Sie sind auf Seite 1von 41

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/259221644

Ultimate Strength of Reinforced Concrete Circular Members Subjected to Axial


Force, Bending Moment, and Shear Force

Article  in  Journal of Structural Engineering · June 2013


DOI: 10.1061/(ASCE)ST.1943-541X.0000724

CITATIONS READS

10 310

2 authors:

Pier Paolo Rossi Antonino Recupero


University of Catania Università degli Studi di Messina
60 PUBLICATIONS   553 CITATIONS    61 PUBLICATIONS   344 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Seismic design of tied braced steel structures View project

In.CAM.M.I.N.O. View project

All content following this page was uploaded by Pier Paolo Rossi on 14 April 2015.

The user has requested enhancement of the downloaded file.


ULTIMATE STRENGTH OF REINFORCED CONCRETE CIRCULAR MEMBERS

SUBJECTED TO AXIAL FORCE, BENDING MOMENT AND SHEAR FORCE

P. P. Rossi1
2
A. Recupero

ABSTRACT3

Over the last few decades a significant effort has been made to achieve accurate evaluation of the

resistance of reinforced concrete elements subjected to pure shear or combined internal forces

including shear. As regards the latter, continuum models characterized by simplified stress fields

have recently been applied by one of the Authors for the evaluation of the ultimate capacity

interaction diagram of rectangular reinforced concrete cross-sections undergoing combined axial

force, bending moment and shear force.

This paper illustrates the natural progress of these studies and describes an analytical tool for the

calculation of the ultimate strength of reinforced concrete columns with circular cross-section. The

proposed method is based on the application of the static theorem of limit analysis and takes into

account both truss and arch effects. To ascertain the accuracy and reliability of the method, the

relations developed are applied with reference to a large number of tests reported in the literature

and a comparison between the theoretical and experimental results is drawn. Finally, the predictions

of the proposed method are compared with those of other simplified methods present in the

literature.

Keywords: Circular cross-section, shear force, axial force, bending moment, reinforced concrete

1
Associate Professor, Dept. of Civil and Environmental Engineering, Univ. of Catania, Catania 95125, Italy
(corresponding author). E-mail: prossi@dica.unict.it
2
Assistant Professor, Dept. of Civil Engineering, Univ. of Messina, Messina, S. Agata (Messina) 98166, Italy
E-mail: antonino.recupero@unime.it
INTRODUCTION

According to a traditional approach, the design shear strength VRd of beams and columns is often

calculated by means of the relation

VRd  VcRd  VsRd (1)

as the sum of two contributions: the strength of the concrete shear-resisting mechanism VcRd and the

strength of the truss mechanism VsRd.

The first term, VcRd, takes into account the shear force transmitted across the compression zone of

the element, the dowel action of the longitudinal reinforcement and the vertical component of the

shear force transferred across the crack by aggregate interlock effect. As also reported by some

researchers (Turmo et al. 2009) , its evaluation is controversial and relies on empirical methods.

The second term, VsRd, represents the shear force corresponding to yielding of the transverse

reinforcement. In the case of rectangular, T and I beams with stirrups orthogonal to the longitudinal

member axis this contribution is usually calculated by means of the expression

Asw
VsRd  z f ywd cot θ (2)
s

where z is the internal lever arm, i.e. the distance between the compressive and tensile resultant

forces, Asw is the total transverse reinforcement area per layer, s is the spacing of the transverse

reinforcement along the longitudinal member axis, fywd is the yield strength of the transverse

reinforcement and  is the angle of inclination between the compressive stresses of the concrete in

the web and the longitudinal member axis.

A relation similar to Equation (2) has also been derived by Ang et al. (1989) for reinforced concrete

elements with circular cross-section. In particular, these researchers assumed that diagonal cracks

were characterised by an inclination of 45 degrees with respect to the longitudinal member axis and

that cracks were able to mobilize the transverse reinforcement along a length extending the full

width of the confined core of the concrete. The truss mechanism contribution was evaluated as
Asw π
VsRd  D ' f ywd (3)
s 4

where D’ is the diameter of the circular hoops. This relation has subsequently been modified in

Priestley et al. (1994a; 1996) and the term VsRd calculated by the relation

Asw π
VsRd  D ' f ywd cot θ (4)
s 4

where the angle is assumed equal to the maximum between 30 degrees and the corner-to-corner

inclination of the member.

Recently, Kowalsky and Priestley (2000) have modified Equation (4) so as to consider that cracks

are closed in the compression zone of the member and thus that shear cannot be transferred across

cracks by tension strain in the transverse reinforcement. The steel truss mechanism is evaluated by

means of the expression

 D 2
2
D' 2
Asw
'
 y2
VsRd   f ywd cot θ dy (5)
xn  D 2
s D' 2

where xn is the depth of the compression zone and D is the diameter of the cross-section. In the

same paper Kowalsky and Priestley also propose the following more simple relation where the

integration is removed and the effective transverse reinforcement area is approximated as Asw /4


VsRd   D 2  D ' 2  xn 
Asw
f ywd cot θ (6)
s 4

More recently, Turmo et al. (2009) have proposed a new expression for the calculation of the

strength of the truss mechanism VsRd

Asw
VsRd  χ z f ywd cot θ (7)
s

This expression is formally similar to that provided by Priestley et al. (1994a; 1996) but it is

modulated through the efficiency factor . The value of this parameter is variable with the extension

of the compression zone of the cross-section (i.e. depending on the position of the compressive and

tensile resultant forces) and is calculated by means of the following integral


2
1 z zX 
χ 1  4  0 '  dX (8)
0
 D 

where z0 is the distance from the centroid of the tensile forces to the geometric centre of the

cross-section. It is worth noting that Equation (7) is equal to Equation (4) if the height of the hoops

is considered to be fully effective. In fact, in this case the lever arm of the internal forces is equal to

the diameter of the hoops, the ratio z0/D’ is equal to 0.50 and thus the efficiency factor is equal to

 /4.

Turmo et al. (2009) recognize that the use of Equation (8) is not simple in real situations because

the internal lever arm and the efficiency factor are not constant throughout the length of the member.

To overcome this problem, they propose a design oriented formula of the efficiency factor in which

z/D’ is assumed equal to 0.9 and z0/D’ is equal to 0.45. These simplifications lead to a value of the

efficiency factor equal to 0.85 and to the following expression of the term VsRd

Asw
VsRd  0.765D ' f ywd cot θ (9)
s

The same researchers also affirm that the design oriented formula leads to inaccurate results for

columns with low z/D’ ratios and high z0/D’ ratios and that, in these cases, the more complex

relation of the efficiency factor [Eq. (8)] should be used.

In the attempt to overcome some drawbacks of the traditional approach and achieve more accurate

estimates of the shear strength many researchers have also addressed the study of continuum

structural models. As attention is usually focused on the ultimate load capacity of structural

members, the approach based on continuum models often makes use of the fundamental theorems

of limit analysis and considers simplified continuous stress fields to simulate the response of steel

and concrete. Ever since the first studies on this topic (e.g. Nielsen et al. 1978; Nielsen 1984) the

use of continuum models has gradually been perfected and allowed the achievement of important

results with moderately onerous computational tools. In particular, continuum models have often

been used to compensate for the lack of valid code provisions for the assessment of the resistance of
concrete members subjected to the combined action of forces including shear. In this very context,

this approach has been adopted by one of the Authors to define the ultimate capacity interaction

diagrams of reinforced (Recupero et al. 2003) and prestressed concrete members (Recupero et al.

2005) characterised by rectangular, T or I shaped cross-section and subjected to axial force, shear

force and bending moment.

The present paper constitutes the natural progress of these studies and leads to the evaluation of the

ultimate strength of reinforced concrete circular columns subjected to axial force, bending moment

and shear force. Unlike similar methods proposed in the past (Recupero et al. 2003; 2005), the

method described in this paper considers both truss and arch mechanisms. This method is in step

with the guidelines reported in the Fib Structural Concrete Textbook (2009) which, in the case of

structural elements subjected to axial forces, proposes the superposition of statically allowable

partial models.

