary because of disintegration of the arch

concrete due to a combination of freeze-thaw action and leaching.

Fig. 3.9. Gem Lake dam

Page 65: Dam Engineering GRV BUTT

73

In 1966, the extrados of all arches were refaced with an average of about three inches

of gunite, reinforced with heavy wire mesh. A polysulphide rubber coating was then

applied to the surface for additional protection.

Coolidge Dam consists of three egg-shaped domes supported by two massive

intermediate buttresses and the canyon walls (Figure 3.10). The dam has a structural

height of 76.2 m, a hydraulic height of 70 m, and a crest length of 280 m.

Appurtenances to the dam include two uncontrolled spillways with super elevated

chutes, an outlet works consisting of two conduits located partially in the buttress, and

a power plant located in the central dome between the two buttresses.

Fig. 3.10. Coolidge dam

CROSS SECTION

Page 66: Dam Engineering GRV BUTT

74

The domes are 1.30 m thick at the crest and have a maximum thickness of 8.20 m at

the foundation. The span of the domes from center to center of the buttresses is 60 m.

Buttress thickness ranges from a minimum of 7.90 m to a maximum of 19.60 m. The

buttresses are unreinforced except for the area near the spring line of the domes; the

domes are heavily reinforced. Two inclined contraction joints were built in each

buttress to facilitate contraction and expansion of the concrete during temperature

changes. The joints are approximately parallel to the trajectories of the minimum

principal stress.

Seepage control is afforded by a continuous double line grout curtain, which extends

beneath the structure and the spillway weir sections. Each line consists of holes 6.50

m deep on 3-m centers, staggered to produce an effective 1.70 m center-to-center

spacing along the curtain.

Flooding and seismic events have been recorded since construction of the dam, but no

significant damage has occurred to the dam structure.

Multiple-arch dams evolved at approximately the same time as the slab and buttress

dam, but at a much slower rate. Unlike the articulated Ambursen dam, multiple arches

are continuous, monolithic structures where loss of an important structure component

could lead to loss of the entire dam. For this reason, these structures normally require

a better foundation than the Ambursen type.

The arches are oriented normal to the upstream faces of the buttresses. However, due

to the inclination of the arch, the water load is not distributed uniformly from crown

to abutment, since the crown is always at a higher elevation than the corresponding

abutments.

Use of the theory of elasticity could adjust for this effect by allowing for increased

arch thickness at the arch abutment or modifying the circular shape to some other

geometric shape to some other geometric shape to achieve the same purpose.

The design for the buttresses of multiple arch dams essentially was identical to that

developed for the flat slab and buttress dam. A major innovation, initially proposed by

Noetzli was the hollow or double walled buttress. This concept featured a buttress

comprising two relatively thin walls with the interior faces of the walls stiffened by

reinforced concrete diaphragm elements. A substantial savings in concrete, and

reduction of uplift forces made for an overall cost savings. Utilization of the double-

wall buttress had a further advantage in eliminating or minimizing the need for

exterior struts between buttresses.

3.5. MASSIVE HEAD BUTTRESS DAM– STRUCTURAL FEATURES

The structural features of massive head buttress dam are presented on the basis of

several examples. The first one is Gura Raului dam built in Romania between 1973

and 1980 (Figure 3.11). The dam has a total height of 73.50 m, a crest length of 330

m, and a concrete volume of 300,500 m3.

The dam structure is divided into 22 blocks of 15 m wide. The buttresses have a

variable thickness, from 4.50 m at the dam crest to 8.0 m at the base. Also a gradual

Page 67: Dam Engineering GRV BUTT

75

thickening was provided from buttress head to the downstream face. In order to

increase the economics of the dam the upstream face slope is very large (0.57

Horizontal to 1 Vertical), to allow that a part of the sliding stability required block

weight to be replaced by the vertical water pressure on the upstream face. Inclined

construction joints, parallel to the downstream face are provided into each buttress.

Fig.3.11. Gura Raului dam

Water- tightening of the joints between the buttresses was achieved by means of a 1.5

mm thick cooper strip doubled by a PVC strip. Foundation water tightening was

achieved by a grout curtain consisting in 2 rows of grouted drillings at 1.5 m span,

with a depth of 45 m for the upstream row and 40 m for the downstream row.

Drainage underneath of the buttress heads was provided by 30 m deep drillings, two

for each block. Drainage of the rock mass is achieved by two groups of horizontal

drillings on each abutment.

