Next Article in Journal
A Review of Hygienization Methods of Herbal Raw Materials
Next Article in Special Issue
Time Evolution of CO2 Diffusivity of Carbonated Concrete
Previous Article in Journal
The Impact of Data Filtration on the Accuracy of Multiple Time-Domain Forecasting for Photovoltaic Power Plants Generation
Previous Article in Special Issue
Evaluation of the Structural Performance of a Novel Methodology for Connecting Modular Units Using Straight and Cross-Shaped Connector Plates in Modular Buildings
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Evaluation of Structural Behavior of Hysteretic Steel Dampers under Cyclic Loading

1
Head of R&D Center, HUB Structural Engineering Co., Ltd., Seoul 06302, Korea
2
Department of Architectural Engineering & Urban System Engineering, Kongju National University, Cheonan 31080, Korea
*
Author to whom correspondence should be addressed.
Appl. Sci. 2020, 10(22), 8264; https://doi.org/10.3390/app10228264
Submission received: 27 October 2020 / Revised: 18 November 2020 / Accepted: 19 November 2020 / Published: 21 November 2020

Abstract

:
This study proposes a relatively simple steel damper with high energy dissipation capacity. Three types of steel dampers were evaluated for structural performance. The first damper with U-shape had two vertical members and a semicircular connecting member for energy dissipation. The second damper with an angled U-shape replaced the connecting member with a horizontal steel member. The last damper with D-shape had a horizontal member added to the U-shaped damper. All the dampers were designed with steel plates on both sides that transmitted external shear force to the energy-dissipating members. To evaluate the structural performance of the dampers, an in-plane cyclic shear force was applied to the specimens. The D-shaped damper showed ductile behavior with excellent energy dissipation capacity after yielding without decreasing in strength during cyclic load. In other words, the D-shaped specimen showed excellent performance, with about 3.5 times the strength of the U-shaped specimen and about 3.8 times the energy dissipation capacity due to the additional horizontal member. Furthermore, the efficient energy dissipation of the proposed D-shaped steel damper was confirmed from the finite element (FE) analytical and experimental results.

1. Introduction

One of most important purposes of building structures is to protect humans from nature, and design methods that reduce structural damage caused by external loads are constantly being developed. There are even methods being developed to mitigate damage caused by natural phenomena that are difficult to predict, such as earthquakes.
Seismic design includes seismic resistance, seismic isolation, and seismic damping to design structures that are safe from earthquakes. Seismic resistance allows a structure to directly absorb seismic energy, and it can protect human life during a strong earthquake, but severe residual deformation due to the inelastic behavior of the structure results in enormous social costs and environmental pollution due to demolition. Seismic isolation is the best way to avoid earthquake damage by separating a structure from the ground, thereby rendering it unaffected by earthquakes. However, seismic isolation is difficult to apply to many buildings because the cost of the seismic isolator can be prohibitive. Finally, seismic damping allows the damping device, instead of the structure, to absorb the seismic energy. Damping devices can be divided into active and passive control systems. The active control system is relatively expensive and complex because it requires several mechanical devices. On the other hand, the passive control system is relatively simple and economical, and it can easily be replaced when retrofitting structures after an earthquake. For these reasons, passive control systems are widely used in building structures.
Recently, many studies [1,2,3,4,5,6,7] have been conducted to induce dissipation of seismic energy by using the hysteretic characteristics of ductile steel. Applying metallic yielding dampers for the seismic retrofit to existing structures is becoming more common due to the convenience of manufacturing and installing them, along with their economic efficiency. The passive yielding damper system is generally composed of two types of components. One component is a rigid body that transmits energy induced by seismic load, and the other component dissipates the transmitted external energy. Since the performance of the damping system depends on the characteristics of the energy dissipation component, most research has been focused on how to develop this damper component.
A study to dissipate earthquake-induced energy using the yielding of steel was first initiated by Kelley et al. [8] and Skinner et al. [9] in New Zealand. The steel dampers were developed in the form of ADAS (added damping and stiffness) dampers [10,11], TADAS (triangular added damping and stiffness) dampers [12], buckling restrained braces [13], honeycomb [14], shear panels [15,16,17], slit [18,19,20,21], and U-shaped dampers [22,23]. These dampers can be applied to various structural members, such as walls, columns, beams, and braces of buildings, to absorb seismic energy. Although there have been many studies so far, efforts to develop better dampers that combine a simple form with excellent performance are still ongoing. In this study, a steel damper with a relatively simple shape and excellent energy dissipation and damping capacity was developed, and its structural performance was experimentally and analytically evaluated.

