<<

Elastic Beams in Three Dimensions

Lars Andersen and Søren R.K. Nielsen

ISSN 1901-7286 DCE Lecture Notes No. 23 Department of

Aalborg University Department of Civil Engineering Structural Mechanics

DCE Lecture Notes No. 23

Elastic Beams in Three Dimensions

by

Lars Andersen and Søren R.K. Nielsen

August 2008

c Aalborg University Scientific Publications at the Department of Civil Engineering Technical Reports are published for timely dissemination of research results and scientific work carried out at the Department of Civil Engineering (DCE) at Aalborg University. This medium allows publication of more detailed explanations and results than typically allowed in scientific journals.

Technical Memoranda are produced to enable the preliminary dissemination of scientific work by the personnel of the DCE where such release is deemed to be appropriate. Documents of this kind may be incomplete or temporary versions of papers—or part of continuing work. This should be kept in mind when references are given to publications of this kind.

Contract Reports are produced to report scientific work carried out under contract. Publications of this kind contain confidential matter and are reserved for the sponsors and the DCE. Therefore, Contract Reports are generally not available for public circulation.

Lecture Notes contain material produced by the lecturers at the DCE for educational purposes. This may be scientific notes, lecture books, example problems or manuals for laboratory work, or computer programs developed at the DCE.

Theses are monograms or collections of papers published to report the scientific work carried out at the DCE to obtain a degree as either PhD or Doctor of Technology. The thesis is publicly available after the defence of the degree.

Latest News is published to enable rapid communication of information about scientific work carried out at the DCE. This includes the status of research projects, developments in the labora- tories, information about collaborative work and recent research results.

Published 2008 by Aalborg University Department of Civil Engineering Sohngaardsholmsvej 57, DK-9000 Aalborg, Denmark

Printed in Denmark at Aalborg University

ISSN 1901-7286 DCE Lecture Notes No. 23 Preface

This textbook has been written for the course Statics IV on spatial elastic structures given at the 5th semester of the undergraduate programme in Civil Engineering at Aalborg Uni- versity. The book provides a theoretical basis for the understanding of the structural behaviour of beams in three-dimensional structures. In the course, the text is supplemented with labora- tory work and hands-on exercises in commercial structural finite-element programs as well as MATLAB. The course presumes basic knowledge of ordinary differential equations and struc- tural mechanics. A prior knowledge about plane frame structures is an advantage though not mandatory. The authors would like to thank Mrs. Solveig Hesselvang for typing the manuscript.

Aalborg,August2008 LarsAndersenandSørenR.K.Nielsen

— i — ii

DCE Lecture Notes No. 23 Contents

1 Beams in three dimensions 1 1.1 Introduction...... 1 1.2 Equationsofequilibriumforspatialbeams ...... 1 1.2.1 Sectionforcesandstressesinabeam ...... 3 1.2.2 Kinematicsanddeformationsofabeam ...... 5 1.2.3 Constitutiverelationsforanelasticbeam ...... 10 1.3 Differentialequationsofequilibriumforbeams ...... 12 1.3.1 GoverningequationsforaTimoshenkobeam ...... 13 1.3.2 GoverningequationsforaBernoulli-Eulerbeam ...... 14 1.4 Uncouplingofaxialandbendingdeformations...... 15 1.4.1 Determinationofthebendingcentre ...... 15 1.4.2 Determinationoftheprincipalaxes ...... 21 1.4.3 Equations of equilibrium in principal axes coordinates ...... 25 1.5 Normalstressesinbeams ...... 28 1.6 Theprincipleofvirtualforces ...... 29 1.7 Elasticbeamelements...... 33 1.7.1 AplaneTimoshenkobeamelement ...... 34 1.7.2 Athree-dimensionalTimoshenkobeamelement ...... 41 1.8 Summary ...... 44

2 Shear stresses in beams due to and 45 2.1 Introduction...... 45 2.2 Homogeneoustorsion(St.Venanttorsion) ...... 46 2.2.1 Basicassumptions ...... 47 2.2.2 Solutionofthehomogeneoustorsionproblem ...... 48 2.2.3 Homogeneous torsion of open thin-walled cross-sections...... 57 2.2.4 Homogeneous torsion of closed thin-walled cross-sections ...... 59 2.3 Shearstressesfrombending ...... 67 2.3.1 Shear stresses in open thin-walled cross-sections ...... 69 2.3.2 Determinationoftheshearcentre ...... 75 2.3.3 Shearstresses inclosedthin-walledsections ...... 82 2.4 Summary ...... 91

References 93

— iii — iv Contents

DCE Lecture Notes No. 23 CHAPTER 1 Beams in three dimensions

This chapter gives an introduction is given to elastic beams in three dimensions. Firstly, the equations of equilibrium are presented and then the classical beam theories based on Bernoulli- Euler and Timoshenko beam kinematics are derived. The focus of the chapter is the flexural de- formations of three-dimensional beams and their coupling with axial deformations. Only a short introduction is given to torsional deformations, or twist, of beams in three dimensions. A full de- scription of torsion and shear stresses is given in the next chapters. At the end of this chapter, a stiffness matrix is formulated for a three-dimensional Timosheko beam element. This element can be used for finite-element analysis of elastic spatial frame structures.

1.1 Introduction

In what follows, the theory of three-dimensional beams is outlined.

1.2 Equations of equilibrium for spatial beams

An initially straight beam is considered. When the beam is free of external loads, the beam occupies a so-called referential state. In the referential state the beam is cylindrical with the length l, i.e. the cross-sections are everywhere identical. The displacement and rotation of the beam is described in a referential (x,y,z)-coordinate system with base unit vectors {i, j, k}, the origin O placed on the left end-section, and the x-axis parallel with the cylinder and orientated into the beam, see Fig. 1–1. For the time being, the position of O and the orientation of the y- and z-axes may be chosen freely. The beam is loaded by a distributed load per unit length of the referential scale defined by the vector field q = q(x) and a distributed load vector per unit length m = m(x). A differential beam element of the length dx is then loaded by the external vector qdx and external moment vector mdx as shown in Fig. 1–1. The length of the differential beam element may change during deformations due to axial strains. However, this does not affect the indicated load vectors which have been defined per unit length of the referential state. Measured in the (x,y,z)-coordinate system, q and m have the components

qx mx q = q , m = m . (1–1)  y   y  qz mz     — 1 — 2 Chapter 1 – Beams in three dimensions

qdx mdx

F + dF

M + dM

idx −M − y F

j

x i

k dx z l

Figure 1–1 Beam in referential state.

As a consequence of the external loads, the beam is deformed into the so-called current state where the external loads are balanced by an internal section force vector F = F(x) and an internal section moment vector M = M(x). These vectors act on the cross-section with the base unit vector i of the x-axis as outward directed normal vector. With reference to Fig. 1–2, the components of F and M in the (x,y,z)-coordinate system are:

N Mx F = Q , M = M (1–2)  y   y  Qz Mz     Here, N = N(x) is the axial force, whereas the components Qy = Qy(x) and Qz = Qz(x) sig- nify the shear force componentsin the y- and z-directions. The axial component Mx = Mx(x) of the section moment vector is denoted the torsional moment. The components My = My(x) and Mz = Mz(x) in the y- and z-directions represent the bending moments. The torsional moment is not included in two-dimensional beam theory. However, in the design of three-dimensional frame structures, a good understanding of the torsional behaviour of beams is crucial. Assuming that the displacements remain small, the equation of static equilibrium can be established in the referential state. With reference to Fig. 1–1, the left end-section of the element is loaded with the section force vector −F and the section moment vector −M. At the right end-section, these vectors are changed differentially into F + dF and M + dM, respectively. Force equilibrium and moment equilibrium formulated at the point of attack of the section force vector −F at the left end-section then provides the following equations of force and moment

DCE Lecture Notes No. 23 1.2 Equations of equilibrium for spatial beams 3

Qy y

My

x

Mx N

z Mz

Qz

Figure 1–2 Components of the section force vector and the section moment vector. equilibrium of the differential beam element:

−F + F + dF + qdx = 0 ⇒ dF + q = 0 (1–3a) dx −M + M + dM + idx × (F + dF)+ mdx = 0 ⇒ dM + i × F + m = 0 (1–3b) dx From Eqs. (1–1) and (1–2) follows that Eqs. (1–3a) and (1–3b) are equivalent to the following component relations: dN dQ dQ + q =0, y + q =0, z + q =0, (1–4a) dx x dx y dx z dM dM dM x + m =0, y − Q + m =0, z + Q + m =0. (1–4b) dx x dx z y dx y z At the derivation of Eq. (1–4b), it has been utilised that i × F = i × (Ni + Qyj + Qzk)= Ni × i + Qyi × jQzi × k =0i − Qzj + Qyk. (1–5)

Hence, i×F has the components {0, −Qz,Qy}. It is noted that a non-zero normal-forcecompo- nent is achieved when the moment equilibrium equations are formulated in the deformed state. This may lead to coupled lateral-flexural instability as discussed in a later chapter.

1.2.1 Section and stresses in a beam On the cross-section with the outward directed unit vector co-directional to the x-axis, the normal σxx and the shear stresses σxy and σxz act as shown in Fig. 1–3. These stresses must be

Elastic Beams in Three Dimensions 4 Chapter 1 – Beams in three dimensions statically equivalent to the components of the force vector F and the section moment vector M as indicated by the following relations:

N = σxxdA, Qy = σxydA, Qz = σxzdA, (1–6a) ZA ZA ZA

Mx = (σxzy − σxyz)dA, My = zσxxdA, Mz = − yσxxdA. (1–6b) ZA ZA ZA

y Qy

My

x σxy Mx N dA σxx

z Mz σxz

Qz

Figure 1–3 Stresses and stress resultant on a cross-section of the beam.

y

σyy

σyx

σyz σxy dy

σzy σxx σ xz x σzx σ zz dz

dx z

Figure 1–4 Components of the stress tensor.

On sections orthogonal to the y- and z-axes, the stresses {σyy, σyx, σyz} and {σzz, σzx, σzy} act as shown in Fig. 1–4. The first index indicates the coordinate axis co-directional to the outward normal vector of the section, whereas the second index specifies the direction of action of the stress component. The stresses shown in Fig.1–4 form the components of the stress tensor

DCE Lecture Notes No. 23 1.2 Equations of equilibrium for spatial beams 5

σ in the (x,y,z)-coordinate system given as

σxx σyx σzx σ = σ σ σ . (1–7)  xy yy zy  σxz σyz σzz   Moment equilibrium of the cube shown in Fig. 1–4 requires that

σxy = σyx, σxz = σzx, σyz = σzy. (1–8)

Hence, σ is a symmetric tensor.

1.2.2 Kinematics and deformations of a beam The basic assumption in the classical beam theory is that a cross-section orthogonal to the x-axis at the coordinate x remains plane and keeps its shape during deformation. In other words, the cross-section translates and rotates as a rigid body. Especially, this means that Poisson contrac- tions in the transverse direction due to axial strains are ignored. Hence, the deformed position of the cross-section is uniquely described by a position vector w = w(x) and a rotation vector θ = θ(x) with the following components in the (x,y,z)-coordinate system:

wx θx w = w , θ = θ . (1–9)  y   y  wz θz     Further, only linear beam theory will be considered. This means that the displacement compo- nents wx, wy and wz in Eq. (1–9) all small compared to the beam length l. Further the rotation components θx,θy and θz are all small. Especially, this means that sin θ ≃ tan θ ≃ θ, (1–10) where θ represents any of the indicated rotation components measured in radians. The various displacement and rotation components have been illustrated in Fig. 1–5. The rotation component around the x-axis is known as the twist of the beam. Now, a material point on the cross-section with the coordinates (x,y,z) in the referential state achieves a displacement vector u = u(x,y,z) with the components {ux,uy,uz} in the (x,y,z)-coordinate system given as (see Fig. 1–5): ux(x,y,z)= wx(x)+ zθy(x) − yθz(x), (1–11a) uy(x,y,z)= wy(x) − zθx(x), (1–11b) uz(x,y,z)= wz(x)+ yθx(x). (1–11c) It follows that the displacement of any material point is determined if only the 6 components of w(x) and θ(x) are known at the beam coordinate x. Hence, the indicated kinematic constraint reduces the determination of the continuous displacement field u = u(x,y,z) to the determina- tion of the 6 deformation components wx = wx(x), wy = wy(x), wz = wz(x), θx = θx(x), θy = θy(x) and θz = θz(x) of a single spatial coordinate along the beam axis.

Elastic Beams in Three Dimensions 6 Chapter 1 – Beams in three dimensions

θz −θy y z

dwy dwz dx dx

wy wz

x x z y

w w x y x

θx

z x

Figure 1–5 Deformation components in beam theory.

The strains conjugated to σxx, σxy and σxz are the axial strain εxx and the angular strains γxy =2εxy and γxz =2εxz. They are related to displacement components as follows: ∂u dw dθ dθ ε = x = x + z y − y z , (1–12a) xx ∂x dx dx dx ∂u ∂u dw dθ γ = x + y = y − z x − θ (x), (1–12b) xy ∂y ∂x dx dx z ∂u ∂u dw dθ γ = x + z = y + y x + θ (x). (1–12c) xz ∂z ∂x dx dx y

From Eq. (1–12) follows that γxy = γxy(x, z) is independent of y as a consequence of the presumed plane deformation of the cross-section. Then, the σxy = σxy(x, z) must also be constant over the cross-section. Especially, σxy 6= 0 at the upper and lower edge of the cross-section as illustrated in Fig. 1–6a. However, if the cylindrical surface is free of surface shear tractions, then σyx = 0 at the edge. Hence, σxy 6= σyx in contradiction to Eq. (1–8). In reality σxy = 0 at the edges, corresponding to γxy = 0. This means that the deformed cross-section forms a right angle to the cylindrical surface as shown in Fig. 1–6b. The displacement fields Eq. (1–11) are only correct for beams with cross-sections which are circular symmetric around the x-axis. In all other cross-sections, the torsional moment Mx will

DCE Lecture Notes No. 23 1.2 Equations of equilibrium for spatial beams 7

y y σyx γxy σxy dwy σxy dwy dx dx

wy wy

x x z z

wx wx (a) (b) Figure 1–6 Shear stresses on deformed beam section: (a) Deformation of cross-section in beam theory and (b) real deformation of cross-section. induce an additional non-planar displacement in the x-axis, which generally can be written in the form ux(x,y,z) = ω(y,z)dθx/dx. This is illustrated in Fig. 1–7. Hence, the final expression for the axial displacement reads dθ u (x,y,z)= w (x)+ zθ (x) − yθ (x)+ ω(y,z) x . (1–13) x x y z dx

The expressions for uy and uz in Eq. (1–11) remain unchanged, and ω(y,z) is called the warp- ing function. Whereas y and z in Eq. (1–13) may be considered as shape functions for the deformations caused by the rotations θz(x) and θy(x), the warping function is a shape function defining the axial deformation of the cross-section from the rotation component. The definition and determination of the warping function is considered in a subsequent section.

Section A–A A B Deformed top flange Undeformed state

Deformed bottom flange A B Section B–B Figure 1–7 Warping deformations in an I-beam induced by homogeneous torsion. The cross sections A–A and B–B are shown with the top flange on the left and the bottom flange on the right.

As a consequence of the inclusion of the warping, the strain components in Eq. (1–12) are modified as follows: ∂u dw dθ dθ d2θ ε = x = x + z y − y z + ω x , (1–14a) xx ∂x dx dx dx dx2 ∂u ∂u dw ∂ω dθ γ = x + y = y − θ + − z x , (1–14b) xy ∂y ∂x dx z ∂y dx   ∂u ∂u dx ∂ω dθ γ = x + z = z + θ + + y x . (1–14c) xz ∂z ∂x dx y ∂z dx  

Elastic Beams in Three Dimensions 8 Chapter 1 – Beams in three dimensions

θz −θy y z

dwy dwz dx dx

wy wz

x x z y

wx wx Figure 1–8 Kinematics of Bernoulli-Euler beam theory.

Bernoulli-Euler beam kinematics presumes that the rotated cross-section is always orthogo- nal to the deformed beam axis. This involves the following additional kinematical constraints on the deformation of the cross-section (see Fig. 1–8):

dw dw θ = − z , θ = y . (1–15) y dx z dx

Assuming temporarily that θx ≡ 0 in bending deformations, i.e. disregarding the twist of the beam, Eqs. (1–14) and (1–15) then provide:

γxy = γxz =0. (1–16)

Equation (1–16) implies that the shear stresses are σxy = σxz = 0, and in turn that the shear forces become Qy = Qz =0, cf. Eq. (1–6). However, non-zero shear forces are indeed present in bending of Bernoulli-Euler beams. The apparent paradox is dissolved by noting that the shear forces in Bernoulli-Euler beam theory cannot be derived from the kinematic condition, but has to be determined from the static equations.

The development of the classical beam theory is associated with names like Galilei (1564–1642), Mariotte (1620–1684), Leibner (1646–1716), Jacob Bernoulli (1654–1705), Euler (1707–1783), Coulomb (1736–1806) and Navier (1785–1836), leading to the mentioned Bernoulli-Euler beam based on the indicated kinematic constraint. The inclusion of transverse shear deformation was proposed in 1859 by Bresse (1822–1883) and extended to dynamics in 1921 by Timoshenko (1878– 1972). Due to this contribution, the resulting beam theory based on the strain relations Eq. (1–12), is referred to as Timoshenko beam theory (Timoshenko 1921).

The first correct analysis of torsion in beams was given by St. Venant (1855). The underlying assumption was that dθx/dx in Eq. (1–13) was constant, so the warping in all cross-sections become identical. Then, the axial strain εxx from torsion vanishes and the distribution of the shear strains γxy and γxz are identical in all sections. Because of this, St. Venant torsion is also referred to as homogeneous torsion.

DCE Lecture Notes No. 23 1.2 Equations of equilibrium for spatial beams 9

Whenever the twist or the warping is prevented at one or more cross-sections, dθx/dx is no longer constant as a function of x. Hence, axial strains occur and, as a consequence of this, axial stresses arise and the shear strains and shear stresses are varying along the beam. These phenomena were systematically analysed by Vlasov (1961) for thin-walled beams, for which reason the resulting theory is referred to as Vlasov torsion or non-homogeneous torsion. Notice that the shear stresses from Vlasov torsion have not been included in the present formulation. These will be considered in a subsequent chapter.

Seen from an engineering point of view, the primary advantage of Vlasov torsion theory is that it explains a basic feature of beams, namely that prevention of warping leads to a much stiffer struc- tural elements than achieved in the case of homogeneous warping, i.e. a given torsional moment will induce a smaller twist. Warping of the cross-section may, for example, be counteracted by the inclusion of a thick plate orthogonal to the beam axis and welded to the flanges and the web. The prevention of torsion in this manner is particularly useful in the case of slender beams with open thin-walled cross-sections that are prone to coupled flexural–torsional . Obviously, Vlasov torsion theory must be applied for the analysis of such problems as discussed later in the book.

Next, the deformation of the cross-section may be decomposed into bending and shear com- ponents. The bending components are caused by the bending moments My and Mz and deform as a Bernoulli-Euler beam. Hence, the bending components are causing the rotations θy and θz of the cross-section. The shear components are caused by the shear forces Qy and Qz. These cause the angular shear strains γxy and γxz without rotating the cross-section. Further, the dis- placement of the beam axis in shear takes place without curvature. Hence, the curvature of the beam axis is strictly related to the bending components, see Fig. 1–9. With reference to Fig. 1–10, the radii of curvatures ry and rz are related to the rotation increments dθz and −dθy of the end-sections in the bending deformations of a differential beam element of the length dx as follows

rydθz = dx κy = −1/rz = dθy/dx ⇒ (1–17)   −rzdθy = dx   κz = 1/ry = dθz/dx κ Here, κy and κzdenote the components of the curvature vector of the x-axis. Especially, for a Bernoulli-Euler beam the curvature components become, cf. Eq. (1–15), d2w d2w κ = − z , κ = y . (1–18) y dx2 z dx2 From Eqs. (1–14) and (1–17) follows that the axial strain may be written as d2θ ε (x,y,z)= ε(x)+ zκ (x) − yκ (x)+ ω(y,z) x , (1–19) xx y z dx2 where ε(x) denotes the axial strain of the beam along the x-axis given as dw ε(x)= x . (1–20) dx

Here, ε(x), κy(x) and κy(x) define the axial strain and curvatures of the beam axis, i.e. the (x)-axis.

Elastic Beams in Three Dimensions 10 Chapter 1 – Beams in three dimensions

y

x z = y

Mz

x z

+ γ y xy

Qy

x z γxy

Figure 1–9 Decomposition of cross-section deformation into bending and shear components.

y z

dθz −dθy ry rz

wy ds ≈ dx wz ds ≈ dx x x z y

Figure 1–10 Definition of curvature.

