Sunteți pe pagina 1din 16

Engineering Structures 24 (2002) 719734 www.elsevier.

com/locate/engstruct

Seismic performance of a 3-story RC frame in a low-seismicity region


Han-Seon Lee *, Sung-Woo Woo
Department of Architectural Engineering, Korea University, Seoul 136-701 Received 16 April 2001; received in revised form 7 November 2001; accepted 14 November 2001

Abstract The objectives of the research stated herein are to investigate the seismic performance of a 3-story reinforced concrete (RC) ordinary moment-resisting frame, which has not been engineered to resist earthquake excitations, and to evaluate the reliability of the available static and dynamic inelastic analysis techniques. A 1:5 scale model constructed according to the Korean practice of nonseismic detailing and the similitude law was subjected to a series of the shaking table motions simulating Taft N21E component earthquake ground accelerogram. Due to the limitation in the capacity of the used shaking table, a pushover test was performed to observe the ultimate capacity of the structure after earthquake simulation tests. The model showed the linear elastic behavior under the Taft N21E motion with the peak ground acceleration of 0.12g, representing the design earthquake in Korea. The model revealed fairly good resistance to the higher levels of earthquake simulation tests though it was not designed against earthquakes. The main components of its resistance to the high level of earthquakes appear to be (1) the high overstrength, (2) the elongation of the fundamental period, (3) the minor energy dissipation by inelastic deformations, and (4) the increase of the damping ratio. The drifts of the model under these tests were approximately within the allowable limit. Analysis of the results of the pushover test reveals that the model structure has the overall displacement ductility ratio of 2.4 and the overstrength coefcient of approximately 8.7. The evaluation of the accuracy of analytical simulation by IDARC-2D leads to the conclusion that while global and local behaviors can be, in general, simulated with limited accuracy in the dynamic nonlinear analysis, it is easy to obtain a fairly high level of accuracy in the prediction of global behavior in the static nonlinear analysis. 2002 Elsevier Science Ltd. All rights reserved.
Keywords: Reinforced concrete; Earthquake simulation test; Pushover test; Nonlinear analysis; Overstrength; Nonseismic detailing

1. Introduction Recently, minor earthquakes have occurred over 20 times a year in Korea. The earthquake of December 13, 1996 at Yeongweol in Korea, which is known to have a magnitude of 4.5 on the Richter scale imposed signicant nonstructural damages [1]. These recent earthquakes indicate that the Korean Peninsular is no longer safe from seismic hazard. If a severe earthquake such as the 1995 Kobe earthquake should occur in Seoul, the damage would be tremendous. Most building structures in Korea, which are normally medium- to low-rise reinforced concrete (RC) frames,

* Corresponding author. Tel.: +82-2-3290-3337; fax: +82-2-9217947. E-mail address: hslee@korea.ac.kr (H.-S. Lee).

have not been engineered to resist major or moderate earthquakes. Therefore, should any major earthquake occur, the damage or collapse of not only general commercial buildings, but also public-service buildings such as police ofces, communication centers and hospitals, would implement very large life and economic losses as well as cause the critical interference with the function of the nation. Several researchers in the eastern and central United States have already performed research, including earthquake simulation tests, on the seismic capacity of gravityload-designed RC structures [25]. However, attention concentrated mainly on the detail practice in North America, particularly according to American Concrete Institute (ACI) 318 code [6]. The practice in reinforcement detailing and construction in Korea is somewhat different from that of the US, which will be explained in the section on the design of the model.

0141-0296/02/$ - see front matter 2002 Elsevier Science Ltd. All rights reserved. PII: S 0 1 4 1 - 0 2 9 6 ( 0 1 ) 0 0 1 3 5 - 3

720

H.-S. Lee, S.-W. Woo / Engineering Structures 24 (2002) 719734

The objectives of this research are; (1) to observe the actual response of this kind of low-rise RC ordinary moment-resisting frame with nonseismic detailing when subjected to various levels of earthquake ground motions, (2) to get the information on the ultimate capacities (strength, deformability and so on) of the structure, and (3) to provide the calibration to, or to check the reliability of, the available static and dynamic inelastic analysis techniques. Considering the capacity of the shaking table to be used, the reduction scale for the model was determined as 1:5. Using the techniques for manufacturing the model according to the similitude requirements developed through other researches [7,8], a 1:5 scale 2bay 3-story RC frame model was constructed. This model was then subjected to the shaking table motions simulating Taft N21E component earthquake ground motions, whose magnitude of peak ground acceleration (PGA) was modied to approximately 0.12, 0.2, 0.3 and 0.4 g. Before and after each earthquake simulation test, free vibration tests were performed to determine the change in the natural period and the damping ratio of the model. The global behavior and damage pattern were observed. The lateral accelerations and displacements at each story and the local deformations at the critical regions of the structure were measured. The base shear was measured using self-made load cells. Since the ultimate capacity of the structure could not be found due to the limitation in the capacity of the shaking table, a pushover test was performed to observe this capacity of the structure after a series of earthquake simulation tests. Based on all the test results, the interpretation on the response of the model is carried out. The computer code IDARC-2D [9], one of the codes widely used in the world for the nonlinear dynamic and static analyses of RC frame structures, was adopted for the analysis to investigate the correlation between the results of (earthquake simulation and pushover) tests and analyses. Although the time history analyses corresponding to all the earthquake simulation tests were performed, the correlation of experiment and analysis will be investigated only for the case of earthquake simulation test TFTF04 in this paper.