2. Experimental Study

2.1. Design of Specimens

In this study, relatively simple steel dampers that have excellent energy dissipation capacity were planned. As a result of the tension test, the yield strength of the steel used for the dampers was 245 MPa, the tensile strength was 420 MPa, and the strain at yielding was 0.00147. As shown in Table 1, according to the shape of the damper, two types of specimens were designed: open and closed. Each specimen had four identical steel dampers to avoid eccentric deformation.
As shown in Figure 1, all the steel dampers were divided into energy dissipation members and shear plates that transferred external shear force. The size of each shear plate was 75 mm × 110 mm and the thickness was 20 mm. The shear plates were located on each side, and the distance between them was 90 mm. The steel dampers were designed in opened U- and angled U-shapes and a closed D-shape according to the shape of the energy dissipation parts. The thickness of all dampers was 20 mm, and the width of the member for energy dissipation was planned to be 15 mm. The spacing between the shear plates, the thickness of the dampers, and the width and height of the energy dissipation members were designed in consideration of the strength and deformation capacities of the specimens. The corners of the steel dampers were rounded to prevent stress concentration.
As shown in Figure 1, the U-shaped energy dissipation part of the SD1 damper specimen had a height of 90 mm. The energy dissipation part had two vertical members, each with a height of 30 mm, and a semicircular member with an inner diameter of 90 mm. SD2 consisted of horizontal and vertical steel members in order to observe the change in structural behavior when the connecting member was changed from semicircular to horizontal. The vertical members were planned to be 45 mm high, and other details of the specimen were the same as for SD1. The SD3 damper had an additional horizontal member added to the lower part of SD1 to increase the load resisting capacity. The lower horizontal member had a cross-sectional area of 15 mm × 20 mm. The dimensions of the vertical and connecting members of SD3 were the same as those of SD1.

2.2. Test Setup and Loading Procedure

As shown in Figure 2, four identical dampers were installed on each specimen to prevent eccentric deformation of the damper device and to distribute the force acting on the bolt fastened to the shear plates. The dampers were connected to the tops and bottoms of the fronts and rears of T-shaped loading plates, which were connected to L-shaped steel loading frames. As shown in Figure 2, two L-shaped loading frames were used in this experiment: one fastened to the bottom steel frame and the other to a loading actuator of a universal testing machine (UTM) with a capacity of 2000 kN. When a load was applied to the top of the L-shaped loading frame, the loading frame moved vertically, and the shear force was transmitted to the dampers via loading plates.
To maintain a constant distance between the two loading frames, each roller with a capacity of 300 kN was installed on the upper and lower portions of the L-shaped steel loading frames. As shown in Figure 3, the repeated load with the target displacement of ±15mm was applied to the specimens at a speed of 0.06 mm/s to evaluate whether there were changes in the strength, stiffness, and energy dissipation capacity of the specimens during the test.
To measure the relative displacement of the specimens, each LVDT (linear variable differential transducer) was installed on the front and rear surfaces of the loading plates. To check the yielding and deformation characteristics of the specimens, strain gauges were attached to the ends of vertical and horizontal members where the maximum moment occurred. In addition, a strain gauge was attached to the top of each semicircular member to evaluate the deformation characteristics of the connecting members of SD1 and SD3.

