Sunteți pe pagina 1din 79

Chapter 1

STABILITY – GENERAL ASPECTS

1.1. EXAMPLES OF INSTABILITY

In many cases, instability is the most important limit state for structural members
made of steel or aluminium alloys. This phenomenon can affect a part of the
member, the entire member, a part of the structure or the entire structure.

1.1.1. Instability of bars

This is the type of loss of stability that affects bars in compression or compressed
parts of bars. At a certain value of the load, the bar in compression or the
compressed part of the bar (involving the un-compressed part too) finds equilibrium
in a deformed shape, in the neighbourhood of the straight one.

1.1.1.1. Straight bars in compression

This problem is probably the most studied one in the history of stability problems.
Beginning with Euler, 1744 [1], different researchers tried to express the equilibrium
and the failure mode of a perfectly straight member subject to axial compression [2].
When subject to an axial compression force, a straight member may lose its stability
in one of the following forms (Fig. 1.1):
• flexural buckling (v ≠ 0; ϕ = 0) (Fig. 1.1a);
• torsion buckling (v = 0; ϕ ≠ 0) (Fig. 1.1b);
• flexural-torsion buckling (v ≠ 0; ϕ ≠ 0) (Fig. 1.1c);
where v means the lateral displacement in the plane of the cross-section and ϕ is the
rotation of the cross-section in its plane.

Symbols – Widgit Symbols (c) Widgit Software 2002-2013 www.widgit.com

1
Fcr Fcr Fcr

ϕ ϕ

v v

(a) (b) (c)


Fig. 1.1. Forms of buckling of a straight bar in compression [2]

The buckling load is the critical force Fcr at which a perfectly straight member in
compression assumes a deflected position (Fig. 1.1). Buckling is a limit state, in the
meaning that once the force Fcr is reached, the deflection increases until the collapse
of the bar is reached. The member should be subjected only to loads inferior to the
critical force (F < Fcr) [2].

1.1.1.2. Straight bars in bending

In the same way as for any member in compression, the buckling problem appears
for the compressed flange of a beam. Generally, buckling (Fig. 1.2) may not occur in
the plane of the web, as the compressed flange is continuously connected through
the web material to the tensioned part of the cross-section, the tension flange. The
stabilizing effect of the tension zone transforms free transverse buckling into lateral-
torsional buckling, causing lateral bending and twisting of the beam [2].

2
Lateral buckling
of the flange

Torsion
(twisting
of the
beam)
span
Fig. 1.2. Lateral-torsional buckling of a beam [2]

Lateral-torsional buckling may be prevented either by performing checks using


suitable relations, or by introducing lateral bracings whose purpose is to reduce the
distance on which this phenomenon may occur.

1.1.1.3. Curved bars in compression and bending

The arch is a very efficient structural member for resisting symmetrical loads acting
in its plane, normally to the line of its supports. Its important strength comes from the
arm lever exiting between the compressed part (the arch) and the tensioned part (the
line of supports) (Fig. 1.3), which is much bigger than the distance between the
compressed flange and the tension one in the case of a beam.

Fig. 1.3. The working principle of an arch

3
The horizontal stiffness of the supports (the abutment stiffness) is vital for the
performance of the arch; the smaller this one is, the closer the behaviour is to a
“curved beam” and strong values of the bending moment can be found along the
arch.

There are several forms of buckling of an arch (Fig. 1.4):


• in-plane symmetrical (Fig. 1.4a);
• in-plane asymmetrical (Fig. 1.4b);
• out-of-plane (Fig. 1.4c).

(a) (b) (c)


Fig. 1.4. Forms of buckling of an arch

1.1.2. Instability of plates and shells

1.1.2.1. Local buckling of plates

Local buckling of a plate may occur as a result of the action of in-plane normal
compression stresses (σ), of shear ones (ττ), or of their combination (Fig. 1.5).

Fig. 1.5. Local buckling of plates [3]

4
1.1.2.2. Local buckling of shells

Similarly to the case of plates, in-plane compression stresses (σ) can lead to local
buckling of shells (Fig. 1.6).

Fig. 1.6. Local buckling of shells [3]

1.1.3. Instability of structures

1.1.3.1. “Snap-through” buckling

Because of the big values of the angles among bars (Fig. 1.7), strong compression
forces result in the bars. Because of the strains in these bars, the geometry of the
structure changes and, in some circumstances, equilibrium can be found in tension.

Fig. 1.7. “Snap-through” buckling [4]

5
1.1.3.2. Instability of structures containing members in compression

A part of a structure or the entire structure can go unstable, as the end supports or
the joints of members in compression or in compression and bending have limited
stiffness. Figure 1.8 shows the case of a plane frame and of the reticulated structure
of a roof.

Fig. 1.8. Instability of frames or of roof structures [4]

1.1.3.3. Instability of tension structures

Tension members do not generate static instability of structures. However, they can
be subject of dynamic instability of structures like vibrations, resonance, “flutter” etc.
A very well-known example is the failure of Tacoma Narrows Bridge on November 7,
1940 (Fig. 1.9).

Fig. 1.9. Tacoma Narrows Bridge

6
1.2. COMMENTS

Instability is a more severe problem than overcoming strength:


1. Generally speaking, in the case of most strength checks, the
capable loads are expressed based on the yielding limit of the
Important material (except for checks for phenomena that involve brittle
fracture). Compared to the behaviour of the actual member, at
least two conservative aspects can be noticed:
• The nominal yielding limit, used in calculation, is smaller than
the average value (γov = 1,40 – S235; γov = 1,30 – S275; γov =
1,25 – S355 according to P100-1/2013 [5]);
• There is a certain reserve from the yielding limit till the
ultimate strength (about 1,53 – S235; 1,56 – S275; 1,44 –
S355 according to EN 1993-1-1 [6]).
2. In the case of stability checks, the capable loads are expressed
based on the yielding limit and, depending on the values of the
slenderness, failure can occur even before reaching the yielding
limit.
3. Using a material with higher strength increases the strength
capacity but, in some cases, it has no influence on the stability
capacity.
4. In many cases instability is the most important limit state for
structural members.

There are well known situations when the recorded value of the
snow load, for instance, were much higher than the design ones and
no failure was noticed, as no instability problems were involved.

Example Such an example occurred on February 1st 2014 in Austria, when a


snow load of 4,35kN/m2 was measured.

7
1.3. CLASSES OF CROSS-SECTIONS

1.3.1. Definition of the classes of cross-sections

Generally, given the strength of steel and aluminium alloys, failure of a metal
member subjected to loads other than tension occurs by buckling or by local
buckling. Depending on the slenderness of the element, this can happen either in the
elastic range (0 – Y in figure 1.10) or in the plastic range (Y – F in figure 1.10). To
manage this, EN 1993-1-1 [6] defines four classes of cross-sections of structural
members. They are best expressed for members in bending. In these definitions, the
behaviour of the material is presumed perfectly elastic up to the yielding limit and
perfectly plastic for elongations superior to the strain corresponding to the yielding
limit (Fig. 1.10). This model is known as the Prandtl model.

σ
real

Y Prandtl F
fy

0 εy εu ε

Fig. 1.10. The Prandtl model for steel behaviour

Depending on the stress state that causes local buckling, cross-sections of structural
members are classified as [6] (Fig. 1.11):
Class 1 – cross-sections that can form a plastic hinge with sufficient rotation
capacity to allow redistribution of bending moments. Only class 1 cross-
sections may be used for plastic design.
Class 2 – cross-sections that can reach their plastic moment resistance but local
buckling may prevent development of a plastic hinge with sufficient
rotation capacity to permit plastic design (redistribution of bending
moments).

8
Class 3 – cross-sections in which the calculated stress in the extreme
compression fibre can reach the yield strength but local buckling may
prevent development of the full plastic bending moment.
Class 4 – cross-sections in which it is necessary to take into account the effects of
local buckling when determining their bending moment resistance or
compression resistance.

For practical reasons, the limits between these classes are expressed in terms of
slenderness. Tables 1.1, 1.2, 1.3 show the requirements for different cross-sectional
classes. The class of a cross-section is the maximum among the classes of its
components.

z σmax = 0 σmax < fy σmax = fy σmax = fy σmax = fy


fy

(–)
y y x x

(+)

z class 4 class 3 class 2 class 1


z εmax = 0 εmax < εy εmax = εy εy εmax > εy εmax >> εy

y y x x

z
Fig. 1.11. Possible stress distribution, depending on the cross-section class

The plastic hinge is a concept. It is a model of a cross-section where all the fibres
reached the yielding limit in tension or compression (Fig. 1.11) generated by a
bending moment, presuming a bi-linear (Prandtl) behaviour diagram for the material,
while in the neighbour cross-sections the stress state is elastic. In reality, the stress
and strain state is more complex (Fig. 1.12): the material behaviour is not ideally
elasto-plastic and the plastic deformations extend on a certain length.

9
Fig. 1.12. The stresses in the region of a plastic hinge

Tab. 1.1. Limitations for the slenderness of internal walls (EN 1993-1-1 [6] Tab. 5.2)

c c c c
Bending axis
t t t t

t t t
t
c c c
c
Bending axis

Class Wall in Wall in Wall in bending and compression


bending compression
fy fy fy
Stress distribution αc c
c
c fy fy
fy
c 396ε
when α > 0,5: ≤
1 c c t 13α − 1
≤ 72ε ≤ 33ε
t t c 36ε
when α ≤ 0,5: ≤
t α
c 456ε
when α > 0,5: ≤
2 c c t 13α − 1
≤ 83ε ≤ 38ε
t t c 41,5ε
when α ≤ 0,5: ≤
t α

10
fy fy
Stress distribution fy
c c c
fy c/2
ψfy

c 42ε
when ψ > –1: ≤
3 c c t 0,67 + 0,33ψ
≤ 124ε ≤ 42ε
t t
≤ 62ε (1 − ψ) (− ψ)
c
when ψ ≤ –1:
t

235 fy (N/mm2) 235 275 355 420 460


ε=
fy ε 1,00 0,92 0,81 0,75 0,71
Notes:
1. (+) means compression.
2. ψ ≤ –1 applies where either the compression stress σ ≤ fy or the tensile strain εy >
fy/E.
3. “Internal walls” means walls that are stiffened at both ends; it is not only the case
of the webs, but of the central zones of box cross-sections too.

Tab. 1.2. Limitations for the slenderness of flanges (EN 1993-1-1 [6] Tab. 5.2)

t t t t
c c c
c

Compressed Tension and compressed flange


Class flange
Compressed edge Tension edge

Stress distribution αc αc

c c c

1 c c 9ε c 9ε
≤ 9ε ≤ ≤
t t α t α α
2 c c 10ε c 10ε
≤ 10ε ≤ ≤
t t α t α α

11
Stress distribution
c c c

3 c c
≤ 14ε ≤ 21ε k σ
t t

235 fy (N/mm2) 235 275 355 420 460


ε=
fy ε 1,00 0,92 0,81 0,75 0,71
Note: (+) means compression

Tab. 1.3. Limitations for the slenderness of the walls of round tubes (EN 1993-1-1 [6]
Tab. 5.2)

Class Cross-section in bending and/or compression


1 d/t ≤ 50ε2
t
2 d/t ≤ 70ε2 d
3 d/t ≤ 90ε2
fy (N/mm2) 235 275 355 420 460
235
ε= ε 1,00 0,92 0,81 0,75 0,71
fy
ε2 1,00 0,85 0,66 0,56 0,51

In a similar manner, the American code ANSI/AISC 360-16 [7] distinguishes between
compact, noncompact and slender cross-sections. The Japanese code JSCE [8]
makes reference to EN 1993-1-1 [6] and associates compact to class 1 and class 2
cross-sections, noncompact to class 3 and slender to class 4 ones, using its own
limits.

