Skip to main content
Erschienen in:
Buchtitelbild

Open Access 2018 | Supplement | Buchkapitel

4. Application of the MAST System for Collapse Experiments

verfasst von : Riadh Al-Mahaidi, Javad Hashemi, Robin Kalfat, Graeme Burnett, John Wilson

Erschienen in: Multi-axis Substructure Testing System for Hybrid Simulation

Verlag: Springer Singapore

Aktivieren Sie unsere intelligente Suche, um passende Fachinhalte oder Patente zu finden.

search-config
loading …

Abstract

This chapter presents the results of a range of experiments, including switched/mixed load/deformation mode quasi-static cyclic and hybrid simulation tests to highlight the unique and powerful capabilities of the MAST system, specifically for the assessment and mitigation of the collapse risk of structures.
Hinweise
The original version of this chapter was revised: See the “Chapter Note” section at the end of this chapter for details. The erratum to this chapter is available at https://​doi.​org/​10.​1007/​978-981-10-5867-7_​6

4.1 Introduction

One of the main goals of structural/earthquake engineering is to improve the resilience and performance of structures to withstand collapse due to extreme events. Recent devastating events including a number of large magnitude earthquakes around the world (e.g., Northridge 1994, Kobe 1995, Chile 2010, East Japan and Christchurch 2011) demonstrated that their intensity could reach to more than two times the design level according to the regional code provisions. In this context, it is becoming increasingly important to quantify the reserve capacity of structures against extreme events beyond the design level to the levels approaching collapse. Although there have been many advancements in the mathematical models employed in finite-element methods, many of these analytical models are calibrated using experimental observations. Therefore, experimental research remains critical toward better understanding and predicting the response of structures. However, there are also challenges in conducting laboratory tests for a number of reasons.
Firstly, actions on structures during extreme events such as earthquakes are generally multi-directional and continuously varying, due to the time-dependent nature of the input motion. For instance, variations of the axial loads during a seismic excitation may influence the response of the vertical structural components (e.g., bridge piers and building columns) since the response of such elements when combined with flexural, shear and torsional actions may differ from the cases when they are not subjected to the same axial load changes. Simulation of such highly coupled multi-directional loading conditions using conventional structural testing methods can be expensive, time-consuming and difficult to achieve. As a result, advanced and innovative experimental techniques and control strategies are under development by researchers (Nakata 2007; Wang et al. 2012; Hashemi et al. 2014; Hashemi and Mosqueda 2014).
Secondly, the experiments should be conducted large or full scale to accurately capture the local behavior of the elements. Certain types of local behavior such as bond and shear in reinforced concrete (RC) members, crack propagation, welding effects and local buckling in steel structures are well known to have size effects. Conducting large-scale experiments, however, may not be feasible due to the limited resources available in many laboratories including the number and capability of available actuators, the dimensions and load capacity of the reaction systems, difficulties in the actuator assemblies and testing configuration to reliably simulate the boundary conditions (Hashemi et al. 2016). Consequently, the specimen may be tested small scale or under uni/biaxial loading configurations, which do not necessarily represent the actual action or demand on the structural elements and the corresponding nonlinear response of the prototype system.
Finally, conducting multi-directional loading including axial load effects requires switched-/mixed-mode control strategy. The application of axial loads has been mainly considered by researchers using a combination of force-control actuators in the vertical direction that are decoupled from displacement-control actuators in the lateral direction of the specimen (Lynn et al. 1996; Pan et al. 2005; Del Carpio Ramos et al. 2015). In those tests, independent of lateral actuators, only the vertical force-control actuators apply the axial forces, while under large deformations, lateral actuators have a force component in the vertical direction that needs to be accounted for. Therefore, versatile and generally applicable switched-/mixed-mode control algorithms are required to take into account instantaneous three-dimensional coupling in the control systems.
In this chapter, the unique and powerful capabilities of the MAST system are presented in application for collapse assessment of reinforced concrete (RC) building components. The first experiment is a switched-mode quasi-static test of the RC wall subjected to pure axial demands through displacement-controlled tension and force-controlled compression loads. The next two experiments are conducted on two identical RC columns that are tested through mixed-mode quasi-static and hybrid simulation tests through collapse. In the quasi-static test, the specimen is subjected to constant axial load combined with bidirectional deformation reversal with increasing amplitude that follows a hexagonal orbital pattern. In hybrid simulation, the RC column serves as the first-story corner-column of a half-scale symmetrical 5 × 5 bay 5-story RC ordinary moment frame building that is subjected to bidirectional sequential ground motions with increasing intensities. Finally, to assess the effectiveness of the carbon fiber-reinforced polymer (CFRP) on rehabilitation of RC building columns, the earthquake-damaged RC column was repaired and retested under the same loading condition in hybrid simulation. A simplified collapse risk assessment study was then conducted to compare the response of the RC columns obtained from the quasi-static and hybrid simulation tests as well as the responses of the initial and repaired RC columns.

4.2 Switched-Mode Quasi-Static Test

In the first experiment, the MAST system was used to assess the performance of an RC wall in a building that collapsed during the 2011 Christchurch earthquake. The building was a 5-story RC structure with a lateral load-resisting system comprised of RC walls forming a large shear core, centrally located in the floor plan. The collapse mechanism of the building was hypothesized to be an out-of-plane buckling instability of one of the RC core walls.
Due to the configuration of doorways and wall openings in the shear core, one of the walls at the ground floor had no adjacent return segments. It was theorized that this resulted in the wall being subjected to cyclic tension-compression axial loads due to the overturning moments on the shear core during the earthquake. Large cyclic axial loads eventually caused high enough tensile strains on the wall that it became extremely susceptible to out-of-plane buckling instabilities on the reversed cycle, where the wall was subjected to axial compression.
To simulate the out-of-plane buckling of the critical wall, a 3:4 scale model was constructed with a cross section of 1050 mm × 148 mm and a height of 2680 mm. The wall had a single central layer of longitudinal and transverse reinforcement, with, respectively, 0.25 and 0.35% reinforcement ratios. This resulted in two N16 (normal ductility with 16 mm diameter) longitudinal bars and 14-mm-diameter transverse bars placed at 300 mm centers. The reinforcement had a yield stress of 570 MPa and an ultimate stress of 670 MPa. The wall was cast using a Grade 40 concrete mix (i.e., a 40 MPa characteristic compressive strength at 28 days). Despite the concrete being a standard Grade 40 mix, the mean compressive strength on the test day, determined from 10 concrete cylinders with 100 mm diameter, was approximately 65 MPa.
To determine the approximate loading profile of the axial load on the wall, a finite-element model of the full building was built and subjected to the recorded ground motions. The loading profile was applied to the test specimen using switched-mode control. In this method, the applied loading automatically switches between force-controlled and displacement-controlled loading protocols. When the test specimen is being ‘pulled up’ in tension, the concrete cracks and the reinforcement yields and undergoes plastic deformation. Therefore, displacement-controlled loading is suitable to load the specimen up to the required tension displacement. When the test specimen is being ‘pushed down’ in compression, the concrete undergoes very small displacement increments for the associated amount of force, which sometimes are beyond the resolution of the measurement system. Therefore, force-controlled loading should be implemented to accurately reach to the desired level of compressive load. The loading protocol procedure is graphically illustrated in Fig. 4.1. Note that, when switching back and forth between force and displacement, the specimen is brought back to zero force to avoid any control instabilities.
Figure 4.2a shows the wall tested under the MAST system along the Z axis, while the remaining 5 DOFs (i.e., X, Y, Rx, Ry and Rz) are commanded to zero deformation to allow for pure application of axial load on the specimen. Figure 4.2b shows the out-of-plane failure of the RC wall occurred at the base. The command and measured signals are also presented in Fig. ‎4.3. The switching points can be seen in force time history, where the applied load changes from tension to compression and vice versa. Also, it can be seen that the measured displacement and force signals precisely matched the respective command signals in tension and compression. In terms of the specimen response, it is observed that the fracture of a longitudinal rebar (at ~1630 s) caused a sudden force reduction in tension, which was followed by the total failure of the specimen in compression.

