Sei sulla pagina 1di 98

Shear Strength of Slab-Corner Column Connections

by Claudia Correa Agudelo

June 2003

Department of Civil Engineering and Applied Mechanics McGill University Montreal, Canada

A Thesis submitted to the Faculty of Graduate Studies and Research in partial fulfillment of the requirements for the degree of Master of Engineering

/
Claudia Correa Agudelo, 2003

't

'

To my Mom and Dad

Shear Strength of Slab-Corner Column Connections

Abstract
The 1994 CSA Standard requires that slab-column connections be designed for one-way shear and two-way shear action. Three full-scale slab and corner column specimens were constructed and tested to failure to investigate the influence of the size of corner column on the slab shear capacity. The

current design provisions of the 1994 CSA Standard and EC2-02 were compared with the experimental results of the three specimens. The experimental results indicate that the predictions for one-way shear action are more accurate than the predictions for two-way shear action. The beneficial effects of increasing the column size in improving the shear capacity and mode of failure are demonstrated.

Resistance en Cisaillement de connexions dalle-colonne de coin

Resume
La norme CSA 1994 exige que les connexions dalles-colonnes soient congues I'effet de cisaillement uni et bidirectionnel. Trois specimens de dalle et colonne de coin en grandeur reelle ont ete construits et testes jusqu'a la rupture afin d'evaluer I'influence de la taille des colonnes sur la resistance en cisaillement des dalles. Les resistances calculees selon les normes CSA 1994 et EC2-02 ont ete comparees avec les resultats experimentaux des trois specimens. Les resultats experimentaux demontrent que les predictions pour I'effet de cisaillement en une direction sont plus precises que les predictions pour I'effet bidirectionnel. Les essais demontrent que la resistance en cisaillement

augmente avec la taille des colonnes.

Acknowledgements
The author would like to express her deepest thanks to Professor Mitchell for his expert guidance and his continued encouragement, advice and support throughout this research programme. The author would also like to thank Dr. William Cook for his assistance and his helpful suggestions. The research was carried out in the Jamieson Structures Laboratory at McGill University. The author wishes to thank Ron Sheppard, John Bartczak, Marek Przykorski and Damon Kiperchuk for their assistance in the construction and testing of the specimens. The assistance of Zhenyu Li in the laboratory is also appreciated. The valuable help of the secretaries of the Civil Engineering Department, particularly Ann Bless, Sandy Shewchuk-Boyd, Franca Delia Rovere, and Anna Dinolfo was much appreciated during the course of this research project. The author would also like to express her deepest gratitude to her parents, and her sister Marisabel for their constant support. Finally, the author would like to thank the Engineer Louis Brunet for his continuous encouragement, help, patience and understanding.

Claudia Correa May, 2003

Table of contents

Abstract Resume Acknowledgements List of Figures List of Tables List of Symbols Chapter 1: Introduction and Literature Review 1.1 1.2 1.3 1.4 1.5 Introduction Key Research Studies North American Codes European Codes Research Objectives

v v ix 1 1 2 12 16 18 20 20 20 22 22 22 23 23 24 24 24 27 27 28 46 47 49 51 68 68 72

Chapter 2: Experimental Programme 2.1 Description of Prototype Structure 2.2 Design of Test Specimens 2.3 Details of Test Specimens 2.3.1 Specimen S1 2.3.2 Specimen S2 2.3.3 Specimen S3 2.4 Construction Sequence 2.5 Material Properties 2.5.1 Reinforcing Steel 2.5.2 Concrete 2.6 Testing Procedure 2.6.1 Test Setup and Loading Apparatus 2.6.2 Instrumentation Chapter 3: Experimental Results 3.1 Specimen S1 3.2 Specimen S2 3.3 Specimen S3 Chapter 4: Analyses and Comparison of Results 4.1 Comparison of Experimental Results 4.2 Comparison of Experimental Results with Code Predictions

IV

Chapter 5: Conclusions 5.1 Conclusions References

80 80 82

List of Figures

Chapter 2 2.1 2.2 2.3 2.4 2.5 2.6 2.7 2.8 2.9 2.10 2.11 2.12 2.13 2.14 2.15 2.16 2.17 2.18 Chapter 3 Load versus average deflection of Specimen S1 Strain distribution of top reinforcement in Specimen S1 at key load stages 3.3 Flexural cracking on top surface of Specimen S1 at a shear of 75.5 kN 3.4 Flexural shear cracks at a shear of 111.8 kN in Specimen S1, north face 3.5 Maximum crack width for Specimen S1 3.6 Shear failure in Specimen S1, north face 3.7 Shear failure in Specimen S1, west face 3.8 Top surface after shear failure in Specimen S1 3.9 Overall view of shear failure in north face in Specimen S1 at shear of 125.6 kN 3.10 Load versus average deflection of Specimen S2 3.11 Strain distribution of top reinforcement in Specimen S2 at key load stages 3.12 Flexural cracking on top surface of Specimen S2 at a 3.1 3.2 53 53 54 54 55 56 56 57 57 58 58 Plan of prototype flat plate structure Reinforcement details for bottom steel in S1 Top reinforcement in Specimen S1 Detailed view of the joint in Specimen S1 Bottom reinforcement details for Specimen S2 Arrangement on top steel for Specimen S2 Detailed view of the joint in Specimen S2 Bottom reinforcement details for Specimen S3 Top steel for specimen for Specimen S3 Detailed view of the joint in Specimen S3 Overall view of the specimens before casting Stress-strain curves for reinforcing bars Compressive stress-strain responses for concrete Shrinkage strain curve Test Setup for specimens Overall view of loading setup Location of LVDTs Position of strain gauges 30 31 32 33 34 35 36 37 38 39 39 40 40 41 42 43 44 45

VI

3.13 3.14 3.15 3.16 3.17 3.18 3.19 3.20 3.21 3.22 3.23 3.24 3.25 3.26

shear of 75.2 kN Flexural shear cracks at a shear of 113 kN in Specimen S2, west face Maximum crack width for Specimen S2 Shear failure in Specimen S2, north face Shear failure in Specimen S2, west face Top surface after shear failure in Specimen S2 Overall view of shear failure in north face in Specimen S2 at shear of 116.7 kN Load versus average deflection of Specimen S3 Strain distribution of top reinforcement in Specimen S3 at key load stages Flexural cracking on top surface of Specimen S3 at a shear of 67.1 kN Flexural shear cracks at a shear of 120.9 kN in Specimen S3, west face Maximum crack width for Specimen S3 Flexure-shear failure in Specimen S3, north face Flexure-shear failure in Specimen S3, west face Top surface after flexure-shear failure in Specimen S3

59 59 60 61 61 62 62 63 63 64 64 65 66 66 67

Chapter 4 4.1 4.2 4.3 4.4 Load versus average deflection Load versus average deflection Variation of strength predicted by the 1994 CSA Standard with the column size Variation of strength predicted by the 1994 CSA Standard and EC2-02 with the column size 77 77 78 79

VII

List of Tables

Chapter 2 2.1 2.2 2.3 Chapter 4 Summary of reinforcing bar properties Concrete mix design Concrete properties 24 25 26

4.1 4.2
4.3 4.4

Summary of key load stages for Specimens S1, S2 and S3 ....69 Maximum crack widths at peak loads for Specimens S1, S2 and S3 71 Predicted failure loads for Specimens S1, S2 and S3 according to the 1994 CSA Standard 73 Predicted shear strength for the three specimens with EC2-02 75

VIII

List of Symbols
Ac Asw b b0 c C Ci area of critical section area of shear reinforcing steel perimeter of loaded area critical shear perimeter size length of the square column nominal shear stress according to plastic limit analysis size of the edge column perpendicular to the edge of the slab (CT > c2) size of the edge column parallel to the edge of the slab effective depth maximum aggregate size diameter of the column eccentricity of the applied load from the centroid of the column eccentricity of the applied load in the x direction (ex > ey) eccentricity of the applied load in the y direction concrete compressive strength jd k h J fr fsp Fsw fsw fuit fy fyk
pun

vertical component of the forces in inclined prestressing tendons punching load modulus of rupture of concrete splitting strength of concrete vertical component of the forces in the shear reinforcement stress of shear reinforcement ultimate steel strength of reinforcing

c2 d da d st e ex ey f'c

yield stress of steel reinforcement characteristic yield reinforcing bars thickness of the slab polar moment of inertia of the critical section about its centroidal axis flexural lever arm size effect factor stress of

k^ k 2 empirical constants in Eq. (1.12) k2 ls empirical constant in Eq. (1.28) shear span

fc20o compressive strength measured on a 200 mm diameter cylinder fck strength below which 5% of all possible strength measurements for the specified concrete may be expected to fall cube strength of

lx, ly lengths of projections of the critical section for two-way shear m Mt Mu n ultimate resisting moment design moment transferred to the column factored moment ratio of the area of the reinforcing steel in the loaded area over the total area of tensile reinforcement stress concentration factor to account for the concrete strength under a multiaxial state of stress

fck cube compressive concrete Fct

vertical component of the concrete tensile stresses the flexural nc

Fdow dowel-force of reinforcing bars

IX

p V Vc vr

applied axial force on slab shear force nominal punching shear strength provided by the concrete factored shear stress resistance

X X A.0 (Ltg p ai c o t, \\i

factor accounting density

for

concrete

ratio of the diameter of the column to the effective depth empirical parameter tension reinforcement ratio flexural reinforcement ratio principal tensile stresses in the concrete mechanical reinforcement ratio size effect factor rotation

Vfiex shear force at flexural ultimate strength Vr factored shear stress resistance Vest observed shear force at shear failure Vu vu Vu1 Vu2 x punching shear resistance ultimate shear stress cracking shear force of the conical shell in Coulomb's law cracking shear force of the catenary shell in Coulomb's law height of compression zone at flexure in tagential direction when punching occurs inclination of the punching shear crack factor which adjusts vc for support dimensions ratio of the long to the short dimensions of the column yield strain of reinforcing steel strength reduction factor for shear relationship between shear and moment capacity in the vicinity of the loaded area resistance factor for concrete fraction of unbalanced moment transferred by flexure fraction of unbalanced moment transferred by eccentricity of shear reduction factor accounting for the influence of torsion

a as pc ey (j) <o > t

<> fc yf yv r\

Chapter 1

Introduction and Literature Review

1.1 Introduction
Flat plates are very popular structural systems due to their architectural versatility and the reduced construction time through the use of flying forms. The shear strength of slabs in the vicinity of columns is governed by the more severe of two conditions, either one-way shear or two-way shear action. According to the 1994 CSA Standard for the "Design of Concrete Structures" (CSA 1994) it is necessary to check for both of these types of failure modes when designing flat plates. The objective of this thesis is to investigate the behaviour of two-way slabs failing in shear around corner columns and to propose a method for assessing the types of shear failures. In this chapter, the design approaches of the 1994 CSA Standard (CSA 1994), the American Concrete Institute "Building Code Requirements for Structural Concrete" (ACI 318-02), the German Design Standard (DIN 1045 1988), the British Standard 8110 (1997) and the Eurocode 2 (EC2-02) are discussed and compared. In addition some of the key research studies leading up to the code developments are also discussed.

