cyclic behaviour of interior post-tensioned flat plate...

13
Cyclic behaviour of interior post-tensioned flat plate connections S. W. Han*, S.-H. Kee*, T. H.-K. Kang†, S.-S. Ha*, J. W. Wallace† and L-H. Lee* Hanyang University; University of California In high seismic regions, post-tensioned (PT) slab column frames are commonly used to support gravity loads in conjunction with a lateral-force resisting system (LFRS) such as a core wall. The LFRS is designed to resist 100% of the design lateral forces as well as to limit lateral displacements to an acceptable level, whereas the slab column frame must sustain the gravity loads under the expected (design) displacements. Given the relatively sparse data on the seismic performance of PT flat plate slab column frames, cyclic tests of four interior PT slab column connections were conducted. Primary test variables were the level of gravity shear at the slab column connection and the slab tendon arrangement. Test results indicate that both the test variables strongly influence the cyclic behaviour of the PT connections, and that the use of slab bottom reinforcement at the slab column connection was effective in resisting positive moment developed under lateral loading as well as improving the hysteretic energy dissipation capacity. Notation A ps area of tendons A sm area of continuous bottom bonded reinforcement b 0 perimeter of critical section b 1 width of critical section parallel to loading direction b 2 width of critical section perpendicular to b 1 c distance from the centroid of critical section to the perimeter of critical section c 2 column dimension perpendicular to the loading direction d effective slab depth d b bar diameter d p effective slab depth of tendons f 9 c peak concrete compressive stress f pc average compressive stress in concrete owing to effective post-tensioning force only f ps nominal stress in an unbonded tendon (limited to lesser of f py and f se + 210) f py tendon yield stress f se effective tendon stress f y yield stress of slab bottom reinforcement h slab thickness J c polar moment of inertia of critical section l 1 centre-to-centre spans parallel to the loading direction l 2 centre-to-centre spans perpendicular to the loading direction M n, c þ 3 h nominal moment for a slab width of c 2 + 3h M n,fs nominal moment for a full slab width M u,ub peak unbalanced moment V c nominal concrete shear strength V g gravity force to be transferred from a slab to a column V p vertical component of effective post- tensioning force at section V u factored shear force to be transferred from a slab to a column â c ratio of long side to short side of a column ª f factor used to determine the unbalanced moment transferred by flexure at slab– column connections ª v fraction of unbalanced moment transferred by eccentric shear í c nominal shear strength provided by concrete í u maximum shear stress Ł u drift ratio at punching Ł y drift ratio at yielding * Division of Architectural Engineering, Hanyang University, Seoul 133–791, Korea. † Department of Civil and Environmental Engineering, University of California, Los Angeles (UCLA), California 90095, USA. (MCR 51489) Paper received 29 November 2005; accepted 7 August 2006 Magazine of Concrete Research, 2006, 58, No. 10, December, 699–711 699 www.concrete-research.com 1751-763X (Online) 0024-9831 (Print) # 2006 Thomas Telford Ltd

Upload: others

Post on 23-Mar-2020

10 views

Category:

Documents


0 download

TRANSCRIPT

Cyclic behaviour of interior post-tensioned flat

plate connections

S. W. Han*, S.-H. Kee*, T. H.-K. Kang†, S.-S. Ha*, J. W. Wallace† and L-H. Lee*

Hanyang University; University of California

In high seismic regions, post-tensioned (PT) slab–column frames are commonly used to support gravity loads in

conjunction with a lateral-force resisting system (LFRS) such as a core wall. The LFRS is designed to resist 100%

of the design lateral forces as well as to limit lateral displacements to an acceptable level, whereas the slab–

column frame must sustain the gravity loads under the expected (design) displacements. Given the relatively sparse

data on the seismic performance of PT flat plate slab–column frames, cyclic tests of four interior PT slab–column

connections were conducted. Primary test variables were the level of gravity shear at the slab–column connection

and the slab tendon arrangement. Test results indicate that both the test variables strongly influence the cyclic

behaviour of the PT connections, and that the use of slab bottom reinforcement at the slab–column connection was

effective in resisting positive moment developed under lateral loading as well as improving the hysteretic energy

dissipation capacity.

Notation

Aps area of tendons

Asm area of continuous bottom bonded

reinforcement

b0 perimeter of critical section

b1 width of critical section parallel to loading

direction

b2 width of critical section perpendicular to b1c distance from the centroid of critical section

to the perimeter of critical section

c2 column dimension perpendicular to the

loading direction

d effective slab depth

db bar diameter

dp effective slab depth of tendons

f 9c peak concrete compressive stress

fpc average compressive stress in concrete

owing to effective post-tensioning force only

fps nominal stress in an unbonded tendon

(limited to lesser of fpy and fse + 210)

fpy tendon yield stress

fse effective tendon stress

fy yield stress of slab bottom reinforcement

h slab thickness

Jc polar moment of inertia of critical section

l1 centre-to-centre spans parallel to the

loading direction

l2 centre-to-centre spans perpendicular to the

loading direction

Mn,c þ 3 h nominal moment for a slab width of c2 + 3h

Mn,fs nominal moment for a full slab width

Mu,ub peak unbalanced moment

Vc nominal concrete shear strength

Vg gravity force to be transferred from a slab

to a column

Vp vertical component of effective post-

tensioning force at section

Vu factored shear force to be transferred from a

slab to a column

�c ratio of long side to short side of a column

ªf factor used to determine the unbalanced

moment transferred by flexure at slab–

column connections

ªv fraction of unbalanced moment transferred

by eccentric shear

�c nominal shear strength provided by concrete

�u maximum shear stress

Łu drift ratio at punching

Ły drift ratio at yielding

* Division of Architectural Engineering, Hanyang University, Seoul

133–791, Korea.

† Department of Civil and Environmental Engineering, University of

California, Los Angeles (UCLA), California 90095, USA.

