Anda di halaman 1dari 10

Vienna Congress on Recent Advances in

Earthquake Engineering and Structural Dynamics 2013 (VEESD 2013)


C. Adam, R. Heuer, W. Lenhardt & C. Schranz (eds)
28-30 August 2013, Vienna, Austria
Paper No. 152

Consideration of aging effects on the time-dependent seismic vulnerability


assessment of RC buildings
S.T. Karapetrou1, S.D. Fotopoulou1, K.D. Pitilakis1
1

Unit of Soil Dynamics and Earthquake Engineering, Aristotle University of Thessaloniki, Thessaloniki, Greece

Abstract. The present study aims at the assessment of the seismic vulnerability of reinforced concrete (RC)
buildings considering performance degradation over time due to aging effects. Chloride induced corrosion is
taken into account based on probabilistic modeling of corrosion initiation time and corrosion rate. Twodimensional incremental dynamic analysis (IDA) is performed to assess the seismic performance of the initial
uncorroded (t=0 years) and corroded (t= 25, 50, 75 years) RC fixed base frame structures designed based on
different seismic code levels. The time-dependent fragility functions are derived at the various time periods in
terms of spectral acceleration corresponding to the fundamental mode of the structure Sa(T1,5%) for the
immediate occupancy (IO) and collapse prevention (CP) limit states. Results show an overall increase in seismic
vulnerability over time due to corrosion indicating the significant effect of deterioration due to aging effects on
structural behaviour.
Keywords: RC buildings; seismic vulnerability; time-dependent fragility curves; aging effects; IDA

1 INTRODUCTION
Traditionally in seismic vulnerability assessment it is implicitly assumed that the structures are
optimally maintained during their lifetime neglecting any deterioration mechanism that may adversely
affect their structural performance. On this basis, the impact of progressive deterioration of the
material properties caused by aggressive environmental attack is not accounted for. One of the most
important environmental degradation mechanisms of RC buildings is corrosion due to chloride
penetration, leading to the variation of the mechanical properties of steel and concrete over time.
Consequently, both the safety and the serviceability of RC structures may be affected under the action
of seismic loading, compromising the ability of the structures to withstand the loads they are designed
for. The severe uncertainties involved in corrosion phenomena point out the need for a probabilistic
approach to predict degradation phenomena (Duracrete 2000). Recognizing the importance of this
issue, several probabilistic models have recently been introduced into the time-variant vulnerability
assessment of corroded bridges and RC frame buildings (e.g. Ghosh and Padgett 2010, Choe et al.
2010, Fotopoulou et al. 2012).
Based on the above considerations, the aim of this study is the development of time-dependent
fragility curves taking into account deterioration due to aging effects. To demonstrate the
methodology for the time-dependent vulnerability assessment, three RC moment resisting frames,
designed according to different seismic code levels, are selected as case studies. The consideration of
aging is achieved by including probabilistic models of chloride induced corrosion deterioration of the
RC elements within the vulnerability assessment framework. Fragility curves are developed for the
initial (t=0 years) and the corroded buildings (t=25, 50, 75 years) based on the statistical exploitation
of the results of incremental dynamic analysis of the given structures. The derived curves are
compared to gain insight into the potential effects of deterioration due to aging on seismic
vulnerability highlighting the importance of their incorporation in fragility modeling.

