Sie sind auf Seite 1von 11

Soil Mechanics and Foundation Engineering, Vol. 51, No. 2, May, 2014 (Russian Original No.

2, March-April, 2014)

EARTHQUAKE-RESISTANT CONSTRUCTION
IMPACT OF FOUNDATION NONLINEARITY ON THE CRACK
PROPAGATION OF HIGH CONCRETE DAMS

UDC 624.131.552.2
M. A. Hariri-Ardebili
Department of Civil, Environmental and Architectural
Engineering, University of Colorado, Boulder, USA.

In the present paper the seismic stability of gravity and arch dams is evaluated under
earthquake ground motion considering the foundation nonlinearities. For this purpose,
the finite element models of the 103 m Koyna gravity dam and the 203 m Dez arch dam
are prepared. A fixed smeared crack model with the Willam-Warnke failure criterion is
used for modeling the concrete cracking. Also the elasto-plastic model with the DruckerPrager yield criterion is used for foundation cracking. Viscous boundary models are
used for the far-end areas of the foundation medium to absorb the outgoing waves. It
was found that considering the nonlinear foundation model increases the cracked areas
on the dam body, especially in the vicinity of the foundation, and reduces the total stability of the coupled system.

1. Introduction

There are some factors that affect the numerical analysis of a concrete dam, i.e., the semiunbounded size of the reservoir and foundation rock domains, the dam-reservoir interaction, the wave
absorption at the reservoir boundary, water compressibility, dam-foundation rock interaction, spatial
variations in ground motion at the dam-rock interface, the complex nature of material and loads, and
also their interaction in the dam-reservoir-foundation coupled system [1]. Dam-foundation interaction
effects are usually considered by modeling the semi-infinite foundation as a massless medium. Appropriate boundary conditions are required in the exterior boundaries in order to absorb the outgoing
waves. The dam-foundation interaction problem was investigated by Chopra and Chakrabarti [2], Leger
and Boughoufalah [3], Nuss et al. [4], Bayraktar et al. [5], Lemos and Gomes [6], Saleh and Madabhushi [7], Lebon et al. [8], Saouma et al. [9], Hariri-Ardebili and Mirzabozorg [10], Burman et al. [11],
and Hariri-Ardebili and Mirzabozorg [12].
In this paper, an arch dam and a gravity dam are studied for the sake of investigating the
impact of the foundation rock nonlinearities on the seismic stability and crack propagation of the concrete dams in two- (2D) and three-dimensional (3D) cases. The viscous boundary model is implemented at the outer edges of the massed foundation in the numerical model of the dam-reservoirfoundation system. A fixed smeared crack model with a five-parameter failure criterion is used for
crack simulation in mass concrete, while the elasto-plastic model with a two-parameter yield criterion is applied for the foundation rock. Lastly, results are compared in terms of principal stresses, displacements, and crack profiles.

Translated from Osnovaniya, Fundamenty i Mekhanika Gruntov, No. 2, p. 14, March-April, 2014.
72

0038-0741/14/5102-0072

2014 Springer Science+Business Media New York

2. Material Constitutive Models

For crack propagation analysis of a structural system, the proposed model should be able to simulate the behavior of the material in the pre-softening phase, the softening phase, and the cracking behavior.
2.1. Linear Elastic Behavior
In general, the relationship between the stress and strain vectors at the pre-softening phase is given by

= D0 ,

(1)

where D0 is the elastic modulus matrix, and are the vector of stress and strain components. Assuming isotropic linear behavior at the pre-softening phase, the isotropic modulus matrix can be written as
Eq. (2) below. It should be noted that most brittle materials, such as concrete and rock, show anisotropic behavior even at the pre-softening phase. However, modeling such behavior requires information on
Young's modulus, Poisson's ratio, and the shear modulus corresponding to three orthogonal directions.
Due to lack of data for case studies, the behavior of the material is assumed to be isotropic at the presoftening phase.
E (1 v )

(1 + v )(1 2 v )

E (1 v )

Ev
Sym.

