Download as pdf or txt
Download as pdf or txt
You are on page 1of 45

SHEAR AND FLEXURE IN

STRUCTURAL CONCRETE BEAMS

Viktor Sigrist, Manuel Alvarez, Walter Kaufmann


Institute of Structural Engineering
ETH Hönggerberg
CH-8093 Zurich, Switzerland

Reprint from
Comité Euro-International du Béton (CEB)
Bulletin d’Information No. 223
“Ultimate Limit State Design Models”
A state-of -art report
June 1995
Preface

The present report was published in CEB Bulletin d’Information No. 223, “Ultimate
Limit State Design Models – A state-of-art report by CEB Task Group 2.3”. It was pre-
pared by Dr. Viktor Sigrist, a member of CEB Task Group 2.3, in collaboration with
Manuel Alvarez and Walter Kaufmann. The three authors are research associates in my
group at the Institute of Structural Engineering at ETH Zurich.
The report provides a comprehensive overview on limit state design procedures for
structural concrete beams, including truss model, discontinuous stress field and failure
mechanism approaches. Such procedures have been developed primarily at the Technical
University of Denmark and at ETH Zurich during the past three decades.
The three authors are involved in a research project entitled “Deformation Capacity of
Structural Concrete”. This project, established in 1991 and directed by myself, has con-
tinually been financed by the Swiss National Science Foundation and the Association of
the Swiss Cement Manufacturers. The support from these organizations is gratefully
acknowledged.

Zurich, December 1995 Prof. Dr. Peter Marti


Shear and Flexure in Structural Concrete Beams

Viktor Sigrist, Manuel Alvarez, Walter Kaufmann,


Institute of Structural Engineering, ETH Zurich
CH-8093 Zurich, Switzerland

1 Introduction

The CEB-FIP Model Code (1990) provides a synthesis of scientific and technical devel-
opments in the analysis and design of concrete structures. The verification of ultimate
limit states requires the determination of the resistance of a structure as well as of its
components based on physical models. These models should represent continuous equi-
librium systems of internal forces under the design ultimate conditions and should
approximately consider compatibility of deformations. For the design of linear members
such models consist of longitudinal chords connected by web lattices comprising diago-
nal concrete struts and reinforcement in one or more directions. For any given cross-sec-
tional geometry and loading configuration the members may be subdivided into several
wall elements which can then be designed for their individual action effects. This proce-
dure for the design of members in shear and flexure reflects the classical truss model
approach.

The idea of using truss models for following the flow of internal forces in reinforced
concrete structures goes back to Ritter (1899). Based on careful observations of the
behaviour of structures as well as a thorough understanding of basic principles of struc-
tural mechanics, Mörsch (1922) greatly advanced this concept and introduced the classi-
cal 45˚ truss model whose simplicity and transparency are striking even today.

The application of the theory of plasticity to structural concrete beams and walls
started in the late 1950s. Basic work in this field was mainly carried out by two groups of
researchers around Nielsen in Copenhagen and Thürlimann in Zurich. Within the frame-
work of limit analysis the truss model concept of Ritter and Mörsch could be generalized
and established on a sound theoretical basis. The idea of discrete trusses was extended to
discontinuous stress fields which were shown to adequately represent the ultimate limit
state behaviour of wall elements such as webs of beams. Furthermore, based on the kine-
matic or upper-bound method failure mechanisms were developed which provide addi-
tional information in determining the load carrying capacity of a structure. While the
kinematic method has been widely used for slab design, it has not received the same
attention for other applications. Comprehensive monographs on the application of limit
analysis methods in the design of reinforced and prestressed concrete structures were
published by Thürlimann et al. (1983) and Nielsen (1984).

1
Discontinuous Stress Fields

This contribution attempts to give an overview on limit analysis methods for struc-
tural concrete beams subjected to shear and flexure and, thereby, to provide background
information on the relevant provisions of CEB-FIP Model Code (1990).

The discontinuous stress fields presented in chapter 2 form the basis of the well-
known truss models. While it is possible to determine internal forces with a truss model,
it is necessary to consider stress fields to check concrete dimensions and to ensure proper
force transfer from the reinforcing bars to the surrounding concrete. For practical design
it is convenient to use fans and parallel stress bands for following the flow of forces,
thereby including all relevant parts of the structure in the load transfer. For complex
structural geometry and loading configurations it may be easier to apply stress fields in
the form of individual struts and ties. While discontinuous stress fields are applicable to
any structural geometry and general loading, simple section-by-section design proce-
dures can be developed for situations in which geometric and static quantities vary only
gradually along the beam axis. The applicability of the method is illustrated by various
design examples including deep beams, continuous girders and discontinuity regions.

Failure mechanisms involving discrete collapse cracks and failure lines are treated in
chapter 3. Basic principles of the kinematic method as well as their application to beams
subjected to shear and flexure are considered. Upper-bound methods are particularly use-
ful for the analysis of existing structures or for checking essential dimensions or details
of a design. Upper-bound methods complement stress field analyses and are therefore
included here although they are not treated in the CEB-FIP Model Code (1990).

2 Discontinuous Stress Fields

2.1 Basic Concepts of Stress Field Analysis

The basic concepts of discontinuous stress field analysis can be explained with the exam-
ple illustrated in Fig. 2.1. A single-span girder with an overhang is subjected to a uni-
formly distributed load q in the span and a single load Q at the free end of the overhang.

A possible stress field solution is shown in Fig. 2.1 (b), for an assumed constant effec-
tive shear depth dv = 620 mm. Note that dv represents the distance between the top and
the bottom chord centres. The inclination of the concrete diagonals within the span is
chosen to ϑ = 41.5˚; in the overhang, ϑ = 44.4˚ is selected. The discontinuous stress field
consists of fans and nodal zones at locations where concentrated loads are introduced, a
fan in the maximum moment region in the span and parallel stress bands in the remaining
parts of the girder. Note that the stress field could be built up merely by fans, i.e. no par-
allel stress bands need to be included. However, to keep the analysis clear it is generally
convenient to subdivide the span into a number of equal portions, which results in at
least part of the structure being modelled by a sequence of parallel bands of compressed
web concrete. The discontinuous stress field describes the state of stress in the web of the

2
Basic Concepts of Stress Field Analysis

girder and it defines the variation of the chord forces. As an illustration the uniformly
compressed band of web concrete ABED in Fig. 2.1 (b) may be considered. The vertical
load q acting along line DE and the vertical stress band BCED (representing uniformly
distributed stirrups) are balanced by the vertical components of the diagonal compressive
stresses in the web concrete ABED; on the other hand, the horizontal components of the
diagonal compressive stresses cause a linear increase of the compressive force in the top
chord of the girder along DE.

The stress field analysis represents an equilibrium solution for the distribution of the
internal forces and yields the reinforcement requirements given in Fig. 2.1 (b). The chord
forces vary linearly along boundaries of stress bands and hyperbolically or parabolically
along fan boundaries, depending on the type of fan, as will be discussed below. The stir-

(a) Q = 360 kN 800


q = 150 kNm-1
180
400
40
A = 540 kN B = 1080 kN 180
8.4 m 2.1 m
110 180 110

1161 948
(b) 626 580
212
-856 -1023 -58 top chord
-374 -1449 -1515 -600 required resistance force [kN]
-1212 -1328 of top reinforcement
-1568

3600 4800 2100


620
ϑ = 41.5˚
ϑ = 44.4˚
D E
ϑ discontinuous
A B C stress field

150 725 4 x 700 400 5 x 700 750 3 x 633 100


300

transverse
86
150
236 stirrup forces
300
450 386 [kN/m]
580 required resistance 536 569
of stirrup reinforcement 686
780

required resistance -1161


of bottom reinforcement -948
-626 -212
-580 bottom chord
374 1212 1023 58 force [kN]
1515 600
856
1449 1328
1568

Fig. 2.1: Single-span beam with an overhang: (a) Geometry and loading; (b) Develop-
ment of a discontinuous stress field and resulting reinforcement requirements.
(Forces in kN, dimensions in m and mm)

3
Discontinuous Stress Fields

rup requirements are represented by a stepped diagram, each step corresponding to the
contribution of the respective vertical stress band.

The magnitude of forces resulting from the discontinuous stress field can be deter-
mined using an equivalent truss model as illustrated in Fig. 2.2. In the figure the free-
body diagram of the segment of the example girder between the left support and the sec-
tion of maximum bending moment is shown. The flanges of the girder form the chords,
the stirrups act as posts and the web concrete provides the compression diagonals. In this
idealization diagonal and vertical truss elements constitute the lines of action of associ-
ated fans and bands. The distributed load q is replaced by statically equivalent single
loads acting on the nodes of the truss. The underlying discontinuous stress field indicates
that the small portion of the load acting directly above the support is transferred to the
support by an additional fan-shaped stress field. The fan at the support and the adjacent
fan overlap somewhat and hence, in a strict sense, the yield criterion may locally be vio-
lated, i.e. the compressive stresses σc may exceed the effective concrete compressive
strength fc (see section 2.3). For practical design these effects have little significance
[Muttoni et al. (1989)]; normally, they can be neglected.

Figs. 2.2 (a) and (b) show two possible truss models for chosen inclinations of the
concrete diagonals ϑ1 = 41.5˚ and ϑ2 = 31.8˚, respectively. The chord forces at the sec-
tion of zero shear force can easily be calculated from the equation 0.5·540 kN·3.6 m /
0.62 m = 1568 kN, independently of the choice of ϑ. The diagrams in Fig. 2.2 (c) repre-
sent the variation of the stirrup and bottom chord forces according to the underlying dis-
continuous stress fields. The chord forces according to the truss models coincide with
those resulting from the discontinuous stress fields at sections where the stirrup require-
ments are stepped.

The necessary cross-sectional areas of longitudinal bars and stirrups, As , can be deter-
mined from the corresponding tensile forces, Ft , by

Ft
A s ≥ ----- , (2.1)
fy

where fy denotes the yield strength of the reinforcing steel. The amount of steel may be
split up in accordance with common design practice: minimum and maximum bar spac-
ings should be taken into consideration and the arrangement should facilitate the placing
of bars and stirrups. From the diagrams in Fig. 2.2 (c) it is clear that the longitudinal bars
in the bottom chord may be curtailed. Using nominal values for development and
anchorage lengths of reinforcing bars as given e.g. in CEB-FIP Model Code (1990), and
considering shear lag effects for bars in flanges, cutoff locations of longitudinal reinforc-
ing bars can be determined.

The web width bw must be checked by comparing the principal compressive stresses
in the web concrete, σc , with the effective (reduced) concrete compressive strength, fc ,

4
Basic Concepts of Stress Field Analysis

(a) 3600

438 712 700 700 700 350 ϑ1 = 41.5˚

11 109 105 105 105 105 1568


-374 -856 -1212 -1449

40
420

315

210

105
75 17 58
8

-6
620
-64

-4 -3 -1

374 856 1212 1449


1568
540
(b) 588 1012 1000 750 250 ϑ2 = 31.8˚

11 154 150 150 75 1568


-502 -1114 -1477
375

225

8 7 18 620
75
29

-71 -42 -1
-7

502 1114 1477


1568
540
ϑ = ϑ2 required resistance
(c) 75 of stirrup
225 reinforcement
366 150
300
[kN/m]
450 ϑ = ϑ1
580
required resistance
374 of bottom
502
856 reinforcement
1212 [kN]
1449
1114
1477 1568

Fig. 2.2: Truss models and discontinuous stress fields for (a) ϑ = ϑ1 and (b) ϑ = ϑ2 ;
(c) Reinforcement requirements.

Fc
– σ c = --------------------------------- ≤ f c (2.2)
b w ⋅ d v cos ϑ
for parallel stress bands, where Fc denotes the value of the compressive forces in the
respective concrete diagonals and bw the web width. For a beam with a constant web
width the stresses in the concrete next to the fan at the support may be used to determine
the corresponding dimension bw . Tests have shown [Cerruti and Marti (1987), Sigrist
and Marti (1993)] that the effective concrete compressive strength fc can be determined
using the procedure outlined in Fig. 2.5 (see section 2.3). Principal compressive stresses
in the fan region may exceed fc but generally need not to be checked.