members and the details regarding transverse steel, anchorage and splice are shown in Fig. 1(ch). The important characteristics in the Korean detailing practice are as follows: (1) the splice is located at the bottom of the column, (2) the spacing of hoops is relatively large, (3) seismic hooks are not used, (4) connement reinforcements are not used in beamcolumn joints, and (5) the special style of anchorage in the joints. That is, the length of tension and compression anchorage are usually 40 and 25 db respectively, from the critical section, where db means the nominal diameter of reinforcement. Moreover, the length of the tail in the hook is included in this anchorage length and the tails of the anchorage of the bottom bars in beams usually direct downward into the exterior columns as shown in Fig. 1(h). 2.2. Model reinforcement and model concrete It is essential to maintain the similitude in the material properties between prototype and model reinforcement. However, it was difcult to make the cross sections of the model reinforcement conform exactly to the similitude law. So, the yield forces rather than yield stresses were selected as the target to be achieved in annealing the model reinforcement. An electric furnace with a 3zone vacuum tube was designed and used. The deformation on the surface of the model reinforcements was made using a deforming device. Reinforcing bars D22 and D10 in the full-scale structure match D3 and D2 in the model, respectively. The target yield forces derived from similitude requirements are shown in Table 1. The achieved average yield force of 5.0 kN is approximately 10% less than the target yield force in the case of D3. The model concrete was made using type I Portland cement. The average strength of the model concrete at the time of testing was about 21.6 MPa. 2.3. Instrumentation and experimental setup The used shaking table is 35 m with one degree of freedom. The data were acquired simultaneously at the rate of 300 Hz in 32 channels. Displacement transducers, accelerometers and load cells were used to measure the lateral displacement and the angular rotations in some ends of beams and columns, acceleration at each story, and shear forces on the columns of the rst story. Fig. 2(a) and Plate 1 show the view of the experimental setup and instrumentations for the shaking table test. The instrumentation was conned to only one frame due to the limitation in the number of available channels except the load cells at the rst-story columns. To measure story drifts, a reference frame was used as shown in Fig. 2. The white-noise (random-vibration) test, which was performed before the earthquake simulation tests, indicated that the natural frequency of the reference frame is

2. Experiment 2.1. Design of the model The prototype of this test model was adopted from a building structure for the police ofce, actually built and in use in Korea. The plan and elevation of the 1:5 scale model are shown in Fig. 1(a,b). The compressive strength of concrete, f c, in the prototype structure is assumed to be 20.6 MPa and the nominal yield strength of reinforcement, 294.2 MPa. The typical sections of

H.-S. Lee, S.-W. Woo / Engineering Structures 24 (2002) 719734

721

Fig. 1. Plan, elevation and details (unit: mm). (a) Plan; (b) Elevation; (c) Section C1C1 ; (d) Section C2C2 ; (e) Section C3C3 ; (f) Section B1 B1 ; (g) Section B2B2 ; and (h) Anchorage detail of beam bars in exterior joint.

approximately 40 Hz. Therefore, the reference frame was considered to have the sufcient rigidity to measure accurately the drifts of the model. The load cells were designed and manufactured following the reports of ElAttar et al. [4] and Bracci et al. [2] and calibrated by using a universal testing machine (UTM). The volume of the model is reduced to 1/53 of the prototype while the similitude law, premising the agreement in the stressstrain relation of the material between the prototype and the model, requires the mass to be

reduced to 1/52 [10]. Therefore, compensation for the difference in the mass was articially made by adding concrete blocks as shown in Fig. 2(a). The effective weight of the model with concrete blocks is estimated to be 100.6 kN while the weight ideally required by the similitude law is 96.1 kN. The error is approximately 5%. The hinges between the model and the concrete blocks were designed to transmit only the vertical and horizontal forces, excluding the moments, due to the mass of the concrete block.

722

H.-S. Lee, S.-W. Woo / Engineering Structures 24 (2002) 719734

Table 1 Similitude requirements for reinforcement Member and use Beam and column Prototype (stress) Bar Yield strength Stress (MPa) Stirrup and hoop Main reinforcement
*

1/5 Model (force) Tensile strength Stress (MPa) 530 530 Force (kN)(2) 37.8 205.3 D2 D3 Bar Yield strength Stress (MPa) 325 784 Tensile strength (2)1/25 (kN) 1.51 8.21

Force (kN)(1) 25.5 138.5

(1)1/25* Stress (kN) (MPa) 1.02 5.54 481 1161

D10 D22

358 358

Target yield force in annealing.

Fig. 2.

Instrumentation and experimental setup. (a) Earthquake simulation test; and (b) Pushover test.

The experimental setup for the pushover test is shown in Fig. 2(b). Steel plates, each of which has the dimension WDL=9435 cm (0.097 kN) were used as the articial mass. The effective weight of the model with steel plates is estimated to be 100.7 kN. The roof drift was obtained by averaging the two measured values in both Frame A and Frame B. Two different data acquisition systems were used due to the limitations in the number of available channels in the data acquisition sys-

tems, but the measured data were synchronized with each others system by comparing the displacements of transducers T5 and T6 installed at the same location in Fig. 2(b) but belonging to the different data acquisition system. The experimental results were interpreted, assuming that the behavior of Frame A represents that of the whole model structure in both earthquake simulation test and pushover test. The displacements at the second oor (T2) and third oor (T3) were those meas-

H.-S. Lee, S.-W. Woo / Engineering Structures 24 (2002) 719734

723

Plate 1.