4.3 Mixed-Mode Quasi-Static Test

The second experiment is a three-dimensional mixed-mode quasi-static cyclic test conducted on a large-scale limited-ductility RC column. Figure ‎4.4a and b, respectively, shows the design details of the column and the 6-DOF axes of crosshead movements. The specimen is attached to the strong floor from the base and to the crosshead from the top through rigid concrete pedestals. The RC column is 2.5 m high, has a square 250 mm × 250 mm cross section and is reinforced with four longitudinal bars of N16 and tied with R6 stirrups spaced at 175 mm with 30 mm cover thickness. The material properties of the specimen, obtained from laboratory tests, are also presented in Table 4.1.
Table 4.1
Material properties of the RC column
Concrete
Steel
\(\varepsilon_{c}\)
0.002
\(\varepsilon_{y}\)
0.0035
\(f_{c}^{{\prime }}\)
35.1 (MPa)
\(f_{y}\)
633.95 (MPa)
\(\varepsilon_{cu}\)
0.0063
\(\varepsilon_{u}\)
0.073
\(f_{u}^{{\prime }}\)
0.0
\(f_{u}\)
712.0 (MPa)
The loading protocol consists of simultaneously applying a constant gravity load, equal to 8% of ultimate compressive load capacity in force control, while imposing bidirectional lateral deformation reversals in displacement control that follows the hexagonal orbital pattern suggested in FEMA 461 (Federal Emergency Management Agency 2007), as shown in Fig. ‎4.5‎. The sequence of loading in QS testing started with applying the gravity load on the specimen along Z axis. The specimen was then pushed to the initial uniaxial drift ratio toward point ‘a,’ followed by the orbital pattern ‘a-b-c-d-e-f-a.’ The reversal from point ‘a’ accompanies an orthogonal drift at points ‘b’ and ‘c’ equal to one-half the maximum drift ratios at points ‘a’ and ‘d.’ The entire loading cycle was then repeated at the same amplitude. Once the specimen reached point ‘a’ for the second time, the amplitude value for the next two cycles was increased and the next two biaxial load cycles were applied on the specimen. The process continued until the failure of the specimen. The remaining DOF axes (roll, pitch and yaw) were controlled in zero angle forming a double-curvature deformation of the column.
The results of the QS test, including the hysteretic behavior of the RC column in X and Y axes, axial load time history in Z axis, biaxial lateral drifts in X and Y axes and biaxial bending moments in Rx and Ry axes are presented in Fig. ‎4.6‎. The force relaxations observed in the hysteresis were due to pausing of the test in order to collect photogrammetry data at peak deformations in the X axis. The failure of the specimen occurred when the specimen was subjected to the maximum of 7.0 and 3.5% drift ratios in Y and X axes, respectively. These are large drifts for a limited ductile column, but effective of the relatively low axial loads applied to the column (Wibowo et al. 2014).