1.3.2. Practical procedure for determining the class of the cross-section

Establishing the class of a given cross-section might generate discussions: except


for the case of round tubes (Tab. 1.3), where it depends only on the slenderness of
the walls, it involves the loading state too. The stress state that defines the limits
between the classes of cross-sections is associated to the resistance ultimate limit

12
state of the cross-section in compression and bending which is defined using the
following relations:
• for class 1 and class 2 cross-sections, the check for bending moment is done
with the following relation:
M Ed ≤ M N , Rd (EN 1993-1-1 [6] rel. (6.31)) (1.1)

where MN,Rd is the design plastic moment resistance reduced due to the axial
force NEd; therefore, to determine the class of the cross-section, one must
establish the stress distribution associated to MN,Rd (Fig. 1.13).
For doubly symmetrical I- and H- sections or other flanges sections, no
reduction of the plastic resistance moment MN,y,Rd about the y-y axis needs to
be done when both the following criteria are satisfied:
N Ed ≤ 0,25 ⋅ N pl, Rd (EN 1993-1-1 [6] rel. (6.33)) and (1.2)

0,5 ⋅ h w ⋅ t w ⋅ f y
N Ed ≤ (EN 1993-1-1 [6] rel. (6.34)) (1.3)
γ M0
Otherwise, the following reduction is done:
1− n
M N , y , Rd = M pl, y , Rd ⋅ but M N , y, Rd ≤ M pl, y, Rd (EN 1993-1-1 [6] rel. (6.36)) (1.4)
1 − 0,5 ⋅ a
where:
N Ed
n=
N pl, Rd

A − 2 ⋅ b ⋅ tf
a= but a ≤ 0,5
A
A ⋅ fy
N pl,Rd = (EN 1993-1-1 [6] rel. (6.6)) (1.5)
γ M0
tw – web thickness;
hw – web height;
b – flange width;
tf – flange thickness;
Wpl,y ⋅ f y
M pl, y,Rd = (EN 1993-1-1 [6] rel. (6.13)) (1.6)
γ M0
Wpl– plastic section modulus;
γM0 – partial factor for resistance of cross-sections whatever the class is; the
recommended value in EN 1993-1-1 [6] is γM0 = 1,0;

13
• for class 3 cross-sections, the check for bending moment is done with the
following relation:
fy
σ x ,Ed ≤ (EN 1993-1-1 [6] rel. (6.42), (6.43)) (1.7)
γ M0
therefore, to determine the class of the cross-section, one must establish
the stress distribution corresponding to reaching the yielding limit fy in
the farthest fibre from the centre of gravity of the cross-section (Fig.
1.13).

Depending on the class of cross-section and on the amount of axial force NEd,
relations (1.1) and (1.7) define an interaction curve as the one given in figure 1.13.

In practical situations, the stress state on the cross-section is inferior to the limit one.
The question that arises is how to reach the limit, as several options are possible
(Fig. 1.13):
• increasing the axial force (Fig. 1.13(a));
• increasing the axial force and the bending moment according to a variation
law (Fig. 1.13(b));
• increasing the bending moment (Fig. 1.13(c));

(c)

(b)

(a)

Fig. 1.13. Establishing the class of the cross-section

14
The approach in figure 1.13(c), by increasing the bending moment only, is consistent
to the present day version of EN 1993-1-1 [6]. In this idea, using the notations and
the stress distributions given in tables 1.1 and 1.2, the values of α, for the distinction
between class 1 and class 2, and ψ, for the distinction between class 3 and class 4,
should be determined. In these tables, (+) means compression.

1.3.2.1. Determine the class of the flange

In the case of cross-sections subjected to axial force NEd and bending moment about
the y – y axis, My,Ed, (Mz,Ed = 0), there is no normal stress (σ) variation with the width
of the flange (σ is constant on the entire width of the flange). Therefore, the class of
the flange is determined by simply comparing the slenderness of the wall to the
corresponding limits given in table 1.2:
• “non-internal” wall – for flanges of I, H and channel cross-sections and for side
parts of the flanges of box cross-sections;
• “internal” wall in constant compression – for central parts of the flanges of box
cross-sections.

1.3.2.2. Determine the class of the web

In the case of cross-sections subjected to axial force NEd and bending moment about
the y – y axis, My,Ed, the limits for the class of the web depend on the value of the
axial force NEd and they are determined in two main step:
• determine α – the fraction of the web in compression – for classes 1 and 2;
• determine ψ – the ratio between the stress in the extreme fibre opposite to the
one (possibly) yielding in compression and the yielding limit – for class 3 (and 4).

This is generally done by following these detailed steps:


1. Determine the height of the web used by the compression force (considering
a plastic stress distribution)

15
N Ed
hN = (1.8)
nw ⋅ tw ⋅fy

where nw is the number of webs of the cross-section.


Note: In the case of box cross-sections, it must not be neglected that there are
two webs.
2. If hN > c (for the web, c = hw – 2r1, where r1 is the radius of round zone or the
leg of the fillet weld seam), the entire web is in compression and the values
for compression must be used (Tab. 1.1); it is also the case of the central
zone of the flanges of box cross-sections in mono-axial bending

3. If hN < c, the height of the tensioned part (for plastic stress distribution) of the
web is

c − hN
hT = (1.9)
2
4. α – the fraction of the web that is in compression (Tab. 1.1)

hT + hN
α= (1.10)
c
5. The obtained value for α is used for determining the limits (for class 1 and 2)
in table 1.1.

6. If the slenderness of the web is bigger than the limit for class 2, the limit
between class 3 and class 4 must be checked.

7. Determine the stress generated by the compression force on the area of the
cross-section (A) (Tab. 1.1)

N Ed
σN = (1.11)
A
8. Determine the stress range “available” for bending (Tab. 1.1)

σM = f y − σ N (1.12)

9. ψ – the ratio between the stress in the extreme fibre opposite to the one
(possibly) yielding in compression and the yielding limit (Tab. 1.1)

− σM + σ N
ψ= (1.13)
fy

16
1.3.2.3. Establish the class of the cross-section

For each component of the cross-section, the check is done successively for class 1,
2 and 3, till the class is established (if the slenderness is bigger than the limit for
class 3, then it is class 4). The class of the cross-section is the maximum of the
classes of its components.

In many cases, instability is the most important limit state for


structural members made of steel or aluminium alloys. This
phenomenon can affect a part of the member, the entire member, a
Summary part of the structure or the entire structure.
Instability is generally associated to compression but shear stresses
(associated or not with compression stresses) can lead to local
buckling too. Even members in tension can be subject of dynamic
instability.
Instability is a more severe problem than overcoming strength.
EN 1993-1-1 [6] defines four classes of cross-sections of structural
members to manage local buckling. Similar approaches can be
found in the American code ANSI/AISC 360-16 [7] and in the
Japanese code JSCE [8].

Instability, buckling, flexural buckling, torsional buckling, lateral


torsional buckling, in-plane buckling, out-of-plane buckling, snap-
through buckling, local buckling, classes of cross-sections, beam,
Keywords arch, plate, shell, structure, frame

17
Chapter 2

INSTABILITY OF BARS

2.1. TORSION

Generally, torsion is avoided in structural metal (steel or aluminium alloy) members.


There are basically two types of torsion:
• St. Venant torsion (torsiunea cu deplanare liberă);
• warping torsion (torsiunea cu deplanare împiedicată).

As a simplification, in the case of a member with a closed hollow cross-section, such


as a structural hollow section, it may be assumed that the effects of torsional warping
can be neglected; similarly, in the case of a member with open cross section, such
as I or H, it may be assumed that the effects of St. Venant torsion can be neglected.

2.1.1. St. Venant torsion

It occurs when all the following assumptions are accomplished (Fig. 2.1):
• the torsion moment is constant along the bar;
• the area of the cross-section is constant along the bar;
• there are no connections at the ends or along the bar that could prevent
warping.

the flanges remain rectangles

Fig. 2.1. St. Venant torsion

18
2.1.1.1. Stress and strain state

The following aspects can be noticed:


• there is no increase or reduction of the length of the fibres (as there is no
longitudinal force):
εx = 0 → σx = 0 (2.1)
• warping (deplanarea) of the cross-section is a result of the assumption εx = 0
(in order to keep the geometry);

T Ed =  τ × r ⋅ dA (2.2)
A

Fig. 2.2. St. Venant torsion – stress state

• each cross-section rotates like a rigid disk (it goes out of plane but the shape
does not change);
• the rotation between neighbour cross-section is the same along the bar.

θ= = const. (2.3)
dx

2.1.2. Warping torsion

It occurs anytime when at least one of the St. Venant assumptions is not fulfilled
(Fig. 2.3).

19
Fig. 2.3. Warping torsion

2.1.2.1. Stress and strain state

The following aspects can be noticed:


• there are longitudinal stresses and strains (Fig. 2.4):
εx ≠ 0 → σx ≠ 0 → σw; τw (2.4)
• the rotation between neighbour cross-section is variable along the bar.

θ= ≠ const. (2.5)
dx

Fig. 2.4. Warping torsion – stress state

20
2.1.2.2. Equilibrium equations

The following aspects can be noticed:


• there is no axial force acting on the bar:

X Ed ,i = 0  N Ed = 0   σ w dA = 0 (2.6)
A

• there are no bending moments acting on the bar:

M y , Ed ,i = 0  M y , Ed = 0   σ w ⋅ zdA = 0 (2.7)
A

M z , Ed ,i = 0  M z , Ed = 0   σ w ⋅ ydA = 0 (2.8)
A

• in each cross-section, the torsion moment is the sum of the St. Venant
component and the warping component (Fig. 2.5):
TEd =  τ ⋅ r ⋅ dA + Vw ⋅ h e (2.9)
A

TEd = Tt , Ed + Tw , Ed (2.10)

where:
Tt,Ed – the internal St. Venant torsion;
Tw,Ed – the internal warping torsion.

Fig. 2.5. St. Venant torsion and warping torsion

2.1.3. Torsion and bending

2.1.3.1. Bi-symmetrical cross-section subject to bending moment and shear force

21
For a I- or H- cross-section, the force F, acting in the plane xOz, generates only
bending moment about the y – y axis (and shear force) and no torsion moment, as
the resultant forces Ff on the flanges are balanced (Fig. 2.6).

Fig. 2.6. Shear stresses in a bisymmetrical cross-section in bending

2.1.3.2. Mono-symmetrical cross-section subject to bending moment and shear force

A force F, acting in the plane xOz in the centre of gravity of a mono-symmetrical


cross-section, generates not only bending moment about the y – y axis (and shear
force) but torsion moment too (Fig. 2.7).

Fig. 2.7. Shear stresses for force acting in the centre of gravity

TEd = Ff ⋅ h e + Fw ⋅ e (2.11)

The shear centre (centrul de tăiere, centrul de încovoiere-răsucire) is the point


through which the applied loads must pass to produce bending without twisting. A
force F, acting in the plane xOz in the shear centre of a mono-symmetrical cross-
section, generates only bending moment about the y – y axis (and shear force) and
no torsion moment (Fig. 2.8).