The non-linear problem is described from the mathematical point of view and applied to a large

number of columns tested in the past. The comparison between the numerical and laboratory results

is drawn to assess the accuracy and reliability of the approach. The resisting shear forces resulting

from the proposed method are also compared with those from other methods in order to highlight

the advantages for adopting the proposed method instead of existing and more simplified ones.

APPROACH TO THE PROBLEM AND PHYSICAL MODEL

The proposed method is based on the application of the static theorem of limit analysis and

considers both truss and arch actions. The mathematical programming problem resulting from the

application of the static theorem is nonlinear and defined by means of equilibrium equations and

constraint conditions for geometric and mechanical parameters. All these relations are reported later

assuming that the longitudinal axis of the reinforced concrete member is parallel to the z-axis of a

reference system and that the external shear force is oriented along the positive y-axis.

To achieve accurate evaluations of the ultimate strength of the column, the generic cross-section of

this member is divided into three parts, named F1, F2 and F3 (Fig. 1). In each of these parts the
response of the basic materials (concrete and steel) is defined by means of simplified stress fields

which are described later for truss and arch mechanisms separately considered. The stress-strain

constitutive behaviour of both concrete and steel is considered to be rigid-perfectly plastic; however,

while steel bars are assumed to resist both compression and tension, concrete is assumed to resist

compression only. The geometry of the zones F1, F2 and F3 of the cross-section is completely

defined once the position of the separation lines of the central part F3 is known. The position of

these two lines is determined through the angles, 1 and 2, that the x-axis forms with the segments

connecting the geometric centre to the separation line endpoints (Fig.1a).

The column is endowed with longitudinal and transverse reinforcements. In particular, the area Asl

of the longitudinal steel bars is assumed to be distributed over the circumference passing through

the centroids of the longitudinal steel bars. The longitudinal reinforcement  per unit of length of

this circumference is

Asl A
  sl (10)
2  R  c  2Rl

where R is the radius of the column cross-section, Rl is the radius of the circle passing through the

centroid of the longitudinal steel bars and c is the mechanical cover of the longitudinal

reinforcement, i.e. the distance between the external surface of the cross-section and the centroid of

the longitudinal bars. The geometry of the distributed reinforcement  included in the zones F1, F2

and F3 is defined through the angles  and  subtended at the geometric centre by the circle arc

delimited by the x-axis and the separation lines (Fig.1b)

The transverse reinforcement is constituted by circular hoops orthogonal to the longitudinal member

axis and placed at a distance cs from the external surface of the column cross-section. The spacing

of hoops and the mechanical properties of steel of hoops are assumed to be constant along the

length of the column. The geometry of the transverse reinforcement included in the zones F1, F2 and

F3 is defined through the angles  and subtended at the geometric centre by the hoop segment

delimited by the x-axis and the separation lines.


Later, angles  and  are often expressed as a function of the angle  Equations which represent

these relations are:

1  arcsin  rl sin 1  (11)

 2  arcsin  rl sin  2  (12)

1  arcsin  rs sin 1  (13)

2  arcsin  rs sin  2  (14)

where

rl  R ( R  c) (15)

rs  R ( R  cs ) (16)

Angles ,  and  are assumed positive if anticlockwise.

ANALYTICAL FORMULATION OF THE TRUSS ACTION

The Authors hypothesize that in the outermost parts of the cross-section (called F1 and F2)

longitudinal reinforcement and concrete are subjected to “stress fields” parallel to the longitudinal

member axis and that these stresses are constant within the single part of the cross-section. In the

central part (called F3) longitudinal and transverse reinforcements are assumed to experience

stresses which are constant and parallel to the axis of the steel bars. Still in the central zone,

concrete is assumed to experience principal compressive stresses inclined at an unknown angle 

with respect to the longitudinal member axis. Similar stress fields have already been suggested by

many other researchers (e.g. Nielsen et al. 1978; Hsu 1993) and are particularly justified in

reinforced concrete members in which steel bars are closely spaced.

Equilibrium conditions: parts F1 and F2 of the cross-section

To define the equilibrium conditions relative to the cross-section parts named F1 and F2, let us

consider the elements E1 and E2 (Fig. 2a) obtained by cutting the reinforced concrete member by
means of two couples of planes which are parallel and orthogonal to the longitudinal member axis.

In addition, let us consider that those planes which are parallel to each other are set apart at an

infinitesimal quantity. The elements E1 and E2 have, therefore, size equal to dy and dz along the y

and z-axes and depth equal to 2Rcosalong the x-axis. Further, they are subjected to the stresses l

and c due to longitudinal reinforcement and concrete (Fig. 3).

The equilibrium equation of the element E1 along the z-axis may be written as


l1  1 R cos   c1 R cos   0 (17)
cos

where 1 is the equivalent normal stress due to the response of steel and concrete. It is worth noting

that in all the equations in this paper the normal stress of concrete is assumed to be positive when

compressive and that the normal stress of longitudinal and transverse reinforcements is assumed to

be positive when tensile.

Starting from Equation (17), the normal stress 1 may be defined as


1  l1  c1 (18)
R cos  1  rl 2 sin 2 

As is evident, this stress is not constant within the part F1 because it depends on the angle .

Equation (17) simplifies further when referring to elements which are within the mechanical cover

of the longitudinal steel bars and takes the form

 1 2 R cos    c1 2 R cos   0 (i.e.  1   c1 ) (19)

Similar mathematical expressions may be obtained with reference to the element E2 in the part F2 of

the cross-section. Specifically, the equilibrium along the z-axis may be expressed by means of the

relation


l 2  2 R cos   c 2 R cos   0 (20)
cos

and thus the stress 2 may be evaluated as



 2  l 2  c 2 (21)
R cos  1  rl 2 sin 2 

When referring to elements within the mechanical cover of the longitudinal steel bars, Equation (20)

simplifies into the following relation

 2 2 R cos   c 2 2 R cos   0 (i.e.  2   c 2 ) (22)

Equilibrium conditions: part F3 of the cross-section

In the central part (F3) of the cross-section let us consider two elements (E3 and E4) which are

obtained by cutting the reinforced concrete member by means of three planes parallel to the x-axis

(see Fig. 2b). Like E1 and E2, the elements E3 and E4 have depth equal to 2Rcosalong the x-axis.

First, let us focus attention on the element E3 which is obtained by cutting the member with one

plane parallel to the compressive stress of the concrete and with other two planes parallel and

orthogonal to the longitudinal member axis. This element is subjected to the stress l3 of the

longitudinal reinforcement, to the stress s3 of the circular hoops, to the tangential stress and to the

equivalent normal stress 3 of concrete and steel bars (Fig. 4a). The stress of the longitudinal and

transverse reinforcements is assumed constant in this part of the cross-section. However, unlike

rectangular or I-shaped cross-sections, in circular cross-sections the normal stress of the hoop is

inclined to the y-axis at an angle  which depends on the position of the hoop cross-section

considered.

The equilibrium equations along the y and z-axes (see Fig. 4a) are

Asw
(transl. along the y-axis) s3 cos  cos    2 R cos sin   0 (23)
s

2
(transl. along the z-axis) l 3 sin    2 R cos  cos   3 2 R cos  sin   0 (24)
cos 

If the angle  is expressed as a function of the angle , Equation (23) gives the shear stress value

acting on the central part of the cross-section


Asw 1  rs sin 
2 2
  s3 cot  (25)
s 2 R cos 

Hence, the stress 3 may be defined by substituting Equation (25) into Equation (24)

Asw 1  rs sin 
2 2

3  l 3  s3 cot 2  (26)
R cos  1  rl sin 
2 2 s 2 R cos 

As is evident in Equations (25-26), the equivalent normal stress 3 and the shear stress  are not

constant in F3 but variable as a function of the angle .

To evaluate the normal stress c3 of concrete, let us consider the element E4 which is obtained by

cutting the member with two planes parallel and orthogonal to the compressive stress of the

concrete and with a plane parallel to the longitudinal member axis (Fig. 2b). This element is

subjected to the compressive stress c3 of concrete and to the component s3 of the stress of the

circular hoops along the y-axis (Fig. 4b).

