guidelines for design of dams for earthquake

Upload: shachen2014

Post on 02-Jun-2018

271 views

Category:

Documents


7 download

TRANSCRIPT

  • 8/11/2019 Guidelines for Design of Dams for Earthquake

    1/127

    guidelines for

    design of dams

    for earthquake

    - lu#_,

    AUSTRALIAN NATIONA L

    COM MITTEE ON

    LAR GE DAMS

  • 8/11/2019 Guidelines for Design of Dams for Earthquake

    2/127

    GUIDELINES FOR

    D E S IG N O F D A M S

    FOR EARTHQUAKE

    AUG UST 1998

    AUSTRA LIAN NATIONA L

    COM MITTEE ON

    LARGE DAM S

  • 8/11/2019 Guidelines for Design of Dams for Earthquake

    3/127

    A U S T R A L IA N N A T IO N A L C O M M IT T E E O N L A R G E D A M S

    G U ID E L IN E S FO R D E S IG N O F

    D A M S FO R E A R T H Q U A K E

    AUGUST 1998

    IM P O R T A N T D IS C L A IM E R

    "ANCOLD and its Members, and the Convenor, Members and Assistants of the

    Working Group which developed these Guidelines do not accept responsibility

    for the consequences of any action taken or omitted to be taken by any person,

    whether a purchaser of this publication or not, as a consequence of anything

    contained in or omitted from this publication. No persons should act on the basis

    of anything contained in this publication without taking appropriate professional

    advice in relation to the particular circumstances".

    ANCOLD Guidelines for Design of Dams for Earthquake

  • 8/11/2019 Guidelines for Design of Dams for Earthquake

    4/127

    T A B LE O F C O N T E N T S

    Page No.

    F O RE W O RD

    ANCO LD WORKING GRO UP MEMBER SHIP L IST

    NTRODUCTON1

    EARTHQUAKES AND THEIR CHARACTERISTICS2

    2.1Earthquake Mechansms and Termnoogy2

    22EarthquakeGoundMoon2

    23SuaceRupue3

    24MagntudeandInensty3

    2.5Changes to Seismic Waves Near the Ground Surface4

    2.6Attenuation and Amplification of Ground Motion4

    27Reservor InducedSesmcty5

    EARTHQUAKE HAZARD IN AUSTRALIA 6

    31Gnea6

    32Mechansmo Earhquakes8

    33EahquakeDphs9

    34Evauaono Sesmc Hazad10

    35Aenuaon11

    3.6MaximumCredbe Earthquake Magntude11

    3.7Estimates of Ground Motion and Response Spectra at a Site12

    38EarthquakeHazardMaps12

    SELECTION OF DESIGN EARTHQUAKE 16

    41Denons16

    42Seection of the Design Earthquake20

    4.3 Selection of the Operating Basis Earthquake (OBE) 32

    44Concurrent Load Combinations32

    45Earthquakes Induced by the Reservor33

    46Response Spectra and Acceerograms *33

    D E S I GN OF E M BA N K M E N T D A M S A N D A N A LY S IS OF

    LQUEFACTON33

    5.1 Effect of Earthquake on Embankment Dams3 3

    5.2 General ("Defensive") Design Principles for Embankment Dams 34

    5.3Liquefaction of Dam Embankments and Foundations36

    ANCOLD Guidelines for Design ofDams for Earthquake

  • 8/11/2019 Guidelines for Design of Dams for Earthquake

    5/127

    6. SEISMIC STABILITY ANALYSIS OF EMBANKMENTS 57

    61Peame57

    62Pseudo-Sac Anayss57

    6.3 Simplified Methods of Deformation Analysis59

    6.4Post Liquefaction Stability and Deformation Analysis63

    65Numerca Mehods65

    66PoposedGudenes67

    7.ANALYSIS AND DESIGN OF CONCRETE DAMS69

    7.1 Past Performance of Concrete Dams in Earthquakes69

    72DeensveDesgnMeasues70

    73AnayssMhods71

    7.4Design Earthquake and Hydrodynamic Loads82

    75DsgnCea83

    76Dynamc Maea Popetes86

    8APPURTENANT STRUCTURES87

    81noducon87

    82PerormanceRequremens87

    83nakeTows89

    RE F E RE NC E S

    APPEND IX A

    TERMS OF REFERENCE

    APPENDIX B

    TYPICAL EASTERN AUSTRALIAN PEAK

    G RO UND A C C E L E RA T IO N V S A E P

    RE S PO NS E S PE C TR UM FO R 1 in 1000 A E P

    MODIFIED MERCALLI SCALE

    APPENDIX C

    EXTRACTS FROM CANADIAN DAM SAFETY

    GUIDEL INES

    APPENDIX D

    ADDITIONAL INFORMATION ON ACCEPTABLE

    RISKS

    ANCOLD Guidelines for Design ofDams for Earthquake

  • 8/11/2019 Guidelines for Design of Dams for Earthquake

    6/127

    F O R E W O R D

    Even in the matter of earthquakes, Australia can be considered the "Lucky Country" in not being on the

    edge of major tectonic plates. Neighbouring countries like New Zealand and Indonesia are renowned for

    their volcanoes and frequent earthquakes. Australia is relatively earthquake free by comparison and

    earthquakes were seldom considered in early dam designs.

    Certainly there were some zones of known activity such as the Adelaide Hills and the Western

    Australian wheat belt, and whilst major damage had occurred, it was not on the same scale as in other

    countries.

    In 1979, the Standards Association of Australia produced the "Earthquake Code" AS2121. It showed

    zones of seismic activity and recommended methods of determining loads on building structures. The

    development of this code was based largely on statistics of historic earthquakes, for which there were

    relatively short term records.

    However, several major earthquakes subsequently occurred in areas indicated by the code as having

    negligible earthquake risk, the most notable being the 1989 earthquake at Newcastle (Magnitude 5.6) in

    which 12 people died and the Tennant Creek Earthquake in 1988 (Magnitude 6.8). This led to the

    introduction of a new earthquake code (AS 11 70.4-1993)which included data from m ore widespread and

    reliable seismographs and furthermore considered the all important geological situations.

    In parallel with these developments, analytical methods used by dam engineers were improving beyond

    the simplistic application of a horizontal force equating to seismic acceleration. Improvements were

    based on the observed fact that earth dams subjected to earthquakes had slumped vertically rather than

    fail by slipping of a face as indicated by the simplistic analyses.

    Methods of analysing slumping were developed, and further supplemented by sophisticated finite

    element analyses which, by utilising modem computer power, give an ability to undertake rigorous

    analyses of dams where necessary.

    This ANCOLD Guideline brings together improved appraisals of the earthquake loadings that a dam

    may suffer and then describes appropriate methods for analysis and evaluation. Whilst specific to the

    Australian considerations, the majority of this guideline could be applied to dam structures throughout

    the world. The IC O LD Bulletins No. 46(198 3), No. 52(198 6), No. 62(198 8) and No. 72(1989 ) are

    parallel documents in this regard, although not including recent advances.

    This guideline is a major contribution to dam engineering and the voluntary work by the ANCOLD

    subcommittee has unselfishly provided their experience to the dam building community and indeed the

    wider community. Our appreciation goes to Prof Fell and his team for producing this valuable guideline.

    This guideline is not a design code, and dam designers must continue to apply their own considerations,

    judgements and professional skills when designing dams to resist earthquakes. As time goes on there

    will no doubt be improved data and design tools to help the designer and it is intended that this guideline

    will be updated as circumstances dictate. ANCOLD welcomes contributions to discussion on this

    guideline which will assist with future revisions.

    /

    JOHN PHILLIPS

    Chairman, ANCOLD

    ANCOLD Guidelines for Design of Dams for Earthquake

  • 8/11/2019 Guidelines for Design of Dams for Earthquake

    7/127

    MEMBERSHIP OF THE ANCOLD WORKING GROUP FOR

    G U ID E L IN E S F O R T H E D E S IG N O F D A M S F O R E A R T H Q U A K E

    4

    Robin Fell

    School of Civil Engineering, University of New South Wales

    Gamini Adikari

    Snowy Mountains Engineering Corporation, Victoria

    John Bosler

    Snowy Mountains Engineering Corporation, Cooma, New South Wales

    Brian Cooper

    Dam s and C ivil Section, Public W orks and Services Department, NS W

    Peter Foster

    Works Consultancy Services, Power Engineering, New Zealand

    Gary Gibson

    Seismology Research Centre, RMIT, Melbourne

    Sergio Giudici

    Hydro-Electric Commission, Hobart

    Nasser Khalili

    School of Civil Engineering, University of New South Wales

    Ian Landon-Jones

    Dams S afety Group, Sydney Water, New S outh Wales

    Kevin McCue

    Australian Seismological Centre, Canberra

    Len McDonald

    Dams and Civil Section, Public Works and Services Department, NSW

    Brian Shannon

    Water Resources, Department of Primary Industries, Queensland

    David Stapledon

    Geotechnical Consultant, Adelaide, South Australia

    John Waters

    Geo-Eng Pty Ltd, Perth, Western Australia

    Ron Wyburn

    Halcrow Water Power, Victoria

    ANCOLD Guidelines for Design of Dams for Earthquake

  • 8/11/2019 Guidelines for Design of Dams for Earthquake

    8/127

    1. INTRODUCTION

    Public awareness of the potential for damage

    and loss of life in Australia from earthquakes

    was highlighted by the Newcastle earthquake in

    December 1989. This was a Magnitude 5.6

    (M5.6) event, and because of its proximity to

    Newcastle, and local ground conditions, caused

    approximately $1 billion damage.