1
v
1
2
v
1
v
1
2
v
+

)
)(
) ( )(
(

E (1 v )
E
E

1 + v 1 2v 1 + v 1 2 v 1 + v 1 2 v

,
[D0 ]= ( )( )( )( ) ( )( ) E

0
0
0

2 (1 + v )

0
0
0
0

2 (1 + v )

E
0
0
0
0
0

2 (1 + v )

(2)

where E and v are the isotropic modulus of elasticity and Poisson's ratio of the material, respectively
(concrete or rock in this research).
2.2. Fixed Smeared Crack Model for Concrete
In the smeared crack approach, after initiation of the fracture process, which is determined by a
suitable constitutive model, the pre-crack stress-strain relation of the material is replaced by an orthotropic relation with the reference axis system of the material aligned with the fracture direction [13]. In this
study, the five-parameter Willam-Warnke [14] model is used as the failure criterion (Fig. 1, a). The criterion for failure of concrete due to a multi-axial stress state can be expressed as

fc

0,

(3)

where is a function of the principal stress state, is the failure surface, and f 'c is the compressive
strength of the concrete material.
The behavior of the concrete element is assumed to be linear elastic until it reaches the ultimate
strength. Cracking occurs when the principal tensile stress in any direction lies outside the failure surface. In this research, cracking is permitted in three orthogonal directions at each integration point.
When cracking occurs at the integration point, the stress-strain relation is modified, defining a weak
73

zp

fc

(inner edges)

xp = yp = zp

xp
fc

=0

Octahedral plane

yp
fc

1
2

(outer edges)
= 2/3

= 4/3

Fig. 1. Material failure surface: a) William-Warnke failure surface in principal stress


space for the concrete; b) Mohr-Coulomb (1) and Drucker-Prager (2) yield
surfaces in the deviatoric plane for the rock.

plane normal to the crack direction, which is unable to endure any tensile stresses. The presence of a
crack at an integration point and in the specific direction is represented by the modification of the stiffness matrix exerting shear transfer coefficient on the cracked plane [15]. Based on the fact that the mass
concrete can be cracked in one, two, or three orthogonal directions, the modulus matrix can be updated.
The local orthotropic modulus matrix, Dlcr, for the cracked integration points only in one direction can
be written as follows:
The crack is only in one direction and it is open:
E S

E
Sym.
0

(1 + v )(1 v )

E
E
0

(1 + v )(1 v )(1 + v )(1 v )

l
E
Dcr =
,

O
0
0
0
2 (1 + v )

E
0
0
0
0

2 (1 + v )

E
0

0
0
0
0

+
v
2
1
( )

(4)

where E S represent secant modulus and O is the open shear transfer coefficient.
The crack is only in one direction and it is closed:
E (1 v )

(1 + v )(1 2 v )

E (1 v )
Ev

Sym.
(1 + v )(1 2 v ) (1 + v )(1 2 v )

E (1 v )
Ev
Ev

l
1
1
2
1
1
2
1
1
2
+

v
v
v
v
v
v
)(
)( )(
) ( )(
)
Dcr = (
(5)
,

E

0
0
0
C
2 (1 + v )

E
0
0
0
0

2 (1 + v )

0
0
0
0
0
C

2 (1 + v )

where C is the closed shear transfer coefficients. The value of varies from 0.0 to 1.0, in which 0.0
represents a smooth crack and 1.0 represents a rough crack. The values of the shear transfer coefficient
74

can be estimated roughly for different types of concrete structures using experiments or by a set of sensitivity analysis and comparing the results with the coaxial rotating smeared crack model [16]. The
cracked modulus matrices for the other cracking conditions, i.e., cracking in two and three directions,
can be obtained using a similar approach as discussed in [13]. The above-mentioned cracked modulus
matrices are in the local coordinate, which corresponds to the direction of the principal strains. This
matrix should be transformed to the global coordinate using strain transfer matrix T, as follows [16]:
(6)

g
Dcr
= TT Dlcr T.