The constraint condition y = 0 applied to the element under examination states that

Asw
s3 cos   c 3 2 R cos  sin 2   0 (27)
s

and thus

Asw 1  rs sin  1 1  rs2 sin 2 


2 2
c 3   s 3   0
c3 (28)
s 2 R cos  sin 2  cos 

where 0c 3 is the stress of concrete for =0. As evident from this equation the stress c 3 is

non-uniform in F3. In particular, it is maximum on the y-axis and symmetric with respect to this

axis.

Internal forces

The axial force of the generic cross-section is obtained by the following relation
N   1dA    2 dA   3 dA (29)
A1 A2 A3
  
N1 N2 N3

where N1, N2 and N3 are the axial force contributions of the parts F1, F2 and F3 and A1, A2 and A3 are

the areas of the parts F1, F2 and F3.

The contribution N1 to the axial force may be calculated by means of the relation

2 llim 2
cos 
N1   1 2 R cos  d  l1 2 R  d  c1 2 R 2  cos  d
2 2 2
(30)
1 1* 1  rl sin 
2 2
1

where

1*  min  1 ; llim  (31)

and llim is the angle  corresponding to the separation line passing through the longitudinal steel

bar with minimum y-coordinate, i.e.

llim  arcsin  Rl R   arcsin 1 rl  (32)

The consideration of angles llim and 1* is necessary because the contribution of the longitudinal

reinforcement is possible only for values of the angle  in the range llim to llim .

The second contribution N2 is defined by the relation

*2 2
cos 
N 2 l 2 2 R  1 rl sin 
2 2
dc 2 2 R 2 
 2
cos 2  d (33)
llim

where

*2  max   2 ; llim  (34)

and the third contribution N3 by the relation

1* 
cos  A 1

N 3  l 3 2 R   d  s 3 sw R cot 2   cos  1 rs2 sin 2  d (35)


*2 1 rl 2 sin 2  s 2

The bending moment of the generic cross-section is


M   1 y dA   2 y dA   3 y dA (36)
A1 A2 A3


 

 


M1 M2 M3

The contributions M1, M2 and M3 to the bending moment may be written as

llim 2
cos  sin 
M 1  l 1 2 R   d  c1 2 R  cos  sin  d
2 3 2
(37)
1* 1  rl sin 
2 2
1

*2 2
cos  sin 
M 2  l 2 2 R   d   c 2 2 R  cos 2  sin  d
2 3
(38)
llim 1  rl sin 
2 2
 2

1* 
cos  sin  A 1

M 3  l 3 2 R   2
d   s 3 sw R 2 cot 2   sin  cos  1  rs2 sin 2  d (39)
*2 1  rl 2 sin 2  s  2

Finally, the shear force is evaluated as

2

V  2R   cos  d
2 2
(40)
1

The calculation of the integral above leads to the relation

1 A
V   s 3 sw Rs cot  1  sin 1 cos 1  2  sin 2 cos 2  (41)
2 s

where Rs is the radius of the circular hoops. In particular, it is worth noting that if 1=-2=π/2 the

result of Equation (41) equals that produced by Equation (4).

ANALYTICAL FORMULATION OF THE ARCH ACTION

As explained by some researchers (Priestley et al. 1994; Kowalsky and Priestley 2000), in the

presence of external axial load some lateral force can be transmitted directly to the base of the

member through inclined compressive stresses (this type of arch effect is later named as type 1).

The magnitude of the resisting shear force produced by the external axial load is equal to the

horizontal component of the diagonal compression strut that forms between the top and bottom of

the column (Fig. 5a). However, as remarked by Priestley et al. (1994a) the contribution of this effect

to the axial resistance cannot be greater than the external axial load.
An arch effect may also exist without any contribution being made by the external axial load (this

type of arch effect is later named as type 2) (e.g. see Watanabe and Ichinose 1991). In this case a

diagonal compressive strut can directly transfer the lateral load to the base if the member is

endowed with longitudinal reinforcement able to resist the vertical component of the diagonal

compressive force. The diagonal compressive strut can form whether the shear force is applied to

the lateral surface of the member (Fig. 5b) or to the geometric centre of the cross-section (Fig. 5c).

It is worth noting that the arch effect type 2 can develop both in the presence and in the absence of

axial load. Owing to this, in the subsequent analyses both types of arch actions are simultaneously

considered.

To include the arch effect type 1 in the proposed model, the Authors assume that in single bending

columns the resultant of the diagonal strut passes through the point of application of the external

axial load at the top of the column and through the centre Gc2 of the zone F2 at the base of the

column (Fig. 5a); even if not shown in the figure, in double bending columns the strut is assumed to

enter and leave the column through the centre of the zone F2 at the column top and bottom.

The internal forces (N’, M’, V’) produced by this type of arch action are:

2

N '    'c 2 dAc  'c 2 2 R 2  cos 2  d (42)


A2  2

2

M     y dAc   2 R
' '
c2
'
c2
3
 cos 2  sin  d (43)
A2  2

2

V     cot  dAc   2 R cot 


' '
c2
' '
c2
2 '
 cos 2  d (44)
A2  2

where 'c 2 is the vertical stress produced in concrete at the toe of the column base and ' is the

angle of inclination of the strut with respect to the longitudinal member axis. The latter parameter is

given by the following relation


 2 
     
2
cos sin d
 yGc 2   R  2 
'  arctan    arctan  2  (45)
 LV   LV
 cos  d 
2
 
  2 

where yGc 2 is the y-coordinate of the centroid of the area A2 and LV is the shear span of the

member. In regard to the possible values of the normal stress 'c 2 , it should be noted that the

contribution of this arch mechanism to the axial resistance cannot not be greater than the external

axial load (Priestley et al. 1994a; Kowalsky and Priestley 2000) .

To describe the arch effect type 2, the Authors consider the case in which the shear force is applied

to the lateral surface of the member separately from the case in which the shear force is applied to

the cross-section. Referring to the former case (frequent in laboratory tests), the axis of the strut is

assumed to pass through the centroid of the zone named F2 and the intersection point between the

directions of the lateral force and the resultant of the tensile stresses in the longitudinal bars of the

zone named F1 (Fig. 5b). The geometric characteristics of this strut are completely defined through

the geometric properties of the concrete in the zone F2 and the longitudinal reinforcement in the zone

F1. The internal forces (N’’, M’’, V’’) produced by this type of arch action are

N ''   dAl   dAc  0


'' ''
l1 c2 (46)
A1 A2

llim 2
cos  sin 
M   y dAl   y dAc   2 R   d   2 R  cos 2  sin  d
'' '' '' '' 2 '' 3
l1 c2 l1 c2 (47)
A1 A2 1* 1  rl sin 
2
 2

2

V   cot  dAc   2 R cot   cos 2  d


'' '' '' '' 2 ''
c2 c2 (48)
A2  2

where ''c 2 is the vertical compressive stress in the concrete of the zone F2, l''1 is the tensile stress

in the longitudinal bars of the zone F1 and '' is the angle of inclination of the strut with respect to

the longitudinal member axis.

The angle of inclination '' of the strut is


 2 llim
cos  sin  
 yG  yGl 1 

 R  cos  sin  d
2

R
 1  rl 2 sin 2 
d 

 2 1*
''  arctan  c 2   arctan  2
 llim  (49)
 LV   LV LV cos 
 cos 2  d  d 
 1  rl sin 
2 2 
  2 1* 

where yGl1 is the y-coordinate of the centroid of the longitudinal reinforcement in the zone F1.

In the case the shear force is applied to the cross-section the strut is assumed to pass through the

centroid of the compression zone at the column base and the centroid of the member cross-section

at a distance from the base equal to the shear span.