22
Fig. 2.8. Shear stresses for force acting in the shear centre

TEd = VEd ⋅ c (2.12)

VEd ⋅ c = Ff ⋅ h e + Fw ⋅ e (2.13)

Ff ⋅ h e + Fw ⋅ e
c= (2.14)
VEd
Ff
Notations: α= ; Fw = VEd (2.15)
VEd
α ⋅ VEd ⋅ h e + VEd ⋅ e
c= (2.16)
VEd

c = α ⋅ he + e (2.17)

F acting in the centre of gravity F acting in the shear centre


Fig. 2.9. Effects of a force acting in or outside of the shear centre

23
2.1.4. Torsion – calculation

2.1.4.1. St. Venant torsion

The case of open cross-sections


a) Rectangular cross-section
t TEd ⋅ t
τ max = t = minimum edge (2.18)
IT

1
b IT = ⋅ b ⋅ t3 (2.19)
3
dϕ T
θ= = ϕ′ = Ed = const. (2.20)
dx G ⋅ IT

TEd = G ⋅ I T ⋅ ϕ′ (2.21)

b) Cross-section made of several rectangles


Rigid disk assumptions (simplifying assumptions):
1. each cross-section rotates one about the other;
1
2. the rotation varies from one cross-section to the other but it is constant
i for all the points on the same cross-section; the cross-section does not

n change its shape in plane but it can go out of plane;


3. the rotation occurs around an axis parallel to the axis of the bar.
As a result of assumption 2,
n

TEd ,1 TEd,n T Ed ,i
TEd
θ= = ... = = i
= (2.22)
G ⋅ I T ,1 G ⋅ I T ,n n
G ⋅ IT
G ⋅  I T ,i
1

1 n
IT = ⋅  b i ⋅ t 3i (2.23)
3 1
Remark: For hot-rolled shapes,
α n
IT = ⋅  b i ⋅ t 3i α = 1,1 … 1,3 (2.24)
3 1
TEd ⋅ t max
τ max = tmax = maximum thickness (2.25)
IT

TEd = G ⋅ I T ⋅ ϕ′ (2.26)
The case of hollow sections (Fig. 2.10)

24
TEd = Va ⋅ b + Vb ⋅ a (2.27)

Fig. 2.10. Torsion of hollow sections

It is accepted that: (Bredt relation)


TEd
Va ⋅ b = Vb ⋅ a = (2.28)
2
TEd T
Va = ; Vb = Ed (2.29)
2⋅b 2⋅a
Va TEd
τa = = (2.30)
a ⋅ ta 2 ⋅ b ⋅ a ⋅ ta

Vb TEd
τb = = (2.31)
b ⋅ tb 2 ⋅a ⋅ b ⋅ tb

TEd
τ max = (2.32)
2 ⋅ A ⋅ t min

2.1.4.2. Warping torsion

An exact calculation would consider the bar as a sum of shells (Fig. 2.11).

Fig. 2.11. Shell modelling of a bar in torsion

25
In daily practice a simplified approach is used, based on the Vlasov theory. The
simplifying assumptions are the following ones:
1. rigid disk behaviour:
• each cross-section rotates one about the other;
• the rotation varies from one cross-section to the other but it is constant
for all the points on the same cross-section;
• the rotation occurs around an axis parallel to the axis of the bar (Fig.
2.12);

Fig. 2.12. Axis of rotation of the bar

2. the shear deformations are zero in the mid-line of the cross-section (Fig.
2.13);

mid-line

Fig. 2.13. Mid-line of the cross-section

3. σw and τw are constant on the thickness of the cross-section, because it is


thin (the mid-line is representative for the cross-section);
4. when calculating σw, it is assumed that τw = 0.

26
Based on these assumptions, the cross-section of the bar is reduced to its mid-line
(Fig. 2.14) and the following relations can be written between in-plane strains and
longitudinal ones (Fig. 2.15), considering rotation around point C:
nn ' = dv (2.33)
du dv
= (2.34)
ds dx
nn ' = nn ′′ ⋅ cos α (2.35)
dv = nn ' = nn ′′ ⋅ cos α (2.36)

mid-line

Fig. 2.14. Mid-surface of the member

nn ′′ = Cn ⋅ dϕ (2.37)
dv = nn ' = Cn ⋅ dϕ ⋅ cos α (2.38)
r = Cn ⋅ cos α (2.39)
dv = r ⋅ dϕ (2.40)
du r ⋅ dϕ dϕ
=  du = r ⋅ ds ⋅ (2.41)
ds dx dx
du
ε=  ε = ω ⋅ ϕ′′ (2.45)
dx

du

Fig. 2.15. Geometric relations

27
By definition (Fig. 2.15),
r ⋅ ds
r ⋅ ds = dω = 2 ⋅ (2 × area of the triangle ) (2.42)
2
Notation (Fig. 2.15):

[ ]
s s
ω =  r ⋅ ds =  dω L2 normalised warping function (coordonată sectorială) (2.43)
0 0

it is also known as sectorial area


du = r ⋅ ds ⋅ ϕ′ = dω ⋅ ϕ′  u = ω ⋅ ϕ′ (2.44)
Expressing σw and τw
σ x = σ w = E ⋅ ε = E ⋅ ω ⋅ ϕ′′ (2.46)

σ w ⋅ ω ⋅ dA = E ⋅ ϕ′′ ⋅ ω2 ⋅ dA (2.47)

B =  σ w ⋅ ω ⋅ dA = E ⋅ ϕ′′ ⋅  ω2 ⋅ dA (bimoment) (2.48)


A A

(bimoment de încovoiere-răsucire)

I w =  ω2 ⋅ dA (warping constant [L6]) (2.49)


A

(moment de inerţie sectorial)


Parallel between bending moment and warping torsion
M y, Ed B
σx = ⋅z σw = ⋅ω (2.50)
Iy Iw

Vz , Ed ⋅ S y M w , Ed ⋅ S w
τz = τw = (2.51)
t ⋅ Iy t ⋅ Iw

Sw =  ω ⋅ dA (warping static moment [L4]) (2.52)


A

(moment static sectorial)


The coordinates of the shear centre about the centre of gravity are:

 ω ⋅ z ⋅ dA
yC = A
(2.53)
Iy

 ω ⋅ y ⋅ dA
zC = A
(2.54)
Iz

28
2.1.5. Cross-section characteristics associated to torsion

Considering a mono-symmetrical cross-section (Fig. 2.16), the following can be


calculated:

Fig. 2.16. Mono-symmetrical cross-section (SN030a-EN-EU [11])

• the position of the shear centre S from the bottom fibre of the cross-section:
t2 b3 ⋅ t
z SC = + hs ⋅ 3 1 13 (2.55)
2 b 2 ⋅ t 2 + b1 ⋅ t1

• the St. Venant torsional constant:


b1 ⋅ t13 + b 2 ⋅ t 32 + h w ⋅ t 3w
IT = (2.56)
3
• the warping constant (SN030a-EN-EU [11]):
b13 ⋅ t 1 ⋅ b 32 ⋅ t 2
I w = h s2 ⋅ I z ⋅ (2.57)
(b
3
1 ⋅ t 1 + b 32 ⋅ t 2 )
2

2.2. BUCKLING LENGTH

The first known theoretical approach for solving a bar in compression belongs to
Euler (1744) [1]. He started by writing the following equilibrium equation (Fig. 2.17)
for a pin connected bar axially loaded in compression:

29
d2v
dx 2 M 1
=− = (2.58)
  dv  2 32
 EI ρ
1 +   
  dx  
where:
M = F⋅ v (2.59)

L
F x

e0
Fig. 2.17. The equilibrium of a pin connected bar in compression

The solution he obtained is the very well known:


π2 ⋅ EI
Fcr = (2.60)
L2

for the critical force that generates buckling of the bar and:
π⋅x
z = e 0 ⋅ sin (2.61)
L
for the deformed shape of the bar.

This relation was then extended to other types of restraints at the ends, by inscribing
the bar on an equivalent pin-connected bar (Fig. 2.18). To allow this, the buckling
length was defined as a concept. All these theoretical approaches are based on the
theory of bifurcation of equilibrium.

The system length (EN 1993-1-1 [6] def. 1.5.5) is the distance in a
given plane between two adjacent points at which a member is
braced against lateral displacement in this plane, or between one
Definition such point and the end of the member.

The buckling length (Lcr) (EN 1993-1-1 [6] def. 1.5.6) is the system
length of an otherwise similar member with pinned ends, which has
the same buckling resistance as a given member or segment of
Definition member.

30
It is also defined as the distance between two consecutive inflection
points along the deformed shape of a bar. Sometimes, in practice, it
is replaced by the system length.

Euler’s relation is then expressed as:


π2 ⋅ EI
Fcr = (2.62)
L2cr
where Lcr = kL is the buckling length (Fig. 2.18).
k – end fixity condition.

k = 1,0 k = 0,7 k = 2,0 k = 0,5 k = 1,0


Fig. 2.18. Different values of the buckling length factor

2.2.1. Buckling length of columns

In everyday situations, bars are part of a structure, they are connected to other bars
and so the joints are not purely fixed or purely pinned. As a result, the buckling
length of an element depends on its loading state and on the stiffness of the
neighbour bars. Relations for calculating it are given in different books and
publications like Annexe E (informative) of the previous version of Eurocode 3 – ENV
1993-1-1 [12], or SN 008a-EN-EU [13]. For defining the buckling length of a column,
(parts of) structures are separated in “sway” and “non-sway”, depending whether the
(lateral) displacements of the joints at the end of the bar are permitted or not. This
separation is done by means of stiffness criteria; such a criterion is formulated in
ENV 1993-1-1 [12], where it is stated that a steel structure can be considered as

31
braced if the brace system reduces its horizontal displacements by at least 80%.
Usually, the “non-sway” behaviour is guaranteed by means of bracings. The
distribution factors ηi used in figure 2.19 – 2.22 are calculated using the following
relations:
KC
η1 = (ENV 1993-1-1 [12], rel. (E.1)) (2.63)
K C + K11 + K12

KC
η2 = (ENV 1993-1-1 [12], rel. (E.2)) (2.64)
K C + K 21 + K 22
where:
KC – stiffness of the column (I/L);
Kij – stiffness of the beam ij.
Remark: A more precise formulation for Kij would be “stiffness of the connection”
between beam ij and column, as semi-rigid connections could be used. In this case
a more careful analysis should be carried out.

The buckling length for non-sway buckling mode is presented in figure 2.19 [12] and
it is determined using factors obtained according to figure 2.21 [12].

Fig. 2.19. Non-sway buckling mode (ENV 1993-1-1 [12] Fig. E.2.3)

32
The buckling length for sway buckling mode is presented in figure 2.20 [12] and it is
determined using factors obtained according to figure 2.22 [12].

Fig. 2.20. Sway buckling mode (ENV 1993-1-1 [12] Fig. E.2.3)

Fig. 2.21. End fixity condition, k, for non-sway buckling (ENV 1993-1-1 [12] Fig.
E.2.1)

33
Fig. 2.22. End fixity condition, k, for sway buckling (ENV 1993-1-1 [12] Fig. E.2.2)

This model (for sway and for non-sway buckling) can be expanded to continuous
columns, presuming the loading factor N/Ncr is constant on their entire length. If this
does not happen (which is the actual case) the procedure is conservative for the
most critical part of the column [12]. For continuous columns, the distribution factors
are calculated using the following relations:
K C + K1
η1 = (ENV 1993-1-1 [12], rel. (E.3)) (2.65)
K C + K1 + K11 + K12

KC + K2
η2 = (ENV 1993-1-1 [12], rel. (E.4)) (2.66)
K C + K 2 + K 21 + K 22
where K1 and K2 are the values of the stiffness of the neighbour columns (Fig. 2.23).

34
Fig. 2.23. Distribution factors for continuous columns (ENV 1993-1-1 [12] Fig. E.2.4)

2.2.2. Buckling length of beams

Presuming the beams are not subject to axial forces, their stiffness can be taken
from table 2.1, as long as they remain in the elastic range (ENV 1993-1-1 [12]).

Tab. 2.1. Stiffness of a beam in the elastic range (ENV 1993-1-1 [12] Tab. E.1)

Connection at the other end of the beam Stiffness K of the beam


Fixed 1,0 × I/L
Pinned 0,75 × I/L
Rotation equal to the adjacent one (double curvature) 1,5 × I/L
Rotation equal and opposite to the adjacent one 0,5 × I/L
(simple curvature)
General case: θa rotation at the adjacent end and θb (1,0 + 0,5 × θa/θb) × I/L
rotation at the opposite end

35
For regular buildings with rectangular frames and reinforced concrete floors, subject
to uniform loads, it is accepted to consider the stiffness of the beams given in table
2.2.