The concrete in the dam was zoned, by using 240 kg cement / m 3 of concrete in the

buttress heads and a 200 kg cement / m 3 of concrete the buttresses. In order to avoid

PLAN VIEW

CROSS SECTIONS

Page 68: Dam Engineering GRV BUTT

76

cracking during the warm seasons the aggregates and the cement were cooled, the

concrete temperature at placing not exceeding 16o C.

Ben Metir dam was built in Tunisia between 1954 and 1955. The dam, with a

structural height of 71 m, has several specific constructive measures imposed by a

weak foundation rock (Figure 3.12). The current buttress presents a 14 m wide head

and a 4 m thick buttress slab. Near the foundation line the buttresses are enlarged and

become interconnected thus reducing the vertical stress transferred to the ground.

Fig. 3.12. Ben Metir dam:

a. cross section; b. contraction joint; c. buttress foundation:

1- grouting gallery; 2- drainage gallery; 3- reinforced concrete bulb;

4- bituminous mastic; 5- drainage shaft

Sliding stability of the dam was ensured by the weight of the water on the

significantly inclined upstream face and by sloping the foundation, upward from heel

to toe of the dam up to 15 % in the central zone. Due to the continuous concrete cover

of the foundation an extensive drainage system was provided by means of drainage

galleries and drainage drillings.

Figure 3.13 shows the most representative diamond-head buttress dams built

according to the modern concept. For each dam the profile and the characteristic

vertical and horizontal sections are presented. Some basic features are:

- the upstream and the downstream slopes are very similar;

- on the average, the ratio between the base width and the dam height is in

the range of 0.9 and 1.1;

- the head width varies between 14 and 18 m.

The hollowness coefficient , defined as the ratio between the total hollow spaces and

the block width, is in the range of 1/2 to 2/3. In other words, depending upon the

allowable stresses, somewhere between two-thirds and three-quarters of the volume of

Page 69: Dam Engineering GRV BUTT

77

a gravity dam is required as a counterweight to overcome overturning and shear-

friction forces.

For many dams the buttresses are provided with contraction joints parallel to the

downstream slope in order to avoid the concrete cracking induced by temperature

variation. Contiguous foundation pad is created by enlargement of the buttresses when

the foundation rock is weak.

Fig. 3.13. Representative diamond-head buttress dams

Ancipa dam is a hollow gravity dam having features typical of those dams that have

been designed and constructed principally by Italian engineers (Figure 3.14). The dam

was built during 1949 to 1952 period in Sicily. The maximum dam height is 112 m

and the 253 m wide valley is dammed by 9 blocks of 22 m wide each. The upstream

and downstream face slopes are each 4.5 horizontal to 10 vertical. The base width at

the foundation level is 0.9 of the dam height. This compares with base widths varying

between 0.75 and 0.80 of the height as commonly used for gravity dams.

Page 70: Dam Engineering GRV BUTT

78

The 7.0 m-wide internal cavity creates two buttresses, each having a maximum

thickness of 5.0 m. To increase lateral stability, flared sections at both the upstream

and downstream faces join these buttresses. The stabilizing effect of the vertical load,

introduced by providing a sloping upstream face, more than compensates for the loss

of weight that results from coring the monoliths.

The design concept eliminates the redundant concrete, provides massive unreinforced-

concrete sections, and has adequate lateral bracing for the buttresses.

Fig. 3.14. Ancipa dam:

a. plan view; b. characteristic sections:

1- massive concrete wing; 2- contraction joint; 3- drainage gallery

3.6. MASSIVE HEAD BUTTRESS DAMS– DESIGN

The design criteria used for gravity dams may also be applied to the buttress dam

type. The forces acting on a buttress dam are normally the weight of the concrete, the

reservoir water pressure, uplift pressure (pore pressure for other horizontal section

except foundation), earthquake effects and ice thrust. Mention should be made that

uplift pressure is assumed as decreasing linearly from a value equal to that resulting

from full head at the upstream face to one-tenth (or zero) of this pressure at the

downstream end of the buttress head. Pore pressures are assumed to have the same

distribution as those assumed for the uplift pressures.

Page 71: Dam Engineering GRV BUTT

79

For stress calculation the trapezoidal law can be applied, without substantial error, to

dams in the medium height range. In cases involving significant differences between

the elastic properties of the concrete and foundation materials a more exact analysis

by using the finite element method is required.

For current sliding stability analysis the shear strength is a combination of internal

friction that varies with the normal compressive stress and cohesive strength.