4.4 Mixed-Mode Hybrid Simulation Test

The third experiment is a three-dimensional mixed-mode hybrid simulation using an identical RC column to the one previously tested in the quasi-static cyclic experiment. For this purpose, a half-scale symmetrical 5-story (height of first story h1 = 2.5 m, height of other stories htyp = 2.0 m) 5 × 5 bay (column spacing b = 4.2 m) RC ordinary moment frame building is selected as the hybrid model. The physical specimen serves as the first-story corner-column of the building, considered as the critical element of the structure. The rest of the structural elements, inertial and damping forces, gravity and dynamic loads and second-order effects are modeled numerically in the computer. Figure ‎4.7‎ illustrates the components of hybrid simulation including numerical and experimental substructures.
The structure’s beams and columns were modeled using beam-with-hinges elements, where the nonlinear behavior is assumed to occur within a finite length at both ends based on the distributed plasticity concept (Scott and Fenves 2006) (Fig. ‎‎4.8a). The plasticity model follows a peak-ordinated hysteresis response based on the modified Ibarra–Medina–Krawinkler (IMK) deterioration model of flexural behavior (Ibarra et al. 2005; Zhong 2005). This model was chosen because it is capable of capturing the important modes of deterioration that participate in sidesway collapse of RC frames. The model requires the specification of a range of parameters to control the tri-linear monotonic backbone curve and different modes of cyclic deteriorations. As shown in Fig. ‎‎4.8b, these parameters include \(M_{\text{y}}\), \(M_{\text{c}} /M_{\text{y}}\), \(\uppsi_{\text{p}}\), \(\uppsi_{\text{pc}}\) and \(\uplambda\) that, respectively, represent yielding moment, a measure of maximum moment capacity, plastic curvature capacity, post-capping curvature capacity and cyclic deterioration. The model captures four modes of cyclic deterioration, including strength deterioration of the hardening region (\(\uplambda_{\text{S}}\)), strength deterioration of the post-peak softening region (\(\uplambda_{\text{C}}\)), accelerated reloading stiffness deterioration (\(\uplambda_{\text{A}}\)) and unloading stiffness deterioration (\(\uplambda_{\text{K}}\)). Based on the studies by Haselton et al. (2008), the strength deterioration of hardening region and the post-capping strength deterioration were assumed to be equal in the case study, while accelerated reloading and unloading stiffness deterioration were ignored. This reduced the calibration of cyclic deteriorations to one parameter. Table ‎4.2 presents the IMK parameters for beam and column elements.
Table 4.2
IMK model parameters for RC frame structure
Story no.
Element location
\(M_{\text{c}} /M_{\text{y}}\)
\(M_{\text{y}}\) \(\left( {{\text{kN}}.{\text{m}}} \right)\)
\(EI/EI_{\text{g}}\)
\(\uppsi_{\text{p}}\)
\(\uppsi_{\text{pc}}\)
\(\uplambda_{\text{S}} ,\uplambda_{\text{C}}\)
Story 1
Columns
Corner
1.20
114
0.52
0.12
9.00
14.0
Exterior
1.20
128
0.60
0.10
7.00
10.5
Interior
1.18
467
0.48
0.08
8.70
28.0
Story 2
Columns
Corner
1.18
111
0.41
0.13
9.20
15.4
Exterior
1.20
124
0.48
0.11
7.61
11.9
Interior
1.18
464
0.38
0.08
9.13
30.1
Story 3
Columns
Corner
1.18
109
0.39
0.13
9.76
16.1
Exterior
1.20
117
0.44
0.12
8.12
13.3
Interior
1.18
450
0.35
0.09
10.0
32.9
Story 4
Columns
Corner
1.18
106
0.37
0.13
9.96
16.8
Exterior
1.18
112
0.41
0.13
9.18
15.4
Interior
1.18
435
0.35
0.09
10.2
35.0
Story 5
Columns
Corner
1.18
102
0.36
0.14
10.4
18.2
Exterior
1.18
106
0.37
0.14
9.71
16.8
Interior
1.18
419
0.35
0.09
10.6
37.8
Beams
Exterior end span
1.195
160
0.35
0.11
8.99
29.6
Exterior span
152
0.35
0.11
8.83
31.1
Interior end span
247
0.35
0.16
8.99
46.5
Interior span
246
0.35
0.13
9.02
38.2
After developing the numerical model, the elastic fundamental period of vibration and the corresponding first mode shape were obtained through eigenvalue analysis. A nonlinear static pushover analysis was then performed with the lateral force distribution proportional to the fundamental mode of vibration and with the consideration of second-order \({\text{P}} -\Delta\) effects. Figure ‎‎4.9 presents the results of the pushover analysis that show most of the energy dissipation occurs in the lower two stories.
For the HS test, the two horizontal components of the 1979 Imperial Valley earthquake ground motions recorded at El Centro station with peak ground acceleration of 0.15 g were used. Figure ‎‎4.10‎ shows the acceleration, displacement and acceleration–displacement response spectra for the two ground-motion components. Based on incremental dynamic analysis, four levels of intensities were considered to capture the full range of structural response from linear elastic range to collapse. The selected scale factors were 0.6, 4.0, 8.0 and 9.0, which pushed the structure to nearly 0.25% (elastic), 2, 4 and 6% inter-story drift ratios, respectively.
Prior to conducting the actual HS test with the physical subassembly in the laboratory, a series of FE-coupled numerical simulations (Schellenberg et al. 2009) was conducted to evaluate the integration scheme parameters for the actual experiments. Accordingly, Generalized Alpha-OS (Schellenberg et al. 2009) was used as the integration scheme, and the integration time step was optimized to preserve the accuracy and stability of the simulation, while allowing the completion of the entire test during the regular operation time of the laboratory. Five percentage Rayleigh damping was specified to the first and third modes of vibration, corresponding to the primary translational modes in X and Y directions. Additional damping was also assigned to free vibration time intervals between the forced vibrations in order to quickly bring the structure to rest.
The hybrid simulation started with applying the gravity load on the specimen, using a ramp function, followed by sequential ground motions. The entire sequence of loading was performed and automated using OpenSees. Considering the 117-ms delay in the hydraulic system, 500 ms was specified as the simulation time step in xPC-Target predictor-corrector to provide sufficient time for integration computation, communication, actuator motions and data acquisition. This scaled the 60 s of sequential ground motions to 6 h in laboratory time. Similar to the QS test, the rotational axes (roll, pitch and yaw) were controlled in zero angle, forming a double-curvature deformation of the column.
Figure ‎4.11‎ compares the responses of RC column including the hysteretic behavior of the RC column in X and Y axes, axial load time history in Z axis, biaxial lateral drifts in X and Y axes and biaxial bending moments in Rx and Ry axes for the QS and the HS tests. The specimen was pushed to a maximum of 6.4 and 2.7% drift ratios in Y and X axes, respectively. The maximum time-varying axial load applied on the specimen was 553 kN in compression and 161 kN in tension. By comparing the hysteresis plots from the HS test, it can be seen that the column was damaged as the structure progressively moved in one direction, while in the QS test the pattern of damage was symmetrical due to load reversals in cyclic deformations. Figure ‎‎4.12 shows the flexural failure of columns for both tests by comparing the plastic hinges developed at the top and the base of the columns from different angles.

4.5 Comparison of Mixed-Mode Quasi-Static and Hybrid Simulation Tests

4.5.1 Numerical Model and Calibration

In order to investigate the influence of the selected experimental method on assessing the collapse risk, a comparative collapse fragility analysis for a substructure of the RC building was conducted using the results of QS and HS tests, respectively. As illustrated in Figure ‎‎4.13, the numerical model selected for incremental dynamic analysis includes only the first-story corner-column and the overhead mass portion of the upper five floors, which is equivalent to a single-degree-of-freedom (SDOF) system with a natural period of 0.6 s. This allows the study of the response of the critical element (i.e., the first-story corner-column) purely based on experimental results without the influence of other elements’ responses.
The experimental results were used to calibrate the SDOF numerical model. As previously shown in Fig. ‎‎‎4.8, the moment-curvature behavior of the plastic zones follows the IMK hysteresis model. Although this model can generally simulate most of the important behaviors, including strength and stiffness degradation, effects such as the interaction between axial, flexure and shear failure cannot be captured. Accordingly, a unidirectional numerical model of the column was selected, and the hysteresis parameters of IMK model were calibrated to the response of the specimen in the main axis (i.e., the Y axis of the MAST system), along which it experienced maximum deformation. Consequently, the influence of axial loads and out-of-plane movements in the experiments were implicitly taken into account by using the calibrated numerical models. Note that the use of fiber-based plasticity models could be an alternative. However, only the most basic aspects such as material constitutive relationships are modeled, while the degradation parameters that have a significant impact on collapse behaviors are not included (Deierlein et al. 2010).
A close view of the hysteretic responses of the RC columns in the QS and HS tests is presented in Fig. ‎‎‎4.14. It can be clearly observed that the flexural strength, the capping point and post-capping negative tangent stiffness (in-cycle strength degradation) are significantly different. Accordingly, following the procedures given in Haselton et al. (2008), the numerical SDOF model was calibrated to the QS and HS test results, with particular focus on precisely mimicking the plastic and post-capping deformation capacities as well as the cyclic deteriorations that are known to have important influence on collapse prediction (see Fig. ‎‎4.15). However, it is noted that some portions of unloading and reloading phases in the experiments, especially for the QS test, resulted in higher strength and larger hysteretic loops in the experiment compared to the more pinched behavior in calibrated models.
Table ‎4.3 compares the IMK model parameters for QS- and HS-based numerical models. The higher flexural strength and significant reduction in the drift capacity observed in the HS test may be due to the higher levels of axial load in this test, as previously addressed (Lynn et al. 1996; Nakamura and Yoshimura 2002; Sezen 2002). Figure ‎4.16 specifically shows that the rapid drop in shear strength occurred immediately after the maximum compressive axial load in the HS test, which is 2.9 times larger than the uniform axial load applied in the QS test. This clearly shows the significant impacts of the axial load level and its variation on the performance of RC structures and the ultimate drift capacity. Studies conducted by Wibowo et al. (2014) reported similar findings based on an experimental program performed to develop a generic backbone pushover curve for lightly reinforced concrete columns.
Table 4.3
Comparison of IMK model parameters calibrated to QS and HS tests
Experiment method
\(M_{\text{c}} /M_{\text{y}}\)
\(M_{\text{y}}^{ + } \left( {{\text{kN}}.{\text{m}}} \right)\)
\(M_{\text{y}}^{ - } \left( {{\text{kN}}.{\text{m}}} \right)\)
\(EI/EI_{\text{g}}\)
\(\uppsi_{\text{p}}^{ + }\)
\(\uppsi_{\text{p}}^{ - }\)
\(\uppsi_{\text{pc}}^{ + }\)
\(\uppsi_{\text{pc}}^{ - }\)
\(\uplambda_{\text{S}} ,\uplambda_{\text{C}}\)
QS
1.12
68.0
58.0
0.258
0.14
0.18
6.0
5.0
6.0
HS
1.12
78.0
54.0
0.258
0.07
0.12
0.9
1.0
9.2
Another observation is that the specimen showed larger cyclic deteriorations in the QS test due to the application of many large cycles and load reversals to the specimen before failure. Note that the values assigned to the cyclic deterioration in the IMK model are inversely related to the level of deterioration.