(MCR 51489) Paper received 29 November 2005; accepted 7 August

2006

Magazine of Concrete Research, 2006, 58, No. 10, December, 699–711

699

www.concrete-research.com 1751-763X (Online) 0024-9831 (Print) # 2006 Thomas Telford Ltd

rp reinforcement ratio associated with tendons

(¼ Aps/l2dp)

� strength reduction factor

øu uniformly distributed design load

Introduction

Slab–column framing is popular because it offers

distinct advantages such as low cost owing to ease of

construction (e.g. slip forms), low floor-to-floor height

and flexible use of space. In particular, post-tensioned

(PT) flat plate slab systems are very efficient, since the

PT flat plate slab systems provide improved crack and

deflection control, and allow relatively large slab span-

to-thickness ratios, in the order of 35 to 45. Slab–

column frames are commonly used to resist gravity

loads in high seismic regions (SDC-D or -E, IBC-031),

where SDC indicates seismic design category; however,

they may be utilised as intermediate moment frames

(ACI 318–052 section 21.12.6) in areas with moderate

seismic demands (SDC-A, -B and -C). Given the broad

potential applications, a detailed understanding of

slab–column frame behaviour subjected to lateral

forces and/or displacements is important.

According to the seismic provisions in ACI 318–05,2

structural systems are either designated to resist earth-

quake forces [i.e. be part of the lateral-force resisting

system (LFRS)] or they are referred to as ‘non-partici-

pating’ systems or ‘gravity’-force resisting systems

(GFRS). In high seismic regions, post-tensioned slab–

column frames are commonly used for GFRS, particu-

larly for residential and office buildings where the LFRS

consists of shear walls or moment-resisting frames at

the building perimeter. When subjected to earthquake

ground motions, the LFRS undergoes lateral deforma-

tions, which are imposed on the GFRS through the floor

diaphragms. The lateral displacements imposed on the

slab–column frame are likely to introduce significant

unbalanced moments on the slab–column connections,

increasing the potential for punching failures. Before

the introduction of ACI 318–05,2 no specific require-

ments existed to avoid punching failures at slab–column

connections in GFRS owing to the lateral displacement

of the LFRS. However, the ability to maintain gravity

loads after punching failure (i.e. post-punching resis-

tance) could be justified owing to the ACI 3182 chapter

7 requirement for integrity reinforcement (continuous

bottom reinforcement through the column cage).

According to the eccentric shear stress model in ACI

318–05,2 punching failures at slab–column connec-

tions subjected to shear (Vu) and unbalanced moment

(Mub) occur where the shear stress owing to direct shear

and eccentric shear exceeds the nominal shear stress on

the slab critical section. The eccentric shear stress is a

result of moment ªvMub. The remaining portion of Mub

equal to ªfMub is assumed to be resisted by flexure by

providing slab reinforcement over a slab flexural trans-

fer width of c2 + 3h. Prior studies3–5 indicate that the

eccentric shear stress model gives conservative results

for the punching shear capacity of the post-tensioned

interior slab–column connections.

Available seismic provisions for the flat plate slab

systems (e.g. ACI 318–052 section 21.11) focus on

conventional non-prestressed slab–column connections:

similar provisions for prestressed post-tensioned slab–

column connections do not exist. In particular, very

limited testing has been conducted on the PT interior

connections subjected to reversed cyclic loads

(Qaisrani,5 Pimanmas et al.6). For PT flat plate slab

systems with shear reinforcement, existing data are

limited to the shake table tests of an approximately

one-third scale, two-storey, post-tensioned flat plate

frame (Kang and Wallace7). Therefore, seismic per-

formance of the PT connections, which tend to have

relatively high gravity shear ratios owing to the sub-

stantial spans, have not been adequately studied.

Furthermore, the impact of the tendon arrangement,

which typically involves the use of banded tendons in

one direction and distributed tendons in the other direc-

tion (ACI 423.3R-968), has not been systematically

assessed, and the influence of bonded bottom reinforce-

ment on the behaviour of PT flat plate systems sub-

jected to moment reversal has not been addressed.

Test programme

Prototype and specimen design

Figure 1 depicts the prototype building selected to

assist in the determination of specimen proportions and

details. The prototype building is a ten-storey, post-

tensioned flat plate slab system with 3.5 m storey

heights with special reinforced concrete (RC) shear

walls (R ¼ 6) designed according to IBC-031 and ACI

318–052 requirements for SDC-E (high seismic risk).

A slab thickness of 200 mm and a span length of 8 m

were selected for the prototype building, resulting in a

span-to-depth ratio of 40, which is within the range of

35 to 45 typically used for PT flat plate construction.

The shear walls were designed to resist the lateral

earthquake forces, whereas the PT flat plate slab sys-

tems were proportioned to support gravity forces.

Given that the prototype building is used for typical

residential/office construction, the unit weight of the

building is estimated as 23.5 kN/m3. Additional dead

loads (e.g. partition walls) of 0.5 kPa and live loads of

2 kPa were chosen based on recommendations in IBC-

03.1 Slab moments and shear forces owing to factored

gravity loads were determined based on the results of

an elastic analysis using MIDAS/Gen (MIDAS IT,9

Version 6.3.2).

The specimens were approximately two-thirds scale

representations of a typical interior slab–column con-

nection within the prototype building (Fig. 1). Inflection

points were assumed to occur at or near slab mid-span,

S. W. Han et al.

700 Magazine of Concrete Research, 2006, 58, No. 10

resulting in a span length of 4.6 m for the loading direc-

tion. A slightly shorter span length of 3.6 m was used in

the transverse direction. A 300 3 300 mm column

cross-section and a 130 mm slab thickness were selected

based on the two-thirds scale factor.

Slabs were post-tensioned with 12.7 mm diameter,

seven-wire strands with design yield stress of

1861 MPa. The arrangement of slab reinforcement is

provided in Fig. 2. Minimum concrete clear cover was

12 mm for both top and bottom slab reinforcement.

The post-tensioning strands were greased and placed in

polyethylene tubes with a diameter of 16 mm; there-

fore, the tendons were ‘unbonded’. The number of

post-tensioning tendons and the post-tensioning force

per tendon were selected such that approximately 100%

of slab dead weight was balanced (Table 1). The result-

ing average compressive stress ( fpc) of 1.21 MPa is

within the allowable range of 0.88 MPa and 3.44 MPa

as specified by ACI 318–05.2 According to section

18.9 of ACI 318–05,2 minimum bonded top reinforce-

ment ( fy ¼ 352 MPa) was placed within an effective

transfer width of c2 þ 3h at the connection region.

