You are on page 1of 13

Engineering Structures 75 (2014) 113–125

Contents lists available at ScienceDirect

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

Finite element investigation of the influence of corrosion pattern on


inelastic buckling and cyclic response of corroded reinforcing bars
Mohammad M. Kashani a,⇑, Laura N. Lowes b, Adam J. Crewe a, Nicholas A. Alexander a
a
University of Bristol, Dept. of Civil Engineering, Bristol BS8 1TR, United Kingdom
b
University of Washington, Dept. of Civil and Environmental Engineering, Seattle, WA 98195-2700, United States

a r t i c l e i n f o a b s t r a c t

Article history: An optical surface measurement technique was used to characterise three-dimensional corrosion pattern
Received 19 November 2013 of reinforcing bars subjected to accelerated corrosion. After the optical measurement process was per-
Revised 3 April 2014 formed, the corroded bar specimens were tested under monotonic and cyclic axial loading. The optical
Accepted 20 May 2014
measurement data were used to develop a 3D micro-fibre finite element model developed for simulation
Available online 19 June 2014
of the physical testing and parametric study of the influence of corrosion pattern on stress–strain
response of corroded bars. It was observed that the irregular cross sectional shape of pitted sections
Keywords:
has a significant influence on the inelastic buckling and nonlinear cyclic response.
Reinforcing steel
Corrosion
Ó 2014 Elsevier Ltd. All rights reserved.
Inelastic buckling
Stress–strain relations
Nonlinear analysis
Cyclic behaviour
Low-cycle high amplitude fatigue

1. Introduction Recent experimental testing of corroded RC elements subject to


cyclic loading showed that corrosion significantly reduces plastic
Corrosion of reinforcing bars in aging reinforced concrete (RC) rotation capacity, and thus drift capacity; this is due to premature
structures and bridges is a critical issue in developed countries buckling and/or fracture of the longitudinal reinforcing bars as well
[1–3]. Billions of dollars are spent annually by government and as premature fracture of horizontal hoop reinforcement that con-
other agencies monitoring, repairing, and replacing RC structures fines core concrete and restrains longitudinal bar buckling [7,8].
with corrosion damage [1–3]. There have been incidences in which In the recent years researchers have investigated the influence
corrosion damage resulted in the complete collapse of in-service of corrosion on the stress–strain behaviour of corroded bars in ten-
structures [3]. Because of this, a substantial body of research in sion [9–14]. The outcome of theses researches showed that corro-
the past decades has focussed on corrosion in RC structures. This sion has significant influence on the ductility and plastic
research has addressed a number of topics including the corrosion deformation capacity of reinforcing bars owing to the level of
process, the behaviour of corroded reinforcement, performance non-uniform pitting. To improve understanding of the impact of
assessment and repair of corroded structures, and maintenance corrosion on inelastic buckling and cyclic response, Kashani et al.
optimisation for transportation structures [4,5]. A relatively small [15,16] conducted a comprehensive experimental investigation of
portion of this research has addressed the impact of corrosion on the influence of corrosion on inelastic buckling and nonlinear cyc-
the seismic performance of RC structures, despite the fact that lic behaviour of corroded reinforcing bars. The experimental
there are many RC structures located in regions of high seismicity results show that the non-uniform distribution of pitting along
and exposed to corrosive environments. the length of corroded bars changes the buckling mechanism and
For RC bridges subjected to earthquake loading, the desired results in more rapid strength loss under cyclic loading. Kashani
inelastic response mechanism is typically flexural yielding of the et al. report that when the mass loss ratio of corroded bars is more
columns or piers, and drift capacity (i.e. the drift at which than approximately 25%, pitting corrosion becomes more localised.
significant lateral strength loss occurs) is typically determined by Since the smallest cross section of the bar determines the yield
buckling followed by fracture of longitudinal reinforcing bars [6]. strength; localisation of pitting corrosion results in premature
yielding of the reinforcement at a relatively low tension demand.
⇑ Corresponding author. Additionally, Kashani et al. [17] show that pitting corrosion results
E-mail address: mehdi.kashani@bristol.ac.uk (M.M. Kashani). in variation in the cross sectional shape of the reinforcement along

http://dx.doi.org/10.1016/j.engstruct.2014.05.026
0141-0296/Ó 2014 Elsevier Ltd. All rights reserved.
114 M.M. Kashani et al. / Engineering Structures 75 (2014) 113–125

the length of the corroded bar. This results in a change in the 2.2. 3D optical measurement of corrosion pattern
moment of inertia of the reinforcing bar and potentially affects
the buckling response of the bar under cyclic loading. Twenty-three (23) randomly selected corroded reinforcing bars
This paper explores the influence of the pattern of corrosion were taken out of the total of 120 samples for more refined geo-
along the length of a reinforcing bar on inelastic buckling and cyc- metrical surface analysis of corrosion patterns. The reinforcing bars
lic response. Data from a previously presented experimental study varied in length (L/D = 5 to L/D = 20) and had a range of mass loss
[15,16] are reviewed with the objectives of establishing the ratios. The surface pitting pattern of the corroded bars was mea-
response of corroded bars under monotonic and cyclic loading sured using a white light scanner followed by a stochastic analysis
and characterising the pattern of corrosion along the length of of the results in MATLAB [19]. Fig. 1 shows an example contour
the reinforcing bar. A computational model, employing nonlinear plot of scanned bar. The surface area of the bar is represented by
beam–column elements with fibre-type cross-section models and the length along the bar (0–200 mm) and the angle of rotation
nonlinear material models, is proposed for simulating the response (p to p) from an arbitrary plane; the impact of corrosion is repre-
of a corroded reinforcing bar. The variation in the geometrical sented by the radius from the original centroid of the bar to the
properties of corroded bars is modelled using the optical measure- surface of the corroded bar (2.0–6.0 mm). Fig. 2 also shows an
ment data provided in [17]. Moreover, for the first time, the varia- example of cross sections through the solid model of a corroded
tion in load eccentricity due to pitting corrosion along the buckling bar with 54.23% mass loss. A detailed discussion of the optical
length of bars is also included in the finite element model. The measurement process and analysis of optical data is provided in
model is validated through comparison with experimental data. [17]. Once the corrosion pattern analysis completed the 23
The model and experimental data are used to investigate the scanned bars have been tested under monotonic and cyclic loading.
impact of the corrosion pattern on the buckling strength, tensile
yield strength and cyclic response of a reinforcing bar. Finally, 2.3. Observed response under monotonic loading
experimental data are used to evaluate existing one-dimensional
constitutive models for reinforcing steel, which simulate compres- The experimental results show that the non-uniform pitting
sive strength loss due to buckling and the impact of buckling and corrosion changes the buckling mechanism of corroded bars. The
low-cycle fatigue on cyclic response. pitted sections have irregular shapes that vary over the entire
length of the bar [15]. This irregularity creates strong and weak
2. Corrosion induced mechanical–geometrical degradation of axes, for which the second moments of area is reduced from the
reinforcing bars uncorroded state, and creates load eccentricity that results in a sig-
nificant reduction in buckling capacity for corroded bars. Fig. 3
Kashani et al. [15] tested a series of corroded reinforcing bars shows examples of mean stress versus strain data for corroded bars
under monotonic and cyclic loading. Experimental data character- tested in the laboratory under monotonic compression loading.
ise the average stress–strain response of the bars under monotonic The mean stress presented in Fig. 3 was computed as the applied
and cyclic loading; such data can be used to develop and validate load divided by the average reduced area of the bar assuming uni-
the 1D constitutive models for reinforcing steel used typically in form mass loss over the unsupported length (L) of the bar.
fibre-type models of reinforced concrete flexural elements. Exper- The observed responses in Fig. 3 show that considering only the
imental data characterise also the pattern of corrosion along the average reduction in bar cross sectional area is not sufficient for
length of the bars, from which variation in the bar area and first characterising the impact of corrosion on stress–strain response.
and second moments of area are computed. Other parameters such as the minimum area, centroid of the sec-
tion, and minimum second moment of area must be considered
2.1. Summary of experimental programme in constitutive modelling of corroded bars.

