Collapse Mechanisms at The Foundation Interface of Geometrically Similar Concrete Gravity Dams

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Engineering Structures 32 (2010) 1304–1311

Contents lists available at ScienceDirect

Engineering Structures
journal homepage: www.elsevier.com/locate/engstruct

Collapse mechanisms at the foundation interface of geometrically similar


concrete gravity dams
Gabriella Bolzon ∗
Department of Structural Engineering, Politecnico di Milano, piazza Leonardo da Vinci 32, 20133 Milano, Italy

article info abstract


Article history: A computationally effective numerical procedure is introduced to evaluate safety factors and collapse
Received 31 August 2009 mechanisms of geometrically similar structures with dilatant frictional joints like concrete gravity dams.
Received in revised form The envisaged methodology, alternative to the more traditional step-by-step integration of the governing
11 January 2010
rate equations, rests on the solution of a constrained minimization problem endowed with a closed-form
Accepted 14 January 2010
Available online 10 February 2010
algebraic formulation, under the hypothesis that damage evolves along structural joints while the bulk
material remains linear elastic.
Keywords:
© 2010 Elsevier Ltd. All rights reserved.
Structural failure
Collapse mechanisms
Concrete gravity dams
Frictional joints
Mathematical programming

1. Introduction some decimeters (for centrifuge experiments) to a few meters, ver-


sus full-scale dimensions of dozens of meters; see, e.g. [16–20].
The overall structural strength of concrete dams is controlled by Geometrical similitude can be recovered also in real scale applica-
the presence of natural or artificial discontinuity surfaces like con- tions for some dam typology [2]. However, simple energy consid-
struction joints, foundation and abutment interfaces, cracks, which erations show that scaling of the structural response is non-linear,
significantly affect the load-carrying capacity and the failure mode in this context, due to the different spatial dimension of the loca-
of these large structures. The experimental determination of mean- tions where reversible and irreversible phenomena take place; in
ingful mechanical characteristics of the relevant joints is therefore particular, elastic strain energy is stored in the bulk material while
particular desirable for the purpose of construction, retrofitting or dissipation occurs along the joint surfaces [21,22].
repair. At the present time, however, the evaluation of these pa-
An effective description of the structural response of concrete
rameters is rather troublesome and parametric studies are com-
dams rests on the assumption of linear elastic behavior outside the
monly carried out in this engineering field to account for large
discontinuity surfaces like joints and cracks, which are endowed
uncertainties [1–4].
with softening constitutive laws relating tractions to displacement
Mechanical parameters like fracture energy, cohesion and fric-
tion along sliding planes play an important role in dam engineering discontinuities, in the spirit of the discrete cohesive crack approach
as well as in geomechanics, soil-structure interaction problems, for introduced by Hillerborg [23,24]. Quite interestingly, the cohesive
slope stability analysis and landslide control, where local variabil- crack approach can also be interpreted from micromechanical
ity, hydrometric conditions and climate changes suggest statisti- perspectives [25].
cal approaches based on recursive computations; see, e.g. [5–10]. The discrete crack approach, together with the simplifying but
Recursive calculations are also required by parameter identifica- reasonable, in this context, assumption of proportional loading,
tion procedures based on laboratory testing and inverse analy- monotonically increasing with the structural degradation up to
sis [11–15]. failure, leads to a convenient numerical procedure, different to the
The experimental testing of dam structures with their re- step-by-step integration of the governing rate equations, for the
lated joints, if available, is necessarily performed on scale mod- evaluation of safety factors and failure modes of quasi-brittle ge-
els, the main dimensions of laboratory specimens ranging from ometrically similar structures with dilatant frictional joints at re-
duced costs.
This approach has been exploited in [26] considering pure
∗ Tel.: +39 02 2399 4319; fax: +39 02 2399 4220. opening fracture mode for brittle interfaces, where the actual
E-mail address: gabriella.bolzon@polimi.it. shape of the cohesive law is immaterial. In the hypothesis that the
0141-0296/$ – see front matter © 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.engstruct.2010.01.008
G. Bolzon / Engineering Structures 32 (2010) 1304–1311 1305

onset of crack propagation corresponds to structural failure (un-


stable crack propagation), with negligible friction effects, the only
important parameter is the specific fracture energy. These assump-
tions permit one to derive semi-analytical solutions, useful also for
comparison purposes.
In the present paper, the methodology is exploited to under-
stand the role of mixed-mode fracture on the overall response of
concrete gravity dams. No a priori assumption is made about the
failure mode, which is returned by the computations.

