You are on page 1of 124

DEGREE PROJECT, IN CONCRETE STRUCTURES , SECOND LEVEL

STOCKHOLM, SWEDEN 2015

Evaluation of Failure Modes for


Concrete Dams

LISA BROBERG & MALIN THORWID

KTH ROYAL INSTITUTE OF TECHNOLOGY

SCHOOL OF ARCHITECTURE AND THE BUILT ENVIRONMENT


Evaluation of Failure Modes for Concrete
Dams

Lisa Broberg & Malin Thorwid

June 2015
TRITA-BKN. Master Thesis 455, 2015
ISSN 1103-4297,
ISRN KTH/BKN/EX–455–SE
Lisa
c Broberg & Malin Thorwid 2015
Royal Institute of Technology (KTH)
Department of Civil and Architectural Engineering
Division of Concrete Structures
Stockholm, Sweden, 2015
Abstract

The safety of a concrete dam is ensured by designing according to failure criteria, for
all combinations of loads using safety factors. Today in Sweden, RIDAS, the Swedish
power companies’ guidelines for dam safety, is used for the design of dams and is
based on BKR, the National Board of Housing, Building and Planning. Swedish
dams are designed to resist two global failure modes; sliding and overturning. Com-
bination of failure modes, that should be considered in the design of concrete dams,
is however fairly unknown. Since 2009 the Eurocodes was adopted and came into
force 2011. The Eurocodes have replaced BKR in the design of most structures in
Sweden where the partial factor method is used to ensure safety in the design.
The objective of this report was to examine if the design criteria for concrete dams
in today’s condition are enough to describe real failure modes. The other objective
was to analyse if Eurocode is comparable to RIDAS in dam design. The stated
questions were answered by performing a literature study of known dam failures
and analytical calculations for different types of concrete gravity dams, with varying
geometry and loading conditions. The programs CADAM and BRIGADE were also
used as calculation tools to further analyse if failure occurred as expected.
The results from the analytical calculations together with the performed FE anal-
ysis indicate that limit turning does occur and often generate lower safety factors
compared to overturning. Limit turning is similar to overturning failure although it
accounts for material failure in the rock. This design criterion is therefore, highly
dependent on the quality of the rock and requires investigations of the foundation
to be a good estimation of the real behaviour of the dam body.
From the compilation of reported failures the conclusion was that the current de-
sign criteria are adequate. However, the real challenge lies in ensuring that the
construction of dams is correctly performed to fulfil the stated criteria. A transition
to Eurocode appears to be reasonable for the stability criterion. A modification of
the partial factors is nevertheless necessary to obtain result corresponding to
RIDAS, especially for the overturning criteria.
Keywords: Gravity dams, concrete, design criteria, RIDAS, Eurocode, limit turn-
ing

iii
Sammanfattning

För att uppnå säkra dammkonstruktioner, för alla lastkombinationer, dimensioneras


dammar enligt bestämda brottvillkor som ska uppfylla en viss säkerhetsfaktor. Idag
används RIDAS, för dimensionering av dammar i Sverige. RIDAS Kraftföretagens
riktlinjer för dammsäkerhet, är baserat på BKR, Boverkets konstruktionsregler. I
Sverige dimensioneras dammar för att motstå de två globala brottmoderna glidning
och stjälpning. Frågan som behöver besvaras är om det finns eller kan finnas några
kombinationer av brottmoder som borde beaktas vid dimensionering av dammar.
2009 antogs Eurokoderna och trädde i kraft 2011. Eurokoderna har ersatt BKR vid
dimensionering av de flesta konstruktioner i Sverige. I Eurokod används partialko-
efficienter för att garantera säkra konstruktioner.
Syftet med denna rapport var att undersöka om dagens brottkriterium är tillräck-
liga för att beskriva hur dammar går till brott. Rapporten behandlar även möj-
ligheten att övergå från att dimensionera dammar enligt RIDAS till att dimen-
sionera enligt Eurokod. För att besvara detta genomfördes en litteraturstudie av
rapporterade dammbrott. Dessutom genomfördes analytiska beräkningar för flera
olika typer av dammar med varierande geometri och lastfall. Programmen CADAM
och BRIGADE användes som ytterligare verktyg för att analysera brottmoderna.
Enligt resultat från de analytiska beräkningarna tillsammans med FE-beräkningarna
ansågs limit turning inträffa och genererade i högre grad en lägre säkerhetsfaktorer
i jämförelse med stjälpning. Limit turning kan förklars som delvis stjälpande och
inkluderar brott av bergmassan. Brottmodet är dock beroende av kvalitéten hos
berget och det krävs undersökningar av grunden för att kunna göra en god uppskat-
tning av dammens verkliga beteende.
Sammanställningen av tidigare brott visade att nu gällande brottkriterier är lämpliga
och troligtvis tillräckliga. Utmaningen är istället att säkerställa att konstruktion-
erna är korrekt utförda och därmed uppfyller dessa brottkriterier. En övergång
till Eurokod tycks vara möjlig för de globala brottmoderna, dock är det väsentligt
att partialkoefficienterna justeras för att uppnå resultat som överensstämmer med
RIDAS, särskilt för stjälpning.
Keywords: Gravitationsdammar, betong, brottkriterium, RIDAS, Eurokod, limit
turning

v
Preface

This thesis was carried out from January to June 2015 at SWECO Energuide AB
in collaboration with the Division of Concrete Structures, Department of civil and
Architectural Engineering at the Royal Institute of Technology (KTH). The project
was initiated by Dr. Richard Malm, who also supervised the project, together with
Ph.D. candidate Daniel Eriksson and Adjunct Prof. Erik Nordström.
We would especially like to thank Richard Malm for the continuous support which
has been a great encouragement. We would also like to thank Daniel Eriksson for
always finding time to help and guide us throughout this project. We also wish to
thank Erik Nordstöm for his guidance and advice.
We would like to thank the division at SWECO Eneriguide AB for their warm
welcome and for an inspiring job environment. We would like to give special thanks
to Johan Nilsson for the support and help during our work at SWECO.
Stockholm, June 2015
Lisa Broberg & Malin Thorwid

vii
Contents

Abstract iii

Sammanfattning v

Preface vii

1 Introduction 1
1.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Aim of report . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3 Limitations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.4 Structure of report . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

2 Concrete gravity dams 5


2.1 Gravity dams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.1.1 Massive dams . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.1.2 Buttress dams . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.1.3 Gate section . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.1.4 Support methods . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.2 Stability analyses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.2.1 Design loads . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.2.2 Failure modes . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

3 Methods for stability analyses 15


3.1 RIDAS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.1.1 Design loads . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.1.2 Failure modes . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.2 Eurocode . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.2.1 Design loads . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.2.2 Failure modes . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.3 Limit turning . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.3.1 Crushing resistance . . . . . . . . . . . . . . . . . . . . . . . . 28
3.3.2 Failure criteria . . . . . . . . . . . . . . . . . . . . . . . . . . 30

4 Failure modes of concrete dams 33


4.1 Documentation of failures . . . . . . . . . . . . . . . . . . . . . . . . 33
4.2 Compiled failures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
4.2.1 Comparison of properties . . . . . . . . . . . . . . . . . . . . . 35
4.2.2 Failure type . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

ix
4.2.3 Failure mode . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
4.3 Description of failures . . . . . . . . . . . . . . . . . . . . . . . . . . 38
4.3.1 Documentation regarding failures . . . . . . . . . . . . . . . . 39
4.4 Results of the compiled failures . . . . . . . . . . . . . . . . . . . . . 49

5 Stability analyses 53
5.1 Studied dams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
5.1.1 Input data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
5.1.2 Previously studied dams . . . . . . . . . . . . . . . . . . . . . 58
5.2 Stability Calculations . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
5.2.1 Design approaches . . . . . . . . . . . . . . . . . . . . . . . . 59
5.2.2 Parametric study . . . . . . . . . . . . . . . . . . . . . . . . . 62
5.3 CADAM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
5.3.1 Stability calculations . . . . . . . . . . . . . . . . . . . . . . . 63
5.3.2 Modelling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
5.4 FE-analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
5.4.1 Studied dams . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
5.4.2 Model definition . . . . . . . . . . . . . . . . . . . . . . . . . . 67

6 Results and discussion 71


6.1 Analytical analyses . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
6.1.1 Design approaches . . . . . . . . . . . . . . . . . . . . . . . . 71
6.1.2 Parametric study . . . . . . . . . . . . . . . . . . . . . . . . . 79
6.1.3 Previously studied monoliths . . . . . . . . . . . . . . . . . . . 84
6.2 Analyses of limit turning . . . . . . . . . . . . . . . . . . . . . . . . . 84
6.2.1 Analytical analysis . . . . . . . . . . . . . . . . . . . . . . . . 84
6.2.2 FE-analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87

7 Conclusions 91
7.1 Failure modes of concrete dams . . . . . . . . . . . . . . . . . . . . . 91
7.2 Analytical calculations . . . . . . . . . . . . . . . . . . . . . . . . . . 92
7.3 Design guidelines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
7.4 Future studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94

Bibliography 95

A Compiled failures 101

B Results analytical analyses 103

C Output values for dams 105


Chapter 1

Introduction

1.1 Background
Most dams in Sweden were built during 1950 to 1960 on solid good quality rock.
The dams were built under different conditions and safety regulations compared to
the demands stated today (Andersson, 2012). The knowledge of rock mechanics and
material properties of concrete along with the building techniques have improved.
Today the construction of new dams in Sweden is limited by regulations concerning
the preservation of the environment. Therefore, the design of dams mostly involves
maintenance and reparation of existing dams.
Knowledge of why and how dam failure occurs, may help prevent or minimise the
damage. The indication of how a dam behaves prior to failure is therefore of great
importance in order to prevent failure. It is also important since it provides guidance
on how to monitor and measure dams, what types of sensors and where these sensors
should be placed to get early indications of possible failures. Risk and safety are
essential in dam design due to the radical consequences a failure would cause to the
surroundings, as seen in Figure 1.1.

Figure 1.1: Baldwin Hills Reservoir after the disaster 1963 (Wilson, 1963).

1
CHAPTER 1. INTRODUCTION

The consequences of failure could in the worst case scenario lead to lives lost and
economic damage. Therefore, the government through the public utility Svenska
Kraftnät, stated new requirements concerning higher safety demands for the existing
dams (SFS 2014:114). The new requirements also concern the classification of the
dams in Sweden; all dams must be classified, if a failure could result in severe
consequences.
In Sweden, the dam owner is responsible in the event of a failure or an accident.
The Swedish power companies have issued the Swedish power companies’ guidelines
for dam safety, RIDAS, based on the construction rules BKR (2010), the National
Board of Housing, Building and Planning. Since 2011, the Eurocodes (the European
construction standards) together with EKS 9 (2013), have replaced BKR in Sweden.
However, Eurocode does not account for the design of dams (Andersson, 2014).
Today Swedish dams are designed to withstand two global failure modes; sliding
and overturning of the entire monolith. There are questions regarding if there are
or can be any combinations of the failure modes, that should be considered in the
design of concrete dams.

1.2 Aim of report


The main focus of this report is to analyse the different types of failures that have
occurred and can occur in concrete gravity dams, by examining the influence of
different factors. The aim of this study is to answer the following questions:
• Is analytical calculations based on the global failure modes: sliding and over-
turning enough to describe the failure of the dam? Are there other potential
failure modes not covered by these analytical calculations?
• Is a transition to Eurocode possible for dam design? Are the design guidelines
according to Eurocode comparable to RIDAS?
The stated questions will be answered by performing a literature study of reported
dam failures and analytical calculations for several concrete gravity dams. In addi-
tion a FE-analysis will be performed for comparison.

1.3 Limitations
This report only include concrete gravity dams and further limitations for the litera-
ture study are stated in Chapter 4. The analytical calculations only concern dams on
rock foundation. The analyses are limited to Swedish conditions regarding material
properties and design parameters. The influence from seismic loads is not included
due to that it is not considered for design in Sweden, while it internationally may
be of great importance for the design.

2
1.4. STRUCTURE OF REPORT

1.4 Structure of report


Chapter 2 includes a presentation of theory behind the key concepts of concrete
gravity dams. The main features for stability analyses of concrete gravity dams
are presented. The different design loads for the stability analysis are stated and
illustrated. A brief presentation of the causes of concrete gravity dam failure by
explanation of the failure modes is presented.
Stability analyses according to the guidelines RIDAS and Eurocode are presented
in Chapter 3. How the guidelines account for the design loads and explanations of
the analytical calculation methods for the failure modes described in Chapter 2 are
given.
A compilation of reported failures including causes and failure modes is presented
in Chapter 4. Already known facts about the presented failures of concrete dams
are compiled and the sources of failures that has occurred are detected.
The analytical analyses in Chapter 5 describe the stability calculations for several
different dams with varying geometry and loading conditions. A parameter study is
used to determine the most influential parameter and to adapt the design of dams
according to Eurocode with the stability calculations in agreement with RIDAS.
The program CADAM and the software BRIGADE, are used as tools to enable
a comparison to the analytical calculations. The key concepts on how to perform
stability analyses with these tools are presented in this chapter. The results from
the analyses described in Chapter 5, are presented in Chapter 6.
In Chapter 7, the conclusions of the different concrete dam failure analyses are
presented.
Appendix A includes the full compilation about the failures of concrete dams pre-
sented in Chapter 4. Appendix B includes the results from the analytical calculations
described in Chapter 5. The loads and level arms from the analytical calculations
are presented in Appendix C.

3
Chapter 2

Concrete gravity dams

There are different types of concrete dams, which are distinguished by how the
water pressure of the reservoir is transferred down to the ground. Descriptions of
the different types of concrete gravity dams are included in this chapter. The two
most common types are massive and buttress dams, presented in Section 2.1.1 and
Section 2.1.2 respectively. Gate sections mainly consist of pillars and spillways, both
are of gravity dam type and are described in Section 2.1.3. Figure 2.1, shows the
different dam types included in this report.

Figure 2.1: The different types of dams included in this report, a) spillway and pillar,
b) massive and c) buttress.

For other types of dams, not included in this report, concrete can also be used in em-
bankment dams, as a central or upstream membrane, as retaining walls for spillways
or used for many secondary functions. Embankment dams are usually associated
with at least one concrete dam part, either intake and/or discharge facilities. Con-
crete arch dams were introduced relatively late and therefore have a uniform, and
somewhat higher quality. Arch dams are founded on rock and are of slender type
with concaved arches (Kleivan et al., 1994).

5
CHAPTER 2. CONCRETE GRAVITY DAMS

2.1 Gravity dams

2.1.1 Massive dams

Massive or gravity dams are solid structures, designed to resist the external forces
by its dead weight. Today, gravity dams are mainly constructed with concrete,
compared to the previously used method of stone masonry. The development of new
concrete gravity dams is ongoing and the Roller-Compacted Concrete dam (RCC)
is an example of that. The RCC dam has a limited use of formwork, consists of a
drier mix and is placed in a manner similar to paving, i.e. compacted with vibrating
rollers. The benefits are cost beneficial with simple faster construction. The dams
are built with no joints or reinforcement, with low cement content and the use of fly
ash that enable less heat generation while curing (Kleivan et al., 1994).
Solid concrete structures maintain stability against loads due to the geometric shape,
mass and strength of the concrete (Ali, 2012). A gravity dam consists of either a
continuous or a series of concrete monoliths separated by expansion joints (RIDAS,
2011). The monolith cross section is, in principle, triangular with a dam head, an
inclination of the downstream face and a vertical upstream face, which also can have
a small inclination, see Figure 2.2. The benefits of concrete gravity dams are that
they are easily constructed on sites with a foundation of sufficient strength to carry
the weight of the dam (Ali, 2012).

Figure 2.2: Typical cross-section of a massive dam, reproduced from Bergh (2014).

The monoliths are mainly placed in a straight line or sometimes slightly curved, see
Figure 2.3 and are usually of a width between 5 m to 10 m (Bergh, 2014).

6
2.1. GRAVITY DAMS

Figure 2.3: Stadsforsen, massive dam in Sweden (Malm, 2015).

2.1.2 Buttress dams

Over time, there has been a strong effort towards improving concrete quality. There
has been a shift from the previously dominant gravity type dam such as massive
dams, towards a more slender type of dam with reinforcement, known as a buttress
dam (Kleivan et al., 1994).
Buttress dams consist of two rigidly connected elements, the upstream water barrier
(frontplate) and the supporting buttress on the downstream side, which together
form a monolith, see Figure 2.4. The upstream water barrier transfers the hydro-
static pressure over to the buttress, which in turn transfers it down to the foundation.
The water barrier is inclined so the vertical water loads, together with the weight of
the concrete, act in favour for the stability of the monolith (DOI, 2009).

Figure 2.4: Cross-section of a buttress dam (Left), Section of a buttress dam (Right),
reproduced from RIDAS (2011).

A buttress dam consists of a series of monoliths, connected by horizontal struts


acting as contraction joints, connecting the adjacent monoliths, see Figure 2.5. The
casting arrangements and the construction are somewhat more demanding com-
pared to most other types of dams. However, buttress dams are more suitable on

7
CHAPTER 2. CONCRETE GRAVITY DAMS

weak foundations compared to gravity dams, due to the reduced volume of concrete
(Bergh, 2014), while the contact pressure between the buttress and the foundation
is considerably higher.

Figure 2.5: Rätan, buttress dam in Sweden (Vattenkraft.info, 2009).

2.1.3 Gate section

Concrete gravity dams also consist of gate sections to transport water in specific
directions and release water pressure on the dam structure. Concrete functions as
a fastener for many different types of gate installations, with variable functions.
The gate type could be sliding, roller or radial gates, with the function of either an
outlet gate, intake or daft tube (Kleivan et al., 1994). An example of a gate section
in Sweden is shown in Figure 2.6

Figure 2.6: Lima, Spillway dam with two river type gates (Norconsult, 2012).

8
2.1. GRAVITY DAMS

For a spillway section see Figure 2.7, where the overflow is designed with a vertical
upstream face. The water is able to flow over the crest along the inclined down-
stream side with training walls, keeping the water in place and finally the water
reaches the energy dissipating structure, forming a hydraulic jump to avoid erosion
of the riverbed. Pillars, non-overflowing blocks function as enclosures of a number of
overflow sections, see Figure 2.7. Usually spillway sections have gates and typically,
radial gates see Figure 2.7.

Figure 2.7: Spillway section (Left), plane view of spillway (Right), reproduced from
RIDAS (2011).

For hydropower structures, the intake is the connection between the reservoir and
the waterway connected to the turbine of the hydroelectric power plant. Intake
gates are normally designed with a trash rack preventing debris, ice, fish, etc. from
entering the intake. In addition, it also consists of a gate of river or tunnel type, to
close of the conduit, see Figure 2.8.

Figure 2.8: Section of a typical inlet and power station, reproduced from RIDAS
(2011).

The behaviour of a spillway, discharge part, and intake part of the dam is similar
to concrete gravity dams, thus the geometry and loading conditions may be more
complicated (Westberg and Hassanzadeh, 2007). The surface of the concrete is
subjected to very high water velocities as well as abrasion, and therefore must be
steel plated.