Tab. 2.2. Stiffness K of beams – structures with reinforced concrete floors (ENV
1993-1-1 [12] Tab. E.2)

Loading condition of the beam Non-sway buckling Sway buckling mode


mode
Beams supporting directly the 1,0 × I/L 1,0 × I/L
reinforced concrete slabs
Other beams under direct loads 0,75 × I/L 1,0 × I/L
Beams subjected only to 0,50 × I/L 1,5 × I/L
bending moments at the ends

When the beams are subject to axial forces, stability functions must be used for
expressing their stiffness. A simplified conservative approach is proposed in ENV
1993-1-1 [12], neglecting the increase of stiffness generated by tension and
considering only compression in the beams. Based on these assumptions, the
values in table 2.3 can be considered.

Tab. 2.3. Stiffness of beams in compression (ENV 1993-1-1 [12] Tab. E.3)

Connection at the other end of the beam Stiffness K of the beam


Fixed 1,0 × I/L × (1,0 – 0,4 × N/NE)
Pinned 0,75 × I/L × (1,0 – 1,0 × N/NE)
Rotation equal to the adjacent one (double 1,5 × I/L × (1,0 – 0,2 × N/NE)
curvature)
Rotation equal and opposite to the adjacent one 0,5 × I/L × (1,0 – 1,0 × N/NE)
(simple curvature)

where:
π 2 ⋅ EI
NE = (2.67)
L2

36
2.2.3. Empirical relations for the buckling length of columns

ENV 1993-1-1 [12] provides empirical expressions as safe approximations that can
be used as an alternative to the values from figures 2.21 and 2.22. The k coefficient
for the buckling length can be calculated by the following relations:
a. for non-sway buckling mode (Fig. 2.21)
k = 0,5 + 0,14 ⋅ (η1 + η2 ) + 0,055 ⋅ (η1 + η2 )
2
([12], rel. (E.5)) (2.68)
or, alternatively,
1,0 + 0,145 ⋅ (η1 + η2 ) − 0,265 ⋅ η1 ⋅ η2
k= ([12], rel. (E.6)) (2.69)
2,0 − 0,364 ⋅ (η1 + η2 ) − 0,247 ⋅ η1 ⋅ η2
b. for sway buckling mode (Fig. 2.22)

1,0 − 0,2 ⋅ (η1 + η2 ) − 0,12 ⋅ η1 ⋅ η2 


0,5

k=  ([12], rel. (E.7)) (2.70)


1,0 − 0,8 ⋅ (η1 + η2 ) + 0,60 ⋅ η1 ⋅ η2 

2.2.4. Comments on the buckling length of beams

If the buckling length is generally easy to identify for members subject to axial
compression forces, the effective lateral buckling length is a more delicate subject,
given the complexity of the deformed shape (combination of flexural buckling and
torsion). This leads to a temptation to simplified approaches, like considering the
effective lateral buckling length as equal to the distance between points of zero (Fig.
2.24) in the bending moment diagram, or between inflection points of the deformed
shape of the beam about the strong axis of its cross-section [14].

In order to prevent this, the American code ANSI/AISC 360-16 [7]


states in the 6.3 commentary: “In members subjected to double
curvature bending, the inflection point shall not be considered a
Important brace point.”

37
Fig. 2.24. Zero bending moment points along a beam [14]

2.2.5. Present day practical buckling design

The previous approaches are valid only in the elastic range, using the theory of
bifurcation of equilibrium. Several researchers tried to express buckling when at least
one of the following Euler’s requirements for the bifurcation of equilibrium is not
fulfilled:
• the axis of the member is rigorously straight;
• the compression load acts strictly in the centre of gravity of the cross-section;
• the cross-section is bi-symmetrical;
• the moment of inertia of the cross-section is constant all along the bar;
• the deflected shape is a sinusoid;
• the material is homogenous and has a perfectly elastic behaviour (E = constant).

Several researchers focused on buckling in the elasto-plastic range; the following


ones can be mentioned [2]: Engesser and Considère (1889), Tetmayer (1890), von
Kàrmàn and Iassinski at the same time with Engesser (1910), Shanley (1946).

A different approach, based on the theory of divergence of equilibrium, was


considered by ECCS (European Convention for Steel Structures) which conducted
an experimental analysis in seven countries (Belgium, France, Germany, Great

38
Britain, Italy, Netherlands and Yugoslavia) during the decade 1960 – 1970. Following
this, at the beginning one buckling curve and then several ones were drawn,
depending on:
• the shape of the cross-section;
• the plane in which buckling occurs – the axis of the cross-section;
• the yield limit of the steel grade.

Present day practical buckling design codes approach is a compromise of two


distinct concepts:
• buckling length, usually based on the equilibrium bifurcation model, built
around the perfect member idea;
• reduction (buckling) factors, based on the equilibrium divergence model,
taking into account the actual imperfect member.
In fact, the old classical stability analysis consists of two main steps: a stress and
deflections analysis (often using a computer program) and a conventional code
analysis on isolated members [14].

2.2.6. Buckling length of bars with changes in cross-section

There are situations in practice when a bar in compression does not have the same
cross-section on its entire length. This could be the case of a column with a change
in cross-section (Fig. 2.25) or of a pin-connected member, like a bar connected on
gusset plates at its ends (Fig. 2.26). In such cases, an equivalent buckling length can
be used, based on the theoretical approaches of Timoshenko [15].

I2 I1
P

L2 L1
Fig. 2.25. Column with a change in cross-section

Considering the notations in figure 2.25, Timoshenko writes the following equilibrium
equations:

39
d 2 y1
E ⋅ I1 ⋅ = P ⋅ (δ − y1 )
dx 2
(2.71)
d 2 y2
E ⋅ I2 ⋅ = P ⋅ (δ − y 2 )
dx 2
where:
δ – the displacement at the free end of the column;
y1 – the deformed shape of part 1;
y2 – the deformed shape of part 2.
L = L1 + L2
Using the notations,
P P
k 12 = k 22 = (2.72)
E ⋅ I1 E ⋅ I2
the following transcendental equation is obtained for determining the critical load:

tan (k 1 ⋅ L1 ) × tan (k 2 ⋅ L 2 ) =
k1
(2.73)
k2
Replacing the following (Fig. 2.26):
a L−a
L2 → and L1 → (2.74)
2 2
the critical force can be expressed as (in this case, δ is the displacement in the
middle of the length of the bar):
m ⋅ E ⋅ I2
Pcr = (2.75)
L2
where the values of the factor m are given in table 2.4.

I1 I2 I1
P P

Fig. 2.26. Pin-connected member (bar connected on gusset plates)

As a result, the buckling length can be expressed as:


L cr = β × L (2.76)

40
where:
π
β= (2.77)
m

Tab. 2.4. Values for the m factor [15] for figure 2.26

a/L
0,2 0,4 0,6 0,8
I1/I2
0,01 0,15 0,27 0,60 2,25
0,1 1,47 2,40 4,50 8,59
0,2 2,80 4,22 6,69 9,33
0,4 5,09 6,68 8,51 9,67
0,6 6,98 8,19 9,24 9,78
0,8 8,55 9,18 9,63 9,84

The previous equations are written using the elastic approach, by writing elastic
equilibrium equations. An alternative solution can be obtained using the energy
method. In this idea, the deformed shape in figure 2.25 is considered [15]:
 π⋅x 
y = δ ⋅ 1 − cos  (2.78)
 2⋅L
As a result, for the problem in figure 2.25, the critical force is obtained [15]:
π2 ⋅ E ⋅ I2 1
Pcr = ⋅ (2.79)
4⋅L 2
L 2 L1 I 2 1  I 2  π ⋅ L2
+ ⋅ − ⋅  − 1 ⋅ sin
L L I1 π  I1  L

and the β factor in relation (2.76) is:

L 2 L1 I 2 1  I 2  π ⋅ L2
β = 2⋅ + ⋅ − ⋅  − 1 ⋅ sin (2.80)
L L I1 π  I1  L

Similarly, for the problem in figure 2.26, the critical force is [15]:
π2 ⋅ E ⋅ I 2 1
Pcr = ⋅ (2.81)
a L − a I2 1  I2  π⋅a
2
L
+ ⋅ − ⋅  − 1 ⋅ sin
L L I1 π  I1  L

and the β factor in relation (2.76) is:

a L − a I2 1  I2  π⋅a
β= + ⋅ − ⋅  − 1 ⋅ sin (2.82)
L L I1 π  I1  L

41
It is known that the energy approaches generally overestimate the value of the
critical force, leading to values superior to the actual ones.

2.2.7. Buckling length of bars with continuous varying cross-section

There are situations in practice when a bar in compression has a continuous


variation of the cross-section on its entire length (Fig. 2.27).

L a
x
Fig. 2.27. Column with a continuous variation of the cross-section

Using the notations in figure 2.27, the variation of the cross-section is described by:
n
x
I(x ) = I1 ⋅   (2.83)
a
where I1 is the second moment of the area (moment of inertia) of the top cross-
section of the column (where x = a).
Timoshenko [15] writes the following equilibrium equation:
n
x d y
2
E ⋅ I1 ⋅   ⋅ 2 = −P ⋅ y (2.84)
 a  dx
In the general case, the equation can be solved using Bessel functions for any value
of n. For the particular case of n = 2, that would correspond to an I-shape with
constant flanges and continuous variation of the height of the web, the critical force
can be expressed [15] as:
m ⋅ E ⋅ I2
Pcr = (2.85)
L2
where I2 is the second moment of the area (moment of inertia) of the bottom cross-
section of the column (where x = a + L). The buckling length can be expressed using
relations (2.76) and (2.77) and values of the factor m given in table 2.5.

42
Tab. 2.5. Values for the m factor [15] for figure 2.27 and n = 2

I1/I2 0,0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8 0,9 1,0
m 0,250 1,350 1,593 1,763 1,904 2,023 2,128 2,223 2,311 2,392 π2/4

The behaviour of a solid conical bar is described by making n = 4 in relations (2.83)


and (2.84).

2.3. STABILITY BRACING

Bracings are essential components for the structural stability control. They can
provide the lateral supports that are needed for preventing buckling.

Various criteria are used for classifying bracings, as: nodal or relative; punctual or
continuous; against translation, against rotation or against both, specific for buckling
or for lateral buckling etc. Usually, the main supports have also stability functions. A
standard case is that of the classical fork support (Fig. 2.28). Some usual bracing
cases are presented in figure 2.28.

Two main requirements apply on bracing systems:


• strength – the bracing system must be able to resist the forces (or moments)
generated by the buckling trend;
• stiffness – the displacements of the “supports” provided by the bracing system
must be low, otherwise buckling occurs.
In many cases, the stiffness requirement is more severe than the strength one.

43
Fig. 2.28. Usual bracing systems [14]

Complex bracings systems could be investigated using the spring composition


procedure. Two strategies are available in dealing with stability bracings:
• requiring minimum values for the strength and the stiffness of the bracings to
allow the consideration of the analysis unbraced length as the distance between
two consecutive braces (ANSI/AISC 360-16 [7], appendix 6);
• establishing the load capacity of the structure taking into account the bracing
system as it is (EN 1993-1-1 [6]).