5
Discontinuous Stress Fields

Fig. 2.2 (c) demonstrates that choosing a smaller value of ϑ results in less stirrup rein-
forcement and in an increase of the longitudinal reinforcement in the bottom chord. Tak-
ing into account the anchorage length of the longitudinal bars the overall requirements of
reinforcing steel are roughly the same in both cases. Therefore, as pointed out by Cerruti
and Marti (1987), the selection of ϑ should primarily be based on practical considera-
tions rather than on an optimization of the total steel amount.

Fig. 2.3 shows the discontinuous stress field and the equivalent truss model for the top
flange corresponding to the stress field given in Fig. 2.2 (b). While the CEB-FIP Model
Code (1990) suggests to determine the transverse reinforcement in the flange from a
simple truss model a general stress field solution is discussed here. The compressive
chord force at the section of zero shear force is assumed to be uniformly distributed over
the flange width, i.e. 1568 kN / 0.8 m = 1960 kN/m. The force flow into the web is ena-
bled by transverse tie forces. With a selected inclination ϑf of the diagonal stress bands
in the flange the variation of the width of the longitudinal stress bands in both halves of
the flange results from equilibrium requirements. In the example, ϑf = 45˚ is assumed.
Consequently, the necessary resistance of the transverse reinforcement can be computed;
it varies with the shear flow into the web as shown in Fig. 2.3 (c). As the shear flow

(a) 150 1025 1000 1000 500

11 154 150 150 75 1568


-502 -1114 -1477
375

225

8 7 18 620
75
29

-71 -42 -1
-7

502 1114 1477


1568
540
897 843 908 477
(b)
1960 kN/m
ϑf
502 612 363 91
251

306

181 800

(c) required resist-


ance of trans-
74
200 191 verse flange
363 reinforcement
485 [kN/m]

Fig. 2.3: Compression flange: (a) Discontinuous stress field for web; (b) Stress field and
truss model for top flange; (c) Reinforcement requirements.

6
Section-by-section Design

between web and flange corresponds to the derivative of the chord force the determina-
tion of the transverse reinforcement requirements is straightforward. Other transverse
distributions of the chord force may be assumed but do not change the principles of the
analysis. The solution described here is valid for infinitely narrow webs whereas for a
web with finite dimensions somewhat smaller transverse reinforcement forces can be
found. In analogy to the procedure presented above stress fields for tension flanges can
be determined as well. Apart from transverse reinforcement requirements such an inves-
tigation yields cutoff locations of the longitudinal bars.

2.2 Section-by-section Design

Discontinuous stress fields are applicable to any structural geometry and to arbitrary
loading configurations. For situations in which all static and geometric quantities vary
only gradually along the beam axis a simple section-by-section design procedure can be
developed. Fig. 2.4 (a) shows the support region of the example girder of Fig. 2.1. To
determine the selected quantities FA, FB and asw · fy at section AB the free-body diagram

(a) 1600 2100

q = 150 kNm-1
Q = 360 kN
FA
A
ϑ
dv = 620
asw·fy
B
FB
B = 1080 kN

dv · cot ϑ
700 750 300 1900 100

(b) (c) VA
q a sw ⋅ f y = 686 kN/m = ---------------------
d v ⋅ cot ϑ

MA ϑ MB M VA
VA F A = – 58 kN = – -------A- + ---------- ⋅ cot ϑ
dv dv 2
VB
M VB
F B = 58 kN = -------B- + ---------- ⋅ cot ϑ
dv 2

Fig. 2.4: Section-by-section design: (a) Free-body diagram with discontinuous stress
field; (b) Free-body diagram and sectional forces; (c) Design equations.

7
Discontinuous Stress Fields

in Fig. 2.4 (a) may be used from which the results given in Fig. 2.4 (c) are found. It
should be noticed that sectional forces are introduced as indicated in Fig. 2.4 (b); they
can be gained from ordinary structural analysis. Fig. 2.4 shows that the flexural moment
M is statically equivalent to the chord forces while the shear force V is assigned to a uni-
form diagonal compressive stress band in the web. The horizontal component V · cot ϑ of
the resultant diagonal compression must be compensated by tensile forces V · cot ϑ / 2 in
the chords, while the vertical component V has to be balanced by the stirrup force asw · fy
within the length dv · cot ϑ.

The expressions given in Fig. 2.4 (c) can be generalized and additional effects of axial
forces N can be considered. This results in
M N V
F bottom = ------ + ----- + ------ ⋅ cot ϑ (2.3)
dv 2 2
for the bottom chord and
M N V
F top = – ------ + ----- + ------ ⋅ cot ϑ (2.4)
dv 2 2
for the top chord. Inserting the above expressions into Eq. (2.1) reinforcement require-
ments can directly be calculated. For the stirrup reinforcement requirements we find the
design equation
V
a sw ⋅ f y ≥ ---------------------- , (2.5)
d v ⋅ cot ϑ
asw being the cross-sectional area of the stirrups per unit length. Stresses in the diagonal
concrete bands and hence web dimensions can be checked by introducing sectional
instead of truss forces in Eq. (2.2) which yields
V
– σ c = --------------------------------------------- ≤ f c . (2.6)
b w ⋅ d v ⋅ sin ϑ cos ϑ
By introducing appropriate load and strength-reduction factors, Eqs.(2.3) through (2.6)
can be applied for a section-by-section design of constant-depth girders with vertical stir-
rups.

In a strict sense, section-by-section considerations are inadequate for shear design. In


the example shown in Fig. 2.4 the shear force VA has to be balanced by the stirrups
within the distance dv · cot ϑ from section A. Generally speaking, for beams that are uni-
formly loaded on the top the stirrup reinforcement required within a length dv · cot ϑ may
be determined by using the lowest value of V that occurs within this length. This so-
called staggering effect [Marti (1986a)] will be discussed in section 2.7.

2.3 Effective Concrete Compressive Strength

In the stress fields presented above it is assumed that the diagonally compressed web
concrete and the uniformly tied stress bands of the stirrups are superimposed. Diagonal
cracking of the web results in a decrease of the concrete compressive strength to a frac-

8
Effective Concrete Compressive Strength

tion of the cylinder strength fck . The effective concrete compressive strength fc should
thus be determined depending on the state of strain in the web. For design practice sim-
plified reduction rules which provide conservative values of fc are given below.

Considering the effect of transverse tension from reinforcement as well as the need to
transmit forces across cracks, the CEB-FIP Model Code (1990) suggests for the effective
concrete compressive strength in cracked zones
f ck
f c = 0.6 ⋅ 1 – ---------  ⋅ f cd , (in MPa units) (2.7)
 250 
where fck and fcd denote characteristic and design values of the concrete cylinder
strength, respectively. In Eurocode 2 (1992) a similar expression is given,
f ck
f c = 0.7 ⋅ 1 – ---------  ⋅ f cd , (2.8)
 140 
and a minimum value of fc ≥ 0.5·fcd may be assumed.

Based on research carried out in Toronto [Collins and Mitchell (1987)] the Canadian
Code (1984) suggests
φ c ⋅ f c′
f c = -------------------------------- ≤ φ c ⋅ f c′ (2.9)
0.8 + 170 ⋅ ε 1
with
2
ε 1 = ε x + ( ε x + 0.002 ) ⋅ cot α , (2.10)
where φ c ⋅ f c′ denotes the factored specified concrete compressive strength and εx the
longitudinal strain at mid-depth of the web. In order to compare these results with those
of Eq. (2.7), fcd instead of φ c ⋅ f c′ may be inserted into Eq. (2.9). In lieu of determining
εx from a strain compatibility analysis (considering actions M, N and V · cot ϑ ) the code

f c ⁄ ( φ c ⋅ f c′ )
1.0

0.8 00
0.0
=
εx 1
.00
0.6 =0 εx
εx dv
2 fc
.00
=0
0.4 εx ϑ

0.2
recommended dv·cot ϑ
range of ϑ
0
15 25 35 45 55
ϑ [˚]
Fig. 2.5: Effective compressive strength of diagonally cracked web concrete.

9
Discontinuous Stress Fields

allows to use a constant value εx = 0.002. Fig. 2.5 shows the effects of ϑ and εx on fc.
Note that for prestressed members and members subjected to axial compression, there is
a significant benefit due to the lower values of εx .

As indicated in Fig. 2.5 a range 25˚ ≤ ϑ ≤ 45˚ is usually recommended for the inclina-
tion of the diagonal compressive stress bands. In the presence of a substantial axial com-
pressive force even lower values of ϑ may be selected. The CEB-FIP Model
Code (1990) therefore allows for a choice of ϑ in the range from 18.4˚ to 45˚. The selec-
tion of ϑ should generally be based on practical considerations since typically, the total
amount of reinforcement is rather insensitive to this parameter. Minimum requirements
for transverse web reinforcement and the need for appropriate anchorage of longitudinal
reinforcing bars practically exclude possible savings that theoretically result from low
values of ϑ. While a lower value of ϑ allows for larger stirrup spacings, which may facil-
itate casting of the concrete, more longitudinal reinforcement is required, which may
cause problems for the anchorage of the reinforcing bars at support. Furthermore, low
values of ϑ may necessitate very large web widths in order to prevent potential web
crushing failures.

2.4 Additional Considerations

As stated in section 2.1, the chord forces vary linearly along boundaries of parallel stress
bands. However, along fan boundaries non-linear force variations occur. Fig. 2.6 (a)
shows a uniformly loaded corbel which may also be considered as the region of an inter-
mediate support of a continuous girder. The geometry of the curve CB of the nodal zone
follows from solving the differential equation of equilibrium for the vertical forces. This
yields the hyperbolic function
a–b
( η + d ’ ) 2 = d ’2 + ξ 2 ------------ + 2ξ ( b + e ) , (2.11)
b
where the notation given in Fig. 2.6 is used. In determining Eq. (2.11) it is assumed that
fc is reached along line CB as well as along lines AB and AC. Note that in the case of a
continuous girder the distributed load p includes the contribution of the stirrups, i.e. the
force asw · fy acting within length a. Moment equilibrium for point F or substituting d’ by
(dv - c/2) in Eq. (2.11) results in the depth of the nodal zone
( a + 2e + b ) ⋅ b
c = ------------------------------------- . (2.12)
2 ⋅ dv
This derivation implies that the effective shear depth dv has been previously estimated,
as it is usually done in design practice. With the calculated values of b and c , minimum
requirements for the dimensions of the bearings and, in the case of an intermediate sup-
port region, for the depth of the bottom flange are obtained.

The decrease of the tensile force along line DE corresponds to the increase of η, i.e.
Ft ( x ) = F D – f c ⋅ bw ⋅ η . (2.13)

10
Additional Considerations

(a) Ft truss model


stress field

x
(b)
p
D D E
p⋅a
FD G b = ----------------
d’ = dv - c f c ⋅ bw
dv 2
d
( a + 2e + b )b
C C F
ξ c = ---------------------------------
c 2 ⋅ dv
A FAC A B

η FAB

bw b e a

Fig. 2.6: Fan action in a uniformly loaded corbel: (a) Variation of the tensile force in the
top chord; (b) Stress field and equivalent truss model.

Using the relation x = ξ · ( fc · bw)/p in Eq. (2.11), η can be expressed as a function of the
distance x from point D. This results in

( a – b )b 2 ( b + e )b
η ( x ) = – d ’ + d ’2 + x 2 -------------------
2
- + x -------------------------- , (2.14)
a a
and hence the force variation plotted in Fig. 2.6 (a) is obtained. In this figure the equiva-
lent truss model is shown as well. In contrast to the stress field analysis the truss model
yields a stepped force diagram.