9
CHAPTER 2. CONCRETE GRAVITY DAMS

2.1.4 Support methods

There are different support methods for concrete gravity dams. Common methods
are to secure the dam body to the foundation by rock bolts or tendons. Rock
bolts are non-pre-stressed reinforcement bars installed in the interface between the
foundation and the dam body. There are different types of fastners for the rock
bolts and they can be secured to the foundation through anchors, cables, dowles or
by grouting. The bolts are anchored in the dam body by adhesion. Rock tendons
consist of pre-stressed cables or rods, anchorage and corrosion-inhibiting coating.
The tendons can be unbonded or bonded to the surrounding concrete. The tendon
is inserted into a casing and grouted after the tendon is stressed (PTI, 2000).
Another common method used to support concrete dams is earth support fill on the
downstream side of the dam, giving rise to stabilising forces. A grout curtain helps
to stabilise the dam by decreasing the uplift pressure. The grout curtain is achieved
by inserting cement into pores and cavities in the ground. The grout curtain might
deteriorate over time and therefore the decreased uplift is usually not accounted for
in the design of dams (Ferc Engineering Guidelines, 2002). The uplift pressure could
also be decreased by inserting drainage pipes.

2.2 Stability analyses


Concrete dams are massive large structures since they are designed to fulfil the
requirements for stability. The design should also fulfil requirements for long life
spans and water tightness to withstand the permanent water pressure (Bond, 2014).
The safety of the concrete dam is assured by designing according to failure criteria,
for all combinations of loads using safety factors. The safety criterion is the definition
of the stress level when failure occurs. The safety factors are chosen to provide for
all underlying uncertainties. Their magnitude should reflect the probability of the
occurrence for the particular load, the accuracy of conditions and the method of
analysis. The factor of safety is thereby higher for foundation studies, because of
the greater amount of uncertainty in assessing the load-resistance capacity of the
foundation.

2.2.1 Design loads

The loads included in the stability analysis should represent the actual loads acting
on the concrete dam during operation. Many of the loads are unable to be exactly
determined, the engineer is then responsible for estimating these loads based on
available measurements, judgement and experience (Ferc Engineering Guidelines,
2002).
The required loads acting on a dam for a stability analysis are shown in Figure 2.9.

10
2.2. STABILITY ANALYSES

Figure 2.9: Loads acting on a dam, reproduced from (Bergh, 2014).

1. Hydrostatic pressure (P1-P2) – depends on the water level in the dam


2. Tailwater pressure (P3-P4) – depends on the tailwater level
3. Uplift pressure (P5) – hydrostatic pressure acting vertically, assumed to vary
linearly from hydrostatic pressure at the heel to the tailwater pressure at the
toe
4. Dead weight (P6) – the weight of the concrete
5. Ice pressure (P7) – load acting on the face of the dam due to an ice cover
6. Silt pressure (P8) – settled sediments exerting active pressure towards the dam
7. Seismic loads (P9-P11) – horizontal and vertical accelerations caused by earth
quakes

2.2.2 Failure modes

The definition of dam failure can differ between individuals, the general definition
could be expressed as: "Collapse or movement of part of a dam or its foundation,
so that the dam cannot retain water” (ICOLD, 1995).
Concrete dams have various failure behaviour. Sliding or shear failure is the most
common failure for dams constructed on rock. The dam may fail due to crushing,
i.e. the failure of its materials when the compressive stresses exceed the acceptable
stresses. Concrete cannot withstand sustained tensile stress and if the tension that
develops in the concrete exceed its tensile strength, it could lead to ultimate failure.
The dam may also fail due to overturning where it rotates about the toe (Ali, 2012).

11
CHAPTER 2. CONCRETE GRAVITY DAMS

Overturning

Overturning occurs when the forces acting on the dam causes rotation of the dam,
see Figure 2.10. Overturning is analysed by calculating a factor of safety, which
is defined as the ratio of stabilising and overturning moment. These moments are
calculated around its toe or another weak point in the structure or foundation. For
the overturning failure, it is also important that the resultant is located in the mid
third of the base area since this will assure that the whole base of the dam is under
compression. If tensile stress can be avoided it will reduce crack propagation in
the concrete. The criterion is verified by application of the Navier equation for a
cantilever action under combined axial and bending load (Bergh, 2014).

Figure 2.10: Overturning failure around the dam toe.

Sliding

Sliding occurs when the horizontal forces exceed the frictional resistance. Sliding
can be divided in to three different kinds of failures (Gustafsson et al., 2008):
1. Failure in the interface between the concrete and the foundation
(Figure 2.11 a).
2. Failure in weak planes of the foundation, such as cracks (Figure 2.11 b).
3. Failure in the solid foundation (Figure 2.11 c).

Figure 2.11: Different types of sliding failures, reproduced from Gustafsson et al.
(2008).

12
2.2. STABILITY ANALYSES

There are different views on whether cohesion between the concrete and the foun-
dation can be accounted for in the sliding stability. The reason is that it is difficult
to quantify through borings’ and testing. Higher allowable safety factors may be
applied, if cohesion is included in the calculations for sliding stability (Ferc Engineer-
ing Guidelines, 2002). If cohesion is not accounted for, the concrete-rock interface
is treated as unbonded giving a conservative method that may result in expensive
and over-strong structures (Krounis, 2013).
Failure in the interface between the foundation and the dam body is normally ac-
counted for in the design of dams. Though failure more often occurs in weak planes
of the foundation, see Figure 2.11 (DOI, 2012). Sliding can also occur in the dam
body at weak planes such as lift joints or along cracks, this failure is seldom analysed
except for high dams (Ali, 2012).

Limit turning

According to Fishman (2007), the classical failure modes sliding and overturning,
do not account for material failure. Classical overturning failure is unrealistic as
it requires infinitely strong rock and concrete. Fishman infers that the one failure
mode, either limit turning or sliding, giving the lowest stability factor should be
used for the design of the structure and decisions regarding interface preparation
(Fishman, 2009).
Fishman states that the stresses developing below the upstream side will result in
a tensile crack along the rock, see Figure 2.12. A compressive zone will be formed
in the rock, underneath the toe, due to the applied forces on the dam. When the
stresses exceed the crushing resistance of the rock, a crushing zone is formed. The
size of the crushing zone depends on the strength of the rock, for a weaker rock the
crushing zone will extend further to the upstream side. The turn axis appears where
the tensile crack and the compressive crack meet. The concrete and rock will act as
a single body and failure will occur when they rotate about the new rotation point.
This failure mechanism is called limit turning, which is similar to the overturning
method although it also accounts for the strength of the rock. The result gives a
more reliable and conservative safety factor (Fishman, 2009).

Figure 2.12: Limit turning failure.

13
Chapter 3

Methods for stability analyses

In this chapter, the design methods for stability analyses are presented. There
are different methods for analysing failure of dams, varying between countries or
methods. In Sweden, RIDAS (2011), is used for the design of dams and is based
on BKR. 2011 BKR was replaced with EKS 9 (2013). Today, Eurocode is used and
function as guidelines providing a common structural design in European countries.
A revision of RIDAS, to incorporate Eurocode instead of BKR has begun. The
transition work is led by Svensk Energi, responsible for RIDAS. The research com-
pany, Energiforsk, previously known as Elforsk until January 2015, has also started
to work with the transition from BKR to Eurocode. The investigations has so far
mostly been focused on cross-section design (Andersson, 2014).
For the stability analyses, there is an ongoing project financed by Energiforsk, where
a structural reliability based method is under development. In this report, the
focus will be on RIDAS, using the deterministic method, compared to Eurocodes
semi-probabilistic method, resulting in the use of partial factors, for calculations of
stability (Westberg, 2014).
The denotations from RIDAS and Eurocode were used in Section 3.1 and Section
3.2, respectively.

3.1 RIDAS
RIDAS (2011), is based on BKR, with adjustments for specific requirements for con-
crete dams. BKR, as mentioned before, is not valid today however it may still be
used if the contents do not conflict with the Eurocodes (Andersson, 2014). Guide-
lines are given for the design of dams together with control and reconstruction of
existing dams. RIDAS states requirements for stability, strength and durability of
the dam and what criteria to fulfil. The requirements and criteria concern gravity
dams, where RIDAS have included the most common types; massive and buttress
dams, including spillway, inlet dams and pillars.

15
CHAPTER 3. METHODS FOR STABILITY ANALYSES

3.1.1 Design loads

RIDAS (2011) includes guidelines to determine design loads acting on concrete dams.
It states how to account for the loads presented below. RIDAS also gives guidance
how to account for rock anchors (described below), temperature effects, creep and
shrinkage, and traffic loads (included if unfavourable), which are not described in
this report.

Dead weight

For design of new concrete dams, the dead weight for reinforced concrete is assumed
to be 23 kN/m3 if no material tests are available. For existing dams, the dead weight
should be determined from material tests or from information about the design.

Hydrostatic pressure

Both the water pressure on the up- and downstream side should be accounted for.
The most unfavourable combinations of up- and downstream water levels applied to
the dam determine the water pressure to be used in the calculations.

Uplift pressure

The uplift pressure distribution varies for different dam types and designs with or
without drain pipes. For massive dam structures without drainage where the whole
foundation area is under compression, the uplift pressure distribution varies linearly
from the upstream to the downstream side.
For massive dams the uplift pressure can be reduced by the use of drain pipes as
shown in Figure 3.1 and Figure 3.2.

16
3.1. RIDAS

Figure 3.1: Uplift distribution for a dam with a drainage pipe in the rock and a
drainage tunnel by the rock surface, reproduced from RIDAS (2011).

The uplift distribution in Figure 3.1 will vary from H to 0.3 · (H − h) + h closest
to the drainage tunnel and varies linearly to h at the toe of the monolith, with no
uplift beneath the drainage tunnel. H is the headwater level and h is the tailwater
level.

Figure 3.2: Uplift distribution for a dam with a drainage pipe in the rock and
drainage tunnel in the concrete, reproduced from RIDAS (2011).

The uplift distribution in Figure 3.2 will vary from H to 0.5 · (H − h) + h closest
to the drainage tunnel and varies linearly to h at the toe of the monolith. H is the
headwater level and h is the tailwater level.

17
CHAPTER 3. METHODS FOR STABILITY ANALYSES

For buttress dams, the uplift distribution is assumed to vary linearly over the
thickness of the frontplate. If the buttress is thicker than 2 m, the uplift pressure
underneath the buttress should be included as shown in Figure 3.3.

Figure 3.3: Distribution of uplift pressure for a buttress dam with a buttress thicker
than 2 m, reproduced from RIDAS (2011).

For spillways, the uplift distribution is assumed similar to massive dams. The uplift
distribution in Figure 3.4 is assumed for pillars, where w is the width of the pillar.
The uplift varies from full uplift pressure to zero at the distance w from the spillway.

Figure 3.4: Uplift distribution for pillars, reproduced from RIDAS (2011).

The effect from cement grouting is not considered in the uplift pressure distribution,
due to the strength of the cement decrease with time due to deterioration. The grout
curtain is only considered as extra safety and should not be accounted for unless
re-grouting is possible. This is seldom the case due to difficulties incorporating
re-grouting tunnels in the dam body design.

18
3.1. RIDAS

Ice load

The intensity of the horizontal ice pressure depends on the geographic location,
altitude and local conditions for the dam. RIDAS suggest the horizontal ice pressure
50 kN/m with an ice thickness of 0.6 m, for dams located in the southern part of
Sweden. Dams located in the middle part, should be designed for an ice load of
100 kN/m with an ice thickness of 0.6 m. An ice load of 200 kN/m with an ice
thickness of 1 m, is suggested for the rest of Sweden. The resultant of the ice
pressure is assumed to be located at one third of the ice thickness, calculated from
the top surface of the ice.

Rock anchors

For lower dams, it can be hard to achieve stability, and according to RIDAS in these
cases it is allowed to assign a load capacity of 140 MPa to the rock bolts. This is
applied for dams that have a headwater level less than 5 m and do not belong to
any of the two highest safety classes.
For all other dams, rock anchors should not be considered in the stability calcu-
lations, due to the complications of verifying their strength. However, it is stated
that the installation of rock anchors of the dimension φ25 − 32 is a good preventive
measure.

Earth pressure

Soil may be added as downstream support fill to increase stability. The earth pres-
sure should be determined as a at-rest pressure and it should be calculated as the
lowest theoretical pressure that may occur. The soil density and earth pressure co-
efficient should be obtained from in-situ tests. If testing is not possible, the values
in Table 3.1 may be used.

Table 3.1: Example values for unit weight and coefficients for earth pressure RIDAS
(2011).

Unit weight density [ kN/m3 ] Friction angle [ ◦ ] Coefficient for earth pressure [-]
Material Un-saturated Saturated φ At-rest K0 Active Ka
Rockfill 17.5 11 42 0.33 0.20
Gravel 18 11 35 0.43 0.27
Sand 18 11 32 0.47 0.31
Moraine 21 13 34 0.45 0.29

19
CHAPTER 3. METHODS FOR STABILITY ANALYSES

3.1.2 Failure modes

According to RIDAS (2011) there are three failure modes that need to be analysed
for stability; sliding, overturning and the bearing capacity of the concrete and the
foundation.
Dam stability analyses are performed using safety factors for overturning and allow-
able friction coefficients for sliding, to achieve a safe design.
RIDAS has listed different load combinations to be analysed, these are divided in
to; normal load combinations, exceptional load combinations and accidental load
combinations. The loads are calculated without partial factors and are analysed for
individual monoliths.

Overturning

For overturning the requirement of the safety factor s is defined according to


Table 3.2.

Table 3.2: Saftey factor for overturning.

Load case Safety factor (s)


Normal 1.50
Exceptional 1.35
Accidental 1.10

The safety factor defines the relation between stabilising and overturning moment,
see Equation (3.1), and should not be lower than the safety factor in Table 3.2.

Mstab
s= (3.1)
Mover

Sliding

RIDAS states that sliding should be analysed between the interface of the concrete
and in the foundation, along potential weak planes and in weak points in the dam
body. Stability against sliding is achieved if the sum of the horizontal forces divided
by the vertical forces, see Equation (3.2), does not exceed the maximum allowed
friction coefficient, see Table 3.3.

RH tan δg
µ= ≤ µmax = (3.2)
RV sg

20
3.2. EUROCODE

where
RH is the resultant of the horizontal forces.
Rv is the resultant of the vertical forces.
tan δg is the friction angle.
sg is the safety factor.

Table 3.3: Maximum friction coefficient, µmax

Normal load Exceptional Accidental


Foundation
case load case load case
Rock 0.75 0.90 0.95
Gravel, sand and moraine 0.5 0.55 0.60
Silt 0.40 0.45 0.50

The maximum friction coefficient can be calculated according to Equation (3.2)


where the values for the safety factor sg are presented in Table 3.4.

Table 3.4: Saftey factors sg for calculations of µmax .

Normal load Exceptional Accidental


Foundation
case load case load case
Rock 1.35 1.10 1.05
Gravel, sand and moraine 1.50 1.35 1.25
Silt 1.50 1.35 1.25

Cohesion between the dam and the foundation should not be considered in the
calculations for the resistance against sliding according to RIDAS.
When calculating stability against sliding, the value for the maximum allowed fric-
tion coefficient (µmax ) is calculated according to Equation (3.2), with values of the
safety factor from Table 3.4 and the friction angle from geotechnical investigations.
The values in Table 3.3 can be used for dams constructed on a foundation of good
quality, when calculating stability against sliding.

3.2 Eurocode
Eurocode is the European standard for technical rules in construction work, pro-
viding a common structural design tool in European countries. Eurocode clearly
states that their guidelines do not cover the design of dams. This is due to the high
safety required for dams and that other aspects than for usual design need to be
considered. This section is therefore solely based on the authors assumptions on
how to apply Eurocode to dam design. Therefore, in this report a compilation of
information from the listed Eurocodes below was performed in order to obtain a
method applicable for stability analyses of dams.

21
CHAPTER 3. METHODS FOR STABILITY ANALYSES

• EN 1990 “Basis for structural design”.


• EN 1991 “Actions on structures”.
• EN 1992 “Design of concrete structures”.
• EN 1997 “Geotechnical design”.
The same failure modes as described in Section 2.2.2 were analysed. The values and
equations in the following sections were based on the assumption that dams can be
considered as comparable with retaining wall structures.

3.2.1 Design loads

The safety is applied on the loads by partial factors. The load acting on the structure
is multiplied with the partial factor γ to define the design load. The loads are
classified according to variation in time: permanent or variable loads and whether
the load is favourable or unfavourable, which would result in different partial factors.
When designing geotechnical structures, different approaches are used. In accor-
dance with retaining wall structures, the concrete dams were assigned the design
approach 3 (DA 3). The different approaches give different values for the partial
factors; for load and load effects, soil parameters and the strength (EC 7, 2011).
Design approach 3 states that different partial factors should be used for geotech-
nical actions and structural actions. Geotechnical actions are defined as actions
transmitted to the structure by the ground, fill, standing water or groundwater.
For structural actions the strength of the material is significant. For the structural
actions Equation (3.3) and (3.4) are used. For geotechnical actions, Equation (3.5)
is used (EC 0, 2002). The partial factors for the equations are listed below in Table
3.5.
Ed = Σγd · γG · Gk + Σγd · γQ · ψ0,i · Qk,i (3.3)
Ed = Σγd · ξ · γG · Gk + γd · γQ,1 · Qk,1 + Σγd · γQ,i · ψ0,i · Qk,i (3.4)
Ed = Σγd · γG · Gk + γd · γQ,1 · Qk,1 + Σγd · γQ,i · ψ0,i · Qk,i (3.5)

where
γG is the partial factor for permanent actions.
γd is the partial factor depending on safety class.
Gkg is the characteristic value of a permanent action.
γQ,i is the partial factor for variable action.
ψ0,i is the factor for combination value of a variable action.
Qk,i is the characteristic value of a single variable action.
ξ is the reduction factor.

22
3.2. EUROCODE

Table 3.5: Partial factors according to Eurocode (EC 7, 2011).

Load combination equation 3.3/3.4 3.5


For an unfavourable permanent load γG = 1.35 γG = 1.1
For a favourable permanent load γG = 1.0 γG = 1.0
For an unfavourable variable load γQ = 1.5 γQ = 1.4
For a favourable variable load γQ = 0 γQ = 0
Reduction factor ξ = 0.89/- -

Structures are classified into different safety classes depending on the harm a failure
would cause, the definitions are stated in Table 3.6. For calculations of stability,
according to the partial factor method in EC 0 (2002), the partial factor γd is applied
and this value depends on the safety class of the structure. The partial factor for
the different safety classes is shown in Table 3.7.

Table 3.6: Consequence classes (EC 0, 2002).

Consequences class Description


CC3 High consequence for loss of human life, or
economic, social or environmental
consequences very great.
CC2 Medium consequence for loss of human life,
economic, social or environmental
consequences considerable.
CC1 Low consequence for loss of human life, and
economic, social or environmental
consequences small or negligible.

Table 3.7: Partial factors γd according to safety class (EC 0, 2002).

Safety class Partial factor, γd


1 0.83
2 0.91
3 1.0

Dead weight

The dead weight stabilises the dam and therefore acts as a favourable, permanent
load and is a structural action. For reinforced concrete with normal weight, the
density γc = 24 kN/m3 should be used for calculations of the dead weight (EC 1,
2013).

23
CHAPTER 3. METHODS FOR STABILITY ANALYSES

Hydrostatic pressure

The horizontal water load (HW) acting on the upstream face, shown in Figure 3.5,
is a permanent and unfavourable load while horizontal tailwater load (TW) on the
downstream side and the vertical load (VW) are permanent and favourable loads.
The water loads are categorised as geotechnical actions. For analyses where the
hydrostatic pressure is increased above the headwater level, the hydrostatic pressure
can be classified as a variable load.