To develop the corroded reinforcement specimens for testing, a 2.4. Observed response under cyclic loading
total of eight reinforced concrete prisms were cast. Each specimen
was dimensioned 250  250  700 mm and incorporated 8 No. The experimental data of the cyclic tests support the following
12 mm diameter reinforcing bars. An accelerated corrosion simula- observations of interest to the current work:
tion technique (known as anodic corrosion) was employed to accel-
erate the corrosion time in the laboratory environment. Different  It was found that the cyclic degradation of the reinforcing bars
prisms were subjected to different corrosion times to achieve differ- is very sensitive to the strain history. The buckling strength is
ent levels of reinforcement mass loss due to corrosion. After corro- also significantly affected by the strain history and the tension
sion simulation, the concrete specimens were broken open and the strain amplitude.
corroded bars were carefully removed from the concrete and  Corroded bars subjected to cyclic loading exhibit reduced buck-
cleaned in accordance with ASTM G1-03 [18]. A total of 120 corroded ling strength in comparison with uncorroded bars but also in
bars were prepared. Twenty-three (23) randomly selected corroded comparison with corroded bars subjected to monotonic com-
bars with different mass loss ratios were selected for optical mea- pressive loading. Strength reduction under cyclic loading is
surement and statistical analysis of the corrosion patterns. Fifty-
seven (57) bars were used for monotonic buckling tests and 40 were
used for cyclic tests. A detailed discussion of specimen preparation,
optical measurement and testing is provided in [15–17].
Mechanical testing was conducted for a series of different slen-
derness ratios. The slenderness ratios were chosen based on the
range of horizontal tie spacings commonly used in RC column con-
struction. The slenderness ratio (L/D) is defined as the unsupported
length of the bar (L) divided by the diameter of the reinforcing bar
(D). The L/D ratios tested were 5, 8, 10, 15, and 20 for the monotonic
buckling tests and 5, 10 and 15 for the cyclic tests. For each slender-
ness ratio, three control (uncorroded) specimens were also tested. Fig. 1. Contour plot of corrosion pattern in a corroded bar with 36.40% mass loss.
M.M. Kashani et al. / Engineering Structures 75 (2014) 113–125 115

Fig. 4. View of the critical pitted section in solid model and experimental testing of
a corroded bar with L/D = 15 and 34.26% mass.

finite element analysis of individual reinforcing bars, with different


L/D and yield strengths, to investigate their buckling response.
Fig. 2. Cross section through the solid model of a corroded bar with 54.23% mass
Nakamura and Higai [23] performed nonlinear finite element anal-
loss.
yses studying the stress–strain behaviour of reinforcing bars under
attributed to premature tensile yielding of the corroded bar. monotonic and cyclic loading. In all cases, analyses and experi-
Premature tensile yielding results from a reduction in bar area ments considered isolated bars (i.e. bars were not embedded in
due to corrosion (Fig. 4). concrete); in all cases, good agreement between the experimental
 Under cyclic loading, corroded bars exhibit more rapid strength data and finite element results was observed. Here nonlinear finite
loss with increasing strain demand than do bars without element analysis was employed to investigate the impact of corro-
corrosion. sion on the behaviour of reinforcement under monotonic and cyc-
 Corrosion reduces the strain at which bars fracture in tension lic loading. An approach similar to that employed by Dhakal and
(Fig. 5(a)). It was observed that for bars with higher slenderness Maekawa [21] for analysis of bars without corrosion was
ratios the amount of corrosion needed to cause tension fracture employed.
under cyclic loading is reduced (Fig. 5(a)). This is due to the
effect of inelastic buckling. 3.1. The finite element model
 Corrosion acts to reduce the effective slenderness ratio, L/D, of
the bar. For example, uncorroded bars with L/D = 5 displayed The open-source three-dimensional nonlinear finite element
a symmetric hysteresis response. However, as the level of corro- analysis platform OpenSees (Open System for Earthquake
sion increased the hysteresis response of this group of bars Engineering Simulation) [24] and the force-based distributed
showed a pinching effect in a similar way to the bars with plasticity beam–column (FBBC) element formulation available in
higher slenderness ratios (Fig. 5(b)). the OpenSees platform were used for the analyses [25,26]. This
element formulation assumes a linear moment distribution and
Further discussion of the experimental results is provided in [16]. constant axial force distribution along the length of the element.
Nonlinear material response under flexural-type loading is simu-
3. Micro-fibre nonlinear finite element analysis of reinforcing lated at the section level. For this study, a fibre-type model of
bars the reinforcing bar cross section was used; fibre response was
defined using a nonlinear constitutive model. Moment and axial
Nonlinear finite element analysis has been used by other load at a section of the beam–column element are determined by
researchers to study the behaviour, including buckling, of integrating the fibre stresses over the section. Using the FBBC
reinforcing bars without corrosion under monotonic and cyclic element formulation, generalised nodal displacements are
loading. Mau and El-Mabsout [20] developed a special beam– determined from section deformations using a Gauss-Labotto inte-
column element for nonlinear finite element analysis of inelastic gration scheme. Previous research conducted by [27,28] showed
buckling of reinforcing bars. Using a fibre-type model of the bar that co-rotational formulation is computationally very efficient
cross section, Dhakal and Maekawa [21,22] conducted a nonlinear when used in beam–column elements to model the geometrical