Generally it is not necessary to determine the variation of the tensile force within the
fan region in detail. Curtailing of the longitudinal reinforcement in accordance with code
requirements for bar anchorage can be based on a rough estimate of the force variation.
Practical considerations are more important than too detailed calculations. For example,
the non-centered fan of Fig. 2.6 can be replaced by a fan centered at point F. As a result,
the concrete compressive stress at point F would tend to infinity (b = 0, c = 0) and the top
chord force would vary parabolically along DE, i.e.

F t ( x ) = F D – ---------- e + ---  .
px x
(2.15)
dv  2

This simplification is rather useful in design practice.

Note that the uniformly distributed load p usually extends right up to the support and
consequently e can then be set to zero in the above equations.

In Fig. 2.7 the support region of a beam without bottom flange is treated. In the nega-
tive moment regions of T-beams or beams with rectangular cross-section the compres-
sive chord forces cannot be transferred to the outer regions of a bottom slab as in the case

11
Discontinuous Stress Fields

q
F F G I

FF ϑ FI
d1 dv1 dv d
c1 FDE asw·fy
E E
C
D c D
H
A FAD A B FH
FAB

bw b a dv·cot ϑ

Fig. 2.7: Negative moment region over support: Fan shaped stress field and nodal zone
of a beam without bottom flange (T-beam).

of hollow-box or I-beams (see Fig. 2.6 (b)). From the free-body diagram shown in
Fig. 2.7 it is apparent that

1
F AD = F H + --- ( q + a sw f y ) cot ϑ ( a + d v cot ϑ ) (2.16)
2
and

F F = F AD + F DE = c ⋅ b w ⋅ f c , (2.17)

where the notation given in the figure is used. The load transferred by the fan ECGF has
to be resisted by the forces FAB and FDE at the nodal zone DCE while the additional
depth (c - c1) has to withstand the bottom chord force FAD . Assuming a certain value of
dv and knowing FAB from overall equilibrium considerations, FH and FAD are readily
determined from Eqs. (2.3) and (2.16). Iteratively, dv can be adjusted until the nodal
zone ABCD fits into the girder and the depth d1 of the fan ECGF is found. Replacing c
and dv in Eq. (2.12) by c1 and (d1 - c1/2), respectively, and setting e = 0 the depth c1 of
the nodal zone CDE can be determined as

c 1 = d 1 – d 12 – b ( a + b ) . (2.18)

Finally, the total depth c of the compression zone and, considering Eq. (2.17), the tensile
force FF are found. The result is in accordance with a direct computation of FF via the
equilibrium condition for the support moment, in which the lever arm (d - c/2) of the
internal forces has to be determined iteratively as well. However, the example of Fig. 2.7
clearly illustrates that only by considering the geometry of nodal zones the proper force
flow can be traced.

For practical reasons, the assumption of a constant effective shear depth dv within the
span of the girder is usually maintained although more general solutions can be devel-
oped, see section 2.5. Further details on nodal zones are given there and in section 2.6.

12
Strut and Tie Action

2.5 Strut and Tie Action

Strut and tie models as well as stress field analyses are appropriate methods for describ-
ing the force flow in discontinuity regions where static, geometric or material conditions
vary abruptly. Compared to general discontinuous stress fields, strut and tie models rep-
resent a simplified approach, since only uniformly compressed struts and biaxially com-
pressed nodal zones are used to model the contribution of the concrete. Starting from a
simple strut and tie model that visualizes the basic load transfer, more sophisticated con-
ceptions of the force flow may be envisaged by incorporating general tools of stress field
analysis: models including fans, arches and bands allow for a proper consideration of
distributed loads and reinforcement. This section attempts to illustrate the connection
between the well known strut and tie models and related stress fields, pointing out the
wide potential of stress field analysis.

Fig. 2.8 shows one half of a symmetric deep beam with a rectangular cross-section
bw · h carrying a single load Q [Nielsen et al. (1978), Marti (1980)]. The load is directly
transferred over the distance a to the support through a single inclined strut. The state of
equilibrium under consideration does not require any transverse ties and since the tensile
strength of the concrete is neglected, the entire tensile force in the longitudinal reinforce-
ment must be anchored behind the support. This can be achieved by means of a rigid
plate as shown in Fig. 2.8 or by providing hairpin-shaped reinforcing bars enclosing the
nodal zone, see section 2.6. Note that the effective depth d of the beam is usually
restricted to a maximum value which is attained, when the distance between the resultant
of the reinforcement and the bottom of the beam equals half the depth of the nodal zone
at the support.

The equilibrium conditions


Fc = c ⋅ bw ⋅ f c = As ⋅ f y = Ft , c ⋅ bw ⋅ f c ⋅ ( h – c ) = Q ⋅ a (2.19 a,b)

yield the following expressions for the unknowns c and As

Q
Q
b = ----------------
Fc c bw ⋅ f c
As ⋅ f y
h - 2c
d h
c = ω ⋅ d , ω = ------------------------
bw ⋅ d ⋅ f c

Ft c Fc = Ft = ω ⋅ d ⋅ bw ⋅ f c
Q
bw
b
a+b

Fig. 2.8: Deep beam with single load: Direct load transfer through strut and tie action.

13
Discontinuous Stress Fields

2
h h Q⋅a fc
c = --- – ----- – ---------------- , A s = c ⋅ b w ⋅ ----- , (2.20 a,b)
2 4 bw ⋅ f c fy
where fc denotes the effective concrete compressive strength. The depth of the compres-
sion zone, c, is usually expressed as a fraction of the effective depth d of the beam
fy fy As
c = ω ⋅ d = ρ ⋅ ----- ⋅ d , with ω = ρ ⋅ ----- and ρ = -------------- , (2.21 a,b,c)
fc fc bw ⋅ d
where ω and ρ are the so-called mechanical and geometrical reinforcement ratios,
respectively. Substituting c = ω · d and h / d = 1 + ω / 2 in Eq. (2.20 a) one gets
2
bw ⋅ f c ⋅ h ω ⋅ ( 1 – ω ⁄ 2 )
Q = -------------------------- ⋅ --------------------------------
2
- for ω ≤ 2 ⁄ 3 , (2.22)
a (1 + ω ⁄ 2)
and
2
bw ⋅ f c ⋅ h
Q = -------------------------- for ω ≥ 2 ⁄ 3 . (2.23)
4⋅a
As illustrated in section 3.3, mechanisms that are compatible with the strut and tie model
in Fig. 2.8 can be found. Therefore, the above expressions render the exact values of the
ultimate load Qu according to limit analysis. Note that Qu reaches a maximum for ω =
2/3 and cannot be increased by a further rise of ω.

The flow of internal forces in the uniformly loaded deep beam shown in Fig. 2.9 can
be visualized either by strut and tie action or by means of the more elaborated associated
discontinuous stress fields [Marti (1985), (1980)]. Since no transverse reinforcement is
provided a direct load transfer to the supports is mandatory for equilibrium and no allow-
ance can be made for curtailment of the longitudinal reinforcement. Depending on the
slenderness of the beam, the reinforcement ratio and the loading history, fan or arch
action will prevail. Formulating equilibrium of the sectional forces Fc and Ft yields the
expression
1–ω⁄2
2 ⋅ h ⋅ b w ⋅ f c ⋅ ω ⋅ ---------------------------2- = q ⋅ a ⋅ 1 – ---------------- 
2 2 q
(2.24)
 bw ⋅ f c
(1 + ω ⁄ 2)
for the determination of ω. Points A through E in all four figures coincide and the nodal
zones ABC are all biaxially compressed to fc . In Figs. 2.9 (a) and (c) the distributed load
q is replaced by two statically equivalent single loads q a / 2 which are transferred to the
supports by struts, where they are balanced by the support reactions and the tie force.
The transition to the fan-shaped stress field shown in Fig. 2.9 (b) is achieved by subdi-
viding the span into differential elements dx and considering infinitely thin struts carry-
ing loads q dx whose ends are bounded by the nodal zone ABC and the compression
zone DEF. Formulating equilibrium for the struts results in two differential equations,
whose solutions provide the functions for the curves AC and DF, which are found to be
quadratic parabolas. The principal compressive stresses in the fan vary hyperbolically
along their straight trajectories. The principal stresses in the compression zone DEF are

14
Strut and Tie Action

(a) (b)
a/4 a/2 a/4 a
qa / 2 qa / 2 q
E F E
Fc ω d Fc
D D

d h
h - 2ω d

A A

Ft ωd Ft
G
B C B C
qa qa
bw
qa / (bw fc) qa / (bw fc)

(c) (d)
a/4 a/2 a/4 a
qa / 2 qa / 2 q
E F E
Fc ω d Fc
D D

d h
h - 2ω d

A A

Ft ωd Ft
G
B C B C
qa qa
bw
qa / (bw fc) qa / (bw fc)

Fig. 2.9: Uniformly loaded deep beam without transverse reinforcement. Strut and tie
models and discontinuous stress fields: (a) and (b) Fan action; (c) and (d) Arch action.

σ1 = -q/bw and σ2 = -fc . A similar procedure leads to the formula of the two quadratic
parabolas AE and CD shaping the arch in the discontinuous stress field shown in
Fig. 2.9 (d). The principal compressive stresses in the arch are 0 ≥ σ1 ≥ -q/bw and
σ2 = -fc ; σ1 varies hyperbolically along the straight trajectories perpendicular to the
parabola CD. The area AEF is uniaxially compressed to σ2 = -q/bw.

Bending mechanisms with rotational centre in point D and one or more collapse
cracks within the area CDG are compatible with the stress fields in Figs. 2.9 (b) and (d).
Therefore, the following expressions, derived from Eq. (2.24),

bw ⋅ f c  8 ⋅ h ω ⋅ ( 1 – ω ⁄ 2 )
2
q = ---------------- ⋅ 1 – 1 – ------------
- ⋅ --------------------------------- for ω ≤ 2 ⁄ 3 , (2.25)
2  a
2
(1 + ω ⁄ 2) 
2

and

bw ⋅ f c  2⋅h 
2
q = ---------------- ⋅ 1 – 1 – -------------  for ω ≥ 2 ⁄ 3 , (2.26)
2  a 
2

provide the exact values of the ultimate load qu according to limit analysis.

15
Discontinuous Stress Fields

(a) (b)
a a
Q Q
H
F E F E
Fc ω d G Fc
D D

d h
h - 2ω d

A
A
I J
Ft ωd Ft
C’
B C B C
Q Q
bw
b = Q / (bw fc) b = Q / (bw fc)

(c) Q (d) Q
H
F E F E
G Fc ω d G Fc
D D

d h
h - ω d - ca

A A I J
Ft ca Ft
B C B C
Q Q
bw

(e) Q (f) Q
F E F E
H Fc ω d G Fc
G
D D
K
d h
h - ω d - ca
H

A A K
I I J
Ft ca Ft
B C B C
Q Q
bw

(g) (h)
x x
(c) (d)

(e) (f)

(a) (b)
Ft Ft

Fig. 2.10: Deep beam with transverse reinforcement. Strut and tie action and discontinu-
ous stress fields for given geometry bw , h, a and factored load Q : (a) Direct load transfer;
(b) through (d) Fan action; (e) and (f) Combined arch and fan action; (g) and (h) Varia-
tion of longitudinal tensile forces for the different models.

16
Strut and Tie Action

Figs. 2.10 (a) through (f) show a deep beam with transverse reinforcement carrying a
single load Q over an effective length a. The support and load plates in all examples have
the same width b = Q / (bw fc), i.e the illustrated states of equilibrium result in the same
lower-bound value of the ultimate load.

The force flow envisaged by the strut and tie model in Fig. 2.10 (a) ignores the pres-
ence of shear reinforcement and therefore requires that the longitudinal reinforcement
must be fully anchored behind the supports. On the other hand, the strut and tie models in
Figs. 2.10 (c) and (e) as well as the related stress fields in Figs. 2.10 (d) and (f) and the
stress field in Fig. 2.10 (b) allow for a curtailment of the longitudinal reinforcement as
illustrated in Figs. 2.10 (g) and (h). In analogy to Fig. 2.9 fan action and combined arch
and fan action are shown in Figs. 2.10 (c), (d) and (e), (f). Note that the quantities d, c, ca
and hence ω = c / d vary from the examples (a), (b) to (c), (d) and (e), (f).