2. Theoretical background

2.1. Cohesive crack models for dilatant frictional joints


a
Frictional contact and quasi-brittle fracture are peculiar aspects
of the mechanical response of concrete and rock joints that can
be described (both in total terms and in rates) by softening in-
terface laws relating tractions to displacement discontinuities.
The rate problem formulation and time-integration procedures
are required to accurately describe history-dependent irreversible
phenomena. Still, path-independent (holonomic, fictitiously re-
versible) models are admissible under proportional external ac-
tions, leading to failure with monotonically increasing material
degradation, as hypothesized by the classical ‘‘deformation theory’’
of plasticity [27–31]. This is the reference context of the present
work.
The mechanical contribution of the joint to the overall struc-
tural behavior can be described starting from the components tn
and tt of the interface traction vector t, in normal and tangential
direction, respectively. Let tnu and ttu represent the initial strength
of the undamaged joint in pure tensile and pure shear mode. Inter-
facial resistance is supposed to be progressively reduced by damag-
ing phenomena such as wear and cracking, so that t̄n (0 ≤ t̄n ≤ tnu )
and t̄t (0 ≤ t̄t ≤ ttu ), collected by vector t̄, represent the current
strength values, tnu and ttu being their upper limits. Shear strength
t̄t is supposed to be increased or decreased by friction, depending
on the sign of the normal stress tn , here assumed to be positive in
tension. Hence, interface tractions must satisfy the restrictions:

− t̄t − µtn ≤ tt ≤ t̄t − µtn



tn ≤ t̄n ; (1)
where µ represents the friction coefficient, here assumed to be a
constant. This restriction could be removed to account for different
frictional effects occurring under static and kinetic, or dry and wet,
conditions without substantial changes of the mathematical struc-
ture of the present problem; see e.g. [30]. The coefficient µ might
also depend on internal variables, e.g. for cycling loading, consid-
b
ered in [11,12] but ruled out from the present context, where exter- Fig. 1. Graphical representation of an exponential interface model.
nal actions are supposed to increase monotonically up to structural
failure. The interface degradation is described through some softening
The initial strength domain defined by relationships (1) as function t̄ (λ), as in [30,31]. Among the possible choices, the
t̄n = tnu , t̄t = ttu , is represented by the thick lines of Fig. 1(a). This exponential decay illustrated in Fig. 1(b) will be considered in
domain evolves, with the interface degradation, toward the pure the following application; functions t̄ (λ) are therefore analytically
Coulomb’s frictional limit situation represented by dashed lines; defined by:
an intermediate state is visualized by dash-dotted lines.
t̄n = tnu e−αλn −γ (λt +λ−t )

Interface degradation can be related to the total amount of
(3)
unrecoverable opening and sliding displacements wn and wt , t̄t = ttu e−δλn −β(λt +λ−t )
which can be expressed in terms of the non-negative kinematics
‘‘internal’’ variables λn and λt (or λ−t when reversing the sliding where parameters α and β can be related to the specific fracture
direction) collected by vector λ, as follows: energies GI = tnu /α and GII = ttu /β , graphically represented by the
dashed areas in Fig. 1(b). Parameters α and β can be alternatively
wn = λn + µ̃ (λt + λ−t ) related to the strength decay rate; in particular, for uncoupled

wt = λt − λ−t
(2) degradation (i.e., for γ = δ = 0), the residual strengths t̄n , t̄t equal
5% their original tnu , ttu values when λn = 3/α , λt = 3/β . These
where µ̃ denotes the dilatancy associated to the joint roughness; quantities can be derived from specific experiments carried out
see, e.g., [11] and references therein. on concrete to rock block assemblies; see e.g. [12]. However, real
1306 G. Bolzon / Engineering Structures 32 (2010) 1304–1311

Relations (1) can be expressed in matrix form, more suited to


computations, as follows:

1 0   1 0  
" # " #
t̄n tn
0 1 − µ 1 = aT t̄ − cT t ≥ 0 (5)
t̄t tt
0 1 µ −1
where, clearly:

µ µ
   
1 0 0 1
a≡ ; c≡ . (6)
0 1 1 0 1 −1
Analogously, relation (2) reads:
 ( λn )
wn µ̃ µ̃
  
1
w≡ = λt = c̃λ (7)
wt 0 1 −1
λ −t
with evident definition of matrix c̃.
Independently of the analytical expression of functions t̄ (λ),
the relationships governing the assumed interface behavior can
be cast in the following compact format of a complementarity
problem (CP), which entails orthogonality between the two sign-
constrained vectors, λ and ϕ:

ϕ ≡ aT t̄(λ) − cT t ≥ 0 λ ≥ 0 ϕT λ = 0. (8)
The CP (8) implies that relations (1) and, hence, (5) hold true as
strict inequalities as λ = 0 and t̄n = tnu , t̄t = ttu ; correspondingly,
Fig. 2. Benchmark problem [26,32].
no irreversible displacement wn or wt develops at the interface in
the normal and tangential direction, respectively; see (2) or (7).
joint conditions are rather difficult to reproduce in laboratory and As any component λi (i = n, t , −t ) is positive, then the relevant
large uncertainties are expected. Parametric studies are therefore traction component ti equals the interfacial strength t̄i . Notice also
of paramount importance in most related applications. that CP (8) implies λ−t = 0 if λt 6= 0, and vice-versa, when tt 6= 0.
Relationships (8) return the actual description of the joint be-
2.2. Structural response haviour when the internal variables λi increase monotonically in
time, so that the structural response to external actions corre-
Linear elasticity is assumed to describe the bulk material re- sponds to progressive damage growth without local unloading.
sponse, outside the joint location, in the hypothesis of small strains This holonomy assumption, which implies a fictitiously reversible
and displacements, well justified in the quasi-brittle context to be system response, represents a reasonable simplification of the real
investigated herein. This linear background permits one to exploit problem when effectiveness in terms of computing time and cost
the superposition principle, in particular to express the relation- is a priority, whereas accuracy can be sacrificed to some extent.
ship between traction and displacement discontinuity distribution Comparison between the results of previous holonomic and non-
at the interface 0 , i.e. between the vector functions t(ξ) and w(ξ0 ), holonomic (rate-dependent) analyses [31,35] showed that reliable
respectively, by the integral equations: results in terms of overall load carrying capacity and failure mode
Z are usually gathered under the assumption of reversibility even if
local effects are not always well captured.
X
t(ξ) = sk tEk (ξ) + z(ξ, ξ0 )w(ξ0 )dξ0 (4)
k 0
3.1. Discretised problem
where: ξ and ξ0 are coordinate systems spanning the interface 0
(see e.g. Fig. 2); vectors tEk (ξ) denote elastic tractions generated at The primary unknowns of the present problem are the traction
the interface by unit external actions in the absence of displace- and the displacement discontinuity fields, t(ξ) and w(ξ), defined
ment discontinuities, i.e. in a fictitious elastic regime; scalars sk over the interface 0 . These fields can be approximated by the in-
indicate the amplification factor of each external action; matrix terpolation, through the functions collected by matrices Mw (ξ),
z(ξ, ξ0 ) collects the set of the influence functions pertaining to a Mt (ξ), of the corresponding discretisation (usually nodal) vari-
distribution of distortions assigned along 0 in the fictitious elastic ables, say T and W, as common in FE and BE approaches [31,33,
body. 36,37,34]:
Relations (1)–(4) define the dependence of the traction vector t
on the displacement discontinuities w and, hence, on the here as- w(ξ) = Mw (ξ)W, t(ξ) = Mt (ξ)T. (9)
sumed internal variables λ. The final aim is to determine the max- Interpolation functions collected by matrices Mw (ξ) and Mt (ξ)
imum amplification factor sk associated to variable loads (e.g., the can be profitably defined over 0 in interrelated way; for instance,
maximum water level in the dam reservoir) compatible with the following [37]:
structural strength, and the corresponding failure mode. Z
MTw Mt dξ = I (10)
3. Computing issues 0
where I represents the identity matrix of order (nD · nN )×(nD · nN ),
The discrete crack approach can be profitably employed in nu- nD being the spatial dimension of the problem (nD = 1, 2, 3) and nN
merical analyses based on finite element (FE) or boundary element the number of the discretisation points (nodes) along the selected
(BE) modeling [33,34], as follows. interface.
G. Bolzon / Engineering Structures 32 (2010) 1304–1311 1307

Relations (9) and (10) show that variables T coincide with of the bulk material, account taken of the compatibility relation-
equivalent nodal forces since: ship (7), which can also be re-written as:
Z Z
T= MTw (ξ)t(ξ)dξ. (11) W = C̃3, C̃ = MTt (ξ)c̃Mw (ξ)dξ (19)
0 0
With these provisions, the discrete (weak) counterparts of the to return:
integral equations (4) read: X
X 8 = AT T̄(Λ) − CT ZC̃3 − sk CT TEk ≥ 0 3 ≥ 0
T= sk TEk + ZW (12) k (20)
k 8T 3 = 0.
where T and TEk are nodal forces, in accordance with (10) and (11), Notice that, generally, matrix CT ZC̃ is neither symmetric nor sign
while the influence matrix Z, symmetric and semi-definite in sign, definite.
is formally obtained as
Z Z 3.2. Solution strategies
Z= MTw (ξ)z(ξ, ξ0 )Mw (ξ0 )dξdξ0 . (13)
0 0
The CP (20) can be solved for any given set of load multipli-
In a standard displacement-based FE context, matrix Z can be ers sk by dedicated software; see, e.g. [29] and references therein.
simply derived from the usual stiffness matrix K by condensation Usually, some factors sk in relations (20) are fixed (e.g., those rel-
of all the out-of-0 discretisation variables; see [36,37]. evant to self-weight) while variable loads are amplified by some
It is worth recalling that: common term, say: sk = ηŝk , with given ŝk . The maximum sustain-
(i) due to the assumed linear elastic background, matrix K can be able load can then be sought by increasing η in repeated analyses,
formed, partitioned and inverted once for all at the beginning with the maximum allowable η multiplier singled out by lack of
of the analysis to derive the corresponding Z matrix; solution.
More consistently, the maximum η-value (say ηopt ) and the cor-
(ii) matrix K, and therefore Z, are independent of the actual dimen-
responding deformation mode are the result of the following con-
sions for self-similar and similarly discretised structures; strained optimization problem (OP):
(iii) in self-similar structures, vectors TE are proportional to DnL ,
 