Fig. 3. Observed buckling response of corroded bars under monotonic loading: (a) compression response with L/D = 8 and (b) compression response with L/D = 15.
116 M.M. Kashani et al. / Engineering Structures 75 (2014) 113–125

Fig. 5. Observed nonlinear cyclic response of corroded bars: (a) L/D = 15 and 36.35% mass loss and (b) L/D = 5 and 40.75% mass loss.

nonlinearity. Therefore, co-rotational formulation is implemented material model is used to model the uniaxial stress–strain response
nonlinear beam–column element in the OpenSees and is used in of steel. The GMP model employs a smooth transition curve that
this study. Further detail is available in [27,28]. asymptotically approaches bilinear tension and compression
A displacement control with an adaptive solution algorithm response envelopes (Fig. 5); this enables simulation of the Bauschin-
code is developed using the Tcl code in the OpenSees to run the ger effect. Variables in Fig. 6 are defined in Table 1. Table 1 provides
nonlinear analysis. To check the solution convergence, the norm also the values used in the analyses, which were determined from
of displacement increment is check against a defined tolerance. coupon tests conducted per ASTM E8/E8M-09. It should be noted
This command is implemented in the OpenSees to construct a con- that the ultimate strain (eu) is the strain at the ultimate strength (ru).
vergence test which uses the norm of the left hand side solution
vector of the matrix equation to determine if convergence has been 3.3. Model validation
reached. The tolerance considered is 108 over the maximum of 35
iterations. The analysis starts with Newton–Raphson solution algo- For validation of the computational model, computed and
rithm and if the convergence is not achieved the displacement experimental average stress versus strain results for uncorroded
increment is cut by a factor of 0.1 and if the convergence is not bars were compared. Fig. 7(a–d) shows a comparison between
achieved again the displacement increment is cut by another factor the computed and observed experimental response of monotonic
of 0.1. If the convergence is not achieved by cutting the step sizes tests for uncorroded bars with L/D = 8, 10, 15 and 20 respectively.
twice, then the solution algorithm is changed. The solution algo- A very small deviation between the computed and experimental
rithms used in the adaptive solution strategy in this study are results is seen in Fig. 7(a and b) which is due to the strain harden-
including Newton–Raphson, modified Newton–Raphson, New- ing and spread of plasticity at the plastic hinge location of the bars
ton–Raphson with Line Search and Krylov–Newton algorithms. with L/D < 10 which may not be captured in the fibre model. How-
Further details of algorithm commands are available in [24]. ever, as the buckling length of bars increases the computed and
A series of preliminary analyses were conducted to determine experimental results are almost identical (Fig. 7(c and d)).
model parameters for use in the subsequent analyses. A mesh Fig. 8(1-c) shows a comparison between the computed and
refinement study was conducted at the section and component observed experimental response of cyclic tests for uncorroded bars
levels. It was found that a section mesh comprising 6 fibres in with L/D = 5, 10 and 15 respectively. As shown in Figs. 7 and 8 the
the radial direction and 35 fibres in the circumferential direction numerical model accurately simulates the monotonic and cyclic
resulted in a converged section response. It was found that use of response of reinforcing bars out to relatively large strain demands.
10 mm long elements with three Gauss–Lobatto integration points The simulation results show that the computational model is able
per element resulted in a converged component response. An ini- to capture the pinching effect seen in the bars with L/D > 10 under
tial imperfection of 0.001 mm (1.0  105 to 4.16  106 of L) the large displacement demand. Furthermore, the compressive
was introduced at mid height of the bar to initiate buckling (the
initial imperfect shape was linear). To emulate the boundary con- σ
ditions during the laboratory tests, the rotations and displace-
ments of the bottom node of the bars were fully restrained in the
Eh
model. The axial displacement history applied in the laboratory
experiments was applied in the analyses.

3.2. Steel material model

Dhakal and Maekawa [22] modelled the path-dependent Es


behaviour of isolated reinforcing bars exhibiting buckling. They con-
cluded that the average stress–strain and buckling behaviour of bars
εy ε
in compression and the unloading/reloading behaviour of cyclic
loops are greatly influenced by the accuracy of the basic steel model
used for the analysis. Therefore, to model the cyclic behaviour of
bars correctly, an accurate steel model must be used that is able to
simulate the unloading/reloading behaviour of reinforcing steel
including the Bauschinger effect. In this study the Giuffre–Menegot-
to–Pinto (GMP) model [29] implemented in OpenSees as the steel02 Fig. 6. The Giuffre–Menegotto–Pinto model.
M.M. Kashani et al. / Engineering Structures 75 (2014) 113–125 117

Table 1 of area, an equivalent rectangular cross section was used in the


Mechanical properties of uncorroded reinforcement. fibre model of a corroded bar. The equivalent section had the same
Reinforcement type 12 mm (B12) area and the minimum second moment of area as calculated from
Yield strength ry (MPa) 540 optical measurement data. To validate this assumption, initially,
Modulus of elasticity Es (MPa) 210,000 equivalent circular and rectangular cross section shapes were
Hardening ratio Eh/Es 0.008 considered for an uncorroded bar. It was observed that the cross
Ultimate strength ru (MPa) 616 section shape has no effect on the computational response (Fig. 9).
Ultimate strain at ultimate strength eu 0.06
Strength ratio ru/ry 1.14
As shown in Fig. 2 the cross section shape of corroded bars can
be considered as s a simple a non-self intersecting polygon. There-
strength degradation of bars under cyclic loading due to the influ- fore, using Green’s Theorem the area of a simple polygon with n
ence of tension history observed in the experiment is also captured vertices can be calculated as defined in Eq. (1).
accurately in the computational simulation. For bars with L/D = 10
1Xn1
 
and 15 subjected to cyclic loading, crack propagation caused by A¼ y zi1  yiþ1 zi ð1Þ
2 i¼0 i
low-cycle fatigue in bars resulted in degradation of the tension
envelope and ultimate tensile strength; the numerical model was where y and z are the coordinate of the vertices.
not able to capture this response as strength degradation due to Subsequently by taking moments about the centre of the
low-cycle fatigue is not incorporated in the steel02 material model. polygon, the centroid and second moment of area about y and z
axis can be calculated using Eqs. (2)–(6).
3.4. Modelling the effect of corrosion pattern
1 X
n1
 