The stress fields in Figs. 2.10 (b) and (d) may be compared with that in Fig. 2.7: the
section at the intermediate support in Fig. 2.7 is analogous to the section at midspan in
the examples of Fig. 2.10. In Fig. 2.7 a constant effective shear depth dv is assumed
although a horizontal chord of infinite dimensions represents a simplified model. On the
other hand, the stress fields in Figs. 2.10 (b) and (d) take into account the constant web
width of the beam and consequently develop the finite dimensions of the compression
zone and thus consider a variable effective shear depth dv with a maximum value dv = d
at section HI and a minimum value dv = d · (1 - ω /2) at midspan.

Q
H
F E
Fc,1 c1
E’
G

Q
G
E’
Fc, 2 c2
D
asw · fy
A

Ft, a h - (c 1 + c 2 + ca)
B C
Q

I J
Ft, a Ft ca
C

b e

Fig. 2.11: Free-body diagrams of the deep beam shown in Fig 2.10 (d).

17
Discontinuous Stress Fields

Formulating equilibrium of the sectional forces in the free-body diagram shown in


Fig. 2.10 (d)

c ⋅ b w ⋅ f c ⋅ h – --- – ----a-  = Q ⋅ a
c c
Fc = c ⋅ bw ⋅ f c = As ⋅ f y = Ft , (2.27 a,b)
 2 2
in addition with equilibrium requirements for the free-body diagrams in Fig. 2.11
F c, 1 = c 1 ⋅ b w ⋅ f c = c a ⋅ b w ⋅ f c = F t , a (2.28)
a sw ⋅ f y ⋅ ( a – b – e ) = Q = b ⋅ b w ⋅ f c (2.29)

c 1 ⋅ b w ⋅ f c ⋅ ( h – c 1 ) = a sw ⋅ f y ⋅ ( a – b – e ) ⋅  -------------------- + e + --- 
a–b–e b
(2.30)
 2 2
and a condition ensuring that the boundary trajectories CG and GI of the two fans lie on
the same line
e a–b
------------ = -------------- (2.31)
ca ⁄ 2 h – c1
yields the unknown geometric quantities c, c1 = ca, e, As and asw for the deep beam in
Fig. 2.10 (d). In particular, the equation
2 b ⋅ (a + b) h⋅a⋅b
c 1 – 2h ⋅ c 1 + h + ------------------------ ⋅ c 1 – ------------------ = 0
3 2
(2.32)
 4  2
is obtained for c1 = ca, which corresponds to the fraction of the tensile force at midspan
that has to be anchored behind the support.

The quantities for the examples shown in Figs. 2.10 (b) and (f) can be determined
considering analogous free-body diagrams as the one depicted in Fig. 2.11 resulting in
expressions similar to Eqs. (2.27 a,b) through (2.32). In particular, substituting
(h - c /2 - ca /2) by (h - c) in Eq. (2.27 b) and (h - c1) by (h - c /2 - c1 /2) in Eqs. (2.30)
and (2.31) leads to the solution of the example shown in Fig. 2.10 (b). Replacing the
right hand sides of Eqs. (2.29), (2.31) and (2.30) by ∆Q = b’ · bw · fc , ca / b and
bw · fc · [b’ · (a + e) /2 + (b - b’) · (a - b’ /2)], respectively, yields the unknown quantities of
the example in Fig. 2.10 (f); b’ denotes the distance GE’ in the free bodies analogous to
the ones in Fig. 2.11 and ∆Q represents the portion of the load Q transferred by the fan
DGIJ and can freely be chosen.

The fans DGIJ in Figs. 2.10 (b), (d) and (f) are bounded by the curved line DG and
the straight line IJ. Curve DG is a hyperbola as illustrated in the example corresponding
to Fig. 2.6. The fans ACGH (AC’GH) correspond to the type shown in Fig. 2.9 (b) and
hence both border lines AC (AC’) and GH are quadratic parabolas. The compressive
stresses in the fans are balanced along IJ by uniformly distributed vertical stirrup forces
asw · fy and by bond stresses which reduce the tensile forces in the longitudinal reinforce-
ment; along GH they meet the principal stresses σ1 = - asw · fy / bw and σ2 = - fc of the com-
pression zone GFH.

The location of the vertical ties in Figs. 2.10 (c) and (e) and that of the resultants of
the uniformly distributed tensile forces of the fans in Figs. 2.10 (d) and (f) coincide, i.e.
the vertical ties bisect the length IJ. From the stress fields in Figs. 2.10 (b), (d) and (f) it

18
Strut and Tie Action

can be seen that the tensile stresses connecting both fans or the fan and the arch, respec-
tively, extend over a length smaller than (a - b); transverse reinforcement is needed only
along IJ.

The provision of transverse reinforcement in the examples shown in Figs. 2.10 (b)
through (f) allows the longitudinal reinforcement to be curtailed; the tensile force that
has to be anchored behind the support is smaller than that for direct load transfer. Hence,
the depth of the nodal zone ABC, ca , is smaller than the depth of the compression zone at
midspan, c ; however, ca must always be large enough to accommodate the distributed
reinforcement whose resultant is represented by the horizontal tie. The effective depth of
the beam, d, is bigger than that for direct load transfer, resulting in reduced compressive
and tensile forces at midspan and thus reduced mechanical reinforcement ratios ω = c / d
(Note: d(a),(b) < d(e),(f) < d(c),(d) , c(a),(b) > c(e),(f) > c(c),(d) and ω(a),(b) > ω(e),(f) > ω(c),(d) in
Fig. 2.10). In Fig. 2.10 (b), the longitudinal reinforcement is curtailed, but no use is
made of a possible increase of the effective depth d. Therefore, the required tensile force
at midspan is the same as that for the direct load transfer shown in Fig. 2.10 (a). In
Figs. 2.10 (c), (d) and(e), (f) the tensile forces at midspan are smaller, since the effective
depth d has been maximized in order to be in agreement with the required depth ca of the
nodal zone ABC.

In Figs. 2.10 (e) and (f), a portion of the load Q is transferred by the inclined strut
FHKG and the arch FHKG, respectively. In Figs. 2.10 (c) and (d), the entire force ini-
tially goes to the bottom of the beam where it is tied up to continue its way to the sup-
port. Hence, the force flow in Figs. 2.10 (e) and (f) requires less transverse but more
longitudinal reinforcement than the one in Figs. 2.10 (c) and (d).

The variation of the longitudinal tensile forces along the line IJ can be determined
from Eqs. (2.13) and (2.14) by inserting d’ = d · (1 - ω) and considering that a in these
equations corresponds to the distance IJ in the examples in Fig. 2.10. The quantity p in

(a) (b)
a a

Q Q

Fc ωd Fc

h - ω d - ca

Ft ca Ft
Q Q
Q / (bw fc) Q / (bw fc)

Fig. 2.12: Further development of discontinuous stress field in Fig. 2.10 (d): (a) reduced
length of transversely reinforced region; (b) reduced total amount of transverse rein-
forcement (combined strut and fan action).

19
Discontinuous Stress Fields

Eqs. (2.13) and (2.14) corresponds to the uniformly distributed vertical tensile stresses
acting along IJ, i.e. asw · fy in Fig. 2.11.

The stress fields in Figs. 2.10 (d) and (f) and the equivalent strut and tie models in
Figs. 2.10 (c) and (e) can be adapted to various distributions of transverse reinforcement.
In Fig. 2.10 (d) the portion of the beam along which vertical tensile stress bands act was
selected such that the two fans have a common boundary. As shown in Fig. 2.12 (a), the
model can be adapted to a reduced length of the transversely reinforced region, leading
to separated fans. Combined strut and fan action can be considered as well, see
Fig. 2.12 (b): the direct transfer of a portion of the load Q to the support through a single
inclined strut, located in between both fans, results in a smaller amount of transverse
reinforcement.

The arch in Fig. 2.10 (f) was chosen such that point I coincides with that of
Fig. 2.10 (d). Other transverse reinforcement layouts can be considered also within this
model. In particular, the limit of direct load transfer in Fig. 2.10 (a) is reached by contin-

(a) (b)

(c)
(d)

(f)

(e)

Fig. 2.13: Strut and tie action: (a) Knee joint under closing moment; (b) and (c) Knee
joints under opening moments; (d) and (e) Beam with dapped end; (f) Corbel.

20
Nodal Zones

uously increasing the portion of the load transferred by the arch, thus reducing the arch
curvature and the length of the fan region.

Fig. 2.13 shows further examples of strut and tie models for situations with static
and/or geometric discontinuities. Again, the models could be refined by incorporating
fans, arches and bands in order to properly simulate the presence of distributed reinforce-
ment. Figs. 2.13 (a), (b) and (c) show knee joints subjected to pure bending moments;
situations with general sectional forces can be treated similarly [Thürlimann et
al. (1983), Muttoni et al. (1989)]. Compared to Fig. 2.13 (b) the relatively weak diagonal
reinforcement in Fig. 2.13 (c) results in benefits for the anchorage of the vertical and lon-
gitudinal reinforcing bars; however, the corresponding tensile forces are increased, since
the lever arm of the internal forces is slightly reduced. Figs. 2.13 (d) and (e) show two
possible solutions for the support region of a dapped-ended beam; similar to knee joints,
solutions with diagonal reinforcement may be envisaged [Schlaich, Schäfer (1984)].
Finally, Fig. 2.13 (f) illustrates the case of load transfer in a corbel.

2.6 Nodal Zones

The force flow envisaged by a stress field can only develop if the nodal zones are capa-
ble of transferring the applied forces. Therefore, the dimensions of the nodes must be
checked and the reinforcement has to be properly detailed.

Consider the three intersecting struts shown in Fig. 2.14 (a). The compressive forces
FA, FB and FC are in equilibrium, i.e. the strut axes are intersecting in one point and the
force polygon is closed. For freely chosen strut widths, the stresses σA, σB and σC in the
uniaxially compressed struts will in general be different. The state of stress within the tri-
angular nodal zone defined by the intersection points A, B and C is determined graphi-
cally in Fig. 2.14 (b) [Marti (1985)]: The poles QA, QB and QC of the Mohr’s circles
corresponding to the individual struts are found by drawing lines parallel to the strut axes
through point O. The states of stress at the faces of the nodal zone are determined by the
intersection points SA, SB and SC of lines through the poles QA, QB and QC, parallel to the
sides BC, CA and AB, with the corresponding Mohr’s circles. The three points SA, SB
and SC define the Mohr’s circle for the state of stress within the biaxially compressed
nodal zone. The lines QASA, QBSB and QCSC must all intersect in the pole Q of this circle.
The maximum compressive stress σ2 in the nodal zone will be higher than the highest
compressive stress in the struts if the side of the nodal zone adjoining the strut carrying
the highest compressive stress (AB, σC in the example) is not perpendicular to this strut.
Thus, in the case of different stresses in the struts, the stresses within the nodal zones
have to be checked in order to assure that the maximum compressive stress does not
exceed the compressive strength fc of the concrete.