S.T. Karapetrou, S.D. Fotopoulou, K.D. Pitilakis/ VEESD 2013

2 METHODOLOGY
2.1 Description of the structural models
Three moment resisting frames (MRFs) have been selected as reference structures designed according
to different seismic code levels (Figure 1) to illustrate the effect of aging in the time-dependent
seismic vulnerability assessment. Syner-G (www.syner-g.eu) taxonomy for RC structures is used to
describe the different building configurations under study. The first one is a three storey three bay
frame model originally designed for the purpose of an experimental study (Bracci et al. 1992) that is
representative of low rise buildings designed for gravity loads only with no seismic provisions (Low
rise-No code). The second is a four storey three bay frame structure (Kappos et al. 2006) that
represents mid rise buildings designed following the provisions of the Greek modern seismic code
(Mid rise-High code). Finally the third is a nine storey frame model with three bays (Kappos et al.
2006) that is considered typical of high rise buildings with low level of seismic design according to
the prescription of the 1959 seismic code of Greece (High rise-Low code). For all the structural
models under study fixed base conditions have been considered. Table 1 presents some of the main
characteristics of the reference models such as the total mass, the fundamental period and the strength
of concrete and steel. The analytical modeling of the structures is conducted using OpenSees finite
element analysis platform (Mazzoni et al. 2009). Inelastic force-based formulations are employed for
the simulation of the nonlinear beam-column frame elements considering four Gauss-Lobatto
(Noenhofer and Filippou 1997) integration points along each members length. The applied
formulations allow both geometric and material nonlinearities to be captured. Distributed material
plasticity along the element length is considered based on the fiber approach to represent the crosssectional behaviour. The modified Kent and Park model (Kent and Park 1971) is used to define the
behavior of the concrete fibers, yet different material parameters are adopted for the confined (core)
and the unconfined (cover) concrete. The uniaxial Concrete01 material is used to construct a uniaxial
Kent-Scott-Park concrete material object with degraded linear unloading/reloading stiffness according
to the work of Karsan-Jirsa (Karsan and Jirsa 1969) with zero tensile strength. The steel reinforcement
is modeled using the uniaxial Steel01 material to represent a uniaxial bilinear steel material with
kinematic hardening described by a nonlinear evolution equation.
Table 1. Characteristics of the studied buildings
RC building

Total mass [t]

Initial fundamental period T [sec]

fc [MPa]

fy [MPa]

Low-rise/No code

207.0

0.98

24.0

276.0

Mid-rise/High code

130.0

0.66

20.0

400.0

High-rise/Low code

334.0

0.89

14.0

400.0

2.2 Corrosion modelling


Several models have been proposed to quantify and account for corrosion in the design, construction,
fragility analysis and maintenance of RC structures. A summary of these models can be found e.g. in
DuraCrete (1998). In the present study the corrosion of reinforcing bars due to the ingress of chlorides
is considered, as it is reportedly one of the most serious and widespread deterioration mechanisms of
RC structures. The probabilistic model proposed by FIB- CEB Task Group 5.6 (2006) is adopted to
model corrosion initiation time due to chloride ingress that is expressed as:

2
Tini
n

4 ke kt DRCM ,0 t0

C
erf 1 1 crit

Cs

1 n

(1)

S.T. Karapetrou, S.D. Fotopoulou, K.D. Pitilakis/ VEESD 2013

Figure 1. Reference MRF models used for time dependent vulnerability assessment

where Tini=corrosion initiation time (years), =cover depth (mm) Ccrit.=critical chloride content (wt %
cement); Cs = the equilibrium chloride concentration at the concrete surface (wt % cement); t0=
reference point of time (years); DRCM,0=Chloride migration Coefficient (m2/s); ke=environmental
function; kt=regression parameter; erf=Gaussian error function and n=aging exponent.
The statistical quantification of the model parameters describing the chloride induced corrosion
adopted for the present study is given in Table 2 in accordance with FIB- CEB Task Group 5.6 (2006)
prescriptions and the available literature (e.g. Choe et al. 2009; Ghosh and Padgett 2010). An
atmospheric exposure environment (e.g. ke=0.67, Choe et al. 2009) with water-to cement ratio of the
concrete material equal to 0.5 is assumed for the considered chloride induced deterioration scenario. It
is noted that the adopted corrosion rate implies a relatively high corrosion level (Stewart, 2004).
First Order Second Moment (FOSM) reliability analysis is conducted to assess Tini that varies for the
different models under study depending on their structural characteristics and more specifically on the
cover depth considered for each case (see Table 2). Based on the applied probabilistic model mean
Table 2 Statistical characteristics of parameters affecting the chloride induced corrosion deterioration of RC
elements adopted in the present study
Parameter

Mean

COV

Distribution

Cover Depth (mm)

20/25

0.40/0.32

Lognormal

0.67

0.17

Normal

Chloride migration Coefficient (DRCM,0) (m /s)