Although the slab–column frame is designed for

gravity loads, it is subjected to the lateral deformations

imposed on it by the lateral system (shear walls). To

investigate the influence of the lateral deformations of

the LFRS on the slab moment distribution, three-

dimensional elastic analyses of combined PT flat plate

frame and shear wall system were conducted for var-

ious levels of seismic demand. Resulting slab moments

are depicted in Fig. 3. The slab moment diagrams

indicate that the positive slab moment eventually devel-

ops on one side of the slab–column connection as the

seismic demand is increased; therefore, sufficient bot-

tom reinforcement should be provided to avoid the

formation of large cracks. Requirements for minimum

bonded bottom reinforcement for the post-tensioned

connections do not exist; therefore, it is common to

provide structural integrity reinforcement per ACI

318–052 section 7.13.2.5. This provision was devel-

oped primarily based on common practice for RC

slab–column frames, which typically have slab span-to-

thickness ratios closer to 20. Owing to the larger spans

of typical of PT construction, bottom reinforcement

satisfying ACI 3182 (section 7.13.2.5) and ACI-ASCE

35210 requirements [equation (1)] was provided

Asm ¼ 0:5øu l1 l2

� f y(1)

where � ¼ 0.9. Accordingly, five D10 (db ¼ 10 mm)

bottom bars were placed within a column width of c2for both directions, whereas no bottom reinforcement

was provided outside of c2.

Material properties

Normal Portland cement concrete with a design con-

crete strength of 29.8 MPa and a maximum aggregate

size of 19 mm were used. The measured slump of the

concrete ranged from 100 to 125 mm. As indicated in

Table 2, mean compressive test-day strength of the con-

crete was 32.3 MPa based on compressive test results

of five, 100 3 200 mm concrete cylinders in accor-

dance with KS F 2404 (KS-0211). Seven-wire strands

[steel wire for prestressed concrete (SWPC) 7B11,

db ¼ 12.7 mm] and D10 mild reinforcement were used

for slab reinforcement, whereas D25 reinforcement

(db ¼ 25 mm) was used for column longitudinal rein-

forcement. The SWPC 7B strands ( fy ¼ 1555 MPa)

and reinforcement ( fy ¼ 466 MPa) were fabricated in

accordance with KS D 7002 and KS B 0802, respec-

��������� ��� ���

���

���

���

���

���

���������� ��������

���

�������

����

�����

���

��������� ��� ���

Fig. 1. Prototype building

Cyclic behaviour of interior post-tensioned flat plate connections

Magazine of Concrete Research, 2006, 58, No. 10 701

tively (KS-0211). Properties determined from the mate-

rial test programme are summarised in Table 2, whereas

the stress–strain relations obtained from representative

samples are shown in Fig. 4.

Testing and instrumentation

The four interior post-tensioned slab–column con-

nections were subjected to uni-directional, reversed

cyclic loading using the test set-up illustrated in Fig. 5.

The primary variables of the test programme were the

level of gravity shear, which could be varied using

the hydraulic jack at the base of the column, and the

arrangement of slab tendons (Table 1). For two of the

specimens (PI-B50 and PI-B30), tendons were banded

in the loading direction (E–W) and approximately uni-

formly distributed in the transverse direction (N–S),

whereas for the other two specimens (PI-D50 and PI-

D30), the tendons were uniformly distributed in the

loading direction (E–W) and banded in the transverse

direction (N–S), as shown in Fig. 2. The name PI-B50

refers to the following variables: Post-tensioned, Inter-

ior, Banded, with 50% gravity shear ratio. The gravity

shear ratio is calculated as Vg/�Vc , where Vg is the

factored gravity shear determined from analysis, Vc is

calculated according to ACI 318–052 (equation (11–

36)), which includes the impact of the prestress, and

� ¼ 0.75.

The base of the column was pinned, and the slab

edges were pin-supported at the four corner points by

struts (steel bars) with a diameter of 10 cm (Fig. 5). To

evaluate the appropriateness of the boundary condi-

���������� �������������

���

��������

��� !"#��$%��&�

!'��

���

� �

����

(��

(��

���������� �����

���

���

���

)��

#�

����

����

���$

���)

��

���

�&�

(��

(����(�

���

����

)��

��(�

!"#��$%

��&�

!"#��$%�$��

!'������

!"#��$%�$�� ������� #

���� ���$ ���) ����

����

���&

����

���

��*� ��+

� &���

� &� �

����

���&

��*� ��+

���

�!"#��$%

�$��

�!"#��$%

�$�

� &����

� &����

���

���

Fig. 2. Details of unbonded tendons and bonded top bars of

specimens: (a) PI-B50 and PI-B30; (b) PI-D50 and PI-D30

Table 1. Properties of specimens

Mark c2: mm d†: mm rp: % f 9c: MPa fpc: MPa Vc: kN Vg: kN

PI-B50 300 104 0.21 32.3 1.21 343 132

PI-B30 300 104 0.21 32.3 1.21 343 82

PI-D50 300 104 0.16 32.3 1.21 343 132

PI-D30 300 104 0.16 32.3 1.21 343 82

c2: column dimension in the direction perpendicular to loading

d†: effective depth based on average of d ps in two directions

rp: ratio of post-tensioning tendons

f 9c: mean compressive concrete strength at the time of testing

fpc: average compressive stress in concrete owing to effective post-tensioning

Vc: nominal shear strength

Vg: gravity force to be transferred from slab to column

e.g. PI-B50, (P): PT, (I): interior, (B): banded, (50): Vg/�Vc � 0.50 (� ¼ 0.75)

�)��

� ��

�&��

&��

��

)��

,��

����-.

��

'�+��� �/����0����

0� %� ��

� & ) � � �)0� %� ��

���'���&������"'�!%����'���&������"'�!�����'���&������"'�!'����'���&������"'�!��

Fig. 3. Slab moment distribution for both gravity and lateral

loads

S. W. Han et al.

702 Magazine of Concrete Research, 2006, 58, No. 10

tions, an elastic finite element analysis was conducted

using MIDAS/Gen.9 This analysis showed that, for a

given drift ratio, the difference in the unbalanced mo-

ments between the test condition (pinned at four cor-

ners) and the complete building system was less than

3%.

Gravity loads were simulated by applying an axial

load at the base of the column using a 5000 kN hydrau-

lic jack, as well as by placing loading blocks on the

slab. Given that the ratio of shear force and moment at

the critical section of the slab–column connection sig-

nificantly influences the behaviour of the con-

nections,12,13 the location of the loading blocks was

determined based on an finite element analysis to

match the ratio of shear force and moment for the

specimens and the prototype building. As shown in Fig.