The internal forces (N’’, M’’, V’’) are

N ''   dAl   dAc  0


'' ''
l3 c2 (50)
A3 A2

** 2
cos  sin 
M    y dAl    y dAc   2 R   d  c'' 2 2 R3  cos 2  sin  d
'' '' '' '' 2
l3 c2 l3 (51)
A3 A2  ** 1  rl sin 
2
 2

2

V   cot  dAc   2 R cot   cos 2  d


'' '' '' '' 2 ''
c2 c2 (52)
A2  2

In Equation (51), ** is the angle which identifies the area of the longitudinal bars subjected to the

stress l''3 due to the arch effect type 2. The angle ** is given by the following relation supposing

that the longitudinal bars subjected to the stress l''3 are symmetric with respect to the longitudinal

axis of the member

**  min  1 ;  2   0 (53)

The angle of inclination '' of the strut is

 2 
     
2
cos sin d
 yGc 2   R  2 
''  arctan    arctan  2  (54)
 LV   LV
 cos 2  d 
 
  2 
MECHANICAL AND GEOMETRIC CONSTRAINT CONDITIONS

The constraint conditions reported in this section refer to geometric and mechanical parameters. In

particular, in regard to the discretization of the generic cross-section the solution of the

programming problem must ensure that the zones F1 and F2 do not overlap. To this end, angles 1

and 2 must verify the inequality

1   2 (55)

As evident from Equation (28), diagonal stresses caused by truss action can develop in the concrete

of a layer orthogonal to the bending plane and belonging to the central part F3 if the circular hoop

intersects this layer. The maximum extension of the central zone is thus equal to the diameter of the

circular hoop, i.e.

-lim
s  1   s
lim
(56)

-lim
s  2   s
lim
(57)

In these relations the angle lim


s corresponds to the separation line passing through the point of the

circular hoop with minimum y-coordinate and is calculated by means of the relation

lim
s  arcsin 1 rs  (58)

The constraint conditions involving mechanical parameters refer to the simplified stress fields

considered for concrete and steel bars. The relations reported for longitudinal and transverse

reinforcements are simple and rigorous because the stresses produced by truss and arch actions are

parallel to the steel bars. The maximum normal stresses must be lower than the yield strength of

steel and, therefore, the following relations must hold

 f yl  l1  l''1  f yl (59)

 f yl  l 2  f yl (60)

 f yl  l 3  l''3  f yl (61)

 f yw   s 3  f yw (62)
fyl being the yield strength of the longitudinal reinforcement and fyw the yield strength of the circular

hoops.

Different and slightly approximate constraint conditions are considered for concrete in zones F1, F2

and F3. Specifically, in zones F1 and F2 the normal stress of concrete is assumed to be positive and

bounded above by the compressive strength of concrete fc1 , as shown in the following inequalities

0  c1  f c1 (63)

0  c 2  'c 2 cos 2 '  ''c 2 cos 2 ''  f c1 (64)

As evident in the last relation, to simplify the solution of the nonlinear programming problem, the

maximum normal stress considered in the zone F2 is assumed equal to the sum of the maximum

normal stresses caused by truss and arch effects, i.e. the three maximum normal stresses are added

ignoring they have different directions.

In the central zone F3, instead, the normal stresses of concrete must be positive and bounded above

by the reduced compressive strength of concrete under biaxial state of stress fc2 . The normal

stresses relative to this zone of the cross-section must verify the relations

c 3,min  0 (65)

c 3,max  ''c 2 cos 2 ''  f c 2 (shear force applied to the lateral surface of the column) (66)

c 3,max  f c 2 (shear force applied to the centroid of the cross-section) (67)

where c 3,min and c 3,max are the minimum and maximum diagonal stresses produced by truss

action in the central zone of the cross-section. The first of these relations is intended to ensure that

the minimum normal stress is positive. It is worth noting that in this relation only the contribution

due to truss action is considered, i.e. to be on the safe side, no compressive stress produced by arch

mechanism is added. As apparent from Equation (28), the stress c 3,min produced by truss action is

developed on either separation line of the central zone F3. The calculation of the maximum normal

stress considers, instead, the contributions of both truss and arch actions. In particular, the normal
stress c 3,max produced by truss action is obtained in the concrete layer orthogonal to the bending

plane and closer than others to the geometric centre of the cross-section. If the shear force is directly

applied to the lateral surface of the member the principal stress produced by the arch mechanism

type 2 is added to the truss contribution because the strut caused by this arch mechanism in squat

columns is often characterised by fairly high angles of inclination with respect to the longitudinal

member axis. Therefore, this strut entirely crosses the central part of the member close to the

column base, where the bending moment is practically equal to the maximum value. Conversely,

the normal stress produced by the other arch actions considered in this paper are not included in the

inequality because the strut produced by these arch actions is generally characterised by low angles

of inclination with respect to the longitudinal member and thus crosses the zone F3 in a part of the

column where the bending moment is low. Similarly to the inequality (64), the approximate

maximum value of the compressive stress in F3 is calculated ignoring the difference between the

directions of the normal stresses produced by truss and arch mechanisms.

Still with reference to the arch effects the solution of the mathematical problem must ensure that the

normal stresses produced by the arch mechanisms are compressive, i.e.

'c 2  0 (68)

''c 2  0 (69)

and that the contribution of the arch effect type 1 to the resisting axial force is not greater than the

external axial load Nexp (Priestley et al. 1994a; Kowalsky and Priestley 2000), i.e.

'c 2 A2   N exp (70)

Further, to avoid that in the longitudinal reinforcement the normal stresses caused by truss and arch

effects may have opposite sign the following relations must be verified

l1  l''1  0 (71)

l 3  l''3  0 (72)
Finally, to consider that the angle of inclination of failure cracks in the zone F3 does not coincide

with that of first cracks and depends on geometric and mechanical parameters, the cotangent of the

angle of inclination of the compressive stress c 3 is assumed to be variable between two minimum

and maximum values, i.e.

cotg   cotg min  0 (73)

cotg   cotg max  0 (74)

VALIDATION OF THE PROPOSED METHOD

Comparison with results of laboratory tests

To verify the reliability and accuracy of the proposed method, the maximum shear forces resisted

by 67 columns are compared with those derived from the application of the numerical method. The

columns under examination were tested in the past in either single or double bending under a

constant value of the axial force. The laboratory tests and the relative results are described in detail

in Ang et al. 1985, 1989; Priestley et al. 1994a, 1994b, 1994c; Kowalsky and Priestley 2000; Stone

and Cheok 1989; Wong et al. 1990, 1993; Chai et al. 1991; Ascheim and Moehle 1992; Priestley

and Benzoni 1996; Kunnath et al. 1997; Hose et al. 1997; Lehman and Moehle 1998; Calderone et

al. 2000; Berry et al. 2004. Some geometric and mechanical characteristics of the members are

reported in Table 1 to show that the selected set of laboratory tests represents a suitable basis for the

validation of the proposed method. In particular, the columns considered are characterized by very

different geometric and mechanical properties as well as by different shear span to diameter ratios;

also, some of the columns were tested in the absence of the axial force. It should be noted that the

maximum internal forces recorded during the tests have been slightly corrected in Berry et al. (2004)

to take account of P- effects and that, in this paper, the modified values are considered for

comparison.

The ultimate internal actions of these columns have been numerically predicted by means of the

method described in the previous sections. The variables of the mathematical problem
are c1 , l1 , c 2 , l 2 , 0c 3 , l 3 ,  s , , 1, 2,  (regarding the truss action), 'c 2 (arch action type 1)

and either ''c 2 l''1 (in case the arch action type 2 is due to an external shear force applied to the

lateral surface of the member) or ''c 2 l''3 (in case the arch action type 2 is due to shear force applied

to the cross-section). The programming problem is constituted by one objective function and

equality and inequality conditions. The total shear force (sum of the shear forces due to truss and

arch actions) is considered as the objective function of the mathematical programming problem and

is maximized to obtain the maximum positive value of this internal force. Therefore, the function to

be maximized is

V  V '  V '' (75)

where the truss and arch contributions to the resisting shear force are obtained by Equations (40, 44,

48, 52). As is evident from these equations the objective function is non-linear in the variables of

the problem.