Figure 3.5: Hydrostatic pressure.

According to EC 1 (2013) the density for fresh water is set to γw = 10 kN/m3 and
the loads caused by water should be determined with respect to the water level.
No combination factor is given for the water load; therefore, in this report, a value
in the interval between the value for the highest snow load and the value for imposed
loads on buildings was assumed. The combination factor ψ0,w = 0.75 can therefore
be used for variable water loads.

Uplift pressure

The vertical uplift pressure is considered as an unfavourable permanent load and


geotechnical action. The density of water and the partial factor can be set according
to the hydrostatic pressure. An additional horizontal uplift pressure will be present
for monoliths without horizontal bottom surfaces. The horizontal uplift pressure can
act as both an favourable and an unfavourable load, depending which direction the
monolith is inclined. The horizontal uplift pressure resultant can also, depending on
the location, vary between unfavourable and favourable for sliding and overturning.

Ice load

The ice load is an unfavourable variable load and is a structural action. In this
report, the combination factor ψ0,ice = 0.8 was chosen for the load combinations,
based upon the highest value for snow loads.

24
3.2. EUROCODE

Rock anchors

Rock anchors are considered as permanent favourable loads and geotechnical actions.
The design strength of the reinforcing rock anchors may be calculated according to
Equation (3.6).

fyk
fyd = (3.6)
γs

where
fyk is the characteristic strength of the reinforcement.
γs is the partial factor for the reinforcement.
The partial factor applied to untensioned and tensioned reinforcement bars is in this
report assumed to be γs = 1.15 based on EC 2 (2011).

Earth pressure

Earth pressure is both a favourable and an unfavourable permanent load and is in


this report considered as a geotechnical action. Eurocode states that the soil prop-
erties should be chosen from investigations or by theoretical or empirical correlation
or from other relevant documentation. If standard values from tables are used, the
characteristic values should be chosen with great care. According to the design of
retaining wall structures, the determination of the earth pressure should be taken
as at-rest pressure, if no movement of the wall relative the ground takes place. The
lateral earth pressure coefficient, KO is calculated according to Equation (3.7) for
a horizontal backfill and according to Equation (3.8) for an inclined backfill and
depend on the friction angle (EC 7, 2011).
Horizontal backfill: √
KO = (1 − sin ϕ0 ) · OCR (3.7)

Inclined backfill:
KO;β = KO · (1 + sin β) (3.8)

where
ϕ0 is the effective friction angle.
OCR is the overconsolidation ratio.
β is the slope of the soil.
Values for the unit weight density and friction angle in Table 3.8, was obtained from
Trafikverket (2011).

25
CHAPTER 3. METHODS FOR STABILITY ANALYSES

Table 3.8: Material properties for ground materials.

Unit weight density [ kN/m3 ] Friction angle [ ◦ ]


Material Un-saturated Saturated ϕ0
Rockfill 18 11 45
Gravel 19 12 37
Sand 18 10 35
Gravelly moraine 20 13 38
Sandy moraine 20 12 35
Silty moraine 20 11 33

Eurocode also states that the earth pressure should be calculated according to the
chosen design approach, as shown in Equation (3.9). The design value of the earth
pressure is:

Xk
Xd = (3.9)
γM

where
Xk is the characteristic value of the material property.
γM is the partial factor of the material property.
The partial factor for material properties was chosen in accordance with design
approach (DA3) to γM = 1.3. In most cases the earth pressure acts as a stabilising
force, leading to a decreased force, resulting in a more conservative value for the
earth pressure. If the earth pressure is active, the diagram in Figure 3.6 can be used
to obtain the active earth pressure coefficient, Ka .

26
3.2. EUROCODE

Figure 3.6: Active earth pressure (EC 7, 2011).


Figur C.1.1 – Aktiva, effektiva jordtryckskoefficienter Ka (den horisontella delen): stöttad horisontell
markyta (E = 0)

3.2.2 Failure modes

EC 7 (2011) defines how to perform the design of geotechnical structures. The


calculation model should describe the behaviour of the foundation and be reliable
and give an error on the safe side.
It should be verified that ultimate limit state is not exceeded for:
• Internal failure or excessive deformation of the structure or structural elements,
128 in which the strength of the structural material is significant in providing
resistance
Tillägg och kommentarer (STR).har gjorts av Emma Persson. Teknikområde Grundläggning.
i detta dokument Senaste revidering 2013-02-25.
Uppdatering enligt EKS9 har gjorts av Anette Sjölund och Elizaveta Pronina. Senaste revidering 2015-04-16.
• Failure or excessive deformation of the ground, in which the strength of soil
or rock is significant in providing resistance (GEO).

Overturning

The failure criterion presented in Equation (3.10) for ultimate limit state from
EC 0 (2002), defines the safety against overturning as:

Md,dst ≤ Md,stb + Td (3.10)

27
CHAPTER 3. METHODS FOR STABILITY ANALYSES

where
Mdst is the overturning moments.
Mstb is the stabilising moments.
Td is the shearing resistance.
If the shearing resistance, Td , is included, it should not have a considerable effect on
the result.

Sliding

The failure criterion presented in Equation (3.11) for ultimate limit state according
to EC 7 (2011), defines the safety against sliding as:

Hd ≤ Rd + Rp;d (3.11)

where
Hd is the design value of unfavourable horizontal forces.
Rp;d is the design value of favourable horizontal forces
Rd is the design shear resistance.
The design shear resistance is calculated by Equation (3.12).

tan δd
Rd = Vd0 · (3.12)
γM

where
Vd0 is the design value of the effective vertical load.
tan δd is the design friction angle.
According to Eurocode, the friction angle tan δd , should be determined based on
geotechnical investigations.

3.3 Limit turning

3.3.1 Crushing resistance

An important parameter in the limit turning calculations is the crushing resistance


of the rock mass, Rcr . This is a better estimation of the resistance to shear loading
compared to the shear strength parameters friction angle and cohesion. The crushing
resistance of the rock mass should be obtained from geotechnical investigations and
depends on the peak shear and normal stresses acting on the rock. When the
crushing resistance is exceeded and the crushing zone is formed, a limit turning

28
3.3. LIMIT TURNING

failure will occur (Fishman, 2007). From experiments performed by Fishman the
relationship between the crushing resistance and uniaxial stress was obtained, shown
in Equation (3.13).

Rcr = 1.47 · σc (3.13)

If no investigations are available, values from Table 3.9 may be used in the calcula-
tions (Fishman, 2009).

Table 3.9: Crushing resistance Rcr for different categories of rock mass (Fishman,
2009).

Category of rock Type of foundation Parameter Rcr (MPa)


I Massive, large fragmental, laminated, platy, very 20.0
low and low jointed, unweathered rock
characterised by uniaxial compression strength in a
sample σc > 50 [MPa]
II Medium jointed, inconsiderably weathered rock 10.0
characterised by σc > 50M P a
III Intensively jointed rock with σc = 15 − 50M P a and 5.0
inconsiderably weathered and low jointed rock with
σc = 5 − 15M pa
IV Semi-rock, platy, thin-platy, medium, high and very 2.5
high jointed with σc = 5M P a

29
CHAPTER 3. METHODS FOR STABILITY ANALYSES

3.3.2 Failure criteria

The stability factor is the ratio between the sum of resisting moments and the sum
of turning moments. Including the moment from the force of the peak crushing
resistance S. The moments are calculated relative to the turning axis O, as sown in
Figure 3.7.

Figure 3.7: Principles of limit turning, reproduced from Fishman (2007).

The position of O axis is determined as follows:


N
O = (a, d) = ( , [(h2 + 2 · a · e − a2 )1/2 − h]) (3.14)
t · Rcr

The force of peak crushing resistance is defined in Equation (3.15).

S = (a2 + d2 )0.5 · t · Rcr (3.15)

The moment of the peak crushing resistance will be calculated about the O axis and
added to the resisting moment:

Mp.c = S · bcr · 0.5 (3.16)

Limit turning stability factor, relative to the O axis:


ΣMr
Fs = (3.17)
ΣMt

30
3.3. LIMIT TURNING

where
a is the x-distance from the downstream toe B to the position of the turning
axis O.
d is the y-distance from the downstream toe B to the position of the turning
axis O.
N is the resultant of the vertical forces.
T is the resultant of horizontal forces.
t is the width of the structural section along a projected center-line or the
thickness of the buttress.
Rcr is the crushing resistance of the rock.
h is the lever arm of the horizontal forces T relative the downstream toe B.
e is the lever arm of the vertical forces N relative the downstream toe B.
bcr is the length of crushing plane OB.
P
Mr is the sum of resisting moments.
P
Mt is the sum of turning moments.

31
Chapter 4

Failure modes of concrete dams

The aim of this chapter was to detect the factors that might cause concrete massive
and buttress dams to fail. Especially to consider if other than the currently used
design criteria could be relevant. Dams are designed against failure criteria based
on sliding and overturning. By going back and studying failures it is possible to
detect if the failure criteria are sufficient or if other failure modes may have to be
accounted for in the design process.
This chapter includes a compilation of reported concrete dam failures across the
world; how and what have caused them to fail. This study excludes China since the
documentation there is incomplete. Many failures occurred decades ago, and there-
fore the documentation and important information regarding these failures might
be inadequate.
Greater incidents of concrete dams were also included. Known dam failures without
information about either the foundation or the failure cause were excluded.

4.1 Documentation of failures


The aim of the engineering industry today is to take responsibility of establishing a
global collaboration as well as openness to share and increase the general knowledge
of the industry by creating formalised channels such as registers and organisations.
The Committee on Dam Safety (CODS) in particular is working with this. However,
it is difficult to obtain information about particular failures, especially in cases where
failure took place long ago. Other reasons could be that some dam owners are not
willing to admit failure and do not make the records public, or due to a legal policy
preventing publication of records. This has slowed down the technical development
of the industry, limiting possibilities of understanding earlier generations’ thoughts
behind their solutions and designs. The sharing of information and the ability to
talk about dam failures could help increase our knowledge of the field as well as
provide the opportunity to learn from the experience of others, which would greatly
increase the knowledge of the industry (Isander, 2013).

33
CHAPTER 4. FAILURE MODES OF CONCRETE DAMS

4.2 Compiled failures


The study includes failures of 19 concrete dams. These are divided into 12 massive
dams, one gravity spillway dam and six buttress dams, presented in Table 4.1. The
included failures have, to varying degrees, documented information about the failure
and the dam. For a full compilation of the studied dams, see appendix A.

Table 4.1: Concrete massive and buttress dams included in the study.

Height over
Year com- Year of
Dam name Country Dam type lowest
missioned failure
foundation
Bayless1 USA Massive 17 1909 1910(1911)
Camara2 Brazil Massive 50 2002 2004
Eigiau3 GB Massive 10 1911 1925
Elwha river1 USA Massive 51 1912 1912
(hydro-power)
High Falls6 USA Massive 9 1910 1999
Marquette no 36 USA Massive 10 1924 2003
Shih-Kang dam5 Taiwan Massive 22 1977 1999
(gravity spillway)
St Francis1 USA Massive 62 1926 1928
Torrejon-Tajo1 Spain Massive 62 1967 1965
Upriver dam6 USA Massive 12 1937 1986
Warrensburg6 USA Massive 8 1909 1976
Xuriguera1 Spain Massive 42 1902 1944
Zerbino4 Italy Massive 16 1925 1935
Ashley1 USA Buttress 18 1908 1909
Cascade lake dam8 USA Buttress 5 1908 1982
Komoro1 Japan Buttress 16 1927 1928
Morris Sheppard7 USA Buttress 58 1941 1986
Overholser1 USA Buttress 17 1920 1923
Stony creek1 USA Buttress 21 1913 1914

1
(Douglas, 2002)
2
(Shaffner and Scott, 2013)
3
(J Andrew et al., 2011)
4
(Luino et al., 2014)
5
(Kung et al., 2001)
6
(Reegan, 2015)
7
(Anderson et al., 1998)
8
(Jarrett, 1986)

34
4.2. COMPILED FAILURES

4.2.1 Comparison of properties

In Section 4.2, Table 4.1 show the variation in year commissioned, age at failure and
height, these properties are compared in Figure 4.1 and Figure 4.2. The majority of
the studied dams were commissioned before 1940 according to Figure 4.1.

3 Massive dams
Buttress dams
2

Figure 4.1: Year studied dams were commissioned.

After five years

During first five years

Buttress dams
Massive dams
During first filling *

During construction

0 2 4 6 8 10

Figure 4.2: Variation in age at failure of the analysed buttress and massive dams.

The buttress dams had according to Figure 4.2 slightly a higher tendency to fail
during the first five years while the massive dams generally failed after five years.
The majority of the failures did however occur within the first years. Failure is less
feasible for older dams, where the possibility of failures per year decrease with the
age of the dam.

First filling is the first time the dam was filled

35
CHAPTER 4. FAILURE MODES OF CONCRETE DAMS

4.2.2 Failure type

Information about foundation material and geology of the foundation is presented


in Table 4.2. In some cases information is missing, and this is denoted with the
symbol "-". The failure of each dam is referred to a certain failure type, the used
failure codes are defined below.
F f, failure due to dam foundation.
F b, failure due to the structural behaviour of the dam body.
F a, failure due to appurtenant works.
F m, failure due to dam materials.

Table 4.2: Failure types for massive dams.

Dam name Foundation Geology Failure code


material
Bayless Rock Sandstone horizontal layers with shale and Ff
clay between
Camara Rock Plane of micaceous silty clay Ff
Eigiau Clay Hard blue clay containing boulders of granite Ff/Fm
overlain by a layer of peat
Elwha river Soil/rock Fluvioglacial and conglomerate Ff
High Falls Rock - Fm
Marquette no 3 Rock - Fa
Shih-Kang dam Rock Top deposition layer: unconsolidated gravel, Ffb
sands, silts and clay. On Soft bedrock:
slate-gray, sandy-shale and silty-sandstones
St francis Rock Conglomerate and schist Ff
Torrejon-Tajo - - Fa/Fm
Upriver dam Soil - Fa
Warrensburg - - Fa
Xuriguera Rock - Ff
Zerbino Rock Schist and hornfeld Faf
Ashley Soil Fluvioglacial Ff
Cascade lake dam Soil Glacial terminal-moraine sediments Ffa
Komoro Rock Tuff Ff
Morris Sheppard Rock Shale Ff
Overholser Rock - Ffa
Stony creek Soil - Ff

36
4.2. COMPILED FAILURES

4.2.3 Failure mode

Table 4.3 contain information about the failure type by either information about the
failure mode of the foundation marked with the symbol "x", or the failure mode of
the dam, as listed below. The failure mode is the parameter affecting the incident
mode. The following definitions have been used:
P, piping; water and material passing through the foundation.
SC, scour; when sediment surrounding the dam abutment is removed.
S, sliding.
SH, shear sliding within dam.
EQ, earthquake damage.
T /C, tensile and compressive failure within dam.
ST, structural damage to appurtenant equipment, such as spillway gates.

Table 4.3: Failure mode for massive dams.

Dam name Failure mode Failure Failure Failed due to overtopping


foundation mode type
dam code
(P) (SC) (S)
Bayless x Ff Overtopping due to unknown cause
Camara x Ff Not at highest water level
Eigiau x Ff/Fm Not at highest water level
Elwha river x Ff Filled
High Falls (ST) Fm Overtopping due to unknown cause
Marquette no 3 Fa Overtopping due to unknown cause
Shih-Kang dam (EQ) Ffb No information
St Francis x x Ff Gradual during first fill
Torrejon-Tajo (SH) Fa/Fm Flood during construction
Upriver dam x Fa Overtopping due to unknown cause
Warrensburg (T/C) Fa No information
Xuriguera x Ff No information
Zerbino x x Faf Overtopping due to unknown cause
Ashley x Ff J ust spilling when pipe failed
Cascade lake x Ffa Overtopping due to unknown cause
dam
Komoro x x Ff No suggestions of high water level
Morris x Ff Releases kept within channel capacity
Sheppard
Overholser x Ffa Overtopping due to unknown cause
Stony creek x Ff Not clear if failed at top level or
overtopping

37
CHAPTER 4. FAILURE MODES OF CONCRETE DAMS

4.3 Description of failures


Descriptions about the failures for those dams which are not further discussed in
Section 4.3.1 are given in Table 4.4 and Table 4.5.

Table 4.4: Failure description for massive dams.

Dam name Failure description


High Falls Overtopping led to breach of 23 meter long portion
of concrete crest cap, left half the spillway. Repairs
completed.
Marquette no 3 Overtopping and failure of abutment, due to an
failure of upstream dam.
Torrejon-Tajo Shear sliding within the dam. Failure cause was
traced to organic material present in the aggregate
and filling of the dam by a flood during
construction before the concrete had fully hardened.
Upriver dam Washout of the abutment and the power canal
embankments due to overtopping. Not a complete
failure and reparation of the dam was possible.
Warrensburg Breach of north abutment. Reconstructed in 1998.
Xuriguera Failed by foundation sliding, shear strength and
poor design.

Table 4.5: Failure description for buttress dams.

Dam name Failure description


Ashley Piping failure in fine sand with clay and gravel, 6m
deep below cut-off.
Stony creek Piping in foundation followed by settling of dam,
cracking and collapse of dam.
Cascade lake dam The dam was overtopped before tipping over and
failing. The cause of failure was the hydrostatic
water pressure on the dam and erosion of the
abutments. Stored water was released rapidly due
to short time of breach development and the width
of the breach was large.
Komoro Failure due to softening of volcanic ash in
foundation. Unclear cause, either piping, sliding or
both.
Overholser Overtopping leading to scour of abutment.

38
4.3. DESCRIPTION OF FAILURES

4.3.1 Documentation regarding failures

Bayless Dam, USA

Flaws in the Design of the Dam


A cut-off wall (shear key) was installed to suitable bedrock and steel rods were built
into the wall and secured in the rock (The Engineering News Publishing company,
1910). Against the upstream face, an embankment dam was placed, composed of
compacted disintegrated shale, clay and some loam. The intention of the embank-
ment was to prevent water percolating down to the porous strata beneath the dam.
The engineer desired a deep cut-off wall that was to be constructed down through
the rock strata, but was overruled due to its cost.
When the dam was completed one small vertical crack followed by more cracks,
appeared on one side of the spillway. This was due to contraction since there was
no water in the dam. The Bayless dam prior to failure is shown in Figure 4.3.

Figure 4.3: Bayless Dam prior to failure (The Engineering News Publishing com-
pany, 1910).