2.3. Elasto-plastic Model for Rock


The Drucker-Prager material model based on the elastic-plastic constitutive matrix is used for
modeling the failure of the foundation rock [17]. For an elasto-plastic material, the stress-total strain
relation can be expressed as
= Dep ,
(7)
where Dep is the elasto-plastic constitutive matrix. The total strain tensor can be divided into elastic and
plastic parts. The stress-elastic strain relation can be expressed as

= D e e,

(8)

where De is the elastic constitutive matrix.


Considering the yield function as F( ), we write its differential as dF. In the incremental theory
of plasticity, during the plastic deformation F( ) = 0, and consequently dF = 0 is satisfied over the load
step. This condition can be expressed as
T

F
F e
p
dF = d =
D (d d ) = 0.

(9)

Moreover, considering a non-associative plasticity, the increment of plastic strain will be


d p = d

G( )
,

(10)

where d is a proportional factor and G( ) is a plastic potential function.


Substituting (10) in (9) and after some simplifications we obtain
T
T
F
F De G( ) = 0.
dF = De d d

(11)

Combining (11), (10), and (7), we obtain the general expression for the elasto-plastic constitutive
matrix as
T

De
Dep = De

G( ) F e
D

T

F e G ( )
D

(12)

The Drucker-Prager yield surface represents a cone, whose axis matches the space diagonal of
the coordinate system of principal stresses. The pressure-sensitivity incorporation in this model can be
75

used for accurate modeling of materials such as soil and rock, which are characterized by a strong
dependence of their yield limit on the hydrostatic pressure. The Drucker-Prager yield function is presented as [18]
(13)
F ( ) = 0 I1 + J 2 k ,
where I1 and J2 are the first and second invariants of the stress tensor, respectively, 0 is a material constant, and the parameter k defines the size of the yield surface and can be considered as a constant
parameter for the material without hardening. The last two parameters can be related to the cohesion c
and the friction angle .
Also, assuming associative plasticity, we define the plastic potential by the same function as
G( ) = 0 I1 + J 2 k ,

(14)

where 0 is the material constant parameter related to the dilatancy, representing an inelastic volume
increase.
When the Drucker-Prager surface is matched with the Mohr-Coulomb surface, the parameters 0
and k are given by

0 =

2sin

3 (3 sin )

; k=

6 c cos

3 (3 sin )

(15)

where corresponds to the inner and outer Drucker-Prager surfaces, respectively, as shown in Fig.1, b.
If the inner surface is used, the tensile strength is correctly predicted and the compressive strength is
overestimated, while if the outer surface is used, the compressive strength is correctly predicted but the
tensile strength is overestimated.
3. Foundation-Dam Interaction and Boundary Conditions

Generally there are three types of foundation modeling, i.e., the rigid foundation model, the
massless foundation model, and the massed foundation model. In the case of the massed foundation
model, appropriate boundary conditions at the far-end edges are required in order to absorb the outgoing waves. In this study, the viscous boundary model is used at the foundation exterior surface to prevent wave reflection at the artificial boundaries. The viscous boundary condition is applied at the outer
surface of the foundation in 3D space given as [19]
= f VP u

1 = f VS v ,
2 = f VS w

(16)

where and are the normal and shear tractions, respectively; u , v , and w are the normal and two
tangential particle velocities at the boundary; f is the foundation mass density; VS and VP are the shear
and pressure wave velocities at the foundation medium, respectively, given by [12]

VS =

2(1 f )
1
; VP =
VS ,
f 2(1 + f )
(1 2 f )

Ef

(17)

where Ef and vf are the modulus of elasticity and Poisson's ratio of the foundation, respectively. Radiation damping can be derived from Eq. (16) and applied on the far-end boundary of the foundation using
lumped dashpots that are added to the global damping matrix of the structure as follows
76

Fig. 2. Finite element model of the coupled system for a) 2D


Koyna gravity dam, and b) 3D Dez arch dam.