44
2.3.1. The bow imperfection (imperfecțiunea inițială în arc)

To take into account the fact that the bar is not perfectly straight, EN 1993-1-1 [6]
uses an equivalent bow imperfection (Fig. 2.29).
P P N P
H H

q α
q

e0 e0
L L

Fig. 2.29. Bow imperfection for a pinned member [14]

Considering a sine-function for the buckling shape (Fig. 2.29):


π⋅x
z = e 0 ⋅ sin (2.86)
L
the associated shear force on the supports is:
dz π π⋅ x
H = N⋅ = N ⋅ e0 ⋅ ⋅ cos (2.87)
dx L L
e0
H max = π ⋅ N ⋅ (2.88)
L
If a parabola is considered as initial deformed shape (Fig. 2.29), the value of Hmax
increases a little and the following relation, given in figure 5.4 from EN 1993-1-1 [6],
can be obtained:
e0
H max = 4 ⋅ N ⋅ (2.89)
L
To manage this bow imperfection, EN 1993-1-1 [6] uses an equivalent lateral load
(Fig. 2.29) that would produce a bending moment equal to N × e0.

45
q ⋅ L2
M max = N ⋅ e 0 = (2.90)
8
The values recommended for e0 are given in table 2.6.

Tab. 2.6. Recommended values for bow imperfection (EN 1993-1-1 [6] Tab. 5.1)

Buckling curve Elastic analysis Plastic analysis


e0 / L e0 / L
a0 1 / 350 1 / 300
a 1 / 300 1 / 250
b 1 / 250 1 / 200
c 1 / 200 1 / 150
d 1 / 150 1 / 100

For the buckling curves given in EN 1993-1-1 [6], Hmax results as (0.9...2.7)% of the
compression force.

2.3.2. The sway imperfection (imperfecțiunea inițială datorată abaterii de la axa


verticală)

To take into account the fact that the force does not act exactly on the centroid line
of the bar, EN 1993-1-1 [6] uses an equivalent sway imperfection (Fig. 2.30).
P P
H

Fig. 2.30. Sway imperfection for a pinned member [14]

46
H = φ⋅P (2.91)
where:
φ = φ0 ⋅ α h ⋅ α m (EN 1993-1-1 [6], rel. (5.5)) (2.92)
ϕ0 – basic value, ϕ0 = 1/200;
αh – the reduction factor for height h (h = L in Fig. 2.30) applicable to columns:
2 2
αh = but ≤ αh ≤ 1 (2.93)
h 3
h – “the height of the structure in meters”; in this case, h = L;
αm – “the reduction factor for the number of columns in a row”;

 1
α m = 0,5 ⋅ 1 +  (2.94)
 m
m – “the number of columns in a row including only those columns which carry a
vertical load NEd not less than 50% of the average value of the column in the
vertical plane considered” (Fig. 2.31).

Fig. 2.31. Equivalent sway imperfections (EN 1993-1-1 [6] Fig. 5.2)

According to EN 1993-1-1 [6], for building frames, it may be disregarded where:


H Ed ≥ 0,15 ⋅ VEd (EN 1993-1-1 [6], rel. (5.7)) (2.95)

where:
HEd – “the design value of the horizontal reaction at the bottom of the storey to the
horizontal loads and fictitious horizontal loads”;
VEd – “the total design vertical load on the structure on the bottom of the storey”.

47
2.4. STRUCTURAL STABILITY ANALYSES

The present day computer programs for structural analysis and computing devices
used to run these applications allow analysing a structure by writing the equilibrium
equations on its deformed shape. However, this type of analysis is not always
necessary. In general, two types of analyses can be carried out on a structure:
• first order analysis – equilibrium is expressed on the initial shape of the
structure (efforts do not increase because of the displacements);
• second order analysis – equilibrium is expressed on the deformed shape of
the structure (efforts increase, significantly, because of the displacements).

According to EN 1993-1-1 [6], it is not necessary to perform a second order analysis


in situations when:
Fcr
α cr = ≥ 10 for elastic analysis
FEd
(EN 1993-1-1 [6], rel. (5.1)) (2.96)
F
α cr = cr ≥ 15 for plastic analysis
FEd
where:
αcr – the factor by which the design loading would have to be increased to cause
elastic instability in a global mode;
FEd – the design loading on the structure;
Fcr – the elastic critical buckling load for global instability mode based on initial
values of the elastic stiffness.
These requirements must be fulfilled on each floor of a building.

Provided that the compression forces in the beams are not important, αcr for portal
frames with shallow roof slopes can be calculated with the following approximative
formula (Fig. 2.32):

H   h 
α cr =  Ed  ⋅  

(EN 1993-1-1 [6], rel. (5.2)) (2.97)
 VEd   δ H ,Ed 

48
Fig. 2.32. Notations for determining αcr (EN 1993-1-1 [6] Fig. 5.1)

where:
HEd – the design value of the horizontal reaction at the bottom of the storey to the
horizontal loads and fictitious horizontal loads;
VEd – the total design vertical load on the structure on the bottom of the storey;
h – the storey height;
δH,Ed – the horizontal displacement at the top of the storey, relative to the bottom of
the storey, when the frame is loaded with horizontal loads and fictitious
horizontal loads which are applied at each floor level.

Remarks:
1. In the absence of more detailed information, in the previous formulation, a roof
slope may be considered to be shallow if it is not steeper than 1:2 (26°).
2. In the absence of more detailed information, in the previous formulation, the axial
compression in the beams may be assumed to be significant if:
A ⋅ fy
λ ≥ 0,3 ⋅ (EN 1993-1-1 [6], rel. (5.3)) (2.98)
N Ed

where:
λ – the in-plane non dimensional slenderness calculated for the beam considered
as hinged at its ends of the system length measured along the beam;
NEd – the design value of the compression force.

49
For single storey frames, in the elastic range, if αcr ≥ 3,0, second order sway effects
due to vertical loads may be calculated by increasing the horizontal loads HEd and
equivalent loads φ × VEd due to imperfections by the factor:
1
(EN 1993-1-1 [6], rel. (5.4)) (2.99)
1
1−
α cr

“When performing the global analysis for determining end forces and end moments
to be used in member checks, local bow imperfections may be neglected. However
for frames sensitive to second order effects local bow imperfections of members
additionally to global sway imperfections should be introduced in the structural
analysis of the frame for each compressed member where the following conditions
are met:
• at least one moment resistant joint at one member end;
A fy
• λ > 0,5 (EN 1993-1-1 [6], rel. (5.8)) (2.100)
N Ed

where:
λ – the in-plane non-dimensional slenderness calculated for the member
considered as hinged at its ends;
NEd – the design value of the compression force”.

2.5. THE UNIQUE GLOBAL AND LOCAL IMPERFECTION

EN 1993-1-1 [6] accepts a unique imperfection approach, as an alternative to the two


types of imperfections – bow and sway ones, using the following relation:
N cr e N Rk
ηinit = e 0 ⋅ ⋅ηcr = 02 ⋅ ⋅ηcr (EN 1993-1-1 [6], rel. (5.9)) (2.101)
E ⋅ I ⋅ η′cr′ ,max λ E ⋅ I ⋅ η′cr′ ,max

where:
χ ⋅ λ2
1−
γ M1
e 0 = α ⋅ (λ − 0,2 ) ⋅
M Rk
⋅ for λ > 0,2 (EN 1993-1-1 [6], rel. (5.10)) (2.102)
N Rk 1 − χ ⋅ λ2

50
α ult ,k
λ= is the relative slenderness (EN 1993-1-1 [6], rel. (5.11)) (2.103)
α cr

α – the imperfection factor for the relevant buckling curve;


χ – the reduction factor for the relevant buckling curve for that cross-section;
αult,k – the minimum force amplifier for the axial force configuration NEd in members
to reach the characteristic resistance NRk of the most axially stressed cross
section without taking buckling into account;
αcr – the minimum force amplifier for the axial force configuration NEd in members
to reach the elastic critical buckling;
MRk – the characteristic moment resistance of the critical cross section (Mel,Rd or
Mpl,Rd);
NRk – the characteristic resistance to axial force;
E ⋅ I ⋅ η′cr′ ,max – the bending moment due to ηcr at the critical cross section;

ηcr – the shape of elastic critical buckling mode.

2.6. LATERAL-TORSIONAL BUCKLING OF BEAMS

Lateral-torsional buckling may occur in the case of beams or lattice girders. It can be
prevented either by performing checks using suitable relations, or by introducing
lateral bracings whose purpose is to reduce the distance on which this phenomenon
may occur.

The relation given in EN 1993-1-1 [6] for the lateral-torsional buckling resistance is:
fy
M b,Rd = χ LT ⋅ Wy ⋅ (EN 1993-1-1 [6], rel. (6.55)) (2.104)
γ M1
where:
χLT – the reduction factor for lateral-torsional buckling;
Wy – the appropriate section modulus:
 Wy = W pl,y for Class 1 or 2 cross-sections
 Wy = W el,y for Class 3 cross-sections
 Wy = W eff,y for Class 4 cross-sections
fy – the yielding limit;

51
γM1 – the partial safety factor for resistance of members to instability assessed by
member checks (γM1 = 1,0 in the National Annex of EN 1993-1-1 [6]);
χLT is taken form the appropriate buckling curve, based on the non-dimensional
slenderness:
Wy ⋅ f y
λ LT = (2.105)
M cr

where Mcr is the elastic critical moment for lateral-torsional buckling, whose
expression is not given in EN 1993-1-1 [6]. Under these circumstances, it can be
taken from recognised sources like publications, or computer programs. Some
examples of such sources are the following ones:
• the previous version of Eurocode 3 - ENV 1993-1-1 [12];
• published books (ex. Timoshenko, Gere [15] etc.);
• „Non-Contradictory Complementary Information” documents (NCCI), ex.
SN003b-EN-EU [16];
• computer programs like LTBEAM [17], developed at CTICM (Centre
Technique Industriel de la Construction Métallique).
The relation recommended by SN003b-EN-EU [16] is:
 
π 2 ⋅ E ⋅ I z   k z  I w (k z ⋅ L ) ⋅ G ⋅ I T
2

( )
2
M cr = C1 ⋅   ⋅ + + C ⋅ z
2
− C ⋅ z  (2.106)
(k z ⋅ L)2   k w  I z π2 ⋅ E ⋅ I z
2 g 2 g

 
where:
E – modulus of elasticity (Young’s modulus E =210000N/mm2);
G – shear modulus (G =81000N/mm2);
Iz – the second moment of the area about the weak axis (z – z);
IT – the torsion constant;
Iw – the warping constant;
L – the beam length between points which have lateral restraints;
kz – the effective length factor that refers to the end rotation about the z – z axis;
kw – the effective length factor that refers to the end warping;
zg – the distance between the point of load application and the shear centre;
C1, C2 – coefficients depending on the loading and end restraint conditions.
It is to note that the value of the critical moment is influenced by the position of the
loading point. The load can have a stabilizing or a destabilizing effect (Fig. 2.33).

52
destabilizing stabilizing
compression flange
effect effect

tension flange

Fig. 2.33. Influence of the position of the loading point

The relation (2.104), given in EN 1993-1-1 [6], can be used by means of three
methods given in the code.

2.6.1. The general method

For all cross-sections, unless otherwise specified, the following relation can be used:
1
χ LT = but χ LT ≤ 1,0 (EN 1993-1-1 [6], rel. (6.56)) (2.107)
2
Φ LT + Φ 2LT − λ LT

where:

[ ( ) 2
Φ LT = 0,5 ⋅ 1 + α LT ⋅ λ LT − 0,2 + λ LT ] (2.108)
αLT – imperfection factor, given in the (Romanian) National Annex of EN 1993-1-1
[6]; the recommended values can be taken from table 2.7.

Tab. 2.7. Values for lateral torsional buckling (EN 1993-1-1 [6] Tab. 6.3)

Buckling curve a b c d
Imperfection factor αLT 0,21 0,34 0,49 0,76

53
The proper buckling curve, depending on the type of cross-section, is chosen based
on the recommendations given in table 2.8.

Tab. 2.8. Recommended values for lateral torsional buckling curves for cross-
sections using relation (2.107) (EN 1993-1-1 [6] Tab. 6.4)

Cross-section Limits Buckling curve


Rolled I-sections h/b ≤ 2 a
h/b > 2 b
Welded I-sections h/b ≤ 2 c
h/b > 2 d
Other cross-sections - d

M Ed 2
For values of the slenderness λ LT ≤ λ LT, 0 or for ≤ λ LT , 0 , lateral torsional buckling
M cr
effects may be ignored and only cross sectional checks apply. The maximum

recommended value for λ LT, 0 , which was adopted in the Romanian National Annex
of EN 1993-1-1 [6] is 0,4.

2.6.2. The specific method for rolled sections or equivalent welded sections

This method is a particular case for rolled sections or equivalent welded section. In
this case,
χ LT ≤ 1,0
1 
χ LT = but  1 (EN 1993-1-1 [6], rel. (6.57)) (2.109)
χ LT ≤ 2
2
Φ LT + Φ 2LT − β ⋅ λ LT λ LT

where:

[ ( ) 2
Φ LT = 0,5 ⋅ 1 + α LT ⋅ λ LT − λ LT 0 + β ⋅ λ LT ] (2.110)
αLT – imperfection factor.
λ LT 0 ≤ 0,4 the value in the Romanian National Annex of EN 1993-1-1 [6] λ LT 0 = 0,4
β ≥ 0,75 the value in the Romanian National Annex of EN 1993-1-1 [6] β = 0,75
The values of αLT are taken from table 2.7, depending on the proper buckling curve,
chosen based on the recommendations given in table 2.9.

54
Tab. 2.9. Recommended values for lateral torsional buckling curves for cross-
sections using relation (2.109) (EN 1993-1-1 [6] Tab. 6.5)

Cross-section Limits Buckling curve


Rolled I-sections h/b ≤ 2 b
h/b > 2 c
Welded I-sections h/b ≤ 2 c
h/b > 2 d

2.6.3. The modified specific method

This method is also specific for rolled sections or equivalent welded section. A
correction is introduced, to take into account the bending moment diagram along the
bar. In this idea, a modified reduction factor is calculated, as follows:
χ LT
χ LT ,mod = but χ LT , mod ≤ 1,0 (EN 1993-1-1 [6], rel. (6.57)) (2.111)
f
where:
χ LT – the reduction factor obtained at 2.6.2, using relation (2.109);

[ (
f = 1 − 0,5 ⋅ (1 − k c ) ⋅ 1 − 2,0 ⋅ λ LT − 0,8 )]
2
but f ≤ 1,0 (2.112)

kc is given in table 2.10, depending on the bending moment diagram.

Tab. 2.10. Correction factor kc (EN 1993-1-1 [6] Tab. 6.6)

Moment distribution kc

1,0
ψ=1
1
1,33 − 0,33ψ
-1 ≤ ψ ≤ 1
0,94
0,90
0,91
0,86
0,77

55
0,82

2.6.4. Simplified assessment methods for beams with restraints in buildings

In the case where the compressed flange is provided with discrete lateral restraints
(the recommended situation) at a distance Lc between two consecutive ones, the
beam is not susceptible to lateral-torsional buckling if:
k cLc M c ,Rd
λf = ≤ λ c0 (EN 1993-1-1 [6], rel. (6.59)) (2.113)
i f ,z λ 1 M y,Ed

where:
My,Ed– the maximum design value of the bending moment within the restraint
spacing;
fy
M c, Rd = Wy ⋅ ;
γ M1
kc – a slenderness correction factor for moment distribution between restraints,
given in table 2.10;
if,z – the radius of gyration of the equivalent compression flange composed of the
compression flange plus 1/3 of the compressed part of the web area, about
the minor axis of the section;
I eff ,f
i f ,z = ;
1
A eff ,f + ⋅ A eff , w ,c
3
where:
Ieff,f – the effective second moment of area of the compression flange about
the minor axis of the section;
Aeff,f – the effective area of the compression flange;
Aeff,w,c – the effective areas of the compressed part of the web.
λ c 0 – a slenderness limit of the equivalent compression flange defined as follows; it
is given in the National Annex of EN 1993-1-1 [6]; the recommended value is
λ c 0 = λ LT ,0 + 0,1 ;

56
E
λ1 = π = 93,9ε ;
fy

235
ε= (fy in N/mm2).
fy

If the requirements from relation (2.113) are not fulfilled, the design buckling
resistance moment is expressed as:
M b,Rd = k fl ⋅ χ ⋅ M c ,Rd but M b.Rd ≤ M c.Rd (EN 1993-1-1 [6], rel. (6.60)) (2.114)

where:
χ – the reduction factor of the equivalent compression flange determined with λ f ;
kfℓ – the modification factor accounting for the conservatism of the equivalent
compression flange method.
The recommended value is kfℓ = 1,10 and it was adopted in the Romanian National
Annex of EN 1993-1-1 [6]. χ is determined based on λ f , using curve d for welded
h
sections having ≤ 44ε and curve c for all other situations, where:
tf
h – the overall depth of the cross-section;
tf – the thickness of the compression flange.

2.7. STABILITY REQUIREMENTS FOR BRACING SYSTEMS

Buckling and lateral-torsional buckling checks strongly depend on the position of the
bracing points along the structural member in discussion. The requirements for a
bracing system are not only in terms of strength but stiffness too. In fact, if a bracing
system needs important displacements to resist against forces generated by the
buckling trend of an element, buckling of the braced element could occur.

Following the strength calculation according to relations given in EN 1993-1-1 [6], it


could be considered that, globally, a mean force of about 2% of the axial load (1.6%
corresponding to the bow imperfection and 0.5% for the sway imperfection) is going
to the support of a pin-connected member. On this basis, it could be appreciated that
some required strengths for bracings in ANSI/AISC 360-16 [7] are under-evaluated

57
by EN 1993-1-1 [6], as in the American code requirements are expressed not only as
strength but as stiffness too. Winter, Yura and others [7] showed that the minimum
theoretical stiffness of bracings, evaluated for the ideal member, is not enough for
the actual imperfect member. Figure 2.34 and table 2.11 show examples [14] of
strength and stiffness requirements for bracing systems in ANSI/AISC 360-16 [7].
P P

P
Lb Pr,b
βbr
Lb
Pr,b
βbr
P P Lb
Pr,b Pr,b
βbr βbr Lb Pr,b
Lb Lb Lb Lb
Lb Pr,b βbr
βbr Pr,b
βbr
(a) (b) (c) (d) (e)
Fig. 2.34. Strength and stiffness requirements in ANSI/AISC 360-16 [7]

In table 2.11, the notations have the following meanings:


Pr,b – required brace strength (N);
Pr – required strength in axial compression (N);
Lb – unbraced length (mm);
βbr – required brace stiffness (N/mm);
ϕ – resistance factor (ϕ = 0,75);
Mr – required flexural strength (Nmm);
Cd = 1,0, except in the following case;
= 2,0, for the brace closest to the inflection point in a beam subject to double
curvature bending;
h0 – distance between flange (chord) centroids (mm).

Tab. 2.11. Strength and stiffness requirements in ANSI/AISC 360-16 [7]

Case Required strength Required stiffness


(a) Pr ,b = 0,004 ⋅ Pr ([7], rel. (A-6-1)) 1  2 ⋅ Pr 
β br = ⋅  ([7], rel. (A-6-2))
φ  L b 

58
(b) Pr ,b = 0,004 ⋅ Pr ([7], rel. (A-6-1)) 1  2 ⋅ Pr 
β br = ⋅  ([7], rel. (A-6-2))
φ  L b 
c) Pr ,b = 0,01 ⋅ Pr ([7], rel. (A-6-3)) 1  8 ⋅ Pr

β br = ⋅  ([7], rel. (A-6-4))
φ  L b 
(d) M r ⋅ Cd 1  4 ⋅ M r ⋅ Cd 
Pr ,b = 0,008 ⋅ ([7], rel. (A-6-5)) β br = ⋅   ([7], rel. (A-6-6))
h0 φ  L b ⋅ h 0 
(e) M r ⋅ Cd 1  10 ⋅ M r ⋅ C d 
Pr ,b = 0,02 ⋅ ([7], rel. (A-6-7)) β br = ⋅   ([7], rel. (A-6-8))
h0 φ  L b ⋅ h 0 

In figure 2.34 and table 2.11, case (d) refers to relative brace while case (e) is for
nodal brace. Relative brace – brace that controls the relative movement of two
adjacent brace points along the length of a beam or column or the relative lateral
displacement of two stories in a frame (ANSI/AISC 360-16 [7]). Nodal brace – brace
that prevents lateral movement or twist independently of other braces at adjacent
brace points (ANSI/AISC 360-16 [7]). Bracing – members or system that provides
stiffness and strength to limit the out-of-plane movement of another member at a
brace point (ANSI/AISC 360-16 [7]).

2.8. MEMBERS IN COMPRESSION AND BENDING

If checking stability of a member in compression or in bending is quite well controlled


by means of mathematical approaches and test data that are available, the
behaviour of a member in compression and bending is much more difficult to handle.
The checking relations try to merge buckling of member in compression with lateral-
torsional buckling of members in bending and, generally, this is not done in a single
formula; in most of the cases, codes use a pair of two stability relations. However, all
these “built-up” formulae have their limits, as they try to describe a complex
phenomenon which cannot be seen as a simple sum of the two ones previously
mentions. Moreover, there are situations when the most severe check for a member
in compression and bending is the strength check, therefore they are briefly
presented here.

59
2.8.1. Resistance of cross-sections in compression and bending

The resistance checking relations given in EN 1993-1-1 [6] for members in


compression and bending are as follows:
• for class 1 and class 2 cross-sections, the general relation is:
M Ed ≤ M N , Rd (EN 1993-1-1 [6], rel. (6.31)) (2.115)

where MN,Rd is the design plastic moment resistance reduced due to the axial
force NEd;
• for class 3 and class 4 cross-sections, in the absence of shear force, the general
relation is:
fy
σ x ,Ed ≤ (EN 1993-1-1 [6], rel. (6.42), (6.43)) (2.116)
γ M0
where σx,Ed is the design value of the local longitudinal stress due to moment
and axial force taking account of fastener holes where relevant.
For bisymmetrical I- and H- sections or other flanges sections, no reduction of the
plastic resistance moment MN,y,Rd about the y-y axis needs to be done when both the
following criteria are satisfied:
N Ed ≤ 0,25 ⋅ N pl, Rd and (EN 1993-1-1 [6], rel. (6.33)) (2.117)

0,5 ⋅ h w ⋅ t w ⋅ f y
N Ed ≤ (EN 1993-1-1 [6], rel. (6.34)) (2.118)
γ M0
For bisymmetrical I- and H- sections or other flanges sections, no reduction of the
plastic resistance moment MN,z,Rd about the z-z axis needs to be done if:
hw ⋅ tw ⋅ fy
N Ed ≤ (EN 1993-1-1 [6], rel. (6.35)) (2.119)
γ M0
Otherwise, if the requirements in relations (2.117), (2.118), or (2.119) are not fulfilled,
the following reductions must be done:
1− n
M N , y , Rd = M pl, y , Rd ⋅ but M N , y, Rd ≤ M pl, y, Rd (EN 1993-1-1 [6], rel. (6.36))(2.120)
1 − 0,5 ⋅ a
for n ≤ a: M N ,z , Rd = M pl,z , Rd (EN 1993-1-1 [6], rel. (6.37)) (2.121)

  n − a 2 
for n > a: M N ,z , Rd = M pl,z , Rd ⋅ 1 −    (EN 1993-1-1 [6], rel. (6.38)) (2.122)
  1 − a  
where:

60
N Ed
n=
N pl, Rd

A − 2 ⋅ b ⋅ tf
a= but a ≤ 0,5
A
For bi-axial bending:
α β
 M y , Ed   M z , Ed 
  +  ≤1 (EN 1993-1-1 [6], rel. (6.41)) (2.123)
 M N , y, Rd   M N ,z , Rd 
in which α and β are constants, which may conservatively be taken as unity,
otherwise as follows:

• I and H sections:

α = 2 ; β = 5n but β ≥ 1

• circular hollow sections:

α = 2 ;β = 2

• rectangular hollow sections:

1,66
α=β= but α = β ≤ 6
1 − 1,13 n 2

2.8.2. Buckling checks according to EN 1993-1-1 [6]

EN 1993-1-1 [6] uses a pair of two relations for buckling check of a member in
compression and bending. These relations can lead to different results, as some of
the factors that intervene can be calculated according to two different procedures,
given in Annexes A and B. The checking relations are the following ones:
N Ed M y , Ed + ∆M y , Ed M + ∆M z , Ed
+ k yy + k yz z , Ed ≤1
χ y N Rk M M z , Rk
χ LT y , Rk
γ M1 γ M1 γ M1

(EN 1993-1-1 [6], rel. (6.61)) (2.124)


N Ed M y , Ed + ∆M y , Ed M + ∆M z , Ed
+ k zy + k zz z , Ed ≤1
χ z N Rk M M z , Rk
χ LT y , Rk
γ M1 γ M1 γ M1

(EN 1993-1-1 [6], rel. (6.62)) (2.125)

61
where NEd, My,Ed and Mz,Ed are the design values of the compression force and the
maximum moments about the y-y and z-z axis along the
member, respectively

∆My,Ed, ∆Mz,Ed are the moments due to the shift of the centroidal axis for
class 4 sections

χy and χz are the reduction factors due to flexural buckling

χLT is the reduction factor due to lateral torsional buckling

kyy, kyz, kzy, kzz are the interaction factors that depend on the chosen
method (from Annex A or from Annex B)
In the Romanian National Annex, the method in Annex A is recommended.
Depending on the class of the cross-section, characteristics given in table 2.12
should be used.

Tab. 2.12. Values for NRk = fy Ai, Mi,Rk = fy W i and ∆Mi,Ed (EN 1993-1-1 [6], Tab. 6.7)

Class 1 2 3 4
Ai A A A Aeff
Wy Wpl,y Wpl,y Wel,y W eff,y
Wz Wpl,z Wpl,z Wel,z W eff,z
∆My,Ed 0 0 0 eN,y NEd
∆Mz,Ed 0 0 0 eN,z NEd

Figure 2.35 illustrates the values of the maximum bending moment


that can be acting on a bar in the presence of an axial compression
force having different values. The shape of the bending moment

Example diagram is triangular, with the maximum value at one end and 0
(zero) at the other end. The cross-section of the bar has a 10 × 400
mm web and the flanges are 20 × 300 mm. The length of the bar
considered in this example is 5m.

It can be noticed that there are important differences among the


values obtained using the five different checking relations:
• one relation for resistance;

62
• two buckling relations using Annex A factors;
• two buckling relations using Annex B factors.
The resistance of the bar is obtained as the minimum given by the
three relations, depending on the Annex that is used.

Fig. 2.35. Interaction N – M diagram

2.8.3. Buckling checks according to ANSI/AISC 360-16 [7]

ANSI/AISC 360-16 [7] uses the following relations for buckling check of members in
compression and bending:
a) when Pr/Pc ≥ 0,2:

Pr 8  M rx M ry 
+ ⋅ + ≤ 1,0 (ANSI/AISC 360-16 [7], rel. (H1-1a)) (2.126)
Pc 9  M cx M cy 

b) when Pr/Pc < 0,2:

Pr M M ry 
+  rx +  ≤ 1,0 (ANSI/AISC 360-16 [7], rel. (H1-1b)) (2.127)
2 ⋅ Pc  M cx M cy 

where:
Pr – required axial strength;

63
Pc – available axial strength;
Mr – required flexural strength;
Mc – available flexural strength;
x – subscript relating symbol to major axis bending;
y – subscript relating symbol to minor axis bending.

2.8.4. Commentary

Given the complexity of the interaction formulae for checking for


compression and bending, it is not easy to define the most
unfavourable load combination (design situation). In most of the
Important situations, it is defined either by the maximum value of the bending
moment MEd,max and the corresponding value of the axial force NEd,
or by the maximum value of the axial force NEd,max and the
corresponding value of the bending moment MEd. However, nobody
guarantees that a combination having a smaller value of the
bending moment and a smaller value of the axial force is not
more unfavourable, as the shape of the bending moment diagram
has a major influence on the results. Therefore, the safest
approach is to perform the checks in all the load combinations
that were considered.

2.9. THE USE OF BUCKLING LENGTH FOR ARCHES AND FRAMES

There are situations when simplified relations are useful for the design of arches or
single storey frames. The following relations can be used [18].
• Arches – for two- and three-hinged arches (Fig. 2.36) with the ratio h/L
between 0,15 and 0,5 and essentially uniform cross-section, the in-plane
buckling length may be taken as:
L cr = 1,25 ⋅ s (2.128)
where s is half of the arch length;

64
Fig. 2.36. Buckling length for a two-hinged arch [18]

• Two- and three-hinged frames – if the inclination of the columns is less than
15°, the following relation can be used for the column buckling length (Fig.
2.37):
I⋅s E⋅I
L cr = h ⋅ 4 + 3,2 ⋅ + 10 ⋅ (2.129)
I0 ⋅ h h ⋅ Kr

where Kr is the stiffness of the semi-rigid connection in the joint;

Fig. 2.37. Buckling length for a three-hinged frame [18]

For the buckling length of the rafter, the following relation can be used:
I⋅s E⋅I I ⋅N
L cr = h ⋅ 4 + 3,2 ⋅ + 10 ⋅ ⋅ 0 (2.130)
I0 ⋅ h h ⋅ Kr I ⋅ N0

where N and N0 are the axial forces in the column and in the rafter; in the
case of tapered cross-sections, the cross-sections at 0,65h or 0,65s (Fig.
2.37) respectively should be used;

65
• Columns or rafters with knee bracing – the in-plane buckling length for
columns (Fig. 2.38(a)) and for rafters (Fig. 2.38(b)) can be estimated as:
L cr = 2 ⋅ s l + 0,7 ⋅ s 0 (2.131)

(a) (b)
Fig. 2.38. Portal frame (a) and frame with V-shaped columns (b) [18]

• Torsional buckling of spatial frames – for the rotational buckling of axi-


symmetrical spatial structures (Fig. 2.39), the following approximate relation
can be used for determining the buckling length factor k, provided that 1 < k <
2 and a/s < 0,2.

Fig. 2.39. Rotationally symmetric spatial frame [18]

a 3 ⋅ π2 ⋅ a ⋅ E ⋅ I
k = 1 + 2 ⋅ + 10 ⋅ (2.132)
s 4 ⋅ s 2 ⋅ (1 + a s ) ⋅ K r

where EI is the bending stiffness of the rafter for bending about the vertical
axis and Kr is the rotational stiffness of the connection between the rafter and
the compression ring for bending about the vertical axis.

66
Chapter 3

INSTABILITY OF PLATES

3.1. PLATE BUCKLING MODELS IN EN 1993-1-5 [19]

Local buckling is a very important form of instability of plates that


governs the resistance of cross-sections made of plates subjected to
in-plane compression normal stresses, to shear stresses or to their
Chapter combined effect. The European code that deals with the behaviour of
introduction steel cross-sections subject to local buckling is EN 1993-1-5 [19].

EN 1993-1-5 [19] uses two methods for considering plate buckling effects [20]:
• the effective cross-section approach (metoda secțiunii transversale eficace);
• the reduced strength method (metoda tensiunilor reduse).
The relations are developed presuming the resistance of the cross-section is
governed by the most compressed fibre. The following assumptions are considered
to be fulfilled [20]:
• the individual plate elements are quasi-rectangular (the angle between the
longitudinal edges is less than 10°);
• stiffeners, if any, are placed on orthogonal directions about the direct stresses;
• openings and cut outs, if any, are small;
• the cross-section is uniform;
• flange induced web buckling is prevented by constructional recommendations.
For panels having the edges at angles higher than 10°, it is conservative to consider
a panel having the dimensions equal to the maximum of the actual ones. The
openings are considered to be less than 5% of the element width.

67
Both methods are equivalent for single plate elements; differences can be noticed in
the case of cross-sections composed of several plates. The effective cross-section
approach allows the use of cross-sections with higher slenderness than the reduced
strength method and, as a result, the importance of serviceability limit state
requirements increases in this case [20].

3.1.1. Main aspects of the effective cross-section approach

In this method, a separate check is made for each individual loading situation. Then,
interaction formulae are used to check for the combined effects. The design is
governed by the yield strength in the most compressed fibre of the reduced cross-
section of the member [20].

The main steps of using the effective cross-section method are as follows [20]:
1. Determine the stress distribution based on the full cross-section;
2. Based on this distribution, determine the reduced cross-section of each individual
plate that is part of the cross-section;
3. Determine the stress distribution on the obtained effective cross-section,
consisting of reduced cross-sections of each component plate;
4. Establish the effective cross-section, based on the stress distribution from step 3.
5. Repeat steps 3 and 4 till the stress distribution is compatible with the cross-
section.
The post-buckling strength of a plate depends on the aspect ratio and on the
possible presence of stiffeners. Therefore, three models are used for the behaviour
of a plate element:
• the plate type behaviour;
• the column type behaviour;
• an interpolation zone.

68
3.1.1.1. Plate behaviour

A correct analysis of the elastic critical plate buckling stress can be done by means
of appropriate computer programs. Alternatively, EN 1993-1-5 [19] provides two
simple approaches, depending on the number of longitudinal stiffeners that are
present in the compression zone:
• multiple stiffeners (three or more);
• one or two stiffeners.
In the case of multiple stiffeners, the element is treated like an orthotropic plate with
smeared stiffeners, meaning that the total rigidity of the stiffeners is distributed on
the plate, transforming it into a fictitious one.

The effective cross-section is obtained by using a reduced width for the compression
parts.
b c ,eff = ρ ⋅ b c (3.1)

where:
ρ – the reduction factor for plate buckling;
bc – the actual width of the compressed part;
ρ depends on the stress distribution ψ across the width of the plate and on the
support condition along the longitudinal edges. It is calculated as follows:
• for internal compression elements (two longitudinal edges supported):
ρ = 1,0 for λ p ≤ 0,673 (3.2)

λ p − 0,055 ⋅ (3 + ψ )
ρ= ≤ 1,0 for λ p > 0,673 , where (3 + ψ ) ≥ 0
λ2p

(EN 1993-1-5 [19], rel. (4.2)) (3.3)


• for outstand compression elements (one longitudinal edge supported and the
other one free):
ρ = 1,0 for λ p ≤ 0,748 (3.4)

λ p − 0,188
ρ= ≤ 1,0 for λ p > 0,748 ([19], rel. (4.3)) (3.5)
λ2p

where λ p is the relative plate slenderness, which is determined in a similar manner

as a column slenderness [20], as square root of the ratio between the characteristic

69
resistance of the cross-section and the critical force, considering Ac,eff the effective
area of the compression zone:
A c,eff ⋅ f y fy
λp = = (3.6)
A c,eff ⋅ σ cr σ cr

Considering E=210000 N/mm2, ν=0,3

π2 ⋅ E
2
t
σ cr = k σ ⋅ σ E = k σ ⋅ ⋅ 
12 ⋅ (1 − ν )  b 
2
(3.7)

fy b t
λp = = (3.8)
σ cr 28,4 ⋅ ε ⋅ k σ

where:

235
ε=
fy

kσ – the buckling factor corresponding to the stress ratio ψ and boundary


conditions; for long plates, it is given in tables 3.1 and 3.2.
For simply supported unstiffened compression plates, including wall elements with
longitudinal stiffeners, subjected to uniform compression, kσ is calculated as:

α n 
2

kσ =  +  (3.9)
 n α
where:
a
α=
b
a – length of the considered panel in the direction of the direct stress;
b – width of the considered panel in the direction of the direct stress;
n – number of half sine waves in the direction of compression corresponding to α.

70
Tab. 3.1. Internal compression elements (EN 1993-1-5 [19], Tab. 4.1)

Tab. 3.2. Outstand compression elements (EN 1993-1-5 [19], Tab. 4.2)

71
3.1.1.2. Column behaviour

The previous relations, corresponding to plate behaviour, include a post-buckling


strength reserve. However, for small α ratios (i.e. <1 for unstiffened plates), a column
behaviour may be expected, with less post-buckling reserve. In this case, a reduction
factor χc corresponding to column buckling is used. This reduction is more severe
than ρ. The modelling of column behaviour is obtained by removing the longitudinal
supports. Consequently, for an unstiffened plate, the critical stress is:

π2 ⋅ E
2
t
σ cr ,c ⋅ 
12 ⋅ (1 − ν )  a 
= 2
(EN 1993-1-5 [19], rel. (4.8)) (3.10)

fy
λc = (EN 1993-1-5 [19], rel. (4.10)) (3.11)
σ cr ,c

and the reduction factor χc is obtained using the relations that describe the buckling
curves in EN 1993-1-1 [6], where the factor α is replaced by a factor αe, defined in
EN 1993-1-5 [19].

3.1.1.3. Interaction between plate and column buckling

The final reduction ρc factor is obtained by interpolation between χc and ρ:


ρ c = (ρ − χ c ) ⋅ ξ ⋅ (2 − ξ ) + χ c (EN 1993-1-5 [19], rel. (4.13)) (3.12)

where:
σ cr ,p
ξ= − 1 but 0 ≤ ξ ≤ 1
σ cr ,c

σcr,p – the elastic critical plate buckling stress;


σcr,c – the elastic critical column buckling stress.
The image of the interpolation in relation (3.12) is given in figure 3.1.

72
Fig. 3.1. Interpolation between column behaviour χc and plate behaviour ρ [20]

3.1.1.4. Checks for instability produced by bending moment (and axial force)

The checking relations are as follows:


• for uniaxial bending:
N Ed M + N Ed ⋅ e N
η1 = + Ed ≤ 1,0 (EN 1993-1-5 [19], rel. (4.14)) (3.13)
f y ⋅ A eff f y ⋅ Weff
γ M0 γ M0

• for biaxial bending:


N Ed M y ,Ed + N Ed ⋅ e y, N M z ,Ed + N Ed ⋅ e z, N
η1 = + + ≤ 1,0
f y ⋅ A eff f y ⋅ Wy ,eff f y ⋅ Wz ,eff
γ M0 γ M0 γ M0
(EN 1993-1-5 [19], rel. (4.15)) (3.14)
where:
eN – the shift of the position of the neutral axis.

73
3.1.1.5. Checks for instability produced by shear force

Slender web panels in shear can develop an important post-buckling resistance due
to the tension field that is formed after local buckling. The present day models used
in EN 1993-1-5 [19] are based on Höglund’s model [21].

The theoretical resistance to local buckling is obtained based on the assumption of


the presence of a rigid end post (sufficiently rigid transverse stiffeners are present on
the supports).

Notations a) No end post b) Rigid end post c) Non-rigid end post


Fig. 3.2. End supports (EN 1993-1-5 [19] Fig. 5.1)

The shear panel slenderness is:

f yw 3 f yw
λw = = 0,76 (EN 1993-1-5 [19], rel. (5.3)) (3.15)
τ cr τ cr

and the elastic shear buckling stress of a perfect shear panel is:
2
π2 ⋅ E  t 
τ cr = k τ ⋅ σ E = k τ ⋅ ⋅  
( )
12 ⋅ 1 − ν  h w 
2
(3.16)

where, for a plate between two consecutive transverse stiffeners [20]:


4,00
k τ = 5,34 + when α ≥ 1,0
α2 (3.17)
5,34
k τ = 4,00 + 2 when α < 1,0
α
EN 1993-1-5 [19] recommends that whenever
h w 72
o ≥ ⋅ε for an unstiffened web (3.18)
tw η

74
h w 31
o ≥ ⋅ ε ⋅ kτ for a stiffened web (3.19)
tw η
the resistance to shear buckling should be checked and transverse stiffeners should
be provided at the supports. The recommended values for η are:
o η = 1,20 for S235, S275, S355, S460;
o η = 1,00 for steel grades higher than S460.
and they were adopted in the Romanian National Annex of EN 1993-1-5 [22].

The relation given in EN 1993-1-5 [19] for calculation of the shear buckling
resistance is:
η ⋅ f yw ⋅ h w ⋅ t w
Vb,Rd = Vbw ,Rd + Vbf ,Rd ≤
3 ⋅ γ M1
(EN 1993-1-5 [19], rel. (5.1)) (3.20)
where Vbw,Rd is the contribution of the web:
χ w ⋅ f yw ⋅ h w ⋅ t w
Vbw ,Rd = (EN 1993-1-5 [19], rel. (5.2)) (3.21)
3 ⋅ γ M1
and Vbf,Rd is the contribution of the flanges, considered only when the resistance of
the flanges is not completely used for resisting the bending moment (MEd < Mf,Rd),
where:
χw – the reduction factor for the shear resistance, depending on the web
slenderness λ w ; the values are given in table 3.3;
γM1 – the partial factor for the resistance to instability;
η – coefficient that includes the increase of shear resistance at smaller web
slenderness [20];

Tab. 3.3. Values of χw (EN 1993-1-5 [19], Tab. 5.1)

Rigid end post Non-rigid end post


λ w < 0,83 η η η

0,83 η ≤ λ w < 1,08 0,83 λ w 0,83 λ w

λ w ≥ 1,08 (
1,37 0,7 + λ w ) 0,83 λ w

75
The presence of η bigger than 1,0 is explained by test results [20] that showed an
increase of the shear strength up to 0,7 … 0,8 times the yielding limit for webs with
reduced slenderness; there is also a contribution of the flanges but, as these two
influences could not be separated, η is not allowed for single plates, without flanges.

Vbf,Rd, the contribution of the flanges, is calculated as:

b f ⋅ t f2 ⋅ f yf   M Ed 
2
  (EN 1993-1-5 [19], rel. (5.8))
Vbf ,Rd = ⋅ 1−   (3.22)
c ⋅ γ M1   M f ,Rd 


 
where:
bf – width of the compressed flange, not larger than 15εtf on each side of the web;
Mf,Rd – moment resistance of the cross-section consisting of the effective area of the
flanges only (Fig. 3.3);

Fig. 3.3. Definition of Mf,Rd ([20] Fig. 5.11)

c – distance between the plastic hinges on the flange at the limits of the tension
field (Fig. 3.4).
According to Höglund [21], for steel plate girders, the distance c can be estimated
with the following approximate formula [20]:
 M pl,f   1,6 b f t f2 f yf 
c = a  0,25 + 1,6 ⋅  = a  0,25 +  ([20], rel. (5.25)) (3.23)
 M pl,w   t w h 2w f yw 
   

76
Fig. 3.4. Tension field carried by bending resistance of flanges ([20] Fig. 5.10)

where:
b f t f2 f yf
M pl,f = is the plastic moment resistance of the flanges (3.24)
4
t w h 2w f yw
M pl,w = is the plastic moment resistance of the web (3.25)
4
In the presence of an axial force NEd, the plastic moment resistance Mf,Rd must be
reduced by multiplying with the following factor:
 
 
1 − N Ed  (EN 1993-1-5 [19], rel. (5.9)) (3.26)
 (A f 1 + A f 2 ) ⋅ f yf 
 
 γ M0 

Usually, the contribution of the flanges is small and can be neglected. It may be
important at the end supports, where the flanges are less loaded. The values of c
observed in tests are usually larger [20].

Based on the previous approaches, the contribution of the flanges, Vbf,Rd, can be
calculated as (Fig. 3.4):
4 ⋅ M pl,f ,Rd b f ⋅ t f2 ⋅ f yf
Vbf ,Rd = = (3.27)
c ⋅ γ M1 c ⋅ γ M1
Presuming the bending moment is resisted only by the flanges, leading to axial
forces in the flanges that reduce their plastic moment resistance, a correction factor
is used:

77
  M 
2

1 −  Ed   (3.28)
  M f ,Rd  
   
and so, relation (3.22) is obtained.

The local buckling checking relation for shear is:


VEd
η3 = ≤ 1,0 (EN 1993-1-5 [19], rel. (5.10)) (3.29)
Vb,Rd

1. Dalban C., Dima S., Chesaru E., Şerbescu C. – Construcţii cu


structura metalică, Ed. Didactică si Pedagogică, 1997
2. Dima, Ş., Ştefănescu B. – Steel Structures – basic elements,
Conspress Bucureşti, 2005
Bibliography
3. ESDEP – The European Steel Design Education Programme,
http://www.haiyangshiyou.com/esdep/master/toc.htm
4. Tien T. Lan – Space Frame Structures, Structural Engineering
Handbook, Ed. Chen-Wai Fah, 1999, pag. 24-32
5. P100-1/2013 – Cod de proiectare seismică – Partea 1 –
Prevederi de proiectare pentru clădiri
6. EN 1993-1-1 – Eurocode 3: Design of Steel Structures – Part 1-1:
General rules and rules for buildings
7. ANSI/AISC 360-16 – Specification for Structural Steel Buildings,
AISC, USA, 2016
8. JSCE – Standard Specifications for Steel and Composite
Structures
9. SN007b-EN-EU – NCCI: Torsion, Access Steel
10. Trahair, N.S., Bradford, M.A., Nethercot D.A., Gardner, L. – The
Behaviour and Design of Steel Structures to EC3, 4th Edition.
Taylor & Francis, 2008
11. SN030a-EN-EU – NCCI: Mono-symmetrical uniform members
under bending and axial compression, Access Steel
12. ENV 1993-1-1 – Eurocode 3: Design of Steel Structures and
National Application Document – Part 1-1: General rules and
rules for buildings

78
13. SN008a-EN-EU – NCCI: Buckling lengths of columns: rigorous
approach, Access Steel
14. Diacu I., Ștefănescu B. – Stability bracing in practice, Steel – A
New and Traditional Material for Building, Proceedings of the
International Conference in Metal Structures – Poiana Brașov,
România, September 20-22, 2006, D. Dubină, V. Ungureanu
Editors, Taylor & Francis Group, 2006, pag. 109-117
15. Timoshenko, S.P., Gere, J.M. – Theory of elastic stability, 2nd
Edition. McGraw-Hill, 1961
16. SN003b-EN-EU – NCCI: Elastic critical moment for lateral
torsional buckling, Access Steel
17. LTBEAM – http://www.cticm.com/content/ltbeam-version-1011
18. Blass H.J. – Timber Engineering STEP 1: Basis of design,
material properties, structural components and joints, Almere:
Centrum Hout, 1995
19. EN 1993-1-5 – Eurocode 3: Design of Steel Structures – Part 1-5:
Plated structural elements
20. Johansson B., Maquoi R. , Sedlacek G. , Müller C. , Beg D. –
Commentary and Worked Examples to EN 1993-1-5 “Plated
Structural Elements”, JRC – ECCS Joint Report, First Edition,
October 2007, EUR 22898 EN – 2007
21. Höglund T. – “Shear buckling resistance of steel and aluminium
plate girders” – Thin-Walled Structures, Vol. 29, Nos. 1-4,
pag.13-30, 1997
22. SR EN 1993-1-5/NA – Eurocod 3: Proiectarea structurilor din oțel
– Partea 1-5: Elemente structurale din plăci plane solicitate în
planul lor – Anexa Națională

79

S-ar putea să vă placă și