    Dam engineers in Australia have been

    conscious of earthquakes for many years, but it

    was the earthquakes at Tennant Creek in 1988,

    which were M6.3, M6.4, M6.7, with a total fault

    scarp length of 32km which raised the question

    most acutely as to whether dams in Australia

    could be subject to large earthquakes, and if so,

    could they withstand them without resultant

    loss of the facilities and lives, property, and

    environmental values downstream. Other large

    earthquakes in the M6 to M7 range had

    occurred in Australia, the most notable being in

    Meckering in 1968 (M6.9), but the Tennant

    Creek event was critical because it occurred in

    an area which had previously been regarded as

    virtually free of earthquakes.

    Recent assessments of earthquake ground

    motions for some large Australian dams have

    been based on the assumption that the

    maximum credible earthquake is M7.5, which is

    large by any standards. The seismologists

    involved in these studies indicate that on the

    available evidence, such earthquakes, ie. M7.5,

    could occur anywhere in Australia.

    In general, it is not possible to identify active

    faults which might cause such earthquakes. For

    example, there had been no movement on the

    Tennant Creek fault for more than 200,000

    years (Crone and Machette, 1992), so the

    question arises, can it occur at, or close to any

    damsite? Peak ground accelerations close to a

    M 7.5 earthquake can be very high.

    The past performance of dams in earthquake

    h&s been very good, with few dams suffering

    major damage. Where this has occurred, it has

    been due to liquefaction in the dam or the

    foundation. Very few of these dams have

    breached and released a flood wave. However,

    several might have breached, if the reservoir

    level had been higher at the time of the

    earthquake. Seed (1979), USCOLD (1992),

    ICOLD (1986), NSWDSC (1993) and Hinks

    and Gosschalk(1993) give some details.

    The approach taken by seismologists in

    Australia is to use statistical analysis to predict

    the frequencies of recurrence of ground

    motions. This typically results in 1 in 1000

    AEP peak ground accelerations of 0.15g, 1 in

    10,000 AEP s0.35g, and 1 in 100,000 AEP

    ~0.5g. These are large loadings and it is likely

    that assessment of many of the existing dam s in

    Australia for such loads could indicate some

    deficiencies to either the dam or appurtenant

    structures. New dams would also need (less)

    expensive additional design features to cope

    with earthquake.

    In recognition of the need to provide some

    guidance to dam engineers and owners in

    Australia, ANCOLD established a Working

    Group to prepare Guidelines for the Design of

    Dams for Earthquake. The Working Group was

    established in September 1993, and took over

    from an earlier ANCOLD Working Group

    preparing Guidelines on Seismic Analysis and

    Design of Embankment Dams. The Terms of

    Reference for the Working Group are in

    Appendix A.

    These guidelines are to cover all types of dams,

    including tailings dams, and apply to existing

    and new dams. They cover the selection of the

    design earthquake, analysis and design of

    embankment and concrete dams, and

    appurtenant structures.

    The guidelines are not meant to be used as a

    design code, and of necessity, do not include

    complete details of all the analysis and design

    methods which are recommended. The area is

    rapidly evolving, and those involved in the

    analysis and design of dams for earthquake

    should refer to the references given, and to

    more recent publications so as to be fully

    informed. In some situations it will be

    necessary to seek specialist advice.

    ANCOLD Guidelines for Design ofDams for Earthquake

    1

  • 8/11/2019 Guidelines for Design of Dams for Earthquake

    9/127

    2. EARTHQUAKES AND

    T H EIR CH A R A CT ER IS T ICS

    2.1 Earthquake Mechanisms and

    Terminology

    An earthquake is the motion that is produced

    when stress within the earth builds up over a

    long period of time until it eventually exceeds

    the strength of the rock, which then fails and a

    break along a fault is produced. It may take

    tens, hundreds or thousands of years for the

    stress to build up in a particular area, and it is

    then released in a few seconds. Part of the

    energy is transmitted away as seismic waves

    and part of it as heat.

    The fault displacement in a particular

    earthquake may vary from centimetres up to a

    few metres in a great earthquake. Once

    ruptured, the fault is a weakness which is more

    likely to fail in future earthquakes, so a large

    total displacement may build up from many

    earthquakes over a long period of time. This

    may eventually measure kilometres for thrust

    faults produced by compression, or hundreds of

    kilometres for horizontal strike-slip faults such

    as the San Andreas.

    The point on the fault surface where a

    displacement commences is called the

    hypocentre or focus, and the earthquake

    epicentre is the point on the ground surface

    vertically above the hypocentre. The

    displacement usually propagates along the fault

    in one direction from the hypocentre, but

    sometimes it propagates in both directions.

    Energy release is near but not exactly at the

    hypocentre.

    The hypocentral distance from an earthquake to

    a point is the three dimensional slant distance

    from the hypocentre to the point, while the

    epicentral distance is the horizontal distance

    from the epicentre to the point.

    2.2 Earthquake Ground Motion

    Earthquake ground vibration is recorded by a

    seismograph or a seismogram. Most modem

    seismographs record three components ol

    motion: east-west, north-south and vertical.

    The rupture time for small earthquakes is a

    fraction of a second, for earthquakes of

    magnitude 5.0 it is about a second, and for large

    earthquakes may be up to tens of seconds.

    However the radiated seismic waves travel al

    different velocities, and are reflected and

    refracted over many travel paths, so the total

    duration of vibrations at a site persists longei

    than the rupture time, and shows an exponentia

    decay.

    Several types of seismic wave are radiated from

    an earthquake. Body waves travel in three

    dimensions through the earth, while surface

    waves travel over the two dimensional surface

    like ripples on a pond. There are two types oi

    body wave (P and S waves), and two types oi

    surface wave (Rayleigh and Love waves).

    Primary or P waves are ordinary sound waves

    travelling through the earth. They are

    compressional waves with particle motion

    parallel to the direction of propagation.

    Secondary or S waves are shear waves, with

    particle motion at right angles to the direction of

    propagation. The amplitude of S waves from an

    earthquake is usually larger than that of the P

    waves.

    P waves travel through rock faster than S

    waves, so they always arrive at a seismograph

    before the S wave.

    The frequency content of earthquake ground

    motion covers a wide range of frequencies up to

    a few tens of hertz (cycles per second). Most

    engineering studies consider motion between

    about 0.2 and 25 Hz.

    The amplitude, duration and frequency content

    of earthquake ground motion at a site depend on

    many factors, including the magnitude of the

    earthquake, the distance from the earthquake to

    the site, and local site conditions.

    The larger the earthquake magnitude, the

    greater the amplitude (by definition a factor of

    ten for each magnitude unit), the longer the

    2

    ANCOLD Guidelines for Design of Dams for Earthquake

  • 8/11/2019 Guidelines for Design of Dams for Earthquake

    10/127

    duration of motion, and the greater the

    proportion of seismic energy at lower

    frequencies. A small earthquake has low

    amplitude (unless it is very close), short

    duration, and has only high frequencies.

    The smaller the distance from an earthquake to

    the site, the higher the amplitude. The duration

    is not strongly affected by distance. High

    frequencies are attenuated by absorption within

    the ground more quickly than low frequencies,

    so at greater distances the proportion of seismic

    vibration energy at high frequencies will

    decrease.

    2.3 Surface Rupture

    Surface rupture is a relatively rare phenomenon

    which occurs when a fault break reaches the

    ground surface. It may produce a vertical or

    horizontal offset (or both) with a displacement

    of millimetres to a few metres, and a length

    from metres to tens of kilometres.

    Because rock near the surface is relatively

    weak, few earthquake hypocentres occur in the

    top one or two kilometres. It is common for

    surface sedimentary rocks to be folded in

    response to faulting at depth, giving a

    monocline and scarp at the surface, but without

    a surface fault.

    Most earthquakes, especially most larger

    earthquakes, occur on existing faults. This is

    because faults are weaker than surrounding

    unbroken rock, and are much more likely to fail

    again when stress rebuilds.

    A site will have surface rupture potential if an

    existing fault is found which has been active in

    the recent geological past (perhaps the past few

    million years). This will be quite rare, and

    possibly be difficult to establish. It will usually

    be easier to show that a site with simple surface

    geology has no faulting history, than to show

    that a site with complex geology has suffered

    recent faulting.

    2.4 Magnitude and Intensity

    Earthquakes vary enormously in size. In 1935

    Richter defined a magnitude scale to indicate

    the size of an earthquake. For the Richter local

    magnitude scale, ML, the logarithm of the peak

    ground displacement is taken and an empirical

    correction depending on the distance from

    earthquake to seismograph is subtracted. The

    resulting values are averaged for all the

    seismographs that have recorded the

    earthquake.

    Other magnitude scales have been defined,

    including moment magnitude, and while not

    exactly the same as the R ichter local magnitudes,

    they give similar values that can range from 0.0

    to over 9.0. For each unit of magnitude there is

    a tenfold increase in ground displacement, and a

    thirtyfold increase in seismic energy release.

    Another measure of earthquake size is the fault

    area, or the area of the fault surface which is

    ruptured. The fault area ruptured in an

    earthquake depends on the ma gnitude and stress

    drop in the earthquake. For a given magnitude,

    a higher stress drop will give a smaller rupture

    area. Typically, a magnitude 4.0 earthquake

    ruptures a fault area of about 1 square

    kilometre, magnitude 5.0 about 10 square

    kilometres, and magnitude 6.0 about 100 square

    kilometres (perhaps 10 by 10 kilometres).