TABLE 1
Characteristics
Ec, GPa
c, kg/m3
vc
f 't, MPa
f 'c, MPa
Ef, GPa
f, kg/m3
vf
c, MPa
, deg

Koyna Dam
Dez Dam
Static value Dynamic value Static value Dynamic value
31.03
35.68
40.0
46.0
2,643
2,643
2,400
2,400
0.20
0.14
0.20
0.14
2.40
3.60
3.40
5.10
24.00
36.00
35.00
36.60
16.86
19.18
13-15
15-17.25
2,700
2,700
2,400-2,500
2,400-2,500
0.18
0.13
0.25
0.18
0.6
0.6
0.8-1.0
0.8-1.0
41
41
30~34
30-34

Cni = f VP N i dA ; Csi = f VS Ni dA,


Ae
i

(18)

Ae

where C n and C s are the damping factors in the normal and tangential directions, respectively, and Ni is
the element shape function. A similar approach can be derived for the 2D case.
It is noteworthy that, the impact of the reservoir water is considered using an Eulerian approach
on the upstream face of the dam body and also on the foundation. The general equation of the fluidstructure interaction and different boundary conditions of the reservoir domain can be found in [20].
4. Application to Concrete Dams

The Koyna Dam in India is selected as a case study of gravity dams. The existing Koyna Dam is
a rubble concrete gravity dam of 853 m length and 103 m height, and its thickness at the base and crest
are 70.2 m and 12.1 m, respectively, for the central non-overflow monoliths [21]. The 2D finite element
model of Koyna gravity dam-reservoir-foundation system is shown in Fig. 2, a. Dez Dam is selected as a
case study of arch dams. Dez is a double curvature high arch dam. The total height of the dam is 203 m
but the height above the concrete plug is 194 m. The crest length is 240 m, the thickness of the dam at
the crest is 4.5 m, and its maximum thickness at the base is 21.0 m [22]. The 3D finite element model
of Dez arch dam-reservoir-foundation system is shown in Fig. 2, b. In order to better capture the concrete
cracking under the seismic loads, the vertical contraction joints are not modeled between the arch dam
blocks in the present research. Also none of the horizontal lift joints are considered in this paper.
Table 1 represents the mechanical and strength properties of the mass concrete and the foundation
rock for both Koyna and Dez dams. It should be noted that the dynamic properties are different from
static properties due to the rate dependence of the mechanical and strength properties of the material
[23]. Also the appropriate safety factors [23] are applied to the dynamic values in this table. The applied
loads on the system are dam body self-weight, hydrostatic pressure, and seismic load. The Newmark-
time integration method is utilized to solve the coupled problem of the dam-reservoir-foundation system.
77

0.5
Horizontal component

0
-0.5

10

20

0.5

30

40, sec

Vertical component

0
-0.5
0

10

20

30

40, sec

Acceleration (g) Acceleration (g) Acceleration (g)

Acceleration (g)
Acceleration (g)

0.5

Component L

0
-0.5

10

20

0.5

30

40

50, sec

Component T

0
-0.5 0

10

20

0.5

30

40

50, sec

Component V

0
-0.5

10

20

30

40

50, sec

Fig. 3. Acceleration time history of the scaled ground motions for (a) Imperial Valley-06; (b) Tabas.

The PEER [24] NGA ground motion database is used in order to select the appropriate ground
motions based on the general site characteristics of the dams. Imperial Valley-06 ground motion recorded at Victoria Station is chosen for the Koyna Dam. On the other hand, Tabas ground motion recorded
at the Tabas Station is selected for the Dez Dam. Figure 3 shows the scaled acceleration time history of
the selected motions in both horizontal and vertical directions. Imperial Valley-06 ground motion is
scaled based on peak ground acceleration of the ground motion that destroyed the Koyna Dam; while
Tabas ground motion is scaled based on the maximum design level earthquake at the Dez Dam site
obtained already by seismic hazard analysis [20]. In order to reduce the computational efforts in the
nonlinear analysis of the coupled system, only the strong ground motion part of the records are applied,
which includes at least 90% of the Arias intensity of the motion. Accordingly, the significant duration
for Imperial Valley-06 and Tabas ground motions is calculated as 22.0 and 33.0 sec, respectively.
5. Results and Discussion