1.2 Key Research Studies

Many researchers have studied the behaviour of reinforced concrete elements subjected to shear. In 1902, the German engineer, E. Morsch (Morsch 1902) developed the following equation for nominal shear stress, v, in a beam subjected to flexural cracking: v = bjd lever arm. Morsch also described the resisting mechanisms for shear in the form of a truss idealization with parallel chords and compressive diagonals inclined at 45 degrees with respect to the longitudinal axis. A few years later, A. N. Talbot presented the results of his research study on reinforced concrete footings which has had a significant impact on design practice. Talbot (1913) carried out experiments on 114 wall footings and 83 column footings. He calculated the shear stress in the footings around the (1.1)

where V is the shear, b is the perimeter of the loaded area and jd is the flexural

square columns from the following equation: v= , V , 4(c + 2d)jd (1.2)

where c is the side length of the square column and d is the effective depth of the footing. Talbot found that the shear strength increased as the amount of tensile reinforcement in the slabs was increased. Based on the specimens that failed in shear, he concluded that the critical section was located at a distance of d from the face of the column. In 1933 Graf studied the shear strength of slabs loaded by concentrated loads near the supports. He found that the shear capacity decreased as the load

was moved away from the supports and increased with increasing concrete strength. Graf proposed the following formula for the shear stress in slabs: v = 4ch where h is the thickness of the slab. He suggested that the flexural cracking might have some influence on the shear strength. In 1939 Richart and Kluge (1939) found that an increase of the flexural strength of the slabs would result in increased shear strength. In 1946, Forsell and Holmberg reported results of different tests that were carried out over a period from 1926 to 1928. They assumed that the shear stress distribution was parabolic and utilized a critical section at a distance of 0.5h from the edges of the loaded area. Their expression for the maximum shear stress was:
1

(1.3)

-5V

M A\

bh They also concluded that increasing the slab moments decreased the shear strength. Shear tests on mortar bridge deck slabs were reported at the University of Illinois by Newmark, Siess et al. (1946). Their results indicated that the failure loads were dependent on the same factors as the loads at first yielding. In 1948, Richart (1948) concluded that shearing stresses at failure of slabcolumn connections, calculated on a critical section located at a distance d from the faces of the columns, varied generally from 0.05 f'c to 0.09 f'c. In 1953, based on the results reported by Richart (1948), Hognestad (1953) suggested that the shearing stresses be computed at zero distance from the loaded area or column faces. Hognestad developed the following expression for the shear stress at failure:

v=

V 7 8 bd

'

-_ 0.07A 0.035 + f c +130psi

(1.5)

where <o is the ratio , where Vtest is the observed shear force at shear > |
Vflex

failure and Vfiex the shear force at flexural ultimate strength as computed by yieldline theory. In 1956, Elstner and Hognestad (1956) reported that the concentration of the tension and compression reinforcement in the column region as well as the effect of eccentricity of the applied load had no effect on the ultimate shearing strength. The shearing strength was calculated as: v = = 333 psi + 0 . 0 4 6 ^ 7 4>0 8bd (1.6)

In 1957, an ultimate strength theory for shear was presented by Whitney (1957), based on previous test results but ignoring those that involved bond failure. He assumed that the shear strength is primarily a function of the ultimate resisting moment, m, inside the "pyramid of rupture" whose surface makes 45 degrees with the column. The shear strength at a distance d/2 from the perimeter of the loaded area was given by v = 100psi + 0.75^- P(1.7)

where ls is the "shear span" which is the distance between the support and the nearest edge of the loaded area in the case of a slab supported along the edges. A good approximation of the ultimate shear stress was found by Scordelis, Lin and May (1958), when using equations (1.6) and (1.7). In 1961, Moe (1961) verified that there is an interaction between flexure and shear which Elstner and Hognestad and Whitney had suggested (see Eq. (1.6) and (1.7)).

Based on test results in the literature and on the results of tests on 43 slabs, tested under different conditions, he found that the shear force at the calculated ultimate flexural capacity of the slab, Vfiex, was one of the parameters governing the shear strength of slabs and footings. The concrete strength, f'c, and the ratio of the side length of the loaded area to the slab thickness, c/d, were also directly related to the calculation of the ultimate shearing strength which was expressed as

15 1-0.075v bd 5.25bdJf'c y 1+ Vflex

15 1-0.075-

5.25<|)c

(1.8)

Moe concluded that the inclined cracking load should be determined on the basis of stresses computed at a distance of d/2 from the periphery, while the shear compression failure should be predicted on the basis of the stresses on the periphery of the loaded area or column. Moe suggested that for design, the shearing stresses should be calculated as: v = 9.23-1.12d v = 2.5 + 10 fc for-<3 d for- > 3 d (1.9a) (1.9b)

At the same time, a rational mechanical model for the punching shear failure of slabs was proposed by Kinnunen and Nylander in Europe. Based on 61 tests on circular slabs with circular column stubs, the model predicts the ultimate load for both flexural and punching failures, and is also capable of predicting the deformation of slabs at failure. The model involves the rigid

rotation of slab portions separated by radial cracks. This research was the basis of the 1964 Swedish Building Code (1964) and had significant influence upon European code considerations. In 1975, Hawkins, Mitchell and Hanna (1975) found that concentrating the reinforcement in the column region, within a distance of 1.5 times the slab

thickness on both sides of the column, improved the resistance to punching shear, however the ductility was reduced (Hawkins and Mitchell 1979). During the 1970's, limit analysis was applied to study all types of failure, including brittle failures such as punching shear. One of these studies was

published in 1976 by Braestrup, Nielsen et al. (1976). Based on the theory of plasticity, they assumed concrete as a perfectly plastic material. They used Coulomb's law to predict the punching shear resistance, Vu, by comparing the fracture energy of the conical shell with the work performed by the applied loads, giving: Vu = V u 1 +V u 2 cracking shear of the catenary shell. In 1985 Regan (1985) discussed the punching shear resistance of connections of slabs with edge and corner columns. Assuming that the (1.10)

where V u i describes the cracking shear of the conical shell and VU2 gives the

resistance to nodal forces and torsion at the slab edge are lost when torsion cracking occurs, the pure bending resistance is the ultimate moment provided by the reinforcement perpendicular to the edge. Rankin and Long (1987) proposed a new formula to calculate the punching shear strength, V, which involves the reinforcement ratio, p (equal to As/bd) as follows:

V=1.66 7^7(c + d)d VlOOp

(1.11)

In 1987 several tests were carried out by Bazant and Cao (1987), in order to quantify the size effect on the punching shear resistance. This was based on the concept that the punching load should be predicted by fracture mechanics instead of plastic limit analysis because the failure does not occur simultaneously along the ultimate failure surface. They concluded that the nominal stress at failure decreases as the size increases and regarded this behaviour as a direct confirmation of the size-effect law. The punching load was proposed as:

v =C 1 + V Ma7

(1.12)

where vu is the nominal shear stress; d is the effective depth of the slab; X0 is the empirical parameter characterizing the fracture energy of the material and the shape of the structure; d a is the maximum aggregate size; and C is the value of the nominal shear stress according to plastic limit analysis and is calculated as C = k 1 f c ^1 + k 2 ^ where ki and k2 are empirical constants. In 1989 Georgopoulos (1989) proposed a method for the calculation of the punching shear strength and the angle of inclined cracking of a slab without shear reinforcement. He assumed that approximately 75% of the punching shear was carried by the principal tensile stresses a, in the concrete and could be calculated from the vertical component of the resulting tensile force. Hence, the punching shear was calculated as:

Vu =4.13a 1 d 2 cota - + 0.20 + 0.35cota


v2 J

(1.13)

where GAS given by 0.17 (fck, Cube)2/3; d is the effective depth; a is the inclination of the inclined punching shear crack (tana = 0.056/co + 0.3); c is the mechanical o reinforcement ratio = p fyk/Wubei A, is dst/hm and dst is the diameter of the column. Based on the model proposed by Shehata in 1985, a simplified method was proposed by Shehata and Regan in 1990. They considered that the

punching region of a slab is divided into rigid radial segments that rotate around the center of rotation located at the column face and at the level of the neutral axis. To calculate the ultimate punching shear resistance, they defined three critical states at which the frontal part of the radial segment fails to support the force at the column face: 1. If the angle a of the compressive force reaches 20, there are principal tensile stresses in the compressed front part and failure occurs by splitting

of the concrete. Experimental and numerical data show that the angle of inclination of the internal crack surface can be approximated as 20 from the horizontal. 2. If the average radial strain on the compressed face reaches a value of 0.0035 at the column face, there is a radial crushing of the concrete. 3. If the tangential strain of the compressed face reaches 0.0035 at a distance x from the column face, there is tangential crushing of the concrete. The rotation \\iu at which any of these critical states is first achieved, is required to obtain the punching shear capacity, Vu, with the conditions of equilibrium on a radial segment being satisfied. The resulting expression for the punching shear capacity is: Vu = 27idstxn c f'ctan10 o (1.14)

where dst is the diameter of the column; x is the height of compression zone at flexure in tangential direction when punching occurs; and nc is the stress concentration factor to account for the concrete strength. In the same year, Broms (1990) modified the model of Kinnunen and Nylander. In his theory Broms includes unsymmetrical punching and size effect. At the same time, Bortolotti (1990) applied the theory of plasticity to calculate the punching shear. A three-dimensional axi-symmetrical model and rigid plastic material properties were assumed in using a modified Coulomb yield criterion for the concrete. Alexander and Simmonds (1992) developed a model that combines radial arching action with the concept of a critical shear stress on a critical section (beam action shear). For brittle punching failure, bond strength of the

reinforcement is seen as the significant factor limiting the beam action shear. In 1996, Menetrey (1996) presented an analytical method to predict the punching shear capacity. This method was based on experimental results that show the influence of the punching cone inclination and the difference between

punching shear failures and flexural failures. This method considers the inclined principal tensile stresses in the concrete. Considering also the contribution of each type of reinforcement crossing the punching shear crack, the punching load of a general slab is expressed as: Fpun = Fct + Fdow + Fsw + Fp (1.15)

where Fct is the vertical component of the concrete tensile stresses obtained by integrating the vertical components of the tensile stresses in concrete, Fd0w is the dowel-force contribution of the flexural reinforcing bars, Fsw is the vertical component of the forces in studs, stirrups or bent-up bars which are wellanchored, and Fp is the resultant of the vertical components of the forces in inclined prestressing tendons. Theodorakopoulos and Swamy (2002) presented a method to predict the ultimate punching shear strength of slab-column connections. The model is

applied to lightweight and normal weight concrete and both normal strength and high strength concrete. Failure is assumed to occur when the tensile splitting strength of the concrete is exceeded. Research on the punching shear strength of edge and corner columns was made by Andersson (1966) using yield line theory. Considering a nominal shear stress located at a distance of d/2 from the column, he proposed the following formula for the calculation of the punching shear stress: ^V_ 1 + d2 v= d 0.4e

MS
6 e
+Tf

1 +1^

(1.16)

where V is the shear force; Ci and c2 are sizes of the column; d is the effective depth and e is the eccentricity of the load from the centroid of the column. Zaghlool and de Paiva (1973a) developed a theoretical method to analyse the punching shear strength of a corner column connection considering the effect of the c/d ratio, the MA/ ratio and the steel configuration in the section close to

the column. Zaghlool and de Paiva (1973b) reported on test results on slabs with corner columns these results indicated the conservative predictions the ACI Code. Ingvarsson (1977) suggested a critical section passing through a point located d/2 from the inner corner of the column and projecting at an angle of 45 degrees from the column face. This gives a critical perimeter b0 = d + V2 (bi + b2). Based on one-way shear, an ultimate shear stress at failure was given as:

vu = r,^ (0.126 + 2.24p) Ji\

(1.17)

where rj is a reduction factor accounting for the influence of torsion expressed as


ri = 1

'

ex-ey+0.5(0,-c2)" d +^2(0, +c 2 )

where the ex and ey are the eccentricities of the applied loads (MA/) and ex > ey and C1 > c2; \ = 1.75 - 1.25 d > 1.0. Regan and Braestrup (1985) describe the approach taken by Regan to determine the punching shear resistance of edge and corner slab-column connections. This approach involves considerations of different peripheries for the minimum shear resistance, the maximum shear resistance and the flexural resistance. Walker and Regan (1987) carried out tests on 11 slabs with corner columns, where some of the columns did not extend above the slab. They

observed that yield lines could form over the column so that the width for moment transfer was reduced. The moment distribution obtained from the equivalent frame analysis of the ACI and British Codes and those from the test results were compared. Based on 27 previous experimental results, Moehle (1988) proposed a strength method of analysis assuming that if the total shear force in the connection is less than 75 percent of the pure shear capacity, there is no interaction between shear and moment. The method evaluates the shear

10

strength, which is critical in the section area located at a distance of d/2 from the face of the column. The ultimate shear stress is calculated in MPa units as < 2^ v. =0.17 1 + fc <V0-33fc c

(1.18)

3c

where pc is the ratio of the long to the short dimensions of the column. The moment strength is calculated as the flexural strength provided by the reinforcement placed within a band width of c2 + 2 Ci, where c2 is the size of the column parallel to the edge of the slab and Ci is the dimension of the column perpendicular to the edge (but not to exceed the distance from the inner face of the column to the slab edge). Failure is determined from the critical failure mode, that is shear or moment. This model was adopted by ACI-ASCE Committee 352 (1988) as an optional design model. Test results of six slabs with edge columns were reported by Mortin and Ghali (1991). The purpose of this research was to investigate the effectiveness of providing stud-shear reinforcement on the capacity of slab-column

connections.