1.58E-11

0.20

Normal

Aging exponent n

0.362

cov=0.677 , a=0.0, b=0.98

Beta

Critical Chloride Concentration (Ccr) wt % cement

0.6

cov=0.25, a= 0.2, b=2.0

Beta

Surface Chloride Concentration (Cs) wt % cement

1.283

0.20

Normal

0.25

Normal

Environmental transfer variable ke


2

Rate of Corrosion (icorr) mA/cm

S.T. Karapetrou, S.D. Fotopoulou, K.D. Pitilakis/ VEESD 2013

values for Tini are estimated as 7.01 and 14.11 years respectively for the structural models designed
with no or low seismic provisions and with modern seismic code.
The effects of corrosion are assumed to be distributed uniformly about the perimeter and along the
concrete members. Once the protective passive film around the reinforcement dissolves due to
continued chloride ingress, corrosion initiates and the time-dependent loss of reinforcement crosssectional area can be expressed as (e.g. Ghosh and Padgett 2010):

2
n Di 4
A(t )
2
max n D t ,

for t T
ini

(2)

for t Tini

where n=number of reinforcement bars; Di=initial diameter of steel reinforcement; t=elapsed time in
years and D(t)=reinforcement diameter at the end of (t Tini) years, which can be defined as:

D(t ) Di icorr (t Tini )

(3)

where icorr=rate of corrosion (mA/cm2); =corrosion penetration (m/year) ( =11.6m/year uniform


corrosion penetration for generalized corrosion).
Due to the radial pressure developed along the steel bar surfaces, caused by the increasing volume of
the corrosion products, the tensile stresses in the concrete surrounding the rebars may exceed the
tensile strength leading thus to the cracking of the concrete cover. The cracking and spalling of the
cover concrete are taken into account by reducing the concrete cover strength according to the model
proposed by Coronelli and Gambarova (2004) and applied latter in Simioni (2009):
f c*

fc
1 K

1
co

(4)

where K is a coefficient related to bar diameter and roughness (K = 0.1 for medium-diameter ribbed
bars), co is the strain at the peak compressive stress fc, 1 is the average tensile strain in the cracked
concrete calculated as:

b f bo
bo

(5)

where bo is the section width in the virgin state before corrosion cracking while bf corresponds to the
increased section width due to corrosion cracking and rust expansion. An approximation of the width
increase is given by:

b f bo nbars wcr

(6)

where nbars is the number of the bars in the layer under compression, wcr is the total crack width for a
given corrosion level:

wcr ui ,corr 2 ( rs 1) X

(7)

where rs is the ratio of volumetric expansion of the rust products with respect to the virgin material, X
is the depth of the corrosion attack equal to the reduction in bar radius and ui,corr is the opening of each
single corrosion cracks. In the present study rs is assumed equal to 2 as proposed in Simioni (2009).

S.T. Karapetrou, S.D. Fotopoulou, K.D. Pitilakis/ VEESD 2013

Table 3 Loss in reinforcement, concrete cover strength and steel ultimate deformation reduction for the
considered corrosion scenarios
t (years)

25

50

75

Beam Column

Beam Column

Beam Column

Low-rise/No code

5.0

5.0

12.0

11.0

18.0

16.0

Mid-rise/High code

4.0

3.0

11.0

9.0

19.0

15.0

High-rise/Low code

6.0

5.0

14.0

11.0

21.0

18.0

Low-rise/No code

54.0

40.0

74.0

61.0

82.0

71.0

Cover strength reduction (%) Mid-rise/High Code

41.0

28.0

70.0

56.0

80.0

68.0

High-rise/Low code

53.0

33.0

73.0

54.0

81.0

65.0

Steel area loss (%)

Steel ultimate
deformation reduction (%)