Three different finite element models are provided for seismic analysis:
Model 1: Both the concrete and the rock materials are assumed to be linear elastic.
Model 2: Concrete is modeled using the smeared crack approach, while the rock is assumed to
be linear elastic.
Model 3: Concrete is modeled using the smeared crack approach, and the rock is modeled
based on the elasto-plastic model.
The dam-reservoir-foundation system is excited using the horizontal and vertical components of
the scaled ground motions. The responses of the coupled system due to its gravity load and the hydrostatic pressure are taken as initial conditions in the seismic analysis.
5.1. Gravity Dam
The time history of the crest displacement in the stream direction is shown in Fig. 4 for all the
three models. The positive direction of the displacement is in the downstream direction. The responses
of the nonlinear models before initiation of crack at the first integration point (t = 6.80 sec) coincide
with that obtained from the linear elastic analysis. This means that both the tensile and compressive
stresses of the concrete and rock are less than the strength of the material. Cracking of the materials
leads to separation of the displacement curves from each other. Model 3 fails earlier than the other nonlinear model at t = 13.90 sec. Model 2 fails at t = 14.90 sec, one second later than the previous one.
The type of failure in both cases is full collapse. This means that there is at least one full cracked path
that connects the upstream face to the downstream. Thus, the displacement time history is not shown
after the collapse.
78

Displacement, mm

, sec

Fig. 4. Displacement time history of the crest point at the


stream direction for Koyna Dam.

Max. first principal stress

Min. third principal stress

Max. first principal stress

Min. third principal stress

Fig. 5. Nonconcurrent envelope of the principal stresses up to failure time for the Koyna Dam (Pa):
a) model 2; b) model 3.

Initiation

Initiation

Propagation

Propagation

Failure

Failure

Fig. 6. Crack propagation process in Koyna dam-foundation system: a) model 2 ; b) model 3.

Figure 5 shows the nonconcurrent envelope of the principal stresses, i.e., the first principal stress
and the third principal stress, up to failure time of the dam for nonlinear models, i.e., model 2 and
model 3. As seen, the neck area and also some parts of the dam-foundation interface, especially in the
vicinity of the dam heel, are critical locations. It should be noted that the failure of each integration
point depends on the combined resultant of the principal stresses and also the status of the considered
point with respect to the failure surface. Based on this figure, the foundation nonlinearity has a direct
effect on increasing the overstressed areas, especially at the neck.
Figure 6 shows the crack propagation process in the Koyna Dam under the earthquake loading.
In model 2, the crack is initiated at the downstream face near the slope discontinuity, while the other set
of cracks starts at the heel of the dam. As time passes, the base crack propagates downstream. The crack
79

Displacement, mm

, sec

Fig. 7. Displacement time history of the crest point at the stream


direction for Dez Dam.

Max. first principal stress

Min. third principal stress

Max. first principal stress

Min. third principal stress

Fig. 8. Nonconcurrent envelope of principal stresses on the upstream and downstream faces for theDez
Dam (Pa): a) model 2; b) model 3.

at the neck then propagates in an inclined line towards upstream. After it reaches the middle of the neck
area, it propagates almost horizontally to reach the upstream face. This figure also shows the crack
propagation of the model, including both the concrete and rock nonlinearities. Like the previous model,
cracking starts at the point of slope discontinuity on the downstream face and extends through the width
of the neck. Cracks also appear at the dam-foundation interface and propagate from heel of the dam
towards downstream. Also, there are some cracked sections in the vicinity of the toe of the dam. The
final length of the base crack is higher in the case of the nonlinear rock model than in the linear one.
Assuming the elasto-plastic model for the rock leads to cracking of the foundation near both the heel
and toe of the dam, the dam fails completely in model 3 earlier than in model 2.
5.2. Arch Dam
The time history of the crest displacement for the crown cantilever in the stream direction is
shown in Fig. 7 for all the three models of the Dez Dam. The positive sign of the displacement is in the
downstream direction. The responses of the nonlinear models before initiation of crack at the first integration point (t = 8.70 sec) coincide with that obtained from the linear elastic analysis. The displacement has a higher value for model 3 than for model 2, and both are higher than the linear elastic model.
None of the nonlinear models fail completely in this dam because there is no full cracked path in the
finite element models. A higher displacement for a numerical model means a lower stiffness of the coupled system at the considered load step.
Figure 8 shows the nonconcurrent envelope of the principal stresses on the upstream and downstream faces for arch dam nonlinear models. As seen, the upper parts of the dam body on both faces, especially in the vicinity of the crest, are critical locations for tensile stresses. Also, the upper parts of the cen80