1.2.3 Constitutive relations for an elastic beam

In what follows we shall refer to N(x), Qy(x), Qz(x), Mx(x), My(x) and Mz(x) as generalised stresses. These are stored in the column matrix

N(x) Qy(x)  Q (x)  σ(x)= z . (1–21)  Mx(x)     M (x)   y   Mz(x)      DCE Lecture Notes No. 23 1.2 Equations of equilibrium for spatial beams 11

The internal virtual work of these quantities per unit length of the beam is given as

T δω = Nδε + Qyδγxy + Qzδγxz + Mxδθx + Myδκy + Mzδκz = σ δε, (1–22) where ε(x) γxy(x)  γ (x)  ε(x)= xz . (1–23)  θx(x)     κy(x)     κz(x)      The components of ε(x) are referred to as the generalised strains. The components of σ(x) and ε(x) are said to be virtual work conjugated because these quantities define the internal virtual work per unit length of the beam. Let E and G denote the modulus and the shear modulus. Then, the normal stress σxx and the shear stresses σxy and σxz may be calculated from Eq. (1–14) as follows: dw dθ dθ d2θ σ = Eε = E x + z y − y z + ω(y,z) x , (1–24a) xx xx dx dx dx dx2   dw ∂ω dθ σ = Gγ = G y − θ + − z x , (1–24b) xy xy dx z ∂y dx     dw ∂ω dθ σ = Gγ = G 2 + θ + + y x . (1–24c) xz xz dx y ∂z dx     By integration over the cross-sectional area, it then follows that

dw dθ dθ d2θ N = E A x + S y − S 2 + S x , (1–25a) dx y dx z dx ω dx2   dw dθ Q = G A y − θ + R x , (1–25b) y y dx z y dx     dw dθ Q = G A z + θ + R x , (1–25c) z z dx y z dx     dw dw dθ M = G S z + θ − S y − θ + K x , (1–25d) x z dx y y dx z dx       dw dθ dθ dθ M = E S x + I y − I z + I x , (1–25e) y y dx yy dx yz dx ωz dx   dw dθ dθ dθ M = E −S x − I y + I z − I x , (1–25f) z z dx yz dx zz dx ωy dx   where Ay, Az, Ry, Rz, Sy, Sz, Sω, K, Iyy, Izz, Iyz = Izy, Iωy and Iωz are cross-sectional (or geometrical) constants identified as:

A = dA, Ay = αyA, Az = αzA, (1–26a) ZA Elastic Beams in Three Dimensions 12 Chapter 1 – Beams in three dimensions

∂ω ∂ω R = − z dA, R = + y dA, (1–26b) y ∂y z ∂z ZA   ZA  

Sy = zdA, Sz = ydA, Sω = ωdA, (1–26c) ZA ZA ZA 2 2 Iyy = z dA, Izz = y dA, (1–26d) ZA ZA

Iyz = yzdA, Iωy = ωydA, Iωz = ωzdA, (1–26e) ZA ZA ZA ∂ω ∂ω K = y2 + z2 + y − z dA. (1–26f) ∂z ∂y ZA   Here, A is the cross-sectional area, whereas Ay and Az signify the so-called shear areas. Beam theory presumes a constant variation of the shear stresses in bending, whereas the actual variation is at least quadratic. The constant variation results in an overestimation of the stiffness against shear deformations, which is compensated by the indicated shear reduction factors αy and αz. If the actual distribution of the shear stresses is parabolic, these factors become αy = αz = 5/6. For an I-profile, the shear area is approximately equal to the web area. For GAy → ∞ we have γxy = Qy/(GAy)=0. Bernoulli-Euler beam theory is charac- terised by γxy = 0. Hence, Timoshenko theory must converge towards Bernoulli-Euler theory for the shear areas passing towards infinity. The magnitude of the shear deformations in propor- tion to the bending deformations depends on the quantity (h/l)2, where h is the height and l is the length of the beam. This relation is illustrated in Example 1-3. Ry and Rz are section constants which depend on the warping mode shape ω(y,z) as well as the bending modes via y and z. Further, the section constants Sy and Sz are denoted the static moments around the y- and z-axes. Sω specifies a corresponding static moment of the warping shape function. Iyy and Izz signify the bending moments of inertia around the y- and z-axes, respectively. Iyz is denoted the centrifugal moment of inertia, whereas Iωy and Iωz are the corresponding centrifugal moments of the warping shape function and the bending mode shapes. K is the so-called torsion constant. This defines merely the torsional stiffness in St. Venant torsion. As mentioned above, the additional contribution to Mx from Vlasov torsion will be considered in a subsequent section.

1.3 Differential equations of equilibrium for beams

In what follows, the governing differential equations for Timoshenko and Bernoulli-Euler beams are derived. At this stage, the twist θx and the torsional moment Mx are ignored. With no further assumptions and simplifications, Eq. (1–25) reduces to

N EA −ESy −ESz 0 0 dwx/dx M ES EI −EI 0 0 dθ /dx  y   y yy yz   y  Mz = −ESz −EIyz EIzz 0 0 dθz/dx . (1–27)  Q   0 0 0 GA 0   dw /dx − θ   y   y   y z   Qz   0 0 00 GAz   dwz/dx + θy              DCE Lecture Notes No. 23 1.3 Differential equations of equilibrium for beams 13

The coefficient matrix of Eq. (1–27) is symmetric. When formulated in a similar matrix format, the corresponding matrix in Eq. (1–25) is not symmetric. This is a consequence of the ignorance of the Vlasov torsion in Mx.

1.3.1 Governing equations for a Timoshenko beam Next, Eq. (1–27) is inserted into the equilibrium equations (1–4a) and (1–4b), which results in the following system of coupled ordinary differential equations for the determination of wx, wy, wz, θy and θz:

dN/dx 0 qx dM /dx Q m  y   z   y  dMz/dx = −Qy − mz ⇒  dQy/dx   0   qy         dQz/dx   0   qz             

EA ESy −ESz 0 0 dwx/dx ES EI −EI 0 0 dθ /dx d y yy yz y  −ES −EI EI 0 0   dθ /dx  dx z yz zz z  0 0 0 GAy 0   dwy/dx − θz       0 0 00 GAz   dwz/dx + θy          000 0 0 dwx/dx qx 000 0 GA dθ /dx m  z   y   y  = 0 0 0 −GAy 0 dθz/dx − mz . (1–28)  000 0 0   dθz/dx − θz   qy         000 0 0   dwz/dx + θy   qz              Equation (1–28) specifies the differential equations for Timoshenko beam theory. These should be solved with proper boundary condition at the end-sections of the beam. Let x0 denote the abscissa of any of the two end-sections, i.e. x0 =0 or x0 = l, where l is the length of the beam. At x = x0 either kinematical or mechanical boundary conditions may be prescribed. Kinematical boundary conditions mean that values of wx, wy, wz, θy and θz are prescribed,

wx(x0) = wx,0 w (x ) = w y 0 y,0  wz(x0) = wz,0  , x0 =0, l, (1–29)  θy(x0) = θy,0  θz(x0) = θz,0   whereas mechanical boundary conditions imply the prescription of N, Qy, Qz, My and Mz,

N(x0) = N0 Q (x ) = Q y 0 y,0  Qz(x0) = Qz,0  , x0 =0, l. (1–30)  My(x0) = My,0  Mz(x0) = Mz,0    Elastic Beams in Three Dimensions 14 Chapter 1 – Beams in three dimensions

In Eq. (1–30), the left-hand sides are expressed in kinematical quantities by means of Eq. (1–27). Of the 10 possible boundary conditions at x = x0 specified by Eqs. (1–35) and (1–30), only 5 can be specified. The 5 boundary conditions at x0 =0 and x0 = l can be selected independently from Eq. (1–35) and Eq. (1–30). With given boundary conditions Eq. (1–28) can be solved uniquely for the 5 kinematic quan- tities wx, wy, wz, θy, θz, which make up the degrees of freedom of the cross-section. Although an analytical solution may be cumbersome, a numerical integration is always within reach.

1.3.2 Governing equations for a Bernoulli-Euler beam Next, similar differential equations are specified for a Bernoulli-Euler beam. At first the shear forces Qy and Qz in the equations of equilibrium for My and Mz in Eq. (1–4b) are eliminated by means of the 2nd and 3rd equations in Eq. (1–4a):

2 2 2 2 d My/dx − dQz/dx + dmy/dx =0 d My/dx + qz + dmy/dx =0 ⇒ (1–31) 2 2   2 2 d Mz/dx + dQy/dx + dmz/dx =0   d Mz/dx − qy + dmz/dx =0. Using the Bernoulli-Euler kinematical constraint  Eq. (1–15), the constitutive equations for the resulting section forces may be written as

N EA ESy −ESz dwx/dx 2 2 My = ESy EIyy −EIyz −d wz/dx . (1–32)      2 2  Mz −ESzz −EIyz EIzz d wy/dx       Then, the equations of equilibrium Eq. (1–4a) and Eq. (1–31) may be recasted as the following system of coupled ordinary differential equations d dw d2w d2w EA x − ES z − ES y + q =0, (1–33a) dx dx y dx2 z dx2 x   d2 dw d2w d2w dm ES x − EI z − EI y + q + y =0, (1–33b) dx2 y dx yy dx2 yz dx2 z dx   d2 dw d2w d2w dm −ES x + EI z + EI y + q + z =0. (1–33c) dx2 z dx yz dx2 zz dx2 y dx   The governing equations (1–33) should be solved with 5 of the same boundary conditions as indicated by Eqs. (1–35) and (1–30). The difference is that θy(x0), θz(x0), Qy(x0) and Qz(x0) are represented as, cf. Eqs. (1–4b) and (1–15),

dw (x ) dw (x ) − z 0 = θ , z 0 = θ , (1–34a) dx y,0 dx z,0

dM (x ) dM (x ) − z 0 − m (x )= Q , y 0 + m (x )= Q . (1–34b) dx z 0 y,0 dx y 0 z,0

DCE Lecture Notes No. 23 1.4 Uncoupling of axial and bending deformations 15

With this in mind, the kinematic boundary conditions for Bernoulli-Euler beams are given in the form

wx(x0) = wx,0 w (x ) = w y 0 y,0  wz(x0) = wz,0  , x0 =0, l, (1–35)  dwz(x0)/dx = θy,0  dwy(x0)/dx = θz,0   whereas the mechanical boundary conditions defined in Eq. (1–30) are still valid.

1.4 Uncoupling of axial and bending deformations

Up to now the position of the origin O and the orientation of the y- and z-axes in the cross- section have been chosen arbitrarily. As a consequence of this, the deformations from the axial force and the deformation from the bending moments My and Mz will generally be coupled. This means that the axial force N referred to the origin O will not merely induce a uniform displacement wx of the cross-section, but also non-zero displacements wy and wz of O as well as rotations θy and θz. Similarly, the bending moment My will not merely cause a displacement wy and a rotation θy of the cross-section, but also a non-zero displacement wy and a rotation θz in the orthogonal direction in addition to an axial displacement wy of the origin. The indicated mechanical couplings are the reason for the couplings in the differential equations (1–28) and (1–33). The couplings may have a significant impact on the structural behaviour and stability of an engineering structure and the position of the origin for a given beam element as well as the orientation of the coordinate axes must be implemented correctly in a computational model. In this section, two coordinate transformations will be indicated, in which the axial force re- ferred to the new origin B, called the bending centre, only induces a uniform axial displacement over the cross-section. Similarly, the bending moments My and Mz around the new rotated y- and z-axes, referred to as the principal axes, will only induce the non-zero deformation compo- nents (wz,θy) and (wy,θz), respectively. Especially, the moments will induce the displacement wx =0 of the bending centre, B.

1.4.1 Determination of the bending centre

The position of the bending centre B is given by the position vector rB with the components {0,yB,zB} in the (x,y,z)-coordinate system. In order to determine the components yB and zB, a translation of the (x,y,z)-coordinate system to a new (x′,y′,z′)-coordinate system with origin in the yet unknown bending centre is performed (see Fig. 1–11). The relations between the new and the old coordinates read

′ ′ ′ x = x , y = y + yB, z = z + zB, (1–36)

In the new coordinate system, the displacement of B (the new origin) in the x′-direction (the new beam axis) becomes (see Fig. 1–11):

′ wx = wx + zBθy − yBθz. (1–37)

Elastic Beams in Three Dimensions 16 Chapter 1 – Beams in three dimensions

′ ′ θz,Mz θz,Mz

z z′

′ ′ ′ B y θy,My zB N ′

rB

O y θy,My N yB

Figure 1–11 Translation of coordinate system.

The axial strain of fibres placed on the new beam axis becomes

′ ′ ′ ′ dw dw ε (x )= = = ε + z κ − y κ , (1–38) dx′ dx B y B z where Eq. (1–17), Eq. (1–20) and Eq. (1–37) have been used. The components of the rotation vector θ of the cross-section are identical, i.e.

′ ′ ′ θx = θx, θy = θy, θz = θz. (1–39) In turn, this means that the components of the curvature vector κ in the two coordinate systems are identical as well ′ ′ ′ dθ dθ ′ dθ dθ κ = y = y = κ , κ = z = z = κ . (1–40) y dx′ dx y z dx′ dx z Further, the components of the section force vector F in the two coordinate systems become identical, i.e.

′ ′ ′ N = N,Qy = Qy Qz = Qz. (1–41)

As a consequence of referring the axial force N = N ′ to the new origin B, the components of the section vector in the (x′,y′,z′)-coordinate system are related to the components in the (x,y,z)-coordinate as follows:

′ ′ ′ Mx = Mx,My = My − zBN,Mz = Mz + yBN. (1–42)

′ ′ Equations (1–38) and (1–40) provide the following relation for ε in terms of ε , κy and κz: ′ ′ ′ ′ ε = ε − zBκy + yBκz = ε − zBκy + yBκz. (1–43)

DCE Lecture Notes No. 23 1.4 Uncoupling of axial and bending deformations 17

′ ′ ′ ′ ′ Then the relation between {N,My,Mz} and {N ,My,Mz} and {ε,κy,κz} and {ε ,κy,κz} may be specified in the following matrix formulation:

′ σ = AT σ, (1–44a)

′ ε = Aε , (1–44b) where N ε σ = M , ε = κ , (1–45a)  y   y  Mz κz     N ′ ε′ σ′ ′ ε′ ′ = My , = κy , (1–45b)  ′   ′  Mz κy     1 −zB yB A = 0 1 0 . (1–45c)  0 0 1    The components {N,My,Mz} and {ε,κy,κz} of σ and ε may be interpretedas work conjugated generalised stresses and strains. With reference to Eq. (1–32), the constitutive relation between σ and ε is given as

σ = Cε, (1–46) where C denotes the constitutive matrix,

A Sy −Sz C = E S I −I . (1–47)  y yy yz  −Sz −Iyz Izz   Likewise, the constitutive relation in the (x′,y′,z′)-coordinate system reads

′ ′ ′ σ = C ε (1–48) where the constitutive matrix has the form

A Sy′ −Sz′ ′ C = E S ′ I ′ ′ −I ′ ′ . (1–49)  y y y y z  −Sz′ −Iy′z′ Iz′z′   Obviously, as given by Eqs. (1–47) and (1–49), the cross-sectional area A is invariant to a rotation of the cross-section about the x-axis and a translation in the y- and z-directions. From Eqs. (1–44a), (1–44b) and (1–46) follows that

′ ′ σ = AT Cε = AT CAε ⇒

1 00 A Sy −Sz 1 −zB yB ′ C = AT CA = E −z 1 0 S I −I 0 1 0 ⇒  B   y yy yz    yB 0 1 −Sz −Iyz Izz 0 0 1       Elastic Beams in Three Dimensions 18 Chapter 1 – Beams in three dimensions

A Sy −zBA −(Sz −yBA) ′ 2 C = E Sy −zBA Iyy −2zBSy +zBA −Iyz + yBSy +zB(Sz −yBA) .  2  −(Sz −yBA) −Iyz + yBSy +zB(Sz −yBA) Izz −2yBSz +yBA  (1–50)

The idea is now to use the translational coordinate transformation to uncouple the axial defor- mations from the bending deformations. This requires that Sy′ = Sz′ = 0. Upon comparison of Eq. (1–49) and Eq. (1–50), this provides the following relations for the deformation of the coordinates of the bending centre:

S S y = z , z = y . (1–51) B A B A

With yB and zB given by Eq. (1–51), the bending moments of inertia, Iy′y′ and Iz′z′ , and the centrifugal moment of inertia, Iy′z′ , in the new coordinate system can be expressed in terms of the corresponding quantities in the old coordinate system as follows:

′ ′ 2 2 Iy y = Iyy − 2zB(AzB)+ zBA = Iyy − zBA, (1–52a)

′ ′ 2 2 Iz z = Izz − 2yB(AyB )+ yBA = Izz − yBA, (1–52b)

Iy′z′ = Iyz − yB(AzB) − zB(AyB)+ yBzBA = Iyz − yBzBA. (1–52c) The final results in Eq. (1–52) are known as König’s theorem.

O z x yB

y

Figure 1–12 Single-symmetric cross-section.

If the cross-section is symmetric around a single line, and the y-axis is placed so that it coincides with this line of symmetry, then the static moment Sy vanishes, i.e.

Sy = zdA =0 (1–53) ZA As a result of this, the bending centre B will always be located on the line of symmetry in a single-symmetric cross-section, see Fig. 1–12. Obviously, if the cross-section is double symmet- ric, then the position of B is found at the intersection of the two lines of symmetry.

DCE Lecture Notes No. 23 1.4 Uncoupling of axial and bending deformations 19

Example 1.1 Determination of bending and centrifugal moments of inertia of non-symmetric thin-walled cross-section The position of the bending centre of the cross-section shown in Fig. A is determined along with the bending moments of inertia Iy′y′ and Iz′z′ and the centrifugal moment of inertia Iy′z′ .

2a x O z

2t

2a

t t

a y Figure A Thin-walled cross-section.

The (x,y,z)-coordinate system is placed as shown in Fig. A. Then, the following cross-sectional con- stants are calculated:

A = 2a 2t + 2a t + a t = 7at, (a) · · · t a 1 Sy = 2a 2t a + 2a t + a t = (2t + 9a)+ ta, (b) · · · · 2 · · 2 2 t 1 Sz = 2a 2t t + 2a t (2t + a)+ a t 2t + 2a + = (21t + 8a)ta, (c) · · · · · · 2 2   1 3 1 3 1 3 1 2 2 Iyy = 2t (2a) + 2a t + t a = 2t + 17a ta, (d) 3 · · 3 · · 3 · · 3  2 1 3 1 3 2 1 3 t Izz = 2a 2a (2t) + (2a) t + 2a t (2t + a) + a t + a t 2t + 2a + 3 · · · 12 · · · 12 · · · · 2   1 = 59t2 + 54ta + 20a2 ta, (e) 3

 1 t a 1 2 2 Iyz = 2a 2t t a + 2a t (2t + a) + a t 2t + 2a + = 8t + 25ta + 4a ta. (f) · · · · · · 2 · · 2 · 2 4   Here, use has been made of König’s theorem at the calculation of contributions to Iyy, Izz and Iyz from the three rectangles forming the cross-section. The coordinates of the bending centre follow from Eq. (1–51) and Eqs. (a) to (c). Thus, S 3 4 S 1 9 y = z = t + a, z = y = t + a. (g) B A 2 7 B A 7 14 (continued)

Elastic Beams in Three Dimensions 20 Chapter 1 – Beams in three dimensions

3 4 2 t + 7 a

1 9 7 t + 14 a

B z′ x′

y′ Figure B Position of bending centre in the thin-walled cross-section.

′ ′ ′ Subsequently, the moments of inertia around the axes of the (x ,y ,z )-coordinate system follow from Eq. (1–52), Eq. (1–53) and Eqs. (d) to (f):

2 1 2 2 1 9 1 2 2 I ′ ′ = (2t + 17a )ta t + a 7ta = (44t 108ta + 233a )ta, (h) y y 3 − 7 14 · 84 −   2 1 2 2 3 4 1 2 2 I ′ ′ = (59t + 54ta + 20a )ta t + a 7ta = (329t + 504ta + 368a )ta, (i) z z 3 − 2 7 · 84   1 2 2 1 9 3 4 1 2 2 I ′ ′ = (8t + 25ta + 4a )ta t + a) t + a 7ta = (7t 15ta 22a )ta. (j) y z 4 − 7 14 2 7 · 14 − −    Now, for a thin-walled cross-section the thickness of the flanges and the web is much smaller than the widths of the flanges and the height of the web. In the present case this means that t a. With this in mind, Eqs. (g) to (j) reduce to ≪ 4 9 yB a, zB a (k) ≃ 7 ≃ 14 and

233 3 92 3 11 3 I ′ ′ ta ,I ′ ′ ta ,I ′ ′ ta . (l) y y ≃ 84 z z ≃ 21 y z ≃− 7

It is noted that the error on Iz′z′ estimated by Eq. (l) increases rapidly with increasing values of t/a. Thus, for t/a = 0.1, the error is about 13%. The errors related to the estimated values of Iy′y′ and Iy′z′ are somewhat smaller, i.e. about 5% and 7%, respectively. 

From now on, the origin of the (x,y,z)-coordinate system is placed at the bending centre. Then, the constitutive matrix given by Eq. (1–47) takes the form A 0 0 C = E 0 I −I . (1–54)  yy yz  0 −Iyz Izz   DCE Lecture Notes No. 23 1.4 Uncoupling of axial and bending deformations 21

As a result of this, an axial force N no longer induces deformations in the y- and z-directions, and the bending moments My and Mz do not induce axial displacements. However, the bending moment My will still induce displacements in the y-direction in addition to the expected dis- placements in the z-direction. Similarly, the bending moment Mz induces displacements in both the y- and z-directions.

1.4.2 Determination of the principal axes

In order to uncouple the bending deformations, so that My will only induce deformations in the ′ ′ ′ z-direction, and Mz only deformations in the y-direction, a new (x ,y ,z )-coordinate system is introduced with origin in B and rotated the angle ϕ around the x-axis as shown in Fig. 1–13.

κz,Mz ′ ′ κz,Mz

′ ′ κy,My z′ z y′

ϕ y κy,My B N = N ′

Figure 1–13 Rotation of coordinate system.

′ ′ ′ Let {N,My,Mz} and {N ,My,Mz} denote the components of the generalised stresses in the (x,y,z)-coordinate system and the (x′,y′,z′)-coordinate system, respectively. The two sets of generalised stresses are related as

′ σ = Bσ , (1–55a) where

N ′ N 10 0 σ′ ′ σ = My , = My , B = 0 cos ϕ − sin ϕ . (1–55b)  ′      Mz Mz 0 sin ϕ cos ϕ       Likewise, the components of the generalised strains in the two coordinate systems are denoted ′ ′ ′ as {ε ,κy,κz} and {ε,κy,κz}, respectively. These are related as

′ ε = Bε , (1–56a)

Elastic Beams in Three Dimensions 22 Chapter 1 – Beams in three dimensions where ε′ ε 10 0 ε′ ′ ε = κy , = κy , B = 0 cos ϕ − sin ϕ . (1–56b)  ′      κz κz 0 sin ϕ cos ϕ       The constitutive relation in the (x′,y′,z′)-coordinate system reads

A 0 0 ′ ′ ′ ′ σ = C ε , C = E 0 I ′ ′ −I ′ ′ . (1–57)  y y y z  0 −Iy′z′ Iz′z′   The corresponding constitutive relation in the (x,y,z)-coordinate system is given by Eq. (1–46) with C given by Eq. (1–54). Use of Eqs. (1–55a) and (1–56a) in Eq. (1–46) provides

′ ′ ′ ′ Bσ = Cε = CBε ⇒ σ = BT CBε , (1–58) where it has been utilised that B−1 = BT . Comparison of Eq. (1–57) and Eq. (1–58) leads the following relation between the constitutive matrices

′ C = BT CB 1 0 0 A 0 0 10 0 = E 0 cos ϕ sin ϕ 0 I −I 0 cos ϕ − sin ϕ    yy yz    0 − sin ϕ cos ϕ 0 −Iyz Izz 0 sin ϕ cos ϕ  A 0 0      ′ ′ = E 0 C22 C23 . (1–59a)  ′ ′  0 C32 C33   where

′ 2 2 C22 = cos ϕIyy − 2cos ϕ sin ϕIyz + sin ϕIzz, (1–59b)

′ ′ 2 2 C23 = C32 = − sin ϕ cos ϕ(Iyy − Izz) − (cos ϕ − sin ϕ)Iyz, (1–59c) ′ 2 2 C33 = cos ϕIzz +2cos ϕ sin ϕIyz + sin ϕIzz. (1–59d) From Eq. (1–57) and Eq. (1–59) follows that 1 1 I ′ ′ = (I + I )+ (I − I ) cos(2ϕ) − I sin(2ϕ), (1–60a) y y 2 yy zz 2 yy zz yz 1 I ′ ′ = − sin(2ϕ)(I − I ) − cos(2ϕ)I , (1–60b) y z 2 yy zz yz 1 1 I ′ ′ = (I + I ) − (I − I ) cos(2ϕ)+ I sin(2ϕ), (1–60c) z z 2 yy zz 2 yy zz yz where use has been made of the relations sin(2ϕ) = 2 sin ϕ cos ϕ, cos(2ϕ) = cos2 ϕ − sin2 ϕ, 1 1 cos2 ϕ = (1+cos2ϕ), sin2 ϕ = (1 − cos2 ϕ). 2 2

DCE Lecture Notes No. 23 1.4 Uncoupling of axial and bending deformations 23

′ ′ ′ Uncoupling of bending deformations in the (x ,y ,z )-coordinate system requires that Iy′z′ =0. This provides the following relation for the determination of the rotation angle ϕ:

2Iyz tan(2ϕ)= − for Iyy 6= Izz, 1 Izz Iyy − sin(2ϕ)(I − I ) − cos(2ϕ)I =0 ⇒ (1–61) 2 yy zz yz   cos(2ϕ)=0 for Iyy = Izz.

Note that cos(2ϕ)=0 implies that sin(2ϕ)= ±1. The sign of sin(2ϕ) is chosen as follows:

1 sin(2ϕ)= 1 ⇒ ϕ = 4 π for Iyz < 0 . (1–62) 3 for  sin(2ϕ)= −1 ⇒ ϕ = 4 π Iyz > 0 

Then, Eq. (1–60) provides the following solutions forIy′y′ and Iz′z′ : 1 1 I ′ ′ = (I + I )+ | I |, I ′ ′ = (I + I )− | I | . (1–63) y y 2 yy zz yz z z 2 yy zz yz

For Iyy 6= Izz the solution for tan(2ϕ) is fulfilled for the following two alternative solutions for sin(2ϕ) and cos(2ϕ):

2I I − I sin(2ϕ)= − yz , cos(2ϕ)= yy zz , (1–64a) J J 2I I − I sin(2ϕ)= yz , cos(2ϕ)= − yy zz , (1–64b) J J where

2 2 J = (Iyy − Izz) +4Iyz. (1–65) q The sign definition in Eq. (1–64a) is chosen. This implies that

2ϕ ∈ [0, π] for Iyz < 0 and 2ϕ ∈ [π, 2π] for Iyz > 0. (1–66)

Insertion of the solution for sin(2ϕ) and cos(2ϕ) into Eq. (1–60) provides the following results for Iy′y′ and Iz′z′ :

2 2 1 1 (Iyy − Izz) +4Iyz I ′ ′ = (I + I )+ , (1–67a) y y 2 yy zz 2 J

2 2 1 1 (Iyy − Izz) +4Iyz I ′ ′ = (I + I ) − , (1–67b) z z 2 yy zz 2 J or, by insertion of J, cf. Eq. (1–65), 1 1 I ′ ′ = (I + I )+ (I − I )2 +4I2 , (1–68a) y y 2 yy zz 2 yy zz yz q 1 1 I ′ ′ = (I + I ) − (I − I )2 +4I2 . (1–68b) z z 2 yy zz 2 yy zz yz q Elastic Beams in Three Dimensions 24 Chapter 1 – Beams in three dimensions

z z ′ z y′ y′

ϕ ′ ϕ x = x B y y B x = x′

z′

(a) (b)

Figure 1–14 Position of principal axes: (a) Iyz < 0 and (b) Iyz > 0.

′ ′ The coordinate axes y and z are known as the principal axes of the cross-section, whereas Iy′y′ and Iz′z′ are called the principal moments of inertia. It follows from Eqs. (1–63) and (1–67) that the choices of signs for sin(2ϕ) implies that Iy′y′ becomes the larger of the principal moments of inertia and Iz′z′ is the smaller principal moment of inertia. It is emphasised that this choice is performed merely to have a unique determination of ϕ. Three other choices of ϕ are possible π 3 obtained by additional rotations of the magnitudes 2 , π and 2 π relative to the indicated. If the cross-section has a symmetry line, and the y-axis is placed along this line, then Iyz =0. Hence, a symmetry line is always a principal axis. Since the principal axes are orthogonal, the z-axis is also a principal axis—even if the cross-section is not symmetric around the axis.

Example 1.2 Determination of principal axes coordinate system The cross-section analysed in Example 1.1 is reconsidered. The thin-wall approximation is used, so the moments of inertia are given by Eq. (l) in Example 1.1 and repeated here (without the primes):

233 3 3 92 3 3 11 3 3 Iyy ta 2.7738 ta ,Izz ta 4.3810 ta ,Iyz ta 1.5714 ta .(a) ≃ 84 ≈ · ≃ 21 ≈ · ≃− 7 ≈− · The position of the bending centre relatively to the top-left corner of the cross-section is provided in Fig. A. From Eq. (a) and Eq. (1–65) follows that

233 92 2 11 2 √9769 J = ta3 ta3 + 4 ta3 = ta3 , (b) 84 − 21 − 7 · 28 s    which by insertion into Eq. (1–64) provides:

2 11 ta3 28 88 sin(2ϕ)= · − 7 · = 0.8903, (c) − ta3 √9769 √9769 ≈ ·  233 ta3 92 ta3 28 3780 cos(2ϕ)= 84 − 21 · = 0.4553. (d) ta3 √9769 − 84 √9769 ≈− ·  · (continued)

DCE Lecture Notes No. 23 1.4 Uncoupling of axial and bending deformations 25

≃ 4 z′ yB 7 a 2a

≃ 9 2t zB 14 a

B z x = x′ 2a ϕ

t y′ t

a y Figure A Position of principal axes coordinated system for the thin-walled cross-section.

◦ From Eqs. (c) and (d) it is found that ϕ = 1.0217 radians corresponding to ϕ = 58.5418 . Hence, π 11 3 ϕ [0, 2 ] in agreement with Iyz = 7 ta < 0. Finally,∈ the moments of inertia in the− principal axes coordinate system follow from Eq. (1–67), i.e.

3 I ′ ′ 2 2 5.3423 ta , y y 1 233 92 1 233 92 11 = + + 4 ta3 = (e)   2 84 21 ± 2 s 84 − 21 7   3 Iz′z′         1.8124 ta .   Clearly, Iy′y′ is greater than any of Iyy or Izz, whereas Iy′y′ is smaller than the bending moments of inertia defined with respect to the original y- and z-axes. 

1.4.3 Equations of equilibrium in principal axes coordinates From now on it will be assumed that the (x,y,z)-coordinate system forms a principal axes coor- dinate system with origin at the bending centre. In this case, the system of differential equations (1–28) for a Timoshenko beam uncouples into three differential subsystems. Thus, the axial deformation is governed by the equation d dw EA x + q =0, (1–69) dx dx x   whereas bending deformation in the y-direction is defined by the coupled equations d dθ dw EI z + GA y − θ + m =0, (1–70a) dx z dx z dx z z     d dw GA y − θ + q =0, (1–70b) dx y dx z y    where the double index yy on the bending moment of inertia has been replaces by a single index y in order to indicate that the principal-axes coordinates are utilised.

Elastic Beams in Three Dimensions 26 Chapter 1 – Beams in three dimensions

Similarly, the flexural deformations in the z-directions are determined by d dθ dw EI y − GA z + θ + m =0, (1–71a) dx y dx z dx y y     d dw GA z + θ + q =0, (1–71b) dx z dx y z    where again the double index on the bending moment of inertia has been replaced by a single index. As seen from Eqs. (1–70) and (1–71), {wy,θz} and {wz,θy} are still determined by pairwise coupled ordinary differential equations of the second order. For a Bernoulli-Euler beam, the system of ordinary differential equations (1–33) uncouples completely into the following differential equations for the determination of wx, wy and wz: d dw EA x + q =0, (1–72a) dx dx x   d2 d2w dm EI y − q + z =0, (1–72b) dx2 z dx2 y dx   d2 d2w dm EI z − q − y =0. (1–72c) dx2 y dx2 z dx  

Example 1.3 Plane, fixed Timoshenko beam with constant load per unit length

Figure A shows a plane Timoshenko beam of the length l with constant bending stiffness EIz and shear stiffness GAy. The beam is fixed at both end-sections and is loaded with a constant load qy and a constant moment load mz. The displacement wy(x), the rotation θz(x), the shear force Qy(x) and the bending moment are to be determined.

δqy mz

z x

EIzz,GAy

l

y Figure A Fixed beam with constant load per unit length.

The differential equations for determination of wy(x) and θz(x) follow from Eq. (1–70). Thus,

2 d θz dwy d dwy EIz + GAy θz + mz = 0, GAy θz + qy = 0. (a) dx2 dx − dx dx −      According to Eq. (1–35), the boundary conditions are:

wy(0) = wy(l) = 0, θz(0) = θz(l) = 0. (b)

(continued)

DCE Lecture Notes No. 23 1.4 Uncoupling of axial and bending deformations 27

Integration of the second equation in Eq. (a) provides:

dwy Qy = GAy θz = qyx + c1 (c) dx − −  

Then, the following solution is obtained for θz(x) from the first equation in Eq. (a):

2 d θz 1 3 1 2 EIz = qyx (c1 + mz) EIzθz(x)= qyx (c1 + mz)x + c2x + c3. (d) dx2 − ⇒ 6 − 2

Further, the boundary conditions θz(0) = θz(l) = 0 provide

1 2 1 c3 = 0, c2 = qyl + (c1 + mz)l. (e) − 6 2

Hence, the following reduced form is obtained for θz(x):

1 3 2 2 θz(x)= qy(x xl ) 3(c1 + mz)(x xl) . (f) 6EIz − − −  Next, Eq. (f) is inserted into Eq. (c) which is subsequently integrated with respect to x, leading to the following solution for wy(x):

dwy GAy 3 2 2 GAy = qyx + c1 + qy(x xl ) 3(c1 + mz)x xl) dx − 6EIz − − − ⇒

1 2 GAy 1 4 2 2 1  3 2 GAywy(x)= qyx + c1x + c4 + qy(x 2x l ) (c1 + mz)(2x 3x l) . (g) − 2 6EIz 4 − − 2 −  

The boundary conditions wy(0) = wy(l) = 0 provide the integration constants 1 1 c4 = 0, c1 = qyl mz, (h) 2 − Φy + 1 where EI z (i) Φy = 12 2 GAyl Then, Eq. (a) and Eq. (f) provide the following solutions:

qy qy 2 2 mz x mz Φy 3 2 wy(x)= (l x)x + (l x) x (2x 3x l) 2GAy − 24EIz − − GAy Φy + 1 − 12EIz Φy + 1 − ⇒

qy 2 2 2 mz Φy 3 2 2 wy(x)= (l x) x + Φyl (l x)x (2x 3x l + xl ), (j) 24EIz − − − 12EIz Φy + 1 −  qy 3 2 2 mz Φy θz(x)= (2x 3x l + xl )+ (l x)x. (k) 12EIz − 2EIz Φy + 1 −

The non-dimensional parameter Φy is a measure of the influence of the shear deformations. For a 1 2 5 rectangular cross-section with the height h we have Iz = 12 h A and Ay = 6 A. Then Φy becomes 72 h2 E Φy = . (l) 5 · l2 G Hence, shear deformations are primarily of importance for short and high beams. On the other hand, for long beams with a small height of the cross-section, shear deformations are of little importance, i.e. only the bending deformation is significant. (continued)

Elastic Beams in Three Dimensions 28 Chapter 1 – Beams in three dimensions

For Bernoulli-Euler beams we have γxy = 0, corresponding to GAy = , cf. Eqs. (1–16) and (1–27). ∞ Then, Φy = 0 and Eqs. (j) and (k) reduce to

qy 2 2 wy(x)= (l x) x , (m) 24EIz −

qy 3 3 θz(x)= (2x 3x l + xl). (n) 12EIz −

It is remarkable that the distributed moment load mz does not induce any displacements or rotations in the considered beam with Bernoulli-Euler kinematics. The shear force Qy(x) and bending moment Mz(x) follow from Eq. (c), Eq. (h) and Eq. (k), respec- tively, i.e. 1 1 Qy(x)= qyx + c1 = qy(l 2x) mz, (o) − 2 − − Φy + 1

dθz qy 2 2 mz Φy Mz(x)= EIz = (6x 6xl + l )+ (l 2x). (p) dx 12 − 2 Φy + 1 − For a Bernoulli-Euler beam these results reduce to 1 Qy(x)= qy(l 2x) mz, (q) 2 − −

qy 2 2 Mz(x)= (6x 6xl + l ). (r) 12 − The constant moment load mz only induces a constant shear force of magnitude mz, whereas Mz(x) − is not affected by this load. Especially, for x = 0 and x = l, Eq. (o) and Eq. (p) provide

1 1 1 Φy Qy(0) = qyl mz, My(0) = qyl + mzl, (s) 2 − 12 2 Φy + 1

1 1 2 1 Φy Qy( l)= qyl mz, My( l)= qyl mzl. (t) − 2 − 12 − 2 Φy + 1 The displacement at the midpoint x = l/2 follows from Eq. (k):

4 qyl wy (l/2) = (1 + 4Φy). (u) 384EIz The first and second terms within the parenthesis specify the contributions from bending and shear, re- spectively. Again the parameter Φy reveals itself as a measure of the relative contribution from shear deformations. 

1.5 Normal stresses in beams

For at beam without warping, the normal stress σxx(x,y,z) in terms of the generalised strains follows from Eqs. (1–17), (1–20) and (1–24):

σxx = E(ε − κzy + κyz). (1–73)

In the principal axes coordinate system, where Sy = Sz = Iyz = 0, the generalised strains {ε,κy,κz} are related to the conjugated generalised stresses {N,My,Mz} as determined from

DCE Lecture Notes No. 23 1.6 The principle of virtual forces 29

Eqs. (1–45a), (1–45b), (1–46) and (1–47) for Sy = Sz = Iyz =0, i.e.

N My Mz ε = , κy = , κz = . (1–74) EA EIy EIz Insertion of Eq. (1–74) into Eqs. (1–19) and (1–24) provides the result for the axial stress in terms of the generalised stresses,

N Mz My σxx = − y + z. (1–75) A Iz Iy Equation (1–75) is due to Navier, and is therefore referred to as Navier’s formula. It should be noticed that Eq. (1–75) presumes that the stresses are formulated in a principal axes coordinate system, so Iy and Iz indicate the principal moments of inertia. The relation is valid for both Timoshenko and Bernoulli-Euler beams. This is so because only the relation (1–15), but not the relation (1–18) has been utilised. Hence, Eq. (1–75) is based on the assumption that plane cross-sections remain plane, but not that they remain orthogonal to the beam axis. The so-called zero line specifies the line in the (y,z)-plane on which σx =0. The analytical expression for the zero line becomes

N M M − z y + y z =0. (1–76) A Iz Iy It is finally noted that warping introduces displacements in the axial direction in addition to those provided by bending. However, if the torsion is homogeneous, these displacements will not introduce any normal strains and therefore no normal stresses. Hence, Navier’s formula is also valid in the case of St. Venant torsion, but in the case of Vlasov torsion, or inhomogeneous torsion, additional terms must be included in Eq. (1–75).

1.6 The principle of virtual forces

In this section the principle of virtual forces is derived for a plane Timoshenko beam of the length l. The deformation of the beam is taking place in the (x, y)-plane. In the referential state, the left end-section is placed at the origin of the coordinate system and the x-axis is placed along the bending centres of the cross-sections, see Fig. 1–15. The principle of virtual forces is the dual to the principle of virtual displacements. In the principle of virtual displacements the actual sectional forces and sectional moments are assumed to be in equilibrium with the loads and the reaction forces applied at the end sections. The virtual displacements and rotations are considered as arbitrary increments to the actual displacements and they only need to fulfil homogeneous kinematic boundary conditions, so that the combined field made up by the actual and the virtual fields always fulfils the actual non-homogeneous boundary conditions as given by Eq. (1–35). Further, the generalised virtual strains defining the internal virtual work must be derived from the virtual displacement and rotation fields. In contrast, the principle of virtual forces presumes that the displacements and rotations of the beam are fulfilling the kinematic boundary conditions, and that the generalised internal strains are compatible to these fields. The actual loads on the beam are superimposed with the virtual incremental loads per unit length δqx and δqy, the virtual moment load per unit length δmz,

Elastic Beams in Three Dimensions 30 Chapter 1 – Beams in three dimensions

δqx δqy

z δmz

δMz,1 δN δN2

δN1 x δMz δMz,2 δQy δQy,2 δQy,1

l

y

Figure 1–15 Virtual internal and external forces.

the virtual reaction forces δNj and δQy,j along the x- and y-directions, and the virtual reaction moments δMz in the z-directions, where j =1 and j =2 indicate the left and right end-sections, respectively. Due to the load increments, the internal section forces and section moment achieve increments δN, δQy and δMz, see Fig. 1–15. These variational fields are assumed to be in equilibrium with the variational load fields δqx, δqy and δmz, and to comply with the variations δNj ,δQy,j and δMz,j of the reaction forces and reaction moments. In what follows, δN, δQy and δMz will be referred to as the virtual internal forces, whereas δqx, δqy, δmz, δNj , δQy,j and δMz,j are called the virtual external forces. The starting point is taken in the kinematical conditions provided by Eqs. (1–12) and (1–17) and rewritten in the form dw dw dθ x − ε =0, y − θ − γ =0, z − κ =0. (1–77) dx dx z xz dx z

The virtual internal forces are related to the virtual external loads per unit length δqx, δqy and δmz via the following equations of equilibrium, cf. Eqs. (1–4a) and (1–4b): d(δN) d(δQ ) d(δM ) + δq =0, y + δq =0, z + δQ + δm =0. (1–78) dx x dx y dx y z The first equation in Eq. (1–78) is multiplied with δN(x), the second equation is multiplied with δQy(x), and the third equation with δM(z). Next, the equations are integrated from x =0 to x = l, and the three resulting equations are added, leading to the identity

l dw dw dθ δN x − ε + δQ y − θ − γ + δM z − κ dx =0 (1–79) dx y dx z xy z dx z Z0        Integration by parts is carried out on the first terms within the innermost parentheses, leading to

l l d(δN) d(δQ ) d(δM ) δNw + δQ w + δM Q − w + y w + z θ − δQ θ dx x y y z z dx x dx y dx z y z  0 Z0   l = δNε + δQy · γxy + δMz · κz dx. (1–80) Z0  

DCE Lecture Notes No. 23 1.6 The principle of virtual forces 31

Upon utilisation of Eq. (1–78), this is reduced to

l l δN · wx + δQy · wy + δMz · θz + δqx · wx + δqy · wy + δmz · θz dx 0 Z0 h l i   = δN · ε + δQy · γxy + δMz · κz dx. (1–81) 0 Z   The generalised strains on the right-hand side of Eq. (1–81) are now expressed in mechanical quantities by means of Eq. (1–74). Further, δN, δQy and δMz fulfil the following boundary conditions at x =0 and x = l, cf. Fig. 1–15,

δN(0) = −δN1, δQy(0) = −δQy,1, δMz(0) = −δMz,1, (1–82a)

δN( l)= δN2, δQy( l)= δQy,2, δMz( l)= δMz,2. (1–82b)

Equation (1–81) then obtains the following final form:

2 l δNj · wx,j + δQy,j · wy,j + δMz,j · θz,j + δqx · wx + δqy · wy + δmz · θz dx j=1 0 X   Z   l δN · N δQ · Q δM · M = + y y + z z dx, (1–83) EA GA EI Z0  y z  where wx,j , wy,j and θz,j denote the displacements in the x- and y-directions and the rotation in the z-direction at the end-sections, respectively. Equation (1–83) represents the principle of virtual forces. The left- and right-hand sides represent the external and internal virtual work, respectively. The use of Eq. (1–83) in determining the displacements and rotations of a Timoshenko beam is demonstrated in Examples 1.4 and 1.5 below. Furthermore, the principle of virtual forces may be used to derive a stiffness matrix for a Timoshenko beam element as shown later.

Example 1.4 End-displacement of cantilevered beam loaded with a force at the free end Figure A shows a plane Timoshenko beam of the length l with constant axial stiffness EA, shear stiffness GAy and bending stiffness EIz. The beam is fixed at the left end-section and free at the right end-section, where it is loaded with a concentrated force Qy,2 in the y-direction. The displacement wy,2 at the free end is searched. The principle of virtual forces Eq. (1–83) is applied with the following external virtual loads: δqx = δqy = δmz = 0, δN1 = δQy,1 = δMz,1 = δN2 = δMz,2 = 0 and δQy,2 = 1. Further N(x) = 0. Then, Eq. (1–83) reduces to

l δQy Qy δMz Mz 1 wy,2 = · + · dx. (a) · GAy EIz Z0   (continued)

Elastic Beams in Three Dimensions 32 Chapter 1 – Beams in three dimensions

Qy,2 δQy,2 = 1

EA,GAy,EIz z x z x

l l

y y

Mz δMz

Qy,2 · ℓ l

Qy δQy

Qy,2 1

Figure A Fixed plane Timoshenko beam loaded with a concentrated force at the free end: Actual force and section forces (left) and virtual force and section forces (right).

The variation of the bending moment Mz(x) and the shear force Qy(x) from the actual load Qy,2 has been shown in Fig. A on the left. The corresponding variational moment field δMz(x) and shear force δQy(x) from δQy,2 = 1 are shown in Fig. A on the right. Insertion of these distributions in Eq. (b) provides the solution

3 3 Qy,2l 1 Qy,2l 1 Qy,2l wy,2 = + = (4+Φy) , (b) GAy 3 EIz 12 · EIz

where Φy is given by Eq. (i) in Example 1.3. The deformation contributions from shear and bending are additive. This is a consequence of the additive nature of the flexibilities indicated by Eq. (1–83) in contribution to the fact that the beam is statically determinate, which provides the fields Mz(x) and Qy(x) as well as δMz(x) and δQy(x) directly. 

Example 1.5 End-deformations of fixed beam loaded with a moment at the free end The beam described in Example 1.4 is considered again. However, now the free end is loaded with a concentrated moment Mz,2. The displacement wy,2 and the rotation θz,2 of the end-section is to be found. At the determination of wy,2 from Mz,2, the principle of virtual forces given by Eq. (1–83) is again applied with δQy,2 = 1 and all other external variational loads equal to zero, leading to Eq. (a) in Example 1.4. However, Mz(x) and Qy(x) are now caused by Mz,2, and are given as shown in Fig. A, whereas δMz(x) and δQy(x) are as shown in Fig. A of Example 1.4. Then, wy,2 becomes

l 2 δMz Mz 1 My,2l wy,2 = · dx = . (a) EI 2 EI Z0 z z (continued)

DCE Lecture Notes No. 23 1.7 Elastic beam elements 33

EA,GAy,EIz z x z x

Mz,2 δMz,2 = 1 l l

y y Mz δMz

Mz,2 1

Qy = 0 δQy = 0

Figure A Fixed plane Timoshenko beam loaded with a moment at the free end: Actual moment and section forces (left) and virtual moment and section forces (right).

At the determination of θz,2 the principle of virtual forces Eq. (1–83) is applied with the following external virtual loads δqx = δqy = δmz = 0, δN1 = δQy,1 = δMz,1 = δN2 = δQy,2 = 0 and δMz,2 = 1. Then, Eq. (1–83) reduces to

l δQy Qy δMz Mz 1 θz,2 = · + · dx. (b) · GAy EIz Z0  

The variation of Qy(x) and Mz(x) from My,2 has been shown in Fig. A on the left, and the variation of δQy(x) and δMz(x) from δMz,2 = 1 is shown in Fig. A on the right. Then θz,2 becomes

l 1 Mz,2 Mz,2l θz,2 = · dx = . (c) EI EI Z0 z z

In the present load case, the shear force is given as Qy(x) = 0. Consequently it will not induce any contributions in Eq. (a) and Eq. (c). 

1.7 Elastic beam elements

When frame structures consisting of multiple beams are to be analysed, the establishment of analytical solutions is not straightforward and instead a numerical solution must be carried out. For this purpose, a discretization of the frame structure into a number of so-called beam ele- ments is necessary, eventually leading to a finite-element model. The aim of the present section is not to provide a full introduction to the finite-element method for the analysis of frame struc- tures, e.g. tower blocks with a steel frame as the load-carrying structure. However, a formula- tion is given for a single beam element to be applied in such analyses. Both the Timoshenko and Bernoulli-Euler beam theories are discussed in this context, and plane as well as three- dimensional beams are touched upon.

Elastic Beams in Three Dimensions 34 Chapter 1 – Beams in three dimensions

1.7.1 A plane Timoshenko beam element Firstly, the stiffness matrix and element load vector is derived for a plane Timoshenko beam ele- ment with constant axial stiffness EA, shear stiffness GAy and bending stiffness EIz,cf. Fig.1– 16. The stiffness relation is described in an (x, y)-coordinate system with origin at the left end- section and the x-axis along the bending centres.

z

EA,GAy,EIz Mz,1,θz,1 N2,wx,2 1 2 x N1,wx,1 Mz,2,θz,2 Qy,1,wy,1 Qy,2,wy,2

l

y

Figure 1–16 Plane Timoshenko beam element with definition of degrees of freedom and nodal reaction forces.

At the end-nodes, nodal reaction forces Nj and Qy,j are acting along the x- and y-directions, respectively, and reaction moments Mz,j are applied around the z-axis. Here, j = 1 and j = 2 stand for the left-end and right-end nodes of the beam element, respectively, and the reaction forces and moments are in equilibrium with the remaining external loads on the element for arbitrary deformations of the beam. The element has 6 degrees of freedom defining the displacements and rotations of the end- sections, cf. Fig. 1–16. These are organised in the column vector

we1 T we = = wx,1 wy,1 θz,1 wx,2 wy,2 θz,2 (1–84) we  2    The sub-vector wej defines the degrees of freedom related to element node j. Similarly, the reaction forces Nj, Qy,j and Mz,j, j = 1, 2, at the end-sections, work conju- gated to wx,j , wy,j and θz,j, are stored in the column vector

re1 T re = = N1 Qy,1 Mz,1 N2 Qy,2 Mz,2 (1–85) re  2   

py

px mz

Mz,1 N2 x N1 z Mz,2 Qy,1 Qy,2 y

Figure 1–17 External loads and reaction forces from external loads on a plane beam element.

The equilibrium of the beam element relating the nodal reaction forces to the degrees of freedom of the element may be derived by the principle of virtual displacements as demonstrated

DCE Lecture Notes No. 23 1.7 Elastic beam elements 35 in a subsequent paper. The resulting equilibrium equations on matrix form may be written on the form re = Kewe + fe. (1–86)

The vector fe in Eq. (1–86) represents the nodal reaction forces from the external element loads when we = 0, i.e. when the beam is fixed at both ends as shown in Fig. 1–17. We shall merely consider constant element loads qx and qy per unit length in the x- and y-directions, and a con- stant moment load per unit length mz in the z-direction, see Fig. 1–17. The reaction forces and reaction moments follow from Eqs. (s) and (t) in Example 1.3:

1 − 2 qxl 1 − qyl + mz  2  − 1 q l2 − 1 Φy m l 12 y 2 Φy +1 z fe =  1  . (1–87)  − 2 qxl   1   − 2 qyl − mz   1 2 1 Φy   qyl − mzl   12 2 Φy +1    The matrix Ke in Eq. (1–86) denotes the stiffness matrix in the local (x,y,z)-coordinate system. Let wi denote the ith component of we. Then, the ith column in Ke represents the nodal reaction forces for fe = 0, and with wi = 1 and wj = 0, j 6= i. These forces are obtained following the derivations in Example 1.3 from Eq. (a) to Eq. (t) with qy = mz =0 and with the boundary condition in Eq. (b) replaced by the indicated conditions. Because of the symmetry of the prob- lem, only two such analyses need to be performed. Still, this is a rather tedious approach. Partly because of this, and partly in order to demonstrate an alternative approach, the stiffness matrix will be derived based on the principle of virtual forces.

Undeformed state Undeformed state x x

θz wy

wx

y y (a) (b)

Figure 1–18 Rigid-body modes of a plane beam element: (a) Translation and (b) rotation.

The beam element has 6 degrees of freedom, by which a total of 6 linear independent modes of deformation may be defined. These consist of 3 linear independent rigid body modes and 3 linear independent elastic modes. The rigid modes may be chosen as a translation in the x-direction, a translation in the y-direction and a rotation around the z-direction as shown in Fig. 1–18. Any rigid body motion of the beam element may be obtained as a linear combination of these component modes of deformation. Obviously, the rigid body motions do not introduce stresses in the beam. Hence, the axial force N, the shear force Qy and the bending moment Mz are all zero during such motions. Since, axial elongations are uncoupled from bending deformations, the elastic elongation mode is uniquely defined as shown in Fig. 1–19a. The two bending deformation modes may be

Elastic Beams in Three Dimensions 36 Chapter 1 – Beams in three dimensions

N0 N0 x

u0 u0 2 2 y N

N0

(a)

2Ma l Ms Ms Ma θa x x θs θs θa 2Ma Ma l y y Qy

2Ma Qy = 0 l

Mz Ma

Ms Ma Mz (b) (c) Figure 1–19 Elastic modes and related section forces in a plane beam element: (a) Axial elongation; (b) symmetric bending and (c) antisymmetric bending.

chosen in arbitrarily many ways. Typically, these are chosen by prescribing an angle of rotation at the other end-section. Following an idea by Krenk (2001), a more convenient formulation may be obtained by choice of two bendingmodes symmetric and anti-symmetric around the mid-point of the beam element as shown in Figs. 1–19b and 1–19c. It should be noticed that these modes also apply if the material properties of the beam are not symmetrical around the mid-point. The axial elongation and conjugated axial force related to the axial elongation mode are de- noted u0 and N0, respectively. The symmetric and anti-symmetric bending modes are described by the end-section rotations θs and θa defined in Figs. 1–19b and 1–19c, respectively. The con- jugated moments are denoted Ms and Ma, respectively. The related distributions of the shear force Qy(x) and the bending moment Mz(x) are shown in Figs. 1–19b and 1–19c.

The shear force is equal to Qy = 0 in symmetric bending, because the bending moment is constant. Then, no shear deformations are related to this mode. In contrast, a constant shear force appears in the anti-symmetric bending mode. Hence, the deformations occurring in this mode are affected by bending as well as shear contributions. At first the constitutive relations between the deformation measures and the conjugated gen- eralised strains for the indicated elastic modes are found by means of the principle of virtual forces Eq. (1–83). In all cases, the beam element is unloaded, so δqx = δqy = δmz =0. For the axial elongation mode N = N0, Qy = Mz = 0, and δN = 1, δN1 = −1, δN2 = 1. Further,

DCE Lecture Notes No. 23 1.7 Elastic beam elements 37

1 1 wx,1 = − 2 u0, wx,2 = 2 u0, wy,j = θz,j =0. Then, Eq. (1–83) reduces to 1 1 l 1 · N l dx l (−1) · (− u )+1 · u = 0 dx ⇒ u = N = N . (1–88) 2 0 2 0 EA 0 0 EA EA 0 Z0 Z0 The last statement holds for a beam element with constant axial stiffness EA. If EA varies, the integral in the middlemost statement must be evaluated analytically or numerically. For the symmetric bending mode N = Qz = 0, Mz = −Ms, δMz = −1, δMz,1 = 1, and δMz,2 = −1. Further, θz,1 = θs, θz,2 = −θs and wx,j = wy,j =0. Then, Eq. (1–83) provides l (−1)(−M ) l dx l 1 · θ + (−1) · (−θ )= s dx ⇒ 2θ = M · = M . (1–89) s s EI s s EI EI s Z0 z Z0 z z Again, the last statement only applies for a homogeneous beam, whereas the middlemost state- ment applies for any variation of the bending stiffness EIz along the beam. Ma For the anti-symmetric bending mode N = 0, Qy = 2 l , Mz(x) = (−1+2x/l)Ma, and δQy = 2δMz(x)/l = (−1+2x/l). Further, θz,1 = θa,θz,2 = θa, wx,j = wy,j = 0. Then, Eq. (1–83) provides

l 2 2Ma x x · (−1+2 )(−1+2 )Ma 1 · θ +1 · θ = l l + l l dx ⇒ a a GA EI Z0 y z ! 4 l dx l (−1+2 x )2 2θ = M + l dx ⇒ a a l2 GA EI Z0 y Z0 z ! 1 1 l 1 l EI 2θ = M 4l + = (1+Φ )M , Φ = 12 z . (1–90) a a GA l2 3 EI 3 EI y a y GA l2  y y  y y As discussed in Example 1.3, the non-dimensionalparameter Φy defines the contributionof shear flexibility relatively to the bending flexibility. The flexibility relations provided by Eqs. (1–88), (1–89) and (1–90) may be written in the following equivalent stiffness matrix formulation: r0 = K0w0, (1–91) where EA N0 u0 l 0 0 EIz r0 = Ms , w0 = 2θs , K0 = 0 l 0 . (1–92)  M   2θ   0 0 3 EIz  a a 1+Φy l       The nodal reaction forces re and r0 in Eq. (1–85) and Eq. (1–92) are related via the transfor- mation re = Sr0 = SK0w0, (1–93) where −1 0 0 0 0 2/l  0 1 1  S = . (1–94)  1 0 0     0 0 −2/l     0 −1 1      Elastic Beams in Three Dimensions 38 Chapter 1 – Beams in three dimensions

Similarly, the elastic deformation measures stored in w0 can be expressed by the degrees of freedom of the element stored in we as follows (see Fig. 1–20): u0 = wx,2 − wx,1, (1–95)

1 2 θz,1 = θa + θs + l (wy,2 − wy,1) 2θa = θz,1 + θz,2 − l (wy,2 − wy,1) ⇒ (1–96) 1   θz,2 = θa − θs + l (wy,2 − wy,1)   2θs = θz,1 − θz,2.  

wy,1

wy,2

θa + θs

θz,1 θz,2 θa − θs Figure 1–20 Connection between elastic deformation measures and element degrees of freedom.

Equations (1–95) and (1–96) may be rewritten in the common matrix form

T w0 = S we, (1–97) where S is given by Eq. (1–94). Insertion of Eq. (1–97) into Eq. (1–93) provides upon compari- son with Eq. (1–86):

T T re = SK0S we ⇒ Ke = SK0S . (1–98)

Insertion of K0 and S as given by Eq. (1–92) and Eq. (1–94) provides the following explicit solution for Ke:

Al2 0 0 −Al2 0 0 0 12 I 6 I l 0 − 12 I 6 I l 1+Φy z 1+Φy z 1+Φy z 1+Φy z −  6 4+Φy 2 6 2 Φy 2  E 0 1+Φ Izl 1+Φ Izl 0 − 1+Φ Izl 1+Φ Izl Ke y y y y (1–99) = 3  2 2  . l  −Al 0 0 Al 0 0  12 6 12 6  0 − Iz − Izl 0 Iz − Izl   1+Φy 1+Φy 1+Φy 1+Φy  −  6 2 Φy 2 6 4+Φy 2   0 Izl Izl 0 − Izl Izl   1+Φy 1+Φy 1+Φy 1+Φy    The corresponding result for a plane Bernoulli-Euler beam element is obtained simply by setting Φy = 0. The equivalent element relations for a three-dimensional beam formulated in a (x, y, x) principal axes coordinate system are given in the next section.

DCE Lecture Notes No. 23 1.7 Elastic beam elements 39

Example 1.6 Deformations of a plane Timoshenko beam structure Figure A shows a plane beam structure ABC consisting of two Timoshenko beam elements AB and BC of the lengths l and l/2, respectively. The loads on the structure and the resulting displacements are described in the indicated (x,y)-coordinate system. The shear stiffness GAy and the bending stiffness EIz are constant and the same in both beam elements. The structure is fixed at point A, free at point C, and simply supported at point B. Both beam elements are loaded with a constant load per unit length qy. Additionally, beam BC is loaded with a concentrated load Py = qyl at the free end C.

qy Py = qyl

z x A1 B2 C

GAy,EIz l l/2 y Figure A Plane Timoshenko beam structure consisting of two beam elements.

We want to determine the displacement of point C in the y-direction, and the reaction forces and moments at the support points A and B. The calculations are performed with the shear stiffness given as EI GA = 120 z . (a) y l2 We shall refer to the beam elements AB and BC with the index e = 1 and e = 2, respectively. With reference to Eq. (i) in Example 1.3, the shear flexibility parameters for the two beam elements become: 12EI 12EI z z (b) Φy1 = 2 = 0.1, Φy2 = 2 = 0.4. GAyl GAy(l/2)

θ1,M1 θ2 θ2 θ3

1 2 w1 w2 w2 w3

Q1 Q2

Figure B Global degrees of freedom and reaction forces in the plane Timoshenko beam structure.

Since, the axial deformations are disregarded, each element has 4 degrees of freedom out of which two are common, namely the displacement and rotation at point B. The related global degrees of freedom have been defined in Fig. B. The element stiffness matrices follow from Eq. (1–99) and Eq. (b):

12 6l 12 6l 12 3l 12 3l 2 − 2 2 − 2 EIz 6l 4.1l 6l 1.9l 8EIz 3l 1.1l 3l 0.4l K1 =  − , K2 =  − . (c) 1.1l3 12 6l 12 6l 1.4l3 12 3l 12 3l − − − − − −  6l 1.9l2 6l 4.1l2   3l 0.4l2 3l 1.1l2   − −   −      (continued)

Elastic Beams in Three Dimensions 40 Chapter 1 – Beams in three dimensions

The corresponding element loads become, cf. Eq. (1–87):

1 1 l 2 qyl 6 2 qy 2 12 −1 2 1 −1 ·l 2 1 f 12 qyl l f 12 qy( 2 ) l 1 =  − 1  = qyl   , 2 =  −1 l  = qyl   . (d) 2 qyl − 12 6 2 qy 2 + Py − 48 60 −1 2 − 1 l 2  qyl   l   q ( )   l   12   −   12 y 2   −          The global equilibrium equation, made up of contributions from both elements, has the structure

r = Kw + f, (e)

where

p q Q1 w1 p q 24 r1 M1 θ1 f1 4l  | |       | |    p q Q2 w2 p q 1 36 r = | | = , w = , f = | | = qyl , (f)  x y   0   θ2   x y  − 48  3l   | |       | |   −   r2   0   w3   f2   60   | |       | |     x y   0   θ   x y   l       3               

p q K1  | p | q  K = | |  x y   | |   K2   | |   x y     12 6  12 6 1.1 1.1 l 1.1 1.1 l 0 0 6 4.1 2 − 6 1.9 2 1.1 l 1.1 l 1.1 l 1.1 l 0 0  12 6 12− 96 6 24 96 24  EIz l + + l l 1.1 1.1 1.1 1.4 1.1 1.4 1.4 1.4 (g) = 2 −6 −1.9 2 6 24 −4.1 8.8 2 − 24 3.2 2 . l  1.1 l 1.1 l 1.1 + 1.4 l 1.1 + 1.4 l 1.4 l 1.4 l   − 96 24  −96 24   0 0 1.4 1.4 l 1.4 1.4 l   −  − 2  − 2   0 0 24 l 3.2 l 24 l 8.8 l   1.4 1.4 − 1.4 1.4    The details in the derivation of the matrix equation Eq. (e), including the formation of the column vectors r and f and the global stiffness matrix K by adding contribution from element components, will be explained in a later chapter. The displacement degrees of freedom w1 and w2 at the nodes A and C are both equal to zero. Similarly, the rotation θ1 at the fixed support at A is zero. When these values are introduced in w0, Eq. (e) provides the values of the reaction components Q1, M1 and Q2 if the remaining unconstrained degrees of freedom θ2, w3 and θ3 are inserted. These are determined from the corresponding equations in Eq. (e) using w1 = θ1 = w2 = 0, leading to

4.1 8.8 2 24 3.2 2 ( 1.1 + 1.4 )l 1.4 l 1.4 l θ2 3l 0 EI 24 −96 24 1 − 3 1.4 l 1.4 1.4 l w3 qyl 60 = 0 l  −3.2 2 24 −8.8 2    − 48     ⇒ l l l θ3 l 0 1.4 − 1.4 1.4 −         θ2l 4 0.1453 qyl w3 = 0.1274 . (h)   EIz   θ3l 0.2912     Insertion of Eq. (h) along with w1 = θ1 = w2 = 0 into the remaining equation (e) provides the following solution for the reaction components Q1, M1 and Q2: (continued)

DCE Lecture Notes No. 23 1.7 Elastic beam elements 41

6 Q1 1.1 l 0 0 4 0.1453/l 24 EIz 1.9 2 qyl 1 M1 = 3 1.1 l 0 0 0.1274 qyl 4l   l  6 24 96 24  − EIz   − 48   ⇒ Q2 ( + )l l 0.2912/l 36 − 1.1 1.4 − 1.4 1.4         Q1l 0.2927 2 M1 = qyl 0.1677 . (i)     O2l 2.7927 −     In case of Bernoulli-Euler kinematics, corresponding to Φy1 = Φy2 = 0, the corresponding solutions become:

θ1l 4 0.1354 qyl w3 = 0.1172 (j)   EI   θ2l 0.2812     Q1l 0.3125 2 M1 = qyl 0.1875 (k)     Q2l 2.8125 −     Comparison of Eqs. (h) and (j) reveals that the displacement w3 as well as the rotations θ1 and θ3 are increased by the shear flexibility. This is so because bending and shear deformations in general are coupled for statical indeterminate structures. At the same time, a comparison of Eqs. (i) and (k) shows that Bernoulli-Euler beam kinematics lead to higher stresses than Timoshenko beam theory, which is due to the fact that the shear stiffness in the Bernoulli-Euler beam is infinite. 

1.7.2 A three-dimensional Timoshenko beam element The formulation of the beam-element stiffness matrix is now extended to three dimensions. This involves flexural displacements in two directions, axial displacements and, in addition to this, twist of the beam. It is assumed that the beam element is straight, of the length l and with constant cross-section. The element relation is described in a principal axes (x,y,z)-coordinate system with origin at the left end-section and the x-axis placed along the bending centres of the cross-sections. Only St. Venant torsion is taken into consideration. The axial stiffness EA, the shear stiffnesses GAy and GAz in the y- and z-directions, respectively, the torsional stiffness GK, and the principal inertial bending stiffnesses EIy and EIz around the y- and z-axes are all constant along the beam element. The degreesof freedomof the elementare made up by the 6 components wx,j , wy,j and wz,j , j = 1, 2, providing the displacements of the bending centres, and the 6 components θx,j , θy,j and θz,j defining the rotation of the end-sections. Again, j = 1 and j = 2 refer to the left and right end-sections, respectively. The nodal reaction forces conjugated to the indicated degrees of freedom consist of the axial forces Qy,j and Qz,j in the y- and z-directions, respectively, the torsional moments Mx,j and the bending moment components My,j and Mz,j in the y- and z-directions. Finally, the element loadings consist of constant loads per unit length {qx, qy, qz} and constant moment loads per unit length {mx,my,mz} in the x-, y- and z-directions. No concentrated element forces or moments are considered. The loads qy and qz as well as the shear forces Qy,j and Qz,j are assumed to act through the so-called shear centre, leading to an uncoupling of the flexural and torsional deformations. The definition of the shear centre and further details about uncoupling of torsional and flexural displacements are given in Chapter 2.

Elastic Beams in Three Dimensions A,GA

42 Chapter 1 – Beams in three dimensions

qx

mx Mx,1 N1 N2 Mx,2 x θ w w θ x,1 x,1 EA,GK x,2 x,2

qy mz z

θz,1,Mz,1 θz,2,Mz,2 x

GAy,EIz

wy,1,Qy,1 wy,2,Qy,2 y

qz my y

θy,1,My,1 θy,2,My,2 x

GAz,EIy

wz,1,Qz,1 wz,2,Qz,2 z Figure 1–21 Three-dimensional Timoshenko beam element with definition of degrees of freedom, nodal reaction forces, element loads and sectional properties.

The element equilibrium equations may be expressed on the matrix form, cf. Eq. (1–86),

re = Kewe + fe (1–100)

Here, re and we are 12-dimensional column vectors storing the reaction forces and the element degrees of freedom, respectively, cf. Eqs. (1–84) and (1–85),

N1 wx,1 Q w  y,1   y,1  Qz,1 wz,1  Mx,1   θx,1       My,1   θy,1      re1  Mz,1  we1  θz,1  re = =   , we = =   . (1–101) re  N2  we  wx,2   2     2     Qy,2   wy,2       Qz,2   wz,2       M   θ   x,2   x,2   My,2   θy,2       Mz,2   θz,2          Likewise, fe is 12-dimensional column vector storing the contributions to the reaction forces

DCE Lecture Notes No. 23 1.7 Elastic beam elements 43 from the element loads, given as, cf. Eq. (1–87),

1 − 2 qxl 1 − 2 qyl + mz  1  − 2 qzl − my 1  − 2 mxl   1 q l2 − 1 Φz m l   12 z 2 1+Φz y   1 2 1 Φy  fe  − qyl − mzl  1  12 2 1+Φy  fe = = 1 , (1–102) fe   2  − 2 qxl     1   − 2 qyl − mz   1   − 2 qzl + my   1   − 2 mxl   − 1 q l2 − 1 Φz m l   12 z 2 1+Φz y   1 2 1 Φy   qyl − mzl   12 2 1+Φy    where Φy and Φz are given as, cf. Eq. (i) in Example 1.3,

EI EI z y (1–103) Φy = 12 2 , Φz = 12 2 . GAyl GAzl

Finally, Ke specifies the element stiffness given as, cf. Eq. (1–99),

EA EA l 00 0 00 − l 00 0 00 z z z z 0 k11 0 0 0 k12 0 k13 0 0 0 k14  y y y y  0 0 k11 0 k12 0 0 0 k13 0 k14 0 GK GK  0 0 0 l 00 0 00 − l 0 0   y y y y   0 0 k12 0 k22 0 0 0 k23 0 k24 0   z z z z   0 k12 0 0 0 k22 0 k23 0 0 0 k24  Ke =  EA EA  , (1–104) − l 00 0 00 l 00 0 00   z z z z   0 k13 0 0 0 k23 0 k33 0 0 0 k34   y y y y   0 0 k 0 k 0 0 0 k 0 k 0   13 23 33 34   0 0 0 − GK 00 0 00 GK 0 0   l l   0 0 ky 0 ky 0 0 0 ky 0 ky 0   14 24 34 44   0 kz 0 0 0 kz 0 kz 0 0 0 kz   14 24 34 44    where

z z z z k11 k12 k13 k14 12 6l −12 6l kz kz kz EI (4+Φ )l2 −6l (2 − Φ )l2 22 23 24 z y y (1–105)  z z  = 2   , k33 k34 (1+Φy)l 12 −6l z 2  k   (4+Φy)l   44        y y y y k11 k12 k13 k14 12 −6l −12 −6l y y y 2 2 k k k EIy (4+Φ )l 6l (2 − Φ )l 22 23 24 z z (1–106)  y y  = 2   . k33 k34 (1+Φz)l 12 6l y 2  k   (4+Φz)l   44        Elastic Beams in Three Dimensions 44 Chapter 1 – Beams in three dimensions

The element equilibrium relation provided by Eq. (1–100) presumes that only St. Venant torsion is taken into consideration, corresponding to the torsional equilibrium equations

Mx,1 GK 1 −1 θx,1 1 1 = − mxl . (1–107) Mx,2 l −1 1 θx,2 2 1        The torsional constant K is determined in the next chapter. It will be shown that the inclusion dθx,1 dθx,2 of Vlasov torsion requires the introduction of two extra degrees of freedom dx and dx . The conjugated generalised stresses are the so-called bimoments. Hence, a Timoshenko beam element, where both St. Venant and Vlasov torsion are taken into consideration, is described by a total of 14 degrees of freedom.

1.8 Summary

In this chapter, the basic theory of Timoshenko and Bernoulli-Euler beams in three-dimensional space has been presented. Some of the main topics covered are summarised below.

Beams are one-dimensional structures that may carry loads in three dimensions including axial forces, shear forces in two orthogonal directions and moments around three directions. Bernoulli-Euler beam kinematics assume that cross-sections remain orthogonal to the beam axis during deformation. Hence, no shear deformation occurs. Timoshenko beam kinematics include shear flexibility, but still a cross-section remains plane during deformation. Hence, shear strains and stresses are homogeneousover the beam height. The bending centre of a beam cross-section is defined as the point of attack of an axial force not producing a bending moment. The principal axes of a beam cross-section are defined as the axes around which a bending moment will neither produce an axial force nor flexural displacements in the other direction. The principle of virtual forces can by applied to the analysis of deformations in a beam. In the case of Timoshenko beam theory, both shear and bending deformation occurs, whereas only bending deformation is present in a Bernoulli-Euler beam. Plane beam elements have six degrees of freedom with three at either end, i.e. two displace- ments and one in-plane rotation. In the general case, the rotations and the axial displacements are coupled, but in a principle-axes description, they become uncoupled. Spatial beam elements have 12 degrees of freedom, that is three displacements and three rota- tions at either end. Generally, the displacements and rotations are coupled, but an uncoupling can be achieved by a proper choice of coordinate system.

Thus, a detailed description has been given of the lateral and flexural deformations in a beam. However, in this chapter only a brief introduction has been given to twist and torsion of a beam. A thorough explanation and analysis of these phenomena will be the focus of the next chapter.

DCE Lecture Notes No. 23 CHAPTER 2 Shear stresses in beams due to torsion and bending

In this chapter, a theoretical explanation is given for the shear forces in beams stemming from bending as well as torsion. In this regard, the coupling of torsion and bending is discussed, and the so-called shear centre is introduced. The derivations are confined to homogeneous torsion, or St. Venant torsion. Later, the strains and stresses provided in the case of non-homogeneous torsion, or Vlasov torsion, will be dealt with.

2.1 Introduction

When the beam is exposed to the loads per unit length qy and qz, the beam will generally deform with bending deformations {uy,θz} and {uz,θy}, respectively. Depending on the line of action of these loads, the bending deformations will be combined with a torsional rotation θx around the x-axis as illustrated in Fig. 2–1a and Fig. 2–1c. Under certain conditions, the bending of the beam is not associated with a torsional deformation. This happens if the loads per unit length qy and qz, the reaction forces at the ends of the beam as well as the shear forces Qy and Qz are acting through a special point S, known as the shear centre, as illustrated in Fig. 2–1b. When this is the case, torsion is caused solely by the moment load mx per unit length, which in part includes contributions from the translation of qy and qz to S. These torsional deformations take place without bending deformation as illustrated in Fig. 2–1d. Hence, bending and torsion can be analysed independently. The position of the shear centre dependson the geometry of the cross-section and is generally different from the position of the bending centre B. However, for double-symmetric cross- sections, the positions of the bending and torsion centres will coincide. The shear forces Qy and Qz as well as the torsional moment Mx bring about shear stresses σxy and σxz on the beam section. In what follows these will be determined independently for the two deformation mechanisms. Hence, qy and qz are presumed to be referred to the shear centre. The shear stresses caused by the torsional moment Mx are statically equivalent to the shear forces Qy = Qz =0. The position of the shear centre has no influence on the distribution of shear stresses in this case. Likewise, in the decoupled bending problem, in which the cross-section is exposed to the shear forces Qy and Qz, the position of the shear centre is determined from the requirement that the resulting shear stresses are statically equivalent to Qy and Qz, and produce the torsional moment Mx =0 around S.

— 45 — 46 Chapter 2 – Shear stresses in beams due to torsion and bending

Qy,qy Qy,qy

B B z z S x S x

(a) y (b) y

Mx, mx Qy,qy B B z z S x S x

(c) y (d) y Figure 2–1 Coupled and uncoupled bending and torsion. Coupling exists in cases (a) and (c), whereas cases (b) and (d) involve no coupling.

2.2 Homogeneous torsion (St. Venant torsion)

It is assumed that the torsional moment Mx and the incremental twist per unit length dθx/dx and the warping of the cross-sections remain unchanged along the beam. Then all cross-sections of the beam are exposed to the same distribution of the shear stresses σxy and σxz. For this reason, this case is referred to as homogeneous torsion. Since the solution of the problem was given by St. Venant (ref.), the case is also called St. Venant torsion. Inhomogeneous torsion refers to the case, where either Mx or the material properties vary along the beam. Then dθx/dx or the warping will vary as well. Figure 2–2a shows a cross-section of a cylindrical beam of the length l. The cross-sectional area is A. The curve along the outer periphery is denoted Γ0. The cross-section may have a number N of holes determined by the boundary curves Γj , j = 1, 2,...,N. At the boundary curves arc-length coordinates s0,s1,...,sN are defined. The arc-length coordinate s0 along Γ0 is orientated in the anti clock-wise direction, whereas s1,s2,...,sN , related to the interior boundaries Γ1, Γ2,..., ΓN , are orientated in the clock-wise direction. The outward directed unit

DCE Lecture Notes No. 23 2.2 Homogeneous torsion (St. Venant torsion) 47

n0

s0 n0 Γ0 s0 ΓN

s0

B zS z sj

nj Γj Mx

yS nj S Γj

sj sj

(a) y (b) Γ0

Figure 2–2 Cross-section with holes: (a) Interior and exterior edges; (b) definition of local (x,nj ,sj )-coordinate systems.

vector on a point of the exterior or interior boundaries Γj is denoted nj , j = 0, 1,...,N. The unit tangential vector to a boundary curve is denoted sj and is co-directional to the arc-length coordinate sj , see Fig. 2–2b. Thus, a local (x, nj ,sj )-coordinate system may be defined with the base unit vectors {i, nj, sj }. The indicated orientation of the exterior and interior arc-length coordinates sj , j = 0, 1,...,N, insures that the related (x, nj ,sj )-coordinate system forms a right-hand coordinate system. The beam material is assumed to be homogeneous, isotropic linear elastic with the shear modulus G. In homogeneous torsion, only shear stresses are present for which reason G is the only needed elasticity constant.

2.2.1 Basic assumptions

For convenience the index x is omitted on the twist θx (the rotation angle around the x-axis), i.e. θ ∼ θx. Figure 2–3 shows a differential beam element of the length dx. Both end-sections of the element are exposed to the torsional moment Mx, so the element is automatically in equilibrium. On the left and right end-sections the twists are θ and θ + dθ, respectively. The increment dθ may be written as

dθ dθ = dx. (2–1) dx

Since Mx and the material properties are the same in all cross-sections, dθ/dx must be constant along the beam. Further, the warping must be the same in all cross-sections, i.e. ux = ux(y,z). This implies that the warping in homogeneous torsion does not induce normal strains, i.e.

∂u ε = x =0. (2–2) xx ∂x

In turn this means that the normal stress becomes σxx = Eεxx = 0. Hence, only the shear stresses σxy and σxz are present on a cross-section in homogeneous torsion.

Elastic Beams in Three Dimensions 48 Chapter 2 – Shear stresses in beams due to torsion and bending

Mx

y l

z

Mx dx

θ Mx

θ + dθ x

dx Mx

Figure 2–3 Beam and differential beam element subjected to homogeneous torsion.

The only deformation measure of the problem is the twist gradient dθ/dx. Then, due to the linearity assumptions, the torsional moment Mx must depend linearly on dθ/dx. Further, σxy and σxz (and hence Mx) depend linearly on G. This implies the following relation: dθ M = GK . (2–3) x dx The proportionality constant K with the dimension [unit of length]4 is denoted the torsional constant. The determination of this constant is a part of the solution of the torsion problem.

2.2.2 Solution of the homogeneous torsion problem As for all beam theories, the shape of the cross-section is assumed to be preserved during the deformation. Then, the displacements in the (y,z)-plane are caused merely by the rotation θ around the shear centre S. The warping displacements ux must also be linearly dependent on the strain measure dθ/dx, correspondingto the last term in Eq. (1–13). This impliesthe displacement components dθ u = u (y,z)= ω(y,z) , u = −(z − z )θ, u = (y − y )θ. (2–4) x x dx y s z s Here ω(y,z) is the so-called warping function as discussed in Section 1.2.2, and its determination constitutes the basic part of the solution of the homogeneous torsion problem. Similarly to Eq. (1–14), it follows from Eq. (2–4) that the components of the strain tensor become: ∂u ∂u ∂u ε = x =0, ε = y =0, ε = z =0, (2–5a) xx ∂x yy ∂y zz ∂z

DCE Lecture Notes No. 23 2.2 Homogeneous torsion (St. Venant torsion) 49

1 ∂u ∂u 1 ε = y + z = (−θ + θ)=0, (2–5b) yz 2 ∂z ∂y 2   1 ∂u ∂u 1 ∂ω dθ ε = x + y = − (z − z ) , (2–5c) xy 2 ∂y ∂x 2 ∂y s dx     1 ∂u ∂u 1 ∂ω dθ ε = x + z = + (y − y ) . (2–5d) xz 2 ∂z ∂x 2 ∂z s dx     As seen only εxy and εxz are non-vanishing. Correspondingly, all components of the Cauchy stress tensor become equal to zero, save the shear stresses σxy and σxz. These are given as ∂ω dθ σ = σ (y,z)=2Gε = G − (z − z ) , (2–6a) xy xy xy ∂y s dx   ∂ω dθ σ = σ (y,z)=2Gε = G + (y − y ) . (2–6b) xz xz xz ∂z s dx   Ignoring the volume loads, the equilibrium equations read ∂σ ∂σ ∂σ xx + xy + xz =0, (2–7a) ∂x ∂y ∂z ∂σ ∂σ ∂σ xy + yy + yz =0, (2–7b) ∂x ∂y ∂z ∂σ ∂σ ∂σ xz + yz + zz =0. (2–7c) ∂x ∂y ∂z

With σxx = σyy = σzz = σyz = 0, and σxy and σxz only dependent on y and z, the two last equations are identically fulfilled, and the first equation reduces to ∂σ ∂σ xy + xz =0. (2–8) ∂y ∂z

Equation (2–8) may be formulated at a point on the boundary curve Γj , j = 0, 1,...,N, in the related local (x, nj ,sj)-coordinate system. Ignoring the index j, the non-trivial equilibrium equation then reads ∂σ ∂σ xn + xs =0, (2–9) ∂n ∂s where σxn and σxs denote the shear stresses along the local n- and s-axes. According to Cauchy’s boundary condition (ref.), σxn can be expressed in terms of the shear stress components σxy and σxz as

σxn = σxyny + σxznz. (2–10)

Here ny and nz denote the components of the unit normal vector n along the y- and z-axes. The symmetry of the stress tensor implies that σxn = σnx. Further, since the exterior and all interior surfaces are free of surface traditions, corresponding to σnx =0, it follows that σxn =0 (see Fig. 2–4). Then, Eq. (2–10) reduces to

σxyny + σxznz =0. (2–11)

Elastic Beams in Three Dimensions 50 Chapter 2 – Shear stresses in beams due to torsion and bending

σxn σnx = 0

n x

Figure 2–4 Shear stress in the normal direction at an exterior or interior boundary.

Finally, insertion of Eq. (2–6) into Eq. (2–11) provides the following boundary condition formu- lated in the warping function

∂ω ∂ω n + n − (z − z )n + (y − y )n =0 ⇒ ∂y y ∂z y s y s z ∂ω = (z − z )n − (y − y )n , (2–12) ∂n s y s z where ∂ω/∂n denotes the partial derivative of ω in the direction of the outward directed unit normal. Equation (2–12) must be fulfilled at the exterior and all interior boundaries. The partial differential for ω to be fulfilled in the interior A of the profile follows from insertion of Eq. (2–6) into the equilibrium equation (2–8), leading to

∂2ω ∂2ω + =0. (2–13) ∂y2 ∂z2

If a solution to Eq. (2–13) with the boundary conditions (2–12) is obtained, the shear stresses are subsequently determined from Eq. (2–10). Equation (2–13) is Laplace’s differential equation, and the boundary conditions Eq. (2–13) are classified as the so-called Neumann boundary con- ditions. Notice that the solution to Eqs. (2–12) and (2–13) is not unique. Actually, if ω(y,z) is a solution, then ω(y,z)+ω0 will be a solution as well, where ω0 is an arbitrary constant. Since the shear stresses are determined by partial differentiation of the of the warping function, all these solutions lead to the same stresses. The boundary value problem for the warping function has been summarised in Box 2.1.

DCE Lecture Notes No. 23 2.2 Homogeneous torsion (St. Venant torsion) 51

Box 2.1 Boundary value problem for the warping function The differential equation, representing the non-trivial equation of equilibrium, reads:

∂2ω ∂2ω + = 0, (y,z) A. (2–14a) ∂y2 ∂z2 ∈ The Neumann boundary condition, representing the relevant Cauchy boundary condition, reads:

∂ω =(z zs)ny (y ys)nz, (y,z) Γ0 Γ1 ΓN . (2–14b) ∂n − − − ∈ ∪ ∪···∪

An alternative formulation for the solution of the problem can be obtained by the introduction of the so-called Prandtl’s stress function S with the defining properties ∂S ∂S σ = , σ = − . (2–15) xy ∂z xz ∂y Upon insertion of Eq. (2–15) into Eq. (2–9), the equilibrium equation is seen to be identical fulfilled, i.e. ∂σ ∂σ ∂2S ∂2S xy + xz = − ≡ 0. (2–16) ∂y ∂z ∂y∂z ∂z∂y From Eq. (2–6) follows ∂σ ∂σ ∂2ω dθ ∂2ω dθ dθ xy − xz = − 1 G − +1 G = −2G . (2–17) ∂z ∂y ∂z∂y dx ∂y∂z dx dx     Then, the differential equation for S is obtained by insertion of Eq. (2–15) on the left-hand side of Eq. (2–17), leading to ∂2S ∂2S dθ + = −2G . (2–18) ∂y2 ∂z2 dx Equation (2–18) is a compatibility condition for S in order that the kinematical conditions (2–6) are fulfilled. The boundary condition for S follow upon insertion of Eq. (2–15) into Eq. (2–11), i.e. ∂S ∂S n − n =0 (2–19) ∂z y ∂y z T The tangential unit vector is given as s = [sy,sz] = [−nz,ny], cf. Fig. 2–2b, where {ny,nz} denotes the Cartesian componentsof the outward directed unit normalvector at any of the bound- ary curves Γj , j =0, 1,...,N, cf. Fig. 2–2b. Then, Eq. (2–19) may be written as ∂S ∂S ∂S s + s = =0. (2–20) ∂y y ∂z z ∂s where ∂S/∂s denotes the directional derivative of S in the direction of the tangential unit vector s. Equation (2–20) implies that S is constant along the exterior and the interior boundary curves, i.e.

S = Sj , j =0, 1,...,N. (2–21)

Elastic Beams in Three Dimensions 52 Chapter 2 – Shear stresses in beams due to torsion and bending

Equation (2–18) is a Poisson differential equation (inhomogeneous Laplace equation), and the boundary conditions Eq. (2–21) are classified as the so-called Dirichlet boundary conditions. In principle, the solution to the indicated boundary value problem is unique. The problem is that the constant values Sj of the stress function along the boundary curves are unknown. The determination of these is a part of the problem. For profiles with interior holes this is only possible by the introduction of additional geometric conditions. The boundary value problem for the Prandtl stress function has been summarised in Box 2.2. The formulation in terms of Prandtl’s stress function is especially useful in relation to homogeneous torsion of thin-walled profiles and will be utilised in a number of examples below.

Box 2.2 Boundary value problem for the Prandtl stress function The equilibrium equation is automatically fulfilled. The compatibility condition is represented by the following differential equation:

∂2S ∂2S dθ + = 2G , (x,y) A. (2–22a) ∂y2 ∂z2 − dx ∈ The Dirichlet boundary condition, representing the Cauchy boundary condition, reads:

S = Sj , j = 0, 1,...,N, (x,y) Γ0 Γ1 ΓN . (2–22b) ∈ ∪ ∪···∪

The shear stresses σxy and σxz must be statically equivalent to the shear forces Qy = Qz = T 0 and the torsional moment Mx. Application of Gauss’s theorem on the vector field v = [vy, vz]=[0,S] provides,

∂0 ∂S N N Q = σ dA = + dA = (0 · dz − S · dy)= − S dy, (2–23) y xy ∂y ∂z j A A j=0 Γj j=0 Γj Z Z   X I X I where the circulation is taken anticlockwise along Γ0 and clockwise along Γj , j = 1, · · · ,N. Further it has been used that Sj is constant along the boundary curve, and hence may be trans- ferred outside the circulation integral. Now, dy = 0. Hence, it follows that any solution Γj to the boundary value problem defined by Eqs. (2–22a) and (2–22b) automatically provides a T H solution fulfilling Qy =0. Using v = [S, 0] it can in the same way be shown that

∂S ∂0 Q = σ dA = − + dA ⇒ z xz ∂y ∂z ZA ZA   N N Qz = − (S · dz − 0 · dy)= − Sj dz =0. (2–24) j=0 Γj j=0 Γj X I X I

The torsional moment Mx must be statically equivalent to the moment of the shear stresses around S, i.e.

Mx = ((y − yS)σxz − (z − zS)σxy) dA = (yσxz − zσxy)dA, (2–25) ZA ZA DCE Lecture Notes No. 23 2.2 Homogeneous torsion (St. Venant torsion) 53

z B z S z

y Mx σxz dA yS σxy S

y

Figure 2–5 Static equivalence of torsional moment to shear stresses.

where it has been used that A σxydA = σxzdA =0. Next, insertion of Eq. (2–15) provides ∂S ∂SR R ∂ ∂ M = − y + z dA = − (yS)+ (zS) dA +2 SdA. (2–26) x ∂y ∂z ∂y ∂z ZA   ZA   ZA The divergence theorem with vT = [yS,zS] provides

∂ ∂ N N (yS)+ (zS) dA = (ySdz − zSdy)= S (ydz − zdy). (2–27) ∂y ∂z j A j=0 Γj j=0 Γj Z   X I X I

Let A0 and Aj denote the area inside the boundary curves Γ0 and Γj. Then, use of Eq. (1–4a) in Eq. (2–27) provides

N ∂ ∂ (yS)+ (zS) dA =2A S − 2A S . (2–28) ∂y ∂z 0 0 j j A j=1 Z   X The negative sign of the last term is because the circulation on the interior boundaries is taken clockwise, cf. the discussion subsequent to Eq. (1–4a). Insertion of Eq. (2–28) gives the follow- ing final result:

N Mx = −2A0S0 +2 SdA +2 Aj Sj . (2–29) A j=1 Z X The shear stresses remain unchanged if an arbitrary constant is added to S. Then, without re- striction one can choose S0 =0, which is assumed in what follows. Let the domain of definition for S(y,z) be extendedto the interior of the holes, where S(y,z) is given the same value Sj as the boundary value along the holes. With S0 = 0, Eq. (2–29) determines Mx as twice the volume below S(y,z). Further, with S(x, y) determined along with

Elastic Beams in Three Dimensions 54 Chapter 2 – Shear stresses in beams due to torsion and bending

Γ0,S0 = 0

B z

Γ1

y S(0,z)

S1 z Figure 2–6 Variation of Prandtl’s stress function over a cross-section with a hole.

the boundary values Sj , j =1, 2,...,N, the torsional constant K can next be determined upon comparison of Eqs. (2–3) and (2–29). This is illustrated by examples below. It is remarkable that no reference is made to the position of the shear centre, neither in the boundary value problem (2–22), for the Prandtl stress function, nor in the expression (2–25) for the torsional moment. In contrast, the coordinates of the shear centre enter the boundary value problem (2–14) for the warping function. Defining the same homogeneous torsion problem, the warping function and Prandtl’s stress function cannot be independent function. The relation follows from a comparison of Eqs. (2–6) and (2–15):

∂S ∂ω dθ = − (z − z ) G , (2–30a) ∂z ∂y s dx  

∂S ∂ω dθ − = + (y − y ) G , (2–30b) ∂y ∂z s dx   where it is recalled that θ ∼ θx.

Example 2.1 Homogeneous torsion of infinitely long rectangular cross-section Figure A shows an infinitely long rectangular cross-section with the thickness t exposed to a torsional moment Mx. The torsional moment is carried by shear stresses uniformly distributed in the y-direction. Then, S = S(z) is independent of y, and the boundary value problem (2–22) reduces to

d2S dθ = 2G , S( t/2) = S(t/2) = 0 (a) dz2 − dx − with the solution 1 dθ S(z)= (4z2 t2)G . (b) − 4 − dx (continued)

DCE Lecture Notes No. 23 2.2 Homogeneous torsion (St. Venant torsion) 55

The shear stresses follow from Eq. (2–15):

∂S dθ σxy = = 2zG , σxz = 0. (c) ∂z − dx The shear stresses are linearly distributed in the thickness direction, and has been illustrated in Fig. A.

t

σxs = σxy

dθ τ = G dx t dMx

dy z dy z

y y Figure A Torsion of a infinitely long rectangular cross-section: Distribution of shear stresses (left); torsional moment on differential cross-sectional segment (right).

Due to the independence of the shear stresses on y, the torsional problem can be analysed by merely considering a differential cross-sectional segment of the length dy exposed to the torsional moment dMx, see Fig. A. The increment dMx is related to the stress function by Eq. (2–28), i.e.

t/2 d/2 1 dθ 2 2 1 3 dθ dMx = 2 S(z)dzdy = dyG , (4z t )dz = t dyG . (d) − 2 dx − − 3 dx Z t/2 Z t/2

At the same time dMx = GdKdθ/dx, where dK denotes the torsional constant related to the differential segment. As seen from Eq. (d), this is given as 1 dK = d3dy. (e) 3 The value dK of the torsional constant for the differential segment will be applied below. 

Example 2.2 Homogeneous torsion of a solid ellipsoidal cross-section Figure A shows an ellipsoidal cross-section without holes with semi-axes a and b, exposed to a torsional moment Mx. At first, it is verified that the Prandtl’s stress-function of this problem is given as

a2b2 y2 z2 dθ S(y,z)= + 1 G . (a) − a2 + b2 a2 b2 − dx   (continued)

Elastic Beams in Three Dimensions 56 Chapter 2 – Shear stresses in beams due to torsion and bending

Γ0

b z

a Mx

y Figure A Ellipsoidal cross-section.

The boundary curve Γ0 is described by the ellipsis

y2 z2 + = 1. (b) a2 b2

Hence, S(x,y) = 0 for (y,z) Γ0, so the boundary condition (2–22b) is fulfilled by Eq. (a). The ∈ Laplacian of s(y,z) becomes

∂2S ∂2S a2b2 2 2 dθ dθ + = + G = 2G . (c) ∂y2 ∂z2 − a2 + b2 a2 b2 dx − dx   Then, also the differential equation (2–22a) is fulfilled, from which is concluded that Eq. (a) is indeed the solution to the homogeneous torsion problem. From Eq. (2–29) follows that

a2b2 dθ y2 z2 a3b3 dθ Mx = 2 SdA = 2 G + 1 dA = π G , (d) − a2 + b2 dx a2 b2 − a2 + b2 dx ZA ZA   where the following result has been utilised:

y2 z2 π + 1 dA = ab. (e) a2 b2 − − 2 ZA   From Eq. (2–3) and Eq. (d) follows that the torsional constant for an ellipsoidal cross-section becomes:

a3b3 K = π . (f) a2 + b2 Gdθ/dx follows from Eq. (d), i.e.

dθ 1 a2 + b2 G = M . (g) dx π a3b3 x Then, the shear stresses become, cf. Eqs. (2–15) and (a):

2 2 2 2 ∂S 2z a b 1 a + b 2 Mx σxy = = Mx = z, (h) ∂z − b2 a2 + b2 · π a3b3 − π ab3

2 2 2 2 ∂S 2y a b 1 a + b 2 Mx σxz = = Mx = y. (i) − ∂y a2 a2 + b2 · π a3b3 π a3b (continued)

DCE Lecture Notes No. 23 2.2 Homogeneous torsion (St. Venant torsion) 57

The warping of the cross-section is given by Eq. (2–4). Due to the symmetry of the cross-section, it is observed that ys = zs = 0, i.e. the shear centre coincides with the bending centre. Then, the warping function is determined from, cf. Eq. (2–30):

∂ω 1 ∂S 2a2 b2 a2 = + z = z + z = − z, (j) ∂y Gdθ/dx ∂z − a2 + b2 b2 + a2

∂ω 1 ∂S 2b2 b2 a2 = y = y y = − y. (k) ∂z − Gdθ/dx ∂y − a2 + b2 − b2 + a2 The solution to Eqs. (j) and (k) is given as

b2 a2 ω(y,z)= − yz. (l) b2 + a2 It is left as an exercise to prove that Eq. (l) fulfils the boundary value problem (2–14). The warping follows from Eq. (2–4), Eq. (g) and Eq. (l)

2 2 2 2 2 2 dθ b a 1 a + b Mx 1 b a Mx ux(y,z)= ω(y,z) = − yz − yz . (m) dx a2 + b2 · π a3b3 G π a3b3 G The solution (l) has been chosen so that the warping from torsion provided by Eq. (m) is zero at the bending centre. This will generally be presumed in what follows. Then, the displacement of the bending centre in the x-direction is caused entirely by the axial force N. Finally, it is noted that, especially for a circular profile with a = b, the warping vanishes everywhere on the profile. 

2.2.3 Homogeneous torsion of open thin-walled cross-sections

Figure 2–7 shows an open cross-section of a cylindrical beam. An arc-length coordinate s is defined along the midpoints of the profile wall, where s = 0 is chosen at one of the free ends, and s = L at the other free. Further, L specifies the total length of the profile wall, and the wall thickness at arc-length coordinate s is denoted t(s).

τ s s = L σxs

Mx, mx ds n S

t(s)

s s = 0

Figure 2–7 Open thin-walled cross-section of cylindrical beam exposed to homogeneous torsion.

Elastic Beams in Three Dimensions 58 Chapter 2 – Shear stresses in beams due to torsion and bending

For thin-walled cross-sections it is assumed that t(s) ≪ L. Then, the profile can be consid- ered as built-up of differential rectangles of the length ds, similar to those considered in Exam- 1 3 ple 2.2. Each has the torsional constant dK = 3 t (s)ds, cf. Eq. (e) in Example 2.2. Hence, the torsional constant for the whole profile is given as

1 K = t3(s)ds, (2–33) 3 ZL where the index L indicates that the line integral is extended over the whole length of the profile measured along the profile wall.

The shear stresses are specified in a local (x,n,s)-coordinate system as shown in Fig. 2–7. These coordinates follow from Eq. (c) upon replacing y with s and z with w. Then,

dθ σ =2nG , σ =0. (2–34) xs dx xn

Finally, using Gdθ/dx = Mx/K, the maximum shear stresses for n = t(s)/2 becomes

dθ M τ = G t(s)= x t(s). (2–35) dx K

The computation of the shear stress in a U-profile is considered in the example below.

Example 2.3 Homogeneous torsion of a U-profile

Figure A shows a U-profile exposed to homogeneous torsion from the torsional moment Mx. With the thin-wall approximation t a, the position of the bending centre is as shown in the figure. Further, the ≪ cross-sectional area and the bending moments of inertia around the y- and z-axes become 64 26 A = 10at, I = a3t, I = a3t, (a) y 15 z 3 where the single indices y and z indicate that the corresponding axes are principal axes. The cross- sectional constants given in Eq. (a) have been calculated for later use. With reference to Eq. (2–33), the torsional constant becomes: 2 1 34 K = 2a(2t)3 + 2at3 = a3t. (b) 3 3 5 (continued)

DCE Lecture Notes No. 23 2.2 Homogeneous torsion (St. Venant torsion) 59

2a 2τ 4 a 5 2t

a

Mx Mx B z

t a τ 2t τ

y 2τ

2τ Figure A Homogeneous torsion of U-profile: Dimensions (left) and shear stresses (right).

The distribution of shear stresses follows from Eq. (2–35). The maximum shear becomes τ and 2τ, respectively, where τ is given as M 3 M τ = x t = x . (c) K 34 at2 Hence, in the present case, the shear stresses in the flanges are higher than those in the web. 

2.2.4 Homogeneous torsion of closed thin-walled cross-sections

The boundary value problem for the warping function ω(y,z), defined by Eq. (2–14), has a unique solution (save an arbitrary additive constant) no matter if the profile has an interior hole or not. Although an analytical solution is seldom obtainable, a numerical solution can always be achieved by a discretization of the Laplace operator by a finite-difference or a finite-element approach. In contrast to this, the boundary value problem defined by Eq. (2–22) for Prandtl’s stress function cannot immediately be solved, because the boundary values Sj , j = 1, 2,...,N, are not known. The determination of these values requires the formulation of additional geomet- rical conditions which express that the warping function ω(y,z) shall be continuous along the

Elastic Beams in Three Dimensions 60 Chapter 2 – Shear stresses in beams due to torsion and bending

boundary curves Γj . This may be formulated as ∂ω ∂ω dω = dy + dz =0, j =1, 2,...,N. (2–37) ∂y ∂z IΓj IΓj   Equation (2–37) can be expressed in the stress function by use of Eq. (2–30):

∂S ∂S dθ dy − dz + G ((z − z )dy − (y − y )dz)=0. (2–38) ∂z ∂y dx s s IΓj   IΓj It can be shown that the integrand of the first integral can be rewritten in terms of the coordinates s and n by the substitution ∂S ∂S ∂S dy − dz = − ds. (2–39) ∂z ∂y ∂n The second integral in Eq. (2–38) may be recast as

((z − zs)dy − (y − ys)dz)= − (ydz − zdy)=2Aj . (2–40) IΓj IΓj The change of sign in Eq. (2–40) is because all circulations are taken clockwise. Further, it has utilised that y dz = y dz = 0 and z dy = 0. Insertion of Eqs. (2–39) and (2–40) Γj s s Γj Γj s into Eq. (2–38) provides the following form for the geometrical conditions: H H H ∂S dθ ds =2A G . (2–41) ∂n j dx IΓj Only thin-walled cross-sections are considered. At first a cross-section with a single cell is considered as shown in Fig. 2–8. An arc-length coordinate is introduced, orientated in the clockwise direction. Correspondingly, a local (x,n,s)-coordinate system is defined at each point of the boundary curve with the n-axis orientated inward into the cavity, whereas the s-axis is tangential to the boundary curve and unidirectional to the arc-length coordinate. Then, the shear stresses along the n- and s-directions become, cf. Eq. (2–15), ∂S ∂S σ = , σ = − . (2–42) xn ∂s xs ∂n

Here, S is constant along the exterior and interior boundary curve of the wall, given as S = S0 = 0 and S = S1, respectively. Hence, σxn =0 along these boundaries. If the thickness t(s) of the wall is small compared to a characteristic diameter of the profile, it then follows from continuity that σxn is ignorable in the interior of the wall, i.e.

σxn ≃ 0. (2–43)

The stress function decreases from S = S1 at the inner side of the wall to S = S0 = 0 at the outer side. If t(s) is small, the variation of S(n,s) must vary approximately linearly over the wall thickness t(s), see Fig. 2–8. This implies the following approximation ∂S S − S S σ = − ≃− 1 0 = − 1 . (2–44) xs ∂n t(s) t(s)

DCE Lecture Notes No. 23 2.2 Homogeneous torsion (St. Venant torsion) 61

σxn σxs s s n s n 0 B z

Mx S

t

y S(0,z)

S1 z

Figure 2–8 Closed thin-walled cross section of a cylindrical beam exposed to homogeneous torsion.

Then, the geometrical condition Eq. (2–41) may be written as

ds dθ S =2A G , (2–45) 1 t(s) 1 dx IΓ1 where A1 is area of the cavity. The torsional moment is given by Eqs. (2–3) and (2–29):

dθ M = GK =2A S (2–46) x dx 1 1 Combining Eq. (2–45) and Eq. (2–46) provides the following result for the torsional constant:

S 4A2 ds K =2A 1 = 1 , J = . (2–47) 1 Gdθ/dx J t(s) IΓ1 Equation (2–47) is known as Bredt’s formula. The shear stresses follow from Eqs. (2–44) and (2–46):

S1 Mx σxs(s)= − = − . (2–48) t(s) 2A1t(s)

Notice, that the shear stress σxs(s) is uniformly distributed over the wall-thickness as shown in Fig. 2–8. This is in contrast to an open section, where a linear variation was obtained, as given by Eq. (2–34) .

Elastic Beams in Three Dimensions 62 Chapter 2 – Shear stresses in beams due to torsion and bending

Example 2.4 Homogeneous torsion of a closed thin-walled cross-section Figure A shows the cross-section of a beam with a single cell exposed to homogeneous torsion. The thin-wall approximation t a is assumed to be valid. ≪

1 t 6 τ s

t n a 1 6 τ t 1 Mx Mx 12 τ 1 6 τ 2t 2t a 1 12 τ 2t

1 12 τ a a Figure A Homogeneous torsion of a closed thin-walled cross-section: Geometry (left) and distribution of shear 2 stresses with τ = Mx/(a t) (right).

The area of the cavity and the line integral entering Eqs. (2–45) and (2–47) become

ds 4a 4a 6a A = 3a2, = + = . (a) 1 t(s) t 2t t IΓ1 Then, the torsional constant becomes, cf. Eq. (2–47),

4(3a2)2 K = = 6a3t. (b) 6a/t The shear stresses follow from Eq. (2–48):

Mx σxs = . (c) − 6a2t(s)

2 The distribution has been shown in Fig. A along with the direction of action. Note that τ = Mx/(a t) has been introduced as a normalisation quantity. 

Example 2.5 Comparison of homogeneous torsion of open and closed profiles Figure A shows two cylindrical beams, both with a cylindrical thin-walled cross-section with the side length a and the thickness t. In one case, the cross-section is closed, whereas the cross-section in the other case has been made open by a cut along a developer. The rotation gradient dθ/dx and the shear stresses in two beams due to homogeneous torsion are compared below. (continued)

DCE Lecture Notes No. 23 2.2 Homogeneous torsion (St. Venant torsion) 63

τo

a

a Mx,o,θo

a

τc a Mx,c,θc

Figure A Homogeneous torsion of open and closed cross-sections.

According to Eqs. (2–33) and (2–47), the torsional constants Ko and Kc for the beams with open and closed cross-sections are given as

1 4 K = t3(s)ds = at3, (a) o 3 3 Z 4A2 4a4 K = = = a3t. (b) c ds 4a/t t If the twoH beams are exposed to the same torsional moment Mx, it follows from Eq. (2–3) that the rotation gradients dθo/dx and dθc/dx for the open and closed sections are related as

dθ /dx K 4 a2 o c (c) = = 2 . dθc/dx Ko 3 t

Next, assume that both cross-sections are fully stressed, i.e. the maximum shear stress τo = τc is in both cases equal to the shear stress τy. The corresponding torsional moments that can be carried by the profiles are denoted Mx,o and Mx,c, respectively. With reference to Eqs. (2–35) and (2–48), these torsional moment are related as M K τ /t 2 a x,o o y (d) = 2 = . Mx,c 2a tτy 3 t

Hence, whereas the deformations of the two beams depend on the fraction a2/t2, the torsional moment which can be carried depends on a/t. In conclusion, open sections are extremely ill-conditioned to carry torsional moments in homogeneous torsion compared to closed sections. However, in many cases another mechanics is active, which involves inhomogeneous torsion. This substantially increases the torsional properties of open sections. 

Elastic Beams in Three Dimensions 64 Chapter 2 – Shear stresses in beams due to torsion and bending

As illustrated by Example 2.5, closed thin-walled cross-sections, e.g. tubes and box gird- ers, have a much higher torsional strength and stiffness than open thin-walled cross-section. A mechanical explanation for the highly increased stiffness obtained for a closed cross-section compared with the open cross-section is the fact that warping is hindered by the fact that the displacements in the axial direction must be continuous along the s-direction, i.e. along the wall. Especially, for a circular tube no warping is achieved in homogeneoustorsion, making this profile particularly useful for structural elements with the primary function of carrying torsional loads. Next, consider a thin-walled cellular cross-section with a total of N cavities as illustrated in Fig. 2–9. Each cavity has the cross-sectional area Aj , j = 1, 2,...,N, and the boundary of the cell is denoted Γj . In each cell, a local (x, nj ,sj)-coordinate system is defined with the nj -axis oriented into the cavity as illustrated in Fig. 2–9. The local arc-length coordinate sj is defined along the cell boundary in the clockwise direction, and the wall-thickness is t(sj ).

− SN σxs(s)= t(s) Γ0

n

N s Γj sN x sj B j z n 1 s1 Mx n n s sk k s − s − Sj Sk σxs(s)= t(s)

y

Figure 2–9 Cellular thin-walled cross section of a cylindrical beam exposed to homogeneous torsion.

As discussed in Subsection 2.2.2 and illustrated in Fig. 2–6, Prandtl’s stress function S is constant along each interior boundary. Hence, on the cell boundary Γj , S has the constant value Sj , j =1, 2,...,N. However, different values of Prandtl’s stress function are generally present on either side of a common wall between two adjacent cells, j and k. Because of the thin-wall assumption the variation of S over the wall thickness must be approximately linear. Hence, with reference to Eq. (2–42),

∂S Sj − Sk σxs = − ≃− . (2–51) ∂nj t(sj )

Thus, σxs is uniformly distributed over the wall-thickness as illustrated in Fig. 2–9. The shear stress component σxn vanishes along Γj as well as Γk. It then follows from continuity that σxn is ignorable in the interior of the wall as well, i.e.

σxn ≃ 0. (2–52)

DCE Lecture Notes No. 23 2.2 Homogeneous torsion (St. Venant torsion) 65

Insertion of Eq. (2–51) into Eq. (2–41) provides:

dsj dθ (Sj − Sk) =2Aj G , j =1, 2,...,N (2–53) Γjk t(sj ) dx Xk Z where the summation on the left-hand side is extended over all cavities adjacent to cell j. Notice that Sk = 0 at the part of cell j adjacent to the outer periphery. Then, Eq. (2–53) represents N coupled linear equations for the determination of Sj , j =1, 2,...,N. Subsequently the shear stress σxs in all interior and exterior walls is determined from Eq. (2– 51). According to Eq. (2–29), the torsional moment is given as

N Mx =2 Aj Sj , (2–54) j=1 X where the contribution A S dA can be ignored due to the thin-wall approximation. Then, the torsional constant follows from Eq. (2–3): R N M S K = x =2 A j . (2–55) Gdθ/dx j Gdθ/dx j=1 X

Since Sj turns out to be proportional to Gdθ/dx, the right-hand side of Eq. (2–55) will be independent of this quantity.

Example 2.6 Homogeneous torsion of a rectangular thin-wall profile with two cells Figure A shows a rectangular thin-walled cross-section with two cells exposed to homogeneous torsion from the torsional moment Mx. The wall-thickness is everywhere t a. ≪

8 1 9 σsx = 52 τ σsx = 52 τ σsx = 52 τ

a t 1 Mx 2

n n

s1 s s2 s

a 2a Figure A Homogeneous torsion of a rectangular thin-walled cross-section with two cells. Definition of local coordi- 2 nate systems and distribution of shear stresses with τ = Mx/(a t). (continued)

Elastic Beams in Three Dimensions 66 Chapter 2 – Shear stresses in beams due to torsion and bending

For the two cells, Eq. (2–53) takes the form:

a a a a 2 dθ (S1 0) +(S1 0) +(S1 0) +(S1 S2) = 2a G , (a) − t − t − t − t dx

2a a a 2a 2 dθ (S2 0) +(S2 S1) +(S2 0) +(S2 0) = 2 2a G . (b) − t − t − t − t · dx dθ dθ 4S1 S2 = 2atG , S1 + 6S2 = 4atG . (c) − dx − dx The solution of Eq. (c) reads

16 dθ 18 dθ S = atG , S = atG , (d) 1 23 dx 2 23 dx and by Eq. (2–54), the torsional moment is derived as

2 2 104 3 dθ dθ 23 Mx Mx = 2a S1 + 22a S2 = a tG G = . (e) 23 dx ⇒ dx 104 a3t

Then, S1 and S2 may be written as 8 M 9 M S = x , S = x . (f) 1 52 a2 2 52 a2

The shear stresses σxs follow from Eq. (2–51) and Eq. (f). The distribution has been shown in Fig. A 2 with their direction of action. The quantity τ = Mx/(a t) represents a normalisation quantity for the shear stresses. The torsional constant follows from Eq. (2–55) and Eq. (d) 16 18 104 K = 2 a2 at + 2 2a2 at = a3t. (g) · · 23 · · 23 23 If the interior wall is skipped, the corresponding quantities become:

1 Mx 1 Mx 9 3 S1 = , σsx = , K = a t. (h) 6 a2 − 6 a2t 2

1 Hence, the interior wall increases the shear stress in the exterior cell walls of the right cell from 6 τ to 9 9 3 104 3  52 τ (1.3%). The torsional constant is increased merely from K = 2 a t to K = 23 a t (2.2%).

As indicated by Example 2.6, only a small increase of the torsional stiffness is achieved by the inclusion of internal walls in a cross-section. Hence, from an engineering point of view, the benefits of applying structural members with a cellular cross-sections are insignificant in relation to torsion. Since the advantages regarding flexural deformations are also limited, and profiles with two or more cavities are not easily manufactured, it may be concluded that open cross-sections are generally preferred for beams loaded in unidirectional bending. Likewise, closed cross-sections with a single cavity (pipes or box girders) are preferable when the beam is primarily subjected to torsion and/or bending in two directions. Finally it is noted that the solution method based on Prandtl’s stress-function will only be used in relation to hand calculation for thin-walled cross-sections with at most two or three cells. Otherwise, for thick-walled sections or multi-cell problems, the homogeneous torsion problem will be solved numerically by a finite-element approach or another spatial discretization method based on the Neumann boundary problem specified in Box 2.1.

DCE Lecture Notes No. 23 2.3 Shear stresses from bending 67

2.3 Shear stresses from bending

In the previous section, the shear stresses occurring due to homogeneous torsion of a beam have been analysed. A second contribution to the shear stresses stem from bending or flexural deformations of a beam and with a proper choice of coordinates, the two contributions decouple. Figure 2–10 shows a cross-section exposed to bending without torsion. Correspondingly, the loads per unit length qy and qz as well as the shear forces Qy and Qz have been referred to the shear centre S, the position of which is discussed in this section.

s0 ΓN Γ0 n sN n σ σxz xx x s zS z Mz σxy B s Γ1 s 1 n S Qz qz s yS

y Qy

qy My

Figure 2–10 Cross-section exposed to bending.

The cross-section has the external boundary Γ0 and may have a number of cavities, N, bounded by the interior boundary curves Γj , j = 1, 2,...,N. Again, a local arc-length co- ordinate sj is defined along each boundary, orientated clockwise for all interior boundaries, j = 1, 2,...,N, whereas the arc-length coordinate s0 along the outer boundary curve is orien- tated anti clockwise. At any point along the outer and inner boundary curves, local right-handed (x, nj ,sj )-coordinate systems are defined as shown in Fig. 2–10. The (x,y,z)-coordinate sys- tem with origin at the bending centre B is assumed to be a principal-axes coordinate system. On the cross-section, the normal stress σxx as well as the shear stresses σxy and σxz are acting. With reference to Eq. (2–7), these stresses fulfil the equilibrium equation ∂σ ∂σ ∂σ xx + xy + xz =0, (2–57) ∂x ∂y ∂z where σxx is determined from Navier’s formula (1–75), i.e.

N(x) My(x) Mz(x) σxx(x,y,z)= + z − y. (2–58) A Iy Iz It then follows that

∂σxx dN 1 dMy z dMz y qx y z = + − = − + (Qy + mz) + (Qz − my) , (2–59) ∂x dx A dx Iy dx Iz A Iz Iy

Elastic Beams in Three Dimensions 68 Chapter 2 – Shear stresses in beams due to torsion and bending where the following equilibrium equations have been used in the statement: dN dM dM + q =0, y − Q + m =0, z + Q + m =0. (2–60) dx x dx z y dx y z A stress function T = T (x,y,z) is introduced with the defining properties ∂T ∂T σ = , σ = . (2–61) xy ∂y xz ∂z Hence, T is defined differently from the somewhat similar Prandtl’s stress function, cf. Eq. (2– 15). Insertion of Eqs. (2–59) and (2–61) into Eq. (2–57) provides the following Poisson partial differential equation for T : ∂2T ∂2T q y z x (2–62) 2 + 2 = − (Qy + mz) − (Qz − my) . ∂y ∂z A Iz Iy With reference to Eqs. (2–10) and (2–11), the boundary conditions read

σxn = σxxnx + σxyny + σxznz = σxyny + σxznz =0, (2–63) where nx =0 because the beam is cylindrical. Insertion of Eq. (2–61) into this equation provides the following homogeneous Neumann boundary conditions to be fulfilled on all exterior and interior boundary curves: ∂T ∂T ∂T = n + n =0. (2–64) ∂n ∂y y ∂z z The boundary value problems defined by Eqs. (2–14) and (2–22) for the warping function and Prandtl’s stress function, respectively, are independent of x, i.e. the solution applies for all cross-sections of the beam. In contrast, the corresponding boundary value problem defined by Eqs. (2–62) and (2–64) for T = T (x,y,z) must be solved at each cross-section defined by x, where the shear stresses are determined. This is so, because qx, my, mz, Qy and Qz entering the right-hand side of Eq. (2–62) may vary along the beam. The solution to Eq. (2–62) with boundary conditions given by Eq. (2–64) is unique save an arbitrary function T0 = T0(x) which has no influence on the shear stresses. The method can be applied to thick-walled or thin-walled cross-sections with or without interior cavities. Normally the boundary value problem can only be solved numerically based on a discretization of the Laplace operator entering Eq. (2–62). The boundary value problem for the stress function T has been summarised in Box 2.3.

Box 2.3 Boundary value problem for the stress function determining shear stresses in bending For a given cross-section determined by the abscissa the stress-function T = T (x,y,z) is obtained from the Poisson partial differential equation

∂2T ∂2T q (x) y z x (2–65a) 2 + 2 = (Qy(x)+ mz(x)) (Qz(x) my(x)) , (y,z) A ∂y ∂z A − Iz − − Iy ∈ with the homogeneous Neumann boundary conditions

∂T = 0, (y,z) Γ0 Γ1 ΓN . (2–65b) ∂n ∈ ∪ ∪···∪

DCE Lecture Notes No. 23 2.3 Shear stresses from bending 69

The indicated method is not applicable for hand-calculations. For this purpose a method will be devised in the following sub-sections, which solves the problem for thin-walled open cross-sections and closed thin-walled sections with few cells.

2.3.1 Shear stresses in open thin-walled cross-sections Figure 2–12 shows the same open thin-walled cross-section as shown in Fig. 2–7, when ex- posed to homogeneous torsion. Now the shear stresses σxs and σxn defined in local (x,n,s)- coordinates are requested, caused by the shear forces Qy and Qz acting through the shear centre S. We shall return to the definition and determination of the shear centre later in this chapter.

σxs s s = L

n

S Q z x z Mz B Qy y

t(s) Ω(s) My

s s = 0

Figure 2–11 Cross-section exposed to bending.

1 The shear stress component σxn still vanishes at the surfaces n = ± 2 t(s), cf. Eq. (2–34). For continuity reasons, Eq. (2–43) remains valid in the interior of the wall, i.e.

σxn(x,n,s) ≃ 0. (2–66) Then, with reference to Eq. (2–57), the equilibriumequationof stress componentsin the (x,n,s)- coordinate system reduces to ∂σ ∂σ xx + xs ≃ 0. (2–67) ∂x ∂s

From Eq. (2–58) follows that σxx varies linearly over the cross-section. Thus it must be almost constant over the thin wall and can be replaced by its value at the midst of the wall. From Eq. (2–67) it then follows that σxs must also be approximately constant in the n-direction, i.e.

σxs(x,n,s) ≃ σxs(x, s). (2–68)

The constancy of the shear stress component σxs in the thickness direction has been illustrated in Fig. 2–12. This variation should be compared to Eq. (2–34) for homogeneous torsion of an open cross-section, where a linear variation with n is obtained as illustrated in Fig. 2–7.

Elastic Beams in Three Dimensions 70 Chapter 2 – Shear stresses in beams due to torsion and bending

ds

σxx σxs

H + dH, σsx + dσsx t(s) s dx

H, σsx

σxs + dσxs σ + dσ x xx xx

Figure 2–12 Differential element of a thin-walled cross-section.