4.5.2 Fragility Analysis

Incremental dynamic analyses (IDAs) were performed using the calibrated numerical model in order to capture a range of probable dynamic response behaviors due to record-to-record variability in ground-motion characteristics. For this purpose, three earthquake scenarios including M6.0R28, M6.5R40 and M7.0R90 (M and R stand for magnitude and source-site distance, respectively) were considered. A suite of 20 recorded ground motions was selected from the PEER database (PEER 2013) that are listed in Table ‎4.4 along with the values of peak ground acceleration (PGA) and spectral acceleration at the fundamental natural period of the numerical model S a (T 1) and T 1 = 0.6 s for the SDOF model. The response spectra of the input ground motions are also shown in Fig. ‎4.17.
Table 4.4
List of input ground motions used in incremental dynamic analysis
No.
Record Name
Scenario
\(PGA\left( g \right)\)
\(S_{a} \left( {T_{1} } \right)\left( g \right)\)
1
WHITTIER 10/01/87 14:42 (USC STATION 90017)
M6.0 R28
0.0365
0.0798
2
CHICHI AFTERSHOCK 09/20/99 1757, TCU076, E (CWB)
M6.0 R28
0.0716
0.0719
3
CHICHI AFTERSHOCK 09/20/99 1757, TCU088, E (CWB)
M6.0 R28
0.0584
0.0118
4
CHICHI AFTERSHOCK 09/20/99 1757, TCU129, E (CWB)
M6.0 R28
0.0927
0.0366
5
CHICHI AFTERSHOCK 09/20/99 1803, CHY074, E (CWB)
M6.0 R28
0.0617
0.2227
6
SAN FERNANDO 02/09/71 14:00 (CDWR STATION 269)
M6.0 R28
0.1018
0.0580
7
NORTHRIDGE 1/17/94, 12:31 (CDMG STATION 24305)
M6.5 R40
0.0889
0.1654
8
NORTHRIDGE 1/17/94, 12:31 (CDMG STATION 24307)
M6.5 R40
0.0843
0.1935
9
NORTHRIDGE 1/17/94, 12:31 (CDMG STATION 24521)
M6.5 R40
0.0614
0.1485
10
NORTHRIDGE 1/17/94, 12:31 (CDMG STATION 24644)
M6.5 R40
0.0909
0.2497
11
CHICHI AFTERSHOCK 09/25/99 2352, CHY029, E (CWB)
M6.5 R40
0.2393
0.1636
12
CHICHI AFTERSHOCK 09/25/99 2352, CHY035, E (CWB)
M6.5 R40
0.1709
0.3447
13
CHICHI AFTERSHOCK 09/25/99 2352, HWA020, E (CWB)
M6.5 R40
0.0260
0.0425
14
CHICHI AFTERSHOCK 09/25/99 2352, HWA035, E (CWB)
M6.5 R40
0.0218
0.0520
15
CHICHI AFTERSHOCK 09/25/99 2352, HWA058, E (CWB)
M6.5 R40
0.0465
0.0570
16
CHICHI AFTERSHOCK 09/25/99 2352, TCU048, E (CWB)
M6.5 R40
0.0365
0.0798
17
CHICHI AFTERSHOCK 09/25/99 2352, TCU087, E (CWB)
M6.5 R40
0.0217
0.0399
18
CHICHI AFTERSHOCK 09/25/99 2352, TCU104, E (CWB)
M6.5 R40
0.0282
0.0733
19
CHICHI AFTERSHOCK 09/25/99 2352, TCU136, N (CWB)
M6.5 R40
0.0391
0.0258
20
LOMA PRIETA 10/18/89 00:05 (CDMG STATION 58043)
M7.0 R90
0.0713
0.1294
Each unidirectional ground motion was individually applied to QS- and HS-based calibrated numerical models for the nonlinear simulation. The ground motions were increasingly scaled according to the value of S a (0.6), until reaching the state of collapse. The simulation was based on 5% mass-proportional damping and restricted to sidesway-only collapse with a drift limit of 7% based on the experimental results. The outcome of this assessment is a structural collapse fragility function, which is a lognormal distribution relating the structure’s probability of collapse to the ground-motion intensity, in terms of S a (T 1 = 0.6 s, 5% damping). Figure ‎4.18 presents the results of nonlinear incremental time-history analyses for QS- and HS-based numerical models.
The mean (the \(S_{a}\) level with 50% probability of collapse) and standard deviation (the dispersion of \(S_{a}\)) for each case can be derived from the following equations (Ibarra and Krawinkler 2005):
$${ \ln }\left( \theta \right) = \frac{1}{n}\mathop \sum \limits_{i = 1}^{n} \ln \left( {S_{a} \left( i \right)} \right)$$
(4.1)
$$\beta = \sqrt {\frac{1}{n - 1}\mathop \sum \limits_{i = 1}^{n} \left( {\ln \left( {\frac{{S_{a} \left( i \right)}}{\theta }} \right)} \right)^{2} }$$
(4.2)
where \(n\) is the number of ground motions considered and \(S_{a} \left( i \right)\) is the \(S_{a}\) value associated with the onset of collapse for the \(i\)-th ground motion. In addition, \({ \ln }\left( \theta \right)\) and \(\beta\) are, respectively, the mean and the standard deviation of the normal distribution that represents the \(\ln \left( {S_{a} } \right)\) values. Note that the mean of \(\ln \left( {S_{a} } \right)\) corresponds to the median of \(S_{a}\) in the case that \(S_{a}\) is lognormally distributed.
The computed mean and standard deviation values for QS- and HS-based numerical models show that while the dispersion of \(S_{a}\) is similar in both cases, the \(S_{a}\) level with 50% probability of collapse is significantly over-estimated in the QS-based model (\(\overline{{S_{a} }} = 1.5{\text{g}}\)) compared to the HS-based model \((\overline{{S_{a} }} = 1.2{\text{g}})\). A lognormal cumulative distribution function was then used to define the fragility functions (Porter et al. 2007):
$${\text{P}}({\text{Collapse}}|S_{a} ) =\Phi \left( {\frac{{\ln \left( {\frac{{S_{a} }}{\theta }} \right)}}{\beta }} \right)$$
(4.3)
where \({\text{P}}({\text{Collapse}}|S_{a} )\) is the probability that a ground motion with intensity of \(S_{a}\) will cause the structure to collapse, \(\Phi\) is the standard normal cumulative distribution function (CDF), \(\theta\) is the median of the fragility function and \(\beta\) is the standard deviation of \({ \ln }\left( \theta \right)\).
Figure ‎4.19 indicates that the differences in the collapse probability between fragility curves become larger as the intensity level increases. Specifically, at the intensity level \(S_{a} = 1.5{\text{g}}\), where the QS-based model predicts 50% probability of collapse, the HS-based model predicts 75.2% probability of collapse, which is 1.5 times larger. This shows the significance of the choice of experimental technique and the influence of axial load on collapse risk assessment of the structure.