Low-rise/No code

9.0

20.0

33.0

Mid-rise/High Code

6.0

20.0

33.0

High-rise/Low code

10.0

24.0

35.0

Regarding the effects of corrosion on steel mechanical properties the loss of steel ductility is taken
into account through the reduction of steel elongation at maximum load. Thus the reduction of the
steel ultimate deformation su is calculated based on linear interpolation of the experimental results by
Rodriguez et al. (2001), where the reduction reaches values of 30% and 50% for losses of
reinforcement cross-sectional area of 15% and 28% respectively.
The distribution of the loss of reinforcement area as well as the reduction in concrete cover strength
due to corrosion of the RC elements for the considered corrosion scenarios (t=25, 50, 75 years) are
calculated as a function of the corrosion rate and the corrosion initiation time variables. Table 3
summarizes the mean percentages of reinforcement area loss, cover concrete strength and steel
ultimate deformation reduction due to corrosion within the elapsed time (t-Tini). The estimated COVs
vary from 0.04 to 0.08 and from 0.25 to 0.50 for the loss of reinforcement and cover strength
reduction variables respectively. Overall, for a given corrosion scenario beams are shown to be more
seriously affected compared to columns. An increase of the initial fundamental period of the structures
(Table 1) is expected since corrosion effects cause progressive stiffness degradation as well. The
structural models under study present a percentage increase in the natural period that varies between 68 % for the transition from 0 years to 25 years, 6-9% from 25 to 50 years and 11-18% from 50 to 75
years, depending on the characteristics of the initial and corroded structures.
2.3 Seismic input motion

The next step is the appropriate selection of a representative set of accelerograms that will
subsequently be used for non-linear incremental dynamic analysis and will provide the necessary
response statistics for the time-dependent fragility analysis. The selected scenario earthquake consists
of a suite of 15 real ground motion records obtained from the European Strong-Motion Database
(http://www.isesd.hi.is) (Table 4). They are all referring to outcrop conditions recorded at sites
classified as rock according to EC8 (soil type A) with moment magnitude (Mw) and epicentral distance
(R) that range between 5.8<Mw<7.2 and 0<R<45km respectively. Outcropping records are selected to
avoid uncertainties related to soil effects. Additionally in order to eliminate potential source of bias in
structural response, the selection of pulse-like records has been avoided. The primary selection
criterion is the average acceleration spectra of the set to be of minimal epsilon (Baker and Cornell
2005) at the period range of 0.00<T<2.00sec with respect to a reference spectra defined based on the
ground motion prediction equation (GMPE) proposed by Ambraseys et al. 1996 corresponding to the
median of the Mw and R selection bin. The optimization procedure is performed by making use of the
REXEL software (Iervolino et al. 2010). Figure 2 depicts the mean normalized elastic response
spectrum of the records in comparison with the corresponding median predicted spectrum of
Ambraseys et al. (1996). As shown in the figure, a good match between the two spectra is achieved.

S.T. Karapetrou, S.D. Fotopoulou, K.D. Pitilakis/ VEESD 2013

Table 4 List of records used for the IDA


Earthquake Name

Station ID Date

Mw

R (km)

PGA (m/s2)

Waveform code

Friuli

ST20

6/5/1976

6.5

23

3.499

000055xa

Montenegro

ST64

15/4/1979

6.9

21

1.774

000198xa

Montenegro (aftershock) ST68

24/5/1979

6.2

30

0.667

000234xa

Valnerina

ST225

19/9/1979

5.8

1.51

000242xa

Valnerina

ST61

19/9/1979

5.8

22

0.6

000246xa

Campano Lucano

ST93

23/11/1980

6.9

23

1.363

000287xa

Lazio Abruzzo

ST140

7/5/1984

5.9

0.985

000365xa

Lazio Abruzzo

ST143

7/5/1984

5.9

22

0.628

000368xa

Golbasi

ST161

5/5/1986

29

0.538

000410ya

Golbasi

ST161

6/6/1986

5.8

34

0.167

000412xa

Izmit (aftershock)

ST575

13/9/1999

5.8

15

0.714

001243xa

Mt. Vatnafjoll

ST2483

25/5/1987

42

0.131

005271ya

Kozani

ST1320

13/5/1995

6.5

17

1.396

006115ya

South Iceland

ST2497

17/6/2000

6.5

34

0.386

006269xa

Firuzabad

ST3293

20/6/1994

5.9

39

0.216

007158xa

Figure 2. Normalized average elastic response spectrum of the input motions in comparison with the
corresponding reference spectrum proposed by Ambraseys et al. (1996).