Experimental setup for shaking table tests.

ured in the middle of both frames, and the lateral force distribution was maintained in the shape of an inverted triangle by using the whife tree. 2.4. Experimental program Since there was no recorded strong-motion accelerogram as of the time of this experiment in Korea, the input motion to the shaking table was derived from the recorded Taft N21E component by adjusting the peak ground acceleration (PGA) and compressing the time scale by the factor of 5 according to the similitude law [10]. The most important reason for this selection was that the test results in this study can be compared easily with those of several previous earthquake simulation tests [24,11] which used the same accelerogram, and the other one is that damage potential of Taft N21E component accelerogram seems to be relatively high. The design earthquake dened in Korean seismic code [12] is the event expected in the recurrence period of 475 years (the probability of exceedance of 10% in 50 yr). The elastic response spectra of the input Taft N21E (PGA = 0.12 g) motion and the output table motion are shown in Fig. 3. From this gure, it can be found that

the delity of the shaking table is satisfactory. The comparison of the design spectrum according to the Korean seismic code [12] with the response spectrum of the output motion indicates that the dynamic amplication implicit in the design spectrum of this code is underestimated in the range of short periods whereas it is overestimated in the range of long periods. The nal design spectrum by UBC [13] is quite similar to the Korean design spectrum though the used R factors are different. The program of tests is shown in Table 2. Each test has the signicance stated in the column containing remarks. For earthquake simulation tests, the model structure did not show serious damage even after test TFTF04. However, due to the limitation in the capacity of the shaking table, it was impossible to implement a higher level of earthquake simulation test. Therefore, in order to get the information on the capacities (strength, deformability and so on) of the model structure, a pushover test (or monotonically-increasing lateral-load static test) was conducted.

3. Results of earthquake simulation tests and interpretation 3.1. Global responses Free vibration tests were performed by enforcing an initial lateral displacement of approximately 12 mm at the roof of the model, and releasing. From these tests, the natural periods and damping ratios of the model were obtained by using the Fourier transform and logarithmic decrement method. Table 3 shows that natural periods and damping ratios tend to increase as the model experiences higher levels of ground motions. The drifts at the roof were measured at two locations. The test results indicate that almost negligible torsional behaviors occurred in this model for both earthquake simulation and pushover tests. Therefore, the measured data in one of the two plane frames are assumed to represent the behaviors of both frames. Table 4 summarizes the measured maximum response quantities. Time histories of the oor accelerations for TFTF04 are given in Fig. 4, which indicates that the response has the predominant period approximately equal to the period of the rst mode of the model. Table 4 shows the maximum interstory drift indices (I.D.I.) at the time when the roof undergoes the maximum drift. For the design earthquake of Korea (TFTF012), the model has a maximum I.D.I. of 0.26% which is much less than the maximum allowable value of 1.5%. However, the maximum I.D.I. for the test of TFTF04 appears to be 1.68%. Therefore, regarding drift control, the behavior of the model appears to be satisfactory. From the proles of measured drift and the interstory drift indices in Fig. 5, it can be noted that the

Fig. 3. Response spectrum for input and output table motions and design spectra compressed by the scale of 1: 5 (soil factor S 1.2).

724

H.-S. Lee, S.-W. Woo / Engineering Structures 24 (2002) 719734

Table 2 Test program Identication of test Earthquake simulation test TFTF012 TFTF02 TFTF03 TFTF04 PUSH PGA (g) 0.12 0.2 0.3 0.4 Remarks (return period) Design earthquake in Korea (500 years) Max. earthquake in Korea (1000 years) Max. considered earthquake in Korea (2000 years) Severe earthquake in high seismic regions of the world Ultimate capacity of the structure

Pushover static test

Table 3 Natural period and damping ratio by free-vibration tests Identication of test Before TFTF012 After TFTF012 After TFTF02 After TFTF03 After TFTF04 Natural period (sec) Damping ratio (%) 0.226 0.229 0.265 0.265 0.317 4.1 4.6 4.4 5.8 7.9

less than six stories need not be designed against the earthquake in Korea, it is worthwhile to evaluate the test results with regards to the current Korean seismic code [12]. According to this code, the seismic coefcient, Cw, for the RC ordinary moment-resisting frame is determined as follows: Cw V W AISC R (0.12)(1.0)(1.39) 3.5 0.048 (1)

displaced shapes of the frame remained almost unchanged during the course of the tests and the deformations tend to be concentrated at the second story for the higher levels of motions. The model behaves linear-elastically under TFTF012, which represents the design earthquake in Korea. The developed maximum base shear is 17.64 kN. The maximum base shears for other table motions are given in Table 4. The model has yielded under TFTF04. Here the yielding base shear appears to be 37.14 kN. Though the building structures built before 1988 or of
Table 4 Summary of measured maximum response amplitudes Test Table acceleration (g) 0.138 0.21 0.31 0.4 Roof drift (mm) I.D.I. (%)

where V: base shear, W: 96.1 kN (effective weight of the structure), A: 0.12 (zone factor), I: 1.0 (importance factor), T: 0.23 5 (scale factor) =0.514 s (natural 1 1.162 1.5 period), S: 1.2 (soil factor), C: 1.2 T (dynamic factor), SC: 1.39 ( 1.75), R: 3.5 (response modication factor; ordinary moment frame). Table 4 also gives the coefcient derived from the experiment. For the table motion (TFTF012) simulating the design earthquake in Korea, the actual developed maximum base shear is about 3.8 times the design base shear. It can be noted from Table 4 that this model

Roof acceleration (g) 0.28 0.53 0.61 0.69

Base shear (load cell) (kN) 17.64 30.77 35.08 37.14

V/W

TFTF012 TFTF02 TFTF03 TFTF04

4.5 14.06 17.87 29.88

0.26 0.78 1.08 1.68

0.18 0.32 0.37 0.39

Fig. 4.