Figs. 2.14 (c) and (d) depict the case where the strut widths are chosen such that the
stresses in all struts are equal, σA = σ B = σ C = - fc . The principal stresses in the triangular

21
Discontinuous Stress Fields

nodal zone are σ1 = σ 2 = - fc corresponding to a “hydrostatic” state of stress (in a strict


sense, the term “hydrostatic” is not correct since σ 3 = 0 ). Thus, in the case of equal
stresses in all struts the compressive stresses in the nodes will not exceed the stress in the
adjacent struts. The side faces of the nodal zone are perpendicular to the direction of the

(a) (b) t
FC
bC SA
FA B
bA s- C

s- s-
A
s- 2
1 QA
A
C s s s s s s
2 C A B 1
s- B O
bB

QB
SB
FB Q
FC = -s C· bC
SC
FB = -s B· bB
FA = -s A· bA QC

t
(c) (d)
FC
B bC
FA
bA QA
fc
fc
fc -fc s
fc
C SA=SB=SC=Q
A
fc
bB
QB
FC = fc· bC
FB
FB = fc· bB QC
FA = fc· bA

(e) FE (f)
FC
fc
B FD B
FA
D fc
fc
fc
fc
fc FA
fc C fc
C
A A
fc FD fc
FE FC
FB FB
FB FB
FA FA

Fig. 2.14: Nodal zone: (a) Three struts, unequal stresses; (b) Mohr’s circles correspond-
ing to (a); (c) Three struts, equal stresses; (d) Mohr’s circles corresponding to (c); (e)
Splitting of strut C into two struts; (f) Node with tie force.

22
Nodal Zones

adjoining struts and the nodal zone ABC is affine to the polygon of the strut forces FA,
FB and FC.

If the Force FC is substituted by two forces FD and FE, statically equivalent to FC, the
nodal zone ABC in Fig. 2.14 (c) is complemented by the nodal zone ADB as shown in
Fig. 2.14 (e). The resulting nodal zone ADBC is affine to the polygon of the strut forces
FA, FB, FD and FE. Points A, B and C of the nodal zone remain unchanged. This observa-
tion also holds true if strut C is replaced by a series of smaller struts or a continuous fan
radiating from the nodal zone to balance a distributed load. However, the side face BDA
of the node will now be polygonal or curved, respectively, as demonstrated in
section 2.5. In such cases, the dimensions of nodes may be checked on the basis of the
simpler strut and tie model.

In Fig. 2.14 (f) strut A is replaced by an equivalent tie which has to be anchored
behind node ABC. The state of stress in the node remains unchanged if a proper anchor-
age of the tie force is provided which can be achieved by the use of end plates.

A different solution for the anchorage of tie forces is the application of hairpin rein-
forcement shown in Fig. 2.15 [Marti (1980, 1991)]. The concrete forms compressed
shells spanning between the bends of the hairpins. The action can be explained by a truss
model, and it is seen that tensile stresses are necessary to activate the cover concrete. The
placement of strong dowel bars in the bends of the hairpins will assure a good distribu-
tion of the force to be transmitted.

(a) (c)

A B

(b) (d)

Fig. 2.15: Hairpin Reinforcement: (a) Elevation; (b) Plan; (c) and (d) Truss models.

23
Discontinuous Stress Fields

2.7 Continuous Beams and General Loading

Discontinuous stress fields and truss models can easily be applied to continuous beams
and general loading. First of all the support reactions of the hyperstatic system have to be
determined. Usually, a linearly elastic analysis for the uncracked system is performed
which provides an admissible equilibrium solution. Generally, the elastically determined
support reactions differ from the real ones by certain self-equilibrating or residual forces.
Residual forces depend on the entire loading and restraining history and therefore, in
design practice, they can not be predicted. Some redistribution of forces will occur in any
case and must be enabled through a sufficiently ductile behaviour of the structure.
Indeed, support reactions and internal forces may be assumed to redistribute within
rather wide limits provided that dimensioning and detailing ensure the required ductility.

For known support reactions the beam is subdivided into segments bounded by cross-
sections with zero shear force. By means of discontinuous stress fields or truss models an
equilibrium state of internal forces can be found for each segment. The reinforcement is
then dimensioned and detailed in accordance with the chosen force flow and concrete
dimensions are checked based on a reasonable value of the effective concrete compres-
sive strength. The required redistribution of internal forces is ensured by providing a
well distributed minimum reinforcement and by observing limits for the amount of the
main reinforcement. If necessary, e.g. if nodal zone dimensions require an adjustment of
the assumed lever arm of the chord forces, the model may be modified in order to com-
ply with all static and geometric requirements.

This section illustrates the combination of conventional structural analysis and stress
field analysis on the example of a continuous beam subjected to variable loads. First of
all, a linearly elastic analysis is performed resulting in shear force and moment enve-
lopes; these are subsequently redistributed by superimposing a single state of residual
stresses to all relevant elastic actions. Then required resistances of transverse and longi-
tudinal reinforcements are derived and compared to those resulting from a stress field
analysis for individual (relevant) load cases. Finally, full advantage is taken of redistri-
bution of the internal forces by superimposing different restraint actions to different rele-
vant load cases, thus reducing the design requirements.

The hollow-box girder shown in Fig. 2.16 [Marti (1986)b ] has two equal spans,
l , of 24 m, an effective shear depth, dv , of 1.5 m and a web width, bw , of 0.3 m. Each web
of the symmetric cross-section carries a uniformly distributed factored dead load

q = 50 kNm-1
g = 100 kNm-1

A B C
x dv = 1.5 m
l = 24 m l = 24 m

Fig. 2.16: Continuous beam: Geometry and loading.

24
Continuous Beams and General Loading

g = 100 kNm-1 and a factored live load q = 50 kNm-1 which may be placed in the most
critical position.

A linearly elastic analysis results in the expressions

Vmax = g ( 3/8 l - x ) + q (-x 4/ l 3 + 10 x 2/ l - 16 x + 7l ) /16 , (2.33)


Vmin = (g + q) ( 3/8 l - x ) - q (-x 4/ l 3 + 10 x 2/ l - 16 x + 7l ) /16 for 0 ≤ x ≤ l ;

Mmax = g ( 3/8 l x - x 2/2 ) + q ( 7 l x /16 - x 2/2 ) , (2.34)


Mmin = (g + q) ( 3/8 l x - x 2/2 ) - q ( 7 l x /16 - x 2/2 ) for 0 ≤ x ≤ 4 l /5 ,

Mmax = g ( 3/8 l x - x 2/2 ) + q [ x (-x* 4/ l 3 + 10 x* 2/ l - 16 x* + 7l ) /16 - ( x - x* )2/2 ](2.35)


Mmin = (g + q) ( 3/8 l x - x 2/2 ) - q [ x (-x* 4/ l 3 + 10 x* 2/ l - 16 x* + 7l ) /16 - ( x - x* )2/2 ] ,
with x* = ( 5 l 2 - 4 l 3/x )0.5 for 4 l /5 ≤ x ≤ l ;

for the shear force and moment envelopes shown in Fig. 2.17 (b). While either of
Eqs. (2.34) results from a single load case, each of Eqs. (2.33) and (2.35) originates from
different load combinations; this is easily seen from the influence lines for shear force
and moment. In design practice, linear interpolations would probably be adopted instead
of Eqs. (2.33) and (2.35); this simplification becomes the more conservative the bigger
the ratio q / g gets. Note that except for the cross-sections over the supports the extreme
values of shear forces and moments − Vmax and Mmax, and Vmin and Mmin, respectively –
do not result from the same loading condition.

In order to equalize the bending moments at Support B and in the span a residual
moment of 2400 kNm at Support B and the corresponding shear force of 100 kN are cho-
sen to be superimposed on the sectional forces obtained from the linearly elastic analy-
sis. Detailing of the transverse and longitudinal reinforcements is based on the resulting
adjusted shear force and moment envelopes V max (a) , V (a) and M (a) , M (a) . Unless
min max min
more detailed investigations on the deformational capacity of the structure are under-
taken, the degree of moment redistribution should be chosen in accordance with code
provisions, e.g. CEB-FIP Model Code (1990), section 5.4.3.

Fig. 2.17 (a) shows the variation of discontinuous stress fields for the different load
combinations including the chosen restraint actions; the cross-section with zero shear
force shifts between the extreme locations at x = 8.35 m and x = 11.24 m. Choosing
∆ = 2.0 m results in the inclination ϑ = tan-1 ( dv / ∆ ) = 36.9˚ of the diagonal compressive
stress bands.

The dotted line in Fig. 2.17 (c) indicates the required factored resistance of the trans-
verse reinforcement. This staggered diagram is determined by the larger of the minimum
(a) / ∆ and - V (a) / ∆ within each length ∆.
values V max min

As long as loads act on top of the girder the staggered diagram is inscribed to the
envelope V (a) / ∆ , which is constituted by the larger of the values V max
(a) / ∆ and - V (a) / ∆ .
min
Loads acting below the top chord of the girder have to be hung up by means of additional
stirrups resulting in a staggered diagram that circumscribes the corresponding shear

25
Discontinuous Stress Fields

(a)

1.50 m

A B
∆ ∆ ∆ ∆ ∆ ∆ ∆ ∆ ∆ ∆ ∆ ∆ ∆ ∆=2m

-2250
(b) Vmin -1660
-1952
-2000
-1373
-1094
-822 -1000
-558
-303
100 -57
0 [kN]
53
307
572 1000
846
1130 Vmax
1425

-10800
Mmin -10000
-6718

-3560 -5000
-1350 2400

0 [kNm]
1350
2550
4500
3600 5000
5850 5250 Mmax
6600 6750 6300

(c)
0 [kN / m]
122 147
249 274 300
381 406
518 minimum stirrup 542
reinforcement 600
660 682
825
required resistance of 900
971
stirrup reinforcement V (a) / ∆
1200

required resistance of 5600 6000


(d) top reinforcement 4247

2080 3000
required resistance of
bottom reinforcement M (a) / dv 549
0 [kN]
1017 1413
2653
3000
2949
3897 4129
4748 5205 4914 6000
5302 5305

Fig. 2.17: Continuous beam: (a) Discontinuous stress fields corresponding to the
extreme locations of the cross-section with zero shear force; (b) Shear force and moment
envelopes; (c) and (d) Required resistance of transverse and longitudinal reinforcement.

26
Continuous Beams and General Loading

diagram. In the present example, the portion of the dead load g corresponding to the bot-
tom slab and the webs, g’ = 45 kNm-1, requires additional stirrups, thus raising the dotted
staggered diagram by this value. The resulting diagram is shown in Fig. 2.17 (c) as a
solid line. Note that the staggering concept of shear design is based on stress field analy-
sis and directly complies with the request for uniformly distributed stirrups within each
length ∆ . In the central part of the girder requirements for minimum stirrup reinforce-
ments govern the design.

The diagram for the required factored resistance of the longitudinal reinforcement
shown in Fig. 2.17 (d) follows from Eqs. (2.3) and (2.4). Considering diagonal compres-
sive stress bands with an inclination ϑ = tan-1 ( dv / ∆ ), and assuming that the extreme val-
ues of the adjusted shear forces and moments occur simultaneously results in
M max (a) V (a) M max (a) V (a) ∆
Ft = - + --------- ⋅ cot ϑ =
------------- -------------
- + -------------- (2.36)
dv 2 dv 2 dv
for the bottom and
– M min (a) V (a) – M min (a) V (a) ∆
- + --------- ⋅ cot ϑ = ----------------
F t = --------------- + -------------- (2.37)
dv 2 dv 2 dv
for the top reinforcement. For sections closer than a distance ∆ from Support B the above
equations are not applicable since fan-shaped rather than banded stress fields occur
within this zone. Thus, a parabolic variation of the top chord force between the values
(a)
Ft = 4247 kN and - M min , B / dv = 5600 kN may be assumed, see section 2.4.

Eqs. (2.36) and (2.37) provide conservative design values for the longitudinal rein-
forcement since, in general, V (a) and M max
(a) or M (a) , respectively, do not result from
min
the same load combination. A mere combination of extreme moments and related shear
forces may result in unsafe designs, since bigger values than the adopted design values
can occur. As a matter of fact, an “optimum” design would be based on the maximum
values of Eqs. (2.36) and (2.37) evaluated with (possibly adjusted) shear forces and
moments that belong together. In this case, even a combination of shear forces smaller
than V (a) and corresponding moments smaller than Max ( M max(a) , - M (a) ) could govern
min
the design.