    Earthquake Intensity is a measure of the effect

    of the seismic waves at the surface, and is

    normally given on the Modified Mercalli

    Intensity scale, a copy of which is attached in

    Appendix B. This is an arbitrary scale defined

    by the effects observed (whether sleeping

    people were woken, trees shaken, etc) and on

    the amount of damage caused. Normally the

    maximum intensity occurs near the epicentre of

    the earthquake, and intensity then decreases

    with distance. However, this may be affected

    by the orientation of the earthquake rupture, or

    by local ground conditions such as topography

    or surface sediments.

    The earthquake recurrence or seismicity

    (seismic activity) of an area must take the range

    of earthquake sizes into account. There are

    many more small earthquakes than large. In

    ANCOLD Guidelines for Design of Dams for Earthquake

    3

  • 8/11/2019 Guidelines for Design of Dams for Earthquake

    11/127

    most places around the earth there are about ten

    times as many earthquakes exceeding

    magnitude 3.0 than there are exceeding

    magnitude 4.0, and ten times as many again

    exceeding magnitude 2.0. In seismicity studies,

    the logarithm of this factor is called the b value,

    so a value of 1.0 is typical. The b value may be

    1.3 or higher if there are many small

    earthquakes, or 0.7 or lower if there are few

    small earthquakes.

    2.5 Changes to Seismic Waves Near

    the Ground Surface

    The energy in seismic waves depends upon

    their amplitude and the physical properties of

    the material through which they are passing.

    When waves pass from high stiffness material

    (eg. rock at depth) into lower stiffness material

    (eg. near-surface rock, or sediments) they are

    reflected towards the vertical and their

    amplitude increases. Their amplitude also

    increases as they approach the earth's (free)

    surface, at which they are reflected. The nature

    and extent of free surface amplification varies

    with topography, even in fresh, strong rock.

    Changes in soil thickness above an irregular

    bedrock surface can give complex surface

    amplification that varies with earthquake wave

    duration.

    Resonance in the surface sediments causes

    amplification at particular frequencies,

    especially at the natural frequency of the

    sediments. This depends on the thickness and

    elastic properties of the sediments. Earthquake

    motion recorded on hard rock includes all

    frequencies up to a value that depends on

    magnitude, while that recorded on soft

    sediments is usually dominated by the resonant

    frequency.

    In surface sediments, high frequency vibrations

    are attenuated much more with distance than

    low frequencies. If sediments are very thick,

    much of the high frequency motion will be lost

    and peak surface accelerations will be low, even

    if resonance has amplified motion at the low

    resonant frequency.

    2.6 Attenuation and Amplification of

    Ground Motion

    Earthquake ground motion attenuates with

    increasing distance from the source due to

    radiation and hysteretic damping. High

    frequency motion is attenuated more quickly

    with distance than lower frequency motion.

    For estimates of peak ground acceleration,

    attenuation is allowed for by using an

    attenuation function of the form

    a =b,eb2Mr*3

    where a =acceeration

    R=foca dstance

    M=Magntude

    b jb are constants, which

    vary considerably over the

    world.

    Some earthquake hazard studies use the Esteva

    and Rosenblueth (1969) attenuation functions,

    which give peak ground velocity (mm/s), peak

    ground acceleration (mm/s2) and Modified

    Mercalli Intensity (IMM) at an epicentral distance

    x kilometres from an earthquake at depth z

    kilometres with local magnitude M. The

    equations are:

    R

    =

    Vx2 +z2 +400

    Vpeak

    = 160 e10 M R"17

    Speak

    =

    20000 e0 8 mR20

    Im m

    =

    loge(2980e15MR-23)

    Because of the 400 term in the expression for R,

    corresponding to a minimum R of 20

    kilometres, these equations give low values of

    ground motion at distances closer than a few

    kilometres.

    These relations were determined using

    Califomian data, and should only be used with

    magnitudes determined using a compatible

    function. If the magnitudes computed for

    seismographs at different distances vary, then

    the attenuation function is invalid for the area .

    4

    ANCOLD Guidelines for Design ofDams for Earthquake

  • 8/11/2019 Guidelines for Design of Dams for Earthquake

    12/127

    In selecting attenuation relationships care is

    needed, and attention paid to the mechanism of

    the source earthquake, eg. whether shallow

    intraplate, or deep crustal boundary

    earthquakes.

    Weak surface materials absorb seismic energy

    rather than transmit it unchanged, thus tending

    to reduce amplitudes at the surface. The

    amount of attenuation depends on the properties

    of the materials, and especially on their

    thickness.

    Near-surface layers will vibrate preferentially at

    their own natural frequencies, depending on

    their thickness and elastic properties. The

    earthquake motion at the natural frequencies of

    the near-surface layers is amplified, while

    motion at other frequencies may be little

    affected or even attenuated. The amplification

    effect can be especially pronounced for deep

    soft sediments such as those underlying M exico

    City, but in deep, stiff sediments subject to high

    frequency earthquake, attenuation may result.

    Dams (like all other structures) have natural

    frequencies of their own depending on their

    mass and stiffness, usually in the range from

    about 0.5 hertz to about 5 hertz for embankment

    dams and 2 hertz to 20 hertz for concrete

    gravity dams.

    2.7 Reservoir Induced Seismicity

    Reservoirs may induce seismicity by two

    mechanisms. Either the weight of the water

    may change the stress field under the reservoir,

    or the increased ground water pore pressure

    may decrease the stress required to cause an

    earthquake. In either case, reservoir induced

    seismicity (RIS) will only occur if relatively

    high stresses already exist in the area. If the

    stress has been relieved by a recent large

    earthquake, say in the last few hundred years

    for low seismicity areas like Australia, then RIS

    is unlikely to occur.

    /

    RIS events initially usually occur at shallow

    depth under or immediately alongside a

    reservoir. As years pass after first filling, and

    groundwater pore pressure increases permeate

    to greater depths and distances, the events may

    occur further from the reservoir. This occurs at

    a rate of something like one kilometre per year.

    RIS is experienced under new reservoirs,

    usually starting within a few months or years of

    commencement of filling, and usually not

    lasting for more than about twenty years. Once

    the stress field and the pore pressure fields

    under a reservoir have stabilised, then the

    probability of future earthquakes reverts to a

    value similar to that which would have existed

    if the reservoir had not been built. Most of the

    earthquake energy does not come from the

    reservoir, but from normal tectonic processes.

    The reservoir simply acts as a trigger.

    In areas with horizontal tectonic compression

    and reverse faulting, like Australia, filling a

    reservoir should increase the vertical minimum

    principal stress and reduce the chance of an

    earthquake under the reservoir. This has been

    called reservoir induced a seismicity. However,

    in some cases earthquakes could then be

    induced by later releasing water from the

    reservoir. Alternatively the change in stress

    during filling could induce earthquakes beside

    the reservoir rather than under it, although this

    stress change is less pronounced.

    It has been suggested that filling a reservoir will

    cause compression under it, increasing the pore

    pressure of the existing groundwater, and so

    tend to induce earthquakes even in areas of

    horizontal compression. Stress change induced

    seismicity, either direct or through this indirect

    mechanism, should occur soon after filling. It

    may then cause seasonal variations in

    seismicity, sometimes lagging a few weeks or

    months behind water level.

    Pore pressure induced seismicity is normally

    delayed, and may occur years after filling. Pore

    pressure increases always tend to induce events.

    If there is a major fault near the reservoir, RIS

    can produce earthquakes exceeding magnitude

    6.0 (Xinfengjiang, China, 1962, M6.1; Koyna,

    India, 1967, M6.3). Such events will only occur

    if the fault is already under high stress. A

    number of Australian reservoirs have triggered

    earthquakes exceeding magnitude 5.0

    (Eucumbene, 1959, M5.0; Warragamba, 1973,

    M5.0; Thomson, 1996, M5.2).

    ANCOLD Guidelines for Design of Dams for Earthquake

    5

  • 8/11/2019 Guidelines for Design of Dams for Earthquake

    13/127

    MPP*8-

    It is more common for a reservoir to trigger a

    large number of small shallow earthquakes,

    especially if the underlying rock consists of

    jointed crystalline rock like granite (Talbingo,

    1973 to 1975; Thomson, 1986 to 1995). These

    events possibly occur on joints rather than

    established faults, so are limited in size, and

    only give magnitudes up to 3 or 4. There is no

    hazard from such low magnitude reservoir

    induced earthquakes, even if they occur

    regularly. Their shallow depth means that they

    may often be felt or heard.

    RIS has been observed for over one hundred

    reservoirs throughout the world, and small

    shallow induced events have probably occurred

    under many others. A relatively high

    proportion of reservoirs with RIS seismograph

    networks do record such activity. A high

    proportion of RIS examples occur in intraplate

    areas, with above average rates in China,

    Australia, Africa and India.

    It is not easy to predict whether a reservoir will

    experience RIS because the stress and strength

    at earthquake depths are not easily measured.

    For the same reason, prediction of normal

    tectonic earthquakes has been unsuccessful in

    most parts of the world.

    It seems that RIS with many small events is

    more likely to occur in intraplate areas with

    near surface crystalline rocks like granite, rather

    than sedimentary rocks. A larger magnitude

    RIS event can only occur if there is an existing

    fault of sufficient dimension that is late in its

    earthquake cycle (the stress is already

    approaching the strength of the fault).

    3. EARTHQUAKE HAZARD

    IN AUSTRALIA

    The understanding of the hazard imposed by

    earthquakes in Australia is critical to selection

    and application of design earthquakes for dams.

    Hence, a relatively detailed discussion on the

    topic follows. This is largely taken from Gibson

    (1994).

    3.1 General

    The Australian continent is within a tector

    plate shared with Southern India, so all of

    earthquakes are intraplate. The pla

    boundaries to the north and east are among t

    most active on the earth. Possibly as a result

    this, Australia is one of the most actr

    intraplate areas on the earth. Despite this, tf

    hazard is quite low when compared with acth

    interplate areas.