Time histories of oor accelerations for TFTF04.

H.-S. Lee, S.-W. Woo / Engineering Structures 24 (2002) 719734

725

Fig. 5. Proles of drift envelopes and interstory drift indices in earthquake simulation test. (a) Drift envelopes; and (b) Interstory drift indices (I.D.I.).

resisted the table motions of higher PGAs with a very high lateral strength of up to 0.39 W. In addition, the degradation of stiffness has elongated the natural period of the model as shown in Table 3 and hence caused the decreased dynamic amplication factor. The test of TFTF04 has shown that the model yielded and that the story displacement ductility at the rst story was about 1.3. From these ndings, it can be concluded that the model behaved linear-elastically under the design earthquake (PGA 0.12 g) while the overstrength up to 0.39 W, the degradation of stiffness, the energy dissipation due to inelastic behavior in the model, and the increase of damping ratio were the additional major contributors to its resistance against the higher table motions (PGA=0.2, 0.3 and 0.4 g). The time histories of the total absorbed energy at each story for the cases of TFTF012 and TFTF04 are depicted in Fig. 6. From this gure we can nd that the total amount of the absorbed energy in the case of TFTF04 is about 13 times that in the case of TFTF012, and that most of the energy absorption took place in the lower two stories. The model did not show serious damage even after the test of TFTF04 simulating a severe earthquake in a high-seismicity zone of the world. There was no apparent crack after tests TFTF012 and TFTF02. After test

TFTF03, a minor exural crack could be noticed in beam end at the exterior joint of the second oor. Several cracks occurred after test TFTF04. The beam ends at the exterior joint of the second oor revealed new cracks, which imply a yield in these regions. In particular, the exterior columns at the rst story have revealed both exural and shear cracks. However, it is interesting to note that the interior columns, which have experienced larger rotations, did not show any apparent cracks. 3.2. Local responses The distance over which the rotational angle was measured is the full depth for beams, and the cross-sectional dimension parallel to the shaking direction for columns. Fig. 7 shows the example of some time histories of angular rotations in the ends of members for earthquake simulation test. The bottom end of the interior column at the second story has shown the largest angular rotation of all that were measured. Relatively large rotations occurred for the interior columns (R4, R3, and R8 in Fig. 2), when compared with those for the exterior columns (R6, R7, and R9 in Fig. 2). Since the model has smaller cross-sections in the interior columns than in the exterior columns, these larger rotations at the interior column could be expected. For the exterior joint at the second oor, the beam (R5) has larger rotation than the

Fig. 6. Time histories of absorbed energy in earthquake simulation test.

Fig. 7.

Angular rotations under TFTF04 (exterior joint).

726

H.-S. Lee, S.-W. Woo / Engineering Structures 24 (2002) 719734

connected columns (R6 and R7). On the contrary, the angular rotations at the beams (R1 and R2) are generally much less than those at the columns (R3 and R4) in the case of the interior joint. Fig. 8 shows the distribution of shear forces in the columns. It is interesting to note that column (1) and column (3) have a directional bias in shear forces due to the inuence of the axial compressive force on the shear resistance.

4. Correlation of analytical and experimental dynamic responses 4.1. Analytical model The computer code IDARC-2D [9], one of the codes widely used in the world for the nonlinear dynamic and static analyses of RC framed structures, was adopted. This code utilizes a global Takeda-like model. The objective of this analytical study is to nd the most appropriate values of parameters in the analysis by IDARC-2D to simulate the responses given by experiments, and then to evaluate the degree of accuracy in the obtained simulations. Eventually, this evaluation of reliability for IDARC-2D will lead to the more careful interpretation of analysis results for other types of RC frame structures such as schools, hospitals, and the like. The material models used to derive the relation between moment and curvature at critical sections are shown in Fig. 9. The program RESPONSE, developed by Felber and Andreas [14], was used to obtain the envelope curve for the momentcurvature relations, as depicted in Fig. 10. The effective width (410 mm) of ACI code 318-95 [6] was used to model the relation between the curvature and the moment for the time history analyses. The hysteretic model incorporates stiffness degradation (HC, a), strength deterioration (HBD, HBE, b), non-symmetric response, slip-lock (HS, g), and a trilinear monotonic envelope. The model traces the hysteretic