Two of the specimens had no shear reinforcement and the

remaining four had different arrangements of shear studs. Lim and Rangan (1995) developed a truss theory to predict the punching shear strength in slabs with edge and corner columns. stud-shear reinforcement were tested. Hammill and Ghali (1994) reported on tests of 5 slabs with corner columns. The variables were the amount of shear reinforcement and the loading procedure. Based on experimental results, Elgabry and Ghali (1996) presented a study of the distribution of ultimate stress along the perimeter of the critical section for slabs with interior, edge and corner columns. Using elastic analysis, they suggested a new equation for yv (the portion of unbalanced moment transferred by eccentric shear stresses) that considers different shapes of the critical section. The equation for edge and corner columns was as given by: Nine specimens with

11

j-Vi=0.4 T--=1(2-s\

(1.19a)
(119b)

1+

0.2

where lx, ly are lengths of projections of the critical section for two-way shear on principal axes x and y, respectively. Ghali and Megally (1999) carried out analytical studies and suggested replacing the J x and J y in equation (1.27) by the second moment about the centroidal principal axes of the critical section, lx and ly.

1.3 North American Codes


The first standard specifications prepared by the ACI Joint Committee (ACI 1909) limited the allowable stress on the concrete in pure shear to 0.06 fc. The shear stress was computed by the following formula: v =^ (1.20,

where the critical section was taken along the perimeter b of the loaded or column area and h is the total slab thickness. ACI reports in 1916 and 1917 (1916 and 1917) gave an allowable value for the concrete punching shear stress in slabs of 0.075 f'c The 1920 ACI Standard prescribed an allowable shear stress of 0.10 f c computed at the support or loading faces for members failing in pure shear. For diagonal tension failures the allowable shear stress was computed using formula (1.1) at a distance of d/2 from the periphery of the loaded area or column, was limited to 0.035 f c.

12

In the report of the 1924 ACI Joint Committee the shear stress was calculated at a distance of (h-1!4 in.) from the periphery of the loaded area. The allowable shear stress was given by v = 0.02 f'c (1+n)< 0.03 f'c total area of tensile reinforcement. The 1956 ACI Building Code allowed maximum shear stresses of 100 psi, proportional to f c that was computed on a critical section at a distance of d from the periphery of the loaded area. The limits were calculated as follows: v < 0.03 f'c <100psi (1.22a) (1.21)

where n is the ratio of the area of the reinforcing steel in the loaded area over the

if more than 50% of the flexural reinforcement passes through the periphery; or v < 0.025 f c <85psi loaded area or reaction area. Concentration of the flexural reinforcement in a narrow band across the column was assumed to increase the shear strength. A new procedure was proposed by ACI Committee 326 (1962) for punching shear based on Moe's equation (Moe 1961), with the variable 0 taken as unity for design. From tests by Diaz de Cossio (1960), it was shown that the punching shear strength of slabs with a large ratio of column dimension to depth, c/d, approached a value close to 1.9Jf7, with f'c in psi units. From test results, assuming the ultimate shear stress as 4.0A/f'c and defining the critical section located at a distance d/2 from the periphery of the loaded area, the punching shear stress was expressed as (1.22b)

if only 25% of the flexural reinforcement passes through the periphery of the

v = ^ - = 4.0Vf7
bd

(1.23)

It was concluded that the punching shear strength was affected primarily by three variables: concrete tensile strength, that is Jf\. , the ratio of the side

13

length of the loaded area to the effective depth of the slab, c/d, and the relationship between shear and moment in the vicinity of the loaded area, < o j . > In 1974, ASCE-ACI Committee 426 presented a state-of-the-art report on the shear strength of reinforced concrete slabs. They concluded that the ultimate shear capacity for normal-weight reinforced concrete slabs of 4yjf\ was only

appropriate for concrete with strengths less than 4000 psi (28 MPa). For higher strengths the capacity depended on a power of f'c closer to the cube root rather than the square root. The 1995 ACI Code requires that for members without shear

reinforcement, the punching shear resistance must be sufficient such that: VU<(|>VC (1.24)

where Vu is the factored shear force at the section being considered; ( is the > j strength reduction factor for shear (equal to 0.85), and Vc the nominal punching shear strength provided by the concrete. The Canadian Standard (CSA 1994) has adopted the same approaches for punching shear. The equations given below are taken from the CSA

Standard and hence are in metric units and the resistance factors are treated differently than in the ACI code. Thus, in absence of shear reinforcement, the factored shear resistance is the smallest of v = v = 1 + 0.2A4cVf'c

Pcy

( 1 - 25a )

V
r

= V
v

=
c

^
V
b

+ 0.2 *.<Mf,c
o
J

(L25b) (1.25c)

vr=vc=0.4^cVfc

where the variable pc was first introduced in the 1977 Code and is the ratio of long side to short side of the column, concentrated load or reaction area; b0 is the perimeter of the critical section and J\\ is limited to 100 psi (25/3 MPa).

14

The coefficient X allows for low density concrete and is equal to 1.0, 0.85 and 0.75 for normal-weight, semi-lightweight and lightweight concrete

respectively; <> is the resistance factor for concrete (equal to 0.6); ocs is 4 for tc interior columns, 3 for edge columns, and 2 for corner columns. When the columns or capitals become very large, equation (1.25b) may govern. It is specified in the 1994 CSA Standard that the unbalanced moment at a slab-column connection must be transferred from the slab to the column by eccentricity of shear and by flexure. Shear transfer is assumed to occur in a critical section at a distance d/2 away from the face of the column, while the fraction of unbalanced moment transferred by flexure is resisted by a width of slab equal to the transverse column width c2, plus 1.5h on each side of the column. The fraction of unbalanced moment transferred by eccentricity of shear is Y = 1-Yf v where yf is the fraction transferred by flexure: (1.26a)

Yf -

J r b1 1+ v3y
N (2^

(1.26b)

and bi and b2 are the lengths of the sides of the critical shear perimeter for moment transfer. The unbalanced moment transferred by eccentricity of shear is yv Mf, where Mf is the unbalanced moment at the centroid of the critical section. The factored shear stress on the critical transfer section is the sum of stresses caused by transfer of direct shear and unbalanced moment between the slab and the column as
v = -**- + Y X e ) A c v yx
V
y,
M QA M J y,yvMue^

(1.27)

15

where the subscripts x and y refer to centroidal principal axes of the shear critical sections; (x,y) are coordinates of the point at which vf is maximum; Ac is the area of critical section; J is defined as the polar moment of inertia of the critical section about its centroidal axis.

1.4 European Codes


Many European codes are based on the theory of Kinnunen and Nylander (1960). The punching shear resistance is calculated in most building codes by evaluating the nominal shear stress on a specified control surface around the loaded area or column, and comparing this with a concrete tensile strength. Provisions for punching shear capacity of concrete for four of the European codes are described below. In the German design code (DIN 1045 1988), the effect of moment transfer can be ignored if the spans of a panel do not differ by more than 33%. The application of DIN 1045 is described by Albrecht (2002) and fib Bulletin 12 (2001). The design model for the punching shear resistance without shear reinforcement considers a shear capacity that is dependent on the concrete cube strength and the flexural reinforcement ratio. The punching shear is determined as V c = 0.441 k 2 f Ci200 2/3 ud where k2 = 0.7 (1.28) 0.45Ju and |i g is the tension reinforcement ratio.

i + A.
500

Eurocode 2 (EC2 2002), the CEB-FIP Model Code 90 (CEB 1990) and the 1996 FIP-Recommendations (FIP 1996) consider the concrete cylinder strength,

16

fck, the flexural reinforcement ratio, pt, the size effect (k) as a function of the effective depth, d, and the shear capacity of the shear reinforcement, fsw Asw. EC2-02 calculates the resistance to punching shear for members without shear reinforcement at a critical section located at 2d from the face of the column as Vc = [0.18 k (100 pf c k ) 1 / 3 ]ud (1.29)

where fci< has several values of strength based on the characteristic strength, k = 1 + J < 2.0, p = ^p x p y < 0.02 and d =
x y

, being px, py, dx and d y the

reinforcement ratios and the effective depths, respectively, in the two orthogonal directions. These equations are compared with the North American approaches by Paultre and Mitchell (2003). The British Standard 8110 (BS 8110 1997) considers the same parameters as the EC-2 and the 1990 Model Code. BS 8110 assumes a shear crack inclination of 33, reducing the amount of shear reinforcement (fib Bulletin 12 2001). The punching shear at a critical section located at 1.5d from the face of the column is given as V c =[0.34k(100pf c k ) 1 / 3 ]ud where k = J other codes. Comparing the provisions for the punching shear for corner columns without shear reinforcement for leads to the conclusion that all have a wide range of results for all failure modes. This is due to the safety factors and the different evaluations of the punching shear capacities. Also, European codes consider the size effect and the effect of amount of longitudinal reinforcement, while the 2002 ACI Code (ACI 2002) and the CSA Standard (CSA 1994) do not account for these effects. (1.30)

and p < 0.03. Albrecht (2002) compares the BS 8110 code with

17

With respect to the flexural reinforcement ratio and the concrete cylinder strength, it is observed that the BS 8110-97 Code, the 1990 Model Code and the 1996 FIP-Recommendations give higher capacities than the ACI and CSA for large flexural reinforcement ratios (p>1%) and lower capacities for lower reinforcement ratios {fib Bulletin 12 2001 and Paultre and Mitchell 2003). For relatively high concrete strengths, the German and the North American codes give upper limits on the punching shear stress. The 2002 ACI Code neglects the influence of the flexural reinforcement ratio and therefore gives more conservative predictions for flexural reinforcement ratios greater than 1.5%. The design of reinforced concrete slabs for punching shear is of great importance, but the provisions for the design and analysis differ considerably among the various European and American design codes. Therefore, research and development of models, material and computer methods are needed in this important area.

1.5 Research Objectives


The objectives of this research program are:

1.