5, load cells were mounted on the hydraulic jack at the

base of the column and on each of the four struts at the

slab corners to monitor the applied loads.

The 250 kN capacity horizontal actuator was used to

displace the top of the column (Fig. 5). A typical lateral

displacement history consisting of three cycles at

monotonically increasing drift levels between 0.2% and

6%, as shown in Fig. 6, was used for the tests. The

applied lateral load and displacements were monitored

using a load cell and linear variable differential trans-

ducers (LVDTs), respectively, and additional LVDTs

were used to monitor (and avoid) slab twisting. Strain

gauges were attached on selected bonded slab bars, and

eight load cells were installed on tendons to measure

the tendon forces during post-tensioning as well as

changes in tendon forces during testing.

Table 2. Properties of materials

Concrete

f 9c: MPa �o Es: MPa

32.3 0.00185 29 600

Steel rebar

db: mm fy: MPa �y Es: MPa fu: MPa �u

10 466 0.0024 193 166 698 0.1245

25 465 0.0027 172 222 584 0.0819

SWPC 7B strand

dps: mm fy: MPa �y Es: MPa fu: MPa �u

16 1555 0.0098 182 223 1751 0.049

f 9c: mean compressive concrete strength at the time of testing

fy: yield stress, f u: ultimate stress

�o: mean strain at peak concrete strength, �y: yield strain, �u: ultimate strain

db or dps: rebar or strand diameter

Es: steel modulus of elasticity

� �+� �+����+

&��

��

)��

���

����

�&��

� ��

�)��

����

'���1 ��

'&��1 ��

"#��$%��&�$�����$!2�� �+����

� �+� �+�� ++��,�

� ���& ��� ���) ���� ��� ���& ���

��

��

&�

&�

��

��

����� � �+�� ++��,

����� � �+����+������ ����� ������ �����& �����) ����&� ����& ����&� �����&

�+���&�������,��

�� �� ������������,��

� ���������$��,��

Fig. 4. Stress–strain relations of representative materials

Cyclic behaviour of interior post-tensioned flat plate connections

Magazine of Concrete Research, 2006, 58, No. 10 703

Test results

Crack patterns and observed damage

Crack patterns at the completion of testing are shown

in Fig. 7. In general, for drifts less than 1%, slab

flexural cracks formed adjacent to the column and sub-

sequently extended across the entire slab width. For

drift ratios greater than 3%, significant cracks were

observed to extend approximately 300 mm (ffi 2.5d)

away from the column face (Fig. 7). Based on observa-

tions, as well as the test data presented in the following

subsections, it is concluded that punching failures oc-

curred for all test specimens after flexural yielding of

slab bonded reinforcement.

Lateral load versus lateral drift

Values of unbalanced moments obtained from two

different approaches are plotted in Fig. 8. The first

estimate of the unbalanced moment was obtained as the

lateral load multiplied by a column height of 2.1 m,

whereas the second estimate was derived from the

difference of vertical reactions measured in the struts at

the slab corners times the distance between the struts

(4.8 m). A comparison of the results obtained using the

two approaches is provided in Fig. 8 and indicates that

consistent results were obtained.

Envelopes to the load–displacement relations as well

as an idealised bi-linear relationship4,14,15 fitting to the

envelope relations also are depicted in Fig. 9. The bi-

linear relations are used to determine drift angles asso-

ciated with yielding (Ły) and punching (Łu) as noted in

Table 3. As depicted in Fig. 9, Łu is defined as the drift

ratio at punching (Łu,1) for specimens that experienced

sudden punching failures (PI-B50, PI-B30 and PI-D50),

or the drift angle at which the lateral load experiences

a 20% drop from the peak lateral load (Łu,2) for the

specimen PI-D30, which did not experience a sudden

punching failure.

As noted for reinforced concrete connections,4,14–17

the limited data available have shown that lateral drift

capacity at punching for post-tensioned slab–column

connections is also strongly influenced by the magni-

tude of direct gravity shear stress applied on the critical

section.4,5 A detailed review of the existing database

for the PT connections4 indicates that higher drift ratios

at punching were obtained for the PT connections, in-

part owing to the larger span-to-thickness ratios (40 to

45) compared with the RC connections (� 25), and that

lower drift values were observed for the PT connections

(��

����

���

�����

�������+� ����

1��������-+

���

���

&���-.���3���

���

� �

) �

&���

#� ����������-+ ����������-+

���

����� ���

��������

���+�3� ��+��3���������������+� ����

"����������� � /���

�����-.45��3���6�-

"����������� � � ������2

Fig. 5. Test set-up

�)

�&

&

)

��&�

�(�

�)�

���

��

)�

(�

�&�

�5� +

��&7 ��&�7����7���7 ��$�7���7 �� 7��$�7&�&7

&�$�7���7

��7

)��7

�� �����/��������7

� � �� �� &� &� �� �� � � ��

�� ����+*� �

������

Fig. 6. Loading histories

S. W. Han et al.

704 Magazine of Concrete Research, 2006, 58, No. 10

subjected to reversed cyclic loading relative to the PT

connections subjected to monotonic or repeated lateral

loading (Fig. 10). Since the database for PT interior

connections subjected to reversed cyclic loading is lim-

ited to the four, isolated specimens with relatively high

gravity shear ratios (Vg/Vc ¼ 0.46 to 0.72) tested by

Qaisrani5 and Pimanmas et al.6, the test results pre-

sented for this study provide valuable information to

assess trends for gravity shear ratios Vg/Vc between

approximately 0.25 and 0.40 (actual values for the con-

nections based on actual material properties, versus the

design values of 0.30 and 0.50, respectively).

Figure 10 presents data (Vg/�Vc; Łu) for the four

specimens tested (Table 3), along with the existing test

data of post-tensioned slab–column connections with-

out shear reinforcement. Based on the results pre-

sented in Fig. 10, a fairly consistent trend of

decreasing drift ratio at punching (Łu) for increasing

gravity shear ratio (Vg/�Vc) is observed. The four test

specimens from this study achieved substantially high-

er lateral drifts (3.2% to 5.9%) compared with earlier

test results (1.8 to 2.3%) for interior connections sub-

jected to cyclic loading with higher gravity shear

ratios.5,6). Furthermore, specimens with lower gravity

shear ratios (PI-B30 and PI-D30) achieved 180 and

135% of the lateral drift ratios of the companion

�8!%��

�8!'�� �8!'��

�8!%��

&����&����

�����

&����

&����&����

&����&����

�����&����

&����&����

Fig. 7. Observed damage

� +�����+ 93 ��

����

���

��

���

����8!%��

����

����

���

��

���

����8!'��

����� �����

����-.