First failure
One year after the dam was completed, rapid melting of large amount of snow
occurred and within three days, the dam was filled to maximum capacity. The next
day, a large slice of earth below the dam dropped down 1.2 m and partially slid into
the valley. The water retained within the dam eroded a path beneath where the
earth had slid. Eventually the water began flowing up through the ground in large
quantities, 5-15 m downstream from the dam toe. The result was that the water
flowed under the dam in the embankment through the rock strata (The Engineering
News Publishing company, 1910). During the third day, the flowing water resulted
in that a portion of the dam, at the overflow spillway section, slided 0.4 m at the
top and 8 m at the base, causing the crack widths to increase at the downstream
face which unloaded the cracks at the upstream face. The movement lasted for eight

39
CHAPTER 4. FAILURE MODES OF CONCRETE DAMS

hours resulting in overflowing of the dam. It took 16 hours to completely empty


the reservoir. The total failure time of the dam was approximately two days. The
reservoir was lowered and no repairs were made before the dam was put back into
service (DOI, 2012).
The initial cause of failure was piping, which caused softening of the clay and shale,
lying between two layers of rock, causing the top layer of rock to slip forward onto the
lower layer. The results from the failure were leaks under the dam, transverse cracks
in the main section and movement of the central part downstream (The Engineering
News Publishing company, 1910).
Second failure
At the end of summer, in 1911, the dam failed for the second time, due to high
headwater level, nearly as great as what caused the first slip failure. The dam failed
in 30 minutes with no indications of a gradual failure. Four-fifths of the length of the
dam broke into several large fragments, see Figure 4.4; most of them remained nearly
vertical. The two largest fragments near the centre (spillway section), fragment E &
D in Figure 4.4 shifted downstream and rotated slightly from its original alignment
and, on both sides of these, large gaps were formed. At the west end, 38 m of
the dam was still intact, fragment G in Figure 4.4 and to the east, fragment B
& C the entire dam was broken and displaced. The failure was so extensive that
no conclusion of the initial point of failure could be drawn. The appearance of the
failed dam, however, indicates a sliding failure. Observations from the wreckage may
suggest that the westerly gap with its sections, slit and sheared out at levels above
the foundation, as a secondary effect (The Engineering News Publishing company,
1911).

Figure 4.4: Bayless Dam, resulting fragments and positions after failure (The Engi-
neering News Publishing company, 1911).

40
4.3. DESCRIPTION OF FAILURES

Camara Dam, Brazil

Flaws in the Design of the Dam


The dam failed due to lack of information regarding the foundation. The smooth
foliation surfaces were left untreated leading to failure during the first filling. Indica-
tions from material in the drains, leakage through the concrete and clogging of drain
holes, etc. suggests that failure was about to occur. The designer recommended emp-
tying the reservoir however due to political reasons it was not implemented (Shaffner
and Scott, 2013).

Figure 4.5: Camara Dam, failure at foundation of left abutment (Risk Assessment
International, 2013).

Failure
The reservoir was filled rapidly in about two weeks, due to heavy rains, and the
retained volume gradually increased thereafter. After about five months when the
reservoir was about 3/4 full, it failed at the foundation of the left abutment, due
to erosion from soil-filled discontinues. High pressure gradients developed under
the dam. As the flow rate of the reservoirs’ retained water increased the erosion
and driving forces on the low-shear strength rock slabs, the dam and the abutment
started to slide until a hole was piped beneath the dam causing a flood. The remains
are shown in Figure 4.5. The remaining rock in the left abutment is essentially free
from fractures indicating that piping occurred along the entire length of fracture
(Ferc Engineering Guidelines, 2014).

41
CHAPTER 4. FAILURE MODES OF CONCRETE DAMS

Eigiau Dam, United Kingdom

Flaws in the Design of the Dam


Different contractors were involved in the construction of the dam. The footings
should have been founded 1.8 m below the clay surface, but at the point of failure,
only 0.5 m was embedded into the clay layer. From investigations, it was suggested
that poor quality concrete contributed to the failure. The concrete lacked the correct
volumes of sand and cement, resulting in the aggregates not cementing together. The
stone aggregates were larger than desirable and were placed carelessly; in several
cases with voids under their bed surfaces (J Andrew et al., 2011).

Figure 4.6: Eigiau Dam, remaining breach (Geograph, 2010).

Failure
A 10 m long breach occurred in the concrete at the side leg of the dam, see Figure
4.6. The breach scoured a 20 m wide channel three meters below the ground surface.
Large portions, 1.5 million cubic meters, of water were released in the first hour
and caused Coedty dam 2.5 miles downstream to overtop and the concrete wall to
collapse. In itself, the soft, porous condition of the boulder clay was aggravated
by cracking due to dehydration in the preceding summer when the lake bed was
exposed (J Andrew et al., 2011).

42
4.3. DESCRIPTION OF FAILURES

Elwha Dam, USA

Flaws in the Design of the Dam


The dam, seen in Figure 4.7, was founded on a deep gravel deposit and water there-
fore blew out the foundation (Oakes, 2001). The contractors ignored specifications
from the engineers, resulting in the dam not being secured to the bedrock.

Figure 4.7: Elwha Dam prior to failure (KPLU 88.5, 2011).

Failure
The dam failed during the first fill due to piping failure of the alluvium under the
dam. Lower sections of the dam were removed by the flowing water and large
portions of material eroded under the dam in about two hours resulting in a hole.
Various reparation methods were attempted and finally the hole was filled with de-
bris. The Elwha dam failure started serious environmental investigations, regarding
removal of the dam to restore the river to its natural self-regulating state. The dam
was finally removed 2012.

43
CHAPTER 4. FAILURE MODES OF CONCRETE DAMS

Morris Sheppard Dam, USA

Flaws in the Design of the Dam


The stability of the foundation was based on peak shear strengths and did not
consider uplift pressures (Anderson et al., 1998).
Failure
Sliding of the spillway section, see Figure 4.8, on the shale foundation, most probably
occurred over a period of several years and was discovered during a routine inspection
45 years after construction.

Figure 4.8: Morris Sheppard dam prior to failure, spillway section (Anderson et al.,
1998).

Floatation of the lower portion of the hollow spillway section, together with the
low resistance to sliding, added to the tendency, of the hollow spillway to move
downstream. The reservoirs’ water level was quickly lowered to avoid complete
failure.
The metal survey points, that were installed along a line, had formed a bow, which
indicated that the hollow spillway section had moved downstream on a slippage zone
in the foundation, and from observations, cracks in the footings were found.
Core borings were made, which indicated that the hydrostatic uplift pressure, under
the spillway slab, was 65 % of the retention level. From the borings a longitudinal
crack along the top of the upstream cut-off was located. This may have allowed for
water to enter the foundation causing significant pressure beneath the shale layer.
Drainage wells were installed to reduce the uplift pressures. A network of measur-
ing equipment was installed as to keep track of any movement or change in uplift
pressure.

44
4.3. DESCRIPTION OF FAILURES

Shih-Kang Dam, Taiwan

Flaws in the Design of the Dam


The Shih-Kang dam was designed according to the traditional design concept of a
pseudo static earthquake acceleration (Kung et al., 2001). The effect from the ver-
tical motion was, however, neglected. The original pseudo static horizontal acceler-
ation was less than the real peak horizontal acceleration of the Chi-Chi earthquake,
which, in turn, caused sliding failure.
Failure
The dam was damaged by surface ruptures caused by; an active fault, the large
displacement of ground surface and great ground motion induced by the Chi-Chi
earthquake. The dam moved in a north-west direction, 10 m vertically and 11 m
horizontally. The ground deformation caused the dam body to crack and separate
from the foundation. The stiffness of the structures affected the deformation of the
dam body.
The left side rock gradually rose towards the upper end of the fault-created scarp.
The north part, between the spillway and the abutment, was cracked along the con-
struction joints into several huge blocks, see Figure 4.9. Water then leaked through
the cracks. The entire dam body seemed to have bent towards the downstream side.

Figure 4.9: Shih-Kang Dam post failure (HSS, 2013).

The water level was decreased to reduce pressure on the dam and complete failure
was avoided.

45
CHAPTER 4. FAILURE MODES OF CONCRETE DAMS

St. Francis Dam, USA

Flaws in the Design of the Dam


The dam, see Figure 4.10 was not designed with the correct uplift theory; uplift
pressure acting to destabilise the sloping abutments. The cross section contained
limited seepage relief, only a few uplift relief wells beneath the central core and no
overlapping for expansion joints. To increase reservoir storage the dam height was
increased 6 m without any substantive widening of the dam base width. The dam was
arched upstream but arch action was neglected in the design. During initial filling
several transverse cracks appeared which were filled in and sealed. The engineers
did not understand the concepts of effective stress and uplift. If the foundations had
been deeper, a cut-off wall and a grout curtain, with seepage relief wells would have
been installed, the failure may had been avoided (Rogers, 1995).

Figure 4.10: St:Francis Dam prior to failure (Water and power, 2015).

Failure
The reservoir had been held within 7 mm from the spillway for five days and the dam
became unstable. At the west abutment a new larger leak and soiled discharge were
detected on the morning of the failure, caused by hydraulic piping. The dams’ west
abutment, built up on a fault contact, was unknowingly founded upon massive paleo
mega-slides, causing cracking between the rock in the mid valley and the abutment
(Rogers, 1995). Eventually the entire left hand side failed, inducing a domino effect
of block failure at the right abutment (Veale and Davison, 2011). The blocks cut
off along the transverse crack resulted in large leaks at the toe of the east abutment
slide and 12 hours later, the dam collapsed. Forty minutes prior to the failure, the
water level was rapidly decreasing.
During final filling, a massive land slide occurred along the dams’ left abutment,
carrying blocks and cutting off parts of the dam as the reservoir rose to full pool.
The slide material initially plugged the outflow until the slide material was eroded by

46
4.3. DESCRIPTION OF FAILURES

the increasing flood wave. Full hydrostatic pressures entered the transverse cracks
causing hydraulic uplift to eliminate the dams’ stabilising dead load. A sudden and
dangerous overstressing of the dams’ cantilevered load capacity lead to excessive tilt
and overturning, causing the upstream heel to go into tension, full hydrostatic head
pressure was introduced within the dam structure which then finally cracked. The
central core of the dam was tilted towards the east abutment and tension developed
in pre-existing cracks. Finally the west abutment was scoured away causing the
central core to rotate (Rogers, 1995).
The dam appeared normal shortly before the failure; it failed 7.5 minutes later and
emptied within an hour. All that was left was a single monolith standing in the
middle of the valley, see Figure 4.11.

Figure 4.11: St:Francis, resulting blocks from the abutment failure (Los Angeles
Times, 2013).

Zerbino Dam, Italy

Flaws in the Design of the Dam


The dam was constructed 300 m west of the Main Dam of Bric Zerbino, where
a saddle, formed by two ridges, was at a lower elevation which could have been
overflowed and poured out into the riverbed (Luino et al., 2014). The dam, shown
in Figure 4.12 was built rather hastily and without sufficient geologic investigations.
It was assumed that the saddle consisted of sound rock. The dam stood on highly
jointed schist’s making up a particularly weak zone within the rock mass. Water
leaks were noticed across the rock diaphragm. Attempts to make the rock mass
impervious were made with no satisfactory results. Miscalculation of precipitation
in the area caused the water levels in the dam to differ from the values used in the
design.

47
CHAPTER 4. FAILURE MODES OF CONCRETE DAMS

Figure 4.12: Zerbino Dam prior to failure (Molare.net, 2010).

Failure
Heavy rains caused the level of the reservoir to rise quickly and therefore the bottom
discharge valves were opened to handle these large volumes. The large amount of
mud and debris accumulating at the bottom of the dam caused the valves to stop
working after only a few minutes. The reservoir could then only be discharged at the
surface spillway and the siphons. Water then started to overflow, see figure 4.13 into
both the main and the secondary dam, Zerbino. This caused repercussion on the
jointed rocks that made up the saddle. After just one hour, the large water pressure,
of the overflowing reservoir, displaced large rock blocks causing destabilisation of the
dam which lead to the collapse. The Zerbino dam failed due to scour and sliding
during overtopping. The level of the reservoir went down by some 25 m within a few
minutes.

Figure 4.13: Zerbino Dam, flooded reservoir (Molare.net, 2010).

48
4.4. RESULTS OF THE COMPILED FAILURES

4.4 Results of the compiled failures


The comparison of the properties for the documented failures in Section 4.2.1 in-
dicate that majority of the massive dams failed after five years, in fact in a wide
range of 10-90 years after the dam was commissioned. This corresponds well with
the fact that the majority of the failed dams were built 1900-1940. The majority
of the dams stood for many years, which could indicate ageing of concrete as a
contributing factor. All but one of these dams failed due to overtopping, showing
that the massive dams were sensitive to additional loading. The massive dams could
have been designed against lower loads than they were subjected to at the time of
failure. The majority of the buttress dams failed during the first five years, due to
foundation failure.
The failure types described in Section 4.2.2 are compared in Figure 4.14. The used
failure codes are listed below.
F f, failure due to dam foundation.
F b, failure due to the structural behaviour of the dam body.
F a, failure due to appurtenant works.
F m, failure due to dam materials.

3 Massive dams
Buttress dams
2

0
Ff Ffa Ffb Ff/Fm Fa/Fm Fa Fm

Figure 4.14: Failure types of the studied massive and buttress dams.

Of the studied failures, the majority failed due to, or partly due to, the dam founda-
tion. This indicate the importance of a good foundation, meaning that information
about the material and properties of the foundation play a crucial part in dam de-
sign i.e. dam stability. This could be achieved by geological studies, core samples,
material testing, site visits, etc.

49
CHAPTER 4. FAILURE MODES OF CONCRETE DAMS

Figure 4.14 indicates that the failure type for massive dams are mainly due to
different combinations of foundation and dam failure. Massive dams are heavy
structures with a large amount of concrete volume, which requires a strong enough
foundation and correct casting arrangements. For buttress dams, foundation failure
is the dominant failure mode. The reason could be that buttress dams are light
structures. The casting sequence and construction of buttress dams are complicated
where faults are more prone to occur.
The failure modes descried in Section 4.2.3 are compared in Figure 4.15, with the
definitions listed below.
P, piping failure.
SC, scour failure.
S, sliding failure.
SH, shear sliding within dam.
EQ, earthquake damage.
T /C, tensile and compressive failure within dam.
ST, structural damage to appurtenant such as spillway gates.
N o, no information was available.

2 Massive dams
Buttres dams

0
P S/SC S S/P SH EQ No T/C ST SC
info
Figure 4.15: Failure modes of the studied massive and buttress dams.

The majority of the failures occurred in the foundation, seen in Figure 4.15, which
in some cases could have been avoided if improved geotechnical investigations were
performed. Figure 4.15 also shows that failures in buttress dams are mainly caused
by piping, sliding or both. Since not so many cases of failed buttress dams are
included in this study, it is hard to conclude if there is a coincidence or if sliding and
piping is what mainly causes the buttress dams to fail. From the results, piping affect
buttress dams more than massive dams, not surprising considering that the water

50
4.4. RESULTS OF THE COMPILED FAILURES

only have to travel underneath the frontplate of the buttress. For a massive dam
there is a much longer distance for the water to erode before it causes failure. Sliding
is more common for buttress dams, also not surprising since the less use of concrete
volume, and therefore also the vertical component acting on the sliding surface is
smaller. This is not always the case for the buttress dams since the inclination of
the frontplate results in an additional vertical water load.
The initial, or the main cause of failure is shown in Figure 4.15, although when
the failure starts to propagate it will proceed to fail due to other failure modes or
combinations of failure modes. The dam failures described in Section 4.3.1 above,
give a clear indication that there is a combination of failure modes leading to the
final failure of the dam.
Out of all the studied cases, none failed due to global overturning; this indicates that
it is mostly a theoretical failure or that this failure mode is designed with high safety
margin. The only indications of overturning were found after the initial failure, as a
local failure of parts of the whole structure. This might create questions regarding
the importance of this design criterion, however it is still an adequate and accurate
indicator of the dams’ stability.
As shown in Figure 4.15, only one dam failed within the dam body, which occurred
during construction before the concrete had cured properly, which therefore does not
suggest that the global stability is insufficient, but rather that special considerations
should be made also during construction.
Figure 4.16 shows a summation of failures caused by faults in the dam design or
during construction of the dam.

Inadequate ground
investigations

Poor construction
Number of failures

Unknown

0 5 10 15

Figure 4.16: Faults in dam design or construction for the studied dams.

From the failures described in Section 4.3.1 above it is clear that, to some extent,
the neglect of vital information, or requirements, occurred in all cases. It is hard

51
CHAPTER 4. FAILURE MODES OF CONCRETE DAMS

to know if there were any relevant reasons for these assumptions, especially in the
cases of poor construction.
Failure due to overtopping is shown in Figure 4.17, where it is detected that extreme
floods affect the behaviour of the dam. Even though the height of the water level
is not the decisive factor for the failure, high water levels result in higher water
pressures affecting the concrete.

No information

Not at highest water level

Before first fill Buttress dams
Massive dams

Overtopping

No suggestions of high water
level

0 1 2 3 4 5 6

Figure 4.17: Failure due to overtopping for the studied buttress and massive dams.

52
Chapter 5

Stability analyses

In the analytical analyses, several different dams were studied with varying geometry
and loading conditions. Stability calculations were performed with Matlab R2013a
according to both RIDAS and Eurocode as defined in Section 3.1 and Section 3.2.
Some of the analysed dams do not fulfil today’s failure criteria, due to that the dams
were designed according to different presumptions. Improvements have in many
cases been performed to fulfil the more recent criteria. The structural rehabilitation
were for some cases included, depending on the obtained drawings and information.
The dams were subjected to the first normal load case according to RIDAS, with the
headwater level of the dam, full ice load and all gates closed. For the calculations
according to Eurocode, the loads were combined according to design approach 3,
see Section 3.2.1.
In addition the stability analysis tool, CADAM was used for the massive dams for
comparison. The failure criteria were calculated and compared to the analytical
calculations.
The analytical calculations according to Eurocode were further investigated through
a parametric study. The aim was to establish the most influential parameter, mod-
ify the design loads by adjusting the partial factor and finally obtain results in
agreement with the results from the stability calculations according to RIDAS.
The dams were also analysed analytically for the failure mode limit turning. The
dam with the largest difference in safety factor for overturning and limit turning, in
addition to a dam with a minimal difference in the safety factor for overturning and
limit turning, was analysed with the FEM software BRIGADE Plus 5.2.

5.1 Studied dams


The calculations were performed for a variation of dam types subjected to different
loads. The geometry of the different types of studied dams is shown in Figure 5.1
and Figure 5.2. These figures show the headwater level, the rotation point and the

53
CHAPTER 5. STABILITY ANALYSES

characteristic loads applied to the different dam types, i.e. tailwater level and soil
level.

Figure 5.1: Geometry of the spillway Dam 9 (Left) and pillar Dam 11 (Right)

Figure 5.2: Geometry of the buttress Dam 14 (Left) and massive Dam 2 (Right).

5.1.1 Input data

The required information about the input data for the different dams was collected
from Section 3.1.1 for calculations according to RIDAS and Section 3.2.1 for calcu-
lations according to Eurocode.
Recommended material values, characteristic values and partial factors were chosen
according to Eurocode to the extent it was possible. Standard values from RIDAS
were used if the required information was not found in Eurocode, which was the
case for both the uplift pressure and the ice load. Usually, the uplift pressure for
pillars is calculated with a distribution for massive dams to account for uncertainties
regarding the uplift pressure. This was applied to the analysed pillars in this report,
instead of the distribution suggested in Section 3.1.1 according to RIDAS.
For the calculations performed according to RIDAS, the density of water
ρ = 1000 kg/m3 and the gravitational force g = 9.81 m/s2 was used.