Initiation

Propagation

Final

Initiation

Propagation

Final

Fig. 9. Crack propagation process in the upstream and downstream faces of


the Dez dam-foundation system: a) model 2; b) model 3.

tral blocks near the crest level are critical locations for compressive stresses. Modeling the nonlinear foundation increases the overstressed area for both tensile and compressive stresses. Also modeling the foundation with nonlinear behavior transfers the tensile overstressed area to the dam-foundation interface.
Figure 9 shows the crack propagation process in the upstream and downstream faces of the Dez
Dam. In model 2, crack is initiated near the crest level in both faces and propagates downwards and also
towards the abutment. In model 3, cracking starts at the upper parts of the dam near the crest level.
Cracks propagate toward the dam-foundation interface and joint to other sets of cracks that had already
started at the dam-foundation interface. There are several cracked elements in the foundation medium in
the vicinity of the dam. The number of cracked elements in the final crack profile for model 3 is more
than that of model 2. This means that for the realistic crack evaluation of arch dams, especially at the
interface with foundation, the nonlinear behavior of the rock should be considered.
Conclusion

The objective of this study is to evaluate the seismic stability and crack propagation of concrete
gravity and arch dam-reservoir-foundation coupled systems under the earthquake loading. For this purpose, detailed finite element models of the coupled system are provided. The fixed smeared crack model
is used for modeling the nonlinear behavior of mass concrete, and the Drucker-Prager elasto-plastic
model is used for foundation rock. The viscous boundary approach is used as an artificial absorbing
layer at the exterior surface of the foundation medium. Fluid-structure interaction is considered using
the Eulerian-Lagrangian approach.
It is found that considering the nonlinear behavior of the rock leads to cracking of the foundation beneath the gravity dam, especially in the vicinity of the heel. Also, it leads to cracking of the abutment of the arch dam in the upper elevations. Cracking at the foundation leads to decreasing the total
stiffness of the coupled system and increasing the vulnerability of the dam body. The nonlinear model
for the foundation rock with the ability to crack leads to increasing the cracked area on the concrete
81