where D represent the characteristic structural dimension; see
ηopt = min −η s.t. : 8(η, 3) ≥ 0, 3 ≥ 0, 8(η, 3)T 3 = 0 (21)

e.g. Fig. 2 and nL = 1, 2, 3 for line, surface and volume loads, η,3
respectively.
with the expression of 8(η, 3) derived from (20).
As local equilibrium equations are replaced by their weak form, The optimization problem (21) has the format of a so-called
the interface constitutive law can be reformulated in global terms mathematical program with equilibrium (here, complementarity)
as well, using the same weight and shape functions introduced constraint (MPEC). Details on MPECs and on the relevant solving
before for static and kinematics variables, as follows: procedures can be found, for example, in [38] and in the references
therein.
ϕ(ξ) = MTt (ξ)8, λ(ξ) = MTw (ξ)3. (14)
One possible approach consists of transforming the MPEC (21)
Then, due to the orthogonality condition (10): into a more standard constrained OP by relaxing the rather nasty
Z Z complementarity constraint, namely:
8= MTw (ξ)ϕ(ξ)dξ, 3= MTt (ξ)λ(ξ)dξ. (15)  
ηopt = min −η + κ 8(η, Λ)T 3 s.t. : 8(η, 3) ≥ 0, 3 ≥ 0 . (22)

0 0
η,3
Therefore, the discretised version of the CP (8) reads:
Factor κ introduced with (22) represents a penalty coefficient,
8 = AT T̄(3) − CT T ≥ 0 3 ≥ 0 8T 3 = 0 (16) iteratively adjusted (usually, progressively increased) to force the
non-negative complementarity term to reduce to zero as the
where matrices A and C collect constant terms:
Z Z optimum point is approached [29,35,39].
A multi-criteria optimization problem is defined as two or more
A= MTt (ξ)aMw (ξ)dξ, C= MTt (ξ)cMw (ξ)dξ. (17)
0 0 load factors vary independently one from the other [40]. This situ-
ation is frequently reduced to a standard OP by the combination of
A proper ordering of the nodal variables collected by vector 3 may
the different objectives into a single scalar fitness function, e.g. by
reduce A and C to block matrices, each block coinciding with the
a sum weighted with positive factors ωi :
corresponding local matrix a or c. This characteristic derives also
from the assumed orthogonality between Mw and Mt , relation (10). nv
X
Vector T̄(3) collects the discrete counterpart of the current ηeq = ωi ηi ωi > 0. (23)
interface strength functions t̄(λ), weighted according to (11) or, i=1
equivalently, (15), account taken of the interpolation (14): Optimal sets of ηi multipliers are to be considered for various
Z Z combinations of the weights ωi . Efficient formulations and opti-
T̄(3) = MTw t̄(λ)dξ = MTw t̄(Mw 3)dξ. (18) misation solvers are hence quite mandatory for this class of prob-
0 0 lems [41].
These integrals can often be given an explicit form; this is for in- The first-order sequential quadratic programming (SQP) algo-
stance the case of exponential decay and linear shape functions rithm available in a widely used optimisation toolbox [42] can be
considered for the present application, see [31]. efficiently exploited in the present context, as shown in the next
The final formulation of the structural problem in point is ob- section. SQP considers a sequence of quadratic approximation of
tained by the combination of the finite-dimensional CP (16), gov- the Lagrangian function that incorporates the original function to
erning the interface response, with the discretised form of the be minimised and its non-linear constraints. The solution is then
equilibrium equations (12), relevant to the linear elastic behavior sought by gradient-based techniques. Gradients are also exploited
1308 G. Bolzon / Engineering Structures 32 (2010) 1304–1311

to build up some approximation to the Hessian of the Lagrangian a


function (no second-order derivatives are evaluated).
Significant savings and increased computing efficiency can be
gained by the non-conventional approach followed in this paper
since the integral functions 8(η, 3) are given in explicit form and °
derivatives can be evaluated analytically.
The presence of multiple minimum points cannot be ruled
out in the present context. The problem is in fact non-convex by
its very nature, even in its rate formulation; see e.g. [39]. Local
and global minima can however be discerned by the following
procedure. Once one optimum value of the load multiplier, say η̂opt ,
has been obtained, check whether the CP (20) can return a solution
for the augmented load factor η = η̂opt + 1η. This computation
may be performed by the same software used for the OP (21) since
the CP (20) can be reformulated as:

min 8(Λ)T 3 s.t. : 8(η, 3) ≥ 0, 3 ≥ 0 .