4.6 Post-Earthquake Rehabilitation of Damaged RC Column Through CFRP Repair

The use of fiber-reinforced polymer (FRP) for the repair and rehabilitation of earthquake-damaged structures can be considered as a cost-effective and time-saving alternative to replacement. This is due to many advantages that FRP possesses over traditional materials for strengthening, including its high tensile strength, light weight, resistance to corrosion, durability and ease of installation. While numerous experimental studies have demonstrated the effectiveness of FRP strengthening for improving the seismic behavior of RC beam–column joints (Shan et al. 2006; Barros et al. 2008; Yalcin et al. 2008; Realfonzo and Napoli 2009; Wei et al. 2009; Gu et al. 2010; Ozcan et al. 2010; ElSouri and Harajli 2011; Realfonzo and Napoli 2012; Wang et al. 2012; Wang and Ellingwood 2015), few studies are available on the repair and rehabilitation of previously damaged column connections and the suitability of FRP as a post-earthquake strengthening solution (Ma and Li 2015; Jiang et al. 2016).
The fourth experiment aimed to evaluate the capability of CFRP repair on rehabilitating the earthquake-damaged RC column to its initial collapse resistance capacity. For this purpose, the damaged column was repaired using CFRP wraps in the plastic zones observed at both ends of the specimen. The damage contained localized zones of spalled and fractured concrete, horizontal and inclined cracking and bent longitudinal reinforcements.
The repair methodology involved: (1) removal of all spalled and fractured concrete; (2) epoxy injection of any cracks wider than 0.3 mm; (3) reinstatement of damaged concrete with a suitable repair mortar; and (4) wrapping of the column with CFRP. However, replacement of the yielded/buckled/ruptured rebars was not included in the repair process. Visual inspection and light tapping using a rubber hammer were used to identify and remove fractured concrete. Cracks that required injection were identified and labeled. Epoxy injection ports were drilled into the concrete directly over the crack and bonded to the surface with epoxy resin. The surface of the crack was sealed and the injection carried out using Sikadur® 52 high-strength adhesive (see Fig. ‎4.20).
The injection ports were cut after hardening of the Sikadur® 52. A repair mortar, BASF MasterEmaco® S 5300, which is a polymer-modified structural repair mortar, was then used to replace the damaged concrete (see Fig. ‎4.21). The average compressive strength of the repair mortar was 41.9 MPa, based on the results of three 50 mm × 50 mm cubes obtained on the test date. The mortar was tested in accordance with ASTM C109 (American Society for Testing and Materials 2011).
The CFRP wrapping was applied over a 600-mm length at each end of the column in regions corresponding to the maximum moment, three days after the crack injection was performed. The concrete in these regions was confined using three layers of MBrace CF130 unidirectional carbon fiber sheet. The CFRP was expected to provide a passive confinement pressure, thereby increasing the compressive strength of concrete. Furthermore, the orientation of the fibers was arranged parallel to the existing steel stirrups, and this was expected to significantly increase the shear capacity at the column ends. The total increases in axial and shear capacity of the column as a result of the CFRP wrapping were estimated as 35 and 250%, respectively, when calculated in accordance with ACI440.2R-08 (American Concrete Institute 2008). A summary of the material properties of the CFRP (MBrace CF130) and adhesive (MasterBrace® P 3500 Primer) used in the repair is given in Tables ‎4.5 and 4.6.
Table 4.5
Summary of CFRP material properties
Properties
MBrace
Tensile strength
4900 (MPa)
Tensile modulus
230 (GPa)
Ultimate elongation
2.1%
Thickness
0.227 mm
Table 4.6
Summary of epoxy saturant and primer material properties
Properties
Saturant
Primer
Resin type
Epoxy
Epoxy
Specific gravity
1.12
1.08
Modulus of elasticity
> 3.0 (GPa)
0.7 (GPa)
Tensile strength
> 40 (MPa)
> 12 (MPa)
Compressive strength
> 80 (MPa)
Prior to the application of the CFRP to the concrete surface, the corners of the column were rounded to achieve a minimum radius of 25 mm. A mechanical abrasion technique was used to remove the weak layer of cement laitance adhering to the surface of the concrete and achieve a surface roughness similar to 60 grit sandpaper (see Fig. ‎4.22). The surface was cleaned to remove any dust prior to application of the FRP. The FRP was applied using a wet layup technique, and each layer was thoroughly impregnated with resin prior to application to the column. The repair process was performed while the column was still under the MAST system and supporting an axial load of 130 kN, in order to mimic the actual gravity load in a real structural repair scenario. The CFRP was cured at 50 ℃ for 7 days using heat lamps prior to testing (see Fig. ‎4.23).