    Most people in Australia can expect to feel

    earthquake about every five or ten year

    although many of these may not be recognisf

    as an earthquake. Most Australian earthquake

    that are reported are heard with a noise like

    distant quarry blast or thunder, with possibly

    slight vibration being felt.

    Only a proportion of earthquakes that are fel

    perhaps about one in twenty, will cause som

    damage in their epicentral area. If they occur i

    an inhabited area, most earthquakes larger tha

    about magnitude 4.0 will cause some damage.

    By contrast, in an active interplate area lik

    New Britain or Bougainville in Papua Nev

    Guinea, earthquakes are felt very often, c

    average every week or two. These are normal

    felt rather than heard, with any soun ds being thf

    reaction of a building to the vibration rathe

    than the earthqua ke itself.

    PNC

    havf

    A very small proportion of these

    earthquakes, perhaps about 1 in 500,

    caused any damage in their epicentral area, an

    o

    aO

    q.

    *6

    90o

    . Q

    If

    s >

    O do

    a

    o

    *

    8o

    Q

    < ?

    o 00

    o

    o C O

    o

    o

    a!?

    H

    Australian Earthquakes to 1994

    Magntudes: *405o

    Figure 1. Australian earthquakes with magnitude exceeding M L4.0 since 1850 (Gibson, 19 94).

    Earthquakes on reverse faults usually give a

    high stress drop, where the seismic energy

    comes from a small source volume. High stress

    drop earthquakes radiate a greater than average

    proportion of their energy in higher frequencies.

    Higher frequency motion implies higher

    accelerations for a given energy release or

    earthquake magnitude, but not necessarily

    fewer cycles, than for similar magnitude

    earthquakes in, say, West Coast USA.

    Compression giving reverse and thrust faul

    produces surface uplift. Therefore, earthquakf

    are most likely to occur in areas whei

    mountains are developing, and less likely und0)

    jjquiiocScn

    Pm+mnca\ Mo

    Jogontt daw (

    CtMHdol , A4

    > Iw^BeNan

    2030

    *0

    40

    Figure 14. Relationship between stress ratios causing liquefaction and (N,),# values for clean sa

    magnitude 7.5 earthquakes. (Seed and De Alba, 19 86).

    2030

    < N * 0

    Figure 15. Relationship between stress ratios causing liquefaction and (N,)^ values for silty sa

    magnitude 7.5 earthquakes (Seed and De Alba, 1986).

    (iii) For earthquakes of magnitude other

    than 7.5, correct the values of Tav/a'0 by

    the factors in Table 13.

    44

    ANCOLD Guidelines for Design of Dams for Earthquake

  • 8/11/2019 Guidelines for Design of Dams for Earthquake

    52/127

    Table 13. Representative Number of Cycles and Corresponding Correction Factors (Seed and De

    Alba, 1986).

    Earthquake

    Magnitude (M)

    Number of

    Representative Cycles

    at 0.65 t

    Factor to Correct

    Abscissa of Curve in

    Figures 14 and 15

    8.5

    26

    0.89

    7.5

    15

    1.00

    6.75

    10

    1.13

    6.0

    5-6

    1.32

    5.25

    2-3

    1.5

    It should be noted that the magnitude of

    the earthquake has a significant

    influence on whether liquefaction will

    occur. It may be assumed that

    liquefaction will not occur for

    earthquakes of Magnitude 5 or less

    (there are not sufficient cycles of

    loading for small earthquakes).

    However, where static liquefaction may

    occur, even small earthquakes could

    trigger failure. Morgenstem (1995)

    discusses static liquefaction which is

    likely to occur only in very

    loose/poorly compacted, saturated soils

    including dredged sand, mine

    overburden dumps and mine tailings.

    Salmon (1995) points out that when

    earthquake loading is determined by a

    probabilistic analysis of the history of

    earthquakes in the seismotectonic zone

    around the dam (as is done in

    Australia), it is possible to determine

    the contribution of different magnitude

    earthquakes to the assessed ground

    motion. Figure 16 gives an example

    which shows that the major

    contribution to the estimated ground

    motion for a given PGA (in this case

    the 1 in 1000 PG A) is from small

    magnitude earthquakes near the dam.

    Given the nonlinear relationship

    implied in Tab le 13, and Figures 14 and

    15, this has an important influence on

    the assessment of the probability of

    liquefaction, and should, if practicable,

    be included in the assessment.

    /

    ANCOLD Guidelines for Design of Dams for Earthquake

    45

  • 8/11/2019 Guidelines for Design of Dams for Earthquake

    53/127

    50

    GO

    OS

    I t

    o

    10

    50

    Suniat9 0

    M A G N IT U D E C O N T R IB U IIO N S

    -t-

    Sums5 6

    lb.

    34567

    M A C N I IU O E ( M )

    oisiANcr coNfRinunnN^

    Nnle ConUfbolKjftS /ro

  • 8/11/2019 Guidelines for Design of Dams for Earthquake

    54/127

    - 4

    o

    a.

    c

    o

    2 3

    - 0

    O JomfoUowtki tf oL 09091

    0 MurorrocM and KoboroiM UM 2)

    A (ihihofo and K090 (Oil )

    it Robvrlion (1902)

    V MHehtll (1903)

    O HoriNr > oL (1964)

    ne '

    bkul/lool

    OO 002 003 01 0205 I

    Mean Groin Sit*, O0 -mm

    Figure 17. Variation of ratio with mean grain size (qc) measured in tsf 100 kPa) (Seed and De

    Alba, 1986).

    06

    V 0'

    Z 03

    >

    u

    o

    o.i

    -11-

    M> 7.3 orthquaku

    % riots >33 213 slO 13

    Ojolmm) 01 0.2 0.2302304 08

    le/Ngo 3.3 41 4.4 4,4 4 33

    _1_

    X

    4080120160200

    Modifiad Con* P*n*tralion R*iilanc*. qe|-lf

    240

    Figure 18. Relationship between stress ratio causing liquefaction and cone tip resistance for sands and

    silty sands (1 tsf100 kPa (Seed and D e Alba, 19 86).

    There are several problems in applying the Seed

    et al semi empirical approach:

    (a) It has been developed for level or near

    level ground conditions. For most dam

    applications it can therefore only be

    directly applied to assess whether

    liquefaction would occur without the

    dam.

    ANCOLD Guidelines for Design of Dams for Earthquake

    47

  • 8/11/2019 Guidelines for Design of Dams for Earthquake

    55/127

    Seed and Harder (1990) recommend

    that to allow for the static driving stress

    due to the dam, a correction factor

    should be applied to the calculated

    (Tav/CT'0). This is calculated using

    (xayCT'0)a^=(xav/a'0)a^ Ka

    where Ka is a correction factor

    determined from Figure 19. To

    determine Ka, the relative density and

    a (the ratio of static driving shear stress

    on a horizontal plane to the initial

    effective overburden stress, (tav/a'0) has

    to be determined, eg. from finite

    element analyses.

    The correction only applies to soils

    where Tav/a'0

  • 8/11/2019 Guidelines for Design of Dams for Earthquake

    56/127

    (c) In most natural soil deposits the SPT

    values are variable (over that increase

    which occurs with the increasing

    overburden stresses). There are no

    clear guidelines given by Seed et al as

    to how this should be accounted for.

    US SR (1989 (a)) discuss this issue and

    present some useful practical points

    which are recommended for use. These

    include:

    The results of each interval of each

    drill hole, with regard to liquefaction

    potential, should be prepared in a

    table and should be presented on

    geologic cross sections and profiles

    to allow examination of the

    frequency and continuity of those

    intervals indicating liquefaction

    susceptibility. From such a

    presentation a judgement is drawn

    as to whether or not the continuity

    of potential liquefaction intervals

    indicated is great enough to be of

    concern.

    If the deposit being sampled is

    known to contain, or may contain

    gravel, these coarse particles may

    increase the blow count, implying a

    more dense, less potentially

    liquefiable soil. To check for this,

    USBR (1989(a)) recommend

    recording SPT blow counts for each

    300mm of penetration, and

    correcting for the erratic effects of

    gravel (as a minimum, the three sets

    of 150mm blow counts should be

    checked to see whether this potential

    irregularity is present). Where

    gravel is extensive, shear wave

    velocity methods should be used to

    assess liquefaction potential.

    X

    }*

    *x

    V

    X

    X

    X

    5s .* *

    X

    t

    X

    35%

    =0 ifFC

  • 8/11/2019 Guidelines for Design of Dams for Earthquake

    58/127

    It is important to note that a positive result from

    the Seed method only indicates that a soil may

    be susceptible to liquefaction. It does NOT

    mean that just because (part of) the dam

    foundation will liquefy, the dam will fail. What

    should follow such an outcome, is an

    assessment of the extent and continuity of

    potentially liquefiable soil, and the residual

    undrained strength of that soil, for use in post

    liquefaction analysis.

    Liao, Venziano and Whitman (1998) provide

    some useful information which will help in the

    application of liquefaction in a probabilistic

    framework.

    5.3.4 Shear Wave Velocity Methods for

    Assessing Liquefaction Potential

    Shear wave velocity is affected by many of the

    variables which influence liquefaction, eg.

    (relative) density, confining pressures, stress

    history, geologic age. Hence, it has some use as

    an indicator of potential for liquefaction.

    The shear wave velocity may be obtained by

    downhole, crosshole or surface to downhole

    seismic methods, or by a seismic cone

    penetration test (a modification of the

    piezocone test) developed by Campanella et al

    (1986).