Cy ¼ ai yi þ yiþ1 ð2Þ
As it was shown in Figs. 2 and 4, pitting corrosion creates irreg- 6A i¼0
ular cross section shapes along the length of corroded bars. This
irregularity creates strong and weak axes, for which the second 1 X
n1

moments of area is reduced from the uncorroded state, and creates Cz ¼ ai ðzi þ ziþ1 Þ ð3Þ
6A i¼0
load eccentricity at the cross section. These factors must be
included in analysis of isolated corroded reinforcing bars.
1 X
n1
 
Iy ¼ ai z2i þ zi ziþ1 þ z2iþ1 ð4Þ
3.4.1. Modelling the cross section with equivalent rectangular 12 i¼0
As it was discussed in previous sections, nonuniform corrosion
affects both area of cross section and second moment of area due to 1 X
n1
 
the change in the cross section shape. Circular section has only one Iz ¼ ai y2i þ yi yiþ1 þ y2iþ1 ð5Þ
12 i¼0
parameter (radius), so that, it is not possible to fit two parameters
(area and second moment of area) in a circular section. Therefore,  
ai ¼ yi zi1  yiþ1 zi ð6Þ
to model the impact of corrosion on both area and second moment

Fig. 7. Comparison of fibre model simulation with monotonic test results: (a) L/D = 8, (b) L/D = 10, (c) L/D = 15 and (d) L/D = 20.
118 M.M. Kashani et al. / Engineering Structures 75 (2014) 113–125

Fig. 8. Comparison of fibre model simulation with cyclic test results: (a) L/D = 5, (b) L/D = 10 and (c) L/D = 15.

were used as rigid link elements (RGE) to connect the nonlinear


elements.
The load eccentricity is considered as the product of the cen-
troid in y and z axes as defined in Eq. (7):
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
eðxÞ ¼ cy ðxÞ2 þ cz ðxÞ2 ð7Þ

where e(x) is the coefficient of load eccentricity ratio as a function of


length x and cy(x) and cz(x) are the section centroid in the y and z
axes respectively. The load eccentricity is used together with the
minimum principal second moment of area (considering axis
rotation) and applied to the weak axis of the equivalent rectangular
section in the fibre model. Fig. 10 shows an idealised view of the
finite element model in the OpenSees [24] including the RGEs
arrangement. The elastic beam–column elements were defined to
have an axial and flexural stiffness (EA and EI) twenty times higher
Fig. 9. Comparison of the nonlinear response of beam–column element models than the axial and flexural stiffness of the original uncorroded
with circular and equivalent rectangular fibre sections. reinforcing bars (a 12 mm diameter bar).

The above equations are used to calculate the geometrical proper-


3.4.3. Mesh sensitivity and averaging length
ties of corroded bars and used in the fibre model.
The optical measurement data defined the corroded bar cross
section at 0.5 mm intervals. However, the results of stochastic
3.4.2. Modelling the load eccentricity corrosion pattern analysis using autocorrelation and cross
Pitting corrosion results also in shifting of the centroid of the correlation functions showed that averaging the data over 10 mm
section along the length of the bar. Since load is applied through segments yields a correlation factor above 70% (the detailed dis-
the average centroid of bar, at any location along on the length cussion is available in [17]). This suggests that the variation of
of the bar, the axial load may be applied eccentrically. This load the cross section within 10 mm of any given point along the bar
eccentricity creates an imperfection in the corroded bars and sig- is relatively small and that meshing the corroded bar using
nificantly affects the inelastic buckling response. 10 mm long elements with constant cross sections could be
To simulate the shifting centroid of the corroded bar, the nodes expected to yield accurate results. Since the computational
of the nonlinear beam–column elements were located at the mea- demand of modelling with 0.5 mm is significantly greater than
sured centroid of the section and elastic beam–column elements modelling using 10 mm long elements, the impact of mesh refine-
M.M. Kashani et al. / Engineering Structures 75 (2014) 113–125 119

P 4. Discussion of computational results and comparison with


observed experimental results

4.1. Buckling response under monotonic loading

To investigate the factors that affect the buckling strength of


corroded bars, experimental and simulation data were compared
for specimens with different L/D ratios and mass loss ratios and
for models with and without representation of section eccentricity
Rigid link elements resulting from corrosion. Models varied only in representation of
to represent the eccentricity; all numerical models represented the loss in section
eccentricity area and second moment of area resulting from corrosion.
The data in Fig. 12(a) shows load–displacement response for a
bar with L/D = 20 and a mass loss ratio of 34.91%. These data show
10mm fibre that if the reduction in cross section area and second moment of
beam-column
area are included in the simulation, a 46% reduction in peak buck-
elements
ling load is predicted; if these factors as well as load eccentricity
are included in the simulation, a 73% reduction in peak buckling
strength is predicted. The data show also that even with load
eccentricity included in the model, predicted buckling strength is
18% greater than measured strength. This is attributed to irregular
cross sectional shape of pitting sections. This will result in stress
concentration at pitted sections and subsequently premature
Fig. 10. Schematic representation of fibre beam–column model of a corroded bar yielding of the section. This phenomenon then results in an addi-
with eccentricity. tional imperfection in the bar under compression which affects
the buckling load capacity. As expected, it was observed that the
influence of cross section shape is a function of mass loss and slen-
ment was investigated. The nonlinear buckling responses of two
derness ratio. Finally, the data in Fig. 12(a) shows that while the
finite element models, one with 10 mm and one with 5 mm ele-
model over predicts maximum buckling strength, post-buckling
ments, was compared (Fig. 11). In both models, a constant cross
response is accurately predicted using the model with or without
section was assumed for the entire element length, cross section
simulation of load eccentricity.
geometric properties (area, centroid and second moment of area)
Fig. 12(b and c) shows simulated and experimental response for
were defined as the average over the element length, and three
bars with L/D = 15 and 10, respectively; in these figures, simulation
Gauss Lobatto integration points per element were used. As it is
results include consideration of the impact of corrosion on area,
shown in Fig. 11, reducing the element length from 10 mm to
second moment of area and load eccentricity. For these tests, max-
5 mm did not significantly affect the nonlinear buckling response
imum simulated strength exceeds measured strength by 16% for L/
of a corroded bar with L/D = 20 and 34.91% mass loss. It is also find
D = 15 and 10% for L/D = 10. This shows that the difference
very small element lengths results in numerical instability at post-
between the simulated and experimental results reduced by reduc-
buckling response (Fig. 11). Therefore, using 10 mm long elements
ing the slenderness ratio of the corroded bars. However, as the
are reasonably accurate and also computationally efficient.
slenderness ratio of bars reduces the contribution of cross section
shape to the overall reduction of buckling load increases. In other
words, as the slenderness ratio of the bars increases the impact
3.4.4. Reference model of imperfection induced by the combined cross sectional shape
Preliminary analyses resulted in the development of a reference effect and second moment of area on the buckling load increases.
model used for all subsequent analyses. This reference model Comparison of Fig. 12(b and c) with Fig. 12(a) shows that bigger
employed (i) geometric properties (area, centroid and second mass loss ratio results in more sever localised pitting corrosion and
moment of area) were averaged over a 10 mm length, (ii) 10 mm subsequently degrades the geometrical properties of pitted sec-
long force-based beam–column elements with three Gauss– tions more significantly. This will increases the imperfection in
Lobatto integration points, and (iii) RGE to link element end nodes the bar due to effect of second moment of area and cross section
and represent eccentricity of the corroded section. shape that has more significant effect on the bars with bigger slen-
der ratio. Whereas in the shorter bars, the influence of second
moment of area is less visible and the main parameter governing
the buckling load reduction is the premature yielding of pitted sec-
tions due to stress concentration. Finally, the data in Fig. 12(a–c)
shows that while corrosion reduces buckling strength, it does not
significantly affect the shape of the post-peak response curve and
that the numerical model, with or without simulation of eccentric-
ity, accurately simulates post-peak response.