3 INCREMENTAL DYNAMIC ANALYSIS IDA


3.1 Performing IDA

The IDA procedure is used to determine the seismic performance and finally to assess the timedependent seismic vulnerability of the initial and corroded structures.
IDA is an emerging analysis method which involves performing a series of nonlinear dynamic
analyses under a suite of multiply scaled ground motion records whose intensities should be ideally
selected to cover the whole range from elasticity to global dynamic instability (Vamvatsikos and
Cornell 2002). IDA curves of the structural response, which provide a relationship between a damage
measure quantity (i.e. engineering demand parameter EDP) and a scalable intensity measure (IM) of
the applied scaled accelerograms, are then constructed by interpolating the resulting EDP-IM discrete

S.T. Karapetrou, S.D. Fotopoulou, K.D. Pitilakis/ VEESD 2013

Figure 3. IDA curves for the high-rise, low code structure for the different points in time.

points. Within the framework of this study, the damage measure is expressed in terms of maximum
interstory drift ratio, max, which is known to relate well to dynamic instability and structural damage
of frame buildings. The intensity measure is described by the 5%-damped spectral acceleration at the
fundamental period of the structure [Sa(T1,5%)]. The latter is found to be both efficient and sufficient
for first-mode dominated, moderate period structures (Shome and Cornell 1999) as the ones presented
in this research. IDA for the three structural models is conducted by applying the 15 progressively
scaled records for the selected time periods t=0, 25, 50 and 75 years. An advanced tracing algorithm,
namely the hunt & fill (Vamvatsikos and Cornell 2002; 2004), which ensures that the records are
properly scaled with the minimum required computational effort, is used to perform the IDA. In
particular, we allow a maximum of 12 runs for each record with an initial step of 0.1 g, a step
increment of 0.05g and a first elastic run at 0.005g and 0.01g for the low and high code structures
respectively. By interpolating the derived pairs of Sa(T1,5%) and max for each individual record we
get 15 continuous IDA curves for each structural model and time scenario.
Figure 3 presents indicative plots of the derived IDA curve for each record and the corresponding
summarized across all records IDA curves at 16%, 50% and 84% fractiles for the high-rise, low code
structure for the different analyzed time periods (0, 25, 50 and 75 years). Despite the fact that
Sa(T1,5%) is proven to be a more efficient IM compared to others (e.g. PGA), we observe that IDA
curves still display a considerable record-to-record variability with dispersion in the order of 20%40% depending on the characteristics of the initial and corroded structures.
3.2 Limit damage states

The selection of well-defined and realistic damage (or performance) states is a key issue in the
development of seismic fragility curves. For the purpose of the present study, two limit states are
defined in terms of maximum interstory drift ratio, max, representing the immediate occupancy IO and
collapse prevention CP (or near collapse) performance levels. The first limit state is defined at 0.5%
according to HAZUS prescriptions (NIBS 2004) for RC moment resisting frame structures whereas

S.T. Karapetrou, S.D. Fotopoulou, K.D. Pitilakis/ VEESD 2013

Table 3 CP limit state max values defined on the IDA curve for each structural model over time
Time (years)

Low rise MRF-no code

High rise MRF-low code

Mid rise MRF-high code

0.028

0.0225

0.039

25

0.025

0.021

0.037

50

0.024

0.020

0.033

75

0.021

0.017

0.030

the second is assigned on the median (50%-fractile) IDA curve. The main idea is to place the CP limit
state at a point where the IDA curve is softening towards the flatline but at low enough values of max
so that we still trust the structural model (Vamvatsitkos and Cornell 2004). Different CP limit state
values are thus chosen on the IDA curve for the same structure depending on the considered timedependent corrosion scenario. Table 3 summarizes the time-variant CP limit state values adopted for
each structural model.

