Stability Lect Main Mukki 2
Stability Lect Main Mukki 2
Stability of Structures
Prof. Dr.-Ing. Gerhard Müller
Corinna Treimer, M.Sc.
For academic use only. Reproduction and distribution are not permitted without the permis-
sion of the author.
Stability of Structures
Winter Term 2022/2023
III
2.4.2 Linearized Buckling Analysis with Geometrical Stiffness Matrices . . 55
2.4.3 Geometrical Nonlinear Analysis and Eigenvalue Problem: Example . 57
2.4.4 Effective Length of Beam Columns of Plane Frames . . . . . . . . . 64
Bibliography 138
Martin Buchschmid
Structural analysis aims to prove that a structure can safely support the anticipated loadings
and provide sufficient serviceability. This includes the investigation of relations between ap-
plied forces and moments (due to self weight, wind load, traffic, etc.), imposed displacements
Stability of Structures
Winter Term 2022/2023
1
(support settlements etc.), or strains (due to thermal expansions, shrinkage, creep, etc.), and
the internal forces and support reactions, stresses, strains, and displacements caused by these
actions.
Quantities of class b) only occur due to applied actions, whereas quantities of classes a), c),
and d) can be both imposed and caused.
The relations between the different classes can be illustrated as follows:
external internal
forces (1)
p σ
(4) (2)
displacements u ε
or strains (3)
Figure 1.1
Relation (1) involves equations of equilibrium usually relating externally applied forces and
moments. These equations can be derived considering either the undeformed or the
deformed system, according to the assumptions of the mechanical model employed. Both
the differential considerations and suitable energy methods (e.g., principle of virtual
work) yield the same results because they are essentially equivalent.
Relation (2) is governed by one or more different constitutive laws depending on the used
material. These laws relate strains to corresponding stresses or internal forces and mo-
ments, respectively. It is, as per definition, required that measures of strain correspond
to suitable measures of stress to give valid expressions of work.
Relation (3) between displacements and strains is obtained from the kinematic assumptions
of a mechanical model. Whether these assumptions are justified can be examined using
a more accurate model or via experiment.
If all of them are linear, then relation (4) will also be linear. In this case, we can always use
linear superposition, e.g., to split up loads in symmetric and anti–symmetric components. The
solutions of analyses of linear systems are always clearly defined.
If one or more of the relations are nonlinear, then relation (4) will be nonlinear. This implies
that the principle of superposition is not valid in general. Furthermore, there might exist
several equilibrium paths to certain problems.
Depending on the degree of linearization, we have the choice between various theories and
methods of computation (see table). On which theory we can base the structural analysis,
and which one is necessary or most advantageous, is determined by the structure itself and its
loadings (i.e., the validity of linearizations in any of the relations (1) to (3)), and by economical
and practical considerations. A higher degree of linearization simplifies the computation, but
at the same time we get a less realistic model of the behaviour of a structure.
b1
linear in general
(1): Equilibrium
non-linear
linear
linear
linear
(4): Applied Actions –
Displacements
d linear
c1 κy , κz , ψ, κω in general
u
v (3): Kinematics
c2
w φ(xi )
non-linear in plastic
ϑ
hinges
Figure 1.2: Applied and internal forces respectively stresses, strains and displacements in the analy-
sis of beam-like structures
1 1
2 2 AV
Virtual displacements are imposed to the loaded system in order formulate the equilibrium
with the help of the virtual work theorem.
After cutting the left support and introducing the vertical reaction force AV , the system shows
one single degree of freedom of motion; the deflected beam and the applied loads can rotate
about the right hand pin support as the centre of rotation to a neighboring state.
1
e.g. in frame structures
wA (δφ)
wP (δφ)
AV (δφ)
Figure 1.4: Finite displacement
If we apply a finite angle of rotation, the support reaction and the internal forces will change
while moving from the deformed position to the neighboring state. Hence, we will have a
completely non–linear problem.
Thus, we conclude:
1. To keep the real state of stress constant while imposing virtual displacements, a purposive
virtual displacement needs to be infinitesimally small.
P
δφ
δwA (δφ)
δwP (δφ)
AV
3. Because of the resulting linear kinematic relations, virtual displacements may be written
in any size of order.
4. Further, the displacements of the structure in the deformed state may be disregarded.
δWi + δWe = 0
δwP δwA
=
1
2
1
1
δwP = δwA
2
1
− AV δwA + P̄ δwA = 0
2
1
(−AV + P̄ ) δwA = 0
2
1
AV = P̄
2
fh
Figure 1.6: Imperfect structural system
fh = l sin(φ0 + φ) (1.2)
During an imposed virtual displacement, internal work is performed in the elastic spring and
external work is done by the applied forces P and H.
The virtual quantities δφ, δfv and δfh depend upon each other by the prescribed kinematic
relations of the system. Because they are infinitesimal small, they may be calculated as linear
terms of a Taylor series expansion:
df 1 d2 f
f (φ + dφ) = f (φ) + dφ + dφ2 + . . .
dφ 2 dφ2
dfv
δfv = δφ = l sin(φ0 + φ)δφ (1.6)
dφ
dfh
δfh = δφ = l cos(φ0 + φ)δφ (1.7)
dφ
δφ
In contrast to real movement, the virtual displacement does not occur on a circular arc but
on a straight line being at right angels to the radius ray (see Section 1.3.1).
Because of the assumed arbitrary and non-vanishing virtual displacements, the sum of
the terms in brackets vanishes.
kφ φ − P l sin(φ0 + φ) − H l cos(φ0 + φ) = 0
kφ φ − P l sin(φ0 + φ)
=Hl (1.9)
cos(φ0 + φ)
a(P, φ, φ0 ) = P
If we allow only vertical loads, the above relation (1.9) converts into:
kφ φ − P l sin(φ0 + φ) = 0 (1.10)
kφ φ
P = (1.11)
l sin(φ0 + φ)
P (φ = 0) = 0
dP kφ
(φ = 0) =
dφ l sin φ0
dP
→∞
dφ
kφ φ
P = (1.13)
l sin φ
kφ φ kφ kφ
lim P (φ) = lim = lim =
φ→0 φ→0 l sin φ φ→0 l cos φ l
dP kφ sin φ − φ cos φ
=
dφ l sin2 φ
dP kφ sin φ − φ cos φ
lim = lim
φ→0 dφ φ→0 l sin2 φ
kφ cos φ − (cos φ − φ sin φ)
= lim
φ→0 l 2 sin φ cos φ
kφ φ
= lim =0
φ→0 2 l cos φ
Model of Analysis 2: Linearization for the purpose of a second order theory of small displace-
ments
kφ φ − P l (φ0 + φ) − H l = 0
A(P)V = P
In particular: H = 0
kφ φ
P = (1.16)
l (φ0 + φ)
dP kφ (φ + φ0 ) − φ kφ φ0
= =
dφ l (φ + φ0 ) 2 l (φ0 + φ)2
φ=0: P =0
dP kφ kφ
= <
dφ l φ0 l sin φ0
Perfect structure: φ0 = 0
kφ
P crit = (1.18)
l
Model of Analysis 1: Analysis without taking displacements into account (first order analy-
sis).
kφ φ − H l = 0 KV = P (1.19)
A(P, φ, φ0 ) = P
Model of Analysis 5: Linearization for a second order theory of small displacements, taking
into account material non-linearities
K(P, φ) V = P(φ0 )
kφ (φ) φ − H l = 0 (1.22)
K(φ) V = P
P = kφ φ
l sin(φ)
P = kφ φ
l sin(φ0 +φ)
P = kφ
l
P = kφ φ
l φ0 +φ
(a) φ = 0 − 2 π
P̄ klφ
(b) φ = 0 − 58 π
Figure 1.8: Results for P̄ sketched for different models of analysis
EA → ∞
EA
EA
P
2
w
l h
dα l + ∆l
α0 α
Constitutive Law
1q 2
σ = εE = a + (h − w)2 − 1 E
l
1.4.3 Equilibrium
P.V.D.:
Z
σ δε dV − P δw = 0
P , σ: applied load and stresses in the deformed state; they are assumed to be constant on
virtual displacements and virtual strains
δε, δw: infinitesimally small, starting from the deformed state, which is equilibrated
Virtual displacement:
dε
δε = δw
dw
1 1h 2 i− 1
= a + (h − w)2 2 2 (h − w) (−1) δw
l 2
h−w
=− q δw
l a2 + (h − w)2
l l + ∆l δw
h α
h−w
∆α δα
α
From Fig. 1.11, we read (bearing in mind that the elongation of the bars is related to the
unstrained length):
δw sin α
−δε =
l
δw (h − w)
=
l (l + ∆l)
h−w
= q δw compare above
l a2 + (h − w)2
2 snap through
P =0
w
tensile force
snap back
P =0
−P
−P P =0
1. The central axis (x-axis) of a beam coincides with its centroidal axis.
2. The axes of a cross section (y- and z-axes) are principal axes. Moreover, the x- and
z-axes lie within the plane of the structure.
3. The structure consisting of individual beams, or the structural system itself, may have
imperfections according to the pertinent standards (indicated by the subscript 0).
4. Local failures or out-of-plane failures are either prevented by constructive measures or
separate analysis is performed.
5. The effects of strains caused by shearing stresses or shear forces are negligible (Bernoulli
hypothesis).
6. Linear relations according to the engineering beam theory are assumed between strains
(axial strains and curvatures) of the central axis and its displacement.
Hence, we will develop a theory of small displacements.
7. A homogeneous and isotropic material is assumed. Furthermore, according to the en-
gineering beam theory, the normal stresses are governed by a one-dimensional linear
constitutive law.
Thus, only small strains are admissible.
We will discuss, however, the effects of a nonlinear constitutive law in Elastic-Plastic Analysis of Fr
.
8. The structural system under investigation is generally free of stresses in the undeformed
configuration.
9. The loads act within the plane of the structure.
10. The loads do not change their directions while causing structural deformations.
Stability of Structures
Winter Term 2022/2023
20
11. Any distributed axial load may be applied on a beam, which may cause a varying normal
force diagram in a first order analysis. However, for the iteration process of a second
order analysis, we will assume a constant (possibly averaged) normal force in a beam
defined by two nodes.
Here, the relevant relations of bending are compiled. We confine our interest to a second order
theory of bending because we assume that the basic relations for the elongation of bars under
normal forces hold unchanged.
Kinematic relations:
φ = −w′ (2.1)
κ = φ′ (2.2)
Constitutive law:
M = EI(κ − κ) (2.3)
∆T
κ = αT + κS + κK + κpl (2.4)
h
pz (x)
x
w0 (x) w0 (x + dx)
z, w0 , w w(x)
w(x + dx)
L
M (x) T (x) L
T (x + dx) M (x + dx)
kw (x) · w(x)
kφ w′ (x)
Equilibrium:
L′ = 0 (2.5)
T ′ − kw w + pz = 0 (2.6)
Taking the derivative of (2.7) with respect to x and substituting (2.6) yields
Normal and shearing forces and longitudinal, respectively transverse forces (within the range
of validity of the 2nd order analysis for small quantities w′ + w′0 and with |(w′ + w′0 ) · T | ≪ |L|
and |(w′ + w′0 ) · Q| ≪ |N |(Q ≪ N and T ≪ L)):
N ≈L (2.9)
Relevant quantities:
φ = −w′ (2.12)
M = −EI(w′′ + κ) (2.13)
Governing equation:
!
M ′ = −EIw′′′ − EI ′ w′′ − EIκ′ − EI ′ κ = Q(x) − kφ w′ − my = 0
For beams without a spring bed and without a distributed bending moment, the extreme value
of the bending moment M (x) coincides with zero values of the shearing force Q(x).
Remark 2: All terms of eq. (2.16) have units of a distributed force pz : [force/length]
Thus, this relation describes the statically indeterminate split of pz (as well as of the quantities
m′y , L w′′0 and (EIκ)′′ ) to the different components of the resistance of the structural member:
(EIw′′ )′′ (Bending), − Lw′′ (Th.II.O.), kφ w′′ (spring bed), kw w (spring bed).
Remark 3: Concerning the mechanical meaning of destabilizing forces of a second order theory:
N w′′
N w′ N x
′
w
may hold:
1 −(w′ + w′0 )
Q T
≈ ′
N w + w′0 1 L
1 w′ + w′0
T Q
≈
L −(w′ + w′0 ) 1 N
Thus, with the additional assumptions |(w′ + w′0 ) · T | ≪ |L| and |(w′ + w′0 ) · Q| ≪ |N |
(Q ≪ N und T ≪ L), we have:
N ≈L (2.17)
Stiffness matrices, vectors of fixed end forces and moments, transfer matrices, and load vectors
are denoted as element matrices in this context.
To determine the stiffness matrix of a bar we only need the homogeneous part of eq. (2.16):
cT = [C1 C2 C3 C4 ] (2.23)
w(x) = NF c (2.24)
The functions of the row matrix NF represent a fundamental set of solutions of eq. (2.20).
Hence, we can present the displacements and stress resultants of equations (2.12), (2.13), and
(2.14) from page 22 as functions of the constant column matrix c (the entries of which can
not be interpreted in a mechanical sense).
w(x) NF C1
−N′F
φ(x) C
2 Nu
= =
c
T (x) −EI NF − EI NF + (L + kφ )NF
′′′ ′ ′′ ′ C Ns
3
M (x) −EI N′′F C4
These four constants in (2.25) can be expressed in dependence of the four displacements and
rotations of the member ends.
w(xa ) NF (xa )
wa C1
−w ′ (x ) −N′ (x ) C Nu (xa )
φ
a a F a 2
= = = v = [G]−1 c (2.26)
c
wb w(xb ) NF (xb ) C3 Nu (xb )
φb −w′ (xb ) −N′F (xb ) C4
c = Gv (2.27)
w(x) = N v (2.30)
The functions of the row matrix NF are the exact shape functions of the member.
We obtain the exact stiffness matrix in formulating the transverse forces and bending moments
at the ends of the element (in sign convention 2):
−T (xa )
Ta
−M (x ) −Ns (xa ) −Ns (xa )
M
a a
= = G v=k v =⇒ G=k (2.32)
Tb T (xb ) Ns (xb ) Ns (xb )
Mb M (xb )
To remember: A column of the stiffness matrix contains the end forces and end moments if
you set the related end displacement equal to +1 and the rest of them are 0.
My
Pz sign convention 2
my1 my2
pz1 pz2
φa φb
z
wa wb Ma Ta Tb Mb
ls
a l
Positive sense of direction of end displacements, loads and end forces and moments.
Governing equation:
wh (x) = C̃1 1 + C̃2 x + C̃3 sin ε̃x + C̃4 cos ε̃x (2.34)
s
|N |
ε̃ = (2.35)
EI
s
|N |
Characteristic parameter: ε=l (2.38)
EI
Remark 2: Expression (2.38) denotes the only mechanical descriptive quantity for the charac-
teristic parameter. (The same result is obtained by inserting ε̃ = εl in (2.34).)
Remark 3: Stiffness matrix in a local coordinate system:
EA EA
0 0 − 0 0
l l
EI EI EI EI
0 [2(A′ +B ′ )−D′ ] 3 −(A′ +B ′ ) 2 0 −[2(A′ +B ′ )−D′ ] 3 −(A′ +B ′ ) 2
l l l l
′ EI ′ EI ′ EI ′ EI
′ ′
0 −(A +B ) 2 A 0 (A +B ) 2 B
l l l l
k =
EA EA
− 0 0 0 0
l l
EI EI EI EI
−[2(A′ +B ′ )−D′ ] 3 (A′ +B ′ ) 2 [2(A′ +B ′ ) − D′ ] 3 (A′ +B ′ ) 2
0 0
l l l l
′ EI ′ EI ′ EI ′ EI
′ ′
0 −(A +B ) 2 B 0 (A +B ) 2 A
l l l l
(2.39)
ε (ε − sin ε) |N |
B′ = D ′ = ε2 = l 2
2 (1 − cos ε) − ε sin ε EI
π 2 EI
Pcrit =
l2
Pz
Px
stress problem
− Normal Force (≈ L = −P x )
Governing equation:
wh (x) = C1 e−ε2 x cos ε1 x + C2 e−ε2 x sin ε1 x + C3 eε2 x cos ε1 x + C4 eε2 x sin ε1 x (2.42)
Boundary disturbances of a beam column on a continuous elastic support are fading if the bar
is sufficiently long:
P
e
If N vanishes, the characteristic parameters are taking the form of the beam on a spring bed.
EI
The exponential terms e±ε2 x take a value of 1 and we receive the homogenous solution:
s √
|N | ± N 2 − 4 EI kw
3. N > 4 EI kw
2
ε1,2 = (2.46)
2 EI
Remark 1: We can derive the critical compressive force of a simple beam column on hinge
supports in a different approach:
w(0) = 0 w(l) = 0
nπx
Trial function: w = C sin (2.48)
l
π nπx
w′ = C n cos (2.49)
l l
2
π nπx
w = −C n
′′
sin (2.50)
l l
All boundary conditions are satisfied. If we substitute the above trial function in the governing
differential equation, we have:
4 2
nπ nπx nπ nπx nπx
EI C sin − |N | C sin + kw C sin =0
l l l l l
4 2
nπ nπ
EI − |N | + kw = 0 (2.53)
l l
Hence,
" #
l2 n4 π 4 n2 π 2 l2
Pcrit = EI +kw = EI + kw
n2 π 2 l4 l2 n2 π 2
s !2 #
1
" # "
l4 kw EI l4 kw EI
= n2 + 2 4 π2 2 = n2 + π2
n π EI l nπ 2 EI l2
EI
= φ28 π 2 (2.54)
l2
s 2
1 l4 kw
φ28 (n) = n2 + 2 (2.55)
nπ EI
n=4
n=5
n=3
q 2
n = 2 φ28 (n) = n2 + 1
n·π 2
l4 kw
EI
n=1 q
l 4 kw
EI
q
l4 kw
Figure 2.1: The system parameters l, kw , EI are included in the term EI on the x-axis. On the
y -axis there is the critical load Pcrit =φ28 π 2 EI
l2in dependency of the parameter φ28 .
For a given system, corresponding to one point on the x-axis, one can determine the
critical load of the system by reading the lowest value of φ28 , given by the red envelope
curve. Additionally q on can deduce the failure mode n that occurs for this combination of
4
l, kw , EI . E.g. for l EI
kw
= 40 the system will collapse with n = 2 sinus half waves, as
Pcrit for n = 2 is smaller as Pcrit for n = 1.
2.2.3.6 Beam with Constant Width and Linearly Varying Height, EI = var.; N =0; kw = 0
b a a l b b
x
x h = ha (2.58)
ha x ξ hb a
x3
EI(x) = EIa (2.59)
a3
Ma Mb
Qa Qb
d 1 d
g= h(x) = h(x) ≤ 0.2
dx l dξ
Governing equation:
EIa h 3 ′′′′ i
3
x w +6x 2 ′′′
w +6x w ′′
= pz + m′y −[EI(x) κ]′′ (2.60)
a
1
wh (x) = C1 1+ C2 x+ C3 + C4 ln x (2.61)
x
Remark 1: Stiffness matrix (bending) for a beam of constant width and linearly varying depth
(linear theory of 1st order):
l
α= β =1+α η = (2 + α) ln β − 2α
a
α3 (α+2) α3 α3 (α+2) β α3
− 2 − −
l3 l l3 l2
Ta
wa
2β ln β − 3α − 2α
2 2
α 3
β (2β ln β − α2 − 2α)
Ma EIa
− ϕa
=
l l2 l
α (α+2)
3
β α3
Tb
η
wb
l3 l2
Mb
ϕb
β 2 (2 ln β − α (2−α))
symm.
l
p = k Iexakt v
h ba x x bb h
ba a l bb
a b
ba ha x x bb hb
ba a l bb
a b
ha x x bb hb
d ah
ba ah l bb
a b tF
tF
ts ts ha x x ξ bb hb ts ts
tF
tF
d ah
ba ah l bb
2.2.3.7 Beam Column with Constant Width and Linearly Varying Height,
EI = var.; N = const.>0; kw = 0
x
h = ha (2.62)
a
x3
EI(x) = EIa (2.63)
a3
Governing equation:
EIa h 3 ′′′′ i
x w + 6x 2 ′′′
w + 6x w ′′
− N w′′ = pz + m′y − [EI(x) κ]′′ (2.64)
a3
√ 2ε √ 2ε
! !
wh (x) = C1 1 + C2 x + C3 x J1 √ + C4 x N1 √ (2.65)
x x
s
|N |
Characteristic parameter: ε= a3 (2.66)
EIa
Bessel and Neumann functions of the first order compared with trigonometric functions:
sin(x)
cos(x)
Governing equation:
Governing equation:
wh (x) = C1 e−ε1 x cos ε2 x + C2 e−ε1 x sin ε2 x + C3 eε1 x cos ε2 x + C4 eε1 x sin ε2 x (2.72)
s s
kw N
N = 4EI kw ε= = (2.73)
2 4
EI 2 EI
v
√
− 4EI kw
u
u |N | ± N 2
N 2 > 4EI kw ε1,2 = (2.75)
t
2EI
wh (x) = C̃1 cosh ε1 x + C̃2 sinh ε1 x + C̃3 cosh ε2 x + C̃4 sinh ε2 x (2.76)
We start with equation (2.16) (Govering Differential Equation of bending of a beam acc. Th
II. Order with elastic bedding):
Z
(EIw ) − (L + kφ ) w + kw w − pz −
′′ ′′ ′′
m′y − L w′′0 + (EIκ) = 0 | · δw(x)|
′′
. . . dx
Z bh i
(EIw′′ )′′ − (L + kφ ) w′′ + kw w − pz − m′y − L w′′0 + (EIκ)′′ δw(x)dx = 0
a
Partial Integration:
Z b Z b
g δw dx = −
′
g δw′ dx + [g δw]ba
a a
Z b Z b Z b
g IV δw dx = − g III δwI dx + [g III δw]ba = g II δwII dx − [g II δwI ]ba + [g III δw]ba
a a a
Z
EI w′′ δw′′ dx − [EI w′′ δw′ ]ba + [(EI w′′ )′ δw]ba
Z
+ L w′ δw′ dx − [L w′ δw]ba
Z
+ kφ w′ δw′ dx − [kφ w′ δw]ba
Z Z Z
+ kw w δw dx − pz δw dx + my δw′ dx − [my δw]ba
Z
+ EI κ δw′′ dx − [EI κ δw′ ]ba + [(EI κ)′ δw]ba
Z
+ L w′0 δw′ dx − [L w′0 δw]ba = 0
Z
[EI w′′ δw′′ + (L + kφ ) w′ δw′ + kw w δw] dx
Z
+ [−pz δw + (my + L w′0 ) δw′ + EI κ δw′′ ] dx
Z
− [EI w′′ δw′′ + (L + kφ ) w′ δw′ + kw w δw] dx
Z
b
+ [pz δw − (my + L w′0 ) δw′ − EI κ δw′′ ] dx + [T (x) δw(x)]ba − [M (x) δw′ (x)]a = 0
. . . . . . − T (a) δw(a) + T (b) δw(b) + M (a) δw′ (a) − M (b) δw′ (b) = 0
Z
− [EI w′′ δw′′ + (L + kφ ) w′ δw′ + kw w δw] dx
Z φa φb
+ [pz δw − (my + L w′0 ) δw′ − EI κ δw ] dx
′′
wa wb Ma Ta Tb Mb
+Ta δwa + Tb δwb + Ma δφa + Mb δφb = 0 (2.78)
2.2.5 Element Matrices for Bending on the Basis of Cubic Test Functions
(Finite Element Method)
Starting from eq. (2.78) (weak formulation) a finite element should be derived for a beam
column with a varying height on an elastic support. As test functions we choose the shape
functions of a prismatic (uniform) beam without elastic support.
Z
− [EI w′′ δw′′ + (L + kφ ) w′ δw′ + kw w δw] dx
Z
+ [pz δw − (my + L w′0 ) δw′ − EI κ δw′′ ] dx
+ δvT p = 0 (2.87)
Z
kp + kgk = N′′T EI(x) N′′ dx (2.88)
Z
kgL =L N′T N′ dx (2.89)
Z
kgφ = kφ N′T N′ dx (2.90)
Z
kgw = NT kw (x) N dx (2.91)
Z
pp = pz (x) NT dx (2.92)
Z
pm =− my (x) N′T dx (2.93)
Z
pL = −L w0 (x)′ N′T dx (2.94)
Z
pκ = EI κ(x) N′′T dx (2.95)
n o
δvT − [kp + kgk + kgL + kgφ + kgw ] v + pp + pm + pL + pκ + p = 0
p0 = −(pp + pm + pL + pκ ) (2.98)
p = k v + p0 (2.99)
Evalutation of eqs. (2.89) to (2.91) yields the desired matrices based on cubic form functions,
for instance:
L
wb − wa Ta = (36 wa − 3 l φa − 36 wb − 3 l φb )
φa = − ; φb = φa ; 30l
l
L wb − wa wb − wa
= 36 wa + 3 l − 36 wb + 3 l
30l l l
+L
+L L
Ta = (36 wa + 3 wb − 3 wa − 36 wb + 3 wb − 3 wa )
30l
Tb L wb − wa
= (30 wa − 30 wb ) = −L · ;
30l l
156 −22 · l 54 13 · l
4 · l2 −13 · l −3 · l2
−22 · l
kw · l
kgw =
420 54 −13 · l 156 22 · l
13 · l −3 · l2 22 · l 4 · l2
For a beam of constant width and linearly varying height, we get the following for the matrices
kp + kgk :
kp+gk = kp + kgk
12
− l62 − 12 − l62
l3 l3
4 6 2
l l2 l
kp =
EIa
12 6
l3 l2
4
l
kgk = EIa ·
12 15
l3 10 α + 65 α2 + 7 3
20 α − l62 α+ 7 2
10 α + 51 α3 − 12
l3
15
10 α + 65 α2 + 7 3
20 α − l62 2α + 17 2
10 α + 12 α3
4 3 2 2 1 3 6 7 2 1 3 2 3 13 2 2 3
4α + 5α + 10 α α+ 10 α + 5α 2α
+ 10 α
+ 5α
l l2 l
12 15
+ 65 α2 + 7 3 6
2α + 17 2 1 3
l3 10 α 20 α l2 10 α + 2 α
4 9 19 2 11 3
l 4α + 10 α + 20 α
The following methods are available for a first/second order analysis of plane frames:
1. Differential Equation Methods
These methods directly solve the governing differential equations by substituting a suitable
fundamental solution. The constants have generally no mechanical interpretation and are
determined by boundary conditions. This approach is only limited to very simple structural
systems and therefore not recommended in engineering practice.
2. Iterative Hand Methods
These methods are only applicable for simple systems (e.g., cantilever beams with a vary-
ing cross section). (See e.g., Chr. Petersen, Statik und Stabilität der Baukonstruktionen,
Chapters 3.1. (Verfahren iterativer Annäherung) and 3.2. (Verfahren approximativer Verfor-
mungsaffinität)).
The destabilizing forces of a geometrically nonlinear beam theory according to a second order
analysis may be included in the local relations of a beam element with exact or approximate
1. The theoretical error of the element formulations within the scope of the underlying
mechanical model is zero for all quantities.
2. It follows that in the discretization process, every member defined by the geometry of
the structure can be assembled into the calculation model as one single element. The
global results are theoretically exact and independent of how the structure is divided
into its elements. However, this requires that not only the stiffness matrices but also the
fixed-end forces and moments are obtained in the same manner.
3. The relations between the various quantities at a point of the element remain theoreti-
cally exact, which allows checks for conditions that must also be satisfied exactly.
4. Any type of distributed loads can be modeled theoretically exact, independent of the
discretization.
5. In contrast to a merely numerical approach, all the elements of the characteristic matrices
remain valid, even if they are isolated from their context. Therefore, they may be used
any time for checks.
6. Exact solutions provide the best imaginable basis for estimating the accuracy and con-
vergence of approximate solutions.
7. Analytical solutions usually allow a direct examination of functional dependencies and
therefore the behavior of structural members and structures.
1. Members subjected to tension and compression require their own respective set of stiff-
ness matrices, which must be provided by the computer program.
2. Exact solutions may contain complex analytical functions that are not available as stan-
dard mathematical functions in common high-level programming languages.
3. Rigorous solutions sometimes do not allow the numerical transition e.g. from a beam
subjected to compressive forces to a beam without axial forces. In this case, analytically
undefined expressions may occur with unforeseeable numerical consequences (computer
programs that assemble exact matrices must avoid these cases by performing suitable
checks).
4. Analytical solutions are eventually also evaluated numerically, and are therefore sub-
jected to errors inherent in the numerical algorithms.
p1
p2
The principle of superposition holds for the second order theory only if the resulting normal
force is already known for the computation of the individual load cases.
EI w′′′′ + |N | wII = p1 + p2 + . . .
However, this approach is not applicable in practice, because the resulting normal force is
usually not known.
p
P
The principle of superposition is
V a) valid
P
p
P1
b) not valid
V
P2
p
W
c) not valid
d) load rearrangement
̸= + not possible
In a first order analysis, all internal forces and moments increase linearly if a given combination
of loads is increased proportionally. In a second order analysis, the stiffness of a beam element
is nonlinearly dependent on the normal force in the element. The stiffness of elements or
structural members subjected to compression decreases whereas the stiffness of elements or
structural members subjected to tension increases. This causes, in general, a redistribution
of internal forces and moments in the entire structure. If selected internal forces are plotted
over the load parameter λ, we can observe either a proportional or an inversely proportional
increase. In some cases, even a sign change in the service load domain is possible.
Example: Plane frame, bearing a horizontal load, changing direction of action
19011,1489 −8031,3371
−5460,6574 6131,5397
W = ±6600 kN ≈ 75 % Wkrit
M II
W
moment-curves for are scaled with −1
For poorly designed structures, it may occur that individual parts of the structure reach or
exceed their critical loads upon increasing the load combination, although the entire struc-
ture is still far from reaching its lowest critical load. These parts of the structure carry the
corresponding normal forces but they also apply their destabilizing forces on the remaining
structure. With respect to the numerical computation, this may require a higher number of
λ·P
1
1 2
2 2m
3 2m Beam:
I = 0,0001627 m4 ; A = 0,00727 m2
4
Column:
4 2m
I = 0,0000264 m4 ; A = 0,00478 m2
5 Load:
P = 100 kN
5 2m
4m
λ = 14
Displacements (oversized)
IInd Order Theory Buckling Shape λkrit = 17,3
λ = 10
λ = 10 λ = 20 λ = 20
λ = 14 λ = 14
Moment Curves
IInd Order Theory
The computation (Beam 5 and Column 10 finite elements, iteration of normal forces) results
in:
EA → ∞
P EI = 900 kNm2
kφ = 5000 kNm
P = 120 kN
2 M = 80 kNm
4,0
M
kφ kφ
1 3
4,0
4
=
ˆ
kφ
4,0 4,0
z
x
−17,48
−19,11
−20,88 −22,51
−9,22 −9,94
x x x x
19,11
17,48
9,19 9,91
z
z z z z
x
20,90 22,54
z
kφ = 1500 kNm
−20,80
−19,19
−6,63
x x
20,80
6,61
z
z z
19,21
x
λ · PV λ · pV
λ · PH
The entire load on a plane frame consisting of distributed loads and point loads is applied
gradually with a load paramater λ starting from λ = 0 and λ = 1.
The fixed-end forces of a beam elementq are nonlinearly dependent on the load parameter λ
|λ N (ε)|
and the buckling coefficient ε(λ) = l EI
, which must be determined iteratively:
A solution of this nonlinear system of equations can only be obtained iteratively. If K and
0
P are expressed as functions of the buckling coefficient, which is obtained from the normal
forces of the previous iteration step, then we can find a unique solution if
det K ̸= 0 (2.102)
If the determinant of K vanishes, then there is no unique solution to this equation system.
This leads to a method for the determination of the critical load parameter λi . The (generally
infinite number of) roots of the function det(K) of λi yield corresponding load magnitudes, at
which equilibrium is satisfied for a corresponding displacement shape.
det(K)
λkrit,i
λkrit,i+1 λ
A displacement vector Vi (λi ) is assigned to each eigenvalue λi , whose components are deter-
mined from
K(λi ) · Vi = 0, (2.103)
if one component of Vi is given. With λi , only the shape of the displacements, but not the
actual magnitudes of the displacements are determined.
Kb + K̃g (λ) · V = 0 (2.105)
Assuming that the relative distribution of the normal forces in the entire structure remains
constant if the load combination is increased proportionally over the load parameter λ, we can
extract the parameter λ from K̃g :
(λ) (λ=1)
K̃g = λ · Kg (2.106)
Kb +λ · Kg · V = 0 (2.107)
Setting
Kg = −Kg (2.108)
we get:
(Kb −λ · Kg ) · V = 0 (2.109)
The eigenvalue problem given with this equation possesses the following properties:
1. both Kb and Kg are symmetrical.
2. If Kb and Kg are banded, then they possess the same band structure.
3. Kb is always positive definite, i.e., VTi Kb Vi > 0 for Vi ̸= 0
4. Kg is generally positive semi-definite, i.e., VTi Kg Vi ≥ 0 for Vi ̸= 0
λ · P2 λ · P3
3
w3 a(1 − cos φ1 )
a[1 − cos(φ1 + φ2 )]
a φ1
φ2
kφ2 2
w2 a(1 − cos φ1 )
φ1
a
kφ1 1
M2 = kφ2 φ2 (2.113)
EI ⇒ ∞ EA ⇒ ∞
∂w2 ∂w2
δw2 = δφ1 + δφ2
∂φ1 ∂φ2
h i
kφ1 φ1 − λ P 2 a sin φ1 − λ P 3 a (sin φ1 + sin(φ1 + φ2 )) δφ1 +
h i
kφ2 φ2 − λ P 3 a sin(φ1 + φ2 ) δφ2 = 0 (2.118)
h h ii
kφ1 φ1 − λ a (P 2 + P 3 ) sin φ1 + P 3 sin(φ1 + φ2 ) δφ1 +
h i
kφ2 φ2 − λ P 3 a sin(φ1 + φ2 ) δφ2 = 0 (2.119)
h i
g1 (φ1 , φ2 , λ) = kφ1 φ1 − λ a (P 2 + P 3 ) sin φ1 + P 3 sin(φ1 + φ2 ) = 0 (2.120)
A linearisation regarding 2nd order theory where sin φ ≈ φ and cos φ ≈ 1 yields:
h i
kφ1 φ1 − λ a (P 2 + P 3 ) φ1 + P 3 (φ1 + φ2 ) = 0 (2.123)
resp.
Remark.:
Both functions h i
g1 (φ1 , φ2 , λ) = kφ1 φ1 − λ a (P 2 + P 3 ) sin φ1 + P 3 sin(φ1 + φ2 ) = 0
and
g2 (φ1 , φ2 , λ) = kφ2 φ2 − λ P 3 a sin(φ1 + φ2 ) = 0
represents for arbitrarily defined λ a surface-curve in the φ1 − φ2 − plane spanned by the
variables. The intersections with the horizontal zero-plane mark the states of equilibrium for
The points of intersection of both the contour lines mark the solution for the problem. The
solution of the linearized equation system is obtained analogously to eq. (2.125).
a = 1,5 m
P 2 = 4 kN kφ1 = 20 kNm
P 3 = 2 kN kφ2 = 10 kNm
λ · P2 λ · P3
K = Kb − λ · Kg 3
a φ1
20 0 12 3
φ2
Kb = ; Kg = ;
0 10 3 3 kφ2 2
φ1
a
20 − λ · 12 −λ · 3
K =
kφ1 1
−λ · 3 10 − λ · 3
φ1
Vλ = K(λ) V = 0
φ2
Eigenvectors:
1. λ1 = 1,4088
φ1 = 1
−(20 − 12λ1 )
φ2 =
−3λ1
1 1.Eigenform
= +0,7322 V1 =
0,7322
2. λ2 = 5,2578
φ1 = 1
−(20 − 12λ2 )
φ2 =
−3λ2
1
2.Eigenform
= −2,732 V2 =
−2,732
π 2 EI P
I
Pcrit = simply supported
l2
π 2 EI P
II
Pcrit = 2 clamped
l
2
π 2 EI left: clamped P
III
Pcrit = 2 right: simply supported
√l
2
π 2 EI left: clamped P
IV
Pcrit =
(2 · l)2 right: free end
sk = β · l
we can consistently express the critical load independent of the boundary conditions:
π 2 EI
Pcrit =
s2k
π 2 EI
Pcrit =
β 2 l2
In the case of a linear constitutive law, we find the stresses under the critical load to be:
π 2 EI
σcrit =
A β 2 l2
i2
σcrit = π 2 E
β 2 l2
βl
λ= (slenderness ratio)
i
π2 E
σcrit =
λ2
The deflection line in the first eigenmode must be a solution of this homogenous equation:
s
|λ1 · L|
w(x) = C1 + Ĉ2 x + C3 cos ε̂1 x + C4 sin ε̂1 x ε̂1 = (2.127)
EI
x x x
resp: w(x) = C1 + Ĉ2 l + C3 cos ε̂1 l + C4 sin ε̂1 l
l l l
s
x x x |λ1 · L|
= C1 + C2 + C3 cos ε1 + C4 sin ε1 ε1 = l · (2.128)
l l l EI
ε21
|λ1 · L| = EI C2 = Ĉ2 · l (2.129)
l2
1 ε1 x ε1 x
w(x)′ = +C2 − C3 sin ε1 + C4 cos ε1
l l l l l
1 ε21 ε1 x ε1 x
+ λ1 L C2 − 2 EI −C3 sin ε1 + C4 cos ε1
l l l l l l
λ1 · L
T (x) = + C2 (2.132)
l
Using:
T
C2 = + l (2.133)
λ1 · L
T x x
w(x) = C1 + x + C3 cos ε1 + C4 sin ε1 (2.134)
λ1 · L l l
The coordinates of the inflection points may be calculated in the local coordinate system of
the bar from the following condition:
ε2 x ε2 x !
w(x)′′ = −C3 2
cos ε 1 − C 4 2
sin ε1 = 0
l l l l
Hence:
ε1 xwi
C3 = −C4 tan (2.135)
l
Here, the quantities xwi are the desired local coordinates of the inflection points.
Introducing a new coordinate system, whose origin is in the neighboring inflection point (left
x = x − xwi (2.136)
x = x + xwi (2.137)
xwi ≤ 0
xw2 x
βR · lR βR · lR
lR
xw1 x
T
wc′
λ1 · L λ1 · L
T
Rem.: From eq. (2.132) we get λ1T·L = Cl2 = wc′ , being the inclination of the chord of inflection
points of the deformed bar (see chapter 2.2.5, destabilizing forces).
Using eqs. (2.137) and (2.135) we get from eq. (2.134):
T x + xwi x + xwi
w(x) = C1 + (x + xwi ) + C3 cos ε1 + C4 sin ε1
λ1 · L l l
T xw x + xwi x + xwi
= C1 + (x + xwi ) − C4 tan ε1 i cos ε1 + C4 sin ε1
λ1 · L l l l
"
T xw x xw x xw
= C1 + (x + xwi ) + C4 − tan ε1 i cos ε1 cos ε1 i − sin ε1 sin ε1 1
λ1 · L l l l l l
#
x wwi x xw
+ sin ε1 cos ε1 + cos ε1 sin ε1 i
l l l l
1
" #
T x xwi xwi x xwi
= C1 + (x + xwi ) + C4
xw
sin ε1 sin ε 1 sin ε 1 + sin ε 1 cos2
ε 1
λ1 · L cos ε1 i l l l l l
l
T 1 x
w(x) = C1 + (x + xwi ) + C4
x
sin ε1
λ1 · L cos ε1 i
w l
l
T ε1 x
w(x)′ = + + C4
x
cos ε1
λ1 · L l cos ε1 i
w l
l
ε2 x
w(x)′′ = −C4 1
xw
sin ε1
l2 cos ε1 i l
l
xw1 = 0
2. point of inflection:
ε1 xw2 π·l π
=π xw2 = =β·l Thus: β =
l ε1 ε1
Rem.: The effective length of a (structural) member is identical with the distance of neigh-
boring inflection points of the function of the deflection line of the bar in the first eigenmode.
This definition assumes straight bars of constant cross-section and constant axial compressive
forces.
Having calculated the least λ1 of the overall system, we can compute the respective compressive
forces λ1 · L of every member. The related buckling coefficient is:
s
|λ1 · L|
ε1 = l ·
EI
If we equate this compressive force with the critical force of an Euler 2 column (having the
same member length and stiffness), we develop:
v
ε2 2
! π EI π π 2 EI
u
|λ1 · L| = 21 EI = 2 2 β= β=
u
t
l β l ε1 |λ1 · L| l2
In these relations, the longitudinal force L is related to λ = 1 (from which we started the
eigenvalue analysis).
2.4.4.4 Example
kw kw · w b
b
Tb
Ma
kφ
a
kφ · φa
EI C ′ EI C ′ − D′ wb
Tb = φa +
l l l l l
EI ′ EI ′ wb
Ma = C φa + C
l l l
C ′ − D′ C′
+ kw 0
EI 3
EI 2 wb
l l =
C′ C′
φ 0
EI 2 EI + kφ a
l l
C ′ − D′ C′ C ′2
" #" #
EI + kw EI + kφ − EI 2 4 = 0
l3 l l
C ′ (C ′ − D′ ) (C ′ − D′ ) C′ 2 C
′2
EI 2 + EI kφ + EI kw + kw kφ − EI =0
l4 l3 l l4
(C ′ − D′ ) C′
" # " #
l3 l
C (C − D ) + kφ l
′ ′ ′
+ kw l 3 + kw kφ − C ′2 = 0
EI EI EI EI
kφ l kw l3
Parameter: γ= δ=
EI EI
C ′2 − C ′ D′ + γ (C ′ − D′ ) + δ C ′ + δ γ − C ′2 = 0
− C ′ D′ + γ (C ′ − D′ ) + δ C ′ + δ γ = 0
ε2 sin ε
C = ′
sin ε − ε cos ε
D ′ = ε2
ε3 cos ε
C ′ − D′ =
sin ε − ε cos ε
Buckling equation (compare. Petersen, Statik und Stabilität der Baukonstruktionen, Table
5.4, p. 341)
3.1 Introduction
1.Prandtl’s problem
SA
Cantilever in the undeformed state (side view
and cross–section)
λ·F
Applied load (acting in the symmetry plane of
the beam) and bending moment
M
λ·F
Deflection line and applied load (1st order ana-
lysis)
w
F crit = λcrit · F
Stability of Structures
Winter Term 2022/2023
72
λ · px
2.Torsional buckling of a thin–walled beam in compres-
sion
Cross–section and Rotation of
yoke support in cross-sections in
the ends of the midspan
column
F F
2h
l b
M X = F · h · sin ϑl
h ϑl
M X = GIT
l
GIT
In the special case of = F · h for the
Mx ( GIl T < F · h) l
linearised problem of (sin ϑl ≈ ϑl ), we can
Mx find a neighbouring state of deformation
in equilibrium (in the sense of a stability
M x = F h sin ϑl
problem, which essentially is an eigenvalue
problem):
ϑn.l. ϑ
GIT
F · h− ϑl = 0
l
P P P
P yP P zP
Yoke support
P yP Yoke support
P zP
P P P
z z z
Stress problem Buckling, Lateral–torsional
Torsional buckling buckling
y y y
z z z
Lateral–torsional Buckling, Lateral–torsional
buckling Torsional buckling buckling
S S S
y y y
M M M
z z z
Stress problem Lateral–torsional Lateral–torsional
buckling buckling
My My
x
y y
z
z
Stress problem Lateral buckling
3.2.1 Assumptions
1. The beam elements under investigation here are assumed to have a straight and prismatic
shape, i.e., the cross-section is constant along the central axis.
2. The individual members (beam elements) are made of a homogeneous, isotropic material.
3. In the reference configuration, we shall assure that the beams are unstressed and possess
their initial geometry. A local Cartesian coordinate frame is assigned to each individual
member, where the x-axis and the central axis coincide.
Applying external loads causes deformations and displacements of the structure, and the
external loads are in equilibrium with the internal forces and moments.
4. The equilibrium equations are derived using the principle of virtual displacements (the
Euler equations of the integral principle) according to a first order analysis, i.e., without
taking the deformations of the beam into account.
5. The shape of the cross sections in the y-z-plane does not change besides a rotation.
Axial strains, shearing strains, and stresses in the cross-sectional plane are negligible,
particularly for warping torsion of thin-walled bars.
6. We will adopt the Bernoulli hypothesis for the shearing deformations in bending, where
the effect of shearing strains on the state of stress is neglected, which implies that the
sections remain plane and orthogonal to the central axis.
7. Depending on the geometry of a section, the boundary conditions and the applied load,
we can employ the following mechanical models for the computation of the effects of
torsion:
a) Bars with a solid cross section or thick-walled sections with or without holes are
amenable to the general uniform torsion theory, provided that warping is not re-
strained.
b) Sufficiently accurate results for thin-walled tubular or multicellular sections can be
obtained with Bredt’s theory.
c) Thin-walled open sections with negligible warping restraints can be computed using
the specialized uniform torsion theory for thin-walled sections.
d) Thin -walled open-closed cross sections (with mainly open segments) are in a state
of stress that is a superposition of those described above.
v(x, y, z) = v − ϑ · z (3.2)
w(x, y, z) = w + ϑ · y (3.3)
y
P (y,z)
ϑ
ϑ·y
ϑ·z
Please note that in the above expressions, the functions u(x, y, z), v(x, y, z) and w(x, y, z)
refer to any point in the interior or on the surface of the bar under investigation, whereas the
displacements u, v, w and ϑ (without explicitly mentioning their dependence on x) are related
to the axis of the bar.
φy = −w′ (3.4)
φz = −v ′ (3.5)
ψ = −ϑ′ (3.6)
we get:
The strains in the interior of the beam may in general be expressed as a function of the
displacement fields as:
The first and the third term on the right side of (3.12) correspond to combined bending and
shearing in the x-y-plane. They vanish according to assumption 6.
Because of the hypothesis of Wagner, the remaining terms are related to uniform torsion
(which are not produced by changing axial warping stresses with respect to x). To emphasize
this fact, we introduce an index V (de St. Venant) with ω and get:
v′ φy resp. uz
u x w z
Similarly, the first and the third term on the right side of (3.13) correspond to combined
bending and shearing in the x-z-plane. They vanish according to assumption 6.
Again because of the hypothesis of Wagner, we may associate the remaining terms to uniform
torsion:
κz ≈ φ′z (3.16)
κy ≈ φ′y (3.17)
κω ≈ ψ ′ (3.18)
Here, we can rewrite the strains of the central axis in terms of the derivatives of the displace-
ments as:
ε = u′ (3.19)
κz = −v ′′ (3.20)
κy = −w′′ (3.21)
κω = −ϑ′′ (3.22)
ε(x, y, z) = ε + κz y + κy z + κω ω0 , (3.23)
δε = δu′ (3.25)
respectively
The terms in brackets on the right sides are strains of the central axis, and therefore a function
of x and dependent on the loading. The strains of the second equation may be combined to a
vector κ(x) as:
resp. κT (x) = [ε κz κy κω ]
The factors without brackets in the product expression on the right side are merely a function
of the coordinates of the section and are independent of the loading. They may be interpreted
as the unit warping corresponding to stretching, rotation about the y- and z-axis, and torsion.
Using the above conventions, we obtain the strain at an arbitrary point of the beam:
Because of the adopted hypothesis (5.) (shape of the cross section), (6.) (Bernoulli) and (7.a)
(Wagner), we can only afford to introduce a constitutive law for normal stresses (produced by
axial stretching, biaxial bending and constrained warping effects) and for shearing stresses,
which does not depend on the rate of change of normal stresses.
In any case we adopt linear elastic constitutive laws.
We assume a one-dimensional Hooke’s law for the longitudinal stresses resulting from normal
strains, bending, and warping torsion without taking the lateral strains into account (fiber
model):
σ(x, y, z) = E ω T κ (3.38)
The shearing stresses from the uniform torsion theory are computed according to the general-
ized Hooke’s law of three-dimensional elasticity theory as:
Here, again we attach an index V to both shearing stresses and shearing strains.
In General:
− δW = −δ (Wi + Wa ) = 0 (3.41)
Z Z
− δWi = [σx δεx + σy δεy + σz δεz + τxy δγxy + τyz δγyz + τzx δγzx ] dA dx (3.42)
l A
Based on the assumption that the shape of the cross sections in the y-z-plane does not change
as well on the hypotheses of Bernoulli and Wagner (see lecture notes) equation (3.42) can be
written as:
Z Z
− δWi = [σx δεx + τxy,V δγxy,V + τzx,V δγzx,V ] dA dx (3.43)
l A
The cross-sectional values (3.48) to (3.49) vanish, if in the system of principal axes the axis of
the center of shear is considered separately. Loads and internal forces have to be defined as
depicted below:
Reference Frame for thin-walled cross sections:
x
y
S (y S ,z S )
y x
α
y M (y M ,z M )
α
D (y D ,z D )
z
z z
In the system of principal axes where the center of shear is the axis of rotation (3.52) is
simplified:
Z
−δWi = {δε EAε + δκz EIz κz + δκy EIy κy + δκω EIω κω + δϑ′ GIT ϑ′ } dx (3.53)
l
With
one obtains:
Z
−δWi = {N δu′ − My δw′′ − Mz δv ′′ − Mω δϑ′′ + MxV δϑ′ } dx (3.56)
l
yp
ypM yM
y S
zM
M
Qy φy v ϑ
zp
zpM
ϑ
P
my py px
mx pz
mz z
w
φz
Qz
Z
− δW = [N δu′ + MxV δϑ′ − My δw′′ − Mz δv ′′ − Mω δϑ′′ ] dx
Z h
− px (δu − yp δv ′ − zp δw′ − wp δϑ′ ) + py (δv − δϑ zpM ) + pz (δw + δϑ ypM ) +
+ mx δϑ + my δφy + mz δφz ] dx
h ib
− N δu + Qy δv + Qz δw + M x δϑ + M y δφy + M z δφz + M ω δψ =0 (3.60)
a
Z h i
− δW = [−N ′ − px ] δu + −MxV
′
+ py zpM − pz ypM − px ωp − mx − Mω′′ δϑ
h i h i
+ −My′′ − pz − px zp − m′y δw + −Mz′′ − py − p x y p − m′z δv dx
h ib
+ N − N δu
a
h ib h ib
+ MxV + Mω′ + px ωp − M x δϑ + −Mω + M ω δϑ′
a a
h ib h ib
+ My′ + my + px zp − Qz δw + −My + M y δw′
a a
h ib h ib
+ Mz′ + mz + px yp − Qy δv + −Mz + M z δv ′ =0 (3.61)
a a
Decisive differential equations, boundary conditions (for loads) and internal forces:
1. ′
MxV + Mω′′ + mx − py zpM + pz ypM + p′x ωp = 0
my + px zp
pz
x My + My′ dx + . . .
V : Q′z + pz = 0 (3.64)
X
M : − Qz + My′ + my + px zp = 0 (3.65)
X
Based on the assumptions given in Chapter 3.2, in this chapter, we shall develop the theory of a
second order analysis of beams and beam structures subjected to three–dimensional loads. The
derivation of the governing equations is performed for a straight beam by using the principle
of virtual displacements.
Starting with this virtual work principle, we will eventually obtain the governing differential
equations in form of Euler’s equations. The mechanical interpretation of the relevant terms
in these equations is discussed subsequently. The internal forces and moments of the second
yM ypM b
yBM x
L(b)
l u(b)
a S Internal forces, displace-
M y (b) ments and loads at a
M z (b) beam element of length l.
px (x) u(a)
ϑ(b)
M y (a) M ω (b)
S φz (b) M x (b)
L(a)
M z (a) φy (b) M T y (b)
y
w(b)
v(b)
φz (a) T z (b) B
zM
φy (a) T y (a)
M
w(a) mz (x)
M x (a) y(a)
M ω (a)
ϑ(a) pz (x)
zBM B
T z (a)
p
zpM
p
mx (x) py (x) my (x)
The principle of virtual displacements also holds for a second order analysis. However, we
need to impose the virtual displacement on the deformed configuration.
According to assumption in Chapter 3.2 (the loads are defined in the initial configuration and
do not change their direction) as well as requirement 9 (the virtual work integral is formulated
Unlike in Chapter 3.3, here, we will assume that the distributed longitudinal load px acts in
the centroidal axis of the beam and that the contribution of the virtual work px ω (0, 0) δϑ′ to
the integral expression of the external loads may be neglected.
Please note that although the loads are generally defined in the undeformed configuration, they
follow the displacements of the structure from the undeformed configuration to the deformed
configuration without changing their direction (only in this configuration we can have an
equilibrium condition). Because we formulate the following relations based on the reference
frame of the initial configuration, it is also necessary to consider the resulting loads in this
reference frame.
The forces and moments applied at the ends of the beam are positive according to the internal
forces and moments along the beam (i.e., definition 1).
−δ(Wi + Wa ) = 0 (3.67)
Z
−δWi = [ L δu′ + Mx δϑ′ − My δw′′ − Mz δv ′′ − Mω δϑ′′
Z
δWa = px δu + py δ(v − ϑ zpM ) + pz δ(w + ϑ ypM )
+ mx δϑ + my δφy + mz δφz ] dx
A torsional moment causes twisting of cross sections adjacent at a distance dx about an angle
dϑ about the shear center axis. With rM equal to the distance from the shear center to an
individual fiber, we get for the inclination β of the fiber with respect to the longitudinal axis
of the beam:
dϑ
β = rM
dx
Therefore, the differential force σ dA acting in the direction of the inclined fiber possesses a
component parallel to the cross–sectional plane of magnitude
dF = rM ϑ′ σ dA
dMxσ = rM
2
ϑ′ σ dA
It follows:
!
Z
N My Mz MωM
Mxσ = ϑ ′
+ z+ y+ ωM 2
rM dA (3.70)
A Iy Iz Iω
" #
N Z 2 My Z Mz Z MωM Z
=ϑ rM dA +
′
z rM dA +
2
y rM dA +
2 2
ωM rM dA
A Iy Iz Iω
If we compute the radial distance rM from the shear center in the principal axes system, we
2
rM = (y − yM )2 + (z − zM )2 (3.71)
= y 2 − 2 y yM + yM
2
+ z 2 − 2 z zM + zM
2
= y 2 + z 2 − 2 (y yM + z zM ) + yM
2
+ zM
2
1 Z
rSy = z (y 2 + z 2 ) dA (3.73)
Iy
1 Z
rSz = y (y 2 + z 2 ) dA (3.74)
Iz
1 Z
rM ω = ωM (y 2 + z 2 ) dA (3.75)
Iω
We refer to IP S as the polar moment of inertia of the cross section. The cross–sectional distances
rSy and rSz possess the unit of length.
rSy vanishes for cross-sections that are symmetrical about the y–axis, whereas rSz vanishes for
cross-sections that are symmetrical about the z–axis. The cross-sectional property rM ω does
not have a unit. For cross-sections that are symmetrical to both the y– and the z–axis, rSy ,
rSz and rM ω are equal to zero.
Using the definitions (3.72)–(3.75), we can write the integral expressions in (3.70) as
Z Z
2
rM dA = y 2 − 2 y yM + yM
2
+ z 2 − 2 z zM + zM
2
dA = IP S + yM
2
A + zM
2
A
Z Z
z 2
rM dA = z y 2 − 2 y yM + yM
2
+ z 2 − 2 z zM + zM
2
dA
Z Z
= z (y 2 + z 2 ) dA − 2 zM z 2 dA = rSy Iy − 2 zM Iy
Z Z
y 2
rM dA = y y 2 − 2 y yM + yM
2
+ z 2 − 2 z zM + zM
2
dA
Z Z
= y (y + z ) dA − 2 yM
2 2
y 2 dA = rSz Iz − 2 yM Iz
IP S
=ϑ N ′
+ yM
2
+ zM
2
+ My (rSy − 2 zM ) + Mz (rSz − 2 yM ) + MωM rM ω
A
IP S
i2P S = (3.76)
A
i2P M = i2P S + yM
2
+ zM
2
(3.77)
by = rSy − 2 zM (3.78)
bz = rSz − 2 yM (3.79)
we obtain
h i
Mxσ = ϑ′ N i2P S + yM
2
+ zM
2
+ My (rSy − 2 zM ) + Mz (rSz − 2 yM ) + MωM rM ω
or
h i
Mxσ = ϑ′ N i2P M + My by + Mz bz + MωM rM ω (3.80)
respectively.
Mxσ = KT σ ϑ′ (3.82)
Remark: Expressions for the cross-sectional properties rSy and rSz , as well as for the distances
from the shear center yM and zM can be found in Der Stahlbau, Chr. Petersen, page 375.
symmetrical y =0 ̸= 0 ̸= 0 =0 = rSy − 2 zM =0
to z – axis
z
symmetrical ̸= 0 =0 =0 ̸= 0 =0 = rSz − 2 yM
y
to y – axis
z
symmetrical y =0 =0 =0 =0 =0 =0
to both axes
z
Pz
Ty
y
My
x
Qỹ
Mx
ỹ Mω
Mỹ
Mx̃ L
z x̃
Tz z̃
Mω̃
Qz̃
Nx̃
Mz Mz̃
p
pz
v
M
w
py
S
M y
B
S
ỹ
z
z̃
y v′ w′
w′ ϑ
w′ ϑ x
ỹ
z̃
Inclination of the gravitational axis compared to the shear axis because of ϑ ̸= const.:
w′
y S
Ñ
zM · ϑ′
zM · dϑ
zM
dϑ zM M yM
Ñ
x dϑ yM · dϑ yM · ϑ′
yM z
ϑ
v′
x̃ Mz̃
M Nx̃
w′
Mx̃
z̃ Qz̃
z
y − yM
ϑ
Qỹ
Mỹ
ϑ
S x Qz̃
v ′ + ϑ′ zM
Mỹ
M x̃ Nx̃ z − zM Mz̃
v′
Mx̃
ỹ Qỹ
y
1 −v ′ −w′ 0 0 0
L Nx̃
v + zM ϑ′ 1 0 0 0
′
Ty
−ϑ Qỹ
1 0 0 0
′
Tz w − yM ϑ′ ϑ Qz̃
=
1 −(v ′ + zM ϑ′ )
′ ′
M
x
∗ ∗ ∗ w − yM ϑ Mx̃
My ∗ 0 0 v′ 1 ϑ Mỹ
Mz ∗ 0 0 −w′ −ϑ 1 Mz̃
The above given entries to the transformation matrix result from a split of the vectors of stress
resultants (actual state) with respect to the reference frame (undeformed state).
In the actual configuration, the normal force (N ≈ L) acts in the relocated centre of gravity
S̃.
v S
y
w zM
S̃ ϑ zM
M
ϑ yM
(S) w
ϑ
zM
∼ Qỹ
ϑ M̃ v
ỹ ∼ Qỹ
z
ϑ
z̃
yM yM
1 −v ′ −w′ 0 0 0
L Nx̃
v + zM ϑ′ 1 0 0 0
′
Ty
−ϑ Qỹ
1 0 0 0
′
Tz w − yM ϑ′ ϑ Qz̃
=
1 −(v ′ + zM ϑ′ )
′ ′
M
x
∗ −w v w − yM ϑ Mx̃
My w − ϑ yM 0 0 v′ 1 ϑ Mỹ
Mz v + ϑ zM 0 0 −w′ −ϑ 1 Mz̃
In the coordinate system of the reference configuration, the normal force Ñ (of the deformed
state) can be decomposed into its component Ñ · (v ′ + zM ϑ′ ) which is in parallel to the y–axis
and in another component Ñ · (w′ − yM ϑ′ ) which is in parallel to the z–axis.
− Ñ · (v ′ + zM ϑ′ ) · (w − zM − ϑ yM ) ≈ Ñ · (v ′ + zM ϑ′ ) · zM
S
y
zM
Ñ (v ′ + zM ϑ′ )
S̃ ϑ zM
M
ϑ yM
(S) w
ϑ
zM
Ñ (w′ − zM ϑ′ )
M̃ v
z
yM yM
1 −v ′ −w′ 0 0 0
L Ñx
Ty
v ′ + zM ϑ′ 1 −ϑ 0 0 0 Q̃y
Tz w′ − yM ϑ′ ϑ 1 0 0 0 Q̃z
=
z (v ′ + zM ϑ′ ) − yM (w′ − yM ϑ′ ) −w 1 −(v ′ + zM ϑ′ )
′ ′
M
x M v w − yM ϑ M̃x
My w − ϑ yM 0 0 v′ 1 ϑ M̃y
Mz v + ϑ zM 0 0 −w′ −ϑ 1 M̃z
sA = TGA
⃗ sG (3.83)
Remark: The portion K̊T σ ϑ′ of the complete torsional moment, specified in Chapter 3.4.3, is
included in Mx̃ .
Effect of applied distributed forces of the actual configuration related to the co-
ordinate system of the reference state; influence of displacements; displacements
of the axis of elastic supports
v P zpM
ϑ
w
(P)
ypM ϑ
P̃
py zpM ϑ
pz z
1
px . . . . . px̃
py
. 1 . . . .
pỹ
pz . . 1 . . . pz̃
= (3.84)
−(w + ypM ϑ) v − zpM ϑ 1 .
m . . mx̃
x
my w − ϑ yM . . . 1 . mỹ
mz v + ϑ zM . . . . 1 mz̃
Using eqs. (3.83) and (3.84) we can rewrite eq.(3.67) (to recall: displacements are related to
the coordinate system of the reference state):
Z h i
Ñ − Q̃y v ′ − Q̃z w′ δu′
+ Ñ (zM (v ′ + zM ϑ′ ) − yM (w′ − yM ϑ′ )) − Q̃y w + Q̃z v
+ M̃ xV + M̃ xσ − M̃y (v ′ + ϑ′ zM ) + M̃z (w′ − ϑ′ yM ) dϑ′
h i
+ −Ñ (w − yM ϑ) − M̃x v ′ − M̃ y − M̃z ϑ δw′′
h i
+ −Ñ (v + zM ϑ) − M̃x w′ − M̃y ϑ − M̃ z δv ′′
h i
+ M̃ ω δϑ′′
h ib
− L δu + T y δv + T z δw + M x δϑ − M y δw′ − M ω δϑ′ =0 (3.85)
a
The above present problem is obviously non–linear cause the stress resultants Ñ , Q̃y , Q̃z , M̃x
a. s. o. are depending upon the displacements u, v, w and ϑ.
Thus, we have to choose a solution procedure in steps: in terms, which include products of
displacements and stress resultants (both existing in the deformed state), we accept internal
forces and moments resulting from a first order analysis to be sufficiently exact. We denote
these stress resultants by a small circle, for instance M̊y . (In German literature, the complete
set of circled internal forces and moments is denoted as deformed state).
Those internal forces, not being multiplied by displacements, are substituted by the respective
Ñ = EA u′ M̃y = −EIy w′′ M̃z = −EIz v ′′ M̃ω = −EIω ϑ′′ M̃xV = GIT ϑ′
Thus, we get:
Z h i
EA u′ − Q̊y v ′ − Q̊ w′ δu′
+ N̊ (zM (v ′ + zM ϑ′ ) − yM (w′ − yM ϑ′ )) − Q̊y w + Q̊z v
+ GIT ϑ + K̊T σ ϑ − M̊y (v + ϑ zM ) + M̊z (w − ϑ yM ) δϑ′
′ ′ ′ ′ ′
h i
+ −N̊ (w − yM ϑ) − M̊x v ′ + EIy w′′ − M̊z ϑ δw′′
h i
+ −N̊ (v + zM ϑ) + M̊x v ′ + M̊y ϑ + EIz v ′′ δv ′′
+ kv (v − zBM ϑ) δv + kw (w + yBM ϑ) δw
h i
+ kv (−zBM v + zBM
2
ϑ) + kw (yBM w + yBM
2
ϑ) + kϑ ϑ δϑ
h ib
− L δu + T y δv + T z δw + M x δϑ − M y δw′ − M z δv ′ − M ω δϑ =0 (3.86)
a
b b
+ [{. . .}δu]ba + [{. . .}δϑ(x)]ba + [{. . .}δϑ(x)′ ]a + [{. . .}δw(x)′ ]a
b
− [{. . .}δw(x)]ba + [{. . .}δv(x)′ ]a − [{. . .}δv(x)]ba
Having achieved some auxiliary calculations (including rearrangement of terms) we get the
final form of the principle of virtual displacements:
" #
Z
−δW = ′ ′ ′ ′
− {EA u − Q̊y v − Q̊z w } − px δu(x)
"
h i′
+ {EIω ϑ′′ }′′ − {GIT ϑ′ }′ − N̊ (i2P M ϑ′ + zM v ′ − yM w′ )
h i′ h i′ h i′
− M̊y by ϑ′ + M̊y v ′′ − M̊z bz ϑ′ − M̊z w′′ − M̊ω rM ω ϑ′
+ kv (−zBM · v + zBM
2
ϑ) + kw (yBM · w + yBM
2
ϑ) + kϑ ϑ
#
+ mz w′ − my v ′ + py ypM ϑ + pz zpM ϑ − mx + py zpM − pz ypM δϑ(x)
"
+ {−N̊ (w − yM ϑ)′ }′ + {−M̊x v ′ + EIy w′′ − M̊z ϑ}′′
#
+ kw (w + yBM ϑ) − m′y − pz δw(x)
"
+ {−N̊ (v + zM ϑ)′ }′ + {−M̊x w′ + M̊y ϑ + EIz v ′′ }′′
#
a
"
h i
+ N̊ i2P S ϑ′ + zM v ′ − yM w′ + M̊y by ϑ′ − M̊y v ′ + Q̊z v
#b
+ M̊z bz ϑ′ + M̊z w′ − Q̊ w + GIT ϑ′ − (EIω ϑ′′ )′ − M x ) δϑ
a
" #b
+ {EIω ϑ + M ω } δϑ(x)
′′ ′
a
" #b
+ {−N̊ (w − yM ϑ) − M̊x v + EIy w − M̊z ϑ + M y } δw(x)
′ ′′ ′
a
" #b
− ({−N̊ (w − yM ϑ) − M̊x v ′ + EIy w′′ − M̊z ϑ}′ − my − px (w − yM ϑ) + T z ) δw(x)
a
" #b
+ {−N̊ (v + zM ϑ) + M̊x w + EIz v + M̊y ϑ + M z } δv(x)
′ ′′ ′ ′
a
" #b
− ({−N̊ (v + zM ϑ) + M̊x w + EIz v + M̊y ϑ} − mz − px (v + zM ϑ) + T y ) δv(x)
′ ′′ ′
=0
a
The resulting axial force in point x in the reference configuration follows from the boundary
terms:
The normal force in the actual configuration we obtain applying a constitutive law:
Ñ = EA ũ′ ≈ EA u′ (3.90)
The normal force being relevant for the dimensioning we calculate from:
Remark 1: For any point x the barred stress resultants represent internal forces of the actual
configuration, being transformed to the coordinate frame of the reference configuration.
Thus we omit the bar henceforth.
Remark 2: Internal forces of the actual configuration are relevant quantities for dimensioning.
Henceforth they are denoted by a tilde, for instance M̃y .
Remark 3: In accordance with many authors we denote by Grundsatz all internal forces
calculated according to a first order analysis and mark them by a small circle (for instance
M̊y ).
3.4.6.2 Torsion
h i′
D.equ.: {EIω ϑ′′ }′′ − {GIT ϑ′ }′ − N̊ i2P M ϑ′ + zM v ′ − yM w′
′ ′
+ M̊y v ′′ − M̊y by ϑ′ − M̊z w′′ − M̊z bz ϑ′ − M̊ω rM ω ϑ′
+ kv − zBM · v + zBM
2
ϑ + kw yBM · w + yBM
2
ϑ + kϑ ϑ
Boundary conditions:
The torsional moment of the deformed configuration we get using the constitutive laws of
uniform of non–uniform torsion:
Using these relations and neglecting the last term in eq.(3.96) we can rewrite for the torsional
moment in the reference frame of the initial configuration
Mx = M̃x + N̊ i2P M ϑ′ + zM v ′ − yM w′ − Q̊y w + Q̊z v
The torsional moment of the actual configuration, which governs the design of the member is
obtained as:
M̃x = Mx − N̊ i2P M ϑ′ + zM v ′ − yM w′ + Q̊y w − Q̊z v
D.equ.:
{EIy w′′ }′′ − {N̊ (w − yM ϑ)′ }′ − {M̊z ϑ}′′ − {M̊x v ′ }′′ + kw (w + yBM ϑ)
− my ′ − pz = 0 (3.101)
BC.1:
− N̊ (w − yM ϑ) − M̊x v ′ + EIy w′′ − M̊z ϑ + M y = 0 (3.102)
Resulting bending moment My and transverse force Tz in point x, related to the reference
frame of the initial state:
= My′ + my + px (w − yM ϑ) (3.105)
Thus, we get:
= My′ + my + N̊ ′ (w − yM ϑ) + px (w − yM ϑ) + N̊ (w − yM ϑ)′ +
Bending moment M̃y and shearing force Q̃z , (which are governing structural dimensioning and
design of members):
D.equ.:
{EIz v ′′ }′′ − {N̊ (v + zM ϑ)′ }′ + {M̊y ϑ}′′ + {M̊x w′ }′′ + kv (v − zBM ϑ) − m′z − py = 0
(3.111)
BC.1:
− N̊ (v + zM ϑ) + M̊x w′ + M̊y ϑ + EIz v ′′ + M z = 0 (3.112)
BC.2:
{−N̊ (v + zM ϑ) + M̊x w′ + M̊y ϑ + EIz v ′′ }′ − mz − px (v + zM ϑ) + T y = 0 (3.113)
Resulting bending moment Mz and transverse force Ty in point x, related to the reference
frame of the initial state:
= Mz′ + mz + px (v + zM ϑ) (3.115)
Thus, we get:
= Mz′ + mz + N̊ ′ (v + zM ϑ) + px (v + zM ϑ) + N̊ (v + zM ϑ)′ +
Bending moment M̃z and shearing force Q̃y , (which are governing structural dimensioning and
design of members):
(EIz v ′′ )′′ + kv v
−(N̊ v ′ )′ +(M̊x w′ )′′ −[N̊ (zM ϑ)′ ]′ + (M̊y ϑ)′′ − kv zBM ϑ −m′z − py = 0
−(M̊x v ′ )′′ −(N̊ w′ )′ +[N̊ (yM ϑ)′ ]′ − (M̊z ϑ)′′ + kw yBM ϑ −m′y − pz = 0
−pz ypM
(EIz v ′′ )′′ + kv v
3.4.6.5.3 Governing System of Differential Equations for Cross-Sections, Symmetric to the z-Axis;
without Prestressing, including Imperfections v0 , w0 , ϑ0 ; yM = bz = rM ω = 0
(EIz v ′′ )′′ + kv v
(EIz v ′′ )′′ + kv v
We can collect equations (3.92), (3.120), (3.110), (3.100), (3.109) and (3.119) in matrix notation
and get:
0 v′ w′ 0 0 0 N̊
Ñ L
−(v ′ + zM ϑ′ ) 0 0 0 0 Q̊
Q̃y ϑ Ty
y
−(w′ − yM ϑ′ ) 0 0 0 0
Q̃z −ϑ Q̊z T
z
= +
−z v ′ + y w ′ − i2 ϑ′ 0
M̃
x M M PM w −v v ′ − by ϑ′ ′ ′
−w − bz ϑ M̊x Mx
M̃y −(w − yM ϑ) 0 0 −v ′ 0 −ϑ M̊y My
−(v + zM ϑ) 0 0 0
′
M̃z w ϑ M̊z Mz
sG = TAG
⃗ s̊A + sA (3.121)
Remark: In contrary to eq.(3.83) in the above written equation all parts of the moment
Mxσ = N̊ i2P M ϑ′ + M̊y by ϑ′ + M̊z bz ϑ′ form eq. (3.80) are explicitly written down.
Until now very few analytical (exact) solutions of the system of differential equations presented
in Chapters 3.4.6.5.1–3.4.6.5.4 are known.
One of them holds for the girder and its loads, plotted in Fig. 3.3. The solution is established
in the following
mx p
My My λ py
M
y S
x
kv
B
l kϑ
z
py z
x
l
y
Figure 3.3
First order analysis: loads and stress resultants (1st order analysis)
πx
v0 = v0m sin (3.131)
l
πx
ϑ0 = ϑ0m sin (3.132)
l
Thus the system of governing equations of Sections 3.4.6.5.1 - 3.4.6.5.4 may be rewritten:
These equations only contain derivatives of order 0, 2 and 4 of the unknown displacement
functions. The trial functions
πx πx
v = vm sin (3.135) ϑ = ϑm sin (3.136)
l l
end up in a system of two algebraic equations to determine the constant values vm and ϑm :
A vm = b (3.137)
π2 π4 π2 π2
+ GIT + kv zBM + kϑ + M̊y by
2
−kv · zBM − M̊y l2
EIω l4 l2 l2
A=
π4 π2
EIz l4
+ kv −kv · zBM − M̊y l2
2
mxm − zpM pym + M̊y [v0m − by ϑ0m ] πl2
vm
vm = b=
π2
ϑm pym + M̊y ϑ0m l2
Mz (0) =M̃z (0) + N̊ (0) (v(0) + zM ϑ(0)) − M̊x (0) w′ (0) − M̊y (0) ϑ(0) (3.140)
Analogously
Abbreviations:
π4 π2
B = EIz + kv (3.143) D = Ts + TB + M̊y by (3.146)
l4 l2
π2 π2 π2
TS = EIω + GI T (3.144) M̊ = M̊y + kv zBM (3.147)
l2 l2 l2
TB = kv zBM
2
+ kϑ (3.145)
1
" #
π2 π2
vm = mxm − zpM pym + M̊y (v0m − by ϑ0m ) 2 M̊ + pym + M̊y ϑ0m 2 D
−M̊ 2 + B D l l
(3.148)
1
" #
π2 π2
ϑm = mxm − zpM pym + M̊y (v0m − by ϑ0m ) 2 B + pym + M̊y ϑ0m 2 M̊
−M̊ 2 + B D l l
(3.149)
Internal forces and moments, relevant for design (see section 3.3 ”Bending and Torsion in a
1st Order Analysis”):
π πx
M̃xV = GIT ϑ′ = GIT ϑm cos (3.150)
l l
π3 πx
M̃xW = −EIω ϑ′′′ = EIω ϑm cos (3.151)
l3 l
M̃y ≈ My = M̊y = const. (3.152)
mxB = − ϑ kϑ (3.156)
Remark: To give proof of a correct solution, stress resultants in the coordinate system of the
initial state are calculated using equations (3.121) and (3.83).
p
My My λ py
M
S
x y
z
z
λ py
zM = −8,8 cm
by = 22,2 cm
Span: l = 6 m kv = kϑ = 0
Loads:
M̊y = 30 kNm
π·x
py = 1,5 sin kN
zpM = 0
6 m
1,5 · 64
max v = = 0,01267 m
π 4 1575
1,5 · 6
Ay = = 2,865 kN
π
Mx = 0
Matrix A:
π2
a11 = −30 = −8,228
62
π4 π2 π2
a12 = 14,95 4 + 9,91 2 − 30 (−0,222) 2 = 1,12 + 2,72 + 1,89 = 5,668
6 6 6
π4
a21 = 1576,134 = 118,5
64
π2
a22 = −30 = −8,228
62
Load vector of the system of linear equations A v0 = b
b1 = 0
b2 = 1,5
vm = 0,0148m
ϑm = 0,02044
π
v(x) = 0,01408 sin x
6
π
ϑ(x) = 0,02044 sin x
6
π π π
M̃xV = 0,02044 · 9,91 cos x = 0,1061 · x
6 6 6
Governing differential equations, boundary conditions, first order internal forces (circled quan-
tities) Rem.: First order internal forces (circled quantities, henceforth: f.o.a) are determined in
any special case and depend upon the shape of the cross-section, the supports and the loads.
Torsion:
Additional Assumptions:
Cross–section yM = zM = 0 b y = bz = 0 EIω
py = M̊z = Q̊y = 0
Torsion:
EIω ϑ′′′′ − GIT ϑ′′ − N̊ i2P S ϑ′′ = 0 BC.1: Mx = GIT ϑ′ − EIω ϑ′′′ + N̊ i2P S ϑ′
BC.2: Mω = −Eω ϑ′′
1. Neighbouring state
Buckling in the x–y–plane
z
P x
v 2. Neighbouring state
Mz
Buckling in the x–z–plane
w
My
ϑ 3. Neighbouring state:
Mx Torsional Buckling without side sway
1
2
1. P Di = · GIT + l EIωπ
i2P S
1 π 2 EIz
!
GIT l2 Iω
= 2 · · · + ; Remark: Iz < Iy
iP S l2 EIz π 2 Iz
GIT l2 Iω
c2 = · 2+ ; Radius of gyration
EIz π Iz
1 π 2 EIz 2
⇒ P Di = 2 · ·c
iP S l2
π2
2. P Kiy = · EIz
l2
π2 c2 π 2 EIz
3. P Kiy ≥ P Di : · EI z ≥ · iP S > c
l2 i2P S l2
π 2 EI
For Euler columns we get: Pki =
s2k
Pki π 2 EI
Respective buckling stress: σki = =
A A s2k
s
I
Radius of gyration i=
A
sk
Slenderness ratio: λ=
i
π 2 E i2
Hence: σki =
s2k
π2 E π2 E
σki = 2 ≤ σyield ; λ2 =
λ σki
Remark: Bars, having equal slenderness ratios (but possibly different supports) show equal
buckling stresses (at least in the elastic range).
π2 E π2 E i2P S
λ2V = = =
1 1 GIT Iω
!
π 2 EIω
σDi
GIT + + 2
A i2P S l2 A π2 E l
π2 E
σDi =
λ2V
Bars with similar comparing slenderness ratios have similar torsional buckling loads.
Additional assumptions:
Cross–section: yM , z M
by , b z
EIω
First order analyis: N̊ < 0
pz = M̊y = Q̊z = 0
py = M̊z = Q̊y = 0
mx = M̊x = 0
K̊T σ = N̊ i2P M
Torsion:
BC.1: Mz = −EIz v ′′ + N̊ (v + zM ϑ)
P P
x
position x = 0 : position x = l :
w=0 w=0
v=0 v=0
ϑ=0 ϑ=0
Trial function:
π
v = A · sin ·x
l
3
π π π π
v′ = · A · cos · x v ′′′ = − · A · cos ·x
l l l l
2 4
π π π π
v =−
′′
· A · sin · x v IV
= · A · sin ·x
l l l l
π
w = B · sin ·x
l
π π
w′ = · B · sin · x
l l
π
ϑ = C · sin ·x
l
π π
ϑ′ = · C · cos · x
l l
2
π π
ϑ′′ = − · C · sin ·x
l l
The trial functions, chosen, satisfy all boundary conditions (see eqs. (3.121)). Substituting
the approach into the differential equations yields:
4 2
π π π π
1. EIω · · C · sin x + GIT · C · sin x+
l l l l
h i π 2 π
+ N̊ · iP M 2 · C + zM · A − yM · B · · sin x = 0
l l
2
π
EIω C + GIT · C + N̊ iP M 2 C + zM A − yM B = 0
l
4 2
π π π π
2. EIy · · B · sin x + N̊ · (B − yM · C) · · sin x = 0
l l l l
2
π
EIy · · B + N̊ · (B − yM · C) = 0
l
2
π
3. EIz · · A + N̊ · (A + zM · C) = 0
l
Cross–section: P
y
yM = 0 , z M by , b z = 0
v
First order analyis: ϑ
Mz
z Mx
N̊ < 0 P
mx = M̊x = 0 ẘ
N̊ Reference
state
M̊y
See diagrams 7.2 and 7.3 in: [Petersen
1982, chapter 7: Biegedrillknicken und
Kippen]
Torsion:
h i
EIω ϑ′′′′ − GIT ϑ′′ − N̊ iP M 2 ϑ′′ + zM v ′′ + M̊y (v ′′ − by ϑ′′ ) = 0
pz v
ϑ
Mx
z Mz
x Neighbouring
state Additional assumptions:
ẘ
Q̊z
Reference
state
M̊y
Cross–section:
yM = 0 , zM by , b z = 0
First order analysis:
py = M̊z = Q̊y = 0 iP M 2 = iP S 2 + zM 2
See diagrams 7.2 and 7.3 in: [Petersen 1982, chapter 7: Biegedrillknicken und Kippen]
In Section 3.4.8, we discussed the case of a yoke–supported girder with respect to a second
order analysis.
My My
M
S
y
kv
B
kϑ
z
z
Setting the determinant of eq. (3.137) equal to zero, and solving the quadratic equation
emerging, for M̊y , we receive (using the abbreviations, introduced in Section 3.4.8):
1
!
q
l2
M̊y1,2 = − 2 kv zBM + B by ± B (−4 kv zBM by + B by 2 + 4 TS + 4 TB ) 2
2 π
if we exclude an elastic continuous support, we can rewrite this formula, using cross–sectional
stiffness quantities:
s
1 by π 2 EIz π4 π4 π2
M̊y1,2 = ± EIz 2 by 2 + 4 EIz k ϑ + 4 EIz EIω + 4EIz GI T (3.157)
2 l2 l4 l4 l2
If we assume double symmetric cross–sections without a torsional spring bed we finally get:
s
π2 l2
M̊y,crit = 2 EIω + 2 GIT EIz
l π
π2 π4 π2 π2
−M̊y 2 14,949 4 + 9,908 2 + M̊y 0,22253 2
A= 6 l l 6
4 2
π π
1576,134 4
−M̊y 2
6 6
=
118,4643 −0,2742 M̊y
M
S
y
Rem.: The difference in the absolute value of the crit. moments results exclusively out of
π2
the expression M̊y by in the coeff-matrix A.
l2
For a double-symmetric cross section with (the same) stiffnesses EIω , EIz and GIT one
would obtain:
454,8905
s
M̊y,crit = ± = ±77,796 kNm
0,07516
Additional assumptions:
Cross–section
yM = zM = 0
by = b z = 0
EIω = 0
Torsion:
GIT ϑ′′ − M̊y v ′′ = 0
BC.1: Mx = GIT ϑ′ + Q̊z v − M̊y v ′
BC.2: not applicable
My
My
z
v
ϑ Neighbouring
Qy
Mz state
Mx
Reference
ẘ state
M̊y
Boundary conditions:
position x = 0: position x = l:
ϑ(0) = 0 ϑ(l) = 0
v(0) = 0 v(l) = 0
Mz (0) = −EIz v ′′ (0) − M̊y ϑ(0) = 0 Mz (l) = −EIz v ′′ (l) − M̊y ϑ(l) = 0
In matrix notation:
M̊y −GIT
C1 0
=
2 ·
π
−M̊y C2 0
EIz
l
[Petersen 1982] Petersen, Ch.: Statik und Stabilität der Baukonstruktionen. 2. Vieweg
Verlagsgesellschaft, 1982
[Wunderlich and Kiener 2004] Wunderlich, W. ; Kiener, G.: Statik der Stabtragwerke.
Vieweg+Teubner, 2004
Stability of Structures
Winter Term 2022/2023
138