3. Experimental Results

3.1. Shear Force Versus Displacement Relationships

The shear force versus displacement relationships of the specimens are shown in Figure 4, and the experimental results for the first yielding and maximum load of the specimens are shown in Table 2. As shown in Figure 4, the stiffness of all specimens decreased significantly after first yielding, and gradually increased until the target displacement. For all specimens during the repeated load cycles, the peak load slightly increased or almost remained steady, and no decrease in structural performance of the specimens was found. The experiment of Specimen SD1 was terminated after performing only 2.5 repetitions due to an error with the test equipment, but it showed sufficient behavioral characteristics to allow for a comparison with the hysteresis behavior of the Specimens SD2 and SD3.
As shown in Figure 4, SD1 with U-shaped damper and SD2 with the upper horizontal connecting member showed similar loading history curves. On the other hand, as shown in Table 2 and Table 3, SD2 had a similar load at first yielding as SD1 but showed some differences in maximum load and elastic and plastic stiffness. This is due to the height of the vertical members and the shape of the connecting member. However, the increase in load-resisting capacity of SD2 was not higher than that of SD3. In other words, SD3 with additional horizontal connecting member between shear plates showed a greater load-resisting capacity than SD1 and SD2 did.
As shown in Table 2, with the addition of the lower horizontal connecting member, the yield strength of SD3 was about 2.0 times higher than that of SD1. In addition, the maximum load of SD3 was up to 3.9 times higher in the first cycle and 3.5 times higher in the second cycle than that of SD1. Furthermore, as shown in Figure 4, the overall behavior of SD3 was similar to that of SD1, which means SD3 was ductile despite its high load-resisting capacity. Both SD2 and SD3 showed no decrease in strength at the target displacement during five repetitions.
Table 3 indicates the experimental results for the elastic and plastic stiffness of the specimens. SD2 was 1.29 times and 1.97 times greater than SD1 in the elastic and plastic moduli, respectively, during the first positive loading phase. On the other hand, after showing the Bauschinger effect, the plastic modulus of SD2 was 1.13 times that of SD1, meaning there was no significant difference. SD3 showed a 5.9 times greater elastic modulus compared to U-shaped SD1. A high elastic modulus means the damper reacts more quickly to an external seismic load. In particular, as shown in Table 2, closed-type SD3 had a faster yielding point and higher initial rigidity than open-type SD1 and SD2, which indicates excellent energy dissipation capability. The plastic modulus of SD3 was 9.1 times and 2.4 times greater than that of SD1. In this experiment, no significant decrease in strength, stiffness, and energy dissipation capacity was observed in SD3, which is similar to the experimental results of SD1 and SD2.

3.2. Deformation of Specimens

None of the specimens fractured or buckled until the end of the experiment. Figure 5 shows the deformation of each specimen at the target displacement. As shown in the figure, when the target displacement was reached, the deformation of each specimen was very stable. In all the specimens, the upper connecting member served to induce deformation so the vertical members could dissipate energy as much as possible.
Since both ends of the horizontal member of SD3 were fixed to both shear plates, it received an anti-symmetric moment as vertical displacement occurred. Adding the horizontal connecting member to the lower part of the shear plates contributed to the increase in load resisting capacity while maintaining ductile behavior characteristics. As shown in Figure 5, the semicircular connecting members of SD1 and SD3 helped to make a relatively small lateral deformation at the maximum displacement, which is due to the geometric shape of the circle.

3.3. Strain Distribution

The strains of the specimens as measured by the strain gauges are shown in Figure 6. SD1 showed a maximum strain of about 0.023 after the vertical member reached its first yield strain at 17.9 kN in the forward direction for first cycle. On the other hand, no yielding was observed in the semicircular connecting member. As shown in Figure 6, energy dissipation of SD1 was concentrated in the vertical member.
SD2 reached yield strain at the positive loading phase of the first cycle in both the vertical and horizontal members. The vertical member first reached yield strain at 18.3 kN, and then the horizontal member yielded at 24.5 kN. Like SD1, the energy dissipation of SD2 was concentrated on the vertical member. In the case of SD2, the strain of the vertical member increased up to about 0.046, and the horizontal connecting member was deformed up to about 0.0028.
Similar to SD1, SD3 did not yield in the semicircular connecting member. On the other hand, both the vertical and the lower horizontal members reached yield strain in first cycle. The first yielding was observed at 36.2 kN in the lower horizontal member, and then at 55.5 kN in the vertical member. The vertical member of SD3 showed a maximum strain of about 0.027, showing similar strain characteristics to SD1. It can be seen from Figure 6 that the lower horizontal member, which is unique to SD3, exhibited a maximum strain of about 0.066, and both the vertical and the lower horizontal members greatly contributed to energy dissipation. It is noted that the maximum strain range of the damper can be controlled by planning the geometric dimensions of the energy dissipation members of the damper.