Design three full-scale slab-column specimens to investigate the shear capacity at corner columns in accordance with the 1994 CSA Standard A23.3. The main variable in this test series is the size of the corner

columns. 2. Design a testing apparatus capable of producing positive moments around the free edges of the slab specimens and negative moments around the columns. 3. Construct and test the three corner column-slab specimens.

18

Evaluate the performance of the specimens and recommend an appropriate design approach.

19

Chapter 2

Experimental Programme

2.1 Description of Prototype Structure

The prototype structure was designed in accordance with the Canadian Standards Association A23.3-94 (CSA 1994). The slab was designed for a

relatively high specified live load of 10 kPa and a superimposed dead load of 1.5 kPa. This resulted in a high percentage of reinforcement and a shear-critical slab design. The clear spans in both directions were 4.0 m and a total slab thickness of 150 mm was chosen. Analysis indicated that for an edge panel, the slab

moments are zero at a distance of 0.32 m from the face of the column with the maximum positive in the slab occurring at a distance of 1.82 m from the column face. In choosing the size of the test specimen it was decided to use a corner slab specimen that cantilevered out a distance of 1.150 m from the column faces (see Fig. 2.1).

2.2 Design of Test Specimens


Three full-scale specimens were constructed and tested to failure in the Jamieson Structures Laboratory in the Department of Civil Engineering at McGill University. The slab specimens were identified as Specimens S1, S2 and S3, having square corner columns with sizes, c of 250 mm, 300 mm and 350 mm, 20

respectively. The columns were extended above and below the slab a distance of 1.425m, giving a total column height of 3.0 m. A minimum specified concrete compressive strength of 30 MPa was used for the design of the slab and the column. The specified yield strength used for the reinforcement was 400 MPa. The clear cover on both top and bottom steel reinforcement in the slab was 20 mm, with a 30 mm clear cover used for the ties in the column. All the test specimens contained the same flexural steel in the slab and the same column reinforcement. Eleven 10M bottom reinforcing bars were used in both directions, providing a positive moment capacity of 158.4 kN for slab specimens. Four 15M top bars were used resulting in a flexural capacity of 123.8 kN for the specimens. Steel plates of dimensions 40 x 40 x 9 mm were welded at one end of the bars to ensure that the reinforcement was properly anchored. Additional top and bottom edge steel bars, 400 mm in length, were placed along the loading edges at the same level as the flexural reinforcement. These bars were used to avoid any local failures in the regions of the load application. Five 10M bars in each direction were located in such a way that 3 of them were in the zone of the loading plate and the two remaining covered the rest of the span. Steel plates were welded at one end. No shear reinforcement was provided in any of the slabs. The column reinforcement consisted of four vertical 15M bars and 15 sets of 10M ties at a spacing of 220 mm.

21

2.3 Details of Test Specimens

2.3.1 Specimen S1 The details of Specimen S1 are shown in Fig. 2.2. The column of The

Specimen S1 had cross-sectional dimensions of 250 mm by 250 mm. overall dimensions of the slab were 1400 by 1400 by 150 mm.

Two of the 11-10M bottom bars were anchored inside the column in order to satisfy the structural integrity requirements. Steel plates were welded at both ends. The length of these bars was 1380 mm to allow for a cover of 20 mm in the column. The remaining bars were 1300 mm long and they were equally distributed at a spacing of 125 mm. Both the dimensions and the detailing of the positive moment reinforcement are shown in Fig. 2.2. The 4-15M bars on top were fully anchored with welded plates inside the column allowing for a cover of 20 mm. Figure 2.3 shows the arrangement of the top layer of reinforcement. Figure 2.4 shows the details of the reinforcing steel in Specimen S1 just before casting the concrete.

2.3.2 Specimen S2 Specimen S2 had a 300 mm square column resulting in overall slab dimensions of 1450 mm x 1450 mm x 150 mm. As shown in Fig. 2.5, 2-10M bars were placed on the bottom of the slab into the column in both directions. The structural integrity reinforcement was the same as that used in Specimen S1. Nine additional 10M bottom bars were uniformly distributed in both directions. Figure 2.6 shows the 4-15M top bars for negative moment, with a spacing of 60mm within the column. The photograph in Fig. 2.7 shows a closeup of the reinforcement near the column before casting.

22

2.3.3 Specimen S3

Specimen S3 had a 350 by 350 mm column and a 1500 by 1500 by 150 mm slab. The 11-10M bottom bars were equally distributed at 125 mm spacing. Two of the 10M bars were 1480 mm long and continued into the column to satisfy the requirements for structural integrity. The 15M negative moment bars on top were all anchored into the column resulting in a spacing of 70 mm. Figures 2.8 and 2.9 show the arrangement of both the bottom and top layers of reinforcement, respectively. The detailed view of the joint in Specimen S3 is shown in Fig. 2.10.

2.4 Construction Sequence

The test specimens were constructed simultaneously in the Jamieson Structures Laboratory. The formwork was constructed with plywood. Figure 2.11 shows a photograph of the overall view of the three specimens before casting. The lower columns of the three specimens were cast first with the same concrete batch to provide rigidity to the system. Eight days later, concrete was cast in the three slabs. All slabs were constructed with the same batch of

concrete to ensure that all specimens possessed the same material properties. Six days later, the top columns were cast with the same mix design specified by the supplier but adding 1 liter of flow-mix to the concrete. All specimens were cured and then stored in the formwork until they were moved to the testing frame a few days before testing.

23

2.5 Material Properties

2.5.1 Reinforcing Steel

The reinforcement used in the construction of each of the specimens was in conformance with CSA Standard G30.18 (CSA 1992). All reinforcement was weldable grade, hot-rolled, deformed bars with a minimum specified yield stress of 400 MPa. Three sample coupons of each reinforcing bar size were tested in order to determine their stress-strain characteristics. The test samples were 300 mm in length and the extensometer used to measure strains was 150 mm long. Figure 2.12 shows typical stress-strain curves for these reinforcing bars and Table 2.1 summarizes the average values of the mechanical properties.

Table 2.1

Summary of reinforcing bar properties

BAR SIZE

AREA (mm2)

fy (MPa) (std. Deviation) 450.5 (13.9)

6y (mm/mm) (std. deviation) 0.00225


(...)

fuit (MPa)

(std. deviation) 567.6 (8.03) 591.0 (1.94)

10M

100

15M

200

469.7 (5.38)

0.00235
(...)

2.5.2 Concrete The columns and slabs were cast with normal density concrete. All

batches had a specified 28-day compressive strength of 30 MPa. The maximum aggregate size used was 20 mm, and the components and proportions of the concrete mix design, as specified by the supplier, are presented in Table 2.2. 24

After casting, the specimens were covered with wet burlap and plastic to prevent water loss and maintained at standard room temperatures. The slab specimens were moist-cured for a period of 7 days. After 24 hours, specimens used to determine the mechanical properties were demolded and stored in a humid room.

Table 2.2

Concrete mix design

COMPONENT TypelO cement (kg/m3) Fine aggregate (kg/m3) Coarse aggregate (20mm) (kg/m13) Coarse aggregate (14mm) (kg/mJ) Water (kg/m3) Water-cement ratio Superplasticizer (l/m3) Retarding agent (l/mJ) Air-entraining agent (l/nr5) Water-reducing agent (l/mJ) Slump (mm) Air content (%)

COLUMN AND SLABS 340 784 369 693 160 0.47 2.5 0.32 0.19 1.06 80 6.5

For each mix, a series of tests were conducted on the plastic and hardened concrete. First, volumetric air content and slump were determined.

Then six cylinders of 150 mm diameter and 300 mm length were cast for each specimen and for each batch of concrete to test the physical properties of the corresponding hardened concrete. Three cylinders were used to test the

compressive strength, f'c, after 28-days of curing in a humid room. Typical 2825

day compressive stress-strain relationships for the concrete used in the slabs are shown in Fig. 2.13. Three cylinders per specimen with the same dimensions were cast at the same time in order to determine the splitting tensile strength of the concrete, fsp. The modulus of rupture, fr, was determined from flexural beams with dimensions of 100 x 100 x 400 mm. These specimens were subjected to four-point loading. The results obtained are summarized in Table 2.3.

Table 2.3

Concrete properties

SPECIMEN

f c (MPa) (std. dev.) 35.4 (1.36)

fr (MPa) (std. dev.) 4.26 (0.33) 6.41 (0.12) 4.26 (0.64)

fsp (MPa) (std. dev.) 3.2 (0.18) 3.5 (0.20) 3.4 (0.08)

Lower column

Slab

42.8 (0.90)

Upper column

33.6 (0.47)

A standard specimen of dimensions 80 mm by 80 mm by 280 mm was used to take shrinkage measurements for each batch of concrete cast. Measuring studs were placed at each end of the standard specimen to take readings and determine the shrinkage. The specimen was cured in the humid room to simulate the conditions at the interior of the full-scale specimens. The variation of shrinkage in the specimens in time is shown in Fig. 2.14.

26

2.6 Testing Procedure

The formwork of each specimen was removed a few days before testing to hold them in a stable position and to provide stiffness to the whole system. Each specimen was moved with a crane fixed to the specimen at the center of gravity. The slabs were moved inside a steel testing frame used to laterally restrain the column. After placing in the test frame, strain gauges were calibrated until the gage indicated zero strain.

2.6.1 Test Setup and Loading Apparatus

At the moment of testing, the ages of Specimens S1, S2 and S3 were 54, 69 and 83 days, respectively. The base of the column was placed on a neoprene rubber pad and the top of the column was bolted to the steel frame using threaded rods and angle sections on two faces of the column. From the analysis of the prototype structure, the points of inflexion were located at a distance of 320 mm from the face of the column. To simulate the loading effects that the slab would experience two loads were applied as indicated in Fig. 2.15. Each load was applied through a distribution beam

attached below the slab. For this purpose, a square hollow structural section HSS 152 x 152 x 11 mm was used to transmit the load to the test specimen. Two steel plates with dimensions of 150 by 450 mm and 19.04 mm-thick were placed at the top and the bottom of the slab, 25 mm from the edge of the slab and centered along the face of the column (see Fig. 2.15). Eight holes were drilled for 8 bolts connecting the slab, a spreader beam and the loading beam and causing a positive bending moment along the free edges of the slab. The loading beams were centrally located under the spreader beam in such a way

27

that the resultant forces of P/2 were located along the interior column faces at a distance of 500 mm (see Fig. 2.15). Monotonic static load was applied downward at the two loading points and the load was measured by two load cells that reacted against the hydraulic jacks under the reaction floor. The jacks were connected to a common hydraulic pump such that the loads were equal at the two loading points. Incremental forces were applied in steps until failure was achieved. The magnitude of the

increments was reduced at the higher load levels near failure. Figure 2.16 shows a photograph of the test setup.

2.6.2 Instrumentation

During testing, loads and deflections were recorded and strains were monitored at various locations along the flexural reinforcement to provide detailed data of the behaviour of the specimens. The extension and width of the cracks was measured at each load stage and they were marked by a pen for better visibility. Two load cells were used to measure the applied downward forces on the slab at each loading jack during testing. Vertical displacements were measured at several points on the slab and the column using linear voltage differential transducers (LVDTs). The LVDTs were clamped to a light steel frame attached to the column. LVDTs 1 through 4, having a displacement range of 50 mm, measured the displacements of the slab relative to the column (see Fig. 2.16). The position of the LVDTs is shown in Fig. 2.17. Three LVDTs with a range of 15 mm (LVDTs 5, 6 and 7 in Fig. 2.17) were clamped to the column and measured the movement of the bottom surface of the slab relative to the column at the slab-column interface. Two of these LVDTs were attached to the column at the centre of both interior column faces,

28

100 mm below the slab. Another LVDT was attached to the column and located at the corner of the column as shown in Fig. 2.17. Electrical resistance strain gauges were attached to the flexural

reinforcement to measure the strain in the bars. Strain gauges with a 5 mm gauge length were installed during construction of the specimen on all of the negative moment reinforcing bars and on two of the four structural integrity bars. The strain gauges were located at the column interface in both directions of the slab. The positioning of the strain gauges is shown in Fig. 2.18. A computerized data acquisition system was used to record loads, displacements and strains at frequent intervals during loading.

