Вы находитесь на странице: 1из 10

SEISMIC PERFORMANCE OF HYBRID COREWALL BUILDINGS

Bahram M. SHAHROOZ1, Patrick J. FORTNEY2, Gian Andrea RASSATI3 SUMMARY A common structural system involves the use of reinforced concrete core walls, which are typically formed by coupling individual wall piers, and steel perimeter frames. For low-tomoderate rise buildings up to 2530 stories, the core walls are the primary lateral load resisting system, and the perimeter frames are designed for gravity loads. The use of dual systems is more common in taller buildings, in which the perimeter frames are engaged (through outrigger beams or story-deep outrigger trusses) with the walls/cores as a means of reducing lateral drifts. This paper provides an overview of a number of issues related to design of hybrid structures with reinforced concrete central core walls and perimeter steel frames. A number of design options for steel or steel-concrete composite coupling beams and their connections to core walls, and connections between outrigger beams and core walls are discussed. Preliminary results from a new concept for enhancing the performance of coupling beams are also presented. Keywords: composite construction, core walls, coupling beams, hybrid walls, outrigger beams, seismic design. INTRODUCTION An efficient hybrid structural system is obtained when reinforced concrete core walls are used in conjunction with steel perimeter frames. Core walls can effectively be formed by coupling individual wall piers with the use of reinforced concrete, steel, or hybrid coupling beams. The walls may be reinforced conventionally with longitudinal and transverse reinforcing bars, or may include embedded structural steel boundary columns in addition to conventional reinforcing bars. For low-to-moderate rise buildings up to 25 to 30 stories, the core can be used to provide a majority of the lateral force resistance. For taller buildings, the use of dual systems is more common, where the perimeter frames are engaged with the core. Outrigger beams are framed between the core walls and columns (which may be all steel or composite) in the perimeter frame. The successful performance of such hybrid structural systems depends on the adequacy of the primary individual components which are the core walls, steel frames, and frame-core connections. The focus of this paper is to provide an overview of design issues for (a) steel or hybrid coupling beams and their connections to core walls, and (b) connections between outrigger beams and core walls. Each of these two issues is discussed separately. COUPLING BEAMS Coupling beams are analogous to and serve the same structural role as link beams in eccentrically braced frames. Reinforced concrete coupling beams are subject to significant limitations in terms of the shear stress they have to resist, and often require impractically deep sections to carry the loads demanded of them. Steel placement for diagonally reinforced beams is often impractical for beams having a span-to-depth ratio greater than 1.5. Finally, the displacement capacity of reinforced concrete coupling beams has been shown often be exceeded by the demand. Viable alternatives to reinforced concrete coupling beams are steel beams or steel beams encased in
Professor, Department of Civil and Environmental Engineering, University of Cincinnati, 765 Baldwin Hall, Cincinnati, Ohio 45221-0071, U.S.A., e-mail: Bahram.Shahrooz@uc.edu 2 Doctoral Research Assistant, Department of Civil and Environmental Engineering, University of Cincinnati, 765 Baldwin Hall, Cincinnati, Ohio 45221-0071, U.S.A., e-mail:pfortney@cinci.rr.com 3 Visiting Professor, Department of Civil and Environmental Engineering, University of Cincinnati, 765 Baldwin Hall, Cincinnati, Ohio 45221-0071, U.S.A., e-mail: Gian.Rassati@uc.edu
1

79

concrete having varying levels of longitudinal and transverse reinforcement, which are referred to as composite or hybrid coupling beams. The advantages of steel or hybrid coupling beams become particularly apparent in cases where height restrictions do not permit the use of deep reinforced concrete beams, or where the required capacity, stiffness, or deformation capacity cannot be developed economically with a concrete beam. Steel coupling beams have been used in a number of structures (Ferver, et al., 1974, Lehmkuhl, 2002, Paulay, 2002, Taranath, 1998, Yao, 2003), and more cases with steel coupling beams are currently in the design phase. Similar to reinforced concrete coupling beams, energy dissipation characteristics of steel or composite coupling beams play an important role in the overall performance of the structure. Depending on the coupling beam length, steel or hybrid coupling beams may be detailed to dissipate a major portion of the input energy by flexure or by shear. For most coupling beams, however, it is more advantageous to design the coupling beams as shear critical, or shear yielding members since such members exhibit a more desirable mode of energy dissipation. Such a choice is not possible for reinforced concrete members. Coupling Beam Wall Connections The coupling beam-to-wall connection depends on whether the wall boundary element is reinforced conventionally or contains embedded structural steel columns. Two possible connections for the latter case are shown in Fig. 1. If shear connection the wall boundary moment connection element is reinforced with longitudinal and transverse reinforcing bars, a typical connection shear studs as required end plate end plate involves wall steel not shown wall steel not shown embedding the coupling beam embedment provides embedment provides shear resistance into the wall and moment resistance interfacing it with (a) steel coupling beam attached (b) steel coupling beam attached the boundary to steel boundary column to steel erection column element, as shown in Fig. 2. In Figure 1 Two types of hybrid coupling beam connected to embedded steel columns addition to boundary element reinforcing, embedded steel members may also be provided with welded vertical reinforcing bars, attached to the flanges through mechanical half couplers welded to the A flanges, or with shear studs. The vertical bars have been A found to assist in transfer of bearing stresses around the flanges, and to enhance the overall stiffness of the connection (Shahrooz et al., 1993). Adequate control of the B gap that opens at the beam flanges upon load reversal may B also be provided by (a) placing two-thirds of the required vertical boundary element steel within a distance of one half the embedment length from the face of the wall, and (b) ensuring that the width of the boundary element steel not to exceed 2.5 times the coupling beam flange width (Harries et al., 1997). It is not necessary, nor is it practical, to pass boundary element reinforcing through the web of the embedded coupling beam. It is, however, necessary to SECTION A-A provide good confinement in the region of the embedded web. Confinement may be accomplished using hairpins and cross ties parallel to the web as shown in Fig. 2 (section AA). Additionally, vertical boundary element reinforcement in this region may also be relied upon to provide significant confinement to the embedded region.

SECTION B-B
Figure 2 Steel coupling beam embedment details 80

Calculation of Embedment Length The coupling beam is embedded in the wall such that its capacity can be developed. A number of methods (Marcakis and Mitchell, 1980; Mattock and Gaafar, 1982) may be Le - c c L/2 = a used to calculate the necessary embedment length. These methods assume rigid body rotation of the embedded section and calculate spalled cover the internal moment arm between bearing forces generated at each concrete Cb end of the embedment as shown in Fig. 3. Proposed methods Vu differ simply by the assumed stress distribution in the embedment concrete. For steel coupling beams, Vu is taken as the plastic shear capacity of the steel member, i.e., V p = 0.6 F y (h 2t f )t w where

Fy = yield strength of web steel, h = beam depth, tf = flange thickness, and tw = web thickness. To account for strain hardening and material over strength, it is recommended that Fy be taken as 1.5 times the nominal yield strength. For hybrid coupling beams, shear capacity may be taken as Vn = 1.56 (Vsteel + V RC ) in which and V steel = 0.6 F y t w (h-2t f )
V RC = 0.166 f' c bw d + Av f y d

Cf b xb Cb Le xf

C L
f = 0.003
assumed strain distrib ution assumed stress distrib ution Mattock & Gaafar

where bw = web width of the 0.85fc' s encasing element around steel coupling beam, d = effective depth 3/4(Le-c) of the encasing element, Av = total area of stirrups in the encasing assumed 1xf stress element, fy = yield strength of the stirrups, and s = spacing of the fc' xb 0.85fc' distrib ution stirrups. This method has been calibrated based on a relatively Marcakis & large number of case studies (Gong and Shahrooz, 2001). The Mitchell nominal values of Fy and fc (in MPa) are to be used because the Figure 3 Methods for computing equation has been calibrated to account for strain hardening and embedment length material over strength. Additional transfer bars attached to the beam flanges (as discussed above) can contribute to the capacity of the embedment. The required embedment length can be modified to account for the additional strength. However, to ensure that the calculated embedment length is sufficiently large to avoid excessive inelastic damage in the connection region, it is recommended that the contribution of transfer bars be neglected in the embedment capacity design.
Design and Detailing of Coupling Beams

1/3(Le-c)

Well established guidelines for shear links in eccentrically-braced frames may be used to design and detail steel coupling beams (AISC, 2002). The expected coupling beam rotation angle plays an important role in the required beam details such as the provision of stiffeners. This angle is computed with reference to the collapse mechanism shown in Fig. 4 which corresponds to the expected behavior of coupled wall systems. The value of p is taken as Cd e, where Cd is the deflection amplification factor (ICBO, 2000), and e is the elastic inter-story drift angle computed under code level Plastic lateral loads. Knowing the value of p, Hinges the shear angle, p, is calculated as indicated in Fig. 4.
p

Previous research on hybrid coupling beams (Gong and Shahrooz, 2001) indicates that nominal encasement around steel beams provides adequate resistance against flange and web buckling. Therefore, hybrid coupling beams may be detailed without web stiffeners. Due to inadequate data regarding the influence of encasement on local buckling, minimum flange and web slenderness requirements similar to steel coupling beams should still be used.

p p = p
L Lwall
Figure 4 Determination of coupling beam angle of rotation

Lwall L

81

A pair of stiffener plates (on both sides of the web) placed along the embedment length will mobilize compression struts in the connection region as depicted schematically in Fig. 5. These stiffener plates are commonly referred to as face bearing plates. The first face bearing plate Transfer Bar should be inside the confined core of wall boundary element. The distance between the plates should be such that the angle of the compression struts is approximately 45 degrees (hence, the Coupling Beam distance between the plates should be about equal to the clear distance between the flanges). To ensure adequate contribution of face bearing plates, the width of each face bearing plate should be equal to the flange width on either side of the web. Thickness of the face bearing plates can be established based on available guidelines for detailing of shear links in eccentricallybraced frames (AISC, 2002).

Performance of Steel or Hybrid Coupling Beams

Face Bearing Plates

Recent tests by Gong and Shahrooz (2001) suggest satisfactory performance of coupling beams designed and detailed according to aforementioned design and detailing procedures. The distribution of dissipated energy in Fig. 6 shows that the majority of the input energy is dissipated through inelastic deformations in the beam outside of the connection region. Moreover, the design methodology results in connections that exhibit stable hysteretic response and satisfactory ductility (see Fig. 7).
300

R.C. Wall

Compression Struts

Figure 5 Face bearing plates

Cumulative Dissipated Energy (kN-m)

100 80 60 40 20 0 0 0.025 0.05 0.075 Shear Angle (rad.) 0.1


Input Connection Beam
Shear (kN)

200 100 0 -100 -200 -300 -0.15 -0.10 -0.05 0.00 0.05 0.10 0.15 Shear Angle (rad.)

Figure 6 Distribution of dissipated energy

Figure 7 Hysteretic response of hybrid coupling beams

Despite satisfactory performance in terms of energy dissipation characteristics and hysteretic responses, the connection region around steel or hybrid coupling beams is typically damaged extensively under extreme loading. The repair will be expensive, and will involve expensive Fuse chipping of the wall around the connection and placement of new coupling beams. In an effort to control the level of damage, an ongoing study has been focused on coupling beams with Figure 8 Schematic drawing of fuse for coupling beams fuses where the

82

inelastic deformations are concentrated. The fuses are also detailed to eliminate/limit the damage to the coupling beams and wall piers, and to ensure ease of replacement. A schematic drawing of such a system is shown in Fig. 8. The fuse should be designed to ensure shear yielding. The plastic moment capacity of the fuse is set to be equal or greater than the corresponding value for the coupling beam, and the plastic shear capacity of the fuse is taken as a percentage (less than 100%) of the plastic shear capacity of the coupling beam. A series of preliminary tests were recently conducted to examine this concept. In these tests, steel beams were used to represent wall piers. The fuses performed per design, i.e., the coupling beam remained elastic while the fuse dissipated a significant amount of energy through yielding of the web while undergoing large rotations (Fig. 9).
25 20 15

Rotation (% rad.)

10 5 0 -5 Rotation @ shear yielding

-10 -15 -20 -25 -32,000 -24,000 -16,000 -8,000

Shear strain (micro strain)

8,000

16,000 24,000 32,000

Figure 9 Fuse rotation-shear strain

The concept of coupling beams with fuses is being further evaluated through specimens that incorporate wall piers. Fabrication of these specimens is currently underway, and testing would be completed by early 2004. These tests will provide further data regarding performance of fuses in coupling beams, particularly relationships between damage in the wall piers as a function of plasticity in the fuse will be established.
OUTRIGGER BEAM-WALL CONNECTION

In low-rise buildings, up to about 30 stories, the core is the primary lateral load resisting system, and the perimeter frame is designed for gravity loads, and the connection between outrigger beams and cores is generally a shear connection. A typical shear connection is shown in Fig. 10 in which a steel plate with headed studs is embedded in the core wall during casting. After casting beyond the plate, the web of the steel beam is bolted to the High Strength Bolts shear tab which is already welded to the plate. Variations of this detail are common. Although this connection Steel Shear Tab provides some moment resistance, it is generally accepted that the connection is flexible and does not develop large moments; hence, such connections are considered to be shear Outrigger connections. In taller buildings, story-deep trusses may be Beam used to engage the perimeter columns as a means of reducing the overall deformation of the structure. The connection between the top and bottom chords is essentially similar to that shown in Fig. 10.

R.C. Core Wall

In an effort to develop design procedures for outrigger beamwall connections, a coordinated experimental and analytical Figure 10 Outrigger beam-core wall shear study was recently completed at the University of Cincinnati connection (Shahrooz et al., 2003). The main objectives of this study were to (a) investigate effects of concrete cracking and yielding of wall longitudinal bars on the performance of outrigger beam-wall connections, (b) examine influence of wall boundary elements around the connection region on the connection capacity, (c) evaluate significance of

Embedded Plate w/ Shear Studs

83

floor slab on connection behavior, and (d) develop design guidelines.


Test Specimens and Testing Method

Two -scale specimens were selected. Each specimen had a cantilever wall that represented a portion of a core a consisting of a central I-shaped and two C-shaped walls. #4@95 A rectangular wall, corresponding to the web of the I#3@102 shaped wall of the core, was used in the first specimen. 457 #3 The second specimen had a T-shaped wall, which represented the web and a portion of the flange of the I#2@63.5 c shaped portion of the core wall. Each specimen had two outrigger beam-wall connections (Fig. 11). The first 1295 connection was located at 610 mm above the base, which corresponded to the average length of the expected plastic hinge. The wall around the first connection had boundary 2nd Diaphragm is not Connection element reinforcement. The second connection was shown for clarity. placed at 610 mm above the first connection. At this 610 location, the wall would experience moderate levels of 1st Connection cracking, and the wall did not have boundary element 914 610 reinforcement. The outrigger beam-wall connection was #8@95 similar to that shown in Fig. 10. The connections were designed to resist gravity shear and diaphragm axial load #5@203 457 (Shahrooz et al., 2002). Four 9.5-mm diameter studs were 914 b used for the embedded stud plate, and three 19-mm2743 diameter A490 bolts were used to connect the beam to the embedded stud plate. The studs were 102 mm and 69 mm a = 1143 for rectangular wall; 1270 for T-shaped wall b = 914 for rectangular wall; 1080 for T-shaped wall long in the rectangular wall and T-shaped wall, c = 114 for rectangular wall; 241 for T-shaped wall respectively. The flange in the T-shaped wall increases the concrete pullout capacity of the studs; hence, the Figure 11 Wall reinforcement required stud length for the T-shaped wall is about one half of that in the rectangular wall. Floor diaphragm was included in Seat Anchors the second specimen, refer to Fig. angle on each side The floor metal deck was 12. connected to the top flange of the Concrete outrigger beam by five 6.4 mm Infill diameter x 35 mm long studs spaced at 152 mm on center, with Metal Deck the first stud placed at 95 mm from the face of the wall. The metal deck was tack welded to the top R.C. Core Wall surfaces of two 50.8x50.8x6.35 mm seat angles (one on each side of the outrigger beam) that had been connected to the wall flange Figure 12 Diaphragm-core wall connection by six self-tapping 6.4-mm diameter screws with 38 mm embedment length (3 on each side of the outrigger beam). The connection of the metal deck to the wall was intended to resist only the factored floor gravity loads. The general layout of the test setup is shown in Fig. 13. A constant axial load equal to 0.1fcAg (Ag = wall crosssectional area, fc = concrete compressive strength) was applied to the walls. Up to reaching 2.5% drift, the rectangular wall and outrigger beams were loaded simultaneously. After subjecting the wall to 2.5% drift, loading of the wall was stopped, and each connection was separately subjected to cyclic loading until failure. Due to concerns for the adequacy of the loading system and laboratory schedule, loading of the T-shaped wall was limited to only 0.5% before individual testing of each connection to failure. A constant gravity shear equal to 26.2 kN was applied to the top connection of both specimens. The addition of gravity shear for the bottom

84

connection was determined to have minimal effects on the connection capacity, and hence was not included (Shahrooz et al., 2002).
0.1f c A g

P
Loading Slab

2743 mm

26.2 kN P/2 P/2


Footing

Figure 13 Test setup and loading Test Results (a) Rectangular Wall

At 2.5% drift, the rectangular wall experienced significant cracking and spalling around the bottom connection, and moderate level of cracking around the top connection. The wall longitudinal reinforcing bars were strained at least 10 times the yield strain at the base, and approximately 6 times the yield strain at the location of the bottom connection. The strain gages on the longitudinal bars near the top connection were lost during fabrication of the specimen, but the crack patterns do not suggest significant inelastic action around the top connection. The studs in the bottom and top connections yielded, and were strained to 120% and 187% of the yield strain, respectively, when the wall was pushed to 2.5% drift. The differences in the stud strains in the top and bottom connection are mainly attributed to the presence of boundary element around the bottom connection, which reduced the participation of the studs. The ratio of the average strain in the bottom connection studs to the corresponding value in the top connection was 0.56 at the start of the test, and changed slightly throughout various stages of testing. Events such as yielding of the wall longitudinal bars and studs did not significantly impact this ratio, and the average value of 0.58 remained essentially the same as the original value. The lower values of stud strains in the bottom connection are also apparently due to the level of cracking. The cracks around the bottom connection were considerably more extensive than those around the top connection. Hence, widening of the cracks mostly dissipated the input energy in the bottom connection, whereas the corresponding energy in the top connection was predominately dissipated through yielding of the studs. Despite excessive damage and yielding of the wall longitudinal bars around the bottom connection, at the conclusion of combined wall-connection tests the connection did not fail. The top connection did not fail either, albeit the level of damage and inelasticity around the connection was moderate as expected. When loaded individually, the bottom connection and top connection failed due to stud fracture and stud pullout, respectively. The boundary element around the bottom connection apparently prevented pullout of the studs.
(b) T-Shaped Wall

As mentioned before, combined testing of the wall and connections was stopped after subjecting the wall to a lateral drift of 0.5%. At this drift, the wall longitudinal bars in the flange and web at the base were strained to 4.6 y and 2.3 y, respectively, but the level of cracking in the wall was minimal. The wall longitudinal bars near the bottom and top connections did not yield; the maximum strain in these bars was 0.65 y near the bottom connection. Similar to the rectangular wall specimen, the stud strains in the top connection, which was not surrounded by heavily confined boundary element, were larger than the bottom connection stud strains. At 0.5% drift, the studs in the top connection yielded and reached a strain of twice the yield strain, whereas the bottom connection studs remained in the elastic range (the maximum strain was 0.43y). This difference is attributed to

85

confinement by the boundary element that reduced the participation of the studs in the bottom connection. In contrast to the first specimen, both the top and bottom connections in the specimen with T-shaped wall failed as a result of stud fracture. The participation of the floor diaphragm in transferring forces to the wall was negligible. For example, the strains in the concrete deck were on the average 18% of the strains measured in the outrigger beam, and the slab participation away from the outrigger beam was rather insignificant. The majority of the load was transferred directly to the wall through the outrigger beam-wall connection. Diaphragm-wall connections may, therefore, be designed for gravity loads only unless slab bars are interfaced with walls, or other positive means such as collector anchors are used.
Assessment of Connection Capacity and Mode of Failure

The analytical model shown in Fig. 14 was used to compute capacity and mode of failure of the connections in both specimens. At stud pullout failure, the concrete strain around studs is small, and concrete essentially behaves elastically. Therefore, a linear distribution of concrete stress is assumed in the model. The maximum concrete e kd Vu e in which kd and stress (fc) is computed from f c = It It (transformed moment of inertia) are calculated based on standard techniques for a cracked transformed section analysis of reinforced concrete beams. The studs are Tcap. Tu considered as reinforcing bars, the width of the stud plate is used as the beam width, and for simplicity, the contribution d of the studs in compression is ignored. Knowing kd, the value of C is 0.5bfc kd. In the model, the contribution of kd C friction between the embedded stud plate and concrete is also taken into account. Equation (1) is used to compute Vu the tensile load that can be resisted by the connection. f
c

C + 0.5Tu Tcap.

0.5(Vu C) + =1 4 Vc
2

(1)

In this equation, the stud plate is assumed to behave rigidly, Figure 14 Analytical model and accordingly the gravity shear (Vu) and applied axial load (Tu) are divided equally between the studs. Note that the coefficients of 0.5 and 4 are for the test specimens in which four studs were used. For other cases, these coefficients have to be changed depending on the number of studs (Shahrooz et al., 2002). The coefficient of friction () was taken as 0.4. The design shear capacity of studs, Vc, is computed from the applicable equations in PCI Handbook (PCI, 2001). The tensile capacity of the studs, Tcap., is taken as the smaller of the stud fracture capacity and pullout capacity, which are computed based on PCI equations with the following modifications: 1. For connections in which the studs are surrounded by a boundary element, the compressive strength of concrete is increased by 20% to reflect the confining effects of the boundary element. This increase is in accordance to tests on strength and stiffness of confined concrete (Saatcioglu and Razvi, 1992). The effective length of studs located in a boundary element is taken as the depth of the boundary element. Stud fracture strength is computed based on the ultimate tensile strength of the stud instead of the yield strength. The pullout capacity of stud groups is computed by assuming that the failure cone makes an angle of 55 degrees with the plane of the stud head in accordance to the failure patterns observed in the tests. This angle is different from 45 degrees assumed in PCI equations, but is consistent with provisions in Appendix D of the 2002 edition of ACI Building Code (ACI, 2002) and within the range of 50 to 60 degrees recommended by others (e.g., Rehm et al., 1988). The larger angle of 55 degrees increases the projected area of cone failure by 25.8% over the value stipulated in PCI equations.

2. 3. 4.

86

As evident from Table 1, the analytical model accurately predicts the mode of failure. Moreover, the model establishes a very close estimate of the measured capacities (see Fig. 15); on the average, the experimental capacities are Table 1 Experimental and analytical modes 2% more than the computed values (with a coefficient of of failure variation of 1.3%). Note that without the aforementioned (a) R-Wall adjustments, the computed modes of failure and capacities would have been appreciably different from their Connection Computed Observed experimental counterparts (Shahrooz et al., 2003b). Bottom Stud Fracture Stud Fracture Top Stud Pullout Stud Pullout (b) T-Wall Connection Computed Observed Bottom Stud Fracture Stud Fracture Top Stud Fracture Stud Fracture
1.1 1 0.9

Measured/Computed

0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 R-Wall:Bottom Connection R-Wall:Top Connection T-Wall:Bottom Connection T-Wall:Top Connection

Figure 15 Comparison of measured and computed connection capacities

SUMMARY AND CONCLUSIONS

A number of past and ongoing studies at the University of Cincinnati have examined seismic behavior of hybrid core wall structures. In particular, seismic behavior of steel and hybrid coupling beams, and connections between outrigger beams and core walls have been investigated. Previous research has shown the viability of steel and hybrid coupling beams, and has been instrumental in development of design methods. Adequate capacity and distribution of dissipated energy for steel and hybrid coupling beams are achieved by following these design methods. Ongoing studies are aimed at developing the next generation of steel and hybrid coupling beams in order to eliminate and/or limit damage to wall piers while dissipating a significant amount of energy, and to develop an easily repairable system. Test results from a number of preliminary specimens are promising, and suggest that inelastic deformation can be concentrated in replaceable fuses while ensuring elastic response of the coupling beam. An accurate design model for outrigger beam-core wall connections has been developed. This model predicts the connection capacity and its expected mode of failure. Using this model, a capacity design approach may be followed to prevent brittle failure modes (i.e., stud pullout failure or stud fracture) by dissipating energy through ductile yielding and eventual fracture of the shear tab connecting the outrigger beam to the embedded stud plate. Using this approach, the shear tab will act as an effective energy dissipating fuse.
ACKNOWLEDGEMENTS

The reported research is based on investigations sponsored by the National Science Foundation under grants BCS-9319838, CMS-9632496 and CMS-9714860, with Dr. Shih Chi Liu as the program director. These projects were part of the fifth phase of US-Japan cooperative research program on composite and hybrid structures. Any opinions, findings, and conclusions or recommendations expressed in this paper are of those of the authors and do not necessarily reflect the views of the sponsors. 87

REFERENCES

American Concrete Institute (2002), Building Code Requirements for Structural Concrete (318-02) and Commentary (318R-02), Farmington Hills, MI. American Institute of Steel Construction (AISC) (2002), Seismic Provisions for Structural Steel Buildings, Chicago, IL. Ferver, G.W., Mark, M.H. and Lawrence, J.R. (1974), Practical Design of a Coupled Shear Wall Building, Proceedings of the Regional Conference on Tall Buildings Bangkok, 49-64. Gong, B., Shahrooz, B.M. (2001a), Concrete-Steel Composite Coupling Beams-Part I: Component Testing, Journal of the Structural Division, ASCE, Vol. 127, No. 6, 625-631. Gong, B., Shahrooz, B.M. (2001b), Concrete-Steel Composite Coupling Beams-Part II: Subassembly Testing and Design Verification, Journal of the Structural Division, ASCE, Vol. 127, No. 6, 632-638. Harries, K. A., Mitchell, D., Redwood, R. G., and Cook, W. D. (1997), Seismic Design of Coupling Beams - A Case for Mixed Construction, Canadian Journal of Civil Engineering, Vol. 24, No. 3, 448-459. International Conference of Building Officials (ICBO) (2000), International Building Code. Marcakis, K. and Mitchell, D. (1980), Precast Concrete Connections with Embedded Steel Members, Prestressed Concrete Institute Journal, Vol. 25, No. 4, 88-116. Mattock, A. H. and Gaafar, G. H. (1982), Strength of Embedded Steel Sections as Brackets, ACI Journal, Vol. 79, No. 2, 83-93. Lehmkuhl, E. (2002), Renaissance A Composite Coupled Shear Wall System, Proceedings of the 2002 SEAOC Convention. Paulay, T., (2002), Personal Correspondence, August, University of Canterbury, New Zealand. PCI Design Handbook - Precast and Prestressed Concrete (2001), Precast Concrete Institute, Chicago, IL. Rehm, G., Eligehausen, R., and Mallee, R. (1998), Befestigungstechnik (Fastening Technique), Betonkalender, Verlag Wilhelm Ernst & Sohn, Berlin. Saatcioglu, M., and Razvi, S. R. (1992), Strength and Ductility of Confined Concrete, Journal of Structural Engineering, ASCE, Vol. 118, No. 6, pp. 1590-1607. Shahrooz, B. M., Remmetter, M. A., and Qin, F. (1993), Seismic Design and Performance of Composite Coupled Walls, Journal of the Structural Division, ASCE, Vol. 119, No. 11, 3291-3309. Shahrooz, B.M., Tunc., G., Deason, J. T. (2002), RC/Composite Wall-Steel Frame Hybrid Buildings: Connections and System Behavior, Report No. UC-CII 02/01, Cincinnati Infrastructure Institute. Shahrooz, B.M., Deason, J. T., Tunc, G. (2003a), Outrigger Beam Wall Connections: Part I-Component Testing and Development of Design Model, Journal of the Structural Division, ASCE, in press. Shahrooz, B.M., Tunc, G., Deason, J. T. (2003b), Outrigger Beam Wall Connections: Part II-Subassembly Testing and Further Modeling Enhancements, Journal of the Structural Division, ASCE, in press. Taranath, B.S. (1998), Steel, Concrete and Composite Design of Tall Buildings second edition, Section 6.4.3.1 First City Tower, McGraw Hill. Yao, C.C. (2003), Personal Correspondence, February, Principal, Read Jones Christoffersen Ltd., Vancouver, B.C.

88

Вам также может понравиться