behavior of an element as it changes from one linear stage to another, depending on the history of deformations. For a complete description of the hysteretic model, refer to Park et al. [15]. Aycardi, Mander, and Reinhorn [5] used a 0.5, b 0.04, and g 0.7 based on the experimental test results for elements. Stiffness degradation is severer in the model than in the prototype, and an value between 0.5 and 1.0 has been used in the analytical model, which is scaled as 1/4 or below [9]. In this study, the same value of hysteretic parameters as adopted by Aycardi et al. are used to simulate the test results. The articial weight (concrete blocks in this case) loaded to compensate the mass according to the similitude law acts as concentrated loads on the girders. Hence, the girders are divided into three elements which also properly represent the change in the reinforcement of the sections. 4.2. Global responses The parameters determining hysteretic behavior in the Mf relation have actually been adjusted to simulate most closely the response, particularly the roof drift. Though the roof drift history shown in Fig. 11 is the best simulation among all the trials, there is still a discrepancy of 4.11 mm (about 14%) in maximum values, and in phase in the latter part. Though the peak drift at the roof in the analysis is smaller than in the experiment, the peak response acceleration at the roof of the analysis appears to be a little larger than that of the experiment. Nevertheless, the analysis reveals a smaller maximum base shear than the experiment in Fig. 12. This discrepancy in the maximum base shear between analysis and experiment can be attributed to the stronger effects of the second or higher mode in the analysis. In general, the histories of story drifts are similar in shape in both cases of analysis and experiment, indicating that the rst mode governs, though the peak values differ. 4.3. Local responses The time histories of the column shear in Fig. 13 imply that the analysis could not simulate the bias in the shear force caused by the increase of shear stiffness due to the increase of the axial compressive forces in columns. Fig. 14 compares the time histories of angular rotations at the bottom of the second-story column obtained from both the experiment and analysis. The angular rotations were calculated by multiplying the curvature at the end of members in analysis with the same lengths over which the rotations were measured in the experiment. Generally, the magnitude of the angular rotation in the analysis is much smaller than that in the experiment. The discrepancy in magnitude is considered

Fig. 8.

Time histories of column shears under TFTF04.

H.-S. Lee, S.-W. Woo / Engineering Structures 24 (2002) 719734

727

Fig. 9.

Material model. (a) Concrete; and (b) Reinforcement.

Fig. 10.

Momentcurvature envelope for T beams by RESPONSE [14].

Fig. 11.

Comparison of roof drift history.

Fig. 12.

Comparison of base shears derived from column shears.

to be due to the limitation in the member modeling of spread plasticity in IDARC-2D. 5. Interpretation on results of pushover test and analyses 5.1. Analytical model The model for the pushover analysis is the same as that for the time history analysis except that the articial weight needed to compensate the gravity weight was distributed over the slabs using steel plates rather than being concentrated weights as in the case of earthquake simulation tests. To take into consideration the increase in the contribution of slabs at the ultimate strength of the structure, two types of sections were considered to model T beams (the effective width of T beam, = 410 mm (PUSH-I), =840 mm (PUSH-II)) as shown in Fig. 10.

728

H.-S. Lee, S.-W. Woo / Engineering Structures 24 (2002) 719734

Fig. 13.

Time histories of column shear at rst story (column (1)).

Fig. 14. Time histories of angular rotations in interior column of second story for TFTF04.

5.2. Global responses The relations between the lateral load and the roof drift obtained through the pushover test and analyses (PUSH-I and PUSH-II) are shown in Fig. 15. The ndings regarding this gure are as follows: (1) IDARC-2D could not predict or simulate the brittle failure at the point of ultimate strength. (2) The additional contribution of the enlarged width of the slab to the ultimate strength, which is introduced by using PUSH-II model instead of PUSH-I is only approximately 4 kN. (3)

Fig. 15.

Base shear versus roof drift in tests and analyses.

PUSH-II has indicated a still lower strength than that by experiment. The difference is about 8 kN (17%). In the pushover test, the model structure yielded gradually and underwent a series of abrupt strength drops after reaching the ultimate strength of 51.35 kN through brittle failures in the local regions, the locations of which could not be clearly identied, with the exception of the last failure by concrete crushing in the upper end of the interior column at the rst story. This clearly indicates that the model structure has the characteristics of the structure, which has limited ductility due to non-seismic details. The solid marker, representing the maximum base shear and the corresponding roof drift for each earthquake simulation test, was superposed on the graphs of the pushover test and analyses in Fig. 15. It can be seen from this gure that the markers, in general, comply with the curve given by the analysis PUSH-II. The discrepancy in the initial part between these markers (or the curve of the analysis PUSH-II) and the curve obtained through pushover test implies the stiffness degradation which had occurred due to the damages implemented during the course of previous earthquake simulation tests. The reason for the discrepancy in the latter part between the curves of the pushover test and analysis PUSH-II is considered the strain-aging effect [16,17] in the model reinforcement. Erasmus [17] showed that the effect of strain aging can induce the increase of more than 75% in the discontinuous yielding stress. In other words, the model structure showed the clear yielding phenomenon for the earthquake simulation test of TFTF04. Additionally, the time that elapsed between the earthquake simulation tests and the pushover test was more than one year. These two facts must have affected the increase of the yielding strength of the model reinforcement through the effect of strain aging. However, the hollow markers, which indicate the maximum base shear and the corresponding roof drift derived from the time history nonlinear analyses of TFTF012 and TFTF04 by IDARC-2D, deviate signicantly from the results of the earthquake simulation test and pushover test. This means the dynamic nonlinear analysis does not necessarily provide a higher level of accuracy or reliability than the static nonlinear analysis, as is commonly expected. With the assumption that the results of analysis PUSH-II would be the actual behavior of the model structure if the model structure had not experienced the previous earthquake simulation tests and the effect of strain aging, the displacement ductility ratio of the model structure turns outs to be approximately 2.4 as shown in Fig. 15 when the concept of equal energy and the elastoperfect-plastic (EPP) bilinear model are adopted. The effective yielding strength appears to be 40.0 kN in this gure. The maximum interstory drift at the rst story at the collapse state in pushover test was 28.7 mm (3.7%) as