The constraint conditions of the problem are represented by the equilibrium equation (23); the

equilibrium equation (27) calculated for =0 so that the stress c 3 in this relation is equal to the

variable 0c 3 ; inequalities (55-57) relative to the geometry of the parts F1, F2 and F3; inequalities

(59-72) relative to the stresses produced by truss and arch actions; inequalities (73-74) relative to

the angle of inclination  of the stress c 3 in the truss mechanism; equations (46, 50) and two

equality conditions regarding the total internal forces of the base cross-section. The first of these

latter equality conditions states that the total axial force is equal to the assigned axial force Nexp, i.e.

N  N '  N exp (76)

The second states that the bending moment and shear force resulting from the numerical approach

are in the same proportion as the maximum internal forces of the laboratory test (Mexp and Vexp), i.e.

M  M '  M '' M exp


 (77)
V  V  V 
' '' Vexp
The contributions (N, N’, M, M’, M’’, V, V’, V’’) to the axial force, bending moment and shear

force are obtained by Equations (29, 36, 40, 42-44, 47-48, 51-52).

The mathematical problem is non-linear and, therefore, its solution depends on the starting values of

all the variables of the problem. To find an accurate estimate of the shear strength, the mathematical

problem is solved more than once with reference to different starting values of these variables. The

highest value of the shear force is assumed as the ultimate shear capacity of the member. The Solver

tool of the program Excel has been used to solve the single programming problem.

The mechanical properties of concrete and steel are assumed equal to those reported in papers and

thus derived in the past from laboratory tests. In particular, the concrete compressive strengths

considered in the constraint conditions are derived from the experimental values through the

following relations

f c1  f c (78)

 f 
f c 2  0.6  1  c  f c (79)
 250 

In accordance with other researchers (Walther and Miehlbradt 1990; SIA 1989), the minimum and

maximum values of the cotangent of the angle θ are calculated assuming that the failure crack angle

may be 20-25 degrees lower or higher than the angle of inclination I of the first crack. From this

consideration, the minimum value cot min is defined as

cotgmin  cotg   I    (80)

where I is the angle of inclination of the first crack with respect to the longitudinal axis of the

member (Norme tecniche per le Costruzioni 2008) and  is the maximum excursion allowed for

the angle . The angle I is obtained from the relation cot I   I where  and  are respectively

the shear stress and the principal tensile stress which develop at first crack on the chord parallel to

the neutral axis and passing through the centroid of the uncracked, homogeneous cross-section.

Therefore, the angle I is obtained from the relation


f  0.5 N exp A    0.5 N exp A 
2 2
cotgI  ct f ct (81)

where A is the cross-sectional area of concrete and f ct  0.7  0.3  f c21 3 . The maximum value of the

cotangent of the angle of inclination of the compressive stress is given by the relation

cot max  min cot( I  ); L y3  (82)

In this relation the cotangent of the angle  has been limited below to the value L/y3, where y3 is the

dimension of the zone F3 parallel to the y-axis, to consider that the direction of the compressive

stresses of the zone F3 must intersect the zones F1 and F2 (see Fig. 5). In both Equations (80, 82) the

parameter i.e. the maximum excursion allowed for the angle is assumed equal to 23.2°. So, in

the absence of external axial loads, I =45° and thus the values obtained by Equations (80, 82) are

0.4 and 2.5, as sometimes considered in codes (e.g. see Eurocode 2, 1993).

The parameter adopted for comparison of theoretical and experimental results is

Vexp
RV  (83)
Vnum

where Vnum is the ultimate shear force predicted by means of the proposed method and Vexp is the

maximum shear force recorded during the laboratory test. The values of the parameter RV are first

calculated by ignoring the arch effects and reported in Figure 6 as a function of the ratio M/VD (the

predicted resisting shear forces are also reported in Table 1).

As shown in the figure, in the case of high values of the ratio M/VD the numerical results are in

accord with those obtained from the laboratory tests. Conversely, the accuracy of the numerical

results is poor for members with low values of the ratio M/VD. This finding is not surprising

because the resisting mechanism of members with low aspect ratios is chiefly characterized by

direct travel of the force to the support, i.e. by arch action. To highlight the influence of the arch

action, the values of the parameter RV are then calculated considering both truss and arch

mechanisms in the numerical approach. The resisting mechanisms resulting from the arch actions

type 1 and 2 are considered together with that resulting from the truss action. The comparison
between Figures 6 and 7 shows that in the case of high values of the ratio M/VD the results obtained

by means of the evaluation of the sole truss action are virtually equal to those achieved by means of

the evaluation of both truss and arch mechanisms (see also Table 1). On the contrary, the results

corresponding to low aspect ratios change remarkably when the arch effects are considered. Further,

as is evident in Figure 7, the values RV corresponding to low values of the ratio M/VD are only a

little higher than those corresponding to high values of the ratio M/VD, i.e. the accuracy of the

results is virtually independent of the value of the aspect ratio. The minimum and maximum values

of the parameter RV are 0.99 and 1.38 and the standard deviation is 0.086. Even if not shown in

figure, there is no appreciable trend with increasing external axial force or longitudinal and

transverse reinforcements. In particular, the negligible correlation between the parameter Rv and the

external axial force demonstrates that the method may be successfully applied to columns and

beams as well.

It should be noted that the proposed formulation of the problem is particularly suitable for

estimating the shear strength of the abovementioned columns because the transverse reinforcement

ratio is constant along the length of these columns. In case the transverse reinforcement ratio varies

along the length of the member the shear resistance should be theoretically verified in all the

cross-sections where the transverse reinforcement ratio changes. The stress verification in more

than one cross-section would make, however, the mathematical problem very complicated. This

complication is not justified at all if the transverse reinforcement ratio is only slightly different

along the length of the member because in this case a reliable estimate of the shear strength can be

obtained assuming that the transverse reinforcement ratio of the column is equal to the value at the

column base. If some greater approximation is accepted with regard to the modelling of the column,

the abovementioned complication can be avoided even if the transverse reinforcement ratio changes

very sharply along the length of the member. In this case the proposed formulation can still provide

a noticeable result if the transverse reinforcement ratio of the whole column is defined as

representative of the different transverse reinforcement ratios of the member, e.g. equal to the
average value.

Comparison with results of simplified methods

The accuracy and reliability of the proposed method are finally compared with those of other

existing and less complicated methods. Specifically, the results described in the previous section are

compared with those deriving from the methods proposed by Ang et al. (1978), Priestley et al.

(1994a) and Kowalsky and Priestley (2000). The relations proposed by Turmo et al. (2009) are not

considered for comparison because no particular relation is suggested for the concrete contribution

to the shear strength. In all the comparison methods the shear force corresponding to the column

flexural capacity is assumed as the upper bound of the shear strength.

Table 2 reports the results which are deemed to be significant for the comparison, i.e. the results

relative to columns for which the predicted shear strengths are not all equal or virtually equal to the

shear force corresponding to the flexural capacity of the column. The mean value, the standard

deviation and the coefficient of variation of RV are also reported in the table.

As is evident from the latter two parameters, the proposed method (with arch action) is more

reliable than the other simplified methods. In all truth, the mean values of RV show that the

proposed method is less accurate than the other methods. However, this result is neither surprising

nor alarming. In fact, the predictions of the comparison methods are sometimes unconservative and

thus the higher accuracy is obtained at the expense of the structural safety. Further, it should be

noted that all the simplified methods considered here were adjusted in the past on the basis of most

of the laboratory tests reported in Table 2.

CONCLUSION

The paper proposes a mathematical model for the evaluation of the strength of columns with

circular cross-section subjected to combined axial force, shear force and bending moment. The

model considers both truss and arch effects and hypothesizes simplified stress fields to simulate the

response of steel and concrete. The shear strength is evaluated by means of limit analysis.
The study leads to the following conclusions:

- the proposed method highlights the mutual influence of the internal forces in circular

reinforced concrete members and, in particular, explains the level of importance of the shear

force for the ultimate strength of the member.