4. Discussion

4.1. Energy Dissipation

Figure 7 shows the energy dissipation of the specimens for each cycle. In this study, the energy dissipation was the inner area of the load–displacement relationship curve. In the first cycle, since the elastic and plastic phases existed together, the value was slightly lower than those of the other cycles. The cumulative energy dissipation of Specimens SD2 and SD3 showed almost proportional increase, as shown in Figure 7a.
SD1 showed an energy dissipation capacity of 1154.5 kN∙mm in the first cycle, as shown in Figure 7b. Afterward, the energy dissipation area increased by 11.4% to 1286.3 kN∙mm in the second cycle, and the total accumulated energy dissipation capacity was 2440.8 kN∙mm. As shown in Figure 7b, SD2 had an energy dissipation capacity of 1386.6 kN∙mm, 20.1% higher than that of SD1, in the first cycle. In the second cycle, the energy dissipation capacity of SD2 increased by 10.5% from the first cycle, a similar result to that obtained from SD1. The final fifth cycle ended with a 12.5% increase from the first cycle. The experimental results confirmed that the energy dissipation capacity of SD2 remained almost unchanged after the second cycle.
SD3 had an energy dissipation capacity of 4417.4 kN∙mm in the first cycle. This was 3.8 times higher than that of SD1, which meant that even though only the lower horizontal member was added, the energy dissipation capacity was greatly improved. SD3 showed 12.8% and 14.9% higher energy dissipation capacities from the first cycle in the second and fifth cycles, respectively, as shown in Figure 7b. In other words, there was no decrease in energy dissipation capacity during five cycles, such as SD2. From the experimental results, it was confirmed that SD3 has excellent energy dissipation capacities.

4.2. Equivalent Damping Ratio

The damping capacity of the specimens can be determined from the energy dissipated by damping, E D , which is the area enclosed by the hysteresis loop, and the maximum strain energy corresponding to the shaded area, E S o , as shown in Figure 8 [24]. This study derived the equivalent damping ratio ξ as follows:
ξ = E D 4 π E S o
Figure 9 shows the equivalent damping ratios of the specimens for each cycle. As shown in the figure, both open and closed specimens showed high equivalent damping ratios. That is, the equivalent damping ratios of from 48.7% to 50.4% for SD1 and SD2 and from 45.7% to 49.1% for SD3 were experimentally evaluated. SD1 showed an equivalent damping ratio of 49.1% in the first cycle. The second cycle showed an equivalent damping ratio of 45.9%, which was 3.2%p lower than the ratio obtained in the first cycle. For SD2, the equivalent damping ratios of the first and second cycles were 48.2% and 46.5%, respectively, showing a difference of 1.7%p, a similar result to that obtained from SD1. SD2 showed an equivalent damping ratio of 45.7% in the fifth cycle, which was 2.5%p lower than the ratio obtained in the first cycle. SD3 showed an equivalent damping ratio of 48.7% in the first cycle and then showed a higher equivalent damping ratio of about 1.5%p in the second cycle, which it maintained until the final cycle. From the experimental results, the excellent damping capacity of the proposed damper SD3 was confirmed.

5. FE Analysis

5.1. FE Modeling

This study performed a FE analysis to analytically evaluate the stress distributions and energy dissipation zones of the dampers used in the specimens and to verify the accuracy of the experiment. The FE analysis was performed with the 3D nonlinear FE analysis program MIDAS-NFX [25]. This program can take into account both material and geometric nonlinearities. The kinematic hardening model considering the Bauschinger effect was applied as the hysteresis model of steel subjected to cyclic loading. The elastoplastic tri-linear curve of the steel with strain hardening was used for stress versus strain relationship of the material. Geometric nonlinearity was applied to consider the large deformation of the element.
Figure 10 shows the FE modeling of each steel damper of the specimens. A 3D solid hexahedron element was used for the FE modeling, and the size of the element was about 4 mm. To induce vertical displacement in each specimen, the bolted region of one shear plate was fixed and the other was modeled with a roller.

5.2. Comparison of Experimental and Analytical Results

Figure 11 shows the stress distribution of each damper by FE analysis when the target displacement was reached. In the figure, the red color refers to the plastic region after yielding, and the stress decreases as the color turns blue. From Figure 11 it can be seen that the region reaching yield stress in each damper was almost the same as it was in the experimental result in Section 3.3. In the case of SD1, the yield and energy dissipation zones were concentrated in the vertical members. The analysis confirmed that some yielding occurred in the semicircular connecting member adjacent to the vertical member, which was similar to the experimental results obtained from the horizontal member adjacent to the vertical member in SD2.
As shown in Figure 11c, in SD3 with an additional horizontal member, the yield region was concentrated on the vertical and lower horizontal members, the same as it was in the experimental results. In particular, most of the area of the lower horizontal member reached yield stress. In addition, by adding the lower horizontal member, a larger area of the damper engaged in stress transmission and energy dissipation than did the areas of SD1 and SD2 damper.