H.-S. Lee, S.-W. Woo / Engineering Structures 24 (2002) 719734

729

mechanisms (Mechanism 1 and Mechanism 2), assuming the section properties as given PUSH-I and PUSH-II. With the section properties of PUSH-II, the two collapse mechanisms have a difference of only 0.7 kN (1.8%) in strength, which is almost negligible when the uncertainty in the material and section properties are considered. 5.3. Local reponses Fig. 20(a) shows the maximum angular rotations in the ends of members at the collapse state. Angular rotations in the ends of columns are within the range of approximately 0.04 rad and the largest value was obtained at the bottom of the exterior column located at the longer span of the rst story. The failure occurred in the mode of the exural compressive crushing in the upper end of the interior column at the rst story, and the experiment was stopped at this point. In the beam, the largest value of 0.026 rad was observed in the end of the longer-span beam adjacent to the exterior joint of the second oor. Fig. 21 shows the histories of angular rotations in the ends of beams and columns in the pushover test. The ends of columns (1 and 2) have rotations approximately twice larger than those of beams (3 and 4) under the same load at the interior joint. The end of beam (5) exhibits larger rotations than the ends of beams (3 and 4). The member ends (1, 6, and 7) deformed signicantly more even after the strength drops. This means, again, that the deformations were concentrated at the rst story after the strength drop. The rotational angles in the member ends at the time of the ultimate strength (not at the collapse state) in experiment (roof drift = 47.2 mm) are recorded in Fig. 20(b) to compare with those in analysis PUSH-II. The general trend is that the value obtained through analysis

Fig. 16. Story shear versus interstory drift in pushover test and analysis.

shown in Fig. 16. The increase of drift at only the rst story after the strength degradation indicates that a failure mechanism was formed in the rst story. The displacement ductility ratio at the rst story turns out to be about 1.6 for the ultimate strength, and 2.3 at the time of collapse. In case of analysis PUSH-II, these values should be replaced by 2.4 and 3.4, respectively, if the deformation capacity of the model structure is assumed to be the same as that obtained from the experiment. This clearly represents the characteristics of the RC frame with nonseismic detailing. Fig. 17 depicts the development of cracks in experiment. The nal distribution of plastic hinges observed from experiment, and those obtained through two analyses (PUSH-I and PUSH-II), are shown in Fig. 18(ac), respectively. The collapse mechanism in the experiment is the soft-story mechanism (Mechanism 1), while that of analysis is Mechanism 2 as shown in Fig. 19. Simple plastic analyses were performed for the two collapse

Fig. 17. Development of cracks in pushover test.

730

H.-S. Lee, S.-W. Woo / Engineering Structures 24 (2002) 719734

Fig. 18. Distribution of plastic hinges at the ultimate collapse state in pushover test and analysis. (a) Experiment; (b) Analysis (PUSHI); and (c) Analysis (PUSH-II).

Fig. 19. Collapse mechanism and collapse load in pushover test and analyses.

overestimates the angles in columns while it underestimates those in girders. However, it should be noted that the rotational angles in girders obtained by the analysis are very small when compared with those given by the experiment in the case of the interior joint of the second oor (analysis (rad)/experiment (rad) = 0.0025/0.014 and 0.0048/0.013).

cient, , based on the global behavior of a structure is dened as follows: Cy Cw (2)

6. Analysis of overstrength The overstrength factor is dened on the level of the whole structure as a ratio of the actual structural yield level to the code-prescribed strength demand arising from the application of prescribed loads and forces [18,19]. As shown in Fig. 22, the overstrength coef-

where Cy is the base shear coefcient corresponding to the idealized yield displacement y of the structure using the concept of equal energy; Cw is the code-prescribed unfactored design base shear coefcient. This factor can be further decomposed into two factors, s and y, as follows: Cs Cy Cw Cw
s

(3)

where Cs is the base shear coefcient corresponding to the rst signicant yield of the structure.

H.-S. Lee, S.-W. Woo / Engineering Structures 24 (2002) 719734

731

Fig. 20. Angular rotations (unit: 10 rotations at the maximum load.

rad). (a) Maximum angular rotations at collapse state in pushover test; and (b) Comparison of angular

Fig. 21.

Histories of angular rotations in pushover test. (a) Column ends; and (b) Beam ends.

Fig. 22. Typical global structural response idealized as linearly elastic-perfectly plastic curve.

Bertero et al. [20] indicated that it is due to the overstrength that buildings designed according to the code could resist severe seismic excitations. Shahrooz and Moehle [21] showed through an experimental study on

a 1/4-scale, 6-story RC structure that a structure designed for an unfactored base shear coefcient of 0.091 could theoretically resist 7.5 times as much. Miranda and Bertero [22], based on the study of low-rise buildings in Mexico City, have noted the value of overstrength in the range of 25. Meli [23] further showed that the available overstrength varies widely depending on the type of structure and on the characteristics of the ground motion. In the studies on the variation of overstrength factor for a typical interior frame of an ofce building located in a region of high seismic risk, Uang [19] showed that low-rise buildings usually have higher overstrength, while the overstrength is not very sensitive to the number of bays. Most experimental and analytical research in earthquake engineering is focused on high-risk seismic zones. While formulating the design codes, debate is normally focused on the seismic coefcient for higher zones; the coefcient for lower zones is simply prorated in proportion to the expected ground motion intensity in different zones. So, Jain and Navin [24] carried out a study