4 TIME-DEPENDENT FRAGILITY CURVES

A fragility curve represents a graphical relationship of the probability of exceeding a predefined level
of damage (e.g. IO, CP) under a seismic excitation of a given intensity. The results of the IDA
(Sa(T1,5%) - max values) are used to derive time-dependent fragility curves expressed as twoparameter time-variant lognormal distribution functions:

In IM - In IM t
P DS / IM

(8)

where, is the standard normal cumulative distribution function, IM is the intensity measure of the
earthquake expressed in terms of Sa(T1,5%) (in units of g), IM t and (t) are the median values (in
units of g) and log-standard deviations respectively of the building fragilities at different points in time
along its lifetime and DS is the damage state. The median values of Sa(T1,5%) corresponding to the
prescribed performance levels are determined based on a regression analysis of the nonlinear IDA
results (Sa(T1,5%) - max pairs) for each structural model and time scenario (t=0, 25, 50 and 75 years).
More specifically, in accordance to previous studies (e.g. Cornell et al. 2002), a linear regression fit of
the logarithms of the Sa(T1,5%) -max data which minimizes the regression residuals is adopted in all
analysis cases.
The various uncertainties are taken into account through the log-standard deviation parameter (t),
which describes the total dispersion related to each fragility curve. Three primary sources of
uncertainty contribute to the total variability for any given damage state (NIBS 2004), namely the
variability associated with the definition of the limit state value, the capacity of each structural type
and the seismic demand. The uncertainty in the definition of limit states is assumed to be equal to 0.4
while the variability of the capacity is assumed to be 0.25 and 0.3 for the high and no/low code
structures respectively (NIBS 2004). The third source of uncertainty associated with the demand, is
taken into consideration by calculating the dispersion of the logarithms of Sa(T1,5%) - max simulated
data with respect to the regression fit. Under the assumption that these three variability components
are statistically independent, the total variability is estimated as the root of the sum of the squares of
the component dispersions. The herein computed log-standard deviation (t) of the curves varies from
0.57 to 0.69 for all structural models and time scenarios.
Figure 4 illustrates the derived graphs of fragility curves for the initial (t=0 years) and corroded (t=25,
50, 75 years) buildings. It is clearly seen that the vulnerability of the structures increases over time due
to the corrosion. This increase is much more noticeable for the CP limit state (especially for the low/

S.T. Karapetrou, S.D. Fotopoulou, K.D. Pitilakis/ VEESD 2013

Figure 4. Time-dependent fragility curves for the analyzed structures

no seismic code designed buildings). In particular, the change in the median Sa(T1,5%) value for the
CP limit state after 75 years is in the order of 40% for the structures designed with no and low seismic
code provisions whereas this percentage is around 30% for the mid-rise high code structure. For the no
and low code structures this increase in fragility is more pronounced for the transition from 0 years to
25 years and from 50 years to 75 years whereas for the mid-rise high code building the greater shift in
fragility is expected from 25 years to 50 years time period. This observation could be probably
attributed to the different adopted code design criteria and the different corrosion initiation time
calculated for the no/low and high code structures.

5 CONCLUSIONS

The seismic vulnerability of RC frame buildings has been assessed taking into account the
corrosion of the RC structural members. Three fixed-base MRF structures corresponding to different
code design levels and described based on Syner-G taxonomy, are analyzed using 2D incremental
dynamic analysis. Time-dependent probabilistic fragility functions have been derived for the
immediate occupancy and collapse prevention limit states for different periods in time (t=0, 25, 50, 75
years) in terms of Sa(T1,5%). A significant increase in the seismic fragility of the structures is
observed over time due to corrosion, highlighting the importance of considering the deterioration
effects due to aging on the seismic vulnerability of structures. The fragility functions proposed in this
paper can be used to assess the time-dependent vulnerability of a variety of RC typologies considering
aging. Given the significance of this studys findings, future work should also address time-dependent
fragility taking into account different degradation mechanisms such as the cumulative earthquake
damage under successive earthquake shocks during the structures lifetime.

ACKNOWLEDGEMENTS

S.T. Karapetrou, S.D. Fotopoulou, K.D. Pitilakis/ VEESD 2013

10

The work reported in this paper was carried out in the framework of the ongoing REAKT
(http://www.reaktproject.eu/) project, funded by the European Commission, FP7-282862.