- the comparison between experimental and theoretical results highlights that the proposed

model provides good estimates of the experimental results for low and high aspect ratios. The

accuracy and reliability of the method has been proven on many columns with different

geometric and mechanical properties.

- the comparison with some simplified methods available in the literature demonstrates the

higher reliability of the proposed method.

The Authors underline that the described method can be applied to columns with other cross-sections.

This extension is not difficult to accomplish and only needs that the relations considered in this paper

be particularized to the geometry of the new cross-section.

REFERENCES

Ang Beng Ghee, Priestley M.J.N., Paulay T. 1985. Seismic shear strength of circular bridge piers.

Report 85-5, Department of Civil Engineering, University of Canterbury, Christchurch, New

Zealand.

Ang B.G., Priestley M.J.N., Paulay T. 1989. Seismic shear strength of circular concrete columns.

ACI Structural Journal; 86(1): pp.45-59

Ascheim M.A., Moehle J.P. 1992. Shear Strength and Deformability of RC Bridge Columns

Subjected to Inelastic Cyclic Displacement. Report n° UCB/EERC 92/04, Earth. Eng. Research

Center, University of California at Berkeley, USA.

Berry M., Parrish M., Eberhard M. 2004. PEER Structural Performance Database. Pacific Earth.

Eng. Research Center.


Calderone A.J., Lehman D.E., Moehle J.P. 2000. Behavior of Reinforced Concrete Bridge Columns

Having Aspect Ratios and Varying Lengths of Confinement. PEER Report 2000/08.

Chai Y.H., Priestley M.J.N., Seible F. 1991. Seismic Retrofit of Circular Bridge Columns for

Enhanced Flexural Performance. ACI Structural Journal; 88 (5): pp. 572-584

Eurocode 2. Design of concrete structures – Part 1-1: general rules and rules for buildings. European

Committee for Standardization, 1993.

Fib 2009. Structural Concrete Textbook on behaviour, design and performance – DCC Siegmar Kästl

e K., Germany – ISSN 1562-3610

Hose Y.D., Seible F., Priestley M.J.N. 1997. Strategic relocation of plastic hinges in bridge columns,

Structural Systems Research Project, 97/05, University of California, San Diego, La Jolla.

Hsu T.T.C. 1993. Unified Theory of Reinforced Concrete. CRC

Kowalsky M.J., Priestley M.J.N. 2000. Improved analytical model for shear strength of circular

reinforced concrete columns in seismic regions. ACI Structural Journal, 97 (3): pp. 388-396.

Kunnath S.K., El-Bahy A., Taylor A.W., Stone W.C. 1997. Cumulative Seismic Damage of

Reinforced Concrete Bridge Piers. NISTIR 6075, Building and Fire Research Laboratory,

National Institute of Standards and Technology

Lehman D.E., Moehle J.P. 1998. Seismic Performance of Well-Confined Concrete Bridge Columns.

Pacific Earthquake Engineering Research Center, PEER Report 1998/01, pp. 205.

Nielsen M.P., Braestrup M.W., Bach F. 1978. Rational analysis of shear in reinforced concrete

beams. IABSE Proceedings, pp. 1-16.

Nielsen M.P. 1984. Limit analysis and concrete plasticity. Prentice-Hall Series in Civil Engineering,

Englewood Cliffs, New Jersey.

Norme Tecniche per le Costruzioni. Decreto Ministeriale 14/01/2008. Gazzetta Ufficiale n. 29,

February 4th 2008 - Suppl. Ordinario n. 30 (in Italian).

Priestley M.J.N., Seible F., Xiao Y., Verma R. 1994a. Steel jacket retrofitting of reinforced concrete

bridge columns for enhanced shear strength- Part 1: Theoretical consideration and test design.
ACI Structural Journal; 91: pp. 394 – 405.

Priestley M.J.N., Seible F., Xiao Y., Verma R. 1994b. Steel jacket retrofitting of reinforced concrete

bridge columns for enhanced shear strength- Part 2: Test results and comparison with theory.

ACI Structural Journal; 91: pp. 537 – 550.

Priestley M.J.N., Verma R., Xiao Y. 1994c. Seismic shear strength of reinforced concrete columns.

Journal Structural Engineering, ASCE; 120(8): pp.2310-2329.

Priestley M.J.N., Benzoni G. 1996. Seismic Performance of Circular Columns with Low

Longitudinal Reinforcement Ratios. ACI Structural Journal; 93 (4): pp. 474 – 485.

Priestley M.J.N., Seible F., Calvi M. 1996. Seismic design and retrofit of bridges. John Wiley &

Sons: pp. 333-45.

Recupero A., D’Aveni A., Ghersi A. 2003. N-M-V Interaction Domains for Box I-shaped Reinforced

Concrete Members, ACI Structural Journal; 100 (1): pp. 113-119.

Recupero A., D’Aveni, A., Ghersi, A. 2005. N-M-V Interaction Domains for Prestressed Concrete

Beams, Journal of Structural Engineering; 131(9): pp.1413-1421.

SIA 162/1 1989. Betonbau Ergänzende Festlegungen Structures en béton Spécifications

complémentaires. Eingetragene Norm Der Schweizerischen Normen-Vereinigung Snv Norme

Enregistrée de l’Association Suisse de Normalisation, Ausgabe 1989

Stone W.C., Cheok G.S. 1989. Inelastic behavior of full-scale bridge columns subjected to cycling

loading, NIST BSS 166, Building Science Series, Center for building Technology, National

Engineering Laboratory, National Institute of Standard

Turmo J., Ramos G., Aparicio A.C. 2009. Shear truss analogy for concrete members of solid and

hollow circular cross section. Engineering Structures; 31 (2): pp.455-465.

Walther R., Miehlbradt M. 1990. Dimensionnement des structures en béton: Bases et technologie.

Traité de Génie Civil de l'Ecole polytechnique fédérale de Lausanne, vol. 7. Presses

Polytechniques et Universitaires Romandes (PPUR)


Watanabe F., Ichinose T. 1991. Strength and ductility design of RC members subjected to combined

bending and shear. Proceedings of International Workshop on Concrete Shear in Earthquake,

University of Houston, Houston, Texas.

Wong Y., Paulay T., Priestley M.J.N. 1990. Squat circular bridge piers under multi-directional

seismic attack. Report 90-4, Department of Civil Engineering, University of Canterbury,

Christchurch, New Zealand.

Wong Y., Paulay T., Priestley M.J.N. 1993. Response of Circular Reinforced Concrete Columns to

Multi-Directional Seismic Attack. ACI Structural Journal; 90 (2): pp. 180 – 191.

NOTATION

Greek letters

1 , 2 angles which identify the zones F1, F2 and F3 of the cross-section

1* = min  1 ; llim 

*2 = max   2 ; llim 

** = min  1 ;  2 

llim angle  corresponding to the separation line passing through the longitudinal steel

bar with minimum y-coordinate

lim
s  angle corresponding to the separation line passing through the point of the circular

hoop with minimum y-coordinate

 ,  angles subtended by the hoop segment delimited by the x-axis and the separation

lines of the zone F3

 ,  angles subtended by the arc of the distributed longitudinal reinforcement delimited

by the x-axis and the separation lines of the zone F3

 maximum excursion of with respect to the angle 


 angle of inclination of the diagonal stress of the concrete in the web

 angle of inclination of the first crack

 efficiency factor

' angle of inclination of the strut (arch effect type 1)

'' angle of inclination of the strut (arch effect type 2)

 longitudinal reinforcement per unit of length

1 , 2, 3 equivalent normal stress in F1 , F2 and F3 (truss action)

c1 ,c2 ,c3 normal stress of concrete in the zones F1 , F2 and F3 (truss action)

0c 3 normal stress of concrete in F3 calculated for =0 (truss action)