������� �����

����-.

:�

/����;���<������3�������� +/����� ��������+�3� ��+��3������ +

/����;���<������3�������� +/����� ��������+�3� ��+��3������ +

Fig. 8. Comparisons of unbalanced moments

Cyclic behaviour of interior post-tensioned flat plate connections

Magazine of Concrete Research, 2006, 58, No. 10 705

specimens with higher gravity shear ratios (PI-B50

and PI-D50), respectively, before punching failures

(Fig. 10).

For PI-B30 and PI-D30, relatively large lateral drift

ratios were achieved for both specimens, 5.6 to 5.9%,

respectively, indicating that the tendon arrangement did

not impact the drift at punching failure significantly for

the moderate gravity shear ratio. However, for the high-

er gravity shear ratio, based on the load–displacement

relations (Fig. 9), higher drift capacity (120%) and im-

proved ductility (190%) were observed for PI-D50

compared with PI-B50. The lower drift capacity and

ductility for PI-B50 compared with PI-D50 is likely

owing to larger precompression within the connection

region. In both cases of gravity shear ratios, larger drift

ratio at punching failure was achieved for the case with

distributed tendons. Current practice, which typically

bands tendons in one direction, may result in connec-

tions that have less drift capacity for loads parallel to

the banded direction.

Dissipated energy

The dissipated energy per cycle, which is used to

assess the hysteretic damping characteristics of the

system, was calculated based on the area of a load–

displacement relation for that cycle (Fig. 9). Based on

evaluation of these data, it is concluded that a gravity

shear ratio (Vg/�Vc) has a modest impact of the accu-

mulated dissipated energy for specimens with the same

tendon distribution (for a given cycle). However, as

shown in Fig. 11, the values of the accumulated dissi-

pated energy at failure differ substantially, as specimens

PI-B30 and PI-D30 have values that are approximately

150% and 40% higher than those for PI-B50 and PI-

D50, respectively. The PT connections with higher

gravity shear (PI-B50 and PI-D50) tend to fail at a

����8!%�� �����8!%��

�����8!'�������8!'��

&��7 ���7 &��7 ��(7

���7�� 7

��� 7� ��7

'��/��������7�) � �& � & )

'��/��������7�) � �& � & )

'��/��������7�) � �& � & )

'��/��������7�) � �& � & )

�)�

� �

�&�

&�

)�

�� ������-.

�)�

� �

�&�

&�

)�

�� ������-.

�)�

� �

�&�

&�

)�

�� ���

���-.

�)�

� �

�&�

&�

)�

�� ���

���-.

�* -����&�&�-.

&:���* -����� ���-.�������=3�:������&�����

�� ��

�������=3�:������&�����

�* -��� $���-.

&:���* -�������(�-.�������=3�:������&�����

�� ��

�������=3�:������&�����

�* -��� ��)�-.

&:���* -����&(����-.

�������=3�:������&�����

�� ��

�������=3�:������&�����

�* -���� $���-.

&:���* -����������-.�������=3�:������&�����

����

�������=3�:������&�����

Fig. 9. Lateral load plotted against lateral drift ratio

S. W. Han et al.

706 Magazine of Concrete Research, 2006, 58, No. 10

lower drift ratio, and thus achieve less ductility. In

particular, yielding of the slab reinforcement of PI-B50

was limited prior to punching failure. These results

indicate that the hysteretic energy dissipation capacity

is dependent upon the gravity shear ratio.

Test data indicate that the hysteretic energy dissipa-

tion capacity also is sensitive to the tendon distribution

pattern. Based on the relations plotted in Figs 9(a) and

(b), specimens PI-B50 and PI-B30 exhibited fairly lin-

ear behaviour up to 2% drift with limited yielding of

both top and bottom slab bonded reinforcement, prob-

ably as a result of high local precompression provided

by the banded post-tensioning within the connection

region. On the other hand, for specimens PI-D50 and

PI-D30, a significant drop in lateral stiffness was ob-

served at 1% drift (Figs 9(c) and (d)). As a result, the

values of accumulated dissipated energy for PI-B50

and PI-B30 are approximately 90% and 40% of those

for PI-D50 and PI-D30, respectively (Fig. 11).

Bonded slab reinforcement

As mentioned earlier, the test specimens included

minimum bonded top reinforcement (eight D10 bars) in

accordance with ACI 318–052 and bonded bottom rein-

forcement (five D10 bars) as required for structural

integrity reinforcement based on ACI 352.1R-89.10 Dur-

ing the cyclic tests, both top and bottom bonded reinfor-

cing bars reached the yield strain of approximately

0.002 (Fig. 12). Yield of top reinforcement initiated at

drift ratios of 1.1, 1.7, 0.6 and 1.7% for PI-B50, PI-B30,

PI-D50 and PI-D30, respectively, whereas yield of bot-

tom reinforcement occurred between drift ratios of 2.2%

and 3.5% for all specimens. Figures 12(a) and (b) show

that, before punching, the degree of yielding of top

reinforcement was more extensive than that of bottom

reinforcement. Bottom reinforcement strains are plotted

against the measured drift ratios in Fig. 13 for the bars

located at the column centreline. Owing to gravity load-

ing, bottom reinforcement was in compression before

the application of lateral loading. The strains in bottom

reinforcement began to vary from negative to positive

values (i.e. moment reversal) at drift ratios of 0.8% and

0.5% for PI-B50 and PI-D50, respectively. This result is

consistent with the observation that larger pre-compres-

sion within the connection region may have existed

where banded tendons were parallel to the loading direc-

tion. For PI-B30 and PI-D30, with relatively low gravity

shear stresses, moment reversal was observed at rela-

tively low drift ratio of 0.5%.