4.2. Buckling response under cyclic loading

Experimental and computational results are compared for cor-


roded bars with different corrosion patterns along the length of
the corroded bars. Here the uniformity of the corrosion pattern is
Fig. 11. Comparison of the nonlinear response of fibre models with 5 mm and quantified by considering the ratio of the minimum cross section
10 mm elements. area to the maximum cross section area along the gauge length
120 M.M. Kashani et al. / Engineering Structures 75 (2014) 113–125

Fig. 12. Comparison of the simulated and experimental results of corroded bars: (a) corroded bar with L/D = 20 and 34.91% mass loss, (b) corroded bar with L/D = 15 and
19.93% mass loss and (c) corroded bar with L/D = 10 and 14.44% mass loss.

Fig. 13. Comparison of the uniform and localised pitting: (a) corroded bar with L/D = 10 and 38.97% mass loss, (b) corroded bar with L/D = 15 and 34.26%.

defined in Eq. (8). Given the stress–strain response is the average loss with f = 0.80 and Fig. 13(b) shows a localised pitted specimen
strain over the buckling length, therefore in this study, the buck- with f = 0.67.
ling length is considered as the gauge length. Fig. 14(a) shows the experimental and simulation results for a
corroded with a relatively uniform pattern of pitting corrosion,
Minimum Cross SectionArea ðAmin Þ (specimen in Fig. 13(a)); Fig. 14(b) shows results for a corroded
f¼ ð8Þ
Maximum Cross SectionArea ðAmax Þ with a relatively non-uniform pattern of pitting corrosion, (speci-
men in Fig. 13(b)). Fig. 14 shows that for a bar with a relatively uni-
If f > 0.8, the specimen considered as being uniformly corroded and form pattern of corrosion, the model provides reasonably accurate
if f < 0.8, it is considered as specimen with localised pitted section. simulation of response, with two exceptions. The numerical model
This assumption is in good agreement with the observed experi- over predicts compressive strength by as much as 10%; as with
mental results and the results reported by other researchers monotonic loading, this is attributed to 15%. As cyclic displacement
[10,11]. Fig. 13 shows the variation of the cross section area of demands increase, the experimental data show deterioration in
two specimens with relatively similar mass loss ratio and different tensile strength and, ultimately, complete strength loss due to
corrosion patterns. Fig. 13(a) shows a uniformly distributed area bar fracture. Strength loss is attributed to low-cycle fatigue; exper-
M.M. Kashani et al. / Engineering Structures 75 (2014) 113–125 121

Fig. 14. Comparison of the simulated and experimental results of a corroded bar under cyclic loading: (a) L/D = 10 with 38.97% mass loss and (b) L/D = 15 with 34.26% mass
loss.

imental result shows that corrosion reduces the low-cycle fatigue 5. Critical review of the existing analytical models and
life of the reinforcement. Strength degradation due to low-cycle comparison with the experimental and computational results
fatigue is not captured by the numerical model; thus degradation
of tensile strength is not simulated. In the previous sections the influence of different parameters on
When the irregularity of corrosion along the bar increases, the inelastic behaviour of corroded bars under monotonic and cyclic
differences between the computed and experimental responses loadings are discussed. As it was discussed in the previous sections,
increase. Fig. 14(b) shows the computed and experimental response the irregularity in shape and distribution of pitted sections along
for a bar with severe localised pitting corrosion (f = 0.67). Simula- the bars has a significant impact on the buckling and fracture
tion data are provided for the reference model described above as mechanism of corroded bars. Due to the random nature of corro-
well as for a model in which the cross section of the bar is constant sion phenomenon and significant uncertainties associated with
over the height of the bar and the minimum measured bar area and the corrosion pattern, it is almost impossible to quantify the irreg-
second moment of area are used. Using the reference model, tensile ularity of pitted sections with a deterministic approach. Moreover,
and compressive strength are overestimated. Using the minimum- in the real world, it is very challenging to measure the area loss and
section model, computed response is in a very good agreement with the locations of maximum/minimum pitted sections in inspection
the observed experimental response. This supports the conclusion of bridges using a non-destructive method; i.e. in a bridge pier.
of [16], that the nonlinear cyclic behaviour of corroded bars with However, with the help of recent technologies in bridge inspec-
a localised pitting corrosion is mainly governed by the inelastic tions, it is possible to measure the corrosion rate and any given
response of the minimum pitted section. location in a RC component [3]. Using the corrosion rate measure-
A view of the solid model and fracture mechanism of the cor- ment data it is very simple to calculate the average mass loss of
roded bar analysed in Fig. 14(b) are previously shown in Fig. 4. It reinforcing bars. Therefore, it is very useful to relate all the models
is evident from Figs. 4 and 14(b) that the plastic hinging mecha- to a single measure parameter such as percentage mass loss. To do
nism of the bar starts at the location of the minimum pitted section this, an empirical approach using the experimental results should
which then follows by a severe inelastic buckling in compression. be used to model the average behaviour of corroded bars. Using
This mechanism results in a significant stress concentration at this method it is possible to adopt a pseudo stress–strain curve
the minimum pitted section which subsequently increases the as a function of the average percentage mass loss of corroded bars.
total strain amplitude at this location during the cyclic tests. This Using this method the yield and ultimate strengths are adjusted
will cause the premature fracture of bar in tension due to reduced empirically to account for pitting effects. This is sufficient to inves-
low-cycle high-amplitude fatigue life. tigate the global structural response of bridges/structures under
It should be noted that the location of the minimum section is vehicle and/or earthquake loading. If further detailed modelling
also very important. In the test specimen shown in Figs. 4 and is required, (for example detailed model of corroded RC bridge
14(b), the minimum section is located very close to the mid height pier) it is then possible to model the spatial variability of corrosion
of the bar. This resulted in initiation of the buckling at that location. pattern probabilistically using the models developed in [17].
Another example is shown in Fig. 15 where f = 0.70 and the mini- In this section a comprehensive review of the existing analytical
mum section is located very close to the grip. Despite the localised models of reinforcing bars with the effect of buckling and low-
pitted section in this specimen the response is not governing by cycle fatigue is conducted. A comparison between the observed
the minimum section (Fig. 15(a)). It is clearly shown in Fig. 15(b experimental response, detailed FE analysis and the empirical ana-
and c) that a plastic hinge has been formed at the location of the min- lytical models is conducted and critical issues are reported.
imum section but the bar did not fracture at that location. This cor-
rosion pattern resulted in an unsymmetrical buckling mode shape as
reported previously by Kashani et al. [15,16]. Due to the increase in 5.1. Analytical modelling of post-yield buckling behaviour of corroded
the local strain amplitude at the buckling location the low-cycle fati- bars
gue life of this specimen is shortened. As a result the specimen frac-
tured at the buckling location instead of the minimum section. It is In recent decades the nonlinear analysis of RC framed structures
evident from Fig. 15(a) that in this case the numerical model using subject to seismic loading has received a lot of attention. This has
the minimum cross section is underestimating the capacity and been focused on the development of the fibre element technique
shows more pinching effect in the response. However, the detailed [25,26]. In this approach the member cross section is decomposed
model shows a very good fit to the observed experimental response. into a number of steel and concrete fibres at selected integration
122 M.M. Kashani et al. / Engineering Structures 75 (2014) 113–125