4.7 Repeating Mixed-Mode Hybrid Simulation Test

The repaired column was tested under the same hybrid testing conditions as for the initial column. The intensity levels in hybrid simulation included the same previous four scale factors of 0.6, 4.0, 8.0, 9.0, as well as an additional scale factor of 10.0, in order to push the structure to ~0.25% (elastic), 2.0, 4, 6 and 8% maximum inter-story drift ratios, respectively.
Hybrid simulation was completed with no rupture observed in the CFRP sheets. A detailed comparison of hybrid simulation test results for the initial and repaired columns is presented in Fig. ‎4.24. The results include the hysteretic response in X and Y axes, the axial force time history in Z axis, energy dissipation computed from lateral force deformation in X and Y axes and the biaxial bending moments in Rx and Ry axes.
Figure ‎4.25a shows a closer view of the hysteretic response of the initial and repaired columns in Y axis, along which the column experienced maximum deformation. Two main significant changes can be observed in the behavior of the repaired column. First, the CFRP repair was not able to restore the flexural strength of the initial column, as the maximum resisting force was 32% less in the repaired column. This is mainly due to the fact that the repair process did not include replacement of the yielded, buckled or ruptured rebars, and as a result the loss of strength could not be fully compensated. Second, the repaired column showed significant improvement in ductility due to the confinement effects of the CFRP wraps. As observed in Fig. ‎‎4.25a, the hardening branch of the plastic deformation response of the repaired column is extended to much larger drifts compared to the initial column. Specifically, while applying the maximum compressive axial load on the initial column (553 kN = 23.35% ultimate capacity), a rapid drop occurred immediately after reaching the peak resisting force. However, the repaired column remained in the hardening region while being subjected to the same level of axial load. This is also evident by comparisons of other corresponding cycles from the two experiments. For instance, Fig. ‎4.25b shows, respectively, the capping points ‘A’ and ‘B’ for the initial and repaired column from the same corresponding cycles. The initial column shows stiffness hardening up to 3% drift (point ‘A’), while this value has been extended to 4.5% drift (point ‘B’) for the repaired column. In addition, by comparing the behavior of the RC columns after point ‘C,’ which is located on the same corresponding cycle and at the same level of drift, it is observed that the initial column entered the post-capping negative stiffness region, while the repaired column was still in the stiffness hardening region.

4.8 Comparison of Initial and CFRP-Repaired RC Columns

4.8.1 Numerical Model and Calibration

The same numerical model, as shown in Fig. ‎4.13, was selected for the comparative collapse risk assessment of the initial and repaired columns. Figure ‎4.26 compares the calibrated SDOF model of the column to the hybrid simulation results in Y axis. Particular emphasis was placed on precisely mimicking the plastic and post-capping deformation capacities, which are known to have an important influence on collapse prediction. Table ‎4.7 compares the IMK model parameters for the calibrated models. The main differences include the use of a lower flexural strength but a much higher plastic deformation capacity in the stiffness hardening region for the model of the repaired column.
Table 4.7
Comparison of IMK model parameters calibrated to experimental results
Experiment method
\(M_{\text{c}} /M_{\text{y}}\)
\(M_{\text{y}}^{ + } \left( {{\text{kN}}/{\text{m}}} \right)\)
\(M_{\text{y}}^{ - } \left( {{\text{kN}}/{\text{m}}} \right)\)
\(EI/EI_{\text{g}}\)
\(\uppsi_{\text{p}}^{ + }\)
\(\uppsi_{\text{p}}^{ - }\)
\(\uppsi_{\text{pc}}^{ + }\)
\(\uppsi_{\text{pc}}^{ - }\)
\(\uplambda_{\text{S}} ,\uplambda_{\text{C}}\)
Initial column
1.12
78.0
54.0
0.258
0.07
0.12
0.9
1.0
9.2
Repaired column
1.10
56.0
54.0
0.258
0.34
0.12
0.9
1.0
9.2