29

4000

IT
test specimen

Figure 2.1 Plan of prototype flat plate structure.

30

275 +*

1125
a) Plan view -1150-

100

150

tvi

________

130 20

9-10M bars, L=1300 b) Section 1-1 20 (230 1150

130 20"

- 2-1OM bars, L=1380 c) Section 2-2 rcement details for bottom steel in S1 31

25020 250
> i

1150

50

20

150 50 4-15M@50(typ.)

1150

a) Plan view

20

230

1150

20 130

4-15M bars, L=1380 b) Section 1-1

Figure 2.3 Top reinforcement in Specimen S1

32

Figure 2.4 Detailed view of the joint in Specimen S1

33

325 - * k -

1125 a) Plan view

150 150 *

1150

130 20"

<

9-10M bars, L=1300

b) Section 1-1

2-10M bars, L=1430 c) Section 2-2


n

~**~m reinforcement details for Specimen S2. 34

a) Plan view

20

280 H*-

1150
20

-v-

4-15M bars, L=1430 b) Section 1-1

Figure 2.6 Arrangement on top steel for Specimen S2.

35

Figure 2.7 Detailed view of the joint in Specimen S2.

36

1150
t ~20 350
i i
j

200

1500 1150 1500

1500
a) Plan view -200 150< -A-

1150

' '"*

"ir

y--~nr | jBa

130

20

9-10M bars, L=1300 b) Section 1-1

1150

>

130

2cT

2-10M bars, L=1480 c) Section 2-2 m reinforcement details for Specimen S3. 37

350 -* 20 t ~20 350

1150

70
;

210

70 4-15M@70(typ.)

1150

a) Plan view

20

330 -*\*
-A-

1150

20
-p eoo

1^fT

A-

4-15M bars, L=1480 b) Section 1-1

Figure 2.9 Top steel for Specimen S3.

38

Figure 2.10 Detailed view of the joint in Specimen S3.

Figure 2.11 Overall view of the specimens before casting.

39

700 15M bars

ro a.

400300^

t
gauge length

jjj to

200 4 1000
1 1 1 1 1

I
1

0.02

0.04

0.06

0.08

0.1

0.12

0.14

0.16

0.18

Strain (mm/ mm)

Figure 2.12 Stress-strain curves for reinforcing bars.

50

40 -

30 ro
Q.

J.
20

in in

a>
CO

T
10 Slab Bottom column Upper column
i-

0.001

0.002

0.003 Strain (mm/mm)

0.004

0.005

0.006

Figure 2.13 Compressive stress-strain responses for concrete.

40

0.6

0.5 -

/>: v .
E E E E 0.4 -!

Slab

0.3 4

ro CO

100 Time (days)

120

140

160

180

200

Figure 2.14 Shrinkage strain curve.

41

1150
A-

150-25

i!:l
A-

HSS 152X152X11

I P/2
500 < 475 * 175

a) Elevation view
c

1150
X
o o o o o o

P/2

o o

450

150
' P/2

25

3@90
40 25 < 70
o o o o

I I

"W 450

b) Plan view
Steel plate (150x450) Slab

Steel plate (150X100) Steel plate (150X150)

HSS 152X152X11 L=450 HSS152X152X11 L=725

c) Detailed view of steel sections Figure 2.15 Test setup for specimens. 42

Figure 2.16 Overall view of loading setup.

43

500

550

*-

100

A-

-rAv,
100

5 -V-

I P/2
a) Elevation view

500

550

100

[]

7
o o

-B

3
500

550

o - oA
o

o
o

100

o1

b) Plan view Figure 2.17 Location of LVDTs.

44

North edge

C D U) T3 <D to

^L.

X.

a) Top reinforcement

b) Bottom reinforcement

Figure 2.18 Position of strain gauges.

45

Chapter 3

Experimental Results

This chapter presents the results and observations from the tests performed on the three slab specimens. The data recorded for each specimen included the applied loads, reinforcing steel strains, deflections, and the crack development and crack widths. The examination of the behaviour of each specimen includes the description of the overall behaviour using the load versus deflection response. The loads reported in this chapter are the sum of the two point loads applied to the slab, including the self-weight of the slab and the loading devices, as explained in Chapter 2. The corresponding self-weight plus the weight of the loading devices for Specimens S1, S2 and S3 was 9.6, 10.1 and 10.6 kN respectively. The deflections are the average deflections that were measured in two symmetrical points as shown in Fig. 2.17. During the tests, the loads were increased in small increments until failure was reached. The faces of the test specimens are referenced using specific

identification. The flexural cracking on the north face is controlled by the lower steel of the top mat ("flexural weak direction") in Specimen S1. Hence, the west face corresponds to the cracking in the "flexural strong direction" for this specimen. The north face of Specimen S2 and S3 corresponds to cracks

observed in the "flexural strong direction", while the west face corresponds to cracks observed in the "flexural weak direction". A general description of the progression of the cracks with the total load is presented as well as the maximum crack widths recorded at each load stage. 46

The development of the strains in the top reinforcing bars and in two of the structural integrity reinforcement at the face of the column is discussed.

3.1 Specimen S1
Figure 3.1 shows the load versus deflection responses for Specimen S1. Different load-deflection responses are given for the average of the measured deflections at points 1 and 4 and for the average deflections at points 2 and 3. The first flexural cracks appeared at a total load of 21.3 kN. These cracks appeared on the top surface of the slab close to the inner corner of the column and propagated towards both free edges of the slab. These cracks were visible on the north and west faces at locations of 75 mm and 25 mm, respectively, from the faces of the column. Figure 3.1 shows a significant drop in stiffness starting at a load of 15.5 kN, due to cracking. One positive moment crack appeared at 240 mm from the east edge of the slab near the north face when the applied load was 51.2 kN. At a load of 64.1 kN another positive moment crack developed at a location of 155 mm from the south edge of the slab. With increasing load, new flexural cracks occurred on the top surface of the slab as shown in Fig. 3.3. When the total load was 75.5 kN, the maximum width of the cracks on the exposed north and west edges of the slab were 0.30 mm for flexural cracks and 0.25 mm for the torsional cracks. The strains in the reinforcing bars were measured with strain gauges located in line with the faces of the column. Typical results of the strain in the steel showed that the reinforcing bar located in the outermost edge near the west face (gauge #1) was the first one to reach the yield strain at a total load of 84.4 kN as is illustrated in Fig. 3.2. When the load was increased to 91.7 kN, the steel

47

located closer to the inner corner of the column near the north face, reached the yield strain while the remaining bars were below yield. At 105.9 kN, crushing of the concrete occurred on the bottom surface of the slab near the inner corner of the column. In the next load stage, at 111.8 kN, the first shear cracks appeared on the north face of the slab at a distance of 285 mm from the face of the column, as shown in Fig. 3.4. As apparent from Fig. 3.4, the angle of inclination of the shear crack was 45 degrees from the plane of the slab. The width of the major shear crack was initially 0.10 mm. Splitting of the concrete due to high local in compressive stresses was observed at the inside corner of the column immediately below the slab. When the first shear crack appeared on the north face, all the reinforcing bars in that direction had already reached their yield strain. When the load was increased to 117.6 kN, the first shear crack appeared on the west face but for this case, the two outer bars in the "flexural strong direction" (gauges #1 and #4) had yielded, while the two inner bars (gauges #2 and #3) were just below their yield strain (see Fig. 3.2). As failure approached, the top surface cracks widened and at a load of 121.4 kN, a brittle shear failure occurred (see Fig. 3.1). Figure 3.5 shows the maximum crack width recorded at each load stage and Fig. 3.6, 3.7 and 3.8 show photographs of the north and west faces of the slab at failure and the crack pattern on the top surface. From Fig. 3.6 and 3.7 is observed that the failure shear crack on the north face spread to both the top and bottom surfaces of the slab. On the west face, the shear crack failure propagated towards the bottom face of the slab and reached the bottom surface at a distance of 230 mm from the column face. Figure 3.2 shows that yielding was reached in all but one of the top bars in the "flexural strong direction" (at gauge #3) before failure. From Fig. 3.2, it is also observed that the strains were higher in the reinforcement located near the north face ("flexural weak direction").

48

Figure 3.1 shows that the vertical deflection of the slab relative to the column at the edge of the slab (points 2 and 3), was twice the deflection measured at the points where the load was applied (points 1 and 4). relationship remained relatively constant throughout the test. This

The average

deflections recorded were 17.9 mm and 34.9 mm at the peak load of 121.4 kN. Figure 3.9 gives an overall view of the slab-column specimen after shear failure had occurred.

3.2 Specimen S2

Specimen S2 behaved in a manner similar to that of slab S1. Figure 3.10 shows the load-average deflection curves, with the average deflections calculated from the individual deflection readings for the four symmetrical points in the slab. A gradual reduction in stiffness is observed in Fig. 3.10 after a load of 15.3 kN was reached. One crack formed at a load of 18.2 kN on the top surface of the slab and originated in the inner corner of the column. This crack propagated toward the edge of the west face of the specimen, surfacing on the side face at a distance of 68 mm from the face of the column, and extended down the slab edge a distance of 20 mm. The first positive moment crack occurred on the west face at a load of 66.1 kN and was located 220 mm from the south face of the slab. When the load was increased to 75.2 kN, an additional positive moment crack appeared near the north face of the slab at a distance of 232 mm from the east face. With increased loading, the cracks located in the immediate vicinity of the inner corner of the column increased in number and extended to the free edges of the slab as shown in Fig. 3.12. On the north and west faces of the slab, the flexural cracks became inclined indicating flexural-torsion cracks. At this stage, when the applied load was 75.2 kN, the maximum width of the flexural cracks 49

was 0.75 mm, and the maximum torsional crack was 0.35 mm as it is observed in Fig. 3.14. Figure 3.11 shows the strain distribution of the top reinforcement at the face of the column for Specimen S2. The two bars located at the extremities of the west face (gauges #1 and #4 in the "flexural weak direction") reached their yield strain simultaneously at a load of 82.8 kN. When the load was increased to 98.1 kN, the bar closest to the inner corner of the column (gauge #4 in the "flexural strong direction") yielded while the remaining bars in the same direction were below yield. At this load, all of the bars located in the other direction had yielded. For Specimen S2, the first shear crack started at a load of 113 kN, on the bottom face of the slab at a distance of 350 mm from the face of the column, with the crack becoming flatter as it progressed upwards into the slab (see Fig. 3.13). The width of this shear crack was 0.5 mm, while the maximum width for the flexural cracks was 2.25 mm, as is shown in Fig. 3.14. As for Specimen S1, crushing of the concrete was observed at the inner corner of the column, just below the slab. When the load was maintained at 116.7 kN for 1 or 2 minutes, a punching shear failure occurred. The failure shear crack on the north face extended horizontally parallel to the plane of the bottom bars towards the column (see Fig. 3.15). On the west face of the slab, failure took place with the initial shear crack extending towards the column (see Fig. 3.16). Yielding of the top reinforcing bars was recorded in all the bars before failure was reached, as is shown in Fig. 3.11. None of the bottom reinforcing bars yielded. During testing, the average deflections from points 1 and 4 were twice the average deflections from points 2 and 3. (see Fig. 3.10). The maximum average deflection recorded was 18.9 mm and 37.6 mm at a load of 105.6 kN after shear

50

failure occurred. These deflections were greater than those for S1, due to the higher strains reached in the flexural reinforcement. Figures 3.15, 3.16 and 3.17 show photographs of the slab at failure. The overall view of the deformed shape of Specimen S2 is shown in Fig. 3.18.