Fig. 15. Comparison of the simulated and experimental results of a corroded bar under cyclic loading: (a) L/D = 10 with 15.48% mass loss, (b) fractured bar at the end of the
cyclic test and (c) variation of the cross section area along the bar.

points. The material nonlinearity is represented through a uniaxial


constitutive material model of steel (tension and compression) and
η = σ σy (1,1) (η 2 ,ξ 2
*
)
concrete (confined core concrete and unconfined cover concrete).
Therefore a lot of researchers around the world put efforts to
(η2 , ξ2 )
develop accurate uniaxial constitutive models for reinforcing steel
and concrete to improve the accuracy of response prediction mod-
els. Moreover, buckling of vertical reinforcing bars is one of the
most common observed collapse mechanisms of RC structures in
the past earthquakes which ultimately results in crushing of core 0.2
concrete in RC columns. Therefore, in the recent years several
researchers have put efforts to develop analytical models to cap- ξ = ε εy
ture the buckling behaviour of vertical reinforcing bars within
the RC elements [30–34]. However, most of the previous analytical Fig. 16. The proposed Dhakal and Maekawa buckling model.

models ignore the interaction of the vertical reinforcement and


where ry is the yield strength of reinforcement in MPa.
horizontal tie reinforcement. Dhakal and Maekawa [21] developed
Given non-dimensional stress g = r/ry and strain n = e/ey
a new analytical model to model the post-yield buckling behaviour
(where ry is the yield stress and ey is the yield strain) the following
of reinforcing bars with different L/D ratios. The analytical model
equations define the stress–strain envelope shown in Fig. 16.
has a base curve considering the L/D is equal to the ratio of spacing
8
of horizontal ties to diameter of the vertical reinforcement in RC > n; n61
>
> ðg2 1Þ
columns. The interaction of horizontal tie reinforcement and verti- < ðn  1Þ þ 1; 1 < n 6 n2
ðn2 1Þ
cal reinforcement is then captured by replacing the L/D ratio of the g¼ ð10Þ
>
> g  0:02; ðn  n2 Þ; n2 6 n&g P 0:2
basic curve to the actual effective buckling length considering the >
: 2
stiffness of horizontal ties. The detail of buckling length calculation 0:2; otherwise
is reported by Dhakal and Maekawa in [35].
where, the empirical relationships for (g2, n2) are given below:
In this research the Dhakal and Maekawa buckling model
(Fig. 16) is used for modelling the post-yield buckling behaviour n2 ¼ 55  2:3kp ; n2 P 7 ð11Þ
of corroded reinforcing bars.
In this model, the post-yield buckling response of reinforcement g2 ¼ að1:1  0:016kp Þg2 ; g2 P 0:2 ð12Þ
is defined as a function of a compound variable called the non-

dimensional slenderness ratio kp as defined in Eq. (9): where g 2 is the non-dimensional piecewise stress corresponding to the
rffiffiffiffiffiffiffiffiffi n2. The value of a is a softening coefficient and depends on the strain
ry L hardening of reinforcement. Dhakal and Maekawa found that for elas-
kp ¼ ð9Þ
100 D tic-perfectly plastic reinforcement a = 0.75 and for reinforcement with
M.M. Kashani et al. / Engineering Structures 75 (2014) 113–125 123

linear hardening a = 1.0. The detailed description of the model is avail-


able in [15,21].
To model the post-yield buckling behaviour of corroded bars,
Kashani et al. [15] proposed a methodology to modify the ry and
kp using the experimental data. The detail of the model calibration
is available in [15]. In this paper the results of the calibrated model
based on the experimental data is compared with the results of
computational models and the corresponding experimental
results. It should be noted that experimental data is the result of
the physical testing of the set of corroded bars which were scanned
and used in the computational modelling and were not included in
the calibration of the buckling model. Fig. 16 shows an example
comparison of the corrosion extended Dhakal–Maekawa buckling
model with experimental and computational results. It is evident
from Fig. 17 that the proposed analytical model has a better fit
Fig. 18. Comparison of the experimental response and the proposed analytical
to the experimental results compare to computational results. As
model.
it was discussed before this is due to the influence of cross section
shape of pitted sections. and the second moment of area are included in a so called pitting
Fig. 18 shows a comparison of the analytical model and the coefficient to modify the ry and kp as a function of mass loss ratio.
experimental results of a corroded bar with L/D = 15 and 30.32% This method is much simpler and is very easy to be implemented
mass loss. Despite that the experimental data in Fig. 18 were not in finite element programme for nonlinear analysis of the RC struc-
included in the model calibration; the analytical model can still tures considering reinforcing bar buckling.
predict response accurately.
Due to the complexity of the problem, random nature of the pit- 5.2. Analytical modelling of cyclic behaviour of corroded bars including
ting corrosion and limitation in the computational models the buckling and low-cycle fatigue
influence of cross section shape may not be captured through the
fibre model analysis. The only way to model this complex phenom- There are currently a number of analytical models available in
enon is a 3D solid element finite element model which is compu- the literature for cyclic stress–strain response of uncorroded rein-
tationally very expensive. However, in the analytical model the forcing steel with and without buckling [30–34,36–38]. The effect
influence of cross section loss is considered assuming an average of corrosion on the behaviour of uncorroded and corroded reinforc-
reduced cross section. Then, the influence of cross section shape ing bars subject to low-cycle fatigue loading has also been studied