54
5.1. STUDIED DAMS

In this report dams were classified as safety class 3, since dam failures could result
in the loss of human life and are likely to have great economic consequences. The
partial coefficient γd = 1.0 was applied to the loads according to Eurocode, see Table
3.7.
Eurocode does not mention when rock bolts should be accounted for and therefore
the guidelines from RIDAS were applied. For calculations according to Eurocode,
the common reinforcement steel Ks40 was used, with a characteristic strength fyk =
370 MPa for φ = 25 mm and fyk = 350 MPa for φ = 32 mm (Ljungkrantz et al.,
1994). For the calculations according to RIDAS, the load capacity of 140 MPa was
used for the rock bolts.
In the calculations according to Eurocode, the lateral earth pressure KO was calcu-
lated according to Equation (3.7) from Section 3.2.1 with OCR = 1. The knowledge
of the shear resistance Td for the studied dams was limited and hence it was excluded
from the calculations.
In Eurocode, the geotechnical design is more based on investigations compared to
RIDAS. Since the value of the friction angle is not defined in Eurocode, a standard
value of δd = 45◦ for rock foundations was used for the calculations of the sliding
criterion. For calculations according to RIDAS, values from Table 3.3 in Section
3.1.2 was used for µmax to calculate the stability against sliding.

55
CHAPTER 5. STABILITY ANALYSES

Geometry

Table 5.1 provides an overview of the geometry of the studied dams. In some cases
there is no value assigned to the geometric parameter, and this is denoted with the
symbol "-". Under the column for drainage "x" is placed if the dam has a drainage
system. The following abbreviations are used in Table 5.1 and Table 5.2:

M, massive dam.
B, buttress dam.
S, spillway.
P, pillar.
S + P, spillway and pillar.
GW, width of massive/pillar/spillway.

Table 5.1: Input data for the geometry used in the stability calculations.

Dam Type Height GW Buttress Frontplate Inclination Inclination Inclination Drainage


width width upstream downstream sliding plane
face face
[m] [m] [m] [m] [◦ ]
Dam 1 M 5 10 - - 1:1 10:6.5 0 -
Dam 2 M 8 10 - - 1:1 10:6.5 0 -
Dam 3 B 40 - 2 8 Varies 35:10 0 -
Dam 4 B 12 - 2 8 45:10 14:10 0 -
Dam 5 M 6 1 - - 10:1 2:1 11.3 -
Dam 6 B 6 - 2 8 45:10 14:10 0 -
Dam 7 M 13 1 - - 14:1 32:10 4.1 x
Dam 8 S 7 1 - - 1:1 Varies 4.8 -
Dam 9 S 5 1 - - 1:1 Varies 0 -
Dam 10 P 13 2.25 - - 1:1 Varies 16.4 -
Dam 11 P 7 2.37 - - 1:1 8.5:1 0 -
Dam 12 S 4 1 - - 1:1 Varies 0 -
Dam 13 P 7 1 - - 1:1 35:10 0 -
Dam 14 B 20 - 2 8 Varies 35:10 0 -
Dam 15 P 16 4 - - 1:1 25:10 0 x
(S+P) S 2 17 - - Varies Varies 0 x
Dam 16 P 19 3.75 - - 1:1 Varies 0 -
Dam 17 P 19 2.65 - - 1:1 Varies 0 -
Dam 18 P 18 1.85 - - 1:1 Varies 0 -

Loads

Table 5.2 provides an overview of the applied loads acting on the dams. When a
load was not applied to a dam it is denoted with the symbol "-".

56
5.1. STUDIED DAMS

Table 5.2: Input data for the loads used in the stability calculations.

Dam Type Ice Soil Rock Strength Head Soil level Tailwater Rotation
Load material bolts tendons water level point
φ level
[kN] [mm] [kN] [m.a.s.l]1 [m.a.s.l] [m.a.s.l] [m.a.s.l]

Dam 12 M 200 Moraine 25 - 745.21 744.9 - 740.5


Dam 2 M 200 Moraine - - 745.21 744.9 - 738.1
Dam 3 B 200 - - - 273 - - 235.4
Dam 4 B 200 Rockfill 25 - 221.4 220 - 210.0
Dam 5 M 100 Moraine - - 107.3 107.3 - 103.3
Dam 6 B 200 - 25 - 221.4 - - 216.0
Dam 7 M 100 Moraine - - 107.3 105.5 99.5 96.3
Dam 8 S 25 - 32 - 107.3 - 100.6 98.4
Dam 9 S - - - 500 107.3 - 99.5 97.8
Dam 10 P 100 Rockfill - 500 107.3 98.4 102.2 99.0
Dam 11 P 50 - - - 37.47 - - 31.06
Dam 12 S - Gravel - - 26.45 - - 22.31
Dam 133 P 50 - - - 15.05 - - 10.75
Dam 14 B 200 - - - 273 - - 254.64
Dam 15 P 200 - 25 - 125.3 - 114.3 111.8
(S+P) S - - 25 - 115.5 114.3 111.8
Dam 16 P 200 - - - 181 - - 164
Dam 17 P 200 - - - 181 - - 163.5
Dam 18 P 200 - - - 181 - - 164.5

1 meter above sea level.


2 Rock bolts accounted for according to RIDAS.
3 Additional loads from a gate, including the dead weight as well as the hydrostatic load
acting on the gate.

Limit turning

The safety factor for limit turning was calculated with the crushing resistance
Rcr = 20 MPa for all the studied dams, from Table 3.9. From the results of the ana-
lytical calculations, the dam with the largest difference in the safety factor for over-
turning and limit turning and the dam with the small difference in the safety factor
for overturning and limit turning, was analysed with the FEM software BRIGADE.
Additionally parametric analyses of the crushing resistance Rcr was performed to
obtain the relationship between the crushing resistance and the safety factor for
limit turning for the dams analysed with BRIGADE.

57
CHAPTER 5. STABILITY ANALYSES

5.1.2 Previously studied dams

In the report by Fouhy and Rios Bayona (2014), Dam 4, Dam 6 and Dam 15-
18 in Table 5.1 and Table 5.2, was studied. Their study contains a probability-
based analysis for evaluation of the stability of dams to account for the omission
of uncertainties. The aim of their study was to find a reliability index, β-target
value, applicable to the stability analyses for sliding and overturning of concrete
dams. Fouhy and Rios (2014) also performed a deterministic analysis to enable an
interpretation of the results from the probability-based analysis. They performed
their analysis with three different load combinations, ’Load Combination 1’ which
corresponds to the first normal load case in RIDAS was off interest for this report.

The differences compared to this report is presented below.

• The sliding criterion was based on the Mohr-Coulomb equation which includes
cohesion.
• Different values for concrete density.
• The ice load was based on a report by Adolfi and Eriksson (2013), which in
turn is based on collected values for the annual maximum.
• A higher value for the characteristic yield strength of steel was used.

The results from Fouhy and Rios for load combination 1 are presented in Table 5.3.
A brief comparison of the β-target values with the values stated in Eurocode, see
Table 5.4, was included in the report. The β-target values have a reference period of
one year, which correspond to the probability of dam failure evaluated over a period
of one year (Westberg, 2010).

The highlighted values in Table 5.3 are those who do not pass the criteria, either
according to RIDAS for the deterministic calculations or according to Eurocode,
RC3, for the probabilistic calculations. This enable us to easily note how high
β-values that were obtained and the difficulties in comparing these to Eurocode.
There is a bad correlation between the β-value and the safety factor. For the sliding
criterion only one dam fail according to the deterministic calculations while all but
one fail in the probabilistic calculations. For the overturning criterion all values
obtained from the probabilistic calculations are well above the limit according to
Eurocode.

58
5.2. STABILITY CALCULATIONS

Table 5.3: Results from Fouhy and Rios Bayona (2014).

Probabilistic Deterministic
Dam
β for overturning β for sliding Overturning Sliding

Dam 4 8.57 3.93 1.56 0.57


Dam 6 14.98 5.01 1.21 1.07
Dam 15 8.54 2.12 1.57 0.65
Dam 16 12.37 2.47 2.29 0.74
Dam 17 11.53 5.24 2.1 0.56
Dam 18 11.69 2.77 2.07 0.73

Table 5.4: Recommended minimum reliability values according to Eurocode 1990.

Safety class Minimum values for β Probability of failure/year

RC3 5.2 10−7


RC2 4.7 10−6
RC1 4.2 10−5

5.2 Stability Calculations

5.2.1 Design approaches

Overturning

The failure criterion for overturning was calculated according to Equation (3.1) from
Section 3.1.2. The safety factor s should satisfy Equation (5.1).

Mstab
s= > 1.5 (5.1)
Mover

In the calculations according to Eurocode, the failure criterion for overturning


Equation (3.10) from Section 3.2.2, should satisfy Equation (5.2).

Md,stb
s= ≥ 1.0 (5.2)
Md,dst

59
CHAPTER 5. STABILITY ANALYSES

Sliding

The failure criterion for sliding according to RIDAS, was calculated according to
Equation (3.2) from Section 3.1.2. The friction coefficient µ should satisfy
Equation (5.3), with µmax for the rock foundation, from Table 3.3.

RH
µ= ≤ µmax = 0.75 (5.3)
RV

The failure criterion for sliding Equation (3.11) from Section 3.2.2, according to
Eurocode, should satisfy Equation (5.4).

Hd − Rp;d
≤ 1.0 (5.4)
Rd

Limit turning

The safety factor was set to fulfil the value for overturning according to RIDAS,
from Table 3.2, and should satisfy Equation (5.5).

The safety factor for limit turning, about the O axis:

ΣMr
Fs = > 1.5 (5.5)
ΣMt

Calculation method

The calculations were executed in Matlab where a numerical analysis tool was
developed for the stability analyses. These analyses were performed according to
Figure 5.3.

60
5.2. STABILITY CALCULATIONS

SPILLWAY MASSIVE BUTTRESS PILLAR

Define attributes Plot dam


 Frontplate
 Hole
 Gate Calculate geometry

Calculate Loads

Water pressure Tail water pressure Earth Uplift Rock bolts Ice load
Pressure /tendons
 Horizontal  Vertical tailwater Horizontal and Horizontal
water pressure pressure Vertical and vertical uplift:  Horizontal force:
 Vertical water  Horizontal horizontal  If tailwater force  If not
pressure if tailwater pressure earth pressure:  If hole  Vertical buttress
inclined if inclined  Soil same  If inclined force  else
upstream face downstream face density foundation
 Vertical water  Soil diff. else
pressure if density Vertical uplift
spillway

Stability

If not Buttress Partial factors SLIDING Partial factors If Buttress


Load combinations CRITERION Load combinations

Input: Loads SLIDING Loads Input:


Rotation points’ Horizontal CRITERION Horizontal Rotation points’
coordinates Vertical RIDAS Vertical coordinates

Level arms Rotating Rotating Level arms


moment moment
Stabilising OVERTURNING Stabilising
moment RIDAS moment
Input: Rcr Input: Rcr

 Position of O  Position of O
OVERTURNING
axis Partial EUROCODE Partial axis
 Foundation factors factors  Foundation
crushing crushing
resistance resistance
LIMIT
TURNING

Figure 5.3: Stability calculations in Matlab.

61
CHAPTER 5. STABILITY ANALYSES

5.2.2 Parametric study

In the calculations according to Eurocode, the safety is defined on the loads by a


partial factor. Since these calculations were performed according to design approach
DA 3 for retaining walls, the values for the partial factors may not be adequate for
concrete dams, therefore a parametric study was performed. The aim was to define
the parameter with the greatest influence on the stability by calculation of the
weighting factor, see Equation (5.6). The design load could then be modified by
varying the partial factor to obtain a safety factor in agreement with the results
from RIDAS.

The parametric study was performed for the loads acting on the dams, i.e. the
parameters for the sliding and the overturning criterion were analysed. The impor-
tance of each load was calculated in relation to the total loads affecting the structure,
according to Equation (5.6).

xi
αi = p 2 (5.6)
xi + ... + x2n

where

xi is the value for the load studied.


α is the weighting factor.

Equation (5.7) should then be fulfilled for all the included parameters.

αi2 + ... + αn2 = 1 (5.7)

5.3 CADAM

The computer program CADAM is a tool used to analyse structural behaviour and
safety for concrete massive dams. The program can be used to perform 2D analyses
of a single monolith, assuming a unit thickness of 1 meter (Leclerc et al., 2001).

CADAM was used to analyse the stability for the massive dams included in the
analytical calculations. The intention was to investigate if the program is a usable
tool and equivalent to the analytical calculations. CADAM enables analyses of the
crack length, which was utilised to examine the contact between the dam body

62
5.3. CADAM

and foundation, known from the literature study in Chapter 4 to have a significant
influence on failure.

5.3.1 Stability calculations

In CADAM the gravity method is used to perform the stress analyses to determine
the crack lengths and the compressive stresses. Along with the stability analyses to
obtain the safety margin against sliding and the placement of the resultant for all
forces acting on the structure. This makes it possible to evaluate stability against
sliding and overturning of the dam (Leclerc et al., 2003).

The method is based on rigid body equilibrium to determine the internal forces
acting upon the joints in the dam and the rock-concrete interface and beam theory
to determine the stresses (Leclerc et al., 2001).

In CADAM sliding is defined as:

(ΣV + U ) · tan φ + cAc


SSF = (5.8)
ΣH

where

ΣV is the sum of vertical forces excluding uplift pressure.


U is the uplift pressure force resultant.
φ is the friction angle.
c is the cohesion.
Ac is the area under compression.
ΣH is the sum of horizontal forces.

Overturning is defined as:

ΣMS
OSF = (5.9)
ΣMO

where

ΣMS is the sum of the stabilising moments.


ΣMO is the sum of the destabilising moments.

63
CHAPTER 5. STABILITY ANALYSES

5.3.2 Modelling

Input data

The massive dams presented in Section 5.1.1; Dam 1, Dam 2, Dam 5 and Dam 7
were analysed.

The density of the concrete was set to 2300 kg/m3 . The friction angle was defined
as 45 ◦ and the cohesion was assumed to be zero for all dams. The loads presented
in Table 5.2 were applied to the dams.

Model definition

The geometry was defined as for the analytical calculations and the masses and ma-
terials were defined. The geometry of Dam 2 and the applied loads; water pressure,
ice load and resultants of the earth support fill are shown in Figure 5.4.

Dam 7 have a complicated geometry of the crest and was therefore simplified due
to limitations in CADAM, which only allows solid geometries with straight lines.

Figure 5.4: Section of Dam 2, showing geometry and loading in CADAM.

Load combinations

For dams with earth support fill the force resultants for earth pressure from the
analytical calculations were applied, using the user defined loads as point loads.
The loads where combined according to normal load combination.

64
5.4. FE-ANALYSIS

For Dam 1 with rock bolts, the resultant force of the bolts was applied as two point
loads and positioned where the rock bolts intersect the interface of concrete and the
rock. The force was applied with an elevation of 0.1 m in order for the program to
apply the force on the structure.

Cracking and uplift options

Specification of the cracking option was defined as tensile strengths for crack initia-
tion and propagation. The calculations were performed with constant uplift pressure
and with modified uplift pressure after cracking of the dam initiated.

The drainage system for Dam 7 was calculated with the option USACE 1995, which
enables the reduction of the uplift pressure as stated in RIDAS. In addition the
option to reduce the effectiveness of the drainage after cracking beyond the drain,
was used.

Output variables

CADAM may generate different output reports where the stability drawings were
of interest. The studied output variables were the safety factors for sliding and
overturning, calculated in CADAM according to Equation (5.8) and Equation (5.9),
with and without modified uplift pressure. The program also provides information
about the extent of cracking in the concrete and rock interface.

5.4 FE-analysis

A 2D linear elastic finite element analysis, was performed to evaluate if Fishmans’


assumption that the failure mode limit turning is more likely to occur compared to
an overturning failure. One additional analysis was performed, where the rock was
provided with plastic properties to study the impact of limit turning. The analysis
was performed in BRIGADE.

To prevent that sliding occurred in the FE analysis, a high value for the friction
coefficient was defined between the rock foundation and the concrete dam body.
Thereby the monolith was forced to overturn. The destabilising loads were succes-
sively increased in the analysis to capture the load capacity of the dam.

65
CHAPTER 5. STABILITY ANALYSES

5.4.1 Studied dams

A massive monolith, Dam 2, was analysed since the overturning criterion was not
fulfilled and the difference in the safety factor for overturning and limit turning was
small, later shown in Section 6.2.1. The second studied monolith was the buttress
monolith, Dam 3, due to the relatively large difference in the safety factor for limit
turning and overturning, later shown in Section 6.2.1. The magnitudes of the applied
loads were taken from the analytical analyses in Section 5.1.1, presented in Table
5.2. The material properties are presented below.

Massive monolith, Dam 2

The concrete is of type K300 with similar properties to C25/30, used today. The
used values are presented in Table 5.5.

Table 5.5: Material properties of C25/30 (EC 2, 2011).

Density 2300 kg/m3


Young’s Modulus 31 GPa
Poisson’s ratio 0.2

The rock foundation was assumed to be granite and the material properties are
presented in Table 5.6.

Table 5.6: Material properties of granite (Björnström et al., 2006).

Density 2300 kg/m3


Young’s Modulus 60 GPa
Poisson’s ratio 0.2

The crushing resistance was set to Rcr = 20 MPa. Equation (3.13) was used to
obtain the compressive stress at which the crushing zone would form.

Buttress monolith, Dam 3

The material values for the rock foundation are assumed equal to those presented
in Table 5.6. The concrete strength is assumed to correspond to C20/25 as shown
in Table 5.7.

66
5.4. FE-ANALYSIS

Table 5.7: Material properties of C20/25 (EC 2, 2011).

Density 2300 kg/m3


Young’s Modulus 30 GPa
Poission’s ratio 0.2

For the buttress monolith, the crushing resistance was set to Rcr = 20 MPa. The
yield stress of the rock was defined as 13.6 MPa according to Equation (3.13) and
was assigned to plot a graph showing the stress and strain relationship.

5.4.2 Model definition

Model

The geometry of the monoliths and the foundations was defined by creating parts,
and modelled as solid 2D deformable bodies.

Plane stress elements with different thickness was used to enable the parts to have
different widths. For the massive monolith, the parts were assigned a thickness of
1 m. For the buttress monolith, the foundation as well as the monolith was divided
into two parts. The foundation part connected to the frontplate was defined with
the same width as the frontplate. The foundation part connected to the buttress
was defined with the same width as the buttress.

Interaction

For both monoliths the interaction between the surfaces of the dam body and the
foundation was achieved using a frictional contact definition. The normal behaviour
was described by a penalty formulation to allow for elastic slip. The penalty formula-
tion approximates “hard” interaction, hard pressure-overclosure (Dassault Systèmes,
2007). The friction coefficient was chosen to 10 to disable sliding. The reason for
this is that that otherwise the monolith would fail due to sliding.

For the buttress monolith, the surface connection between the two foundation parts
and the interaction between the frontplate and the buttress was defined with tie
constraints.

67
CHAPTER 5. STABILITY ANALYSES

Load procedure

The analysis was defined in different load steps. In the initial step, the initial inter-
actions regarding contact behaviour between the different parts and the boundary
condition was applied. The boundary condition for the foundation was set to con-
strain the bottom of the foundation. The dead weight of the materials was applied
in the second step to allow the normal force and friction force to stabilise. The
gravity loads were defined with a uniform acceleration in the vertical direction. For
the massive monolith, earth pressure was applied to the monolith in an additional
step before the design loads were applied.

To simulate the failure of the monoliths, the method of overload was used where the
design loads were applied in the first stage and in the second stage the destabilising
loads were increased until failure was reached. The horizontal water pressure and the
uplift pressure were increased by increasing the density of the water. The method
resulted in that the monolith was subjected to an increased load while the lever arm
did not change (Nordström et al., 2015).