6. Conclusions

This study experimentally and analytically evaluated the structural performance of the proposed D-shaped steel damper subjected to cyclic loading, and the following results were obtained.
(1)
As a result of the experiment, SD1 and SD2 in which the vertical members were connected by semicircular and horizontal members, respectively, had similar load history characteristics, energy dissipation capacities, and equivalent damping ratios. Therefore, even if the shape of a connecting member is changed from the semicircular type to the relatively simple horizontal type, there will be no significant degradation in structural performance.
(2)
SD3, which added the horizontal member to U-shaped SD1, showed about 3.5 times more load-resisting capacity and about 3.8 times greater energy dissipation capacity than did SD1 in the experiment. In addition, SD3 had higher elastic stiffness and lower displacement at first yielding compared to those of SD1.
(3)
From the experiment, both SD2 and SD3 showed no decrease in stiffness and strength during five loading cycles, and they showed equivalent damping ratios of about 50%, confirming excellent damping capacity. In addition, SD2 and SD3 almost maintained their original shapes, even after five reversed cyclic loads, and no in-plane or out-of-plane buckling was found.
(4)
The FE analysis results for the energy dissipation zone of each damper were almost consistent with the experimental results. In the case of SD3, the yielding zone was concentrated in the vertical member and the lower horizontal member. Furthermore, the semicircular connecting member of SD3 showed similar stress distribution as SD1. This means the additional horizontal member is an important factor in determining the behavioral characteristics of the proposed D-shaped steel damper.

Author Contributions

S.-W.K. and K.-H.K. conceived and performed the experiments and analyzed the data and wrote the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Priority Research Centers Program through the National Research Foundation of Korea (NRF) founded by the Ministry of Education (No. 2019R1A6A1A03032988); This research was supported by Basic Science Research Program through the National Research Foundation of Korea (NRF) founded by the Ministry of Education (No. 2018R1A2B3001656).