Fig. 2.18 (a) shows a discontinuous stress field and the equivalent truss model for the
load combination leading to the moment envelope Mmax within the range 0 ≤ x ≤ 4 l /5.
The inclination of the diagonal compressive struts equals ϑ = 36.9˚ and the support reac-
(a)
tion at the support A amounts to V max , A = 1525 kN. The cross-section with zero shear
(a)
force is located at x = V max, A / (g + q) = 1525 kN / 150 kNm-1 = 10.17 m. The fans on
either side of this section correspond to uniformly distributed tie-up forces along with
parabolic decreases of the chord forces; they enable the transition to the region of paral-
lel stress bands where linear decreases of the chord forces occur in each section. At the
supports centered fans cause a parabolic variation of the tensile chord forces and infinite
compressive stresses at the fan centres which is an allowable simplification for design
purposes, see section 2.4.

27
Discontinuous Stress Fields

(a)

1.50 m

A B
∆ ∆ ∆ ∆ ∆ ∆ ∆ ∆ ∆ ∆ ∆ ∆ ∆ ∆=2m

0 [kN / m]
57.5 45
207.5 182.5
332.5
300
357.5
507.5 minimum stirrup 482.5
reinforcement 632.5 600
657.5
782.5
required resistance of 900
stirrup reinforcement V/∆ 932.5

(b) 1200

6000
(c) required resistance of 4400
3017
top reinforcement 3000
required resistance of
650
bottom reinforcement
0 [kN]
1017 1317
2650 3000
2883
3883 4050
4717 4817
5150 5168 5168 6000

Fig. 2.18: Continuous beam: (a) Discontinuous stress field for the load combination
leading to the moment envelope Mmax within the range 0 ≤ x ≤ 4 l /5 ;
(b) and (c) Required resistance of transverse and longitudinal reinforcement.

Fig. 2.18 (b) shows the staggered diagram for the required factored resistance of the
transverse reinforcement for the considered load combination. The required resistance of
the longitudinal bottom reinforcement plotted in Fig. 2.18 (c) may be compared with that
resulting from Eq. (2.36) (evaluated in Fig. 2.17 (d) ), since both are based on the same
load combination and superimposed restraint action; however, since Eq. (2.36) combines
shear forces and moments that do not belong to same load cases, Figs. 2.17 (d)
and 2.18 (c) differ within the range 0 ≤ x ≤ 4 l /5 by values ( V (a) - |V |) / 2 · cot ϑ at those
sections where discontinuity lines of the diagonal stress bands are intersected by the
chords. Note that the consideration of discontinuous stress fields implies related shear
force and moment values at each cross-section and thus provides slightly reduced values
for the longitudinal reinforcement requirements. However, a safe design for variable
loads demands consideration of all relevant loading conditions. Therefore, combining
extreme shear force and moment values (according to Eqs. (2.36) and (2.37) ), although
resulting from different load positions, represents a safe design practice.

It should be pointed out that the adjusted shear force and moment diagrams in the pre-
vious examples were obtained by superimposing the same residual forces on the elasti-

28
Continuous Beams and General Loading

cally determined actions for all load combinations. Alternatively, different restraint
actions may be assigned to different load cases (although this is not recommended by
CEB-FIP Model Code (1990) ), provided that shear forces and moments resulting from
the same restraint action are superimposed on the load combination under consideration.
In general, such a procedure will lead to more economical designs.

As an example consider Fig. 2.19 (a). The adjusted shear force and moment enve-
(a) and M (a) (solid lines) have been obtained by superimposing different
lopes M max min
restraint actions on the individual load cases constituting the linearly elastic envelopes
(dotted lines, cf. Fig. 2.17 (b) ). Note that the curvatures of the linearly elastic moment
envelopes within the range 4 l /5 ≤ x ≤ l have been modified by superimposing appropri-

-2100
(a) (a) -1804
-2000
Vmin -1517
-1238
-967
-704 -1000
-450
-204
-17
0 [kN]
104
367
638 1000
917
1204 (a)
1500 Vmax

-10000
(a) -7200
Mmax -4400
-5000
-2000
0
0 [kNm]
0
2700 2700
5000
4800 4800
6300 (a)
7200 7500 7200 6300 Mmin

0 [kN / m]
97 147
228
270 300
364 397
503 minimum stirrup 528
reinforcement 600
647 664
803
required resistance of 900
stirrup reinforcement V (a) / ∆ 947
(b) 1200

required resistance of 6000


(c) top reinforcement 4136 4800
2344 3000
required resistance of
bottom reinforcement M (a) / dv 825
0 [kN]
1000 1011
2603 2625
3000
3811 3844
4625 4669
5044 5136 5100 6000

Fig. 2.19: Continuous beam: (b) Adjusted shear force and moment envelopes;
(c) and (d) Required resistance of transverse and longitudinal reinforcement.

29
Discontinuous Stress Fields

ate restraint actions on the different load cases that constitute the same portion of the
envelopes. This enables an utmost subsequent approximation of both curves. The load
combinations leading to the linearly elastic moment envelopes within the range
4 l /5 ≤ x ≤ l also generate the shear envelopes within 0 ≤ x ≤ l. Therefore, adapting the
curvatures of the parabola which form the moment envelopes within 4 l /5 ≤ x ≤ l and
simultaneously joining both parabola at the middle support by superimposing adequate
restraint actions demands for the corresponding adjustments of the shear force envelopes
within 0 ≤ x ≤ l.

The depicted procedure results in the following expressions for the adjusted shear
force and moment envelopes
(a) =
V max g ( l /2 - x ) + q / l ( l - x ) 2/2 + M B(a) / l , (2.38)
(a) = (g + q) ( l /2 - x ) - q / l ( l - x ) 2/2 + M (a) / l
V min B for 0 ≤ x ≤ l ;

(a) =
M max g ( 3/8 l x - x 2/2 ) + x / l [ g l 2/8 + M B(a) ] , (2.39)
(a) = (g + q) ( 3/8 l x - x 2/2 ) + x / l [ (g + q) l 2/8 + M (a) ]
M min B for 0 ≤ x ≤ l ;

where M B(a) denotes the adjusted moment at the intermediate support B and is chosen to
M B(a) = -7200 kNm (see Fig. 2.17 (a) ). Obviously, different adjusted moments in B for
the maximum and minimum envelopes could be considered as well. The values of the
superimposed shear forces ( V max(a) - V (a)
max ) and ( V min - Vmin ) range along the x-axis from
75 to 0 kN and from 75 to 150 kN, respectively; accordingly, the values of the moment at
the intermediate support of the different superimposed moment diagrams
(a)
( M max (a)
, B - Mmax, B ) and ( M min, B - Mmin, B ) vary from 1800 to 0 kNm and from 1800
to 3600 kNm, respectively.

Based on the adjusted shear force and moment envelopes V max (a) , V (a)
min and
(a) , M (a)
M max min according to Eqs. (2.38) and (2.39), respectively, the reinforcement
requirements illustrated in Figs. 2.19 (b) and (c) are determined in the same way as for
the example shown in Fig. 2.17. Again, the dotted lines in the staggered diagram for
shear design drawn in Fig. 2.19 (b) represent the stirrup demand in the case where all
loads act on top of the web, while the solid lines account for additional stirrups to hang
up the portion of the loads acting beneath the top of the web. Comparing
Figs. 2.19 (b),(c) and 2.17 (c),(d) clearly demonstrates that superimposing different
restraint actions on different load combinations generally results in less conservative
designs than applying a single redistribution to all load cases.

30
3 Failure Mechanisms

3.1 General Considerations

The stress field solutions presented in chapter 2 are all based on the static or lower-
bound method of limit analysis. They lead to safe designs and enable a correct detailing
since the flow of the internal forces is followed consistently throughout the structure.

When analysing existing structures or checking reinforcement layouts determined by


computer programs it is often not straightforward to find suitable stress fields. In such
circumstances, the consideration of failure mechanisms based on the kinematic or upper-
bound method of limit analysis is advantageous. It allows to quickly check essential
dimensions and details even in complicated cases. Upper-bounds of the load bearing
capacity may also be used to estimate the degree of conservatism of a stress field design.

The application of the upper-bound method of limit analysis consists in assuming a


certain failure mechanism and determining the corresponding work W, done by the
external forces, as well as the total dissipation, D, which results from yielding of rein-
forcing bars and crushing of the concrete at lines of discontinuity. The most critical
mechanism is the one providing the lowest ratio D/W.

3.2 Dissipation in the Concrete

While the contribution of the reinforcing bars to the total dissipation can easily be calcu-
lated, the calculation of the dissipation in the concrete deserves some further examina-
tion. Only lines of discontinuity for states of plane stress or strain are considered here
since this appears to be sufficient for most practical applications. In the literature, lines
of discontinuity are usually called “failure lines” or “yield lines”. In the following the
term “failure lines” is used since yield lines are commonly associated with slab design.
The term “collapse crack” is used for a failure line with no dissipation in the concrete.

A theoretically correct formulation would involve displacement and strain rates since
only the increments of deformations at failure are of interest. However, in order to avoid
complex formulations the terms displacements and strains rather than displacement and
strain increments are used below.

Consider the discontinuity at point P shown in Fig. 3.1 (a), where the y-axis is normal
to the failure line and the displacement vector ∆ forms the angle α with the x-axis. The
displacement components u and v (in the x- and y-direction, respectively) are assumed to
vary linearly across the narrow failure zone of thickness d. This corresponds to constant
strains within a zone of homogeneous deformation:
∆ ⋅ y ⋅ cos α ∆ ⋅ y ⋅ sin α
u = ---------------------------- ; v = --------------------------- ; (3.1)
d d

31
Failure Mechanisms

(a) y (b) 2/g


1
2

I B Y
P' I
D D /d

a e 2 e e e 1
e
x y
P 2 1
d
a
xIº
A XºI O

Fig. 3.1: Failure line: (a) Narrow zone of homogeneous deformation; (b) Mohr’s circle
of strains.

∂u ∂v ∆ ∂u ∂v ∆
ε x = ----- = 0; ε y = ----- = --- ⋅ sin α; γ xy = ----- + ----- = --- ⋅ cos α . (3.2)
∂x ∂y d ∂y ∂x d
The principal strains ε1 , ε2 can be determined from Fig. 3.1 (b):
∆ ∆
ε 1 = ------ ⋅ ( 1 + sin α ); ε 2 = – ------ ⋅ ( 1 – sin α ) . (3.3)
2d 2d
The principal directions 1 and 2 bisect the angles between the parallel to the discontinu-
ity (I), and the normal to the displacement direction (II). In these so-called slip-line
directions pure shear strains occur. Except for the special case of α = π/2, the principal
strains ε1 and ε2 always have opposite signs.

Letting d tend to zero, a failure line is obtained. Thus, both ε1 and ε2 tend to infinity
while ε3 has a finite value for plane stress and equals zero for plane strain, respectively,
indicating a state of plane strain in the failure line.

The failure criterion usually adopted for concrete is the modified Coulomb yield cri-
terion (Fig. 3.2) consisting of the sliding criterion
τ + σ ⋅ tan ϕ – c = 0 (3.4)
and the tension cutoff
σ – ft = 0. (3.5)
The sliding criterion can be expressed in terms of principal stresses [Chen and Drucker
(1969)]
1 + sin ϕ 2 ⋅ c ⋅ cos ϕ
k ⋅ σ1 – σ2 = f c , where k = --------------------, f = --------------------------- . (3.6)
1 – sin ϕ c 1 – sin ϕ
τ and σ denote the shear and normal stresses acting on an arbitrary plane, fc and ft are the
uniaxial compressive and tensile strengths, and c and ϕ denote the cohesion and the
angle of internal friction, respectively.

In Eq. (3.6), σ1 is the maximum and σ2 the minimum principal stress; by exchanging
indices five similar equations are obtained. Altogether, these equations define an irregu-

32
Dissipation in the Concrete

lar hexagonal pyramid in the principal stress space, the Coulomb yield surface. The pyr-
amid is cut off by three planes corresponding to Eq. (3.5), and the obtained failure
envelope is the modified Coulomb yield surface. Fig. 3.2 (b) shows its projection into
and its intersection with the (σ1 , σ2)-plane.