In high seismic regions, the lateral (roof) drift de-

mand on the structure is typically limited to 2% for life

safety considerations,18 and some storey drift ratios

typically exceed the roof drift ratio. Bonded bottom

steel with the amount of Asm (equation (1)) is not

anticipated to yield for drifts less than approximately

1.5%;14 however, moment reversal is likely to occur

(Fig. 3). Therefore, bonded bottom reinforcement

should be provided for PT flat plate systems designedTable3.Testandanalyticalresults

Mark

Testresults

GSR

Shearstrength

Flexuralstrength

Moment

Failure

mode

Ppeak:kN

Ł y:%

Ł u:%

Mu,ub:

kN/m

� u:MPa

Vg/V

c� u:MPa

� u/�

cªfM

u,ub:

kN/m

PM

n,cþ3h:

kN/m

PM

n,fs:

kN/m

ªfM

u,ub/

PM

n,cþ3h

Mn,ub:

kN/m

Mu,ub/M

n,ub

12

34

56

78

¼5

/7

910

11

12

¼9

/10

13

14

¼4

/13

PI-B50

43. 6

2. 1

3. 3

91. 6

2. 38

0. 39

2. 01

1. 18

55. 0

78. 7

105. 7

0. 70

80. 2

1. 14

F+P

PI-B30

47. 8

2. 0

5. 9

100. 4

2. 22

0. 24

2. 01

1. 11

60. 2

79. 0

106. 5

0. 76

98. 4

1. 03

F+P

PI-D50

49. 1

1. 4

4. 0

99. 3

2. 51

0. 39

2. 01

1. 25

59. 6

59. 4

104. 1

1. 00

80. 2

1. 24

F+P

PI-D30

52. 2

1. 5

5. 4

109. 6

2. 38

0. 24

2. 01

1. 18

65. 8

59. 6

104. 8

1. 10

98. 4

1. 13

F+P

GSR:gravityshearratio

ªf¼

0. 6

(accordingto

ACI318–05)

Mu,ub¼

(Ppeak)(h1),whereh1¼

2. 1m

Mn,cþ3h:flexuralmomentcapacityofslabsover

thewidth

ofc 2

+3h

�M

n,fs:flexuralmomentcapacityofslabsover

entire

slab

width

F+P:flexuralyielding,followed

bypunchingfailure

Cyclic behaviour of interior post-tensioned flat plate connections

Magazine of Concrete Research, 2006, 58, No. 10 707

to resist gravity loads since tensile strains in the bottom

slab reinforcement was observed before reaching a lat-

eral drift ratio of 1.5% (Fig. 12(b)). The quantity of

bottom reinforcement provided in the specimens was in

compliance with equation (1) and was sufficient to

resist positive moment developed up to lateral drifts of

approximately 1.5% to 2.0% before reaching the yield

strain (Fig. 12(b)), as well as to prevent progressive

collapse. Accordingly, bottom reinforcement required

by equation (1) is sufficient to limit extensive yielding;

therefore, there does not appear to be a need to provide

larger quantities of bottom reinforcement.

Stresses in unbonded tendons

For each specimen, average tendon stresses were

monitored using seven or eight load cells mounted at

����

���&

����

���

����

���)

���$

����

���(�������

������

'��/���������*3�

�;���

>����5�+; ���������:�

� ��� ��& ��� �� ��� ��) ��$ ��� ��( �

0�8����!������/������/������� ��3:����������$��

% +�!/����� /�����

���� �����+2��;�3��+; �� ��/��� � ��+3�6 �� ����

� � �+ ���5���� �������

% +�!/����� /���������� �����+2��;�3�+; �� ��/��� � ��

'��+ ��+�� / � �� ����?������#�

���?����� �@�(

,���� <!��3<��&��!��A�@4���� �@& �!��A�@,���� <!��3<��&��!����� �";��&��!��A�@��2��;�+; ��� ��/@�1���;� ����>;�&��!��A�@��2:�+; ��� ��/@�?������#� �!����/�� �2:�+; ��� ��/��� � ���

";��&��!��A�@

������;�����42-��+&��!�8��@B�+�����!�8��@�����+��� �@)�!�8��@

4���� �@���� + ����!�8��@������;�����42-��+&��!��A�@

C�3��;��� �@&&�!��A�@

Fig. 10. Gravity shear ratio plotted against drift ratio at punching (PT slab–column connections)

��

��

&�

&�

���8!%���8!%���8!'���8!'��

0��3�

3� ����++�*

� �� � ��5��-.:�

� � �� �� &� �� � �� ���&��5� +

Fig. 11. Accumulated dissipated energy

'�+��� �/������3���� ��� ����

"������3� +%��� ����*�� ��/��� � �������� ��� ����+

�8!%���8!%���8!'���8!'��

���7 ���

���7

����

����

&���

���������� ������ #

"����+�����)

��

��������� �)�� � �� �&�� � &�� �� )�� ��� ����

'�+��� �/������3���� ��� ����

�8!%���8!%���8!'���8!'��

���7 ���7

&���

&���

����

���������� ������ #

"����+�����)

�&��� ����� ����� ���� � ��� ���� ���� &���

��

"������3� +������ ��� ����+

%��� ���������� ��/��� � �����

��

Fig. 12. Strain distribution in (a) top and (b) bottom reinfor-

cement passing through the column cage

S. W. Han et al.

708 Magazine of Concrete Research, 2006, 58, No. 10

tendon ends (TC, see Fig. 2). Figure 14 reveals that the

stress in the unbonded tendons increased as a lateral

drift ratio increased and that the tendon distribution

influenced the change in the tendon stress in the load-

ing direction. For the PI-B specimens, tendon stresses

in the loading direction (TC 1, 2 and 3) tended to

increase fairly uniformly irrespective of tendon loca-

tion. However, for the PI-D specimens, only the stresses

in tendons placed close to the column (TC 1 or 2)

increased significantly (Fig. 14).

The peak values of the increase in the tendon stresses

are in the range of 4.7 to 6.4% of the effective tendon

stress ( fse). These values are substantially smaller than

those (� 15%) predicted by using equation (2) (ACI

318–052 equation 18–5)

f ps ¼ f se þ 70þ f 9c

300rpMPa (2)

The smaller increases may be partly attributed to the

smaller total elongation of the tendon obtained for the

interior connection under both gravity and lateral loads

(plotted against the interior connection under gravity

loads only, where the tendon in always in the tension

zone). The lower stresses in the unbonded tendons under

lateral loads result in reduced moment capacities under

combined gravity and lateral load, relative to the nom-

inal moment capacities computed for gravity load alone.