Fig. 17. Comparison of the computational and experimental response with the proposed analytical model: (a) L/D = 20 and 34.91% mass loss, (b) L/D = 15 and 19.93% mass
loss, (c) L/D = 10 and 14.44% mass loss.
124 M.M. Kashani et al. / Engineering Structures 75 (2014) 113–125

Fig. 19. Comparison of the corrosion extended Kunnath et al. model and experimentally observed response of corroded bars: (a) a corroded bar with L/D = 5 and 10.36% mass
loss and (b) a corroded bar with L/D = 10 and 21% mass loss.

by other researchers [39–41]. Kashani et al. [16] developed a meth- critical review of the existing analytical material models and a
odology that accounts for the influence of corrosion on combined comparison with the results of computational modelling and phys-
effect of low-cycle fatigue degradation and inelastic buckling of ical testing has been conducted. The key outcomes from this
corroded reinforcing bars subject to cyclic loading. research can be summarised as follows:
Kashani et al. used the experimental data reported by Apostolop-
oulos and Papadopoulos [41] for model calibration. To model the 1. The nonlinear fibre beam–column element can accurately sim-
cyclic behaviour of corroded bars the reinforcing steel model devel- ulate the nonlinear behaviour of reinforcing bars subject to
oped by Kunnath et al. [34] which is already available in the Open- monotonic and cyclic loading. It should be noted that the basic
Sees have been used. This model consists of a low-cycle fatigue steel02 model is not able to capture the low-cycle fatigue degra-
degradation model in tension and Dhakal–Maekawa buckling dation of the reinforcing bars. Apart from this the qualitative fit
model in compression. It also includes the GMP cyclic rules to of the numerical and experimental models are very good. This
model the Bauschinger effect. The combined effect of low-cycle fati- method can be used as a computational tool in the future
gue degradation and the corrosion extended Dhakal–Maekawa research to generate further data for development and calibra-
buckling model is incorporated in and simulated in OpenSees. Fur- tion of new analytical models
ther details and model calibration is available in [16]. A qualitative 2. It was observed in both experimental and computational
comparison between the experimental and analytical cyclic mean responses that the cyclic degradation of compressive strength
stress–strain response of two corroded bars with different L/D and of reinforcing bars with L/D > 8 is a function of the previous
mass loss ratios is shown in Fig. 19. As it is shown in Fig. 19(a) the strain history in tension. This conclusion agrees with the results
corrosion extended Kunnath et al. model can accurately predict obtained by others [42,43].
the response of a corroded bar with L/D = 5 where buckling of bars 3. As the slenderness ratio of bars increased the contribution of
does not have a significant influence on degradation of the hyster- combined cross section area and second moment of area to
etic cycles. However, as the slenderness ratio of bars increases the the overall buckling load reduction increased. Whereas by
accuracy of corrosion extended Kunnath et al. model reduces. reducing the slenderness ratio the contribution of the cross sec-
Fig. 19(b) is an example comparison of this model with the observed tional shape increased.
experimental response of a corroded bar with L/D = 10. 4. It was found that the load eccentricity caused by nonuniform
As discussed before the pinching effect in the cyclic stress– pitting corrosion has a significant contribution on reduction of
strain curve of reinforcing bars is due to the inelastic buckling the buckling load of corroded bars.
and geometrical nonlinearity. In other words as the slenderness 5. It was observed that the cyclic response of corroded bars with
ratio of the reinforcing bars increases the hysteretic response severe localised pitting is mainly governed by the premature
changes towards a nonlinear beam–column element type response yielding of the minimum pitted section.
(e.g. concentric steel bracing). This has also been reported by other 6. The corrosion extended Dhakal–Maekawa buckling model can
researchers [23,42] and their conclusions are in a good agreement accurately predict the post-yield buckling behaviour of corroded
with the results of this study. This phenomenon is more critical in bars under monotonic compression. However, the pinching
corroded bars where non-uniform pitting corrosion can result in a effect and cyclic degradation of the compressive strength of
significant change to the effective slenderness ratio of reinforcing reinforcing bars with L/D > 8 due to the influence of geometrical
bars and subsequently degradation of hysteretic response. nonlinearity is not included in the existing analytical models.
With reference to the Fig. 19, it is evident that there is need for Therefore, there is a need for further analytical study and devel-
further analytical research to develop a new uniaxial material opment of a new phenomenological hysteretic model to account
model for reinforcing bars that accounts for the influence of inelas- for these effects. The computational platform explained in this
tic buckling on degradation of the hysteretic cycles. paper can be used by other researchers for model calibration.

6. Conclusion
Acknowledgements
The influence of corrosion pattern on inelastic buckling and
cyclic behaviour of corroded bars has been investigated computa- The experimental work is funded by Earthquake Engineering
tionally. The key parameters affecting the inelastic buckling and Research Centre (EERC) of the University of Bristol. The computa-
hysteresis behaviour of corroded bars have been identified. A tional part of this research is conducted in collaboration with the
M.M. Kashani et al. / Engineering Structures 75 (2014) 113–125 125