3.3 Specimen S3

The behaviour of Specimen S3 was different than the other two specimens due to the failure mode. The total load versus the average deflection response of Specimen S3 is shown in Fig. 3.19. The first change in stiffness occurred when the total load was 17.7 kN. A gradual decrease of stiffness was observed when the load was increased to 31.9 kN (see Fig. 3.19). The first flexural crack occurred on the top surface of the slab when the total load was 18.8 kN and it was visible at locations 110 and 176 mm from the face of the column on the north and west faces of the slab, respectively. The load was increased gradually and at 74.5 kN the first positive moment crack appeared near the north face (that is in the "flexural strong direction") at a location of 220 mm from the east face of the slab. At this stage, the maximum crack widths were 0.10 mm for the torsional cracks and 0.65 mm for the wider flexural cracks, as is shown in Fig. 3.23. There were no shear cracks at this load level. The crack pattern of the slab when the total load was 67.1 kN is shown in Fig. 3.21. Figure 3.20 shows the measured strains in the reinforcement bars of Specimen S3 at the same key load stages as for Specimens S1 and S2. Strain gauge readings showed that the bars closer to the inner corner of the column (gauge #4 locations) were the first to reach the yield strain at total loads of 74 and 93.2 kN in the north and west faces, respectively.

51

The first sign of a shear crack, with a width of 0.10 mm, appeared at a total load of 120.9 kN on the west face at a location 395 mm from the face of the column (see Fig. 3.22). Figure 3.20 shows that when the first shear crack

appeared, all the reinforcing bars had yielded in flexure. At a total load of 120.9 kN, the maximum crack widths were 0.10 mm for the torsional cracks and 2.0 mm for the flexural cracks on the top surface (see Fig. 3.23). Close to failure, the shear cracks on the north face tended to become horizontal near the bottom face of the slab near the column while on the west face of the slab the angle of inclination with respect to the slab was larger. The maximum load and the corresponding average deflections registered during testing were 124.9 kN and 30.5 mm for the points 1 and 4 and 15.3 mm for points 2 and 3, as is shown in Fig. 3.19. As for the other two specimens, Fig. 3.19 shows that the average deflections in the edges of the slab (points 1 and 4) were about twice those at the points where the loads were applied (points 2 and 3). The magnitude of the measured strains increased appreciably in the bars oriented in the north face, especially the bar closest to the inner corner of the column (gauge #4). Crushing was also observed at the inner corner of the lower column, and on the bottom surface of the slab, which demonstrated that failure by flexure had occurred. The primary mode of failure was by flexure followed by extensive

shear cracks in the slab that could not extend, probably due to the presence of the plates of the loading apparatus (see Fig. 3.25). Photographs of the specimen at failure are shown in Fig. 3.24, 3.25 and 3.26.

52

140
Load vs. deflection (av. of points 2 & 3)

100

Load vs. deflection (av. of points 1 & 4)

60
DE

20 -

10

15 20 25 Average deflection (mm)

30

35

40

Figure 3.1 Load versus average deflection of Specimen S1.

Ey

0.00235

Strain 0.004 (mm/mm)

First flex, crack A First yielding First shear crack Failure Strain (mm/mm)

A/-

Figure 3.2 Strain distribution of top reinforcement in Specimen S1 at key stages.

53

Figure 3.3 Flexural cracking on top surface of Specimen S1 at a shear of 75.5 kN.

Figure 3.4 Flexural shear cracks at a shear of 111.8 kN in Specimen S1, north face.

54

IHU "

100 -

<
T3

1 1

ro o 60 Flexural crack Shear crack Torsion crack 20 1

-1

11

0.2

0.4

0.6

0.8

Maximum crack width (mm)

Figure 3.5 Maximum crack width for Specimen S1.

55

Figure 3.6 Shear failure in Specimen S1, north face.

1 IIP?
^HH&RISB ifB

S1 s

^f

^^H

9 I: y

Figure 3.7 Shear failure in Specimen S1, west face.

56

Figure 3.8 Top surface after shear failure in Specimen S1

Figure 3.9 Overall view of shear failure in north face in Specimen S1 at shear of 125.6 kN.

57

itu -

Load vs. deflection (av. of points 2 & 3)

100 -

/ ^

Load vs. deflection (av. of points 1 & 4)

Seo
o

3
o2

20 0 1
' 1 1 1 1 1

10

15

20 25 30 Average deflection (mm)

35

40

Figure 3.10 Load versus average deflection of Specimen S2.

Ey

0.00235

Strain 0.004 (mm/mm)

First flex, crack A First yielding First shear crack Failure Strain (mm/mm)

Figure 3.11 Strain distribution of top reinforcement in Specimen S2 at key load stages. 58

Figure 3.12 Flexural cracking on top surface of Specimen S2 at a shear of 75.2 kN.

Figure 3.13 Flexural shear cracks at a shear of 113 kN in Specimen S2, west face.

59

140

100 -

i
ro

60 /y Flexural crack Shear crack Torsion crack

_l

20

0.4

0.8

1.2

1.6

Maximum crack width (mm)

Figure 3.14 Maximum crack width for Specimen S2.

60

Figure 3.15 Shear failure in Specimen S2, north face.

Figure 3.16 Shear failure in Specimen S2, west face.

61

Figure 3.17 Top surface after shear failure in Specimen S2.

Figure 3.18 Overall view of shear failure in north face in Specimen S2 at shear of 116.7 kN. 62

140 -

100

Load vs. deflection \ (av. of points 1 & 4)


ro 60

A A 1 J

Load vs. deflection \ (av. of points 2 & 3)


D2

20
0 1
1

i'

10

20

30 40 50 Average deflection (mm)

60

70

Figure 3.19 Load versus average deflection of Specimen S3.


Strain 0.010 (mm/mm)

0.00235

0.006

Ey

0.00235 <>

0.006

First flex, crack A First yielding First shear crack Peak load

0.010 Strain (mm/mm)

Figure 3.20 Strain distribution of top reinforcement in Specimen S3 at key stages. 63

Figure 3.21 Flexural cracking on top surface of Specimen S3 at a shear of 67.1 kN.

Figure 3.22 Flexural shear cracks at a shear of 120.9 kN in Specimen S3, west face.

64

140

100 -

ro

60 Flexural crack Shear crack Torsion crack 20

-i

1-

-i

1-

0.4

0.8

1.2

1.6

Maximum crack width (mm)

Figure 3.23 Maximum crack width for Specimen S3.

65

Figure 3.24 Flexure-shear failure in Specimen S3, north face.

Figure 3.25 Flexure-shear failure in Specimen S3, west face.

66

Figure 3.26 Top surface after flexure-shear failure in Specimen S3.

67

Chapter 4

Analyses and Comparison of Results


In this chapter, the responses of the three specimens are compared with each other and with the predictions obtained using the equations of the Canadian Standard (CSA 1994). These comparisons provide information on the influence of the size of the column and the length of the critical perimeter on the shear strength. The objective of this chapter is to determine if the provisions of the Canadian Standard (CSA 1994) provide conservative shear strength predictions for slab-corner column connections.

4.1 Comparison of Experimental Results


Load-deflection curves for the Specimens S1, S2 and S3, having the same flexural reinforcement are shown in Fig. 4.1 and 4.2. Figures 4.1 and 4.2 illustrate that a punching shear failure in Specimens S1 and S2 resulted in a sudden decrease of the carrying load after the peak loads had been reached at peak loads of 121.4 and 116.7 kN, respectively. For Specimen S3, the maximum load reached was 124.9 kN corresponding to flexural yielding. The post-peak response of Specimen S3 was characterized by a smooth decrease of the carrying load with increasing displacement. Specimen S3 had a maximum load that was 3 to 7% higher than the maximum loads for Specimens S1 and S2.

68

The total loads recorded during testing with the corresponding average deflections at different key stages, such as first flexural cracking, first yielding, first shear crack and peak load, for the various specimens are summarized in Table 4.1.

Table 4.1

Summary of key load stages for Specimens S1, S2 and S3.

FIRST SPECIMEN FLEX. CRACK S1 Load (kN) Deflection (mm) S2 Load (kN) Deflection (mm) S3 Load (kN) Deflection (mm) 21.3 1.2 18.2 0.6 18.8 0.6

FIRST YIELD

FIRST SHEAR CRACK

PEAK LOAD

84.4 15.7 82.8 13.9 74 11.4

111.8 26.8 113 26.4 120.9 27.9

121.4 34.9 116.7 31.3 124.9 30.5

From Table 4.1, it is observed that Specimen S2 exhibited the smallest load at first flexural cracking. The strain gauges in the top reinforcement of the slab showed that the steel bars in Specimen S3 reached the yield strain at a slightly lower load than for Specimens S1 and S2. Table 4.1 also shows that in Specimens S1 and S2, the failure shear crack appeared at only 92 and 97% of the ultimate load, respectively. For Specimen S3, which exhibited a different failure mode, the first shear crack appeared at 97% of the ultimate load. In both Specimens S1 and S2, the angle of inclination between the failure shear crack and the plane of the slab was approximately 30 69

degrees. In Specimen S3 significant shear cracks with an angle of 45 degrees developed before flexural yielding occurred. Menetrey (1996) presented an

analytical method to predict the punching shear capacity, which is influenced by the inclination of the shear cracks. He concluded that inclined cracks with a low angle indicate punching shear failure, while inclined cracks with larger angles indicate primarily flexural yielding. Specimen S2, with a larger column than Specimen S1, experienced smaller deflections during testing due to the smaller flexural lever arm and the increased stiffness of the slab-column connection. However, the maximum

deflection attained at failure for Specimen S1 was 2% higher than the deflection at peak load (i.e., increasing from 34.9 to 35.5 mm) while the maximum deflection for Specimen S2 increased from 31.3 mm at the peak load to a value of 37.6 mm at failure (i.e., an increase of 20%). In contrast to these two

specimens, Specimen S3 reached a maximum deflection of 60.2 mm, that is almost double the deflection at peak load (see Fig. 4.1). It is evident that

Specimen S3 had considerably greater ductility than the other two specimens. It is surprising that Specimen S2, having a larger column than Specimen S1, experienced a peak load that was slightly less than Specimen S1. This may be due to the very brittle nature of punching shear failures. In all the specimens, yielding of all top reinforcing bars was recorded before failure occurred. Specimen S3 underwent general yielding of the negative moment reinforcing bars before the ultimate capacity was reached. Zaghlool and de Paiva (1973) found that punching shear is a secondary phenomenon that develops only after yielding of the steel at the slab-column interface. They

predicted that the ultimate capacity depends primarily on p fy and the column size, among others, and not on y[f\ . From the results presented in Chapter 3, it is observed that the highest strains occurred for the three specimens in the bar located closest to the inner

70

corner of the column in the "flexural weak direction". This seems to show that the greatest stresses were concentrated at the inner corner of the column. Representative crack patterns for the three specimens at failure loads are shown in Fig. 3.8, 3.17 and 3.26. Figures 3.8 and 3.17 in Chapter 3 show that there was not much difference in the crack pattern and mode of failure for Specimens S1 and S2. For all specimens, the flexural cracks first appeared in the immediate vicinity of the inner corner of the column. difference was the width of the failure cracks. The increased column size of Specimen S3 had a favorable influence on the failure mode, shifting from punching shear to flexural yielding. This The only major

observation is evident from the crack pattern of the three specimens in Fig. 3.26 and in Table 4.2. Table 4.2 shows the maximum crack widths at loads corresponding to general yielding of the flexural steel.