(24)
b
3
°
°
If such a solution does not exist, than η̂opt is the sought absolute
maximum. °

4. Case study

The failure mechanism of frictional-cohesive structures can be


°
rather different depending on a few characteristic parameters, like
cohesion and friction angle, which can be hardly quantified with a
reliable confidence level. The approach considered in this paper is
flexible enough to capture the salient features of each situation at
reduced computing costs. Therefore, parametric studies like those
summarized in this section can be carried out quite easily.
The considered benchmark problem, originally proposed by the
Fig. 3. The maximum sustainable overtopping water level as a function of the
International Committee on Large Dams (ICOLD) [32], analyses a characteristic structural dimension D in the presence (a) and in the absence (b) of
typical section of a concrete gravity dam; see Fig. 2. Fracture at the uplift forces. Continuous lines show predictions obtained from either LEFM (curve
foundation interface is hypothesized, promoted by hydraulic pres- lines) for Mode I crack propagation or limit equilibrium state through rigid body
sure and resisted by the self-weight of the structure and by the consideration (straight lines). Numerical results are given for the following interface
properties: maximum shear strength ttu = 0.7 MPa and friction angle φ = 30°
interfacial strength, enhanced by friction. Both the dam concrete
(empty circles ◦), 35° (empty triangles 4) and 45° (empty diamonds ); ttu = 5 MPa
and the foundation rock are considered to be impervious. Uplift and φ = 45° (filled squares ).
pressure may act on the dam-foundation interface as detaching oc-
curs. Structural safety is quantified by the overtopping water level
sustainable by the structure in the hypothetical case of overflow. be eventually reflected by the entries of matrix Z only. Possible
Dynamic effects are not taken into account. This situation has al- discontinuity planes in the foundation rock could be taken into
ready been considered in [26], where an initial notch (1/10 of account as well, although not done in the present demonstrative
the interface length) has been introduced to the dam geometry example.
proposed by ICOLD [32] to allow comparison with linear elastic Self-similar structures of different characteristic size D are con-
fracture mechanics (LEFM) approaches, which cannot predict crack sidered, having in mind some dam typology [2] and scale models
initiation. Although this limitation is overcome by the cohesive– for laboratory testing. In the present approach, D affects the load-
crack approach pursued here, the notch is kept in the present study ing vectors TEk , while matrix Z remains unaltered. The intensity of
to permit comparison with already available reference results. all applied forces but overtopping water pressure and the corre-
Following [26], uplift forces are either neglected (which would sponding uplift, if any, is fixed. The maximum sustainable overtop-
be possible in engineering practice only in the presence of a perfect ping water level is then assumed as a measure of the overall load
drainage system) or uniformly distributed without any reduction carrying capacity of these similar structures.
of the hydrostatic value, as schematized in Fig. 2 for the initial
Mode I fracture characteristics of the foundation joint are those
notch.
already considered in [26,31], namely: tensile strength tnu = 1 MPa
Computations are based on a plane strain model (nD = 2, see
Section 3.1), which allows for a simpler representation of the and specific fracture energy GI = 90 J/m2 . The influence of the
results. In the model, concrete is assumed linear elastic, charac- remaining interface parameters on the overall structural strength
terized by Young’s modulus E = 24 GPa, Poisson’s ratio ν = 0.15 is the subject of a parametric study.
and weight density γc = 24 kN/m3 . The dam body is discretised Figs. 3–7 summarize the results of analyses performed by as-
by means of about 3000 4-node plane elements but most degrees suming shear strength tnu = 0.7 MPa and fracture energy GII =
of freedom (DOFs) are condensed in a reduced 110 × 110 stiffness 270 J/m2 (or, equivalently, β = 2 mm−1 ) as in ICOLD bench-
matrix (concerning 55 nodes × 2 DOFs), as explained, for example, mark [32], characterized by rather brittle interface behavior, or
in [36]. Rigid foundation is considered due to the lack of informa- tnu = 5 MPa and GII = 2500 J/m2 (maintaining β = 2 mm−1 ).
tion about the soil. However, the overall problem dimension and, Friction angle φ is given the values 30°, 35° or 45°. The coupling be-
hence, the computational burden would not change by a detailed tween interface damaging modes is neglected, setting γ = δ = 0,
representation of an even larger foundation volume, provided it to allow for an easier interpretation of the results; analogously,
were linear elastic, as often assumed, since soil stiffness would dilatancy µ̃ = 0.
G. Bolzon / Engineering Structures 32 (2010) 1304–1311 1309

Fig. 4. Distribution of opening/sliding interface displacement at failure for Fig. 6. Distribution of opening/sliding interface displacement at failure for
D = 40 m in the presence of uplift forces (ttu = 0.7 MPa). D = 40 m in the absence of uplift forces (ttu = 0.7 MPa).