732

H.-S. Lee, S.-W. Woo / Engineering Structures 24 (2002) 719734

on the seismic overstrength of four-bay, three-, six-, and nine-story RC frames designed for seismic zones I to V as per Indian codes. The overstrength against lateral load is very signicantly affected by the factored gravity loads used in design. This results in overstrength being much higher for low seismic zones, for low-rise buildings, and for higher design live load. They showed that the overstrength of a three-story interior frame in zone V is 3.3, while it is as high as 15.0 in zone I. The overstrength factor, , of the model structure can be demonstrated by calculating both s Cs / Cw and Cy / Cs with respect to the exural moment y capacity as follows: First, the coefcient, s, can be calculated through the linear elastic analysis of the model structure up to the occurrence of the rst plastic hinge. It was found through linear elastic analyses that the model structure can meet the exural moment demands under the load cases of 1.4D+1.7L and 0.75(1.4D+1.7L1.87E), with the minimum margin of safety being 25% and 34%, respectively. However, since the gravity load condition during the earthquake simulation tests can be described as 1.0D, the ratios of the demanded exural moment to the capacity for the load case, 1.0D, are recalculated. Then, by comparing the demanded exural moment for the load case of earthquake (1.0E) to the reserved exural capacity (Capacity1.0D), we can obtain the coefcient, s, which is the least value among the values of 1

sis (without strain aging) Cy / Cs (40.00 kN) / (24.33 kN) 1.64 in Fig. 19. Therefore, the overstrength coefcient, , can be obtained by multiplying these two coefcients as 11.1 (with strain aging) or 8.7 (without strain aging). These large values of the overstrength coefcient account for the reason why the low-rise RC building structures have the large reserved strength for severe earthquakes even though they were designed only for the gravity loads in the lower seismic zones.

7. Conclusions 1. Though the model structure in this study was designed only for the gravity loads in zones of low seismicity, the structure could resist not only the design earthquake, which it would be supposed to resist if it were to be designed against earthquake, but also the higher levels of the earthquake excitations. The main components of its resistance to the high level of earthquakes appear to be (1) the high overstrength, (2) the elongation of the fundamental period, (3) the minor energy dissipation by inelastic deformations, and (4) the increase of the damping ratio. The design base shear derived from the linear elastic base shear of the structure divided by the response modication factor, R 3.5, seems to be completely ctitious or misleading because the high overstrength factor, 8.7, implicit in the structure due to the pre-existing overstrength of the materials and section properties and the reduction in the dead and live loads with regards to the reactive weight caused the model structure behave entirely linear elastically under the design earthquake. Therefore, as far as this study alone is concerned, it is more reasonable that the concept of the reduction of the design base shear considering the energy dissipation by the inelastic response under the design earthquake be waived for the low-rise building structures in the low-seismicity regions. However, considering the possibility of unexpected large earthquakes, the structures in lowto-moderate seismicity regions should retain the ductility to some extent, which can be achieved through the implementation of the requirements on the detailing of reinforcement and structural layout of important lateral-load-resisting elements. 2. The nonlinear static and time history analyses become more widely used tools nowadays than in the past due to the development of more advanced design [25] and evaluation [26] procedures utilizing these techniques. Here, the reliable prediction of the demand and supply in the global and local responses of the structure must be a crucial factor for the successful design and evaluation. However, the available experimental data to calibrate the existing nonlinear analysis softwares

1.0D 1.0E / as shown in Fig. 23. This value, Capacity Capacity s, appears to be 5.06 and this is similar to the ratio Cs / Cw derived from pushover analysis, s (24.33 kN) / (4.61 kN) 5.28, as shown Fig. 19. Secondly, y Cy / Cs can be calculated in case of the experiment (with strain aging) Cy / Cs (51.35 kN) / (24.33 kN) 2.11 and, in case of the analy-

Fig. 23. Effective earthquake load factor for rst signicant yield at critical member ends.

H.-S. Lee, S.-W. Woo / Engineering Structures 24 (2002) 719734

733

seem to be rather scarce and lack to ensure their reliability. The correlation study between analysis and experiment performed herein in this context reveals that special care should be taken, when the IDARC2D is used, in the interpretation of the analytical results since the time history analyses sometimes underestimate signicantly the global responses and both time history and static analyses generally tend to underestimate the rotations in some ends of beams though the static pushover analyses give the reliable global responses. Efforts to improve the reliability of the software should be made from the side of developing more sophisticated analytical models as well as from the side of calibrating these developed models through the correlation study whenever more experimental data become available. 3. The new points in this study, which have not been addressed in the previous studies on the nonseismic RC frames [25], are as follows: The detailing practice in the model structure as shown in Fig. 1(h) is different from that of ACI 318 for the nonseismic regions. The exterior joints in the model structure revealed no signicant bond slippage and premature strength degradation due to the failure of anchorage, which were generally observed in the American researches [3,5]. The global ductility ratio and the overstrength factor of this model structure turn out to be 2.4 and 8.7, respectively, though the previous researches [24] did not clarify specically the capacities of nonductile RC building structures such as the ultimate strength and deformability of nonductile RC frames through the experiment. The previous correlation studies generally concentrated on the global responses only [24]. However, the correlation study herein covers the local responses as well as the global responses. In general, the local responses predicted by using IDARC-2D appear to be not so reliable as the global responses.

authors express their deepest gratitude to all of these supports.

References
[1] Jo BG, Kim SK et al. Evaluation of intensity of 13 December 1996 Yeongweol earthquake and attenuation properties of Korean Peninsula. In: Proceedings of the EESK ConferenceSpring 1997. Seoul: Earthquake Engineering Society of Korea; 1997. p. 216 (in Korean). [2] Bracci JM, Reinhorn AM, Mander JB. Seismic resistance of reinforced concrete frame structures designed only for gravity loads: part 1design and properties of a one-third scale model structure. Technical report NCEER-92-0027, State University of New York, Buffalo, 1992. [3] Bracci JM, Reinhorn AM, Mander JB. Seismic resistance of reinforced concrete frame structures designed for gravity loads: performance of structural system. ACI Structural Journal 1995;92(5):597609. [4] El-Attar AG, White RN, Gergely P. Behavior of gravity load designed reinforced concrete buildings subjected to earthquake. ACI Structural Journal 1997;94(2):13345. [5] Aycardi LE, Mander JB, Reinhorn AM. Seismic resistance of reinforced concrete frame structures designed only for gravity loads: experimental performance of subassemblage. ACI Structural Journal 1994;91(5):55263. [6] Building code for structural concrete (ACI 318-95) and commentary (ACI 318R-95). Detroit (MI): American Concrete Institute, 1995. [7] Lee HS, Woo SW et al. Manufacturing technique and material properties for 1/5-scale reinforced concrete frame model. In: Proceedings of Autumn Conference. Korea Concrete Institute; 1997. p. 57580 (in Korean). [8] Lee HS, Woo SW. An experimental study on the similitude in structural behaviors for small-scale modeling of reinforced concrete structures. In: Proceedings of the 6th US National Conference on Earthquake Engineering. Oakland, CA: EERI; 1998. [9] Valles RE, Reinhorn AM, Kunnath SK, Li C, Madan A. IDARC 2D version 4.0: a program for the inelastic damage analysis of buildings. Technical report NCEER-96-0010, State University of New York, Buffalo, 1996. [10] Harris HG, Sabnis GM. Structural modeling and experimental techniques. Boca Raton, FL: CRC Press, 1999. [11] Hidalgo P, Clough RW. Earthquake simulator study of a reinforced concrete frame. Report no. UCB/EERC-74/13, Earthquake Engineering Research Center, University of California, 1974. [12] Regulations concerning the structure standards of building, etc. Republic of Korea: Ministry of Construction and Transportation, 1996. [13] UBC, Uniform building code 1994. International Conference of Building Ofcials, 1994. [14] Felber, AJ. RESPONSE: a program to determine the loaddeformation response of reinforced concrete sections. Department of Civil Engineering, University of Toronto, 1990. [15] Park YJ, Reinhorn AM, Kunnath SK. IDARC: inelastic damage analysis of reinforced concrete frameshearwall structures. Technical report NCEER-87-0008, State University of New York at Buffalo, 1987. [16] Booth E, editor. Concrete structures in earthquake regions: design and analysis. New York: John Wiley and Sons; 1994. [17] Erasmus LA. Cold straightening of partially embedded reinforcing barsa different view. Concete International: Design and Construction 1981;3(6):4752.

Acknowledgements The research stated herein was supported by the Ministry of Construction and Transportation, the Republic of Korea, and several private companies including SSangYong Engineering and Construction Corp., DongBu Corp., Hyundai Construction Corp., and DongYang Structural Safety Consultants Corp. The contributions of graduate students at Korea University, Dong-Woo Ko, Yun-Sup Heo, and Kyi-Yong Kang, were crucial to the success of this research. The technique which is essential to the manufacture of the reduced-scale model was developed by the support of advanced STructure RESearch Station (STRESS) at Hanyang University. The

734

H.-S. Lee, S.-W. Woo / Engineering Structures 24 (2002) 719734

[18] Bertero VV. Ductility based structural design, state-of-the-art report. In: Proceedings of 9th World Conference on Earthquake Engineering, Tokyo, Kyoto, vol. VIII, 1989, pp. 67386. [19] Uang CM. Establishing R (or Rw) and Cd factors for building seismic provisions. Journal of Structural Engineering ASCE 1991;117(1):1928. [20] Bertero VV, Anderson JC, Krawinkler H, Miranda H. Design guidelines for ductility and drift limits: review of the state-ofthe-practice and state-of-the-art in ductility and drift-based earthquake-resistant design of buildings. Report no. UCB/EERC91/15, Earthquake Engineering Research Center, University of California, Berkeley, CA, 1991. [21] Shahrooz BM, Moehle JP. Evaluation of seismic performance of reinforced concrete frames. Journal of Structural Engineering ASCE 1990;116(5):140322. [22] Miranda E, Bertero VV. The Mexico earthquake of September

[23]

[24]

[25]

[26]

19, 1985performance of low-rise buildings in Mexico City. Earthquake Spectra 1989;5(3):57190. Meli R. Code-prescribed seismic actions and performance of buildings. In: Proceedings of the 10th World Conference on Earthquake Engineering, X. Rotterdam: A.A. Balkema; 1992. p. 57838. Jain SK, Navin R. Seismic overstrength in reinforced concrete frames. Journal of Structural Engineering ASCE 1995;121(3):5805. Kappos AJ, Manafpour A. Seismic design of R/C buildings with the aid of advanced analytical techniques. Engineering Structures 2001;23(4):31932. FEMA (Federal Emergency Management Agency, US). NEHRP recommended provisions for seismic regulations for new buildings and other structures (1997 ed.): part 1provisions; part 2 commentary. Washington, DC, BSSC, 1998.

S-ar putea să vă placă și