For the massive monolith, the destabilising loads were increased by applying ampli-
tude to the loads in the same step as the design loads were applied. For the buttress
monolith, the loads were increased by applying additional destabilising loads in a
separate step after the design loads had been applied, shown in Figure 5.5. The
design loads and constrains applied to the massive monolith are seen in Figure 5.6.
The hydrostatic loads and earth pressure were defined as pressure loads and the ice
load was defined as a pressure load but with a uniform distribution.

68
5.4. FE-ANALYSIS

Figure 5.5: Design loads applied to the buttress monolith (Left) and the increased
destabilising loads (Right).

Figure 5.6: Design loads applied to the massive monolith.

Mesh

The buttress monolith was assigned a free mesh, built up out of pre-defined mesh
patterns, due to the complex geometries, see Figure 5.7. The foundations and the
massive monolith, with a simpler geometry, were assigned a structural mesh see
Figure 5.7. Plane stress and plane strain elements were chosen where each node
has two degrees of freedom. The strain components perpendicular to the element
faces are zero and the loading acts in the plane of the element (Patzák, 2014).

69
CHAPTER 5. STABILITY ANALYSES

Plane stress elements of type CPS4, were used for the buttress monolith where the
thickness of the elements is small in relation to the width. An convergence test was
performed, resulting in a mesh size of 0.5 m for the whole buttress monolith model.
The number of degrees of freedom was 9720. Plain strain elements were used for
the massive monolith where the thickness is equal to unity, by using the element
type CPE4R. A mesh size of 0.1 m was used for the whole massive monolith model,
where convergence test was performed. The number of degrees of freedom was 7040.

Figure 5.7: Mesh of the massive monolith (Left) and the buttress monolith (Right).

70
Chapter 6

Results and discussion

In this chapter the results from the analytical calculations are presented. A com-
parison of the stability calculations between RIDAS (2011), Eurocode and CADAM
is performed. The results from the parametric study for the stability calculations
according to Eurocode are presented. The results from the analytical calculations
and the FE-analysis of limit turning are presented.

6.1 Analytical analyses

6.1.1 Design approaches

A compilation of the analytical results is presented in this section.

Overturning

According to RIDAS, the safety factor s should satisfy Equation (5.1), from
Section 5.2.1 where
s > 1.5

According to Eurocode, stability against overturning should satisfy Equation (5.2),


from Section 5.2.1 where
Md,stb
≥ 1.0
Md,dst

The safety factor for overturning according to CADAM, Equation (5.9) from Section
5.3.1, should satisfy Equation (5.1) according to RIDAS.

71
CHAPTER 6. RESULTS AND DISCUSSION

The number of analysed dams that satisfied stability against overturning according
to RIDAS, Eurocode and CADAM are shown in Figure 6.1.

6
Total monoliths
5
Overturning RIDAS
4
Overturning Eurocode
3

2 Overturning CADAM

1 Overturning modified
uplift CADAM
0
Massive Buttress Spillway Pillar Spillway
+ Pillar

Figure 6.1: Total dams that satisfy the failure criterion for overturning according to
the different methods.

As seen in Figure 6.1, there are four dams that satisfied the overturning criterion
according to Eurocode but could not fulfil the criterion according to RIDAS. The
pillars represent the biggest difference in meeting the criterion according to Eurocode
compared to RIDAS.

According to the analysis performed in CADAM there is no difference in the number


of massive dams that satisfy the failure criterion.

The essential part of the results was if RIDAS and Eurocode account for stabil-
ity against overturning with comparable safety factors. However, the majority of
the dams gave the same results for the two criteria, the numerical values accord-
ing to Eurocode always overestimate the safety of the dams compared to RIDAS.
The results gave a clear indication that a parameter study was of interest for the
overturning criterion.

A compilation of how well the results from RIDAS compare with the results ac-
cording to Eurocode is presented in Figure 6.2. This was done by normalising the
safety factors obtained from the two methods. It is clearly shown that the results
from Eurocode overestimate the safety and that the safety factor most comparable
to RIDAS still differ more than 10 %.

72
6.1. ANALYTICAL ANALYSES

± 10%
RIDAS−EC
S R I D A S /S R I D A S .l i m 1.6

1.4

1.2

0.8

0.6

0.4
0.4 0.6 0.8 1 1.2 1.4 1.6 1.8
S E C/S E C.l i m

Figure 6.2: The relationship between safety factors for overturning calculated ac-
cording RIDAS and Eurocode, presented with normalised safety factors.

The obtained numerical values for the safety factors are presented in Appendix B,
Table B.1 and CADAM in Table 6.1.

Sliding

To achieve stability against sliding according to RIDAS, the friction coefficient µ


should satisfy Equation (5.3), from Section 5.2.1:

µ ≤ µmax = 0.75

Stability against sliding, according to Eurocode, is fulfilled if Equation (5.4) from


Section 5.2.1 is satisfied, i.e.
Hd − Rp;d
≤ 1.0
Rd

The safety factor for sliding is defined according to Equation (5.8) from Section
5.3.1 in CADAM. Equation (6.1) was used to enable the results from CADAM to
be comparable with the other analytical results.

1
s= (6.1)
SSF

Figure 6.3 shows the number of analysed dams, that satisfy stability against sliding
according to RIDAS, Eurocode and CADAM.

73
CHAPTER 6. RESULTS AND DISCUSSION

6
Total monoliths
5
Sliding RIDAS
4
Sliding Eurocode
3

2 Sliding CADAM

1 Sliding modified uplift


CADAM
0
Massive Buttress Spillway Pillar Spillway +
Pillar

Figure 6.3: Total dams that satisfy the failure criterion for sliding according to the
different methods.

Sliding according to CADAM resulted in no difference for unchanged uplift pressure.


There was a difference when the uplift pressure was modified after cracking initiated,
due to the increase in uplift pressure resulting in the dams more prone to fail.

From Figure 6.3 only one more dam failed according to Eurocode, this indicates
that the sliding criterion from Eurocode is comparable to the sliding criterion from
RIDAS. The results do however indicate that the safety criterion is slightly harder
to satisfy according to Eurocode.

A compilation of how well the results from RIDAS compare with the results accord-
ing to Eurocode is presented in Figure 6.4. This was done by normalising the safety
factors obtained from both RIDAS and Eurocode. The figure shows that there is
a big difference in the correlation between the safety factors from Eurocode and
RIDAS.

74
6.1. ANALYTICAL ANALYSES

± 10%
RIDAS−EC
S R I D A S /S R I D A S .l i m 1.6

1.4

1.2

0.8

0.6

0.4
0.4 0.6 0.8 1 1.2 1.4 1.6 1.8
S E C/S E C.l i m

Figure 6.4: The relationship between safety factors for sliding calculated according
RIDAS and Eurocode, presented with normalised safety factors.

Overturning and sliding

The analysed dams which satisfy stability against both overturning and sliding fail-
ure are presented in Figure 6.5. The results show that there is no greater difference
between the two methods, even though more dams satisfy the criterion according
to Eurocode for overturning. The reason for the good compliance seen in Figure
6.5 is that the failure criteria are more consistent for sliding according to the two
methods.

75
CHAPTER 6. RESULTS AND DISCUSSION

5
Total monoliths
4
RIDAS
3 Eurocode
CADAM
2
Modified uplift CADAM
1

0
Massive Buttress Spillway Pillar Spillway
+ Pillar

Figure 6.5: Total dams that satisfy both failure criteria according to the different
methods.

The results in Figure 6.5 motivate the performance of a parameter study to obtain
failure criteria applicable to dams.

CADAM

The safety factors from CADAM are presented in Table 6.1, showing the results
for the unchanged uplift pressure and the modified uplift pressure after cracking,
including the analytical safety factors from Section 6.1.1 according to RIDAS.

Table 6.1: Results from analytical calculations and CADAM.

Dam Height [m] Analytical (RIDAS) CADAM CADAM-modified uplift


Sliding Overturning Sliding Overturning Sliding Overturning

Dam 1 5 0.67 0.83 0.68 0.85 0.95 0.78


Dam 2 8 0.48 1.0 0.51 0.92 0.75 0.82
Dam 5 6 0.36 1.03 0.35 1.07 0.48 0.94
Dam 7 13 0.29 1.7 0.22 1.60 0.23 1.54

For overturning, the safety factors in Table 6.1 with modified uplift pressure after
cracking are a bit lower than the safety factors for the unchanged uplift pressure
as well as the analytical safety factors. The results from the analyses gave a lower
stability with modified uplift pressure after cracking, which was expected. Sliding

76
6.1. ANALYTICAL ANALYSES

resulted in higher safety factors for the modified uplift, i.e. increase the chance for
a sliding failure to occur. The result was that only Dam 1 did not fulfil the safety
criterion and Dam 2 was just on the limit, compared to the unchanged uplift where
all the studied dams fulfilled the safety criterion.

The safety factors calculated in CADAM were comparable to the analytical calcu-
lations according to RIDAS. Showing that CADAM is a suitable design tool for
massive dams when analysing the stability.

In addition to calculating the safety factors, the crack length in the contact surface
between the rock and concrete was calculated. The analyses were performed to
obtain additional information about the stability. The percentage of how much of
the joint that cracked is presented in Table 6.2.

Table 6.2: Resulting crack length, presented as crack percentage of the contact sur-
face from CADAM.

Dam Height [m] Crack percentage of the contact surface [%]


CADAM CADAM - modified uplift

Dam 1 5 100 100


Dam 2 8 100 100
Dam 5 6 86.5 100
Dam 7 13 7.5 10.6

For Dam 1, Dam 2 and Dam 5, that did not fulfil the failure criterion for overturning
according to RIDAS, seen in Figure 6.1, the cracking in the concrete and rock
interface was significant. In most of the cases the crack length was 100 %, seen
in Table 6.2, for both the unchanged and modified uplift. From the analyses of
the crack length only Dam 7 resulted in percentages that are reasonable for a dam
that has not failed. For the other dams the percentage of 100 % definitely serve as
indications of insufficient stability.

Since these dams have not failed in reality, there is a possibility that the dams are
subjected to lower loads than the required loads in RIDAS. The stability criterion
for overturning is known to be difficult to fulfil for low dams, confirmed by the
results from the analyses in CADAM. There could also have been faults done in the
simplifications or that the program presents an inaccurate estimation.

77
CHAPTER 6. RESULTS AND DISCUSSION

Effect of rock bolts

Only Dam 1 can according to RIDAS account for rock bolts in the stability criteria.
The steel strength accounted for in the two methods differ significantly. The differ-
ence between the two methods was determined by performing additional stability
calculations for all the dams with rock bolts, presented in Table 5.2. The results did
not show any difference for sliding, it did not affect which dams that fulfilled the
sliding criterion. A difference could be detected for the overturning criterion pre-
sented in Figure 6.6. The safety factors for RIDAS and Eurocode were calculated
according to Section 3.1.1 and Section 3.2.1, with rock bolts included in the stability
calculations. It was easily detected that more dams satisfied the overturning criteria
according to Eurocode. Figure 6.6 shows how well the safety factors calculated with
the two methods compare. As previously stated it was clear that the calculations
according to Eurocode result in much higher safety factors.

± 10%
RIDAS−EC
1.6
S R I D A S /S R I D A S .l i m

1.4

1.2

0.8

0.6

0.4
0.4 0.6 0.8 1 1.2 1.4 1.6 1.8
S E C/S E C.l i m

Figure 6.6: The relationship between safety factors for overturning with rock bolts
calculated according RIDAS and Eurocode, presented with normalised
safety factor.

It was difficult to determine if the steel strength according to RIDAS is reduced


more than necessary or if Eurocode overestimates the strength. In this report the
design criteria according to RIDAS were followed for the calculations including rock
bolts and the corresponding calculations according to Eurocode were modified to
give similar results.

78
6.1. ANALYTICAL ANALYSES

6.1.2 Parametric study

By performing a parametric study it was possible to detect which partial factor


to modify to gain results that coincided with RIDAS. The results are presented in
Figure 6.7 and Figure 6.8.

Uplift pressure
Horizontal water pressure
50%
22%

Ice pressure
28%

Figure 6.7: The most influential destabilising parameters for overturning.

The destabilising parameters with the most influence for the overturning criterion
were the vertical uplift pressure (50 %), followed by the ice pressure (28 %) and the
horizontal water pressure (22 %).

Uplift pressure
18%

Ice pressure
12%

Horizontal water pressure


70%

Figure 6.8: The most influential destabilising parameters for sliding.

The most influential destabilising load parameters for the sliding criterion were
determined. The calculations resulted in that the horizontal water load had the

79
CHAPTER 6. RESULTS AND DISCUSSION

greatest influence (70 %), followed by the vertical uplift pressure (18 %) and the ice
load (12 %).

Overturning

Due to four more dams satisfy the overturning criterion according to Eurocode
compared to RIDAS, one additional analytical calculation was performed, with new
partial factors to obtain results corresponding to RIDAS. Since the uplift is the
most influential parameter for stability against overturning, the partial factor was
modified. Due to uplift being a geotechnical action, the partial factor was changed
to the value for an unfavourable variable load, γQ = 1.4. This resulted in that one
additional dam failed, i.e. in agreement with RIDAS.

Further analytical calculations were then performed where γQ was increased to 1.5
for the uplift pressure. While the partial factors for permanent favourable loads was
changed to γG = 0.9 due to that the stabilising loads had a great influence in the
parameter study.

With these modifications of the partial factors for the overturning criterion, the
results from Eurocode were more comparable to RIDAS. The numerical values for
the calculations with the modifications are presented in Table 6.3.

80
6.1. ANALYTICAL ANALYSES

Table 6.3: Result from parametric study of the overturning criterion according to
Eurocode and RIDAS.

Dam Eurocode from Eurocode RIDAS from


number Section 6.1.1 modified Section 6.1.1

Dam 1 0.77 0.66 0.83


Dam 2 0.78 0.65 1.00
Dam 3 1.75 1.52 1.96
Dam 4 1.34 1.13 1.65
Dam 5 0.79 0.66 1.03
Dam 6 0.71 0.61 0.94
Dam 7 1.50 1.12 1.71
Dam 8 1.07 0.79 1.17
Dam 9 1.79 1.23 2.04
Dam 10 1.69 1.31 2.08
Dam 11 1.15 0.91 1.31
Dam 12 1.63 1.19 1.75
Dam 13 1.17 0.98 1.44
Dam 14 1.35 1.17 1.58
Dam 15 1.49 1.07 1.67
Dam 16 1.23 0.98 1.43
Dam 17 1.38 1.07 1.58
Dam 18 1.32 1.04 1.55

In Figure 6.9 it is possible to see how the modified partial factors result in safety
factors that correspond better to RIDAS. An improvement of the previous results
was obtained and the majority of the results do not differ more than ± 10 % from
the safety factor from RIDAS.

81
CHAPTER 6. RESULTS AND DISCUSSION

± 10%
RIDAS−EC modified
1.6
S R I D A S /S R I D A S .l i m RIDAS−EC
1.4

1.2

0.8

0.6

0.4
0.4 0.6 0.8 1 1.2 1.4 1.6 1.8
S E C/S E C.l i m

Figure 6.9: The relationship between safety factors for overturning calculated ac-
cording RIDAS, Eurocode and Eurocode with modified partial factors,
presented with normalised safety factors.

Sliding

RIDAS and Eurocode were quite consistent for the sliding criterion, but since some
difference between the safety factors was obtained, an attempt to achieve results
equivalent to RIDAS was made.

The horizontal water pressure was the most influential parameter. The partial factor
for the horizontal water pressure was modified to γQ = 1.0, which did not affect the
results for the stability criterion. If γQ for the horizontal water pressure was further
decreased or further modifications of the partial factors for the loads were made,
it resulted in dams already in agreement with RIDAS to change, i.e. an unwanted
result. The results gave that the partial factor for the horizontal water load could
only be modified to γQ = 1.0.

As shown in Figure 6.10 the modified partial factors give a slightly better corre-
spondence to RIDAS. The result is however, still not satisfying and therefore more
studies need to be performed to find partial factors that would result in a acceptable
correspondence to RIDAS.

82
6.1. ANALYTICAL ANALYSES

± 10%
RIDAS−EC modified
1.6
S R I D A S /S R I D A S .l i m RIDAS−EC
1.4

1.2

0.8

0.6

0.4
0.4 0.6 0.8 1 1.2 1.4 1.6 1.8
S E C/S E C.l i m

Figure 6.10: The relationship between safety factors for sliding calculated accord-
ing RIDAS, Eurocode and Eurocode with modified partial factors, pre-
sented with normalised safety factors.

Rock bolts

Including rock bolts gave no difference in the result regarding the sliding criterion
and therefore only the partial factors for overturning are discussed in this section.

The modified partial factors used for the results presented above, γQ = 1.5 and
γG = 0.9, were used to determine if the results from Eurocode and RIDAS would
be more similar. The outcome was not as desired and an additional calculation was
performed where, in addition to the previous adjustments, the partial factor γs was
changed to 1.35. The adjustment of γs affected the strength of the rock bolts. The
modifications resulted in more comparable safety factors to RIDAS, shown in Figure
6.11.

83
CHAPTER 6. RESULTS AND DISCUSSION

± 10%
RIDAS−EC modified
1.6
S R I D A S /S R I D A S .l i m RIDAS−EC
1.4

1.2

0.8

0.6

0.4
0.4 0.6 0.8 1 1.2 1.4 1.6 1.8
S E C/S E C.l i m

Figure 6.11: The relationship between safety factors for overturning with rock bolts
calculated according RIDAS, Eurocode and Eurocode with modified
partial factors, presented with normalised safety factors.

6.1.3 Previously studied monoliths

From the report by Fouhy and Rios Bayona (2014), the results was interpreted as a
big scatter, where the performed comparison with the acceptable values according to
Eurocode was limited. From the probabilistic analyses in their report, no conclusions
could be drawn about if the failure criterion was fulfilled or not. Fouhy and Rios
Bayona (2014) showed that their probabilistic method gave the same β-value for a
dam that fulfil the deterministic failure criterion as for a dam that fails to fulfil the
criterion. By observation and comparison to Eurocode, the probabilistic method
did not result in comparable safety limits. The use of partial factors in this report
shows better correspondence with RIDAS.

6.2 Analyses of limit turning

6.2.1 Analytical analysis

The safety factors from the analytical calculations for limit turning and overturning
according to RIDAS, calculated with Rcr = 20 MPa, are presented in Table 6.4. The
table also includes the difference between the safety factors for the two criteria.

84
6.2. ANALYSES OF LIMIT TURNING

Table 6.4: Results from analytical calculations of stability against limit turning and
overturning.