    As discussed above, some coarse grained

    cohesionless materials (gravels, cobbles)

    suspected of being potentially liquefiable

    cannot be successfully sampled using SPT. If

    crosshole shear wave velocity data have been

    obtained on the materials, and these data

    accurately represent the deposits in plan and

    section, then they provide a viable means for

    making a judgement on liquefaction potential.

    USBR (1989) recommend that:

    if shear wave velocities are >365m/s, the

    deposits may be judged non liquefiable

    if shear wave velocities are between 245 and

    365m/s, the deposit may be considered

    likely to be non liquefiable, but supporting

    evidence should be obtained

    if shear wave velocities are < 245m/s, the

    deposit may be judged liquefiable.

    In practice, this will leave many sites in the

    "grey" zone.

    Shear wave velocity may also be used to assess

    liquefaction by relating peak ground

    acceleration, and a history of performance of

    sites in earthquakes. Bierschwale and Stokoe

    (1984) and USNRC (1985) give details.

    US N RC (1985) also gives details of a method

    which assesses peak strains due to the

    earthquake from the shear wave velocity, and

    compares this to threshold strain. Brief details

    of these methods are given in Fell et al (1992).

    5.3.5 Determination of Residual Undrained

    Strength

    In recent years there has been quite a lot of

    discussion of the post liquefaction condition.

    This is usually discussed in terms of "residual

    (undrained) strength", "field residual strength",

    or "steady state undrained strength". Some of

    these are expressed as plots of residual

    undrained strength versus SPT 'N' value (eg.

    Seed (19 87), Lo and Klohn (1990), Seed and

    Harder (1990). These plots are developed from

    backanalyses of liquefaction failures. Figure 22

    is the plot presented by Finn (1993).

    Finn (1993) indicates that lower bound

    strengths are often used for these analyses,

    although 33rd percentile values have sometimes

    been used. In either case, the lower bound or 33

    percentile gives very low (-zero) residual

    undrained strengths at SPT less than about N =

    6.

    An alternative approach adopted by a numb er of

    authors including Ishihara (1994) and Finn

    (1993), is to relate the normalised residual

    undrained strength Su/ct'vo (where SU5 = residual

    undrained strength and ct'Vo effective vertical

    stress) to either SPT (N,)^ values (N values

    corrected to 100 kPa effective stress, 60%

    energy ratio hammer) or to cone penetration

    resistance qc.

    ANCOLD Guidelines for Design ofDams for Earthquake

    51

  • 8/11/2019 Guidelines for Design of Dams for Earthquake

    59/127

    2000

    I60C

    H

    O

    Z

    tu

    cr

    e

    1.3 and maximum

    compressive stress

  • 8/11/2019 Guidelines for Design of Dams for Earthquake

    87/127

    Figure 34 Sequence of Analysis for Operating Basis Earthquake

    (Guthrie 1986)

    The preceding method could probably be used

    for arch and buttress dams as well. However,

    consideration would have to be given to the

    appropriateness of the damping ratios and

    corresponding limiting tensile strengths used.

    (c) Non-linear Dynamic Analysis

    A non-linear dynamic analysis of a concrete

    dam is a complex analysis and would normally

    be undertaken by specialist numerical analysts

    experienced in such work. It would normally

    be done only for major dams where the cost of

    the new dam or the cost of remedial works for

    80

    ANCOLD Guidelines for Design of Dams for Earthquake

  • 8/11/2019 Guidelines for Design of Dams for Earthquake

    88/127

    an existing dam is sufficient to justify the

    greater expense of this type of analysis.

    Provided peak tensile stresses (static +

    dynamic) do not cause cracking in the dam,

    then a linear analysis will suffice. When

    cracking occurs, stresses will re-distribute.

    Therefore, when cracking is expected or there

    are pre-existing cracks (e.g. open vertical

    contraction joints) and the dam is essentially

    3D m nature then a non-linear analysis should

    be done.

    There are three main ways in which cracks can

    be modelled:

    as distributed (smeared) cracks (zero elastic

    modulus in direction perpendicular to

    cracks)

    as discrete cracks at finite element

    interfaces

    using fracture mechanics.

    The first of the alternatives is computationally

    the simplest. Further details on cracking are

    given by Zienkiewicz, Valliappan and King

    (1968), Rashid (1968), Mohraz, Schnobrich

    and Gomez (1970), Darwin and Pecknold

    (1978), Phillips and Zienkiewicz (1976),

    Bazant and Cedolin (1979), Bazant and Ob

    (1979), Argyris, Krempl and William (1977),

    Gerstle (1981), Kotsovos and Newman (1978),

    William and Warnke (1975), Cedolin, Crutzon

    and Dei Poli (1977), Bicanic and Zienkiewicz

    (1983), Zienkiewicz, Fejzo and Bicanic

    (1983), Zienkiewicz, Hinton, Bicanic and

    Fejze (1980), Pande and Shen (1982), Pal

    (1974) and Chapuis, Rebora and Zimmermann

    (1985).

    A non-linear analysis will by its nature,

    require a time-history analysis. Consideration

    will have to be given to:

    the way the compressibility of the storage

    water is modelled

    the way seepage pressures particularly

    those due to water penetrating cracked

    zones, are modelled.

    The non-linear analysis of concrete dams is

    still a developing and specialised field. Other

    relevant papers include Waggoner, Plizzari

    and Saouma (1993), Gao Lin, Jing Zhou and

    Chuiyi Fan (1993), Greeves and Taylor

    (1992), Cervera, Oliver and Galindo (1992),

    Jing Zhou and Gao Lin (1992), Clough and

    Ghanaat (1993), Fenves and Mojtahedi (1993).

    7.3.4 Analysis of Permanent Deformations

    In some concrete gravity dams subjected to

    say the maximum credible earthquake, it may

    be permissible for the dam to slide on its base

    or within the foundations and be permanently

    deformed after the earthquake. This of course

    assumes that during or after the deforming

    process, the security of the storage is not

    preudiced allowing for the potentially

    increased uplift pressures and lower strength

    which may apply. Researchers such as Chopra

    and Z hang (1991), Leger and Katsouli (1989)

    and Danay and Adeghe (1993) discuss

    calculations which indicate that typical

    permanent displacements for large concrete

    gravity dams subjected to earthquakes having

    peak ground accelerations the order of 0.5g

    can range from tens of centimetres to more

    than half a metre. However, some dams with

    suitable foundations would be able to

    withstand small displacements. Dams relying

    on post-tensioned ground anchors or drain

    holes for normal load static stability, might not

    be able to withstand these sort of movements

    if the displacement was sufficient to shear the

    anchors. However it may be acceptable to

    have the anchors sheared for a low probability

    earthquake provided the dam was stable under

    the post earthquake load case.

    The type of analysis required to compute

    permanent deformations is similar to that

    carried out for embankment dams i.e. Makdisi

    and Seed (1978). The dam is considered as a

    block with a limiting sliding strength along its

    foundations. The dam is subjected to a time

    varying input of acceleration. When the

    acceleration is greater than the limiting

    acceleration (the acceleration causing inertia

    forces which are greater than the sliding

    strength of the foundations) the dam will move

    on its foundations. The parts of the

    accelerogram greater than the limiting

    acceleration are double integrated to obtain

    cumulative displacements.

    The type of analysis just described is a

    requirement of the US Corps of Engineers

    method for dynamic analysis of concrete

    gravity dams when the sliding safety factor is

    less than one. The sliding analysis is carried

    out for horizontal planes through the dam

    where there is cracking. A crack is assumed

    through the dam with suitable slip elements

    along the crack interface.

    ANCOLD Guidelines for Design of Dams for Earthquake

    81

  • 8/11/2019 Guidelines for Design of Dams for Earthquake

    89/127

    In their study of base sliding, Chavez and

    Fenves (1993) found that sliding was not

    likely to occur if the storage is less than half

    full. Other conclusions were :

    vertical ground motion has almost no effect

    on the sliding displacement (it slightly

    increased the maximum stresses). This is

    partly because the vertical and horizontal

    ground motion do not necessarily coincide.

    the assumption of rigid foundation rock can

    significantly overestimate the amount of

    base sliding

    At the present stage of development of the

    computation of permanent deformations and

    the determination of suitable maximum

    displacements there is still much work to be

    done especially regarding local conditions.

    These guidelines therefore counsel that

    considerable care be used if a dam is to be

    allowed a permanent deformation following an

    earthquake. Adequate sliding and overturning

    stability must exist after the earthquake using

    foundation strengths appropriate to the

    displacement (usually the residual strength)

    and uplift appropriate to the displaced

    condition, allowing for opening of joints and

    bedding, and for reduced (or no) drainage

    capacity.

    7.4 Design Earthquake and

    Hydrodynamic Loads

    7.4.1 Earthquake Parameters

    As discussed elsewhere in these guidelines,

    dynamic analysis can be carried out in the

    frequency domain or the time domain. For the

    former, response spectra are required while for

    the latter, accelerograms are required.

    (a) Response Spectra

    A response spectrum shows the extent to

    which any single degree of freedom structure

    with an assumed level of damping would

    respond to particular earthquakes.

    Knowing the natural frequencies of vibration

    and the corresponding mode shape for a

    structure, the spectral accelerations

    corresponding to particular natural frequencies

    and damping ratios can be converted to inertial

    loads.

    The response spectrum used in the analysis of

    a dam should be site specific and relate to the

    peak ground accelerations examined. The

    response spectrum should also reflect the

    frequency mix and duration of the design

    earthquakes. It will therefore be derived from

    a number of earthquakes having various

    epicentral distances from the site and

    consequently, different acceleration

    attenuation functions. The response spectra

    should be obtained from a seismologist as part

    of the assessment of seismicity of the dam site.