dam body. Also, it is concluded that the nonlinear dam-nonlinear foundation interaction should be considered for seismic safety evaluation of high gravity and arch dams, especially in seismic prone areas.
REFERENCES
1.
A. K. Chopra, "Earthquake analysis of arch dams: factors to be considered," Struct. Eng. (ASCE),
138, 205-214 (2012).
2.
A. K. Chopra, and P. Chakrabarti, "Earthquake analysis of concrete gravity dams including damwater-foundation rock interaction," Earthquake Eng. Struct. Dynam., 9, 363-383 (1981).
3.
P. Leger, and M. Boughoufalah, "Earthquake input mechanisms for time domain analysis of damfoundation systems," Eng. Struct., 11, 37-46 (1989).
4.
L. K. Nuss, R. L. Munoz, F. J. Jackmauh, and A. K. Chopra, "Influence of dam-foundation interaction
in seismic safety evaluation of two arch dams," Proc. 12th World Conference on Earthquake Engineering,
Auckland, New Zealand (2000).
..
5.
A. Bayraktar, E. Hancer, and M. Akk o se, "Influence of base-rock characteristics on the stochastic
dynamic response of dam-reservoir-foundation systems," Eng. Struct., 27, 1498-1508 (2005).
6.
J. V. Lemos, and J. P. Gomes, "Modeling seismic failure scenarios of concrete dam foundations,"
Applications of Computational Mechanics in Geotechnical Engineering, Sousa, Fernandes, Vargas, Jr.
and Azevedo (eds.), Taylor and Francis, London, 341-349 (2007).
7.
S. Saleh, and S. P. G. Madabhushi, "Response of concrete dams on rigid and soil foundations under
earthquake loading," Earthquake Tsunami, 4, No.3, 251-268 (2010).
8.
G. Lebon, V. Saouma, and Y. Uchita, "3D rock-dam seismic interaction," Dam Eng., 21, No.2,
101-130 (2010).
9.
V. Saouma, F. Miura, G. Lebon, and Y. Yagome, "A simplified 3D model for soil-structure interaction
with radiation damping and free field input," Bull. Earthquake Eng., 9, No.5, 1387-1402 (2011).
10.
M. A. Hariri-Ardebili, and H. Mirzabozorg, "Effects of near-fault ground motions in seismic performance
evaluation of a symmetry arch dam," Soil Mech. Found. Eng., 49, No.5, 192-199 (2012).
11.
A. Burman, P. Nayak, P. Agrawal, and D. Maity, "Coupled gravity dam-foundation analysis using a
simplified direct method of soil-structure interaction," Soil Dynam. Earthquake Eng., 34, 62-68 (2012).
12.
M. A. Hariri-Ardebili, and H. Mirzabozorg, "A comparative study of the seismic stability of coupled
arch dam-foundation-reservoir systems using infinite elements and viscous boundary models," Int. J.
Struct. Stabil. Dynam., 13, No.6, doi:10.1142/S0219455413500326.
13.
M. A. Hariri-Ardebili, H. Mirzabozorg, and R. Kianoush, "A study on nonlinear behavior and seismic
damage assessment of concrete arch dam-reservoir-foundation system using endurance time analysis,"
Int. J. Optim. Civil Eng., 2, No.4, 573-606 (2012).
14.
K. J. Willam, and E. P. Warnke, "Constitutive model for tri-axial behavior of concrete," Int. Assoc.
Bridges Struct. Eng., Italy (1974).
15.
M. A. Hariri-Ardebili, S. M. Kolbadi, M. Heshmati, and H. Mirzabozorg, "Nonlinear analysis of concrete
structural components using coaxial rotating smeared crack model," App. Sci., 12, No.3, 221-232 (2012).
16.
M. A. Hariri-Ardebili, S. M. Seyed-Kolbadi, and H. Mirzabozorg, "A smeared crack model for seismic
failure analysis of concrete gravity dams considering fracture energy effects," Struct. Eng. Mech., 48,
No.1, 17-39 (2013).
17.
D. Drucker, and W. Prager, "Soil mechanics and plastic analysis or limit design," Quar. App. Math.,
10, No.2, 157-65.
18.
L. R. Alejano, and A. Bobet, "Drucker-Prager Criterion," Rock Mech. Rock Eng., 45, 995-999
(2012).
19.
J. Lysmer, and R. Kuhlemeyer, "Finite element model for infinite media," Eng. Mech. ASCE, 95, No. 4,
859-877 (1969).
20.
M. A. Hariri-Ardebili, H. Mirzabozorg, and M. R. Kianoush, "Seismic analysis of high arch dams
considering contraction-peripheral joints coupled effects," Central Eur. J. Eng., 3, No.3, 549-564
(2013).
21.
S. S. Bhattacharjee, and P. Leger, "Application of NLFM models to predict cracking in concrete gravity
dams," Struct. Eng. (ASCE), 120, No.4, 1255-1271 (1994).
22.
M. A. Hariri-Ardebili, H. Mirzabozorg, M. Ghaemian, M. Akhavan, and R. Amini, "Calibration of 3D
FE model of Dez high arch dam in thermal and static conditions using instruments and site observation,"
Proc. 6th International Conference in Dam Engineering, Lisbon, Portugal (2011).
23.
Federal Energy Regulatory Commission (FERC), Engineering guidelines for the evaluation of
hydropower projects, Chapter 11: arch dams, Washington DC, USA (1999).
24.
PEER ground motion database, http://peer.berkeley.edu/peer_ground_motion_database, Beta version,
University of California, Berkeley, CA, USA (2010).

82

Das könnte Ihnen auch gefallen