Dam Height [m] Limit turning Overturning from Difference[%]


Section 6.1.1

Dam 1 5 0.829 0.831 -0.2


Dam 2 8 1.000 1.004 -0.4
Dam 3 40 1.738 1.955 -11.1
Dam 4 12 1.619 1.653 -2.1
Dam 5 6 1.029 1.033 -0.4
Dam 6 6 0.931 0.942 -1.2
Dam 7 13 1.704 1.707 -0.2
Dam 8 7 1.172 1.175 -0.3
Dam 9 5 2.141 2.043 +4.8
Dam 10 13 1.971 2.082 -5.3
Dam 11 7 1.307 1.314 -0.5
Dam 12 4 1.749 1.751 -0.1
Dam 13 7 1.435 1.440 -0.3
Dam 14 20 1.496 1.576 -5.1
Dam 15 16 1.667 1.667 0
Dam 16 19 1.403 1.429 -1.8
Dam 17 19 1.565 1.579 -0.9
Dam 18 18 1.526 1.549 -1.5

The dam with the greatest difference between the safety factors was Dam 3 where
the difference between the two criteria was 11 %. For the majority of the dams the
difference between the safety factors was insignificant. For all dams under 10 m the
impact of limit turning on the safety factor was negligible. For Dam 9 the safety
factor for limit turning increased, showing that limit turning not always result in
lower safety factors. This is presumed to be due to that the sum of the stabilising
forces are either increased more or decreased less than the sum of destabilising forces.
A high contributed stabilising force from the rock might be an other reason why the
safety factor increased. From the assumption that limit turning should fulfil the
same criterion as overturning, only one dam, Dam 14, failed due to limit turning
while fulfilling the criteria for overturning. By the comparison of Dam 5 and Dam 6
as well as for Dam 4 and Dam 7, the difference between the safety factors is greater
for buttress dams compared to massive dams of the same height, highlighted in the
results presented in Table 6.4.

The impact of the crushing resistance on the safety factor for limit turning can be
seen in Figure 6.12 and Figure 6.13. For Dam 2 in Figure 6.12, the relationship

85
CHAPTER 6. RESULTS AND DISCUSSION

between the safety factor for limit turning and the crushing resistance show that
the limit turning criterion is of importance for rock with poor quality, typically
a crushing resistance less than 5 MPa. However, for Dam 3, in Figure 6.13, the
safety factor for limit turning does not approach the value of the safety factor for
overturning in the same way as for Dam 2, therefore in this case the rock quality
becomes more important. The possibility of a limit turning failure may be worth to
consider for high dams.

Limit turning
Overturning
1.2

1.1
Safety factor [-]

0.9

0.8

0.7

0.6

0 5 10 15 20 25 30 35 40 45
Crushing resistance of rock [MPa]

Figure 6.12: The relationship between the safety factor for limit turning and the
crushing resistance for Dam 2.

3.5
Limit turning
Overturning
3
Safety factor [-]

2.5

1.5

0 5 10 15 20 25 30 35 40 45
Crushing resistance of rock [MPa]

Figure 6.13: The relationship between the safety factor for limit turning and the
crushing resistance for Dam 3

A significant influence of the crushing resistance can clearly be seen which also show
the need for geotechnical investigations in order for limit turning design criteria to be

86
6.2. ANALYSES OF LIMIT TURNING

useful. In dam design today approximated material values are often used. The use
of approximated values for the crushing resistance may lead to inadequate design.

6.2.2 FE-analysis

For both the massive and the buttress monolith, the ultimate loading for the desta-
bilising loads was defined as two times the design load. The ultimate loading was
chosen to ensure the model to reach failure.

Dividing the load at which the monolith goes to failure with the design load made it
possible to extract the safety factor to examine if it corresponded with the analytical
calculations from RIDAS.

By analysing if the compressive stresses that corresponds to the crushing resistance


were reached, it was possible to determine if limit turning failure would occur before
overturning failure. Based on Equation (3.13) a compressive strength of 13.6 MPa
(corresponding to the crushing resistance 20 MPa) was used to detect failure.

For Dam 3, the compressive stresses corresponded to the crushing resistance when
the ratio of total loads/design loads = 1.68. The parts of the rock underneath the
toe started to plasticise, i.e. formation of the crushing zone, indicating that limit
turning would occur before overturning. In Figure 6.14, when the ratio of total
loads/design loads = 1.89, a greater part of the rock beneath the toe had reached
stresses equal to the crushing resistance and the crushing zone was easily detected.
This gives that the analytical safety factor for limit turning 1.74 from Section 6.2.1
appears reasonable.

By extracting the stresses distributed over one element in the rock crushing zone,
plastic deformation of the rock could also be seen, shown in the stress-strain curve in
Figure 6.15. Showing that the stresses in the rock lead to the formation of a crushing
zone before the overturning failure occurred, at a safety factor corresponding to 1.94,
i.e. similar to the analytical safety factor for overturning.

87
CHAPTER 6. RESULTS AND DISCUSSION

Figure 6.14: Compressive stress in Dam 3.

16

14

12

10
Stress [MPa]

0
0 0.0005 0.001 0.0015 0.002 0.0025 0.003 0.0035
Strain [‐]

Figure 6.15: Stress-strain relationship for Dam 3.

88
6.2. ANALYSES OF LIMIT TURNING

The analysis of Dam 2 did not display any signs of limit turning failure, which was
not surprising considering the results from Section 6.2.1 where Dam 2 only had
0.4 % lower safety factor for limit turning compared to overturning. The model
failed when the safety factor was 1.09, which corresponds to the analytical safety
factor for overturning. The rock did not reach the compressive stresses required for
the crushing zone to develop, as seen in Figure 6.16. This indicates that the dam is
more prone to fail according to overturning compared to limit turning.

Figure 6.16: Compressive stresses in Dam 2.

89
Chapter 7

Conclusions

The objective of this report was to examine if the design criteria for concrete dams
used today are enough or if modifications are needed. The other objective was to
analyse if Eurocode was comparable to RIDAS in dam design.

7.1 Failure modes of concrete dams

The majority of the studied massive dams, built before 1940, stood for many years
before failure. The dams were designed under different circumstances and for the
dams that did not fail during the first five years, there must be a different expla-
nation. The explanation could be the unknown parameters such as climate change
affecting the loads acting on the dam. The difficulties in predicting amount of rain
and increasing natural disasters is a big problem with design today, and even more
a problem a few decades ago. Today there are possibilities of computer simulations
for probabilistic variations of different loads affecting the dam, resulting in better
designed structures. However, as seen in Section 4.4, this has not prevented recent
dam failures, which were in some cases caused by faults in the design. Along with
computer simulations, the development of concrete recipes has been ongoing. Today,
different measures help improve the concrete dam design and achieve better stability
such as different types of cement, e.g. Portland cement and improved quality of the
reinforcement, e.g. profiled reinforcement bars.

To continue to learn from the previous dam failures helps the development of dam
structures along with predictions about the future and improved concrete recipes.
This is also true for buttress dams where the majority of the dams failed during
the first five years. Since buttress dams are relatively new types of structures, they

91
CHAPTER 7. CONCLUSIONS

are limited to lessons from only a few previous failures. Additionally the building
techniques and material knowledge have improved for these structures, preventing
failure.

From the analysed failures, one result is that the dam foundation has a crucial part
in a safe dam design. Failure in the dam foundation were often caused by the lack
of information about the foundation materials. Establishing tougher regulations
concerning implementation of actual ground investigations could be a solution. In
Sweden today, many of the calculations are based on standard values of the rock
foundation material instead of actual foundation properties, which could lead to an
overestimated safety.

The failure is often initiated by one specific failure mode resulting in a combination
of failures, which shows that pure failure modes seldom occur. There were failure
modes not covered, indicating that the design criteria are not thorough. The results
in Section 4.4, indicate that the failures where more information was collected,
were either caused by a poor design or poor workmanship shown by the contractor
where short-cuts and lack of understanding are displayed. One method to increase
safe construction design, already established in Sweden, is tougher regulations and
placing responsibility on the dam owners, which makes them prone to design safe
constructions.

The known failures were studied to determine whether the currently used design
criteria would suffice. In conclusion, it was determined from the information regard-
ing the collective failures, that the design criteria is extensive enough. However, the
challenge lies in ensuring that the construction of the dam is correctly performed to
fulfil today’s criteria.

The results are summarised as; the current design criteria are sufficient and safer
dam designs are achieved if the following conditions are considered:

• A thorough geotechnical investigation is performed.


• Engineers with enough knowledge are involved in designing the dam.
• Increased focus on technical issues rather than economic aspects.

7.2 Analytical calculations

Limit turning, which is not a design criterion in Sweden today, produced lower
safety factors for the majority of the dams, indicating that it might need to be

92
7.3. DESIGN GUIDELINES

considered. The simulations in the software BRIGADE showed that limit turning
was important for the highest of the two dams. Increasing height of the dam, result
in greater influence of limit turning. A difference in the effect of limit turning on
buttress dams compared to massive dams were detected.

The results indicated that the quality of the rock had a great influence on the
safety factor. The difficulties with limit turning appeared to be that geotechnical
investigations of the rock quality are essential for the calculations to be sufficient
as a design criterion. Approximated values for the crushing resistance could lead to
inadequate design. Limit turning also requires more extensive calculations making
it less convenient to use. Nevertheless, for buttress dams and high dams, especially
high buttress dams, limit turning is of interest.

7.3 Design guidelines

Comparing the methods of partial factors and probabilistic analyses, the method of
using partial factors appears to be easier to adapt to the requirements defined by
RIDAS.

From the comparison of RIDAS and Eurocode it was clear that modifications were
needed for Eurocode to apply to concrete dams, especially for overturning where to
many dams passed the failure criterion. By modifications of the mentioned partial
factors, Eurocode appeared to give similar safety factors as RIDAS. The sliding cri-
terion appeared to be more difficult to adjust to RIDAS, even though the difference
between the two methods was not that significant. The partial factors are presented
in Table 7.1, where the modified partial factors are highlighted.

93
CHAPTER 7. CONCLUSIONS

Table 7.1: The resulting modified partial factors for the failure criteria according to Eu-
rocode

Load EC EC-modified overturning EC-modified sliding

Dead weight 1.0 0.9 1.0


Horizontal water pressure upstream side 1.1 1.1 1.0
Horizontal water pressure downstream side 1.0 0.9 1.0
Vertical water pressure 1.0 0.9 1.0
Vertical uplift pressure 1.1 1.5 1.1
Horizontal uplift pressure 1.1/1.01 1.1/0.9 1.1/1.0
Ice load 1.5 1.5 1.5
Rock anchors2 1.1 1.1 1.1
Earth pressure 1.1/1.0 1.1/0.9 1.1/1.0

1 unfavourable/favourable
2 the partial factor for the reinforcement was modified from 1.15 to 1.35

The stability calculations will be more complicated in Eurocode compared to RIDAS,


since the direction of the force need to be considered as well as if the load is a
favourable or unfavourable. An additional disadvantage of using partial factors is
that different partial factors would have to be used for sliding and overturning. The
engineer would be responsible for interpreting the different loads affecting the specific
dam analysed. Resulting in more factors to consider, leaving room for mistakes.

7.4 Future studies

Further studies needs to be performed to justify the statement that limit turning is
of interest for buttress dams and high dams. For the method of using partial factors
to work, further studies need to be performed, including if it is possible to account
for the strength of the rock bolts according to Eurocode and which partial factors
that could be reasonable to utilise.

94
Bibliography

Adolfi, E., Eriksson, J., 2013. Islastens inverkan på brottsannolikheten för glidning
och stjälpning av betongdammar. Tech. rep., KTH Royal Institute of Technology.

Ali, M. H., March 2012. Comparison of Design and Analysis of Concrete Gravity
Dam. Natural Resources 03, 18–28.

Anderson, C., Mohorovic, C., Mogck, L., Cohen, B., Scott, G., 1998. Concrete Dams
Case Histories of Failures and Nonfailures with Back Calculations. Tech. rep.,
United States Department of the Interior, Bureau of Reclamation, Dam Safety
Office.

Andersson, C., 2012. SVC-dagarna 2012. SVC - Svenskt vattenkraftcentrum, 2012-


09-25.

Andersson, C., 2014. Summering av workshop om Eurokoder med tillämpning på


betongdammar. Workshop 25/11 2014 om Eurokoder med tillämpning på betong-
dammar. Energiforsk(Elforsk), 2014-11-25.

Bergh, H., 2014. Hydraulic Engineering. KTH Civil and Architectural Engineering,
Sweden.

Björnström, J., Ekström, T., Hassanzadeh, M., 2006. Spruckna betongdammar –


Översikt och beräkningsmetoder. Elforsk report 06:29, Eneriforsk(Elforsk).

Bond, H., 2014. Dimensionering och kontroll av betongdammar enligt RIDAS 2012.
Workshop 25/11 2014 om Eurokoder med tillämpning på betongdammar. Energi-
forsk(Elforsk), 2014-11-25.

Dassault Systèmes, 2007. Abaqus Analysis User’s Manual. Dassault Systèmes Group.

DOI, 2009. 19. Risk Analysis for Concrete Buttress Dams. Bureau of reclamination,
United States Departement of the Interior.

DOI, 2012. 20. Risk Analysis for Concrete Gravity Structures. Bureau of reclamina-
tion, United States Departement of the Interior.

95
BIBLIOGRAPHY

Douglas, K. J., 2002. The Shear Strength of Rock Masses. Ph.D. thesis, The Uni-
versity of New South Wales.

EC 0, 2002. Eurocode - Basis of structural design. EN 1990.

EC 1, 2013. Eurocode 1: Actions on strucutres - Part 1-1: General actions - Densi-


ties, self-weight, imposed loads for buildings. EN 1991-1-1.

EC 2, 2011. Eurocode 2: Design of concrete structures - Part 1-1: General rules and
rules for buildings. EN 1992-1-1.

EC 7, 2011. Eurocode 7: Geotechnical design - Part 1: General rules. EN 1997-1.

EKS 9, 2013. Boverkets föreskrifter om ändring i verkets föreskrifter och allmänna


råd (2011:10) om tillämpning av europeiska konstruktionsstandarder (eurokoder).
BFS 2013:10, Boverket.

Ferc Engineering Guidelines, 2002. Chapter III Gravity Dams. Ferc - Federal Energy
Regulatory Commission.

Ferc Engineering Guidelines, 2014. Risk-Informed Decision Making, Chapter R5.


Ferc - Federal Energy Regulatory Commission.

Fishman, Y. A., 2007. Features of shear failure of brittle materials and concrete
structures on rock foundations. International Journal of Rock Mechanics and Min-
ing Sciences 45 (6), 976–992.

Fishman, Y. A., 2009. Stability of concrete retaining structures and their inter-
face with rock foundations. International Journal of Rock Mechanics and Mining
Sciences 46 (6), 957–966.

Fouhy, D., Rios Bayona, F., 2014. Reliability-Based Analysis of Concrete Dams.
Master Thesis, KTH Royal Institute of Technology.

Geograph, 2010. The breach in the dam wall, Liyn Eigiau reservoir. http://www.
geograph.org.uk/photo/1963073, retrieved 2015-04-23.

Gustafsson, A., Johansson, F., Rytters, K., Stille, H. k., 2008. Betongdammars
Glidstabilitet - Förslag på nya riktlinjer. Elforsk report 8:59, Energiforsk(Elforsk).

HSS, 2013. Shih-Kang Dam Project in Taiwan. http://www.hss.dk/casestudies/


taiwan.aspx, HSS Engineering warning system solutions, retrieved 2015-04-23.

ICOLD, 1995. Dam failures statistical analyses. Bullentin 99. Commission Interna-
tionale des Grands Barrages.

96
BIBLIOGRAPHY

Isander, A. (Ed.), 2013. Dams: Accidents and Incidents What Can We Learn. Eu-
ropean Club of ICOLD, International Comission on Large Dams.

J Andrew, C., Tedd, P., Warren, A., 2011. Lessons from historical dam incidents.

Jarrett, R. D., 1986. Hydrology, Geomorphology, and Dam-Break Modeling of the


July 15 , 1982 Lawn Lake Dam and Cascade Lake Dam Failures. Tech. rep., United
States Department of the Interior.

Johansson, F., 2005. Stability Analyses of Large Structures Founded on Rock. Tech.
rep., Department of Civil and Architectural Engineering, KTH Royal Institute of
Technology.

Kleivan, E., Kummeneje, G., Lyngra, A., 1994. Concrete in Hydropower structures.
Norwegian Institute of Technology, Division of Hydraulic Engineering.

KPLU 88.5, 2011. Elwha River dam removal historic, but not explosive. http:
//www.kplu.org/post/elwha-river-dam-removal-historic-not-explosive,
news for Seattle and the Northwest, retrieved 2015-04-23.

Krounis, A., 2013. Uncertainty in Sliding Stability Analyses of Existing Concrete


Gravity Dams with Bonded Concrete-Rock Interfaces. Ph.D. thesis, Division of
Soil and Rock Mechanics, Royal Institute of Technology.

Kung, C.-S., Ni, W.-P., Chiang, Y.-J., 2001. Damage and Rehabilitation Work of
Shih-Kang Dam. Tech. rep., Sinotech Engineering Consultants, LTD.

Leclerc, M., Léger, P., Tinawi, R., 2001. Cadam User’s Manual. Montréal, version
1.4.3.

Leclerc, M., Léger, P., Tinawi, R., 2003. Computer aided stability analysis of gravity
dams - CADAM. Advances in Engineering Software 34, 403–420.

Ljungkrantz, C., Möller, G., Petersons, N., 1994. Betonghandbok - Material, 2nd
Edition. AB Svensk Byggtjänst och Cementa AB.

Los Angeles Times, 2013. St. Fran Dam collapse. http://framework.latimes.com/


2013/03/12/st-francis-dam-collapse, retrieved 2015-05-02.

Luino, F., Tosatti, G., Bonaria, V., 2014. Dam failures in the 20th century: nearly
1,000 avoidable victims in Italy alone. Journal of Environmental Science and En-
gineering 3, 19–31.

97
BIBLIOGRAPHY

Malm, R. (Ed.), 2015. Hydropower- Technology, Economy, Sustainability. KTH Con-


crete Strucutures / SWECO, course material.

MATLAB, 2013. Matlab r2013a. The MathWorks, Inc.

Molare.net, 2010. The disaster of molare. http://www.molare.net/disaster/the_


disaster_4.html, retrieved 2015-04-23.

Nilsen, B., Thidemann, A., 1993. Rock Engineering. Norwegian Institute of Tech-
nology, Division of Hydraulic Engineering.

Norconsult, 2012. Lima kraftverk. https://www.norconsult.se/referenser/


2014/lima-kraftverk/, retrieved 2015-03-03.

Nordström, E., Malm, R., Johansson, F., Ligier, P.-L., Lier, Ø., 2015. Betong-
dammars brottförlopp - Litteraturstudie och utvecklingspotential. Elforsk report
2015:122, Energiforsk(Elforsk).

Oakes, R., 2001. Historical Background on the Elwha River Dams. American Field
Guide, Lakeridge High School.

Patzák, B., 2014. OOFEM Element Library Manual. http://www.oofem.org/


resources/doc/elementlibmanual/html/elementlibmanual.html, retrieved
2015-04-17.

PTI, 2000. What Is Post-Tensioning. Post Tensioning Institute.

Reegan, P., 2015. Dam Incident Data. Ferc - Federal Energy Regulatory Commission,
USA.

RIDAS, 2011. Kraftföretagens riktlinjer för dammsäkerhet, RIDAS. Avsnitt 7.3 Be-
tongdammar Tillämpningsvägledning. Svensk Energi - Swedenenergy AB.

Risk Assessment International, 2013. Dam failure news. http://www.


risk-assessment.at/news_en.html, retrieved 2015-04-23.

Rogers, D. J., 1995. A man, a dam and a disaster: Mulholland and the St. Fancis
dam. In: Doyance B. Nunis, J. (Ed.), The St. Francis Dam Disaster. Historical
Society of Southern California, pp. 1–109.

Sanscot Technology, 2015. Brigade Plus 5.2. Sanscot Technology AB.

SFS 2014:114. Lag om ändring i miljöbalken. Stockholm: Miljödepartementet.