Table 4.2

Maximum crack widths at peak loads for Specimens S1, S2 and S3.

MAXIMUM CRACK WIDTH AT LOAD AT SPEC. GENERAL YIELDING (% OF PEAK LOAD) S1 121.4 (100) S2 116.7 (100) S3 112.2 (90) 1.75 0.1 2.25 0.5 2.0 0.35 0.3 2.0 Flexure + Punching Flexure + Punching Flexure GENERAL YIELDING (mm) Flexural crack Torsional crack Shear crack MODE OF FAILURE

71

As can be seen from Table 4.2, significant shear cracks were present in Specimens S1 and S2 at the load stage corresponding general yielding of the flexural steel. Both of these specimens failed by combined flexure and punching shear. Specimen S3, with the largest column, exhibited flexural yielding and

large deflections and had no shear crack when general yielding occurred.

4.2 Comparison of Experimental Results with Code Predictions

According to the 1994 CSA Standard it is necessary to check the design for both one-way shear and two-way shear action for corner slab-column connections. For both types of failure, either one-way shear or shear combined with moment transfer, the strength is strongly influenced by the size of the critical section or periphery. For one-way shear, the critical section of a square corner column is assumed to be located along a straight line at a distance of d/2 from the inner corner of the column. Ingvarsson (1973) analyzed shear failures at corner

columns on the basis of the slab action being similar to that of a diagonal beam. This critical one-way shear periphery was also suggested by Hawkins and Mitchell (1979). The 1994 CSA Standard requirements for one-way shear at corner columns was based on the work of Hawkins and Mitchell. This results in a length, bo, of the critical section for a square column of size c of (d + V2 c). In two-way shear action the 1994 CSA Standard assumes that the critical section is located at a distance d/2 from the perimeter of the column, resulting in a critical perimeter, b0, of 2 (c + d/2). The one-way shear strength is determined from the following equation:

72

Vc =0.166A(l)cA/fc b 0 d

(4.1)

where X is a factor accounting for concrete density and the two-way shear strength is calculated using Eq. (1.24) in Chapter 1. The critical perimeters for both one-way and two-way action are shown in Table 4.3. Table 4.3 also compares the strengths of the slab-column

connections observed in the tests with the values predicted by the 1994 CSA Standard using Eq. (4.1) and (1.24) for both one-way and two-way, respectively. In this comparison, a value of (j)c of 1.0 was used in the analysis of the test results, allowing the nominal shear resistance, Vc, to be determined.

Table 4.3

Predicted failure loads for Specimens S1, S2 and S3 according to 1994 CSA Standard.

One-way shear across corner at d/2 Spec. c (mm) h (mm)


f'c
dave

One-way shear at d/2 from column faces

Shear and moment transfer at d/2

(MPa)

(mm)

bo (mm) 822.1 963.5 1104.9

Vc (kN) 102.7 120.3 138

bo (mm) 615 715 815

Vc (kN) 77.1 89.6 102.2

bo (mm) 615 715 815

Vc (kN) 63.8 80.9 99.1

Vu (kN) 121.4 116.7 124.9

S1 S2 S3

250 300 350

150 150 150

42.8 42.8 42.8

115 115 115

Table 4.3 shows that one-way shear resistance at d/2 from the inner corner was found to give the best prediction for all three test specimens. The ultimate strengths of the three specimens are also compared to the values obtained from the 1994 Standard equations and are plotted against the column size in Fig. 4.3.

73

As can be seen from Table 4.3 and Fig. 4.3, the one-way shear predictions with a diagonal critical section located at d/2 from the inner corner provide very good agreement with the experimental results for Specimen S1, with the experimental shear capacity being 18% above the predicted one-way shear capacity. The experimental shear strength for Specimen S2 was 3% below the value predicted by the 1994 CSA Standard for one-way shear. Specimen S3 experienced flexural yielding and hence the predicted value for one-way shear in Table 4.3 is not applicable. The total shear corresponding to the negative and positive flexural nominal resistances provided in the weak flexural direction are 139.3 and 183.9 kN, respectively, for Specimen S3. In this case, the negative moment capacity governs and is used in the analysis. The maximum moment reached in the slab at the column face is 10% below the predicted nominal flexural resistance. It is believed that torsion in the slab resulted in additional tensions in the reinforcement on the side faces of the column, thus resulting in a lower moment resistance. It is noted that the prediction for one-way shear on a critical section at d/2 from the column faces provides a conservative prediction of the failure load. This assumed failure mode provides a simple approach, using the same critical section as for two-way shear, and provides a simpler method than the current CSA method for cases with rectangular columns and rectangular slab panels. Another advantage of using this critical section is that all of the current code requirements for proximity to edge of slab and for presence of openings would also apply. Both Table 4.3 and Fig. 4.3 also show that the predicted strength for each specimen in two-way action is conservative. It is noted that the ACI Code does not consider the one-way shear failure at corner columns. Table 4.4 provides a summary of the shear strength for the three specimens as predicted by EC2-02. It is noted that EC2-02 takes account of the effect of the flexural reinforcement ratio on the shear resistance. Figure 4.4 also

74

compares the experimental results with the failure loads predicted by the 1994 CSA Standard and EC2-02.

Table 4.4

Predicted shear strength for the three specimens with EC2-02.

SPEC.

c (mm)

f'c

Clave

P 0.012 0.011 0.010

bo(u) (mm)

Vc (kN) 93.3 98.2 102.3

Vu (kN) 125.9 122.7 133.1

(MPa) 42.8 42.8 42.8

(mm) 115 115 115 2 2 2

S1 S2 S3

250 300 350

611 661 706

The critical perimeter in EC2-02 is considered at 2d from the face of the column to make the limiting shear stress more uniform for different column sizes. This feature is apparent from Fig. 4.4. The expression used in EC2-02 to predict the shear strength is a function of the cube root of the reinforcement ratio and the cube root of the compressive strength of the concrete and also considers the size effect as is seen in Eq. 1.29 in Chapter 1. It can be observed in Table 4.4 and Fig. 4.4 that the shear strength predicted by the 1994 CSA Standard and by EC2-02 show significant differences and that the equations predicted by EC2-02 have smaller variation than the 1994 CSA Standard for different column dimensions. Table 4.4 shows that EC2-02 conservatively predicts values of the shear strength with the predictions being relatively independent of the column size. From Fig. 4.4 is observed that for small column sizes, up to 200 mm, EC202 gives a higher prediction than the one-way shear prediction of the 1994 CSA Standard. For column sizes between 200 and 350 mm, EC2-02 predicts values 75

that are lower than the ones predicted by the 1994 CSA Standard for one-way shear strength. When the size of the column is 350 mm and more, the critical perimeter assumed in EC2-02 is constant because is taken as the lesser of 1.5 d and 0.5 c. For column sizes greater than 350, the 1.5 d requirement controls. Hence, EC2-02 predicts a maximum shear strength for a column size of 350 mm, with the predicted strength being reduced slightly for column sizes above 350 mm, because the flexural reinforcement ratio is reduced. It is also observed that for larger flexural steel ratios such as Specimen S1, a brittle shear failure occurred accompanied by yielding of the tension steel reinforcement. When the flexural reinforcement ratio was decreased, a ductile flexural failure occurred accompanied by general yielding of the negative moment steel. However, Alexander and Simmonds (1992) found that the shear strength was relatively independent of the steel index p fy when punching shear failure occurs.

76

140

100

g 60

20

30

40

50

Average deflection (mm)

Figure 4.1 Load versus average deflection.

140

toffy
100

\S 3

r S2^ ^ k si

8 60

20

i
' 1 ' 1 1 '

3
c

,2

t '

10

20

30

40

50

60

70

Average deflection (mm)

Figure 4.2 Load versus average deflection. 77

250

200 -

T
c
-^ W
Q)
i

150 S1 .
10C

S3 S2

S
JC

H
One way across corner One-way at d/2 from column faces Two-way Test Results

CO

50.

100

200

300 Column Size (mm)

400

500

600

Figure 4.3 Variation of strength predicted by the 1994 CSA Standard with the column size.

78

250
S '
S

s '
S

200 ,

s
s

^
c
CD

150 -

y
S1 S2 ,-'
S3.'' y'
s
~^*^

^-^"^

co

ro 100 a)
CO

-'"

y^
- - One way across corner, CSA One-way at d/2 from column faces, CSA --- Two-way, CSA EC2-02 Test Results

50 -

'
" " ^ **

^^^
* * *

..''

.'''
1 1

1-

100

200

300 Column Size (mm)

400

500

600

Figure 4.4 Variation of strength predicted by the 1994 CSA Standard and EC2-02 with the column size.

79

Chapter 5

Conclusions

5.1 Conclusions

The results of testing three full-scale slab-column specimens are used to evaluate the shear capacity at corner columns and to compare the experimental results with predictions using the requirements of the 1994 CSA Standard A23.3. The main variable in this test series is the size of the corner columns. The slab thickness and reinforcing details were kept constant for all three specimens. The following observations and conclusions are made, based on the experimental results and the strength predictions:

1.

The one-way shear capacities and the two-way shear capacities were used to predict the strength of the slab-column connections. The one-way shear strengths provided a more accurate prediction of the shear failure than the two-way shear predictions.

2.

A practical approach, suitable for codification, would be to assume a oneway shear failure on a critical section located at d/2 from the column faces. This is the same critical section assumed for two-way shear.

3.

Two of the specimens failed in a brittle manner, exhibiting a one-way shear failure. Although torsional cracking was observed, the dominant failure crack was clearly a one-way shear failure. The specimen with the largest column exhibited ductile flexural yielding, having greater post-peak

80

deflections than the other two specimens. followed by a one-way shear failure. 4.

This flexural yielding was

The size of the column was found to have also influence on the capacity of the specimens although S2, with a larger column, experienced a peak load that was slightly less than Specimen S1. In view of the limited number of tests, it would be desirable to conduct further geometrically similar column size-effect tests with larger slabs.

5.

The code rules for the computation of punching shear for corner columns without shear reinforcement differ considerably with respect to the definition of the control perimeter and calculation of the nominal shear stress. EC2-02 considers the size effect and the effect of amount of

longitudinal reinforcement, while the 1994 CSA Standard does not account for these effects. With respect to the flexural reinforcement ratio, it is observed that EC2-02 gives higher capacities than the 1994 CSA Standard for large flexural reinforcement ratios (p>1.3%) and more conservative shear strength for lower reinforcement ratios. 6. All punching failures occurred at connections where yielding of all top reinforcing bars was recorded. For this and the above reason, it is

believed that the punching shear strength of slabs is a function of flexural reinforcement ratio and the yield stress, p fy. 7. There are considerable difficulties in performing an analysis that accounts for the additional tensions in the reinforcement due to torsion. Therefore, further research and development of models, material and computer methods are needed in this important area.