University of Washington while the first author was on sabbatical [21] Dhakal RP, Maekawa K. Modeling for postyield buckling of reinforcement. J
Struct Eng 2002;128(9):1139–47.
leave in the US. The funding provided by the World Wide Univer-
[22] Dhakal RP, Maekawa K. Path-dependent cyclic stress–strain relationship of
sity Network through Research Mobility Programme to the first reinforcing bar including buckling. Eng Struct 2002;24:1139–47.
author is much appreciated. Any findings, opinions and recommen- [23] Nakamura H, Higai T. Modeling nonlinear cyclic behavior of reinforcing bars.
dations provided in this paper are only based on the author’s view. ACI SP 205-14; 2002, vol. 205. p. 273–92.
[24] OpenSees, the open system for earthquake engineering simulation. Pacific
Earthquake Engineering Research Centre, University of California, Berkeley;
References 2011.
[25] Spacone E, Filippou FC, Taucer FF. Fibre beam–column model for non-linear
[1] Wallbank EJ. The performance of concrete in bridges: a survey of 200 highway analysis of R/C frames. Part I: Formulation. Earthq Eng Struct D
bridges, London; 1989. 1996;25:711–25.
[2] The economic impact of current investment trends in surface transportation [26] Spacone E, Filippou FC, Taucer FF. Fibre beam–column model for non-linear
infrastructure. Economic Development Research Group Inc.: ASCE; 2011. analysis of R/C frames. Part II: Applications. Earthq Eng Struct D
[3] Broomfield JP. Corrosion of steel in concrete: understanding, investigation and 1996;25:727–42.
repair. 2nd ed. Taylor and Francis; 2007. [27] Neuenhofer A, Filippou FC. Geometrically nonlinear flexibility-based frame
[4] Rafiq MM, Chryssanthopoulos MK, Onoufriou T. Performance updating of finite element. J Struct Eng 1998;124(6):704–11.
concrete bridges using proactive health monitoring methods. Reliab Eng Syst [28] Souza RM. Force-based finite element for large displacement inelastic analysis
Saf 2004;86:247–56. of frames. PhD thesis, University of California, Berkeley; 2000.
[5] Vu KAT, Stewart MG. Structural reliability of concrete bridges including [29] Menegotto M, Pinto PE. Method of analysis of cyclically loaded RC plane
improved chloride-induced corrosion models. Struct Saf 2000;22:313–33. frames including changes in geometry and nonelastic behavior of elements
[6] Leman DE, Moehle JP. Seismic performance of well-confined concrete bridge under normal force and bending. Preliminary report IABSE, vol. 13, Zurich;
columns. Pacific Earthquake Engineering Research Centre; 2000. 1973. p. 15–22.
[7] Ou Y, Tsai L, Chen H. Cyclic performance of large-scale corroded reinforced [30] Monti G, Nuti C. Nonlinear cyclic behavior of reinforcing bars including
concrete beams. Earthq Eng Struct D 2011;41:592–603. buckling. J Struct Eng 1992;118(12):3268–84.
[8] Ma Y, Che Y, Gong J. Behavior of corrosion damaged circular reinforced [31] Rodriguez ME, Botero JC, Villa J. Cyclic stress–strain behavior of reinforcing
concrete columns under cyclic loading. Constr Build Mater 2012;29:548–56. steel including the effect of buckling. J Struct Eng 1999;125(6):605–12.
[9] Almusallam AA. Effect of degree of corrosion on the properties of reinforcing [32] Gomes A, Appleton J. Nonlinear cyclic stress–strain relationship of reinforcing
steel bars. Constr Build Mater 2001;15:361–8. bars including buckling. Eng Struct 1997;19:822–6.
[10] Du YG, Clark LA, Chan AHC. Residual capacity of corroded reinforcing bars. Mag [33] Bae S, Mieses A, Bayrak O. Inelastic buckling of reinforcing bars. J Struct Eng
Concr Res 2005;57(3):135–47. 2005;131(2):314–21.
[11] Du YG, Clark LA, Chan AHC. Effect of corrosion on ductility of reinforcing bars. [34] Kunnath SK, Heo Y, Mohle JF. Nonlinear uniaxial material model for reinforcing
Mag Concr Res 2005;57(7):407–19. steel bars. J Struct Eng 2009;135(4):335–43.
[12] Cairns J, Plizzari GA, Du YG, Law DW, Chiara F. Mechanical properties of [35] Dhakal RP, Maekawa K. Reinforcement stability and fracture of cover concrete
corrosion-damaged reinforcement. ACI Mater J 2005;102(4):256–64. in RC members. J Struct Eng 2002;128(10):1253–62.
[13] Apostolopoulos CA, Papadopoulos MP, Pantelakis SG. Tensile behavior of [36] Dodd LL, Restrepo-Posada JI. Model for predicting cyclic behavior of
corroded reinforcing steel bars BSt 500s. Constr Build Mater 2006;20:782–9. reinforcing steel. J Struct Eng 1995;121(3):433–45.
[14] Apostolopoulos CA. Mechanical behavior of corroded reinforcing steel bars [37] Balan TA, Filippou FC, Popov EP. Hysteretic model of ordinary and high-
S500s tempcore under low cycle fatigue. Constr Build Mater strength reinforcing steel. J Struct Eng 1998;124(3):288–97.
2007;21:1447–56. [38] Hoehler MS, Stanton JF. Simple phenomenological model for reinforcing steel
[15] Kashani MM, Crewe AJ, Alexander NA. Nonlinear stress–strain behaviour of under arbitrary load. J Struct Eng 2006;132(7):1061–9.
corrosion-damaged reinforcing bars including inelastic buckling. Eng Struct [39] Chang GA, Mander JB. Seismic energy based fatigue damage analysis of bridge
2013;48:417–29. columns: Part I – Evaluation of seismic capacity. Technical report NCEER-94-
[16] Kashani MM, Crewe AJ, Alexander NA. Nonlinear cyclic response of corrosion- 0006; 1994.
damaged reinforcing bar with the effect of buckling. Constr Build Mater [40] Higai T, Nakamura H, Saito S. Fatigue failure criterion for deformed bars
2013;41:388–400. subjected to large deformation reversals, vol. 237. ACI SP 237-4; 2006. p. 37–
[17] Kashani MM, Crewe AJ, Alexander NA. Use of a 3D optical measurement 54.
technique for stochastic corrosion pattern analysis of reinforcing bars [41] Apostolopoulos CA, Papadopoulos MP. Tensile and low cycle fatigue behavior
subjected to accelerated corrosion. Corros Sci 2013;73:208–21. of corroded reinforcing steel bars S400. Constr Build Mater 2007;21:855–64.
[18] ASTM G1-03 standard practice for preparing, cleaning, and evaluating [42] Prota A, Cicco F, Cosenza E. Cyclic behavior of smooth steel reinforcing bars:
corrosion test specimens. ASTM Int’l; 2011. experimental analysis and modeling issues. J Earthq Eng 2009;13:500–19.
[19] MTLAB R2012b. The MathWorks lnc.; 1994–2012. <www.mathworks.com>. [43] Moyer MJ, Kowalsky MJ. Influence of tension strain on buckling of
[20] Mau ST, El-Mabsout M. Inelastic buckling of reinforcing bars. J Eng Mech reinforcement in concrete columns. ACI Struct J 2003;100(1):75–85.
1989;115(1):1–17.

You might also like