98
BIBLIOGRAPHY

Shaffner, P., Scott, G., 2013. The knowledge of precedents and their influence on
professional development.

Svenska Kraftnät, 2014. Regelverk och riktlinjer. http://www.svk.se/


aktorsportalen/dammsakerhet/regelverk-och-riktlinjer/, retrieved
2015-03-03.

The Engineering News Publishing company, 1910. Partial Failure of a Concrete Dam
at Austin Pa. Engineering News 63, 321–323.

The Engineering News Publishing company, 1911. The Partial Failure of a Concrete
Dam at Austin, Pa., on Jan. 23, 1910. Enineering news v.66 (14), 417–424.

Trafikverket, 2011. TK Geo 11 Trafikverkets tekniska krav för geokonstruktioner.


Trafikverket.

Vattenkraft.info, 2009. Info om Svensk vattenkraft. http://vattenkraft.info/


teori/bilder/ratan.jpg, retrieved 2015-03-03.

Veale, B., Davison, I., 2011. Estimation of gravity dam breach geometry Review of
current practice.

Water and power, 2015. St. francis dam disaster. http://waterandpower.org/


museum/St.%20Francis%20Dam%20Disaster.html, retrieved 2015-04-23.

Westberg, M., 2010. Reliability-based assessment of concrete dam stability. Doctoral


Thesis TVBK-1039, Division of Structural Engineering, Lund University.

Westberg, M., 2014. Stabilitetsanalys av betongdammar - läget för pågåenge utveck-


ling av sannolikhetsbaserad metod. Worshop 25/11 2014 om Eurokoder med
tillämpning på betongdammar. Energiforsk(Elforsk), 2014-11-25.

Westberg, M., Hassanzadeh, M., 2007. Stabilitetsanalys av betongdammar. Elforsk


report 07:11, Energiforsk(Elforsk), seminar 2006-11-27.

Wilson, D., 1963. Baldwin Hills Reservoir disaster. http://jpg3.lapl.org/


pics21/00060040.jpg, los Angeles Public Library Images, retrieved 2015-05-20.

99
Appendix A

Compiled failures

101
APPENDIX A. COMPILED FAILURES

Dam name Co- Type Height Year Year Fund Geology Fail. Failure Fail. Failure due to Failure Description
untry lowest com. failure mat. type mode mode overtopping
found foundatio dam
n
P SC S
Bayless USA M 17 1909 1910 Rock Sandstone horizontal Ff x Overtopping due See Section 4.3.1
(1911) layers with shale and to unknown cause
clay between
Camara Brazil M 50 2002 2004 Rock Plane of micaceous Ff x Not at highest See Section 4.3.1
silty clay water level
Eigiau GB M 10 1911 1925 Clay Hard blue clay Ff/F x Not at highest See Section 4.3.1
containing boulders of m water level
granite overlain by a
layer of peat
Elwha USA M 51 1912 1912 Soil/ Fluvioglacial and Ff x Filled See Section 4.3.1
Rock conglomerate
High Falls USA M 9 1910 1999 Rock No information Fm ST Overtopping due Overtopping led to breach
(NY) to unknown cause of 23 meter long portion
of concrete crest cap, left
half the spillway. Repairs
completed.
Marquette USA M 10 1924 2003 Rock No information Fa Overtopping due Overtopping and failure
no 3 to unknown cause of abutment, due to failure
of upstream dam.
Shih-Kang Taiw M 22 1977 1999 Rock Top deposition layer: Ffb EQ No information See Section 4.3.1
dam an unconsolidated gravel,
sands, silts and clay.
On Soft bedrock:
slate-gray, sandy-
shale and silty-
sandstones
St Francis USA M 62 1926 1928 Rock Conglomerate and Ff x x Gradual during See Section 4.3.1
schist first fill
Torrejon- Spain M 62 1967 1965 No No information Fa/ SH Flood during Shear sliding within the
Tajo info Fm construction dam. Failure cause was
traced to organic material
present in the aggregate
and filling of the dam by a
flood during construction
before the concrete had
fully hardened.
Upriver USA M 11,5 1937 1986 Soil No information Fa x Overtopping due Washout of the abutment
dam to unknown cause and the power canal
embankments due to
overtopping. Not a
complete failure and
reparations of the dam
were possible.
Warrens- USA M 8 1909 1976 No No information Fa T/C No information Breach of north abutment.
burg info Reconstructed in 1998.
Xuriguera Spain M 42 1902 1944 Rock No information Ff x No information Failed by foundation
sliding, shear strength and
poor design.
Zerbino Italy M 16 1925 1935 Rock Schist and hornfeld Faf x x Overtopping due See Section 4.3.1
to unknown cause
Ashley USA B 18 1908 1909 Soil Fluvioglacial Ff x Just spilling when Piping failure in fine sand
pipe failed with little clay and gravel,
6m deep below cut-off.
Cascade USA B 5 1908 1982 Soil Glacial terminal- Ffa x Overtopping due The dam was overtopped
lake dam moraine sediments to unknown cause before tipping over and
failing. The cause of
failure was the hydrostatic
water pressure on the dam
and erosion of the
abutments. Stored water
was released rapidly due
to short time of breach
development and the
width of the breach was
large.
Komoro Japan B 16 1927 1928 Rock Tuff Ff x x No suggestions of Failure due to softening of
high water level volcanic ash in
foundation. Unclear
cause, either piping,
sliding or both.
Morris USA B 58 1941 1986 Rock Shale Ff x Releases kept See Section 4.3.1
Sheppard within channel
capacity
Overhol- USA B 17 1920 1923 Rock No information Ffa x Overtopping due Overtopping leading to
ser to unknown cause scour of abutment.
Stoney USA B 21 1913 1914 Soil No information Ff x Not clear if failed Piping in foundation
creek at top level followed by settling of
dam, cracking and
collapse of dam.
 

102
Appendix B

Results analytical analyses

Table B.1: Safety factors from analytical calculations.

RIDAS Eurocode Eurocode mod.


Dam Type Sliding Overturning Sliding Overturning Sliding Overturning

Dam 1 M 0.6666 0.8313 0.9473 0.7689 0.8820 0.6693


Dam 2 M 0.4780 1.0043 0.9464 0.7774 0.8887 0.6471
Dam 3 B 0.7757 1.9554 1.1092 1.7549 1.0121 1.5157
Dam 4 B 0.5648 1.6531 0.9773 1.3381 0.8927 1.1301
Dam 5 M 0.3619 1.0329 0.7291 0.7943 0.6871 0.65577
Dam 6 B 1.3266 0.9423 2.2461 0.7095 2.1750 0.6133
Dam 7 M 0.2875 1.7074 0.4863 1.5023 0.4387 1.1163
Dam 8 S 0.8709 1.1749 1.3687 1.0666 1.2402 0.7944
Dam 9 S 0.2554 2.0430 0.4904 1.7917 0.4229 1.2301
Dam 10 P 0.0407 2.0823 0.1122 1.6909 0.0765 1.3126
Dam 11 P 0.5708 1.3136 0.8598 1.1477 0.7974 0.9135
Dam 12 S 0.4829 1.7511 0.6297 1.6297 0.6373 1.1884
Dam 13 P 0.5911 1.4398 0.9448 1.1677 0.8996 0.9760
Dam 14 B 0.8161 1.5763 1.1960 1.3468 1.1024 1.1654
Dam 15 P+S 0.6045 1.6667 0.9647 1.4866 0.8971 1.0663
Dam 16 P 0.9583 1.4290 1.4689 1.2314 1.3700 0.9785
Dam 17 P 0.6262 1.5790 0.9786 1.3772 0.9176 1.0684
Dam 18 P 0.7225 1.5493 1.1330 1.3227 1.0657 1.0421

103
APPENDIX B. RESULTS ANALYTICAL ANALYSES

Table B.2: Safety factors from analytical calculations including rock bolts

RIDAS Eurocode Eurocode mod.


Dam Type Sliding Overturning Sliding Overturning Overturning

Dam 1 M 0.6666 0.8313 0.9473 0.7689 0.6222


Dam 4 B 0.5139 1.7919 0.8002 1.5901 1.3115
Dam 6 B 1.0075 1.2350 1.2858 1.1966 0.9719
Dam 8 S 0.6873 1.3374 0.8117 1.3743 0.9896
Dam 15 P+S 0.5763 1.7377 0.8608 1.6269 1.1520

104
Appendix C

Output values for dams

Table C.1: Dam 1, massive dam, height 5 m.

Gravity Water H Ice Uplift V Soil H Soil V Anch x Anch y

Loads [MN] 2.63 1.09 2.00 0.90 0.62 0.90 0.31 0.61
RIDAS
Lever arm [m] 2.57 1.57 4.38 2.59 1.47 0.96 0.00 3.41
Loads [MN] 2.74 1.11 2.00 0.92 0.60 0.86 0.71 1.41
Eurocode
Lever arm [m] 2.57 1.57 4.38 2.59 1.47 0.96 0.00 3.41

Table C.2: Dam 2, massive dam, height 8 m.

Gravity Water H Ice Uplift V Soil H Soil V

Loads [MN] 5.20 2.48 2.00 1.90 1.70 2.50


RIDAS
Lever arm [m] 3.62 2.37 6.78 3.63 2.27 1.50
Loads [MN] 5.43 2.53 2.00 1.94 1.64 2.39
Eurocode
Lever arm [m] 3.62 2.37 6.78 3.63 2.27 1.50

Table C.3: Dam 3, buttress dam, height 40 m.

Gravity Water H Ice Uplift V Water V

Loads [MN] 42.12 55.48 1.60 2.95 34.41


RIDAS
Lever arm [m] 17.56 12.53 37.27 34.33 27.16
Loads [MN] 49.95 56.55 1.60 3.01 35.08
Eurocode
Lever arm [m] 17.56 12.53 37.27 34.33 27.16

105
APPENDIX C. OUTPUT VALUES FOR DAMS

Table C.4: Dam 4, buttress dam, height 12 m.

Gravity Water H Ice Uplift V Soil H Soil V Anch x Anch y Water V

Loads [MN] 5.75 5.10 1.60 0.78 2.31 1.25 0.12 0.54 1.55
RIDAS
Lever arm [m] 8.38 3.80 11.07 12.18 3.33 2.38 0.00 12.04 11.68
Loads [MN] 6.00 5.20 1.60 0.79 2.11 1.29 0.27 1.23 1.58
Eurocode
Lever arm [m] 8.38 3.8 11.07 12.18 3.33 2.38 0.00 12.04 11.68

Table C.5: Dam 5, massive dam, height 6 m.

Gravity Water H Ice Uplift V Soil H Soil V Water V

Loads [MN] 0.25 0.11 0.10 0.08 0.05 0.08 0.01


RIDAS
Lever arm [m] 2.11 0.87 3.80 2.34 1.33 0.67 3.35
Loads [MN] 0.26 0.11 0.10 0.08 0.05 0.08 0.01
Eurocode
Lever arm [m] 2.11 0.87 3.80 2.34 1.33 0.67 3.35

Table C.6: Dam 6, buttress dam, height 6 m.

Gravity Water H Ice Uplift V Anch x Anch y Water V

Loads [MN] 2.07 1.14 1.60 0.26 0.12 0.54 0.26


RIDAS
Lever arm [m] 4.59 1.80 5.07 6.23 0.00 6.45 6.24
Loads [MN] 2.16 1.17 1.60 0.27 0.27 1.23 0.27
Eurocode
Lever arm [m] 4.59 1.80 5.07 6.23 0.00 6.45 6.24

Table C.7: Dam 7, massive dam, height 13 m.

Gravity Water H Ice Uplift H Uplift V Soil H

Loads [MN] 1.65 0.61 0.10 0.06 0.85 0.11


RIDAS
Lever arm [m] 7.09 4.57 11.70 0.60 7.19 6.10
Loads [MN] 1.73 0.61 0.10 0.06 0.85 0.11
Eurocode
Lever arm [m] 7.090 4.570 11.70 0.60 7.19 6.10

Soil H3 Soil V Soil V2 Water V Tailw H Tailw V

Loads [MN] 0.15 0.65 0.08 0.19 0.08 0.06


RIDAS
Lever arm [m] 2.05 2.73 0.96 1.03 0.00 3.64
Loads [MN] 0.15 0.62 0.06 0.19 0.08 0.06
Eurocode
Lever arm [m] 2.05 2.73 0.96 11.35 1.37 0.98

106
Table C.8: Dam 8, spillway, height 7 m.

Gravity Water H Water H2 Ice Uplift H Uplift V Anch x

Loads [MN] 0.76 0.41 0.02 0.03 0.04 0.42 0.02


RIDAS
Lever arm [m] 4.50 2.17 6.20 8.70 0.40 4.35 0.53
Loads [MN] 0.8 0.41 0.02 0.03 0.04 0.42 0.05
Eurocode
Lever arm [m] 4.5 2.17 6.20 8.70 0.40 4.35 0.53

Anch y Water V Tailw H Tailw V

Loads [MN] 0.08 0.02 0.02 0.02


RIDAS
Lever arm [m] 6.35 10.85 0.00 9.88
Loads [MN] 0.19 0.02 0.02 0.02
Eurocode
Lever arm [m] 6.35 6.85 0.73 0.49

Table C.9: Dam 9, spillway, height 5 m.

Gravity Water H Uplift V Water V Water V2 Tailw H Tailw V

Loads [MN] 0.74 0.34 0.61 0.08 0.11 0.01 0.02


RIDAS
Lever arm [m] 6.55 2.09 6.75 10.42 8.32 0.57 1.13
Loads [MN] 0.78 0.34 0.61 0.08 0.11 0.01 0.02
Eurocode
Lever arm [m] 6.55 2.09 6.75 10.42 8.32 0.57 1.13

Tend x Tend x2 Tend y Tend y Tend y2

Loads [MN] 0.03 0.09 0.14 0.19 0.07


RIDAS
Lever arm [m] 4.67 4.01 7.88 5.88 5.88
Loads [MN] 0.03 0.08 0.12 0.16 0.06
Eurocode
Lever arm [m] 4.67 4.01 7.88 5.88 5.88

Table C.10: Dam 10, pillar, height 13 m.

Gravity Water H Water H2 Water H3 Ice Ice 2 Ice 3 Uplift H

Loads [MN] 6.16 1.59 0.07 0.58 0.56 0.23 0.50 0.59
RIDAS
Lever arm [m] 7.12 0.37 8.80 4.60 8.10 8.10 8.10 2.40
Loads [MN] 6.43 1.59 0.07 0.58 0.56 0.23 0.5 0.59
Eurocode
Lever arm [m] 7.12 0.37 8.80 4.60 8.10 8.10 8.10 2.40

Uplift V Soil H Water V Tailw H Tend x Tend x2 Tend y Tend y2

Loads [MN] 2.00 0.04 0.08 0.12 0.78 0.15 1.56 1.74
RIDAS
Lever arm [m] 7.30 2.60 1085.45 480.03 8.16 5.00 4.35 2.05
Loads [MN] 2.00 0.03 0.08 0.12 0.68 0.13 1.36 1.52
Eurocode
Lever arm [m] 7.30 2.60 8.85 1.07 8.16 5.00 4.35 2.05

107
APPENDIX C. OUTPUT VALUES FOR DAMS

Table C.11: Dam 11, pillar, height 7 m.

Gravity Water H Water H2 Water H3 Ice Uplift V

Loads [MN] 1.71 0.48 0.14 0.01 0.12 0.40


RIDAS
Lever arm [m] 2.92 2.14 3.77 5.78 6.21 3.61
Loads [MN] 1.79 0.49 0.14 0.01 0.12 0.41
Eurocode
Lever arm [m] 2.92 2.14 3.77 5.78 6.21 3.61

Table C.12: Dam 12, spillway, height 4 m.

Gravity Water H Uplift V Soil H

Loads [MN] 0.27 0.08 0.08 0.01


RIDAS
Lever arm [m] 2.19 1.38 2.65 0.50
Loads [MN] 0.28 0.09 0.08 0.01
Eurocode
Lever arm [m] 2.19 1.38 2.65 0.50

Table C.13: Dam 13, pillar, height 7 m.

Gravity Water H Water H2 Ice Uplift V

Loads [MN] 0.29 0.09 0.01 0.05 0.05


RIDAS
Lever arm [m] 2.96 1.43 3.73 4.10 3.37
Loads [MN] 0.30 0.09 0.01 0.05 0.05
Eurocode
Lever arm [m] 2.96 1.43 3.73 4.10 3.37

H.W shutter V.W shutter H weight shutter V weight shutter

Loads [MN] 0.01 0.03 0.01 0.02


RIDAS
Lever arm [m] 5.65 1.56 5.65 1.56
Loads [MN] 0.01 0.03 0.01 0.02
Eurocode
Lever arm [m] 5.65 1.56 5.65 1.56

Table C.14: Dam 14, buttress dam, height 20 m.

Gravity Water H Ice Uplift V Water V

Loads [MN] 12.67 13.23 1.60 0.94 6.43


RIDAS
Lever arm [m] 8.71 6.12 18.03 15.82 13.37
Loads [MN] 13.26 13.48 1.6 0.95 6.55
Eurocode
Lever arm [m] 8.71 6.12 18.03 15.82 13.37

108
Table C.15: Dam 15, pillar and spillway, height 16/2 m.

Gravity Water H Water H2 Water H3 Ice

Loads [MN] 49.64 3.28 8.16 5.76 4.20


RIDAS
Lever arm [m] 12.75 4.97 6.97 1.43 13.17
Loads [MN] 51.79 3.28 8.16 5.76 4.20
Eurocode
Lever arm [m] 12.75 4.97 6.97 1.43 13.17

Uplift V Anch x Anch y Water V Tailw H

Loads [MN] 25.48 0.00 1.65 9.34 0.66


RIDAS
Lever arm [m] 14.40 0.00 0.00 0.00 0.00
Loads [MN] 25.48 0.00 3.89 9.34 0.66
Eurocode
Lever arm [m] 14.40 0.00 21.31 20.93 0.83

Table C.16: Dam 16, pillar, height 18 m.

Gravity Water H Water H2 Water H3 Ice Uplift V

Loads [MN] 29.56 4.71 1.41 10.74 4.42 7.36


RIDAS
Lever arm [m] 15.05 6.33 14.00 6.10 16.67 16.67
Loads [MN] 30.84 4.80 1.44 10.94 4.42 7.50
Eurocode
Lever arm [m] 15.05 6.33 14.00 6.10 16.67 16.67

Table C.17: Dam 17, pillar, height 19 m.

Gravity Water H Water H2 Ice Uplift V

Loads [MN] 20.27 3.98 2.82 2.33 5.69


RIDAS
Lever arm [m] 15.71 5.83 15.50 17.17 16.67
Loads [MN] 21.15 4.06 2.87 2.33 5.80
Eurocode
Lever arm [m] 15.71 5.83 15.50 17.17 16.67

Table C.18: Dam 18, pillar, height 19 m.

Gravity Water H Water H2 Water H3 Ice Uplift V

Loads [MN] 15.97 2.62 1.41 2.23 2.49 3.86


RIDAS
Lever arm [m] 14.62 5.17 15.00 5.17 16.17 16.67
Loads [MN] 16.66 2.67 1.44 2.28 2.49 3.93
Eurocode
Lever arm [m] 14.62 5.17 15.00 5.17 16.17 16.67

109
TRITA - BKN. Master Thesis 455, Concrete Structures 2015
ISSN 1103-4297
ISRN KTH/BKN/EX-455-SE

www.kth.se

You might also like