81

References

ACI-ASCE Committee 326, "Shear and Diagonal Tension", Proc, ACI, Vol. 59, 1962, pp. 84-115, and 118-120. Albrecht, U., "Design of Flat Slabs for Punching - European and North American Practices", Cement & Concrete Composites, Vol. 24, 2002, pp.531-538. Alexander, S. D. B., and Simmonds, S. H., "Bond Model for Concentric Punching Shear", ACI Structural Journal, Vol. 89, 1992, pp. 325-334. Alexander, S. D. B., and Simmonds, S. H., "Tests of Column-Flat Plate Connections", ACI Structural Journal, Vol. 89, 1992, pp. 495-502. Alexander, S. D. B., and Simmonds, S. H., "Ultimate Strength of Slab-Column Connections", ACI Structural Journal, Vol. 84, 1987, pp. 255-261. American Concrete Institute, ACI 318-02, Building Code Requirements for Reinforced Concrete, Detroit, Michigan, 2002, 443 pp. American Concrete Institute, ACI 318-95, Building Code Requirements for Reinforced Concrete, Detroit, Michigan, 1995, 369 pp. Andersson, J. L., "Genomstansning av plattor understodda av pelare vid fri kant", Nordisk Betong, 1966:2, pp. 179-200. ASCE-ACI Committee 426, "The Shear Strength of Reinforced Concrete Members: Slabs", Proc, ASCE - Journal of the Structural Divison, Vol. 100, No. ST8, 1974, pp. 1543-1591. ASTM, Significance of Tests and Properties of Concrete and Concrete-Making Materials, P. Klieger and J. F. Lamond, 4th ed., Fredericksburg, VA, 1994, pp. 222-227. Bazant, Z. P., and Cao, Z., "Size effect in Punching Shear Failure of Slabs", ACI Structural Journal, Vol. 84, No. 1, 1987, pp. 44-53. Bortolotti, L., "Punching Shear Strength in Concrete Slabs", ACI Structural Journal, Vol. 87, No. 2, 1990, pp. 208-219. Braestrup, M. W., Nielsen, M. P., Jensen, B. C , and Bach, F., Axisymmetric Punching of Plain and Reinforced Concrete, Report R75, Structural Research Laboratory, Technical University of Denmark, Copenhagen, 1976. Broms, C. E., "Punching of Flat Plates - A Question of ConcreteProperties in Biaxial Compression and Size Effect", ACI Structural Journal, Vol. 87, No. 5, 1990, pp. 292-304. 82

BS 8110, Structural use of Concrete, Part 1, Code of Practice for Design and Construction, British Standards Institution, London, 1997. Canadian Standards Association, CSA Standard A23.3-94, Design of Concrete Structures, Rexdale, Ontario, 1994, 199 pp. Collins, M. P., and Mitchell, D., Prestressed Concrete Basics, Canadian Prestressed Concrete Institute, Ottawa, Ontario, 1987, 614 pp. Comite Euro-International du Beton et Federation Internationale de la Precontrainte, CEB-FIP Model Code for Concrete Structures, Lausanne, Switzerland, 1990, 437 pp. Comite Europeen de Normalisation (CEN), Eurocode 2: Design of Concrete Structures. Part 1 - General Rules and Rules for Buildings, prEN 1992-1, 2002, 211 pp. (Final Draft, July 2002) Correspondence between Committee 326 and R. Diaz de Cossio, National University of Mexico. DIN 1045 (07/88), Beton und Stahlbeton, Bemessung und Ausfuehrung, Berlin, 1998. Elgabry, A., and Ghali, A., "Moment Transfer by Shear in Slab-Column Connections", ACI Structural Journal, Vol. 93, No. 2, 1996, pp. 187-196. Elgabry, A., and Ghali, A., "Transfer of Moments between Columns and Slabs: Proposed Code Revisions", ACI Structural Journal, Vol. 93, No. 1, 1996, pp. 5661. Elstner, R. C , and Hognestad, E., "Shearing Strength of Reinforced Concrete Slabs", Proc, ACI Journal, Vol. 53, No. 1, 1956, pp. 29-58. fib Bulletin 12, Punching of Structural Concrete Slabs. Technical Report, Federation Internationale du Beton, Lausanne, 2001, 307 pp. fib Bulletin 2, Structural Concrete. Textbook on Behaviour, Design and Performance - updated knowledge of the CEB/FIP Model Code 1990, Federation Internationale du Beton, Vol. 2: Basis of design, Lausanne, 1999, pp. 202-210. FIP - Recommendations, Practical Design of Structural Concrete. FIPCommission 3 "Practical Design", Sept. 1996. SETO, London, 1999 (distributed by fib, Lausanne). Forsell, C , and Holmberg, A., "Stampellast pa plattor av betong", Betong Vol. 131, No. 2, Stockholm, 1946, pp. 95-123. Georgopoulos, T., "Durchstanzlast und Durchstanzwinkel punktformig gestutzter Stahlbeton platten ohne Schubbewehrung", Bauingenieur 64, 1989, pp. 187-191.

83

Ghali, A., and Megally, S., "Design for Punching Shear Strength with ACI 31895", ACI Structural Journal, Vol. 96, No. 4, 1999, pp. 539-548. Graf, O., "Tests of Reinforced Concrete Slabs under Concentrated Loads Applied Near One Support", Deutscher Ausschuss fur Eisenbeton, Berlin, Heft 73, 1933,2 pp. Hammil, N., and Ghali, A., "Punching Shear Resistance of Corner Slab-Column Connections", ACI Structural Journal, Vol. 91, No. 6, 1994, pp. 697-707. Hawkins, N. M., and Mitchell, D., "Progressive Collapse of Flat Plate Structures", Proc, ACI Journal, Vol. 76, No. 7, 1979, pp. 775-808. Hawkins, N. M., and Mitchell, D., and Hanna, S., "The Effects of Shear Reinforcement on the Reversed Cyclic Loading Behaviour of Flat Plate Structures", Canadian Journal of Civil Engineering, Vol. 2, August, 1975, pp. 572-582. Hognestad, E., "Yield-line Theory for the Ultimate Flexural Strength of Reinforced Concrete Slabs", ACI Journal, Vol. 24, No. 7, 1953, pp. 637-656. Ingvarsson, H., Betongplattors Hallfasthet och armeringsurformning vid hompelare, Meddelande Nr 122, Institutionen for Byggnadsstatik, Kungliga Teknisha Hogskolan, Stockholm, 1977. Joint Committee of 1924, "Standard Specifications for Concrete and Reinforced Concrete - Joint Committee", Proc, ASTM, Vol. 24, Parti, 1924, pp. 312-385. Kinnunen, S., and Nylander, H., "Punching of Concrete Slabs without Shear Reinforcement", Transactions of the Royal Institute of Technology, Stockholm, No. 158, 1960, 112 pp. Lim, F. K., and Rangan, B. V., "Studies of Concrete Slabs with Stud Shear Reinforcement in Vicinity of Edge and Corner Columns", ACI Structural Journal, Vol. 92, No. 5, 1995, pp. 515-525. MacGregor, J., and Bartlett, F., Reinforced Concrete: Mechanics and Design, Prentice Hall Canada, 1st Canadian ed., Scarborough, 2000, pp. 577-646. Menetrey, P., "Synthesis of Punching Failure in Reinforced Concrete", Cement & Concrete Composites, Vol. 24, 2002, pp. 497-507. Menetrey, P., Walther, R., Zimmermann, T., William, K., and Regan, P., "Simulation of Punching Failure in Reinforced-Concrete Structures", ASCE Journal of Structural Engineering, Vol. 123, No. 5, 1997, pp. 652-659. Moe, J., Shearing strength of reinforced concrete slabs and footings under concentrated loads, Development Department Bulletin D47, Portland Cement Association, Skokie, Illinois, 1961, 130 pp.

84

Moehle, J. P., "Strength of Slab-Column Edge Connections", ACI Structural Journal, Vol. 85, No. 1, 1988, pp. 89-98. Moehle, J. P., Kreger, M. E., and Leon, R., "Background to Recommendations for Design of Reinforced Concrete Slab-Column Connections", ACI Structural Journal, Vol. 85, No. 6, 1988, pp. 636-644. Mortin, J., and Ghali, A., "Connection of Flat Plates to Edge Columns", ACI Structural Journal, Vol. 88, No. 2, 1991, pp. 191-198. Newmark, N. M., Siess, C. P., and Peckham, W. M., Studies of Slab and Beam Highway Bridges, Part 2, Bulletin 375, Engineering Experiment Station, University of Illinois, Urbana, Illinois, 1948, 60 pp. Park, R., and Gamble, W.L., Reinforced Concrete Slabs, Wiley, 2nd ed., New York, 2000, 716 pp. Paultre, P., and Mitchell, D., "Code Provisions for High Strength Concrete. An International Perspective", Concrete International, Vol. 25, No. 5, 2003, pp. 7690. Rankin, G. I. B., and Long, A. E., "Predicting the Punching Strength of Conventional Slab-Column Connections", Proc, Institution of Civil Engineers, Vol. 82, Part 1, 1987, pp. 327-346. Regan, P. E., "Design for Punching Shear", Structural Engineer, Vol. 52, No. 6, London, 1974, pp. 197-207. Regan, P. E., and Braestrup, M. W., Punching Shear in Reinforced Concrete: A State of Art Report, Bulletin d'information No. 168, Comite Euro-International du Beton, Lausanne, 1985, 232 pp. Richart, F. E., "Reinforced Concrete Wall and Column Footings", Proc, ACI Journal, Vol. 45, No. 2 and 3, 1948, pp. 97-127 and 237-260. Richart, F. E., and Kluge, R. W., Tests of Reinforced Concrete Slabs Subject to Concentrated Loads, Bulletin 314, Engineering Experiment Station, University of Illinois, Urbana, 1939. Scordelis, A. C , Lin, T. Y., and May, H. R., "Shearing Strength of Prestressed Lift Slabs", ACI Journal, Vol. 30, No. 4, 1958, pp. 485-506 (Proc. V55). Shehata, I. A. E. M., and Regan, P., "Punching in R. C. Slabs", ASCE - Journal of Structural Engineering, Vol. 115, No. 7, 1989, pp. 1726-1740. Sherif, A. G., Dilger, W. H., "Analysis and Deflections of Reinforced Concrete Flat Slabs", Canadian Journal of Civil Engineering, Vol. 25, 1998, pp. 451-466. Theodorakopoulos, D. D., and Swamy, R. N., "Ultimate Punching Shear Strength Analysis of Slab-Column Connections", Cement & Concrete Composites, Vol. 85

24, 2002, pp. 509-521. Talbot, A. N., Reinforced Concrete Wall Footings and Column Footings, Bulletin 67, Engineering Experiment Station, University of Illinois, Urbana, Illinois, 1913, 114 pp. Vanderbilt, M. D., "Shear Strength of Continuous Plates", Proc, ASCE - Journal of the Structural Division, Vol. 98, No. ST5, 1972, pp. 961-973. Walker, P., and Regan, P., "Corner Column-Slab Connections in Concrete Flat Plates", ASCE - Journal of Structural Engineering, Vol. 113, No. 4, 1987, pp. 704-720 Whitney, C. S., "Ultimate Shear Strenght of Reinforced Concrete Flat Slabs, Footings, Beams, and Frame Members without Shear Reinforcement", Proc, ACI Journal, Vol. 54, No. 4, 1957, pp. 265-298. Zaghlool, R., and de Paiva, H., "Strength Analysis of Corner Column-Slab Connections", Proc, ASCE - Journal of the Structural Division, Vol. 99, No. ST1, 1973, pp. 53-70. Zaghlool, R., and de Paiva, H., "Tests of Flat-Plate Corner Column-Slab Connections", Proc, ASCE - Journal of the Structural Division, Vol. 99, No. ST3, 1973, pp. 551-572.

86

Potrebbero piacerti anche