Acknowledgments

The authors would like to give sincere thanks to Hyung-Man Kim who is the president of HUB Structural Engineering, Co., Ltd. for supporting this work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Benavent-Climent, A.; Morillas, L.; Vico, J.M. A study on using wide-flange section web under out-of-plane flexural for passive energy dissipation. Earthq. Eng. Struct. Dyn. 2010, 40, 473–490. [Google Scholar] [CrossRef]
  2. Teruna, D.R.; Majid, T.A.; Budiono, B. The use of steel damper for enhancing the seismic performance of R/C frame with soft first story. J. Civ. Eng. Res. 2014, 4, 191–202. [Google Scholar]
  3. Teruna, D.R.; Majid, T.A.; Budiono, B. Experimental study of hysteretic steel damper for energy dissipation capacity. Adv. Civ. Eng. 2015, 2015, 631726. [Google Scholar] [CrossRef] [Green Version]
  4. Shen, X.; Wang, X.; Ye, Q.; Ye, A. Seismic performance of transverse steel damper seismic system for long span bridges. Eng. Struct. 2017, 141, 14–28. [Google Scholar] [CrossRef]
  5. Abebe, D.Y.; Kim, J.W.; Gwak, G.; Choi, J.H. Low-cycled hysteresis characteristics of circular hollow steel damper subjected to Inelastic behavior. Int. J. Steel Struct. 2019, 19, 157–167. [Google Scholar] [CrossRef] [Green Version]
  6. Naeem, A.; Kim, J. Seismic performance evaluation of a multi-slit damper. Eng. Struct. 2019, 189, 332–346. [Google Scholar] [CrossRef]
  7. Qiu, C.; Wang, H.; Liu, J.; Qi, J.; Wang, Y. Experimental tests and finite element simulations of a new SMA-steel damper. Smart Mater. Struct. 2020, 29, 035016. [Google Scholar] [CrossRef]
  8. Kelly, J.M.; Skinner, R.I.; Heine, A.J. Mechanisms of energy absorption in special devices for use in earthquake resistant structures. Bull. New Zealand Soc. Earthq. Eng. 1972, 5, 63–88. [Google Scholar]
  9. Skinner, R.I.; Kelly, J.M.; Heine, A.J. Hysteresis dampers for earthquake resistant structures. Earthq. Eng. Struct. Dyn. 1974, 3, 287–296. [Google Scholar] [CrossRef]
  10. Whittaker, A.S.; Bertero, V.V.; Thompson, C.L.; Alonso, L.J. Seismic testing of steel plate energy dissipation devices. Earthq. Spectra 1991, 7, 563–604. [Google Scholar] [CrossRef]
  11. Shih, M.H.; Sung, W.P. A model for hysteretic behavior of rhombic low yield strength steel added damping and stiffness. Comput. Struct. 2005, 83, 895–908. [Google Scholar] [CrossRef]
  12. Tsai, K.C.; Chen, H.W.; Hong, C.P.; Su, Y.F. Design of steel triangular plate energy absorbers for seismic-resistant construction. Earthq. Spectra 1993, 9, 505–528. [Google Scholar] [CrossRef]
  13. Deng, K.; Pan, P.; Li, W.; Xue, Y. Development of a buckling restrained shear panel damper. J. Constr. Steel Res. 2015, 106, 311–321. [Google Scholar] [CrossRef]
  14. Lee, M.; Lee, J.; Kim, J. Seismic retrofit of structures using steel honeycomb dampers. Int. J. Steel Struct. 2017, 17, 215–229. [Google Scholar] [CrossRef]
  15. Nakashima, M. Strain-hardening behavior of shear panels made of low-yield steel. I: Test. J. Struct. Eng. 1995, 121, 1742–1749. [Google Scholar] [CrossRef]
  16. Foti, D.; Diaferio, M.; Nobile, R. Optimal design of a new seismic passive protection device made in aluminium and steel. Struct. Eng. Mech. 2010, 35, 119–122. [Google Scholar] [CrossRef]
  17. Abebe, D.Y.; Jeong, S.J.; Getahune, B.M.; Segu, D.Z.; Choi, J.H. Study on hysteretic characteristics of shear panel damper made of low yield point steel. Mater. Res. Innov. 2014, 19, S5-902–S5-910. [Google Scholar]
  18. Chan, R.W.K.; Albermani, F. Experimental study of steel slit damper for passive energy dissipation. Eng. Struct. 2008, 30, 1058–1066. [Google Scholar] [CrossRef]
  19. Oh, S.H.; Kim, Y.J.; Ryu, H.S. Seismic performance of steel structures with slit dampers. Eng. Struct. 2009, 31, 1997–2008. [Google Scholar] [CrossRef]
  20. Kim, J.; Kim, M.; Eldin, M.N. Optimal distribution of steel plate slit dampers for seismic retrofit of structures. Steel Compos. Struct. 2017, 25, 473–484. [Google Scholar]
  21. Lee, J.; Kim, J. Development of box-shaped steel slit dampers for seismic retrofit of building structures. Eng. Struct. 2017, 150, 934–946. [Google Scholar] [CrossRef]
  22. Bagheri, S.; Barghian, M.; Saieri, F.; Farzinfar, A. U-shaped metallic-yielding damper in building structures, Seismic behavior and comparison with a friction damper. Structures 2015, 3, 163–171. [Google Scholar] [CrossRef]