Shear strength of post-tensioned slab–column

connections (ACI 318–05)

The shear strength of the specimens was evaluated

using the eccentric shear stress model of ACI 318–052

chapter 11

�u ¼Vu

b0d� ªvMu,ubc

Jc< ��c (3)

ªf ¼1

1þ 2=3ð Þffiffiffiffiffiffiffiffiffiffiffiffib1=b2

p (4)

ªv ¼ 1� ªf (5)

��c ¼ � �pffiffiffiffiffiffif 9c

pþ 0:3 fpc þ

Vp

b0d

� �(6)

where �vc is the nominal shear stress capacity of the

post-tensioned connection without shear reinforcement

reduced by the capacity reduction factor in units of

MPa, �p is the smaller of 0.29 or (Æsd/b0 + 1.5)/12 and

Æs is 40 for interior columns.

In Table 3 column 10, the nominal moment for a

slab width of c2 + 3h (Mþn,cþ3 h + M�

n,cþ3 h) is calculated

for the specimens using the actual concrete strength

( f 9c), the actual yield stress of bonded bars ( fy), and the

measured stress in unbonded tendons at failure ( fps),

given that all the slab bonded reinforcement yielded at

the time of the failure. It is noted that the moment

transferred by flexure (ªfMu,ub ffi 0.6Mu,ub) and the

flexural transfer capacity (Mþn,cþ3h + M�

n,cþ3h) are al-

most identical (Table 3, column 12), suggesting that the

assumed fraction of unbalanced moment transferred by

flexure (ªf ¼ 0.6) according to ACI 318–052 is reason-

able for the PI-D specimens. For the PI-B specimens,

most of post-tensioning tendons and all the bonded

reinforcement were placed within c2 + 3h (Fig. 2).

Thus, information provided in column 12 in Table 3 is

not sufficient to assess the fraction of unbalanced mo-

ment transferred by flexure.

The peak values of the applied unbalanced moment

(Mu,ub) monitored from a load cell attached to the

horizontal actuator and the values of the total flexural

moment capacity (Mþn,fs + M�

n,fs) computed using meas-

ured material properties for the specimens are indicated

�����

����

���

����

����

&���

&���

����

�8!%���8!%���8!'���8!'��

"����+������

)

'��/��������7� ��� � ��� & &�� � ���

�5

Fig. 13. Strains in bonded bottom reinforcement versus drift

ratio

�&�

&�

)�

��

����8!%�� �����8!%�� �����8!'�� �����8!'��

���� ����

��������

���&����

���

����

���)

���$���&����

���

����

���)���$���&����

����

�������)

���$

���&

����

�������) ���$

"�� ++����� �

��+����3����� �

� ����+��,�

� � & � �'��/��������7

� � & � �'��/��������7

� � & � �'��/��������7

� � & � �'��/��������7

� � & � �'��/��������7

Fig. 14. Stress increments in unbonded tendons versus lateral drifts

Cyclic behaviour of interior post-tensioned flat plate connections

Magazine of Concrete Research, 2006, 58, No. 10 709

in columns 4 and 11 in Table 3, respectively. Note that

Mn,f s denotes the moment capacity considering full slab

width, not c2 + 3h. As might be expected, the values

obtained using the two different approaches are quite

close (ratios ¼ 0.87 to 1.05), and also consistent with

other independently obtained measurements (load cells

installed at the base of slab end points) (see Fig. 9).

These results support that the shear stress capacities

(vc) of the specimens were greater than the peak shear

stresses (vu) such that the full flexural moment capaci-

ties were reached prior to punching failures that even-

tually occurred at relatively large drifts (3.2 to 5.9%).

The peak shear stress (vu) owing to direct shear and

eccentric shear obtained using equations (3) to (6) with

peak unbalanced moment (Mu,ub) was compared with

the nominal shear strength (�c ¼ 0.29ffiffiffiffiffiffif 9c

p+ 0.3fpc

+ Vp/b0d, where Vp is negligible) for the specimens in

column 8 in Table 3. Based on the results, the eccentric

shear stress model gives conservative results for the

nominal shear strength for the PT specimens, as well as

for the post-tensioned interior slab–column specimens

tested earlier (Fig. 15). Results for this test programme

are consistent with results for earlier test programmes

for gravity loads, although the ratios tend to be a bit

lower, possibly owing to the modest increase in tendon

stress under lateral loads noted earlier.

Conclusions

Experimental studies of four isolated, post-tensioned

interior slab–column connections subjected to both

gravity and cyclic lateral loading were conducted.

Based on the test results, the following conclusions are

reached.

(a) Consistent with observations for reinforced con-

crete slab–column connections, the level of gravity

shear on the slab critical section significantly influ-

ences the cyclic behaviour of the post-tensioned

slab–column connections. As the gravity shear ra-

tio increased, a drift ratio at punching and the

hysteretic energy dissipation capacity for the post-

tensioned connections decreased. The improved

hysteretic energy dissipation for the connections

with lower gravity shear (PI-B30 and PI-D30) was

owing to more extensive yielding of bonded rein-

forcement prior to punching failure.

(b) Results indicate that seismic performance of the

post-tensioned flat plate slab systems is impacted

by the tendon distribution. For the higher gravity

shear ratio (40%), higher drift capacity and im-

proved ductility were observed for PI-D50 com-

pared with PI-B50. For these cases, limited

yielding of both top and bottom slab reinforcement

was noted and the use of banded tendons appears

to create larger precompression within the connec-

tion region.

(c) Moment reversal (change from negative slab mo-

ment due to gravity load to positive slab moment

under lateral load on one side of a connection)

occurred between lateral drifts of 0.5 to 0.8%. In

turn, bonded bottom reinforcement, which was

placed according to ACI 318–052 and ACI

352.1R-89,10 reached yield at lateral drifts between

2.2% to 3.5%. Based on these results, it is con-

cluded that bonded bottom reinforcement should

be provided for the post-tensioned flat plate slab

systems; however, integrity reinforcement required

by chapter 7 of ACI 318–05 appears sufficient to

limit rebar yielding. In addition, the bottom rein-

forcement improves the hysteretic energy dissipa-

tion capacity of the PT interior connections.