4.8.2 Fragility Analysis

Incremental dynamic analyses (IDAs) were performed using the calibrated numerical models in order to capture a range of probable dynamic response behaviors due to record-to-record variability in ground-motion characteristics. The ground motions used for this purpose are the same as those previously listed in Table ‎4.4. Each unidirectional ground motion was individually applied to the calibrated SDOF models and increasingly scaled, until the state of complete collapse. The simulation was based on 5% mass-proportional damping and restricted to sidesway-only collapse with a drift limit of 7%, based on the experimental results. The outcome of this assessment is a structural collapse fragility function for the initial and repaired columns, respectively, which is a lognormal distribution relating the structure’s probability of collapse to the ground-motion intensity level (\(S_{a}\)). Figures ‎4.27 and 4.26 present the results of nonlinear incremental time-history analyses and the associated fragility curves, respectively. It is observed that at an intensity level (\(S_{a}\)) of 1.2 g, the probability of collapse for the initial columns is 50%, while this value for the repaired column is 44%. This shows that CFRP repair can effectively restore the capacity of the column and slightly improve the resistance of the column to sidesway collapse (Fig. 4.28).
Literatur
Zurück zum Zitat American Concrete Institute. (2008). Guide for the design and construction of externally bonded FRP systems for strengthening concrete structures. SN:440.2R-8, Farmington Hills, U.S. American Concrete Institute. (2008). Guide for the design and construction of externally bonded FRP systems for strengthening concrete structures. SN:440.2R-8, Farmington Hills, U.S.
Zurück zum Zitat American Society for Testing and Materials. (2011). Standard test method for compressive strength of hydraulic cement mortars (using 2-in. or [50-mm] cube specimens). SN:C109–11/C109 M-11, West Conshohocken, U.S. American Society for Testing and Materials. (2011). Standard test method for compressive strength of hydraulic cement mortars (using 2-in. or [50-mm] cube specimens). SN:C109–11/C109 M-11, West Conshohocken, U.S.
Zurück zum Zitat Barros, J. A. O., Varma, R. K., Sena-Cruz, J. M., & Azevedo, A. F. M. (2008). Near surface mounted CFRP strips for the flexural strengthening of RC columns: Experimental and numerical research. Engineering Structures, 30(12), 3412–3425.CrossRef Barros, J. A. O., Varma, R. K., Sena-Cruz, J. M., & Azevedo, A. F. M. (2008). Near surface mounted CFRP strips for the flexural strengthening of RC columns: Experimental and numerical research. Engineering Structures, 30(12), 3412–3425.CrossRef
Zurück zum Zitat Deierlein, G. G., Reinhorn, A. M., & Willford, M. R. (2010). Nonlinear structural analysis for seismic design, a guide for practicing engineers. Gaithersburg, MD, U.S.: National Institute of Standards and Technology. Deierlein, G. G., Reinhorn, A. M., & Willford, M. R. (2010). Nonlinear structural analysis for seismic design, a guide for practicing engineers. Gaithersburg, MD, U.S.: National Institute of Standards and Technology.
Zurück zum Zitat Del Carpio Ramos, M., Mosqueda, G., & Hashemi, M. J. (2015). Large-scale hybrid simulation of a steel moment frame building structure through collapse. Journal of Structural Engineering, 142(1), 04015086.CrossRef Del Carpio Ramos, M., Mosqueda, G., & Hashemi, M. J. (2015). Large-scale hybrid simulation of a steel moment frame building structure through collapse. Journal of Structural Engineering, 142(1), 04015086.CrossRef
Zurück zum Zitat ElSouri, A. M., & Harajli, M. H. (2011). Seismic repair and strengthening of lap splices in RC columns: Carbon fiber-reinforced polymer versus steel confinement. Journal of Composites for Construction, 15(5), 721–731.CrossRef ElSouri, A. M., & Harajli, M. H. (2011). Seismic repair and strengthening of lap splices in RC columns: Carbon fiber-reinforced polymer versus steel confinement. Journal of Composites for Construction, 15(5), 721–731.CrossRef
Zurück zum Zitat Federal Emergency Management Agency. (2007). Interim testing protocols for determining the seismic performance characteristics of structural and nonstructural components. Department of Homeland Security (DHS), Washington D.C., U.S. Federal Emergency Management Agency. (2007). Interim testing protocols for determining the seismic performance characteristics of structural and nonstructural components. Department of Homeland Security (DHS), Washington D.C., U.S.
Zurück zum Zitat Gu, D. S., Wu, G., Wu, Z. S., & Wu, Y. F. (2010). Confinement effectiveness of FRP in retrofitting circular concrete columns under simulated seismic load. Journal of Composites for Construction, 14(5), 531–540.CrossRef Gu, D. S., Wu, G., Wu, Z. S., & Wu, Y. F. (2010). Confinement effectiveness of FRP in retrofitting circular concrete columns under simulated seismic load. Journal of Composites for Construction, 14(5), 531–540.CrossRef
Zurück zum Zitat Haselton, C. B., Liel, A. B., Lange, S. T., & Deierlein, G. G. (2008). Beam-column element model calibrated for predicting flexural response leading to global collapse of RC frame buildings. Berkeley, U.S.: Pacific Earthquake Engineering Research Center, University of California. Haselton, C. B., Liel, A. B., Lange, S. T., & Deierlein, G. G. (2008). Beam-column element model calibrated for predicting flexural response leading to global collapse of RC frame buildings. Berkeley, U.S.: Pacific Earthquake Engineering Research Center, University of California.
Zurück zum Zitat Hashemi, M. J., Masroor, A., & Mosqueda, G. (2014). Implementation of online model updating in hybrid simulation. Earthquake Engineering and Structural Dynamics, 43(3), 395–412.CrossRef Hashemi, M. J., Masroor, A., & Mosqueda, G. (2014). Implementation of online model updating in hybrid simulation. Earthquake Engineering and Structural Dynamics, 43(3), 395–412.CrossRef
Zurück zum Zitat Hashemi, M. J., & Mosqueda, G. (2014). Innovative substructuring technique for hybrid simulation of multistory buildings through collapse. Earthquake Engineering and Structural Dynamics, 43(14), 2059–2074.CrossRef Hashemi, M. J., & Mosqueda, G. (2014). Innovative substructuring technique for hybrid simulation of multistory buildings through collapse. Earthquake Engineering and Structural Dynamics, 43(14), 2059–2074.CrossRef
Zurück zum Zitat Hashemi, M. J., Mosqueda, G., Lignos, D. G., Medina, R. A., & Miranda, E. (2016). Assessment of numerical and experimental errors in hybrid simulation of framed structural systems through collapse. Journal of Earthquake Engineering, 20(6), 885–909.CrossRef Hashemi, M. J., Mosqueda, G., Lignos, D. G., Medina, R. A., & Miranda, E. (2016). Assessment of numerical and experimental errors in hybrid simulation of framed structural systems through collapse. Journal of Earthquake Engineering, 20(6), 885–909.CrossRef
Zurück zum Zitat Ibarra, L. F., & Krawinkler, H. (2005). Global collapse of frame structures under seismic excitations. Stanford, U.S.: Stanford University, John A. Blume Earthquake Engineering Center. Ibarra, L. F., & Krawinkler, H. (2005). Global collapse of frame structures under seismic excitations. Stanford, U.S.: Stanford University, John A. Blume Earthquake Engineering Center.
Zurück zum Zitat Ibarra, L. F., Medina, R. A., & Krawinkler, H. (2005). Hysteretic models that incorporate strength and stiffness deterioration. Earthquake Engineering and Structural Dynamics, 34(12), 1489–1511. Ibarra, L. F., Medina, R. A., & Krawinkler, H. (2005). Hysteretic models that incorporate strength and stiffness deterioration. Earthquake Engineering and Structural Dynamics, 34(12), 1489–1511.
Zurück zum Zitat Jiang, S.-F., Zeng, X., Shen, S., & Xu, X. (2016). Experimental studies on the seismic behavior of earthquake-damaged circular bridge columns repaired by using combination of near-surface-mounted BFRP bars with external BFRP sheets jacketing. Engineering Structures, 106, 317–331.CrossRef Jiang, S.-F., Zeng, X., Shen, S., & Xu, X. (2016). Experimental studies on the seismic behavior of earthquake-damaged circular bridge columns repaired by using combination of near-surface-mounted BFRP bars with external BFRP sheets jacketing. Engineering Structures, 106, 317–331.CrossRef