'c 2 , ''c 2 vertical compressive stress in the zone F2 (arch actions type 1 and 2)

 principal tensile stress developed at first crack on the chord parallel to the neutral

axis and passing through the centroid of the uncracked, homogeneous cross-section

l 1 , l 2 , l 3 normal stress of the longitudinal reinforcement in the zones F1, F2 and F3 (truss action)

l''1 tensile stress in the longitudinal bars of the zone F1 (arch effect type 2)

s3 normal stress of the transverse reinforcement in the zone F3

 tangential stress in the zone F3

Roman letters

A cross-sectional area of the column

A1, A2, A3 area of the zones F1, F2 and F3

Asl area of the longitudinal reinforcement

Asw area of the transverse reinforcement per layer

c mechanical cover of the longitudinal reinforcement

cs mechanical cover of the circular hoops

D diameter of the member cross-section


D’ diameter of the circular hoop

fc1 compressive strength of concrete

fc2 reduced compressive strength of concrete under biaxial state of stress

fck characteristic value of the compressive strength of concrete

fct tensile strength of concrete

fyl yield strength of the longitudinal reinforcement

fyw yield strength of the circular hoops

fywd design yield strength of the transverse reinforcement

L length of the member

LV shear span of the member

M bending moment of the cross-section (truss action)

M ' , M '' bending moment (arch actions type 1 and 2)

M1 , M2 , M3 contributions of zones F1, F2 and F3 to the bending moment by means of truss action

Mexp maximum bending moment recorded during the laboratory test

N axial force of the cross-section (truss action)

N ' , N '' axial force (arch actions type 1 and 2)

N1, N2, N3 contributions of zones F1, F2 and F3 to the axial force produced by truss action

N1'' , N 2'' contributions of zones F1 and F2 to the axial force produced by the arch action type 2

Nexp axial force applied to the column during the laboratory test

rl  R ( R  c)

rs  R ( R  cs )

R radius of the member cross-section

Rl radius of the circle passing through the centroids of the longitudinal steel bars

Rs radius of the circular hoops

RV = Vexp/ Vnum
s spacing of the transverse reinforcement

V shear force of the cross-section (truss action)

V ' , V '' shear force (arch actions type 1 and 2)

VcRd design strength of the concrete shear-resisting mechanism

Vexp maximum shear force recorded during the laboratory test

Vnum ultimate shear force predicted by means of the proposed method

VRd design shear strength

VsRd design strength of the truss mechanism

xn depth of the compression zone

yGc 2 y-coordinate of the centroid of the area of concrete in the zone F2

yGl 1 y-coordinate of the centroid of the area of the longitudinal reinforcement in the zone F1

z internal lever arm

z0 distance from the centroid of the tensile forces to the centre of the cross-section
Table 1. Laboratory tests
Reference Label Setup (1) D L/D fyl fyw fc1 l s M/VD Nexp Vexp (2)
Vnum (3)
Vnum
2 2 2
(cm) (kN/cm ) (kN/cm ) (kN/cm ) (%) (%) (kN) (kN) (kN) (kN)
Ang et al. (4) 6 SB 60 1.5 43.6 32.8 3.01 3.20 0.51 1.50 0 392 171 296
Priestley et al. (5) C7A DB 183 3.0 46.9 32.4 3.07 2.53 0.17 1.51 592 787 246 600
Priestley and Benzoni (6) 2 DB 183 3.0 46.2 36.1 3.00 1.04 0.17 1.52 503 579 260 422
Ang et al. (4) 19 SB 60 1.5 43.6 32.6 3.44 3.20 0.38 1.52 432 437 134 384
Priestley and Benzoni (6) 1 DB 183 3.0 46.2 36.1 3.00 0.52 0.28 1.52 503 393 226 285
Ang et al. (4) 12 SB 60 1.5 43.6 32.8 2.86 3.20 1.02 1.53 359 526 340 403
Ang et al. (4) 18 SB 60 1.5 43.6 32.6 3.50 3.20 0.51 1.53 440 505 179 402
Ang et al. (4) 20 SB 70 1.8 48.2 32.6 3.67 3.20 0.38 1.80 807 487 157 428
Ang et al. (4) 21 SB 80 2.0 43.6 32.6 3.32 3.20 0.38 1.99 0 271 156 247
Ang et al. (4) 1 SB 80 2.0 43.6 32.8 3.75 3.20 0.51 1.99 0 321 207 263
Ang et al. (4) 24 SB 80 2.0 43.6 31.0 3.31 3.20 0.77 1.99 0 341 239 266
Ang et al. (4) 23 SB 80 2.0 43.6 33.2 3.23 3.20 0.76 2.00 0 333 240 266
Ang et al. (4) 2 SB 80 2.0 29.6 32.8 3.72 3.20 0.51 2.00 0 219 170 201
Ang et al. (4) 4 SB 80 2.0 43.6 31.6 3.06 3.20 0.51 2.00 0 289 198 246
Ang et al. (4) 22 SB 80 2.0 43.6 31.0 3.09 3.20 0.39 2.00 0 285 148 237
Ang et al. (4) 15 SB 80 2.0 43.6 32.6 3.48 1.92 0.51 2.00 0 230 154 181
Ang et al. (4) 5 SB 80 2.0 43.6 32.8 3.11 3.20 0.76 2.00 0 331 243 267
Ang et al. (4) 14 SB 80 2.0 42.4 32.6 3.37 3.24 0.51 2.00 0 316 204 251
Ang et al. (4) 7 SB 80 2.0 44.8 37.2 3.01 3.20 0.38 2.00 0 281 228 257
Priestley et al. (5) C5A DB 244 4.0 46.9 32.4 3.59 2.53 0.17 2.01 592 610 238 524
Priestley et al. (5) C1A DB 244 4.0 32.4 35.9 3.10 2.53 0.17 2.02 592 567 265 454
Priestley et al. (5) C3A DB 244 4.0 32.4 32.4 3.45 2.53 0.17 2.04 1779 718 344 572
Ang et al. (4) 16 SB 80 2.0 43.6 32.6 3.34 3.20 0.51 2.04 420 352 232 325
Ang et al. (4) 11 SB 80 2.0 44.8 37.2 2.99 3.20 0.51 2.08 751 407 266 355
Ang et al. (4) 13 SB 80 2.0 43.6 32.6 3.62 3.20 1.02 2.09 455 436 294 344
Vu et al. (7) NH4 DB 183 4.0 46.8 43.4 3.50 5.21 2.70 2.10 850 905 638 673
Wong et al. (8) 2 SB 80 2.0 47.5 34.0 3.70 3.20 0.47 2.11 1813 489 229 442
Ang et al. (4) 8 SB 80 2.0 44.8 37.2 2.87 3.20 1.02 2.12 721 445 311 357
Ang et al. (4) 10 SB 80 2.0 44.8 33.2 3.12 3.20 1.02 2.12 784 437 299 361
Wong et al. (9) 1 SB 80 2.0 42.3 30.0 3.80 3.20 1.42 2.14 907 461 328 379
Vu et. al (7) NH3 DB 183 4.0 42.8 43.0 3.94 2.41 1.14 2.15 970 510 365 402
Wong et al. (8) 3 SB 80 2.0 47.5 30.0 3.70 3.20 1.42 2.15 1813 579 380 444
Vu et.al. (7) NH1 DB 183 4.0 42.8 43.0 3.83 2.41 1.14 2.17 1928 535 421 467
Vu et al. (7) NH6 DB 183 4.0 48.6 43.4 3.50 5.21 3.04 2.23 1914 957 674 710
Ang et al. (4) 3 SB 100 2.5 43.6 32.8 3.60 3.20 0.51 2.50 0 276 196 222
Ang et al. (4) 17 SB 100 2.5 43.6 32.6 3.43 3.20 0.51 2.56 431 312 219 273
Ang et al. (4) 9 SB 100 2.5 44.8 37.2 2.99 3.20 1.02 2.75 751 364 252 285
Sritharan et al. (7) IC3 SB 180 3.0 46.1 43.4 3.30 1.92 0.81 3.14 400 433 318 337
Sritharan et al. (7) IC2 SB 180 3.0 44.8 43.1 3.46 1.92 0.54 3.14 400 411 300 331
Sritharan et al. (7) IC1 SB 180 3.0 44.8 43.1 3.14 1.92 0.54 3.17 400 387 295 323
Stone and Cheok (7) 2 SB 457 3.0 47.5 43.5 3.43 1.99 1.49 3.21 4450 2968 2355 2454
Calderone et al. (10) 328 SB 183 3.0 44.1 60.7 3.45 2.73 0.89 3.22 454 525 425 442
Stone and Cheok (11) N4 SB 75 3.0 44.6 44.1 2.44 1.98 1.41 3.24 120 63 57 59
Stone and Cheok (11) N5 SB 75 3.0 44.6 44.1 2.43 1.98 1.41 3.32 239 77 62 64
Stone and Cheok (11) N1 SB 75 3.0 44.6 44.1 2.41 1.98 1.41 3.39 120 59 55 56
Stone and Cheok (11) N2 SB 75 3.0 44.6 44.1 2.31 1.98 1.41 3.45 239 73 60 62
Lehman and Moehle (7) 407 SB 244 4.0 46.2 60.7 3.10 0.75 0.70 4.23 654 172 151 154
Lehman and Moehle (7) 415 SB 244 4.0 46.2 60.7 3.10 1.49 0.70 4.32 654 269 226 231
Lehman and Moehle (12) 430 SB 244 4.0 46.2 60.7 3.10 2.98 0.70 4.32 654 448 364 380
Henry et al. (7) 415s SB 244 4.0 46.2 60.7 3.72 1.49 0.35 4.53 654 259 214 226
Henry et al. (7) 415p SB 244 4.0 46.2 60.7 3.72 1.49 0.70 4.92 1308 277 239 245
Kunnath et al. (13) A4 SB 137 4.5 44.8 43.4 3.55 2.04 0.94 5.05 222 72 62 63
Kunnath et al. (13) A2 SB 137 4.5 44.8 43.4 2.90 2.04 0.94 5.10 200 74 59 60
Kunnath et al. (13) A5 SB 137 4.5 44.8 43.4 3.55 2.04 0.94 5.24 222 77 59 61
Kunnath et al. (13) A3 SB 137 4.5 44.8 43.4 2.90 2.04 0.94 5.25 200 75 57 58
Stone and Cheok (7) 1 SB 914 6.0 47.5 49.3 3.58 1.99 0.63 6.79 4450 1289 1161 1182
Stone and Cheok (11) N6 SB 150 6.0 44.6 47.6 2.33 1.98 0.68 6.93 120 30 27 28
Chai et al. (14) 3 SB 366 6.0 31.5 35.2 3.26 2.54 0.17 7.05 1779 207 178 196
Stone and Cheok (11) N3 SB 150 6.0 44.6 47.6 2.54 1.98 0.68 7.13 120 32 27 27
Hose et al. (15) SRPH1 SB 366 6.0 45.5 41.4 4.11 2.66 0.89 7.48 1780 285 236 240
Calderone et al. (10) 828 SB 488 8.0 44.1 60.7 3.45 2.73 0.89 9.30 454 172 155 156
Lehman and Moehle (12) 815 SB 488 8.0 46.2 60.7 3.10 1.49 0.70 9.40 654 130 106 107
Calderone et al. (10) 1028 SB 610 10.0 44.1 60.7 3.45 2.73 0.89 12.12 912 157 128 129
Lehman and Moehle (12) 1015 SB 610 10.0 46.2 60.7 3.10 1.49 0.70 12.39 654 80 81 81