    (b) Accelerograms

    Where a time-history analysis is to be done, at

    least three different accelerograms appropriate

    to the dam site and for a particular peak

    ground acceleration, should be used. These

    accelerograms may be recorded accelerograms

    which are suitably scaled (accounting for

    change in frequency mix and phase with

    change in peak ground acceleration) or

    synthetic accelerograms which fit the response

    spectra for the site. Care should be taken in

    selecting accelerograms which are similar to

    Australian earthquake conditions, and advice

    should be obtained from a seismologist.

    7.4.2 Hydrodynamic Pressures

    Any movement of the dam and foundation will

    cause movement in the water of the storage

    and in turn, the pressures generated by the

    water will impose forces on the dam.

    Engineers have traditionally used

    hydrodynamic pressures derived by

    Westergaard (1933). These pressures are

    commonly converted into equivalent lumped

    'virtual' masses which are attached to the dam.

    Westergaard's pressure distribution assumes

    that the water in the storage is incompressible

    and that the dam and its foundations are rigid.

    However, this is not always so. In high

    gravity dams and slender arch dams especially,

    where the dam is flexible, there can be

    considerable interaction or coupling between

    the dam and the storage.

    Considerable work on the interaction of

    gravity dams and their storages has been done

    by Chopra and his fellow researchers. Details

    of the work are given in Chopra (1967),

    Chakrabarti and Chopra (1974), Chopra,

    Chakrabarti and Gupta (1980), Chopra and

    Gupta (1981). Other relevant papers include

    82

    AN CO LD Guidelines for Design ofDams for Earthquake

  • 8/11/2019 Guidelines for Design of Dams for Earthquake

    90/127

    Clough and Chang (1980), Dungar (1978),

    Hall (1988), Zienkiewicz, Paul and Hinton

    (1983) and Tsai and Lee (1989).

    Besides the lumped, 'virtual' mass approach,

    the interaction of a dam with the water in its

    storage can be determined by treating the

    water as a 'solid' having zero shear modulus

    but retaining compressibility. This approach,

    although simple in principle, has a number of

    numerical problems. An alternative approach

    is therefore preferable where from the

    beginning, the shear components of stress in

    the fluid are neglected. This latter approach

    includes the effect of water compressibility. It

    also will allow the limiting case of

    incompressible water to be considered.

    Accounting for water compressibility and

    coupling of a dam with the water in its storage

    adds considerable computational effort to

    determining the effects of earthquake loading

    on a dam. Neglecting coupling and water

    compressibility in the simpler "added mass"

    approach is not considered significant for

    excitations at frequencies below the natural

    frequency of the reservoir.

    An estimate for the fundamental frequency of

    a reservoir can be obtained from:

    f =__

    w4H

    eff

    where: fw = the fundamental frequency

    C = the compression wave speed

    in water (l,439m/s)

    Heff = an effective depth of the

    reservoir.

    The above relationship was obtained from

    Duron, Ostrom and Aagaard (1994) and

    applies strictly to an infinite reservoir of

    constant cross section. Duron and Hall (1988)

    indicate that if the ratio of to the

    fundamental frequency of the dam-foundation

    alone (fj) is near unity, water compressibility

    will have a significant effect. For ratios of fw

    to fj much greater than one (e.g. >1.5)

    incompressible fluid behaviour can be

    assumed.

    7.4. i Uplift/Seepage Pressures

    The USBR (1977) considers that the uplift

    pressure within the crack is zero while ICOLD

    (1986) assumes full headwater pressure but

    recognises the need for further research.

    Guthrie (1986) uses the pre-crack uplift

    pressure diagram.

    These guidelines recommend that for the

    duration of the earthquake, the pre-earthquake

    uplift pressure distribution is used for the

    stability analysis. However for the post

    earthquake analysis, consideration should be

    given to the amount of cracking and the post-

    earthquake efficiency of the dam's drainage

    system. In the post-earthquake situation, full

    headwater pressure is assumed to exist in a

    crack at least as far as the line of drains. If the

    drains have sufficient capacity and they have

    not been disrupted by sliding of the dam, then,

    if the crack extends, past the drains, a

    significant reduction in uplift pressure should

    be considered. Typically, the pressure at the

    line of drains in this case might be the

    tailwater pressure plus 50% of the difference

    between headwater and tailwater pressures.

    The pre-earthquake uplift pressure distribution

    might have had a 67% reduction. The lesser

    reduction for the post-earthquake case allows

    for the greater amount of drainage with which

    the drains would have to cope.

    7.5 Design Criteria

    7.5.1 General Approach

    The working group has had two major

    difficulties in preparing suitable design criteria

    for concrete gravity dams.

    (1) The current ANCOLD (1991)

    guidelines for design criteria for concrete

    gravity dams are based on a limit state

    approach with partial factors of safety. This

    method has proven to be difficult to use on

    dams subject to significant earthquake loads.

    The ANCOLD (1991) guideline is under

    review to address these problems. In the

    interim, it is recommended that the design

    loadings and acceptance criteria described

    herein are used. These are based largely on

    the BC Hydro (1995) guidelines.

    (2) Existing guidelines for design of

    concrete gravity dams are not simply applied

    to a risk based approach. As a result it has not

    been practicable to develop these guidelines to

    directly apply to a risk based approach, and

    they are given in terms of a deterministic

    method using OBE and MDE. For

    completeness, flood and static load case are

    also listed. Those wishing to use a risk based

    ANCOLD Guidelines for Design of Dams for Earthquake

    83

  • 8/11/2019 Guidelines for Design of Dams for Earthquake

    91/127

    approach may do so, taking account of the

    general intention of the load cases detailed

    herein.

    7.5.2 Loads and Load Cases

    The loads considered in the assessment of

    concrete gravity dams and their foundations

    should include the following:

    Dead loads of permanent structures,

    equipment and foundation rock (D).

    Water load due to maximum normal

    headwater level combined with the most

    critical concurrent tailwater level (H).

    Water load due to maximum flood

    headwater level based on the Inflow Design

    Flood (IDF) with corresponding tailwater

    levels (H).

    Foundation uplift (U), both at the concrete-

    rock contact and at critical discontinuities

    within the foundation.

    Static and dynamic thrust created by an ice

    sheet, for reservoirs subject to freezing (I).

    Vertical and horizontal loading due to rock

    or soil backfill (both natural or engineered),

    including potential effects of liquefaction

    and loads from silt deposited against the

    dam (S).

    Load due to Operating Basis Earthquake

    (OBE) Q'

    Load due to Maximum Design Earthquake

    (MDE ) (Q).

    Determination of the loads should take into

    account the actual field conditions and

    instrumentation records.

    Foundation uplift assumptions should reflect

    the stress state and condition of the dam and

    foundation. Disruption of the dam and/or

    foundation condition due to an earthquake

    should be recognised in assessing the uplift

    assumptions for the post-earthquake case.

    The dam and foundation should be assessed

    for the following load cases:

    (a) Usual Load Case

    permanent and operating loads should be

    considered for both summer and winter

    conditions including self-weight, ice (where

    applicable), silt, earth pressure, and the

    maximum normal reservoir level with

    appropriate uplift pressures and tailwater level.

    (D + H +1 + S + U)

    (b) Unusual (Flood) Load Case

    Permanent and operating loads of the Usual

    Load Case, except for ice loading, should be

    considered in conjunction with reservoir and

    tailwater levels and uplift resulting from the

    passage of the IDF.

    (D + H' + S + UF)

    where subscript "F" refers to the flood case.

    The potential should also be assessed for

    reservoir levels higher than would result from

    passage of the IDF, such as those due to

    operating failures or other unusual conditions.

    The effects of ice loads should not be

    considered simultaneously with flood

    conditions.

    (c) Unusual (Earthquake) Load Case

    Permanent and operating loads of the usual

    load case except for ice loading, should be

    considered in conjunction with earthquake

    loading associated with the Operating Basis

    Earthquake (OBE). The effect of ice loads

    should not be considered simultaneously with

    OBE earthquake conditions.

    The analysis should be carried out for the dam

    empty case.

    (d) Extreme Load Case

    Permanent and operating loads of the Usual

    Load Case should be considered in

    conunction with seismic loads of the

    Maximum Design Earthquake (MDE).

    (D + H + S + Q + U)

    The effects of ice should be given special

    consideration, recognising the high uncertainty

    associated with ice loading on earthquake

    loading, and its effects on the dam.

    (e) Other Load Cases

    Where earthquake-induced cracking at the

    concrete-rock interface or any weak section is

    identified, a stability analysis should be

    carried out to assess whether the dam in its

    post-earthquake condition is capable of

    resisting loads of the Usual Load Case.

    84

    ANCOLD Guidelines for Design ofDams for Earthquake

  • 8/11/2019 Guidelines for Design of Dams for Earthquake

    92/127

    (D + H + S + UpQ)

    where subscript "pg" refers to the post-

    earthquake case.

    Concurrent ice loading (with the post-

    earthquake condition) may be considered in

    areas where appropriate.

    A landslide generated wave case should be

    considered where existing or potential

    landslides, which may affect the reservoir,

    have been identified. Combinations with other

    loads would be site specific.

    An inoperative drain case assuming plugged

    drains may be assessed and could be

    considered as an Unusual Load Case.

    7.5.3 Acceptance Criteria

    All kinematically feasible failure modes,

    analysed by the single-slice, rigid-body, force

    equilibrium method, should satisfy the

    acceptance criteria shown in Table 18.