Fig. 7. Distribution of opening/sliding interface displacement at failure for


Fig. 5. Distribution of opening/sliding interface displacement at failure for D = 240 m in the absence of uplift forces (ttu = 0.7 MPa).
D = 80 m in the presence of uplift forces (ttu = 0.7 MPa).

part of the problem solution, through the compatibility relation-


Results are given in terms of the maximum load multiplier η̂opt , ships (19). Notice that the graphical representation in Figs. 4–7 in-
which affects the overtopping water level, and of the correspond- cludes the initial notch (node 6 coincides with its tip), although the
ing failure mode described by the distribution of opening/sliding relevant DOFs were initially condensed.
displacements along the foundation interface 0 . Displacement dis- Figs. 3(a), 4 and 5 refer to the loading condition that admits
continuities are recovered from the 3 vector returned by (21) as uplift forces along 0 . For small and medium structural sizes, the
1310 G. Bolzon / Engineering Structures 32 (2010) 1304–1311

present results substantially conform to those calculated in the hy- cally similar structures with dilatant frictional joints. The proposed
pothesis of Mode I fracture, independently of the shear characteris- approach, alternative to the step-by-step integration of the gov-
tics of the interface, both for the maximum load multiplier and for erning rate equation, allows proper transfer of results from model
the failure mechanism; see [26]. In these situations, failure is met dimensions to real scale applications in the hypothesis that dam-
at the early stage of crack propagation and most relative displace- ages are concentrated along some natural or artificial joint while
ments between the structure and its foundation are confined to the the remaining structural material behaves as linear elastic. The
notch, where opening prevails; see Fig. 4. Then, the collapse load method has been illustrated with the aid of some computing ex-
can be inferred by semi-analytical methods based on LEFM [22], ercises, based on a benchmark problem proposed by ICOLD [32],
which actually rests on the implicit assumption of unstable frac- focusing on the structural collapse of a concrete gravity dam due
ture propagation soon after its initiation. to progressive crack propagation at the foundation interface. The
LEFM provides accurate prediction of the safety level of the con- presence of multiple, possibly simultaneously activated joints can
sidered structures for D ≤ 60 m. Above this characteristic di- be accounted for in the same way. Material separation is however
mension range, crack propagation is stabilized by the increasing constrained to occur within the a priori defined potential displace-
self-weight contribution; the water level has to be raised to let ment discontinuity surfaces. Crack path bifurcations can be pre-
cracks advance, and failure is observed as the foundation interface dicted, if the alternative propagation directions are included in the
is almost completely detached. In such a situation, LEFM leads to modeling, as done, for example, in [44]. The actual failure mode is
a dramatic underestimation of the maximum overtopping water then returned as a part of the problem solution.
level. The total execution time of each performed analysis in Matlab
The structural response is dominated by opening displacements [42] environment was of the order of a few seconds on a com-
also for D > 60 m, but ultimate failure is then associated with mon personal computer, the convergence rate of the iterative solu-
the consumption of local shear strength, with a slight dependence tion procedure mainly depending on the choice of the initialization
upon friction. Fig. 5 shows the displacement discontinuity distri- vector. The converged values of the previously performed analysis
bution for D = 80 m, the structural size selected for ICOLD bench- (though with different input) was often efficiently exploited to this
mark [32]: with the assumed input data set, the residual normal purpose.
and shear strength are reduced to less than 5% their original val- Peculiar features of the present approach, competitive with
ues whenever wn > 0.3 mm and wt > 1.5 mm (see Section 2.1); respect to more traditional numerical techniques, can be sum-
therefore, cohesion is almost exhausted at the downstream portion marised as follows:
of the interface. Friction plays a secondary role due to the present
dam geometry, which exerts higher self-weight pressure on the - the bulk material behavior is captured by matrix Z; its size de-
upstream portion of the foundation interface. The load carrying pends on the interface discretisation only and is usually smaller
capacity is better captured by limit state analysis based on rigid- by at least one order of magnitude than the stiffness matrix di-
body equilibrium conditions (specifically, against rotation around mension in classical FE approaches;
the downstream toe) as suggested by some national standards, for - Z is independent of the characteristic dimension for geometri-
instance [43]. However, rigid body predictions become unsafe and cally similar structures;
the load carrying capacity sensitive to the shear characteristics of - holonomic (reversible) interface response is hypothesised, con-
the foundation interface as the structural dimensions increase; see sistently with the assumption of progressive damage increase
Fig. 3(a). without local unloading; the dependence of the loading history
Results are somewhat different when water permeation into on the structural response can therefore be neglected; this in-
the structure is prevented; see Figs. 3(b), 6 and 7. The role of friction troduces some approximation (see, e.g., [31]) which is largely
at the foundation interface is now increased since the compres- compensated by the achieved benefits in terms of computing
sion state induced by the dam self-weight is not reduced by uplift costs;
pressure. Traditional approaches based on LEFM provide a fair es- - the formulation is reduced to a small size constrained optimiza-
timation of the safety margin as D < 20 m only. Above this tion problem; the explicit form given to the non-linear contri-
characteristic dimension, rigid body mechanics returns reliable bution representing the quasi-brittle behavior of the dilatant
prediction of the safety level for good shear interface properties. frictional joints, permits the analytical evaluation of gradients
On the contrary, 5° difference in the friction angle, well compati- in the iterative solution procedures, with obvious computing
ble with experimental uncertainties [20], can substantially change savings;
the collapse load and the corresponding failure mode of a dam of - the envisaged non conventional problem formulation naturally
40 m height when tnu = 0.7 MPa, see Fig. 6, despite the fact that leads to solution techniques that are simple and robust at the
opening displacements keep prevailing as D < 60 m. The same ef- same time, implemented in easily available software packages;
fect, even more pronounced, can be observed also for the reference equivalent results in terms of maximum load carrying capac-
ICOLD size D = 80 m; see Fig. 3(b). In this dimension range, a larger ity and corresponding failure mechanisms would otherwise re-
friction coefficient promotes stable crack propagation, with rather quire the implementation of sophisticated arc-length methods
beneficial effects in terms of structural safety. for the step-by-step integration of the governing equations in
Failure mode progressively changes with the characteristic standard finite element codes.
structural dimension. For high D values, opening displacements are
confined to the notch while sliding displacements at the founda-
tion interface dominate the ultimate structural behavior; see Fig. 7. References
In this situation, the assumption of Mode I failure mechanisms
[1] Ghaemmaghami A, Ghaemian M. Large-scale testing on specific fracture
would lead to unsafe estimation of the load carrying capacity, as
energy determination of dam concrete. Int J Fract 2006;141:247–54.
shown by the graph of Fig. 3(b). [2] De Sortis A, Paoliani P. Statistical analysis and structural identification in
concrete dam monitoring. Eng Struct 2007;29:110–20.
5. Summary and conclusion [3] Ardito R, Cocchetti G. Statistical approach to damage diagnosis of concrete
dams by radar monitoring: Formulation and a pseudo-experimental test. Eng
Sruct 2006;28:2036–45.
The present paper focused on an efficient numerical procedure [4] Ardito R, Maier G, Massalongo G. Diagnostics analysis of concrete dams based
for the evaluation of safety factors and failure modes of geometri- on seasonal hydrostatic loading. Eng Struct 2008;30:3176–85.
G. Bolzon / Engineering Structures 32 (2010) 1304–1311 1311