Notes: (1) SB=single bending; DB=double bending; (2) proposed method without arch action; (3) proposed method with arch action; (4) Ang et al.
1989; (5) Priestley et al. 1994b; (6) Priestley and Benzoni 1996; (7) Berry et al. 2004; (8) Wong et al. 1993; (9) Wong et al. 1990; (10) Calderone et al.
2000; (11) Stone and Cheok 1989; (12) Lehman and Moehle 1998; (13) Kunnath et al. 1997; (14) Chai et al. 1991; (15) Hose et al. 1997
Table 2. Comparison with simplified methods
Reference Label M/VD RV (1) RV (2) RV (3) RV (4)

Ang et al. 6 1.50 1.33 1.09 1.25 1.06


Priestley et al. C7A 1.51 1.31 0.94 1.22 0.97
Priestley and Benzoni 2 1.52 1.37 1.17 1.17 1.17
Ang et al. 19 1.52 1.14 0.98 1.14 0.96
Priestley and Benzoni 1 1.52 1.38 1.21 1.21 1.21
Ang et al. 12 1.53 1.31 1.16 1.16 1.16
Ang et al. 18 1.53 1.26 1.08 1.19 1.05
Ang et al. 20 1.80 1.14 1.06 1.11 1.05
Ang et al. 21 1.99 1.10 0.97 0.96 1.00
Ang et al. 1 1.99 1.22 1.02 1.00 1.01
Ang et al. 24 1.99 1.28 1.09 1.09 1.09
Ang et al. 23 2.00 1.25 1.08 1.08 1.08
Ang et al. 2 2.00 1.09 1.03 1.03 1.03
Ang et al. 4 2.00 1.17 0.99 0.94 0.99
Ang et al 22 2.00 1.20 1.06 1.05 1.10
Ang et al. 15 2.00 1.27 1.13 1.13 1.13
Ang et al. 5 2.00 1.24 1.06 1.06 1.06
Ang et al. 14 2.00 1.26 1.04 1.03 1.03
Ang et al. 7 2.00 1.09 0.92 0.88 0.89
Priestley et al. C5A 2.01 1.16 0.90 0.95 0.97
Priestley et al. C1A 2.02 1.25 1.07 1.07 1.07
Priestley et al. C3A 2.04 1.26 1.06 1.06 1.06
Ang et al. 16 2.04 1.08 0.99 0.99 0.99
Ang et al. 11 2.08 1.15 1.10 1.10 1.10
Ang et al. 13 2.09 1.27 1.23 1.23 1.23
Vu et al. NH4 2.10 1.35 1.33 1.33 1.33
Wong et al. 2 2.11 1.11 1.11 1.11 1.11
Ang et al. 8 2.12 1.25 1.24 1.24 1.24
           
Mean   1.22 1.07 1.10 1.08
Standard deviation   0.09 0.10 0.11 0.10
COV   0.07 0.09 0.10 0.09

Notes: (1) proposed method with arch action; (2) method by Ang et al.
1989; (3) method by Priestley et al. 1994a; (4) method by Kowalsky and
Priestley 2000
1 2
(a)

F1 F2 y
F3
direction of the
external force

1 2
(b)

y
direction of the
external force
1 2

Figure 1. Identification of the angles (a)  (b)  and 


z z

(a) (b)


direction of the E2 direction of the E4
external force dy external force

A A B  B
E1 dz E3

y y
y1 y3 y2 y1 y3 y2

Section A-A  Section B-B



2R cos

2R cos

E1 y y
E3

axis of
dy the element E1 dy

x x

Figure 2. Representation of the elements E1, E2, E3 and E4


Figure 3. Stress fields in the element E1
Figure 4. Stress fields in the elements (a) E3 and (b) E4
(a) (b) (c)
N'

V' V'' V''

LV '  ''

''

Gc2 Gl1 Gc2 Gc2


N' V' = N' tg ' V'' V''
N'' = -N'' N'' =V''cotg '' N'' = -N'' N'' =V''cotg ''

y y y y y y y y y
D D D

Figure 5. Models proposed for arch actions: (a) type 1;


(b) type 2 for columns subjected to shear forces applied to the lateral surface
and (c) type 2 for columns subjected to shear forces applied to the geometric center
RV

3.0

2.0

1.0

0.0
0 3 6 9 12 15
M/VD

Figure 6. Ratio of the experimental shear force to the shear force


resulting from the proposed numerical approach (only truss mechanism)
RV

3.0

2.0

1.0

0.0
0 3 6 9 12 15
M/VD

Figure 7. Ratio of the experimental shear to the shear force


resulting from the proposed numerical approach (truss and arch mechanisms)

View publication stats

Das könnte Ihnen auch gefallen