REFERENCES
Ambraseys, N.N., Simpson, K.A. and Bommer, J.J. (1996). Prediction of horizontal response spectra in Europe.
Earthquake Engineering and Structural Dynamics 25: 371-400.
Baker, J.W. and Cornell C.A. (2005). A vector valued ground motion intensity measure consisting of spectral
acceleration and epsilon. Earthquake Engineering and Structural Dynamics 34:1193-1217
Bracci, J.M., Reinhorn, A.M. and Mander, J.B. (1992). Seismic resistance of reinforced concrete frame
structures designed only for gravity loads: Part I Design and properties of a one-third scale model
structure. Technical Report NCEER-92-0027.
Choe, D.E., Gardoni, P., Rosowsky, D. and Haukaas, T. (2009). Seismic fragility estimates for reinforced
concrete bridges subject to corrosion. Structural Safety 31: 275-283.
Cornell, C.A., Jalayer F., Hamburger R.O. and Douglas A.F. (2002). Probabilistic basis for 2000 SAC federal
emergency management agency steel moment frame guidelines. Journal of Structural Engineering 128(4):
526-533.
Coronelli D. and Gambarova P., 2004. Structural assessment of corroded reinforced concrete beams: modeling
guidelines. Journal of Structural Engineering 130(8): 1214-1224.
DuraCrete (1998). Modeling of Degradation, BRITE-EURAM-Project BE95-1347/R4-5.
DuraCrete (2000). Probabilistic performance based on durability design of concrete structures:
Statistical quantification of the variables in the limit state functions, Report No.: BE 95-1347, 62-63.
CEB-FIB Task Group 5.6 (2006). Model for Service Life Design, fdration internationale du bton (fib).
Fotopoulou, S.D., Karapetrou, S.T. and Pitilakis, K.D. (2012). Seismic vulnerability of RC buildings considering
SSI and aging effects, Proceedings of the 15WCEE International Conference, Lisboa.
Ghosh, J. and Padgett, J.E. (2010). Aging considerations in the development of time-dependent seismic
fragility curves. Journal of Structural Engineering 136(12), 1497-1511.
Iervolino, I., Galasso, C.and Cosenza, E. (2010). REXEL: Computer aided record selection for code-based
seismic structural analysis. Bulletin of Earthquake Engineering 8:339-362.
Kappos, A.J., Panagopoulos G., Panagiotopoulos C.and Penelis G. (2006). A hybrid method for the vulnerability
assessment of R/C and URM buildings. Bulletin of Earthquake Engineering 4:391-419.
Karsan, I. and Jirsa, J. (1969). Behavior of concrete under compressive loadings. ASCE J. of the
Structural Division 95: 2543-2563.
Kent, D.C. and Park, R. (1971). Flexural members with confined concrete, Journal of the Structural Division.
Proc. Of the American Society of Civil Engineers 97(ST7): 1969-1990.
Mazzoni, S., McKenna, F., Scott, M.H. and Fenves, G.L. (2009). Open System for Earthquake Engineering
Simulation User Command-Language Manual. Pacific Earthquake Engineering Research Center,
Berkeley, California.
National Institute of Building Sciences (2004). Direct physical damage General building stock, HAZUSMH Technical manual, Chapter 5. Federal Emergency Management Agency, Washington, D.C.
Rodriquez, J. and Andrade, C. (2001). CONTECVET A validated users manual for assessing the residual
service life of concrete structures. GEOCISA. Madrid, Spain.
Neuenhofer, A. and Filippou, F.C. (1997). Evaluation of nonlinear frame finite-element models. Journal of
Structural Engineering 123(7): 958-966, July.
Shome, N. and Cornell, C. A. (1999), Probabilistic seismic demand analysis of nonlinear structures. RMS
Report-35, Reliability of Marine Structures Group, Stanford University, Stanford.
Simioni, P. (2009). Seismic response of reinforced concrete structures affected by reinforcement corrosion. PhD
thesis. Faculty of Architecture, Civil Engineering and Environmental Sciences University of Braunschweig
Institute of Technology and the Faculty of Engineering University of Florence.
Stewart M. (2004). Spatial variability of pitting corrosion and its influence on structural fragility and
reliability of RC beams in flexure. Structural Safety 26: 453-470.
Vamvatsikos, D. and Cornell, C.A. (2002). Incremental dynamic analysis. Earthquake Engineering and
Structural Dynamics 31:491-514.
Vamvatsikos, D. and Cornell, C.A. (2004). Applied incremental dynamic analysis. Earhquake Spectra 20(2):
523-553.

Anda mungkin juga menyukai