The values of ϕ, fc and ft are determined from tests; it is found that a constant value
tan ϕ = 0.75 may be used, corresponding to k = 4 and c = fc / 4. More details are given by
Nielsen et al. (1978) and Marti (1980).

For isotropic materials following the associated flow rule (the plastic strain vector at
yield is perpendicular to the yield surface), the dissipation per unit volume in a displace-
ment discontinuity can generally be expressed as follows [Marti (1980)]:
dD = σ ⋅ ε = c' ⋅ cot ϕ' ⋅ ε ( 1 ) where ε ( 1 ) = ε 1 + ε 2 + ε 3 . (3.7)
The values of c’ and ϕ’ correspond to a Coulomb yield surface circumscribing the actual
failure criterion at the state of stress under consideration.

The dissipation in a displacement may also be found by calculating the scalar product
of the stress and strain vectors. Since the term σ3 · ε3 does not contribute to the dissipa-
tion because either ε3 = 0 (plane strain) or σ3 = 0 (plane stress), it is sufficient to consider
the projections of the yield surface and of the strain vectors into the (σ1,σ2)-plane.

Using Eq. (3.3), the dissipation per unit area of the failure line is given by

dD ⋅ d = σ ⋅ ε ⋅ d = --- ⋅ [ σ 1 ⋅ ( 1 + sin α ) – σ 2 ( 1 – sin α ) ] . (3.8)
2
For the case of plane strain, the relevant states of stress in Fig. 3.2 are given by line AB
in the principal stress space (point H in the stress plane) for α = π/2, point B (line HI) for

(a) t (b) s 2
L Plane Stress
e
j Plane Strain
A
J F
e
O s 1
I
c e
K H B
s
e
D C

k· s 1- s 2 = fc
E
fc ft fc fc /4

Fig. 3.2: Modified Coulomb yield criterion: (a) Stress plane; (b) Principal stress space,
σ3 = 0 (plane stress: intersection, plane strain: projection).

33
Failure Mechanisms

ϕ ≤ α ≤ π/2 and line BE (line JL) for α = ϕ. Eq. (3.8) can be rewritten [Chen and
Drucker (1969)] by using the coordinates of point B {ft , k·ft - fc }:
π fc sin α – sin ϕ
ϕ ≤ α ≤ --- : dD ⋅ d = ∆ ⋅  ----- ⋅ ( 1 – sin α ) + f t ⋅ -----------------------------  . (3.9)
2 2 1 – sin ϕ 
For α = ϕ and α = π/2, the dissipation is independent of the tensile and compressive
strengths, respectively:
fc
α = ϕ: d D ⋅ d = ∆ ⋅ ----- ⋅ ( 1 – sin ϕ ) ; (3.10)
2
π
α = --- : dD ⋅ d = ∆ ⋅ f t . (3.11)
2
Note that for the case of plane strain the assumption of an associated flow rule excludes
discontinuities with α < ϕ.

For plane stress Eq. (3.9) remains valid, the states of stress now corresponding to line
AB (point H) for α = π/2, point B (line HI) for ϕ ≤ α≤ π/2, and line BC (line IJ) for
α = ϕ. In addition, the state of stress may now correspond to point C (line JK) for
-π/2 ≤ α ≤ϕ and to line CD (point K) for α = -π/2. The dissipation in these cases is given
by:
π fc
– --- ≤ α ≤ ϕ : d D ⋅ d = ∆ ⋅ ----- ⋅ ( 1 – sin α ) . (3.12)
2 2
If the tensile strength ft is neglected, the yield surface in plane stress is reduced to the
square yield criterion OCDF shown in Fig. 3.2, and Eq. (3.12) gives the dissipation per
unit area of a failure line for all possible values of α, i. e. -π/2 ≤ α ≤ π/2. For α = π/2
(pure opening) no dissipation occurs in the concrete, and, as stated above, the term “col-
lapse crack” is used for such failure lines [Müller (1978)].

3.3 Examples

In accordance with common engineering practice, ft is neglected in the following. Fur-


thermore, only examples with a state of plane stress are treated. Thus, Eq. (3.12) applies
for all possible values of α, i.e. -π/2 ≤ α ≤ π/2.

Fig. 3.3 shows the beam without shear reinforcement treated in section 2.5 [Nielsen
et al. (1978), Marti (1980)]. The beam has a rectangular cross-section of height h and
width bw and the resultant of the longitudinal reinforcement acts at a distance ω · d / 2
from the bottom: d is the effective depth of the beam and ω denotes the mechanical rein-
forcement ratio
As ⋅ f y
ω = ------------------------ . (3.13)
bw ⋅ d ⋅ f c

According to Fig. 2.8 and Eq. (2.22), the strut and tie model given in Fig. 3.3 (a)
yields the following expression for the ultimate load Q:

34
Examples

2
bw ⋅ f c ⋅ h ω ⋅ ( 1 – ω ⁄ 2 )
Q = -------------------------- ⋅ --------------------------------
2
-. (3.14)
a (1 + ω ⁄ 2)
As stated in section 2.5, this expression is valid for ω ≤ 2 / 3 ; for higher reinforcement
ratios, the ultimate load is given by Eq. (2.23).

All three mechanisms shown in Fig. 3.3 are compatible with the strut and tie model,
i.e. deformations and corresponding stresses fulfil the plastic flow rule. Hence they yield
the exact solution for the ultimate load in terms of limit analysis given by Eq. (3.14).

For the translational mechanism of Fig. 3.3 (c) the work done by the external forces
and the dissipation are given by:
π
W = Q ⋅ ∆ ⋅ cos α + β – ---  (3.15)
 2
and
h ⋅ bw fc π
D = ------------- ⋅ ∆ ⋅ ----- ( 1 – sin α ) + f c ⋅ d ⋅ b w ⋅ ω ⋅ ∆ ⋅ sin α + β – ---  . (3.16)
sin β 2  2
The two terms on the right hand side of Eq. (3.16) represent the dissipation in the con-
crete along the failure line BF and the dissipation in the longitudinal reinforcement,
respectively. Equating W = D and using the expression h / d = 1 + ω / 2 one gets:

(a) Q (b)
Q
G F E
w ·d

fc D h d
(1-2w )·d
C D
w ·d
q
A B
a H
Q Q
Q
(c) (d)
Q

y
O q
q
b x
Q Q

Fig. 3.3: Beam without shear reinforcement: (a) Strut and tie model; (b) Bending mecha-
nism; (c) Translational mechanism; (d) Rotational mechanism.

35
Failure Mechanisms

( 1 – sin α ) π ω
------------------------- + 2 ⋅ sin α + β – ---  ⋅ --------------------
h ⋅ bw ⋅ f c sin β  
2 1+ω⁄2
Q = ------------------------ ⋅ ---------------------------------------------------------------------------------------------------- . (3.17)
2 π
cos α + β – --- 
 2
In order to find the lowest upper bound for the ultimate load, this expression has to be
minimized with respect to α and eventually, Eq. (3.14) is obtained. Alternatively, the
value of α can be found by using the fact that for a mechanism compatible with the strut
and tie model, the principal stress direction is equal to the principal strain direction in the
failure line. As stated in section 3.2, this direction bisects the angle between the parallel
to the failure line and the normal to the displacement vector. This means that the dis-
placement is orthogonal to line CD. For high reinforcement ratios, ω ≥ 2 / 3 , the displace-
ment is found to be vertical, α + β = π / 2, and there is no dissipation in the
reinforcement.

The bending mechanism shown in Fig. 3.3 (b) results in Eq. (3.14) independently of
the inclination of the collapse crack DH. This confirms that the reinforcement cannot be
curtailed in a beam without shear reinforcement – an obvious fact from the strut and tie
model point of view.

The failure line of the rotational mechanism in Fig. 3.3 (d) is a hyperbola with the
equation x·y = constant using the coordinate system shown in the figure. The center of
rotation O lies on the straight line through points D and C [Marti (1980)]. If the center of
rotation is moved towards infinity along this line, the hyperbola will degenerate to a
straight line and the translational mechanism of Fig. 3.3 (c) is obtained.

In the following, examples of beams with both longitudinal reinforcement and verti-
cal stirrup reinforcement are treated. If concrete dimensions are chosen such that failure
will be governed by yielding of both reinforcements, collapse crack mechanisms as
shown in Fig. 3.4 with no dissipation in the concrete will occur. Calculation of upper
bounds of the ultimate load for variable inclination β of the collapse crack is straightfor-
ward. The optimum angle β is found to have the same value as the inclination of the cor-
responding stress band for exact solutions. The most critical mechanisms (Fig. 3.4) are
found if the ends of the collapse cracks are assumed at locations of stirrups or at curtail-

(a) (b)
B b
b D
dv dv

D N A

V V

Fig. 3.4: Collapse crack mechanisms: (a) Rotation; (b) Translation.

36
Examples

ing points of longitudinal reinforcement [Marti (1986b)]. Note that ordinary flexural fail-
ure mechanisms are collapse crack mechanisms with vertical failure lines.

If the longitudinal reinforcement is strong, failure will be governed by yielding of the


stirrups and crushing of the web concrete while the longitudinal reinforcement remains
elastic (Fig. 3.5). This type of failure is known as web-crushing mechanism [Braestrup
(1974)]. Since the longitudinal reinforcement is not yielding the displacement at failure
is vertical. Upper bound values of the ultimate load for variable inclination β can be
expressed as:
d v ⋅ bw f c
V = ---------------- ⋅ ----- ⋅ ( 1 – cos β ) + d v ⋅ b w ⋅ cot β ⋅ f c ⋅ ω v ; (3.18)
sin β 2
τ V ( 1 – cos β )
----- = -------------------------- = -------------------------- + ω v ⋅ cot β , (3.19)
fc d v ⋅ bw ⋅ f c 2 ⋅ sin β
The first term on the right hand side of these equations corresponds to the dissipation in
the crushing concrete along the failure line AB in Fig. 3.5 (a); the second term corre-
sponds to the dissipation in the stirrups. τ and ωv denote the nominal shear stress and the
mechanical reinforcement ratio in the transverse direction (stirrups), respectively,
A sw ⋅ f y
ω v = ----------------------- . (3.20)
bw ⋅ s ⋅ f c
The expression (3.19) has to be minimized with respect to β which results in
cos β = 1 – 2 ⋅ ω v (3.21)
and
τ
----- = ωv ⋅ ( 1 – ωv ) . (3.22)
fc

The mechanism of Fig. 3.5 (b) which has often been observed in tests can be thought
of as a series of failure lines similar to that in Fig. 3.5 (a); both mechanisms yield the
same value of the ultimate load.

For low ratios of shear reinforcement the inclination of the failure line is restricted to
h
tan β ≥ --- , (3.23)
a

(a) b (b) b
B
1=D D
dv dv

A
b
V V

Fig. 3.5: Web crushing mechanisms. (a) Discrete failure line; (b) Failure zone.

37
Failure Mechanisms

V V
D
h

b b
_ _
a a
V V

Fig. 3.6: Failure mechanism for beam with low shear reinforcement ratio.

where a denotes the clear distance between the loading plates, as shown in Fig. 3.6. The
minimum upper bound for the ultimate load is given by:
τ 1 a 2 a
----- = ---  1 +  ---  – ---  + ω v ⋅ --- ,
a
(3.24)
fc 2 h h h
which is seen to correspond to Eq. (2.23), found for a beam without shear reinforcement,
supplemented by the contribution of the stirrups:
τ
----- = ---------- + ω v ⋅  --- – ----------  , with
h a h
(3.25)
fc 4⋅a  h 4 ⋅ a
2
Q h
a = a – ---------------- = a – ---------- . (3.26)
bw ⋅ f c 4⋅a

The solutions obtained for web-crushing mechanisms are shown in Fig. 3.7. It is seen
that an increase of the amount of shear reinforcement beyond ωy = 0.5 does not increase
the load carrying capacity.