(d ) Test results indicate that tendon stresses for com-

bined gravity and lateral loading were approxi-

mately 35 to 60% of the values predicted by ACI

318–052 provisions. The reduced tendon stress

should be considered for flexural design of the

post-tensioned flat plate slab systems in high seis-

mic regions, as moment reversal is anticipated.

(e) The validity of the eccentric shear stress model

defined in ACI 318–052 was assessed. The test

results indicate that the eccentric shear stress mod-

el gives reasonable predictions for the nominal

shear strength (vc) and the flexural transfer capa-

city.

Acknowledgements

The work presented in this paper was sponsored by

Hanyang University, MOST R01–2006–000–10722–0

and SRC/ERC R11–2005–056–04002–0. The views

expressed are those of authors and do not necessarily

represent those of the sponsor.

References

1. International Code Council. International Building Code

(IBC-03). Whittier, California, 2003.

��&

���

��

���

��)

��$

���

����

� �:�

�� ���:&��,

��

:&

� ��� ��& ��� �� ��� ��) ��$ ��� ��( ��*�:�� ���

�:&��,��:&

>����5�������� +�+�

1 * � ��� ��������� +�+�1 � �+ ���5����������� +�+��� + ���� +��*������

Fig. 15. Shear strength of post-tensioned interior slab–

column connections

S. W. Han et al.

710 Magazine of Concrete Research, 2006, 58, No. 10

2. American Concrete Institute. Building Code Requirements

for Structural Concrete and Commentary (ACI 318-05). ACI,

Farmington Hills, Michigan, 2005.

3. Burns N. H. and Hemakom R. Test of post-tensioned flat plate

with banded tendons. Journal of Structual Engineering, ASCE,

1985, 111, No. 9, 1899–1915.

4. Kang T. H.-K. and Wallace J. W. Punching of reinforced and

post-tensioned concrete slab–column connections. ACI Structur-

al Journal, 2006, 103, No. 4, 531–540.

5. Qaisrani A. N. Interior Post-Tensioned Flat-Plate Connections

Subjected to Vertical and Biaxial Lateral Loading. PhD thesis,

University of California, Berkeley, 1993.

6. Pimanmas A., Warnitchai P. and Pongpornsup S. Seismic

performance of 3/5 scaled post-tensioned interior flat slab–

column connections. Asia Conference on Earthquake Engineer-

ing (ACEE 2004), Manila 2004 (CD-ROM).

7. Kang T. H.-K. and Wallace J. W. Dynamic responses of flat

plate systems with shear reinforcement. ACI Structural Journal

2005, 102, No. 5, 763–773.

8. Joint American Concrete Institute–American Society of

Civil Engineers Committee 423. Recommendations for Con-

crete Members Prestressed with Unbonded Tendons (ACI

423.3R-96). ACI, Farmington Hills, Michigan, 1996.

9. Midas It. MIDAS/GENw User’s Manual. Version 6.3.2, Seoul,

Korea, 2004.

10. Joint American Concrete Institute–American Society of

Civil Engineers Committee 352. Recommendation for Design

of Slab–Column Connections in Monolithic Reinforced Con-

crete Structures (ACI 352.1R-89). ACI, Farmington Hills, Michi-

gan, 1989.

11. Korean Standard Association. Standard Specification (KS-

02). KSA, Seoul, Korea, 2002.

12. Akiyama H. and Hawkins N. M. Response of Flat Plate Con-

nection Structures to Seismic and Wind Forces. University of

Washington, Seattle, 1984. Structures and Mechanics Report

SM 84–1, 1–267.

13. Pan A. D. and Moehle J. P. An experimental study of slab–

column connections. ACI Structural Journal, 1992, 89, No. 6,

626–638.

14. Pan A. D. and Moehle J. P. Lateral displacement ductility of

reinforced concrete flat plates. ACI Structural Journal, 1989, 86,

No. 3, 250–258.

15. Megally S. and Ghali A. Design considerations for slab–

column connections in seismic zones. ACI Structural Journal,

1994, 91, No. 30, 303–314.

16. Hueste M. B. D. and Wight J. K. Nonlinear punching shear

failure model for interior slab–column connections. Journal of

Structural Engineering ASCE, 1999, 125, No. 9, 997–1008.

17. Moehle J. P. Seismic design considerations for flat-plate con-

struction. Mete A. Sozen Symposium, Tarpon Spring, Florida (J.

K. Wight and M. E. Kreger (eds)). American Concrete Insti-

tute, Farmington Hills, MI, 1996, ACI SP-162, pp. 1–34.

18. Federal Emergency Management Agency. Pre-Standard

and Commentary for the Seismic Rehabilitation of Buildings

(FEMA-356). FEMA, Washington DC, 2000.

19. Kang T. H.-K., Lafave J. M., Robertson I. N. and Hawkins

N. M. Post-tensioned slab–column connections: Drift capacity

at punching of connections subjected to lateral loading. ACI

Concrete International, to be published in 2007.

20. Trongtham N. and Hawkins N. M. Moment Transfer to Col-

umns in Unbonded Post-tensioned Prestressed Concrete Slabs.

University of Washington, Seattle, 186 pp., 1977, Report

SM77-3.

21. Shatila M. Prestressed Concrete Slab-edge Column Connec-

tion. MSc thesis, University of Calgary, 1987.

22. Foutch D. A., Gamble W. L. and Sunidja H. Tests of post-

tensioned concrete slab-edge column connections. ACI Struct-

ural Journal, 1990, 87, No. 18, 167–179.

23. Matinez-Cruzado J. A. Experimental Study of Post-tensioned

Flat Plate Exterior Slab–Column Connections Subjected to

Gravity and Biaxial Loading. PhD thesis, University of Califor-

nia, Berkeley, 1993.

24. Han S. W., Kee S.-H., Park Y.-M., Lee L.-H. and Kang

T. H.-K. Hysteretic behavior of exterior post-tensioned flat plate

connections. Engineering Structures, in press, doi: 10.1016/j.

engstruct.2006.03.029.

25. Ritchie M. and Ghali A. Seismic-resistant connections of edge

columns with prestressed slabs. ACI Structural Journal, 2005,

102, No. 2, 314–323.

Discussion contributions on this paper should reach the editor by

1 June 2007

Cyclic behaviour of interior post-tensioned flat plate connections

Magazine of Concrete Research, 2006, 58, No. 10 711