Zurück zum Zitat Lynn, A. C., Moehle, J. P., Mahin, S. A., & Holmes, W. T. (1996). Seismic evaluation of existing reinforced concrete building columns. Earthquake Spectra, 12(4), 715–739.CrossRef Lynn, A. C., Moehle, J. P., Mahin, S. A., & Holmes, W. T. (1996). Seismic evaluation of existing reinforced concrete building columns. Earthquake Spectra, 12(4), 715–739.CrossRef
Zurück zum Zitat Ma, G., & Li, H. (2015). Experimental study of the seismic behavior of predamaged reinforced-concrete columns retrofitted with basalt fiber–reinforced polymer. Journal of Composites for Construction. doi:10.1061/(ASCE)CC.1943-5614.0000572. Ma, G., & Li, H. (2015). Experimental study of the seismic behavior of predamaged reinforced-concrete columns retrofitted with basalt fiber–reinforced polymer. Journal of Composites for Construction. doi:10.​1061/​(ASCE)CC.​1943-5614.​0000572.
Zurück zum Zitat Nakamura, T., & Yoshimura, M. (2002). Gravity load collapse of reinforced concrete columns with brittle failure modes. Journal of Asian Architecture and Building Engineering, 1(1), 21–27.CrossRef Nakamura, T., & Yoshimura, M. (2002). Gravity load collapse of reinforced concrete columns with brittle failure modes. Journal of Asian Architecture and Building Engineering, 1(1), 21–27.CrossRef
Zurück zum Zitat Nakata, N. (2007). Multi-dimensional mixed-mode hybrid simulation, control and applications. Ph.D. Dissertation. U.S.: University of Illinois at Urbana-Champaign. Nakata, N. (2007). Multi-dimensional mixed-mode hybrid simulation, control and applications. Ph.D. Dissertation. U.S.: University of Illinois at Urbana-Champaign.
Zurück zum Zitat Ozcan, O., Binici, B., & Ozcebe, G. (2010). Seismic strengthening of rectangular reinforced concrete columns using fiber reinforced polymers. Engineering Structures, 32(4), 964–973.CrossRef Ozcan, O., Binici, B., & Ozcebe, G. (2010). Seismic strengthening of rectangular reinforced concrete columns using fiber reinforced polymers. Engineering Structures, 32(4), 964–973.CrossRef
Zurück zum Zitat Pan, P., Nakashima, M., & Tomofuji, H. (2005). Online test using displacement-force mixed control. Earthquake Engineering and Structural Dynamics, 34(8), 869–888.CrossRef Pan, P., Nakashima, M., & Tomofuji, H. (2005). Online test using displacement-force mixed control. Earthquake Engineering and Structural Dynamics, 34(8), 869–888.CrossRef
Zurück zum Zitat PEER. (2013). Structural performance database. Berkeley, U.S.: Pacific Earthquake Engineering Research Center, University of California. PEER. (2013). Structural performance database. Berkeley, U.S.: Pacific Earthquake Engineering Research Center, University of California.
Zurück zum Zitat Porter, K., Kennedy, R., & Bachman, R. (2007). Creating fragility functions for performance-based earthquake engineering. Earthquake Spectra, 23(2), 471–489.CrossRef Porter, K., Kennedy, R., & Bachman, R. (2007). Creating fragility functions for performance-based earthquake engineering. Earthquake Spectra, 23(2), 471–489.CrossRef
Zurück zum Zitat Realfonzo, R., & Napoli, A. (2009). Cyclic behavior of RC columns strengthened by FRP and steel devices. Journal of Structural Engineering, 135(10), 1164–1176.CrossRef Realfonzo, R., & Napoli, A. (2009). Cyclic behavior of RC columns strengthened by FRP and steel devices. Journal of Structural Engineering, 135(10), 1164–1176.CrossRef
Zurück zum Zitat Realfonzo, R., & Napoli, A. (2012). Results from cyclic tests on high aspect ratio RC columns strengthened with FRP systems. Construction and Building Materials, 37, 606–620.CrossRef Realfonzo, R., & Napoli, A. (2012). Results from cyclic tests on high aspect ratio RC columns strengthened with FRP systems. Construction and Building Materials, 37, 606–620.CrossRef
Zurück zum Zitat Schellenberg, A. H., Mahin, S. A., & Fenves, G. L. (2009). Advanced implementation of hybrid simulation. Berkeley, U.S.: Pacific Earthquake Engineering Research Center, University of California. Schellenberg, A. H., Mahin, S. A., & Fenves, G. L. (2009). Advanced implementation of hybrid simulation. Berkeley, U.S.: Pacific Earthquake Engineering Research Center, University of California.
Zurück zum Zitat Scott, M. H., & Fenves, G. L. (2006). Plastic hinge integration methods for force-based beam-column elements. Journal of Structural Engineering, 132(2), 244–252.CrossRef Scott, M. H., & Fenves, G. L. (2006). Plastic hinge integration methods for force-based beam-column elements. Journal of Structural Engineering, 132(2), 244–252.CrossRef
Zurück zum Zitat Sezen, H. (2002). Seismic response and modeling of reinforced concrete building columns. Ph.D. Dissertation. Berkeley, U.S.: University of California. Sezen, H. (2002). Seismic response and modeling of reinforced concrete building columns. Ph.D. Dissertation. Berkeley, U.S.: University of California.
Zurück zum Zitat Shan, B., Xiao, Y., & Guo, Y. R. (2006). Residual performance of FRP-retrofitted RC columns after being subjected to cyclic loading damage. Journal of Composites for Construction, 10(4), 304–312.CrossRef Shan, B., Xiao, Y., & Guo, Y. R. (2006). Residual performance of FRP-retrofitted RC columns after being subjected to cyclic loading damage. Journal of Composites for Construction, 10(4), 304–312.CrossRef
Zurück zum Zitat Wang, N. Y., & Ellingwood, B. R. (2015). Limit state design criteria for FRP strengthening of RC bridge components. Structural Safety, 56, 1–8.CrossRef Wang, N. Y., & Ellingwood, B. R. (2015). Limit state design criteria for FRP strengthening of RC bridge components. Structural Safety, 56, 1–8.CrossRef
Zurück zum Zitat Wang, T., Mosqueda, G., Jacobsen, A., & Cortes-Delgado, M. (2012). Performance evaluation of a distributed hybrid test framework to reproduce the collapse behavior of a structure. Earthquake Engineering & Structural Dynamics, 41(2), 295–313.CrossRef Wang, T., Mosqueda, G., Jacobsen, A., & Cortes-Delgado, M. (2012). Performance evaluation of a distributed hybrid test framework to reproduce the collapse behavior of a structure. Earthquake Engineering & Structural Dynamics, 41(2), 295–313.CrossRef
Zurück zum Zitat Wei, H., Wu, Z. M., Guo, X., & Yi, F. M. (2009). Experimental study on partially deteriorated strength concrete columns confined with CFRP. Engineering Structures, 31(10), 2495–2505.CrossRef Wei, H., Wu, Z. M., Guo, X., & Yi, F. M. (2009). Experimental study on partially deteriorated strength concrete columns confined with CFRP. Engineering Structures, 31(10), 2495–2505.CrossRef
Zurück zum Zitat Wibowo, A., Wilson, J. L., Lam, N. T. K., & Gad, E. F. (2014). Drift performance of lightly reinforced concrete columns. Engineering Structures, 59, 522–535.CrossRef Wibowo, A., Wilson, J. L., Lam, N. T. K., & Gad, E. F. (2014). Drift performance of lightly reinforced concrete columns. Engineering Structures, 59, 522–535.CrossRef
Zurück zum Zitat Wang, Z. Y., Wang, D. Y., Smith, S. T., & Lu, D. G. (2012). Experimental testing and analytical modeling of CFRP-confined large circular RC columns subjected to cyclic axial compression. Engineering Structures, 40, 64–74. Wang, Z. Y., Wang, D. Y., Smith, S. T., & Lu, D. G. (2012). Experimental testing and analytical modeling of CFRP-confined large circular RC columns subjected to cyclic axial compression. Engineering Structures, 40, 64–74.
Zurück zum Zitat Yalcin, C., Kaya, O., & Sinangil, M. (2008). Seismic retrofitting of R/C columns having plain rebars using CFRP sheets for improved strength and ductility. Construction and Building Materials, 22(3), 295–307.CrossRef Yalcin, C., Kaya, O., & Sinangil, M. (2008). Seismic retrofitting of R/C columns having plain rebars using CFRP sheets for improved strength and ductility. Construction and Building Materials, 22(3), 295–307.CrossRef
Zurück zum Zitat Zhong, W. (2005). Fast hybrid test system for substructure evaluation. Ph.D. Dissertation. U.S.: University of Colorado Boulder. Zhong, W. (2005). Fast hybrid test system for substructure evaluation. Ph.D. Dissertation. U.S.: University of Colorado Boulder.
Metadaten
Titel
Application of the MAST System for Collapse Experiments
verfasst von
Riadh Al-Mahaidi
Javad Hashemi
Robin Kalfat
Graeme Burnett
John Wilson
Copyright-Jahr
2018
Verlag
Springer Singapore
DOI
https://doi.org/10.1007/978-981-10-5867-7_4