Failure mechanisms of continuous girders shown in Fig. 3.8 usually involve more
than one collapse crack or web crushing zone since the system is hyperstatic. Collapse

/t fc (3.24)

(3.22), β=π/2
0.5

0.4

0.3 (3.22)

0.2

0.1

w y
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Fig. 3.7: Ultimate load for beams with strong longitudinal reinforcement.

38
Examples

crack mechanisms for exterior and interior spans will usually be similar to those of Figs.
3.8 (a), (b), (d) and (e) where the ends of the collapse cracks have been assumed at loca-
tions of stirrups or curtailment points of the longitudinal reinforcement. Such mecha-
nisms are often critical in the case of insufficient extensions of the longitudinal bars
[Marti (1986b)]. The ultimate load may be considerably reduced compared to cases
where ordinary rotational mechanisms as shown in Fig. 3.9 govern the collapse of the
structure.

The web crushing mechanism of Fig. 3.8 (f) is identical to the examples treated in
Figs. 3.5 and 3.6, while the mechanism of Fig. 3.8 (c) is analogous to the rotational
mechanism shown in Fig. 3.3 (d). Since part OACD is rotating about point O no dissipa-
tion occurs in the top chord. The failure line CD is a hyperbola with the equation
x’·y’ = constant in the coordinate system (x’,y’) indicated in the figure. The orientation of
the coordinate system follows from:

(a)
A B

(b)
A B
y y'

O b
x
(c) D

A B
x' C

(d)
A B

(e)
A B

(f)
A B

Fig. 3.8: Failure mechanisms for continuous girders [Marti (1986b)].

39
Failure Mechanisms

2 ⋅ xC ⋅ d v
tan ( 2β ) = ------------------------------ , (3.27)
2 2 2
xC – xD – d v

and the dissipation in the crushing concrete along the failure line CD for a unit rotation
about O equals [Marti (1986b)]:
f c ⋅ bw  xC ⋅ d v 2
D c = ---------------- ⋅  --------------------------- – d v  . (3.28)
2  2 ⋅ ( cos β ) 2 

If shear failures as well as failures caused by insufficient anchorage of longitudinal


bars can be excluded, ordinary rotational mechanisms as shown in Fig. 3.9 will give
good estimates of the ultimate load. Such mechanisms are in complete analogy to yield
line theory used in slab design. They are especially useful in early design stages; the
determination of moment resistances needed to carry a certain load is straightforward.

Consider the rotational mechanism of a two-span girder subjected to a uniformly dis-


tributed load q (Fig. 3.9). The mechanism consists of two plastic hinges, one at support B
and the other in the span at a distance λ·l from support A. External work and dissipation
are determined by:
1
W = --- ⋅ q ⋅ l , (3.29)
2

D = M u ⋅  --------- + -----------------------  + M u ⋅ ----------------------- ,


(+) 1 1 (-) 1
(3.30)
λ ⋅ l ( 1 – λ ) ⋅ l (1 – λ) ⋅ l
where Mu(+) and Mu(-) are the moment resistances in the span and above support B,
respectively. Setting D = W the upper bound
(+) (-) (+)
2 ⋅ Mu 1 1 + Mu ⁄ Mu 
q u = ------------------ ⋅  --- + -----------------------------------  (3.31)
l
2
λ 1–λ 

A B

l ·l (1-l ) · l

1
-----------------------
A 1 B (1 – λ) ⋅ l

1 1
--------- + -----------------------
λ ⋅ l (1 – λ) ⋅ l

Fig. 3.9: Flexural mechanism of a two-span girder.

40
Examples

of the ultimate load is obtained. The location of the plastic hinge in the span is deter-
mined by minimizing Eq. (3.31) with respect to λ. This results in:
1
λ = ---------------------------------- , (3.32)
(-)
Mu
1 + 1 + -----------
(+)
Mu
which corresponds to the exact solution for the ultimate load since the same expression is
found by a lower-bound investigation.

For the special case of equal moment resistances in the span and at the support, i.e.
(+)=
Mu Mu(-)= Mu, the following expressions are obtained:
1
λ = ---------------- = 2 – 1 ; (3.33)
1+ 2
Mu
q u = ( 6 + 4 ⋅ 2 ) ⋅ -------
2
. (3.34)
l

The ultimate load given by Eq. (3.34) can be shown to differ by less than 3% from the
value obtained by choosing λ = 0.5 (plastic hinge at midspan). This is due to the fact that
the shape of the curve defined by Eq. (3.31) is rather flat near its minimum which is the
case in most practical applications. A good estimate of the ultimate load can thus often
be established without minimization.

To achieve the value of the ultimate load given by Eq. (3.34), elastically determined
moments have to be redistributed by a factor β . For the static system in Fig. 3.9, assum-
ing that there are no initial restraints or residual moments, β is given by:
qy
β = 1 – ----- = 31.4 % , (3.35)
qu
where
8 ⋅ Mu
q y = --------------- (3.36)
2
l
corresponds to the yield load. Note that to reach the ultimate load a certain rotation
capacity of the plastic hinge region is required and it has to be ensured that no premature
failure occurs due to insufficient anchorage of the reinforcing bars or crushing of the web
concrete.

41
Notation

Notation

a, a distances p uniformly distributed load


asw cross-sectional area of stirrups per q live load
unit length qu ultimate load
As cross-sectional area of reinforce- qy yield load
ment Q single load
b width; length of nodal zone Qu ultimate load
b’ portion of b s stirrup spacing
bw web width u displacement in x-direction
c depth of nodal zone; v displacement in y-direction
cohesion of concrete V shear force
ca depth of nodal zone at support V (a) adjusted shear force
d effective depth in flexure; Vmax, Vmin extreme shear forces result-
thickness of failure zone ing from a linearly elastic analysis
d’ = dv - c/2 W work done by external forces
dv effective shear depth x, y coordinates
D dissipation
e distance α inclination of displacement vector
fc effective concrete compressive β inclination of failure line;
strength ratio of moment redistribution

f c specified compressive strength of γ shear strain
concrete ∆ staggering length
fcd design value of fc ∆ displacement vector
fck characteristic value of fc ε normal strain
fct tensile strength of concrete ε strain vector
fy yield strength of reinforcing steel εx strain in x-direction
F force ε1, ε2, ε3 principal strains
Fc compressive force ε(1) = ε1 + ε2 + ε3 , first invariant of
Ft tensile force strain vector
Ft, a tensile force at support η, ξ local coordinates
g dead load ϑ inclination of diagonal compres-
g’ dead load of webs and bottom slab sive stress band
h depth of girder λ = x/l
k = (1 + sin ϕ) / (1 - sin ϕ) ρ geometrical reinforcement ratio
l span length σ normal stress
M bending moment σ stress vector
(a)
M adjusted bending moment σc concrete stress
M B(a) adjusted bending moment at inter- σs steel stress
mediate support σ1, σ2 principal stresses
Mmax, Mmin extreme moments resulting τ shear stress
from a linearly elastic analysis ϕ angle of internal friction
Mu ultimate moment φc strength reduction factor for f c′
M u(+) , M u(-) ultimate moments in the ω mechanical reinforcement ratio
span and over the support ωv mechanical reinforcement ratio of
N axial force stirrups

42
References

References

Braestrup, M. W. (1974). “Plastic Analysis of Shear in Reinforced Concrete.” Magazine of Con-


crete Research, Vol. 26., No. 89, Dec. 1974, pp. 221-228.
Canadian Code (1984). Design of Concrete Structures for Buildings (CAN3-A23.3-M84).
Canadian Standards Association, Rexdale, Ontario, 1984, 281 pp.
CEB-FIP (1990). CEB-FIP Model Code for Concrete Structures. Comité Euro-International du
Béton, Bulletin d’Information No. 213/214, Lausanne, May 1993, 437 pp.
Cerruti, L.M., Marti, P. (1987). “Staggered shear design of concrete beams: large scale tests.” Ca-
nadian Journal of Civil Engineering, Vol. 14, No. 2, 1987, pp. 257-268.
Chen, W. F., Drucker, D. C. (1969). “Bearing Capacity of Concrete Blocks or Rock.” Proceed-
ings, ASCE, Engineering Mechanics Division, V. 95, No. 4, 1969, pp. 955-978.
Collins, M. P., Mitchell, D. (1987). Prestressed Concrete Basics. Canadian Prestressed Concrete
Institute, Ottawa, 1987, 614 pp.
Eurocode 2 (1992). Planung von Stahlbeton- und Spannbetontragwerken. Teil 1, Grundlagen und
Anwendungsregeln für den Hochbau. Europäische Vornorm, 1992, 173 pp.
Marti, P. (1980). Zur plastischen Berechnung von Stahlbeton. Dissertation, Institut für Baustatik
und Konstruktion, ETH Zürich, Bericht Nr. 104. Birkhäuser Verlag, Basel, Oktober 1980,
176 pp.
Marti, P. (1985). “Basic Tools of Reinforced Concrete Beam Design.” ACI Structural Journal,
Vol. 82, No. 1, Jan.-Feb. 1985, pp. 46-56.
Marti, P. (1986 a). “Staggered Shear Design of Simply Supported Concrete Beams.” ACI Struc-
tural Journal, Vol. 83, No. 1, Jan.-Feb. 1986, pp. 36-42.
Marti, P. (1986 b). “Staggered Shear Design of Concrete Bridge Girders.” Proceedings, Interna-
tional Conference on Short and Medium Span Bridges, Ottawa, Aug. 1986, Vol. 1, pp 139-149.
Marti, P. (1991). “Dimensioning and Detailing.” IABSE Colloquium, Structural Concrete, Stutt-
gart, 10-12 April 1991. Colloquium Report, International Association for Bridge and Structural
Engineering, Zurich, pp. 411-443.
Mörsch, E. (1922). Der Eisenbetonbau – Seine Theorie und Anwendung. Verlag Konrad Wittwer,
Stuttgart, 5. Auflage, 1.Band, 2. Hälfte 1922.
Müller, P. (1978). Plastische Berechnung von Stahlbetonscheiben und -balken. Dissertation, In-
stitut für Baustatik und Konstruktion, ETH Zürich, Bericht Nr. 83. Birkhäuser Verlag, Basel,
Juli 1978, 160 pp.
Muttoni, A., Schwartz, J., Thürlimann, B. (1989). Bemessen und Konstruieren von Stahlbeton-
tragwerken mit Spannungsfeldern. Institut für Baustatik und Konstruktion, ETH Zürich, Vor-
lesungsunterlagen, 1989, 134 pp.
Nielsen, M. P. (1984). Limit Analysis and Concrete Plasticity. Prentice-Hall Series in Civil Engi-
neering, Englewood Cliffs, New Jersey, 1984, 420 pp.
Nielsen, M. P., Braestrup, M. W., Jensen, B. C., Bach, F. (1978). Concrete Plasticity. Danish So-
ciety for Structural Sciences and Engineering, Lyngby, 1978, 129 pp.
Ritter, W. (1899). “Die Bauweise Hennebique.” Schweizerische Bauzeitung, Vol. 17, 1899, pp.
41-43, 49-52 und 59-61.
Schlaich, J., Schäfer, K. (1984). “Konstruieren im Stahlbetonbau.” Betonkalender, W. Ernst, Ber-
lin, 1984, pp. 787-1005.
Sigrist, V., Marti, P. (1993). Versuche zum Verformungsvermögen von Stahlbetonträgern. Institut
für Baustatik und Konstruktion, ETH Zürich, IBK-Bericht Nr. 202. Birkhäuser Verlag, Basel,
November 1993, 90 pp.
Thürlimann, B., Marti, P., Pralong, J., Ritz, P., Zimmerli, B. (1983). Anwendung der Plastizitäts-
theorie auf Stahlbeton. Institut für Baustatik und Konstruktion, ETH Zürich, Autographie zum
Fortbildungskurs, März 1983, 252 pp.

43

You might also like