However, in order to simplify the preliminary design equations, the ultimate

resistance of the dam to sliding is considered as equal to the product of the total

normal force multiplied by the coefficient of static friction.

Under these simplifying assumptions the triangular profile of the dam is defined by

the upstream slope 1 and by the downstream slope . The hollowness coefficient is

previously selected based on the engineering practice, taking into account the

foundation strength. By applying as design criteria the conditions of zero vertical

stress at the dam heel and the sliding stability along the dam foundation contour the

numerical parametric calculations lead to the diagram in Figure 3.15.

Fig. 3.15. Design diagrams for a buttress dam:

a. loading conditions; b. face slopes

The upstream face slope is strongly dependent on the hollowness coefficient and

on the static friction coefficient f. The downstream slope is defined by the

upstream slope and by the static friction coefficient.

As it was mentioned before, the trapezoidal law on which the above diagrams are

based is not appropriate in cases involving significant differences between the

elastic properties of the concrete and foundation materials. In order to evaluate

how far the “conventional” linear assumption of the trapezoidal law can err, a

finite element analysis of a massive head buttress dam is reproduced, after

Zienkievicz, in the followings.

Figure 3.16 shows the dam section, its foundation in which various

inhomogeneities exist, and the division into finite elements. The heterogeneous

foundation region is subject to plane strain conditions while the dam itself is

considered as a plane stress structure of variable thickness.

Page 72: Dam Engineering GRV BUTT

80

Fig. 3.16,a. Dam cross sections

Fig.3.16,b. Finite element discretization of the dam-foundation system

Figure 3.17 shows the actual vertical stress distribution in two horizontal sections. For

the section B-B located ad mid-height of the dam the finite element stresses are not far

from the linear distribution corresponding to the trapezoidal law. A complete different

situation is encountered in section A-A near the foundation. Due to the weak intrusion

of altered grit under the buttress downstream toe the rock deformability is much larger

and the concrete stresses are reduced. The linear distribution of the vertical stresses is

no longer valid.

Page 73: Dam Engineering GRV BUTT

81

Fig. 3.17. Distribution of vertical stresses due to gravity and water pressure

3.7. MASSIVE HEAD BUTTRESS DAM– ECONOMY

The economic merits of the buttress dams have been systematically analysed by C.

Marcello. The main factor rendered evident was the concrete saving that corresponds

to a massive head buttress dam and to a hollow gravity dam respectively, as compared

to a gravity dam. The percentage saving is shown in Figure 3.18.

Fig. 3.18. Concrete savings in terms of the valley shape and the dam height

Savings

Page 74: Dam Engineering GRV BUTT

82

Three shapes were considered for the dammed valley: triangular, trapezoidal and

rectangular. Based on the diagrams in Figure 3.18 one can notice that:

- regardless the buttress dam type the concrete saving increases as the valley

opens;

- hollow gravity dams are more economic than the massive head buttress

dams;

- the concrete saving becomes significant when the dam height is higher

than 35 m;

- for dams higher than 70 - 80 m the concrete saving is constant 30 % on the

average.

However, the saving in costs is more modest due to the fact that a buttress dam

requires more formwork for the more complicated shape and higher cement content

for a slimmer structure. Further reduction of the cost saving is due to the difficulties in

dam concreting induced by the reduced concreting area in the slimmer zones of the

buttresses. According to Italian practice the actual cost reduction per cubic meter of

concrete in the replaced gravity dam was several percentages.

BIBLIOGRAPHY

Constantinescu, Al., Stematiu, D., Hapau-Petcu, S. Staged implementation of the

remedial works program for Poiana Uzului buttress dam. Transactions of 21

Congress on Large Dams, Q82, R20, June, 2003.

Davis, C.,V., Sorensen, E.,K. Handbook of applied hydraulics. McGraw-Hill, 1969.

ICOLD Bulletin 88. Rock foundation for dams. 1993.

Legas, J. Concrete buttress dams. Development of dam engineering in the United

States. Kolgaard, E.,B., Chadwick, W.,L. editors, Pergamon Press, 1988.

Popovici,A. Water storage dams. (in Romanian), Bucharest, 1992.

Priscu, R. Hydraulic structures. (in Romanian) , Bucharest, 1974.

Romanian National Committee on Large Dams. Dams in Romania, Bucharest 2000.

Schnitter, N. A history of dams – the useful pyramids. A.A.Balkema, 1994.