[5] Collison A, Wade S, Griffiths J, Dehn M. Modelling the impact of predicted [26] Bolzon G. Size effects in concrete gravity dams: A comparative study. Engng
climate change on landslide frequency and magnitude in SE England. Engng Fract Mech 2004;41:2957–75.
Geology 2000;55:205–18. [27] Maier G. Matrix structural theory of piecewise linear elastoplasticity with
[6] Calvetti F, di Prisco C, Nova R. Experimental and numerical analysis of soil-pipe interacting yield planes. Meccanica 1970;5:54–66.
interaction. J Geotech Geoenvironm Engng 2004;130:1292–9. [28] Bolzon G, Maier G, Novati G. Some aspects of quasi-brittle fracture analysis as
[7] Zhu DY, Lee CF, Jiang HD. Generalised framework of limit equilibrium methods a linear complementarity problem. In: Bažant ZP, Bittnar Z, Jirásek M, Mazars J,
for slope stability analysis. Géotechnique 2004;53:377–95. editors. Fracture and damage in quasi-brittle structures. London: E&FN Spon;
[8] Troncone A. Numerical analysis of a landslide in soils with strain-softening 1994. p. 159–74.
behaviour. Géotechnique 2005;55:585–96. [29] Maier G, Bolzon G, Tin-Loi F. Mathematical programming in engineering
[9] Rabczuk C, Areias PMA. A new approach for modelling slip lines in geological mechanics: Some current problems. In: Ferris MC, Mangasarian OL, Pang J-S,
materials with cohesive model. Int J Numer Anal Methods Geomech 2006;30: editors. Applied optimization series, vol. 50. Kluwer Academic Publisher; 2001.
1159–1152. p. 201–32.
[10] Sarma SK, Tan D. Determination of critical slip surface in slope analysis. [30] Cocchetti G, Maier G, Shen XP. Piecewise linear models for interfaces and
Géotechnique 2006;56:539–50. mixed mode cohesive cracks. CMES Comput Model Engng Sci 2002;3:279–98.
[11] Puntel E, Bolzon G, Saouma V. A fracture mechanics based model for joints [31] Bolzon G, Cocchetti G. Direct assessment of structural resistance against
under cyclic loading. ASCE J Engng Mech 2006;132:1151–9. pressurized fracture. Int J Numer Anal Methods Geomech 2003;27:353–78.
[12] Puntel E, Saouma V. Experimental behavior of concrete joint interfaces under [32] ICOLD Theme A2: Imminent failure flood for a concrete gravity dam. Fifth
reversed cyclic loading. J Engng Mech 2008;134:1558–68. international benchmark workshop on numerical analysis of dams. 1999.
[13] Bolzon G, Fedele R, Maier G. Parameter identification of a cohesive crack model [33] Maier G, Novati G, Cen ZZ. Symmetric Galerkin boundary element method for
by Kalman filter. Comput Methods Appl Mech Engrg 2002;191:2847–71. quasi-brittle fracture and frictional contact problems. Comput Mech 1993;13:
[14] Maier G, Bocciarelli M, Bolzon G, Fedele R. On inverse analysis in fracture
74–89.
mechanics. Int J Fract 2006;138:47–73.
[34] Maier G, Frangi A. Symmetric boundary element method for discrete crack
[15] Stavroulakis G, Bolzon G, Waszczyszyn Z, Ziemianski L. Inverse Analysis.
modelling of fracture processes. Comput Assist Mech Eng Sci 1998;5:201–26.
In: Karihaloo B, Ritchie RO, Milne I, editors. Comprehensive structural
[35] Bolzon G, Maier G, Tin-Loi F. Holonomic and non-holonomic simulation of
integrity, vol. 3. Kidlington (Oxfordshire, UK): Elsevier Science Ltd.; 2003.
quasi-brittle fracture: A comparative study of mathematical programming
p. 685–718.
approaches. In: Wittman FH, editor. Fracture mechanics of concrete structures.
[16] Renzi R, Ferrara G, Mazzà G. Cracking in a concrete gravity dam: A centrifugal
Freiburg: Aedificatio Publishers; 1995. p. 885–98.
investigation. In: Bourdarot E, Mazars J, Saouma V, editors. Dam fracture and
[36] Bolzon G. Hybrid finite element approach to quasi-brittle fracture. Comput &
damage. Rotterdam: Balkema; 1994. p. 103–10.
[17] Plizzari G, Waggoner F, Saouma VE. Centrifuge modeling and analysis of Structures 1996;60:733–41.
concrete gravity dams. ASCE J Struct Engng 1995;121:1471–9. [37] Bolzon G, Corigliano A. A discrete formulation for elastic solids with damaging
[18] Barpi F, Valente S. Numerical simulation of pre-notched gravity dam models. interfaces. Comput Methods Applied Mech Engrg 1997;140:329–59.
ASCE J Engng Mech 2000;126:611–9. [38] Ferris MC, Tin-Loi F. Limit analysis of frictional block assemblies as a
[19] Caron P, Léger P, Tinawi R, Veilleux M. Slot cutting of concrete dams: Field mathematical problem with complementarity constraints. Int J Mech Science
observations and complementary experimental studies. ACI Struct J 2003;100: 2001;43:209–24.
430–9. [39] Bolzon G, Maier G, Tin-Loi F. On multiplicity of solutions in quasi-brittle
[20] Rochon-Cyr M, Léger P. Shake table sliding response of a gravity dam model fracture computations. Comput Mech 1997;19:511–6.
including water uplift pressure. Eng Struct 2009;31:1625–33. [40] Statnikov RB, Matusov JB. Multicriteria optimization and engineering. New
[21] Bažant ZP, Planas J. Fracture and size effect in concrete and other quasibrittle York: Chapman and Hall; 1995.
materials. Boca Raton (FL): CRC Press; 1997. [41] Zhu ZQ, Liu Z, Wang XL, Yu RX. Construction of integral objective
[22] Bažant ZP. Scaling of structural strength. London (UK): Hermes Penton Science; function/fitness function of multi-objective/multi-disciplinary optimization.
2002. CMES Comput Model Engng Sci 2004;6:567–76.
[23] Hillerborg A. Application of the fictitious crack model to different types of [42] Matlab. User’s guide and optimization toolbox 3. USA: The Math Works Inc.;
materials. Int J Fracture 1991;51:95–102. 2006.
[24] Elices M, Guinea GV, Gòmez J, Planas J. The cohesive zone model: Advantages, [43] D.M. 24.03.82. Norme tecniche per la progettazione e la costruzione delle
limitations and challenges. Engng Fract Mech 2002;69:137–63. dighe di sbarramento. G.U. n. 212, 04.08.82.
[25] Chandra N, Shet C. A micromechanistic perspective of cohesive zone approach [44] Bolzon G, Cocchetti G. On a case of crack path bifurcation in cohesive materials.
in modeling fracture. CMES Comput Model Engng Sci 2004;6:21–33. Arch Appl Mech 1998;68:513–23.

You might also like