  23. Jamkhaneh, M.E.; Ebrahimi, A.H.; Amiri, M.S. Experimental and numerical investigation of steel moment resisting frame with U-shaped metallic yielding damper. Int. J. Steel Struct. 2019, 19, 806–818. [Google Scholar] [CrossRef]
  24. Chopra, A.K. Dynamics of Structures: Theory and Applications to Earthquake Engineering, 5th ed.; Pearson Education: Cranbury, NJ, USA, 2017; p. 960. [Google Scholar]
  25. MIDAS-NFX. 2020. Available online: https://kor.midasuser.com/nfx/ (accessed on 19 September 2020).
Figure 1. Details of steel dampers used in specimens (unit: mm): (a) SD1; (b) SD2; (c) SD3.
Figure 1. Details of steel dampers used in specimens (unit: mm): (a) SD1; (b) SD2; (c) SD3.
Applsci 10 08264 g001
Figure 2. Test setup of specimen.
Figure 2. Test setup of specimen.
Applsci 10 08264 g002
Figure 3. Loading history.
Figure 3. Loading history.
Applsci 10 08264 g003
Figure 4. Shear force versus displacement curve of tested specimens: (a) SD1; (b) SD2; (c) SD3.
Figure 4. Shear force versus displacement curve of tested specimens: (a) SD1; (b) SD2; (c) SD3.
Applsci 10 08264 g004
Figure 5. Deformation of steel dampers at target displacement: (a) SD1; (b) SD2; (c) SD3.
Figure 5. Deformation of steel dampers at target displacement: (a) SD1; (b) SD2; (c) SD3.
Applsci 10 08264 g005
Figure 6. Strain distributions of tested specimens: (a) SD1; (b) SD2; (c) SD3.
Figure 6. Strain distributions of tested specimens: (a) SD1; (b) SD2; (c) SD3.
Applsci 10 08264 g006
Figure 7. Energy dissipation of tested specimens: (a) Cumulative energy dissipation; (b) Energy dissipation.
Figure 7. Energy dissipation of tested specimens: (a) Cumulative energy dissipation; (b) Energy dissipation.
Applsci 10 08264 g007
Figure 8. Estimation of equivalent damping ratio.
Figure 8. Estimation of equivalent damping ratio.
Applsci 10 08264 g008
Figure 9. Equivalent damping ratios of specimens for each cycle.
Figure 9. Equivalent damping ratios of specimens for each cycle.
Applsci 10 08264 g009
Figure 10. FE modeling of steel dampers used in this study: (a) SD1; (b) SD2; (c) SD3.
Figure 10. FE modeling of steel dampers used in this study: (a) SD1; (b) SD2; (c) SD3.
Applsci 10 08264 g010
Figure 11. Stress distribution of steel dampers obtained from FE analysis: (a) SD1; (b) SD2; (c) SD3.
Figure 11. Stress distribution of steel dampers obtained from FE analysis: (a) SD1; (b) SD2; (c) SD3.
Applsci 10 08264 g011
Table 1. Details of specimens.
Table 1. Details of specimens.
SpecimensSteel Dampers Number of Steel Dampers
TypeShapeEnergy Dissipation Member
Vertical
(mm)
Connecting
(mm)
Bottom
(mm)
Thickness
(mm)
SD1OpenedU-shapeB=15
H=30
do=60
di=45
N/A204EA
SD2OpenedAngled U-shapeB=15
H=45
B=120
H=15
N/A204EA
SD3ClosedD-shapeB=15
H=30
do=60
di=45
B=90
H=15
204EA
SD means steel damper; B and H are the width and height of member, respectively; do and di are the outer and inner diameters, respectively.
Table 2. Experimental results of tested specimens.
Table 2. Experimental results of tested specimens.
SpecimensFirst Yielding Peak Load of Each Cycles at Target Displacement (kN)
CycleLoad (kN)Disp. (mm)Dir.1st 2nd 3rd 4th5th
SD1+1st cycle+17.9+1.52(+)24.929.731.1--
(−)−29.6−31.9---
SD2+1st cycle+18.2+1.20(+) 30.5 35.0 35.8 36.0 36.2
(−)−34.8−36.7−37.1−37.3 −37.5
SD3+1st cycle+36.2+0.52(+) 96.0105.1107.2107.4107.1
(-)−107.2−113.0−113.9−114.2−114.1
Table 3. Elastic and plastic modulus of tested specimens.
Table 3. Elastic and plastic modulus of tested specimens.
SpecimensElastic ModulusPlastic Modulus
+1st CycleAfter +1st Cycle
Stiffness
(kN/mm)
Increase Ratio
to SD1
Stiffness
(kN/mm)
Increase Ratio
to SD1
Stiffness
(kN/mm)
Increase Ratio
to SD1
SD111.8-0.31-0.30-
SD215.21.290.611.970.341.13
SD369.65.902.829.100.722.40
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kim, S.-W.; Kim, K.-H. Evaluation of Structural Behavior of Hysteretic Steel Dampers under Cyclic Loading. Appl. Sci. 2020, 10, 8264. https://doi.org/10.3390/app10228264

AMA Style

Kim S-W, Kim K-H. Evaluation of Structural Behavior of Hysteretic Steel Dampers under Cyclic Loading. Applied Sciences. 2020; 10(22):8264. https://doi.org/10.3390/app10228264

Chicago/Turabian Style

Kim, Sang-Woo, and Kil-Hee Kim. 2020. "Evaluation of Structural Behavior of Hysteretic Steel Dampers under Cyclic Loading" Applied Sciences 10, no. 22: 8264. https://doi.org/10.3390/app10228264

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop