You are on page 1of 12

Continuum Mech. Thermodyn.

(1998) 10: 369–380 c Springer-Verlag 1998

A review of the problem of the shear centre(s)


Ugo A. Andreaus1 , Giuseppe C. Ruta2
1 Dipartimento di Ingegneria Strutturale e Geotecnica, Università di Roma “La Sapienza”, via Eudossiana 18, I-00184 Roma;
e-mail: andreaus@scilla.ing.uniroma1.it
2 Dipartimento di Scienze dell’Ingegneria Civile, Università Roma Tre, via C. Segre 60, I-00146 Roma; e-mail: ruta@fenice.dsic.uniroma3.it

Received November 25, 1997

The problem of defining and determinating the shear centre is present in the literature since
the first decades of this century. It soon interlaced with the definition of the twist centre and
their ways became unextricable in the following decades. A number of researchers dealt with
this subject meeting with variable issues, yet in the authors’ opinion it seems of some interest
to run through it again. In this paper a general survey of the literature is performed and the
two definitions of the shear centre (here named kinematic and energetic) are compared. It is
remarked how the energetic definition for the shear centre makes it coincide with a suitably
chosen twist centre. An example is given to show that the two definitions generally fail to
provide the same place even for simple cross sections. It is shown that the two definitions
provide the same place, using standard approximate formulæ, in the case of thin-walled mono-
connected sections with regularly varying thickness and two-connected sections with constant
thickness. As a new result, it will be shown that the two definitions of shear centre do not
provide the same place for thin-walled sections with connection higher than two even in the
trivial case of uniform thickness; an example is given.

Table of notation: E , U , V , P , D , C are sets (of Euclidean places or of translation vectors); l is the
length of the cylinder; o is an origin at will, b is the centroid of the section, x , y are places, d is the load
place in non uniform bending, r is the centre of rotation, c is the twist centre, g is the kinematic shear centre,
s is the energetic shear centre; Greek and other Latin letters denote scalar quantities or scalar-valued fields;
lower case bold-faced letters denote vector quantities or vector-valued fields; upper case bold-faced letters
denote tensor quantities or tensor-valued fields.

1 Introduction

Saint-Venant’s theory of torsion (torsion, [1, 3]) and non uniform bending (flexion inegale, [2, 3]) of right
cylinders with uniform sections does not give any definition for either the shear centre or the twist centre;
besides, in non uniform bending it is assumed that the load acts at the centroid of an end section [2, 3].
Study of Saint-Venant’s problem on torsion and non uniform bending has been continued in this century and
rigorous solutions for several new shapes of sections have been found. In the case of non uniform bending,
non symmetrical sections have been taken up; the place of application of the load at which ‘no torsion is
produced’ was fixed (e. g. [4]). It was shown that the distance of such place from the centroid of the section
may become large and of great practical importance for thin-walled sections. Timoshenko [5] claims that
Maillart in 1921 and 1922 gave this place the name ‘shear centre’ and showed how it could be found.
Osgood [6] reconsidered the question, pointing out the disagreement in the literature as to the location of
the shear centre: he put into evidence that Timoshenko [5], for example, states that the shear centre depends
370 U.A. Andreaus, G.C. Ruta

on Poisson’s ratio, whereas Trefftz [7] affirms it does not, and they obtain different places for the same cross
sections. Osgood [6] attributed the disagreement to the assumption of different boundary conditions.
Besides of that, there is some confusion in the literature also concerning the relation of the shear centre
with the twist centre. This last is usually assumed to be the centre of rotation of any section of a cylinder
loaded at the bases by opposite torques parallel to the axis. It may be shown that this definition does not
provide a unique place: changing the centre of rotation in the torsion of a Saint-Venant cylinder simply adds
an inessential rigid body motion to the solution [8, 9]. Weber [10] and Southwell [11] made plausible the
coincidence of both shear and twist centres, following a summary approach in terms of Maxwell’s reciprocal
relations. Thus, the shear centre may coincide with the twist centre, but the latter may be any place in the
section of the cylinder.
It is evident that this confusion comes from a gauge choice that has to be made; once made a suitable
gauge choice, it is possible to define the centres of shear and twist so that they coincide [12]. It seems that
the most useful choice is based on Trefftz’ paper [7], while for the twist centre the ideas of Cicala [13] have
been adopted. We will illustrate this in the following sections.

2 Statement of the problem

The formulation used here closely follows Di Carlo [14]. Let E be the Euclidean ambient space and U its
space of translations. Given a unit vector e ∈ U , U admits the orthogonal decomposition

U = span{e} ⊕ V ⇒ ∃α ∈ IR, v ∈ V : ∀u ∈ U , u = αe + v. (2.1)

In a plane P ⊥ e we fix the connected and compact domain (a superposed bar stands for the closure of the
set)
[n [
n
D = D0 \ Λi , ∂D = ∂D0 ∂Λi ; (2.2)
i =1 i =1

D0 is a simply connected region and (if present) the n cavities Λi are pairwise disjoint simply connected
regions. When P is oriented, that induces an orientation for the field of the outward unit normal to ∂D .

Fig. 1.

The Saint-Venant cylinder is the body occupying in E the reference configuration (Fig. 1)
     
C = D × [0, l ] ⇒ ∂C = D × {0} ∪ D × {l } ∪ ∂D × [0, l ] . (2.3)

In (2.3), if ζ ∈ [0, l ], then D × {0} and D × {l } are the bases of the cylinder and D × {ζ} is its ζ-cross
section. If we fix an origin o ∈ P , the places x ∈ C are given by the position vector field

x − o = y − o + ζe, y ∈ D × {0}, ζ ∈ [0, l ]. (2.4)

The displacement field u may be split (see (2.1)) into

u(y, ζ) = v(y, ζ) + w(y, ζ)e, u : C → U , v : C → V , w : C → IR. (2.5)

In the following we will indicate with the same symbol a field and its values, when there is no risk of
confusion.
A review of the problem of the shear centre(s) 371

It is assumed that C is loaded only at the bases by unspecified distributions of external forces. That is,
the lateral surface of the cylinder ∂D × [0, l ] is traction free and volume forces vanish. The problem of linear
elastostatics for C under these assumptions was solved by Saint-Venant [1, 2, 3] by means of a semi-inverse
method. The elongation at any place x ∈ C in the e-direction turns out to be a linear function of the position
x − o. The constant terms present in this function linearly depend on the resultant components of the force
parallel to e (normal force) and of the torque in V (bending moment) acting at the bases of C [3, 8, 9, 15,
16, 17, 18, 19]. It is then possible to find an explicit form for the in-plane component of the displacement v
[3, 8, 15].
The rotation ϕ : D → IR of the substantial fibers of C parallel to e about their axes is [8, 14, 15]

2 ϕ = curl v, (2.6)

where curl is the curl operator in V , providing scalar magnitudes of vectors along e. The torsional curvature
(often referred to as the twist) of the substantial fibers of C along e is given by [8, 14, 15]

∂ϕ 1 ∂v ν
ϕ0 := = curl = ϕ0 (o, 0) − [J−1 (∗q)] · (y − o). (2.7)
∂ζ 2 ∂ζ 2µ(1 + ν) b

In (2.7) µ is the shear elastic modulus; ν is Poisson’s ratio; ∗ is Hodge’s operator in V (π/2 rotation in the
positive orientation of P , [14]); q is the component of the resultant force at the bases of C contained in V
(shearing force); Jb is a central tensor of inertia of the section,
Z
Jb := [∗(y − b) ⊗ ∗(y − b)], (2.8)
D

where b is the centroid of D and the tensor product of two vectors is defined by (a⊗b)c := (a·c)b ∀ a, b, c ∈
U (· is the usual scalar product in U ). In all the integrals that we will write down the measure of integration
is understood and will be omitted. Note that, in view of the decomposition (2.1), the (scalar) component along
e of the Gibbs (cross) product of two vectors x, y ∈ V is (∗x) · y [14].
In (2.7) the field of the twist is written in its general form with respect to an origin o chosen at will. If
we refer the position of y ∈ D with respect to b, as usually done in the applications, the expression for the
twist is
ν ν
ϕ0 (y) = τ T + τ F − [J−1 (∗q)] · (y − b), τ T := ϕ0 (o, 0), τ F := − [J−1 (∗q)] · (b − o). (2.9)
2µ(1 + ν) b 2µ(1 + ν) b

In (2.9) τ T + τ F is the mean of the twist over the section with respect to the measure of area of D ; however,
τ F is present, due to Poisson effect, only when there are shearing forces at the bases of C [8], while τ T is
independent of that and characterizes torsion as follows.

3 Torsion

Torsion may be defined unambiguously as the partial solution of Saint-Venant’s problem in which the (scalar)
component along e of the stress, σ, vanishes at any place of C [8]. This implies the absence of normal
forces, bending moments and, because of global balance equations, also of shearing forces at the bases of the
cylinder.
In torsion there are no explicit solutions for either the shear stress t ∈ V or the component w of the
displacement along e. Indeed, these fields are solutions of two elliptic problems defined over the section [8,
9, 15, 16, 17, 18, 19].
According to a ‘displacement’ approach to torsion, the out-of-plane deformation of the places of D (called
warping) is solution of the Dini-Neumann problem

∆φ = 0 in D , (3.1)
(grad φ) · n = −[∗(y − b)] · n in ∂D . (3.2)
372 U.A. Andreaus, G.C. Ruta

Here ∆ and grad are the laplacian and gradient operators in V respectively; φ : D → IR is known as
Saint-Venant’s warping function; n is the outward unit normal to ∂D . The displacement in torsion is [3, 8,
9, 15, 16, 17, 18, 19]

uT = ζτ T ∗ (y − r) + τ T {[∗(r − b)] · (y − b) + φ}e =: vT + wT e, (3.3)

where the place r ∈ P is called centre of rotation and, because of the absence of resultant shearing forces
q and of (2.9), the constant parameter τ T is the twist of any substantial fiber of C parallel to e.
The first term at the right hand side of (3.3)1 represents an infinitesimal rigid rotation of any section D
about an axis parallel to e intersecting P at r. The location of the axis of rotation is immaterial in torsion
and any place r ∈ P could fit (3.3) [8, 9]. A suitable gauge choice to characterize r - the reasons will be
evident in Sect. 5 - is
Z
wT = 0, (3.4)
Z D

wT ∗ (y − b) = 0. (3.5)
D

(3.3)–(3.5) then provide a particular centre of rotation, named twist centre c:


Z
c − b = J−1b φ ∗ (y − b). (3.6)
D

(3.6) was obtained by Cicala [13] asking the displacement field in torsion to approximate a clamping condition
at one base of C . As a matter of fact, (3.3) shows that wT cannot be prescribed arbitrarily at one base of
the cylinder, if Saint-Venant’s hypotheses shall not be violated. Cicala [13] then suggested that the clamping
condition could be approximated in the sense of least squares, requiring that
Z
(wT )2 = minimum; (3.7)
D

if (3.4)–(3.5) are satisfied, so is (3.7) [13].


According to a ‘stress’ approach to torsion, the shear stress (which may be proved to be uniform in C )
is
tT = −µτ T ∗ grad ψ, ψ : D → IR, (3.8)
where ψ is known as Prandtl stress flow function, solution of the Dirichlet problem

∆ψ = −2 in D , (3.9)
ψ = const. = ψ̃i in ∂Di , (3.10)
I
(grad φ) · n = −2AΛi ∀Λi ∈ D . (3.11)
∂Λi

In (3.10) the ∂Di are all the connected components of ∂D ; in (3.11) AΛi is the area of the cavity Λi .
(3.9)–(3.11) suggest that a continuous extension of Prandtl function, ψ̂ : P → IR, may be defined as follows:

 0, y ∈ P \ D0 ,
ψ̂(y) = ψ(y), y ∈D, (3.12)

ψ̃i , y ∈ Λi .

This implies that one can imagine virtual substantial fibers of C parallel to e located in the cavities for which
the stress flow function can also be defined, so that it is possible to work on the simply connected domain
D0 .
The stress field given by (3.8)–(3.12) is statically equivalent to the resultant actions
Z Z Z
q= tT = 0, TT = [∗(y − b) · tT ] = 2µτ T ψ̂, (3.13)
D D D0
A review of the problem of the shear centre(s) 373

where TRT is the component along e of the resultant torque (twisting moment) with respect to b. Defining by
D := 2µ D0 ψ̂ the torsional stiffness constant, the average twist in torsion τ T (which is also the twist of all
substantial fibers) is, from (3.13)2 ,
TT
τT = , (3.14)
D
that is, uniquely determined in terms of the (uniform) twisting moment T T .

4 Non uniform bending

The bending of C may be either uniform (flexion egale, [3]) or non uniform (flexion inegale, [3]). In the
first case, the resultant action at each section of C is a bending moment - hence the uniform feature of
the bending. In the second case, let D × {0} be loaded by an unspecified shear stress distribution −tF (y, 0)
contained in V with resultant −q. By global balance equations, the stress distribution at any section D ×{ζ}
shall assume the form [16, 17, 18, 19]
tF (y, ζ) + σ(y, ζ)e = tF (y, 0) − ζ[∗J−1
b (∗q)] · (y − b), (4.1)
so that the resultant actions at D × {ζ} are
Z Z Z Z
tF = q, σ = 0, −σ ∗ (y − b) = −ζ ∗ q, {[∗(y − b)] · tF } = T F , (4.2)
D ×{ζ} D ×{ζ} D ×{ζ} D ×{ζ}

where the resultant components of the torque are calculated with respect to b. From (4.2)3 the bending
moment is linear with respect to ζ and we face non uniform bending, or bending in presence of a shearing
force. As for any system of vectors, there exists a load place d such that
T F = [∗(d − b)] · q (4.3)
and non uniform bending is characterized only by the shearing force q, Fig. 2. Since tF is unspecified, so are
d and the value of the twisting moment in non uniform bending T F .

Fig. 2.

The shear stress in non uniform bending has the general form [8, 15, 20]
tF = 2µ(1 + ν) grad [(∗c) · h] − µν ∗ grad (c · k) − µτ F ∗ grad ψ. (4.4)
Here h : D → V is a potential field, solution of the Dini-Neumann problem
∆h = −(y − b) in D , (4.5)
(grad h)n = 0 in ∂D ; (4.6)
k : D → V is a flow field, solution of the Dirichlet problem
∆k = −2(y − b) in D , (4.7)
k = const. = k̃i in ∂Di , (4.8)
I
(grad k)n = −2AΛi (bΛi − b) ∀ Λi ∈ D , (4.9)
∂Λi
374 U.A. Andreaus, G.C. Ruta

where bΛi is the centroid of the cavity Λi ; the kinematic parameter c represents the (linear) variation of the
bending curvature of any substantial fiber of C parallel to e [8, 15, 16]:

J−1
b (∗q)
c=− ; (4.10)
2µ(1 + ν)

ψ is again Prandtl stress flow function (see (3.8)); τ F was defined by (2.9)3 .
The last addend in (4.4) is formally similar to the stress in torsion, (3.8), and depends on the parameter
τ F introduced by (2.9)3 . As o may be any place in P , τ F is left undetermined, just like d (see (4.3) above).
Indeed, there exists a one-to-one relationship between τ F and d via (4.2)4 , (4.3) and (4.4) [8]. In non uniform
bending then we face the superposition of a ‘bending’ aspect, ruled by the vector c in (4.10), and of a ‘torsion’
aspect, ruled by τ F or by d . Since c is unambiguously defined by (4.10), one must define conventionally the
‘torsion’ aspect by making a suitable gauge choice to specify τ F or d . One such choice could be to determine
τ F so that T F vanishes [6, 17, 21]. Otherwise, one could choose a load place d allowing for ‘torsionless
bending’ (torsionsfreie Biegung, [7]) or ‘pure bending’ (reine Biegung, [7]) or ‘bending without torsion’ [8,
19]. Then non uniform bending would be obtained by superposing a torsion to it [9, 22]. The meaning of
such expressions as those quoted above in inverted commas should be precisely defined, however. Indeed, if
a definition like ‘pure bending’ is considered from a kinematic point of view, it would suggest the absence
of twist. From (2.9) when q 6= 0 the twist is a linear function of the position instead. Much of the confusion
in the literature concerning non uniform bending arises from the different definitions of the ‘torsion’ aspect.
We will discuss this in the following section.

5 Separation of ‘torsion’ and ‘bending’ in non uniform bending

As already remarked in the above section, the separation of ‘torsion’ and ‘bending’ features in non uniform
bending involves a convention. This last concerns a precise definition of the ‘torsion’ aspect, so that such
expressions as ‘torsionless bending’ are unambiguous. Two different conventions have been proposed in the
literature, which we will call kinematic and energetic and illustrate in the following.

5.1 Kinematic convention

By defining
tFc := 2µ(1 + ν) grad [(∗c) · h] − µν ∗ grad (c · k), tFτ F := − µτ F ∗ grad ψ, (5.1)
(4.4) takes the form
tF = tFc + tFτ F , (5.2)
where the subscripts c and τ F denote respectively the ‘bending’ and the ‘torsion’ components of the shear
stress in non uniform bending. Timoshenko & Goodier [19] and Sokolnikoff [16] adopted this convention
and suggested the decomposition of non uniform bending into the following simpler problems: i ) ‘torsionless
bending’ (τ F = 0), ii ) ‘pure torsion’ (c = 0; remark that the twist is now due to Poisson effect induced by q,
see (2.9)3 ).
Problem i ) is then characterized by setting the twist at the centroid (or, equivalently by (2.9), the average
of the twist with respect to the measure of area of D ) equal to zero. Indeed, due to Poisson effect, it would
be impossible to pretend to annul the twist all over D (2.9). The location of the load place g corresponding
to the stress field tFc is determined by the static equivalence condition
Z
{[∗(y − b)] · tFc } = [∗(g − b)] · q, (5.3)
D

that must hold for an arbitrary choice of q. The load place g corresponding to τ F = 0 given by (5.3) is called
(kinematic) shear centre [16, 19].
Problem ii ) is characterized by the uniform twist
A review of the problem of the shear centre(s) 375

[∗(d − g)] · q
τkF = (5.4)
D
and by the stress field (5.1)2 , formally equivalent to that of torsion (hence the name of ‘pure torsion’ attributed
to problem ii )); the subscript k in (5.4) suggests that this value for τ F comes from the kinematic convention.
Adopting this convention, it is possible to think of the resultant shearing force q at the place d as being
replaced by q at the (kinematic) shear centre g and by a component of torque parallel to e due to the
transplacement of q. This amounts to splitting T F as given by (4.2)4 –(4.3) into the sum of two addends:
Z Z
{[∗(y − b)] · tc } = [∗(g − b)] · q,
F
{[∗(y − b)] · tFτ F } = [∗(d − g)] · q. (5.5)
k
D D

Cicala [13] adopted this convention and showed that the location of the (kinematic) shear centre g may be
given by the sum of two places, one of which (called principal ) coincides with the twist centre c given by
(3.6). The other part (called secondary) depends upon Poisson’s ratio and the shape of D , so that
Z
ν −1
g − b = gI − b + gII − b, gI = c, gII − b = Jb ψ(y − b); (5.6)
1+ν D

thus the (kinematic) shear centre may be determined by means only of Prandtl stress flow function for torsion.
The kinematic convention was proposed by Timoshenko [19, 23] and followed by several other authors,
including Cicala [13], Sokolnikoff [16], Novozhilov [17], Solomon [24], Ceradini [9], Viola [21], Baldacci
[25]. Its main defect is that the Green strain energy associated with non uniform bending cannot be split into
parts associated only with the ‘bending’ feature and only with the ‘torsion’ feature.

5.2 B. Energetic convention

This convention is based on the choice of a particular load place inducing the splitting of the Green strain
energy in non uniform bending into two parts, one due to the ‘bending’ aspect and the other to the ‘torsion’
aspect. There are two possible approaches to this, one of which is based on Maxwell’s reciprocal relations
and the other on Clapeyron’s inner work theorem.
According to the first approach, the work W (F , T ) spent by the actions in non uniform bending (F ) on
the displacements in torsion (T ) is given by
Z
W (F , T ) = q · vT |y=d + l σwT . (5.7)
D

In view of (3.3)–(3.6), (4.1)–(4.3), W (F , T ) vanishes if d ≡ c, that is, if the load place coincides with the twist
centre defined by (3.4)–(3.6). Maxwell’s reciprocal relations then assure that the work spent by the torque
in torsion on a (yet undefined) ‘generalized twist’ of D in non uniform bending vanishes too. Thus, it is
possible to define the (energetic) shear centre as the load place verifying W (F , T ) = 0, so that it coincides with
Cicala’s twist centre (Southwell [11], Shanley [26], Franciosi [27], Ceradini [9], Carpinteri [28]). Weber [10]
proves the coincidence of the (energetic) shear centre (Querkraftmittelpunkt) with a twist centre (Drehpunkt)
which is not unambiguously defined. The main defect of this procedure is the ambiguity in the definition of
the ‘generalized twist’ of D in non uniform bending. Besides of that, the energetic approach followed by
Weber
R [10] and Southwell [11] is summary because they do not take into account the contribution of the term
D
σw T
into the value of W (F , T ).
The approach based on Clapeyron’s inner work theorem was first used by Trefftz [7], then followed
by several other authors, including Capurso [29], Fraeijs de Veubeke [8], Ceradini [9], Viola [21], Corradi
dell’Acqua [30]. Trefftz [7] defines the (energetic) shear centre (Schubmittelpunkt) so to uncouple the strain
energy in non uniform bending into two distinct parts, one associated with ‘pure bending’ (reine Biegung) or
‘torsionless bending’ (torsionsfreie Biegung) and the other with ‘pure torsion’ (reine Torsion). This approach
then determines a particular value for τ F , which we will denote by the subscript e, standing for ‘energetic’.
As proposed by Weber [20], Trefftz [7] imagines the shear stress in non uniform bending as obtained by the
superposition of a principal field,
376 U.A. Andreaus, G.C. Ruta

tFI := 2µ(1 + ν) grad [(∗c) · h], (5.8)


(Hauptspannungen), which is irrotational, and of a secondary field,

tFII := − µν ∗ grad (c · k) − µτ F ∗ grad ψ, (5.9)

(Zusatzspannungen), which is solenoidal. As already remarked in Sect. 4, to define uniquely the secondary
field it is necessary to fix the parameter τ F . Weber [20] suggested to adjust τ F so that the component along
e of the torque due to tFII vanishes; this approach was followed also by Schwalbe [31], who used a different
decomposition for the shear stress field. Trefftz [7] showed that this choice coincided with the request of
uncoupling the strain energy in non uniform bending.
According to Clapeyron’s inner work theorem, the strain energy E in non uniform bending is
Z h Ti Z
∂wT F t l
2E = σ +t · = [tF · tT ]; (5.10)
C ∂ζ µ µ D
∂wT
indeed, by (3.3) ∂ζ = 0, so that the coupling is due only to shear stresses. By (3.8), (4.4), we obtain [15]
Z
2E = µ ψ̂[τ F + νc · (y − b)]. (5.11)
D0

Annulling E as given by (5.11) provides the uncoupling condition


Z R
ψ̂[νc · (y − b)]
ψ̂[τ F + νc · (y − b)] = 0 ⇒ τeF = − D0
R , (5.12)
D0 D0
ψ̂

which has also a kinematic interpretation [8, 15]. Indeed, (2.9), (4.10), (5.12) state that a weighted mean of
the twist in non uniform bending vanishes:
Z
ϕ0F := τeF + νc · (y − b) ⇒ ψ̂ϕ0F = 0. (5.13)
D0

Since the weight is the extended Prandtl stress flow function (3.12), the mean in (5.13)2 shall include the
twist of the virtual fibers of C along e located in the cavities Λi . Furthermore, there is a plane containing e
subjected to vanishing twist [6]; its intersection with P represents a sort of neutral axis for the twist and is
given by
R
0F
ψ̂[νc · (y − b)]
ϕ = 0 ⇒ − D0 R + νc · (y − b) = 0. (5.14)
D0
ψ̂
As Trefftz [7] remarked, when the strain energy uncoupling condition (5.12) is verified, the static equivalence
to zero of tFII may be proved [8]. It turns out that the (energetic) shear centre s ∈ P is the place of application
of q so to be statically equivalent to the field tFI (5.8) [8]:
Z
{[∗(y − b)] · tFI } = [∗(s − b)] · q. (5.15)
D

It may be proved (Weinstein [12]) that the (energetic) shear centre s defined by (5.15) and the twist centre
c given by (3.4)–(3.6) coincide. This is another advantage of the energetic definition of the separation of
‘bending’ and ‘torsion’ effects in non uniform bending; moreover, s ≡ c is independent of Poisson’s ratio.
Using (2.9), (4.10), (3.6), (5.12), the twist at s ≡ c is
 Z  R
0F −1
ψ̂[νc · (y − b)]
ϕ |s≡c = νc · Jb φ ∗ (y − b) − D0 R . (5.16)
D D0
ψ̂

We can thus think of q applied at d as replaced by the same force applied at the (energetic) shear centre
s ≡ c and by a torque due to the transplacement of the force. That is, if d 6= s ≡ c non uniform bending is
A review of the problem of the shear centre(s) 377

the superposition of a problem, which might be called flexure, in which q is applied at s ≡ c and of a torsion
in which C is subjected to the torque
T T = [∗(d − c)] · q, (5.17)
producing an uniform twist
[∗(d − c)] · q
τT = , (5.18)
D
so that the general expression for the field of the twist is (see (2.9))
R
0 TT 2µ ψ̂[νc · (y − b)]
ϕ (y) = − D0
+ νc · (y − b). (5.19)
D D

6 Applications

In this section we will provide two applications to simple but meaningful sections of Saint-Venant cylinders
that should clarify the above discussion. The first example shall concern a thick section, as suggested by
Osgood [6]. It will put into clear evidence how, even in a trivial case, the kinematic and energetic shear
centres do not coincide. The second subsection will begin with short considerations on what happens in the
case of thin-walled sections. In particular it will be shown, as a new result, that even in the trivial case of
uniform thickness, in sections with connection higher than two (i. e., with two or more cavities) the two
definitions for the shear centre do not provide the same place. Consequently, an example is brought to picture
this result.

6.1 Thick sections

As shown in Sect. 5, the two definitions of shear centre generally fail to provide the same place in the plane
P of the section. As a matter of fact, from (5.3), (5.15) it turns out that in general g 6= s. This fact could
be of some importance when considering dynamic problems, as s splits into different contributions the strain
energy in non uniform bending while g does not, and if they do not coincide it is important to know which
could be the distance between them.

Fig. 3.

To picture this fact, let us consider a simply-connected, ‘thick’ section with the shape of a semicircle
of radius R, as suggested by Osgood [6], Fig. 3. We shall assume that Poisson’s ratio (the only material
parameter of interest, needed to find the shear centre according to the kinematic convention) equals 0.3.
Because of symmetry, whatever the definition adopted, the shear centre will lie on the axis of symmetry of
the section. It turns out that the (kinematic) shear centre g lies at a distance 0.085R from the centroid of the
section b, while the (energetic) shear centre s lies at a distance 0.124R from b. That is, the distance between
the two places amounts up to 4% of the radius of the semicircle, which is not negligigble. The dotted line in
Fig. 3 represents the neutral axis for the twist, as given by (5.14). It has to be remarked (Osgood [6]) that,
if Poisson’s ratio vanishes, the two shear centres coincide.
378 U.A. Andreaus, G.C. Ruta

6.2 Thin-walled sections

When the thickness of the wall of the section D is ‘thin’ compared with a typical length of the section, it
is possible to use approximate formulæ for the shear stress field. In the case of torsion, these are usually
attributed to Kelvin for simply connected sections and to Bredt for sections with multiple connection. In
the case of non uniform bending, the standard formula used for the applications is due to Jouravski and
is based only on a balance condition verified by an average stress tJ (the superscript meaning Jouravski’s
approximation). As far as thin-walled simply connected sections are concerned, both Kelvin and Jouravski
approximate formulæ provide (a dominant component of) shear stress parallel to the middle line Γ of the
section; Kelvin’s field is affine (with zero mean) and Jouravski’s field is uniform along the thickness of the
section. Thus, the energetic coupling between ‘torsion’ and ‘bending’ aspects in non uniform bending vanishes
[9, 21, 26, 29, 30] and the shear centres according to both kinematic and energetic convention provide the
same place, that coincides also with Cicala’s twist centre. Thus, the location of the unique shear-twist centre
c ≡ g ≡ s can be determined by means only of the static equivalence condition [22, 23, 32, 33, 34, 35, 36]
Z
{[∗(y − b)] · tJ } = [∗(c − b)] · q. (6.1)
D

The coincidence of the energetic and kinematic shear centres may be proved also for thin-walled two-connected
sections (i. e., sections with only one cavity) with uniform thickness [21].
The coincidence of twist and shear centres was obtained also by Reissner & Tsai [37] starting from
a different model (a constrained shell, not a Saint-Venant cylinder) and by means of a different approach
(influence coefficients).

Fig. 4.

Unfortunately, this coincidence does not hold for sections with connection higher than two (i. e., with
two or more cavities) even in the trivial case of uniform thickness t. We can imagine any such section as the
join of a number N of (generally curvilinear) regular strips, each indexed by the letter h. At the h-th strip
the extended Prandtl stress flow function (see (3.12)) assumes the linear form
 
j j 1 η
ψ̂h = ψ̃h + (ψ̃h − ψ̃h )
k
− , (6.2)
2 t
where ψ̃hj , ψ̃hk denote the (constant) values attained by the flow function at the boundaries of the h-th strip
(it is agreed that ψ̃hk > ψ̃hj and k > j ); − 2t ≤ η ≤ 2t is a rectilinear abscissa orthogonal to the middle line
of the strip (it is agreed that η is positively oriented from the boundary indexed by j to that indexed by k ).
Substituting (6.2) into (5.6)3 yields
Z
tν X j
N
k −1
gII − b = g − c = (ψ̃h + ψ̃h )Jb (y − b), (6.3)
2(1 + ν) Γh
h=1
A review of the problem of the shear centre(s) 379

where Γh is the middle line of the h-th strip of the section. That is, as in general the distance between g and
c given by (6.3) does not vanish, the energetic shear centre (coinciding with Cicala’s twist centre c) differs
from the kinematic shear centre g. This is illustrated in an example based on the three-connected section in
Fig. 4; the energetic shear centre s lies at a distance .08a from b, while the kinematic shear centre g lies at
a distance .4t from s.

7 Conclusions

In this paper we have made a short survey of the literature on non uniform bending and on the location of the
twist and shear centres in the cross section of a Saint-Venant cylinder. We have put into evidence that the two
conventions (here named kinematic and energetic) used in the literature to provide definitions for all these
objects generally fail to coincide. The energetic convention by Trefftz [7] splits the strain energy associated
with non uniform bending into contributions due only to the ‘torsion’ and ‘bending’ aspects respectively.
Besides, it makes the shear centre coincide with the twist centre as defined by Cicala [13].
We have brought a simple example of a simply connected (‘thick’) section in which the different locations
of the energetic and kinematic shear centres are calculated. We have shown that, even if the energetic and
kinematic shear centres coincide for thin-walled sections with connection one and two, this does not happen
anymore, even in the case of uniform thickness, for thin-walled sections with connection higher than two.
We have brought a simple example also of this fact. Table 1 synthetizes the main results obtained as far as
the two sample cross sections are concerned.
Table 1.
Section |b − o| |s − o| |g − o|
Semi-circular (Fig. 3) .424R .509R .548R
Thin-walled (Fig. 4) .45a .37a .37a + .4t

Acknowledgement. The work was partially supported by Italian MURST 40% Fund.

References

The following references include only few of the important Russian scientific publications on this research field.

1. A. J. C. Barré de Saint-Venant (1855) Mémoire sur la torsion des prismes. Mém. Savants Étrangers 16
2. A. J. C. Barré de Saint-Venant (1856) Mémoire sur la flexion des prismes. Journal Math. Liouville (2) 1
3. R. F. A. Clebsch (1883) Théorie de l’élasticité des corps solides (Traduite par MM. Barré de Saint-Venant et Flamant, avec des
Notes étendues de M. Barré de Saint-Venant). Dunod, Paris [Reprinted by Johnson Reprint Corporation, New York (1966)]
4. A. Eggenschwyler (1910) Beitrag zur Festigkeitslehre. Schweiz. Bauztg. 76 266
5. S. Timoshenko (1953) History of strength of materials: with a brief account of the history of theory of elasticity and theory of
structures. McGraw-Hill, New York
6. W. R. Osgood (1943) The center of shear again. J. Appl. Mech. 10 (2) A-62–A-64
7. E. Trefftz (1935) Über den Schubmittelpunkt in einem durch eine Einzellast gebogenen Balken. ZAMM 15 220-225
8. B. M. Fraeijs de Veubeke (1979) A course in elasticity. Springer-Verlag, New York
9. G. Ceradini (1987) Scienza delle costruzioni 3. ESA, Roma
10. C. Weber (1926) Übertragung des Drehmomentes in Balken mit doppelflanschischem Querschnitt. ZAMM 6 85–97
11. R. V. Southwell (1941) An introduction to the theory of elasticity 2nd ed., Oxford University Press
12. A. Weinstein (1947) The center of shear and the center of twist. Quarterly of Applied Mathematics 5 97–99
13. P. Cicala (1935) Il centro di taglio nei solidi cilindrici. Atti Accademia delle Scienze di Torino (serie Fisica) 70 356–371
14. A. Di Carlo (1993) Il problema di Saint-Venant. Lecture notes, Dottorato di Ricerca in Ingegneria delle Strutture, Roma
15. G. C. Ruta (1995) Il problema di Saint-Venant. Studi e Ricerche del Dipartimento di Ingegneria Strutturale e Geotecnica
dell’Università “La Sapienza” di Roma 7
16. I. S. Sokolnikoff (1946) Mathematical theory of elasticity. McGraw-Hill, New York
17. V. V. Novozhilov (1961) Theory of elasticity. Pergamon, London
18. N. I. Muskhelishvili (1975) Some basic problems of the mathematical theory of elasticity. Noordhoff, Leyden
19. S. Timoshenko, J. N. Goodier (1951) Theory of elasticity. McGraw-Hill, New York
20. C. Weber (1924) Biegung und Schub im geraden Balken. ZAMM 4 334–348
21. E. Viola (1933) Scienza delle costruzioni 3. Pitagora, Bologna
380 U.A. Andreaus, G.C. Ruta

22. C. Gavarini (1996) Lezioni di scienza delle costruzioni. Masson, Roma


23. S. Timoshenko (1947) Strength of materials. D. Van Nostrand & Co., New York
24. L. Solomon (1968) Élasticité lineaire. Masson, Paris
25. R. Baldacci (1983) Scienza delle costruzioni 1. UTET, Torino
26. F. R. Shanley (1968) Strength of materials. MIR, Moscow
27. V. Franciosi (1987) Fondamenti di scienza delle costruzioni 2. Liguori, Napoli
28. A. Carpinteri (1993) Scienza delle costruzioni 1. Pitagora, Bolgna
29. M. Capurso (1971) Lezioni di scienza delle costruzioni. Pitagora, Bologna
30. L. Corradi dell’ Acqua (1992) Meccanica delle strutture. McGraw-Hill, Milano
31. W. L. Schwalbe (1935) Über den Schubmittelpunkt in einem durch eine Einzellast gebogenen Balken. ZAMM 15 138–143
32. V. I. Feodosev (1968) Strength of materials. MIR Publishers, Moscow
33. J. Case, A. H. Chilver (1971) Strength of materials and structures. Edward Arnold, London
34. D. Capecchi (1995) Scienza delle costruzioni. CISU, Roma
35. E. P. Popov (1968) Introduction to mechanics of solids. Prentice-Hall, Englewood Cliffs, N. J.
36. V. Z. Vlasov (1961) Thin-walled elastic beams. Israel Program for Scientific Translations, Monson, Jerusalem
37. E. Reissner (1972) W. T. Tsai, On the determination of the centers of twist and of shear for cylindrical shell beams. J. Appl.
Mechanics 39 1098–1102

You might also like