    Table 18

    Stability Index Acceptance Criteria

    Load

    Sliding Factor

    (Note 1)

    Position of Resultant Force

    (Note 2)

    Minimum Compressive

    Stress Factor

    (Note 3)

    Usual

    1.5-2.0

    Mid-third of surface

    No tension

    4.0

    Unusual (Flood)

    and

    Unusual (OBE )

    1.3-1.5

    Mid-half of surface

    One-quarter tension

    2.7

    Extreme (MDE)

    1.1 - 1.3

    Within surface

    1.3

    Post-earthquake

    1.2-1.4

    Mid-half of surface

    One-quarter tension

    2.7

    Notes: 1. Sliding Factor (Frictional Analysis) = Resisting Forces

    Applied Force

    Lower values of the range apply where the geology and the strength parameters are

    reasonably well known.

    2. Vector summation of all forces, including uplift, acting on the analysis surface

    3. Compressive Stress Factor = Unconfined Compressive Strength

    Compressive Stress Normal to Surface

    To be considered primarily for massive but low strength rock and weak deteriorated

    concrete.

    7.5.4 Post Earthquake Stability

    If a dam is likely to be severely damaged after

    being subjected to the MDE, considerable time

    may elapse before the dam can be repaired or

    the storage lowered. Consequently, all parts of

    the dam will need to remain stable after an

    extreme earthquake event. The stability of the

    dam should therefore be checked for static

    loading conditions. The assumed uplift

    pressure distribution should be as discussed in

    sub-section 7.4.3, i.e. full headwater pressure

    within cracks emanating from the upstream

    face.

    7.5.5 Foundation Stability

    Where there is the possibility of a sliding

    failure along faults, shears and/or joints, the

    stability of the foundations should be

    ANCOLD Guidelines for Design of Dams for Earthquake

    85

  • 8/11/2019 Guidelines for Design of Dams for Earthquake

    93/127

    examined. This examination should be made

    for both the earthquake and post-earthquake

    cases. The dam itself should be examined for

    local overstressing due to foundation

    deficiencies.

    Sliding failure is especially likely when

    discontinuities and/or horizontal or sub-

    horizontal seams close to the foundation

    surface contain clay, or have been previously

    sheared eg. bedding surface shears due to

    stress relief, folding, or associated with faults.

    7.6 Dynamic Material Properties

    7.6.1 Concrete

    The compressive and tensile strengths of

    concrete increase with increased rate of

    loading. The dynamic compressive and tensile

    strengths of concrete can therefore be expected

    to be greater than the static strengths. As

    dynamic compressive stresses are rarely of

    concern, the allowable compressive stress for

    static loading can be used also for dynamic

    loading.

    Raphael (1984) states that the apparent tensile

    strength of concrete under seismic loading

    which should be used with linear finite

    element analyses is given by:

    fr =065 fc 2P

    where fc is the concrete compressive strength

    in MPa and fr is the apparent seismic tensile

    strength in MPa. Values given by this formula

    are some 50% greater than the apparent tensile

    strength for static loading. Raphael suggests

    that fr is twice the splitting strength of the

    concrete under static loading.

    C lough and Ghanaat (19 93) suggest that the

    apparent dynamic tensile strength is about

    25% greater than the measured static value

    which gives apparent tensile strength about

    20% of the standard compressive strength.

    They further suggest that there may be a 15 to

    20% loss of strength across lift joints. These

    figures may be even lower for poorly

    constructed or defective lift surfaces,

    fiowever, the peak dynamic tensile stresses

    only exist during a fraction of a response

    cycle. Even though these peak stresses may

    greatly exceed the tensile strength of the

    concrete, any cracking that might be initiated

    will not have time to fully develop. It is well

    recognised that a single spike of excessive

    localised tension should not be taken to

    represent dam failure.

    In consideration of the above however, these

    guidelines recommend that for sound lift

    surfaces, the apparent tensile strength to be

    used is 16% of the standard compressive

    strength.

    For dynamic modulus of elasticity, Clough and

    Ghanaat (1993) suggest a value 25% greater

    than the static value and these guidelines

    recommend this be adopted. For existing

    dams, the elastic modulus of the concrete mass

    may be determined using geophysical means

    (e.g. derived from measured shear wave

    velocity). Values obtained should be

    compared with static and dynamic small

    sample laboratory test values for credibility.

    7.6.2 Rock

    In most cases the stability of the dam will be

    controlled by sliding in or on the dam rock

    foundation. To carry out static and dynamic

    analyses, it will be necessary to:

    assess and map surface exposure,

    excavations and drill core to define the

    geology of the site. In particular the 3-

    dimensional orientation, continuity and

    detailed nature of bedding, joints, and other

    features such as shears are required.

    the shear strength of the foundation rock

    should be determined using appropriate

    rock mechanics techniques, such as those

    described in Hoek (1983, 1990, 1994),

    Hoek and Brown (1980), Patton (1966),

    Barton and Choubey (1977), Barton and

    Bandis (1991). These require an

    assessment of the orientation, spacing,

    continuity, shape and roughness of

    discontinuities in the rock (e.g. joints), and

    the strength of the rock substance as this

    varies with confining stress. Care should

    be taken in applying these techniques, to

    account for the presence of continuous,

    adversely oriented low strength surfaces

    such as bedding surface shears, faults or

    shears, and to take account of the

    mechanisms of failure.

    The Hoek, and Hoek and Brown methods give

    strengths for "undisturbed rock" and

    "disturbed rock". The latter should generally

    be adopted unless advice from an experienced

    86

    ANCOLD Guidelines for Design of Dams for Earthquake

  • 8/11/2019 Guidelines for Design of Dams for Earthquake

    94/127

    rock mechanics specialist indicates otherwise,

    because of the uncertainty as to the accuracy

    of the Hoek methods, and since the

    undisturbed rock strengths apply to confined

    conditions such as underground openings,

    where dilation or shearing causes increases in

    normal stress.

    It will be noted these stress methods give non

    linear failure envelopes, with high friction

    angle and low cohesion at the normal stresses

    usually applicable to dams. They also allow

    estimation of the modulus of the rock mass

    (Hoek, 1994). The dynamic modulus may be

    higher than the static modulus as discussed in

    Clough and Ghanaat (1993) and Scott and Von

    Thun (1993) and may, as for concrete, be

    obtained by geophysical means, or by relation

    to the static modulus.

    A dam's foundations will normally contain

    joints, shears, and bedding. Consequently, it

    will not be possible to transmit tensile stress

    within the foundations and the allowable

    tensile strength for the foundations will

    therefore normally be assumed to be zero.

    However, if extensive site investigation and

    strength testing is able to prove that the

    foundations for a particular dam site are

    capable of transferring tensile stresses, then

    the tensile strength of the rock may be

    included.

    Where foundation rock strength becomes

    critical, as they often will, advice should be

    obtained from a person expert in rock

    mechanics.

    8. APPURTENANT

    STRUCTURES

    8.1 Introduction

    A number of subsidiary structures associated

    with a dam are essential for the dams

    operation. Consequently damage to or

    destruction of these appurtenant structures

    would be prejudicial to the dam's safety. An

    important facility at a dam is one that allows

    water to be released in a controlled manner. If

    therie has been an earthquake and the dam is

    damaged to the extent that the dam is not

    serviceable then it may be necessary to lower

    the storage so that remedial works can be

    undertaken. It will therefore be necessary that

    not only the outlet structures and their gates

    and valves remain serviceable but also that

    there is proper access to these structures.

    Bridges and roads may need to remain in a

    sound state after an earthquake depending on

    their importance.

    Generally, appurtenant structures should be

    such that:

    they maintain their normal operating

    condition after an operating basis

    earthquake

    they are not damaged to an extent where

    they could allow sudden or uncontrolled

    loss of water from the storage for a more

    extreme earthquake up to the maximum

    design earthquake.

    8.2 Performance Requirements

    This sub-section gives the performance criteria

    for the operating basis earthquake (OBE) and

    the maximum design earthquake (MDE) which

    could be the maximum credible earthquake

    (MCE). Performance requirements are given

    for a number of appurtenant structures

    including intake towers, outlet conduits, outlet

    works, spillway gates, spillway piers, spillway

    gate hoist piers, access bridges and piers,

    access roads.

    Intake Tower

    OBE: Static and dynamic loads to

    induce maximum concrete and

    steel reinforcement stresses

    which satisfy AS3600

    (Concrete Structures Code)

    (i.e. limited amount of

    reinforcement yielding) and

    the tower and its base remain

    stable.

    MDE: Significant amount of

    reinforcement can yield

    horizontal reinforcement

    designed to prevent vertical

    reinforcement from buckling

    and to contain concrete when

    it is in compression (i.e.

    concrete contained between

    inner and outer layers of

    vertical reinforcement will not

    spall away).

    Outlet Conduit

    ANCOLD Guidelines for Design ofDams for Earthquake

    87

  • 8/11/2019 Guidelines for Design of Dams for Earthquake

    95/127

    O B E :

    MDE:

    Outlet Works

    O B E :

    Static and dynamic loads to

    induce maximum concrete and

    steel reinforcement stresses

    which satisfy AS3600

    (Concrete Structures Code).

    Dynamic loads to include

    those induced from the

    earthquake effect on the

    overlying dam.

    Conduit not to collapse or

    rupture - Collapse could lead

    to undermining and

    subsequent failure of an

    overlying embankment.

    Rupture could cause piping or

    destabilise an embankment by

    a marked increase in pore

    pressure.

    All valves to maintain their

    normal operating capabilities.

    MDE: Emergency closure and

    regulating valves (especially

    low level release valves) to

    maintain operating capability -

    storage may have to be

    quickly lowered if parts of the

    dam are damaged and need

    remedial works or relief of

    hydrostatic loads.

    Spillway Gates

    OBE: Gates to maintain normal

    operating capability.

    MDE: Gates retaining permanent

    storage at the time of an MDE

    should not fail to the extent

    where water from the storage

    is released in an uncontrolled

    manner. MDE should not

    cause the gates to distort to an

    extent that they cannot be

    opened or closed.

    S pillway Piers

    O B E :

    Carry out appropriate dynamic

    analysis for the spillway piers

    for earthquake loading in the

    upstream/downstream and

    transverse directions.

    Combined static and dynamic

    loads should satisfy ASS 600

    (Concrete Structures Code).

    Earthquake loads from the

    orthogonal directions may be

    combined on a square root of

    the sum of the squares basis.

    MDE: Carry out dynamic analysis as

    for the OBE. Combined static

    and dynamic loads may cause

    cracking but piers must

    remain stable for overturning

    and sliding. Piers should not

    be permanently displaced to

    the extent where the spillway

    gates become jammed.

    Spillway Gate Hoist Piers

    OBE: Carry out appropriate dynamic

    analysis similar to that for

    spillway piers. Combined

    static and dynamic loads

    should satisfy AS3600

    (Concrete Structures Code).

    MDE: If the gates can be operated

    from an alternative position

    (possibly, in a less efficient

    manner), then the spillway

    gate hoist piers (and hoist

    bridge) can be allowed to fail.

    If the continuing operation of

    the gates depends on the

    continued viability of the hoist

    bridge piers (and hoist bridge)

    then carry out an appropriate

    dynamic analysis similar to

    that for the spillway piers.

    Combined static and dynamic

    loads may cause cracking but

    the piers must remain stable

    for overturning and sliding.

    Note: theappropriate

    dynamic analysis

    required for spillway

    piers and spillway

    hoist piers may be part

    of an overall dynamic

    analysis for a concrete

    dam.

    Access Bridge arid Piers

    OBE: Carry out appropriate dynamic

    / of the pier and bridge

    'The anaysis may be

    88

    AN CO LD Guidelines for Design

    Wke'

    *4

  • 8/11/2019 Guidelines for Design of Dams for Earthquake

    96/127

    part of an overall dynamic

    analysis for an intake tower.

    The pier and bridge system

    should be examined for an

    earthquake in the direction of

    the bridge access and

    perpendicular to the bridge

    access. Earthquake loads may

    be combined on a square root of

    the sum of the squares basis.

    Combined static and dynamic

    loads should satisfy the

    appropriate Structures Code).

    MDE: If there are alternative means

    of access then the access bridge

    can be allowed to fail. If there

    are no alternative means of

    access then the bridge and piers

    should remain capable of

    carrying their design loads.

    Access Roads

    OBE: Access roads to the dam and

    its appurtenant works should

    remain passable immediately

    after the OBE. Therefore any

    likely land slip areas along the

    access roads should be checked

    for stability. There should be no

    land slips either during or after

    an OBE.

    MDE: Access roads to the dam and

    its appurtenant works may

    become impassable during or

    immediately after an MDE.

    However, they should be easily

    cleared. Therefore, while there

    may be land slips onto the

    roads, the roads themselves,

    should not be allowed to

    collapse where there is no easy,

    alternative access route.

    8.3 Intake Towers

    8.3.1 Analysis

    The analysis method for intake towers is

    described here in general terms only. A more

    complete description of the method on which

    the following general description is based, is

    given by Chopra and Goyal (199 1) and Goyal

    and Chopra (1989).

    The method presented here is based on a

    simplified dynamic analysis in the frequency

    domain using a suitable response spectrum. It

    uses an added mass representation of

    hydrodynamic effects due to surrounding

    (outside) water and contained (inside) water

    (in the case of wet towers). In addition, it

    includes the effects of tower/foundation

    interaction.

    The steps of the method are:

    (i) Select suitable response spectrum.

    (ii) Compute the added hydrodynamic

    mass of water using Goyal and Chopra

    (1989).

    (in) Determine the structural properties of

    the tower

    mass per unit height

    flexural stiffness

    modal damping ratios.

    (iv) Compute natural periods and mode

    shapes for the first two modes of

    vibration.

    (v) Determine the spectral accelerations

    for the first two modes of vibration

    from the response spectrum.

    (vi) Compute the generalised mass and

    generalised excitation terms using the

    mass distribution and the mode

    shapes.

    (vii) Compute the inertia forces using the

    spectral accelerations, generalised

    mass and excitation terms, mass

    distribution and mode shapes.

    (viii) Add the inertia loads from the first

    two vibration modes on a square root

    of the sum of the squares basis.

    (ix) Compute bending moments and shear

    forces from (viii).

    (x) Design tower according to AS3600

    (Concrete Structures Code).

    8.3.2 Design Criteria

    As discussed in Sub-section 8.2, an intake

    tower is designed elastically for the OBE.

    Generally, it will not be necessary for the

    ANCOLD Guidelines for Design of Dams for Earthquake

    89

  • 8/11/2019 Guidelines for Design of Dams for Earthquake

    97/127

    tower to behave elastically during the MDE.

    In order to ensure that the tower can survive

    intense ground shaking due to the MDE with

    limited damage, it should possess a ductility

    capacity greater than the ductility

    requirements imposed by the ground motion.

    A suitable method for this is described in

    Chopra and Liaw (1975) where a ductility

    factor of two is recommended (ratio of

    maximum permissible displacement to the

    yield displacement).

    /

    90

    ANCOLD Guidelines for Design of Dams for Earthquake

  • 8/11/2019 Guidelines for Design of Dams for Earthquake

    98/127

    R E F E R E N C E S

    /

    ANCOLD Guidelines for Design of Dams for Earthquake

  • 8/11/2019 Guidelines for Design of Dams for Earthquake

    99/127

    REFERENCES

    Ambraseys, N.N. (1973). Dynamics and response of foundation materials in epicentral regions of strong

    earthquake&Proc. 5th World Conference in Earthquake Engineering, Rome, Italy.

    ANCOLD (1991). Guidel ines on Design Criteria for Concrete Gravity Dams.

    ANCOLD(1994). Guidelines on Risk Assessment.

    ANCOLD (1996). Commentary on ANCOLD Guidelines on Risk Assessment (in preparation).

    Argyris, J.A., Krempl, E. and William, K.J. (1977). Constitutive Models and Finite Element Solution of

    Inelastic Behaviour, in Formation and Algorithm in Finite Element Analysis (eds. K.J. Bathe et al).

    Barton, N. and Bandis, S. (1991). Review of predictive capabilities of JRC-JCS model in engineering

    practice. Pub. No. 182, Norwegian GeotechnicalInstitute, Oslo, Norway, 8p.

    Barton, N. and Choubey, V (1977). The shear strength of rock joints in theory and practice. Rock

    Mechanics 10,1-54.

    Bazant, Z.P. and Ob, B.M. (1979). Blunt crack band propagation in finite element analysis.

    J.Eng.Mech. Div., ASCE, 105, EM2,297-315.

    Bazanl, Z.P. and Cedolin, L. (1981). Concrete fracture via stress strain relations. P.I. Theory, P.2

    Verification. Report 81-10/655c, Centre of Concrete and Geomechanics, Northwestern University,

    Evanston, Illinois.

    BC Hydro (1993). Guidelines for Consequence Based Dam Safety Evaluations and Improvements. BC

    Hydro, Report No. H252B, August 1993.

    BC Hydro (1995). Guidelines for the Assessment of Rock Foundations of Existing Concrete Gravity

    Dams. BC Hydro, ReportNo. MEP67, Vancouver, Canada.

    Bicanic, N. and Zienkiewicz; O.C. (1983). Constitutive model for concrete under dynamic loading.

    J.EarthquakeEng. and Struct Dyn, Vol 11,689-710.

    Bierschwale, J.G. and Stokoe, K.H. (1984). Analytical evaluation of sands subjected to the 1981

    Westmorland earthquake. Geotechnical Engineering Report GR-84-15, Civil Engineering

    Department, University of Texas.

    Brown, R.E. (1977). Vibroflotationcompaction of cohesionless soils. J.Geotech. Eng. Div., ASCE, Vol

    103,NoGT12,1437-1451.

    Campanella, R., Robertson, P.K. and Gillespie^ D. (1986). Seismic cone penetration test, in Use of

    In-Situ Tests in Geotechnical Engineering. ASCE Geotechnical Special PublicationNo. 6.

    Canadian Dam Safety Association (1994). Draft Dam Safety Guidelines for Existing Dams. Canadian

    Dam Safety Association, April 1994.

    Castro, G. (1976). Comments on seismic stability evaluation of embankment dams. Proc. Conference

    on Evaluation of Dam Safety.

    Cedolin, L., Crutzen, Y.R.J, and Dei Poli, S. (1977). Stress-strain relationship and ultimate strength of

    concrete under triaxial loading conditions. Jnl. ofEng.Mechs. Div., Vol 103, No EM3,423-440.

    Cervera, M., Oliver, J. and Calindo, M. (1992). Numerical analysis of dams with extensive cracking

    resulting from concrete hydration; Simulation of a real case. Dam Engineering, Vol III, Issue 3,

    February.

    Chakrabarti, P. and Chopra, A.K. (1974). Hydrodynamic effects in earthquake response of gravity

    dams. J.StructuralEng., ASCE, 100, ST6,1211-1224.

    Chapuis, T., Rebora, B. and Zimmerman, T. (1985). Numerical approach to crack propagation analysis

    in gravity dams during earthquakes. Proc. 15th Int. Congress on Large Dams, Lausanne.

    Chavez, J.W