A differential element with the side lengths dx and ds is cut free from the wall at the axial coordinate x and at the arc-length coordinate s, see Fig. 2–12. On the sections with the arc-length coordinates s and s + ds, the shear forces per unit length H and H + dH are acting, where

t/2 H(x, s)= σsx(x,n,s)dn ≃ σxs(x, s)t(s). (2–69) − Z t/2 Here it has been exploited that σsx = σxs, and Eq. (2–68) has been utilised. On the sections with the axial coordinates x and x + dx, the normal stresses σxx and σxx + dσxx are acting (see Fig. 2–12). Equilibrium in the x-direction then provides

(σxx + dσxx)t(s)ds − σxxt(s)ds + (H + dH)dx − Hdx =0 ⇒ ∂H ∂σ + t(s) xx =0, (2–70) ∂s ∂x where it has been utilised that σxx has an ignorable variation in the thickness direction. It has therefore been replaced with a constant value equal to the value present at the midst of the wall. Alternatively, Eq. (2–70) may be obtained simply by multiplication of Eq. (2–67) with t(s) and use of Eq. (2–69). With H0(x) representing an integration constant, integration of Eq. (2–70) provides the so- lution s ∂σ H(x, s)= H (x) − xx t(s)ds. (2–71) 0 ∂x Z0 Especially, in the open thin-walled section the boundary condition σsx =0 applies at the ends of the profile, corresponding to the arc-length coordinates s =0 and s = L, i.e. H(x, 0) = H(x, L)=0. (2–72)

Thus, for an open thin-walled cross-section, the integration constant in Eq. (2–71) is H0(x)=0. The partial derivative ∂σxx/∂x is given by Eq. (2–59). In what follows it is for ease assumed that qx = my =0. Equation (2–59) then reduces to ∂σ (x, s) Q (x) Q (x) xx = y y(s)+ z z(s), (2–73) ∂x Iz Iy

DCE Lecture Notes No. 23 2.3 Shear stresses from bending 71 where (y(s),z(s)) denotes the principal-axes coordinate of a position on the wall determined by the arc-length coordinate s. Then, insertion in Eq. (2–71) and use of Eq. (2–69) provide the following solution for σxs:

1 Qy(x) Qz σxs(x, s)= H(x, s)= − SΩz(s) − SΩy(s), (2–74a) t(s) t(s)Iz t(s)Iy where it has been utilised that H0(x)=0 for the open section, and

s s SΩy(s)= z(s)t(s)ds = zdA, SΩz(s)= y(s)t(s)ds = ydA. (2–74b) Z0 ZΩ Z0 ZΩ

The quantities SΩy(s) and SΩz(s) denote the statical moment around the y- and z-axes of the area segment Ω(s), shown with a dark grey signature in Fig. 2–12 and defined as

s Ω(s)= t(s)ds = dA. (2–75) Z0 ZΩ Equation (2–74) is known as Grashof’s formula. The formula is valid for Timoshenko as well as Bernoulli-Euler beams with a thin-walled open cross-section. In this context it is noted that Bernoulli-Euler beam theory is based on the kinematic constraint that plane cross-sections or- thogonal to the beam axis in the referential state remain plane and orthogonal to the deformed beam axis. In turn this implies that the angular strains γxy and γxz vanish, and hence that σxy = σxy = 0. Hence, the shear stresses in bending cannot be determined from the beam theory itself. Instead, these are determined from Eqs. (2–66) and (2–74) which are derived from static equations alone and, hence, are independent of any kinematic constraints. The shear strain caused by the shear stress σxs is given as 1 1 ε = σ = H(x, s). (2–76) xs 2G xs 2Gt(s)

The shear strain εxs implies a warping ux0(x, 0) of the cross-section, additional to the dis- placement in the x-direction caused by the axial force and the bending moments. The latter is described by the kinematic conditions defined by Eqs. (1–11a) and (1–15) for Bernoulli-Euler beam theory. Hence, the displacements of the cross-section relative to the principal-axes coordi- nate system can be written as

dw (x) dw (x) u (x, s)= w (x) − y y(s) − z z(s)+ u (x, s), (2–77a) x x dx dx 0 uy(x, s)= wy(x), (2–77b) uz(x, s)= wz(x). (2–77c)

The component us(x, s) of the displacement vector u(x, s) in the tangential s-direction in the local (x,n,s)-coordinate system shown in Fig. 2–12 is given as

T dy(s) T ds ux(x, s) dy(s) dz(s) us(x, s)= s u(x, s)= dz(s) = ux(x, s)+ uy(x, s), (2–78) uy(x, s) ds ds " ds #  

Elastic Beams in Three Dimensions 72 Chapter 2 – Shear stresses in beams due to torsion and bending where sT = sT (s) = [dy/ds, dz/ds] signifies the unit tangential vector. Then, the strain εxs(x, S) follows from Eqs. (2–77) and (2–78):

1 ∂u ∂u 1 dy dw dz(s) dw dy dw dz dw ∂u ε = s + x = y + z − y − z + x0 ⇒ xs 2 ∂x ∂s 2 ds dx ds dx ds dx ds dx ∂s     1 ∂u ε = x0 . (2–79) xs 2 ∂s

The warping ux0(x, s) is next determined by integration of Eq. (2–79):

s ux0(x, s)=2 εxs(x, s)ds + u0(x). (2–80) Z0

The arbitrary function u0(x) is adjusted, so that ux0 = 0 at the bending centre B. When εxs is determined from Eq. (2–76), the warping of the cross-section additional to displacements in the x-direction predicted by Bernoulli-Euler beam theory can be determined by Eq. (2–80).

Example 2.7 Shear stresses and warping due to bending of a rectangular cross-section

Figure A shows a rectangular beam of height h and thickness t exposed to a shear force Qy. The bending moment of inertia, the area segment Ω(s) and static moment SΩz(s) around the z-axis become

1 3 h s Iz = h t, Ω(s)= st, SΩz(s)= st . (a) 12 − 2 − 2   h where the arc-length parameter is defined from the upper edge of the profile, i.e. s = 2 + y. Then, σxs = σxy is determined from Eq. (2–74), that is

Q s s2 Q 3 y2 Q y y y (b) σxy = SΩz(s) = 6 2 = 1 4 2 . −tIz h − h ht 2 − h ht     h Equation (b) specifies a parabolic distribution over the cross-section, where σxy = 0 at the edges y = 2 3 Qy ± in agreement with the boundary conditions, and the maximum value 2 ht is achieved at the bending centre B at y = 0. Next, the warping is determined from Eqs. (2–76) and (2–80) together with Eq. (b):

s s 2 2 3 1 Qy s s Qy s s 1 ux0(x,s)= u0 + σxsds = u0 + 6 ds = 3 2 , (c) G Ght h − h2 Gt h2 − h3 − 2 Z0 Z0     h h where u0 is adjusted, so ux0 = 0 at the bending centre, i.e. at s = 2 . With s = 2 + y, the warping displacements become

3 Qy 2y 2y ux0(y)= 3 . (d) 4Gt h − h   ! Equation (d) specifies a cubic polynomial variation, leading to an S-shape of the warping function. The distribution of σxy(y) and ux0(y) has been shown in Fig. A. (continued)

DCE Lecture Notes No. 23 2.3 Shear stresses from bending 73

1 Qy t 2G t s Ω(s) h 2 3 Qy x z 2 th z z x x

B = S ux0

h Qy 2 σxy

1 Qy 2G t y y Figure A Rectangular cross-section subjected to bending: Area segment (left); shear stresses (centre); warping (right). 

Example 2.8 Shear stresses due to bending of a double-symmetric I-profile

Figure A (left) shows a double-symmetric cross-section exposed to the shear forces Qy and Qz. The wall thickness is everywhere t a, i.e. the thin-wall assumption applies. The cross-sectional area and the principal bending moments≪ of inertia become

2 1 3 1 3 a 1 3 7 3 A = 3at, Iy = 2 a t = a t, Iz = 2 at + a t = a t. (a) · 12 6 · 2 12 12   Due to the symmetry, the bending and shear centres are coinciding. n t s1 s s2

s a n s5 2 t B = S n Qz z s x

a Qy 2 n t s3 s4

a a s s 2 2 n y Figure A Double-symmetric I-profile: Geometry (left); definition of arc-length coordinates and local coordinate systems (right).

At first the shear stresses caused by Qz are considered. For each of the four branches and the web of the profile, arc-length coordinates sj and local right-handed (x, nj ,sj )-coordinate systems are defined as shown in Fig. A (right). At the free edge, the shear stresses are σxs = 0. Hence, the shear force per unit length H(x,sj ) vanishes at this point, and Eq. (2–71) becomes valid for each branch. (continued)

Elastic Beams in Three Dimensions 74 Chapter 2 – Shear stresses in beams due to torsion and bending

Q 1 a2t 3 y SΩy(s1) 8 σxs(s1) 4 at SΩy(s2) σxs(s2)

B = S B = S Qz Qz

SΩy(s4) 3 Qy σxs(s4) SΩy(s3) 1 a2t σxs(s3) 8 4 at Figure B Double-symmetric I-profile: Distribution of statical moment of the area segment Ω(sj ) around the y-axis (left); distribution of the shear stresses σxs(sj ) from Qz (right).

On the five branches, the area segments become Ω(sj )= sj t. These have been shown with a dark grey signature in Fig. A (right). For the flanges, the statical moments of Ω(sj ) around the y-axis become 1 1 SΩy(s1)= ts1(s1 a), SΩy(s2)= ts2(s2 a), (b) 2 − − 2 − 1 1 SΩy(s3)= ts3(s3 a), SΩy(s4)= ts4(s4 a). (c) 2 − − 2 − Further, SΩy = 0 along the web. The distribution has been shown in Fig. B (left).

Qy 1 a2t 3 4 7 at SΩz(s1) σxs(s1) SΩz(s2) σxs(s2)

Q B = S 1 a2t B = S 3 y 8 14 at

Qy Qy Q 1 a2t 6 y 2 7 at

SΩz(s4) σxs(s4) SΩz(s3) σxs(s3) Qy 1 a2t 3 4 7 at Figure C Double-symmetric I-profile: Distribution of statical moment of the area segment Ω(sj ) around the z-axis (left); distribution of the shear stresses σxs(sj ) from Qy (right). (continued)

DCE Lecture Notes No. 23 2.3 Shear stresses from bending 75

The shear stresses σxs(sj ) follow from Eq. (2–74): Q Q Q z z z (d) σxs(s1)= SΩy(s1) = 3 3 s1(a s1), σxs(s2)= 3 3 s2(a s2), − tIy a t − − a t −

Qz Qz σxs(s3) = 3 s3(a s3), σxs(s4)= 3 s4(a s4), σxs(s5) = 0. (e) a3t − − a3t − The distribution has been shown in Fig. B (right). The sign refers to the local x, nj ,sj )-coordinate system. Hence, a negative sign implies that the shear stress is acting in the negative sj -direction and, hence, is co-directional to Qz. Then, the shear stresses in the flanges are distributed parabolically in the same way as stresses in the rectangular cross-section considered in Example 2.7. Each flange carries the shear force Qy/2, and no shear force is carried by the web in the present case. Next, the shear stresses caused by Qy are analysed. The statical moment of the area segment Ω(sj ) around the z-axis becomes: 1 1 SΩz(s1)= ats1, SΩz(s2)= ats2, (f) − 2 − 2 1 1 a 1 SΩz(s3)= ats3, SΩz(s4)= ats4, SΩz(s5)= at ts5(a s5), (g) 2 2 − 2 − 2 − and by Eq. (2–74) the shear stresses σxs(sj ) are determined as Q 6 Q 6 Q y y y (h) σxs(s1)= SΩz(s1)= 2 s1, σxs(s2)= 2 s2, − tIz 7 a t 7 a t

6 Qy 6 Qy 6 2 2 Qy σxs(s3)= s3, σxs(s4)= s4, σxs(s5)= (a + s5a s ) . (i) − 7 a2t − 7 a2t 7 − 5 at The distribution of Ω(sj ) and σxs(sj ) has been shown in Fig. C. 

2.3.2 Determination of the shear centre So far in this section we have only considered double-symmetric cross-sections for which the shear centre S coincides with the bending centre B. However, in the general case, the shear centre will be different from the bending centre as suggested by Fig. 2–1. In any case, the shear forces Qy and Qz must be statically equivalent to the shear stresses integrated over the cross- section, i.e.

Qy = σxydA, Qz = σxzdA. (2–83) ZA ZA In the present case, we are concerned with bending uncoupled from torsion. Hence, the torsional moment Mx stemming from the shear stresses on the cross-section should vanish. According to Eq. (2–25), this implies the identity

((y − ys)σxz − (z − zs)σxy) dA =0. (2–84) ZA Combining Eqs. (2–83) and (2–84) provides the relation for the coordinates of the shear centre:

−Qyzs + Qzys = (−zσxy + yσxz)dA. (2–85) ZA Elastic Beams in Three Dimensions 76 Chapter 2 – Shear stresses in beams due to torsion and bending

For a thin-walled cross-section, this identity may be recast intermsofthelocal (x,s,n)-coordinates, providing

−Qyzs + Qzys = h(s)σxs(x, s)t(s)ds, (2–86) ZL where h(s) is the moment arm of the differential shear force σxs(x, s)t(s)ds,

dz(s) dy(s) h(s)= y(s) − z(s) . (2–87) ds ds It is noted that, with the given definition, h(s) may have positive as well as negative values. Insertion of Eq. (2–74) into Eq. (2–86) provides the identity

Q Q −Q z + Q y = − z S (s)h(s)ds − y S h(s)ds (2–88) y s z s I Ωy I Ωz y ZL z ZL which holds for arbitrary Qy and Qz. This implies the following solutions for the coordinates of the shear centre: 1 1 y = − S (s)h(s)ds, z = S (s)h(s)ds. (2–89) s I Ωy s I Ωz y ZL z ZL

In particular, for a thin-walled section without branching, SΩy(s) and SΩz(s) become

s s SΩy(s)= z(s)t(s)ds, SΩz(s)= y(s)t(s)ds. (2–90) Z0 Z0 Hence, Eq. (2–89) involves a double integration with respect to the arc-length parameter over the length L. We shall arrange the calculation in a way such that only a single line integral needs to be evaluated.

Ω(s) s = 0

ds ω(s) h(s) s

S Qz z

x B Qy y

t(s)

s = L

Figure 2–13 Cross-section exposed to bending.

DCE Lecture Notes No. 23 2.3 Shear stresses from bending 77

At first the so-called sector coordinate ω(s) with pole B is introduced. The sector coordinate ω(s) is equal to the area shown in Fig. 2–13, delimited by the bending centre B and the area segment Ω(s). Thus, ω(s) is given as

s ω(s)= h(s)ds. (2–91) Z0 d It follows that ω(s) may be consideredas an integralof h(s) with respect to s, i.e. ds ω(s)= h(s). Hence, from Eq. (2–90) follows that the derivation of SΩy(s) and SΩz(s) with respect to s become d d S (s)= z(s)t(s), S (s)= y(s)t(s). (2–92) ds Ωy ds Ωz Then, integration by parts of Eq. (2–89) and use of Eq. (2–92) provides

1 L L ys = − SΩy(s)ω(s) − z(s)ω(s)t(s)ds (2–93) Iy 0 0 ! h i Z

Now, SΩy(L) = Sy = 0, because the (x,y,z)-coordinate system has origin in the bending centre. Further, both SΩy(0) = 0 and ω(0) = 0, so the first term within the parentheses vanishes in both limits and, hence,

I L y = ωz , I = ω(s)z(s)t(s)ds = ωzdA. (2–94a) s I ωz y Z0 ZA Similarly, it can be shown that

I L z = − ωy , I = ω(s)y(s)t(s)ds = ωydA. (2–94b) s I ωy z Z0 ZA

The quantities Iωy and Iωz are denoted sector centrifugal moments. An arbitrary constant ω0 can be added to ω(s) in Eq. (2–94) without changing the value of Iωz and Iωy. This is so because

A ω0ydA = ω0 A ydA = 0 and A ω0zdA = 0. Hence, the sector coordinate is determined within an arbitrary constant. R As a third method,R the coordinatesR of the shear centre may be determined from Eq. (2–86), if the shear stress σxs(x, s) is calculated from Qy and Qz separately. Thus,

1 y = σ (x, s)h(s)t(0)ds for Q =0, (2–95a) s Q xs y z ZL 1 z = σ (x, s)h(s)t(s)ds for Q =0. (2–95b) s Q xs z y ZL Obviously, if the profile has a line of symmetry, then the shear centre is placed on this line. Finally, a note is made regarding the notation in this chapter. Previously, ω(x, y) has denoted the normalised warping function or, simply, the warping function. However, we shall later see that for open thin-walled sections the warping function is identical to the sector coordinate, which motivates the naming of the latter quantity.

Elastic Beams in Three Dimensions 78 Chapter 2 – Shear stresses in beams due to torsion and bending

Example 2.9 Determination of the shear centre for an I-profile with a single line of symmetry The thicknesses of the flanges and the web of the profile shown in Fig. A (left) are all t. Due to the symmetry around the y-axis, the y-axis as well as the z-axis become principal axes. The bending centre is placed as shown in the figure and the principal moments of inertia become: 91 1044 I = a3t, I = a3t. (a) y 12 z 13

The profile is exposed to a horizontal shear force Qz, and the position of the shear centre will be deter- mined both by Eqs. (2–89), (2–94) and (2–95). Due to the symmetry, the shear centre is placed on the y-axis, i.e. zs = 0.

2a2t SΩy(s1) SΩy(s2) 4a n2 t

s1 s2 36 n1 13 a Qz

S z 6a B x

t 42 a 13 n4 y t s3 s4

3a n3 S (s3) S (s4) Ωy 9 2 Ωy 8 a t

Figure A Single-symmetric I-profile: Geometry (left); definition of arc-length coordinates and distribution of static moments (right).

For each of the four branches indicated in Fig. A (right), an arc-length parameter sj and a local

(x, nj ,sj ) coordinate system are defined. The statical moments SΩy(sj )= Ω z(sj )t(sj )dsj become: s1 1 1 R SΩy(s1)= (2a s)tds1 = (4a s1)s1t, SΩy(s2)= (4a s2)s2t, (b) − − − 2 − 2 − Z0 s3 3 1 1 SΩy(s3)= a s3 tds3 = (3a s3)s3t, SΩy(s4)= (3a s4)syt. (c) − 2 − − 2 − 2 − Z0   The shear stresses σxs(x,sj ) are positive, when acting in the direction of the arc-length parameter sj . These follow from Grashof’s formula (2–74): Q 12 Q z z (d) σxs(sj )= SΩy(sj )= 3 2 SΩy(sj ). − tIy − 91 a t The distribution of shear stresses has been shown in Fig. B (left). (continued)

DCE Lecture Notes No. 23 2.3 Shear stresses from bending 79

72 2 48 13 a Q1 182 τ n2

s2 36 36 13 a 13 a Qz 1 2 ω(s2) S z z B x B x

42 42 13 a 13 a y y

27 Q2 τ 182 63 2 13 a

Figure B Single-symmetric I-profile: Shear stresses from the shear force Qz with τ = Qz/(at) (left); distribution of the sector coordinate ω(s) (right).

The shear stress resultants Q1 and Q2 in the top and bottom flanges become:

2 48 Qz 64 2 27 Qz 27 Q1 = 4at = Qz,Q2 = 3at = Qz. (e) 3 · 182 · · at 91 3 · 182 · · at 91 The shear centre then follows from Eq. (2–95): 36 42 90 Qzys = Q1 a + Qz a ys = a. (f) − 13 13 ⇒ − 91

Next, employing another approach, the moment arm h(sj ) defined by Eq. (2–87) is negative on the branches described by arc length parameters s1 and s2, and positive along the arc-length parameters s2 and s3. Then, from Eq. (b) it follows that

SΩy(s)h(s)ds ZL 2a 1 36 2a 1 36 = (4a s1)s1t a ds1 + (4a s2)s2t + a ds2 − 2 − − 13 2 − 13 Z0   Z0   3a/2 1 42 3a/2 1 42 + (3a s3)s3t + a ds3 + (3a s4)s4t a ds4 − 2 − 13 2 − − 13 Z0   Z0   96 96 189 189 15 = s4t + a4t a4t a4t = a4t. (g) 13 13 − 52 − 52 2 Then, from Eqs. (2–89) and (a) it follows that

1 15 4 90 ys = a t = a. (h) 91 3 − 12 a t · 2 − 91 This result is identical to the result achieved in Eq. (f). However, it is noted that the result in Eq. (h) has been achieved without the determination of the shear stresses. Hence, the second approach may appear to be simpler than the first approach, but it does not provide any information about the stress distribution. (continued)

Elastic Beams in Three Dimensions 80 Chapter 2 – Shear stresses in beams due to torsion and bending

Finally, the same calculation is performed by means of Eq. (2–94). The distribution of the sector sj coordinate ω(sj )= 0 h(s)ds with the pole B for each of the four branches becomes: 36 s1R 36 36 ω(s1)= a ds1 = as1, ω(s2)= as2, (i) − 13 − 13 13 Z0 42 s4 42 42 ω(s4)= a ds4 = as4, ω(s3)= as4. (j) − 13 − 13 13 Z0 The sector centrifugal moment Iωz = A ωzdA = L ω(s)z(s)t(s)ds becomes:

2a R 2a R 36 3 42 15 4 Iωz = 2 (2a s1) as1 2tds1 + 2 a + s4 as4 2tds4 = a t, (k) − · 13 · − 2 · 13 · − 2 Z0 Z0   where the symmetry of the integrand ω(s)z(s) has been exploited. Then, from Eqs. (2–94) and (a) the shear centre is determined as

1 15 4 90 ys = a t = a. (l) 91 3 − 12 a t 2 − 91 Hence, the three different approaches lead to the same result. 

Example 2.10 Determination of the shear centre for a symmetric U-profile The bending centre of the U-profile has the position shown in Fig. A. Since the profile is symmetric, the indicated (x,y,z)-coordinate system is a principal axis system. The moments of inertia with respect to the principal axes become: 64 26 I = a3t, I = a3t. (a) y 15 z 3

2a 4 a 4 a 5 2t 5

a Q2 B B S Qz S z z x x t a Q1 2t Q2

Qy Qy

112 y 65 a y Figure A Symmetric U-profile: Geometry (left); position of the shear centre (right).

Local arc length coordinates s1, s2 and s3 are introduced as indicated in Fig. B. The distributions of the static moments SΩy(sj ) and SΩz(sj ) are shown in the figure. The analytic expressions are given as (continued)

DCE Lecture Notes No. 23 2.3 Shear stresses from bending 81

s1 6 12 2 SΩy(s1)= a s1 2tds1 = as1 s t, (b) 5 − 5 − 1 Z0     s2 4 2 4 4 SΩy(s2)= a t + a tds2 = at(a s2), (c) 5 − 5 5 − Z0   s3 4 2 4 4 2 8 2 SΩy(s3)= a t + a + s3 2tds3 = a t as3t + s t, (d) − 5 − 5 − 5 − 5 3 Z0   s1 SΩz(s1)= ( a)2tds1 = 2as1t, (e) − − Z0 s2 2 2 1 2 SΩz(s2)= 4a t + ( a + s2)tds2 = 4a t as2t + s t, (f) − − − − 2 2 Z0 s3 2 2 SΩz(s3)= 4a t + a 2tds3 = 4a t + 2as3t. (g) − · − Z0

SΩy(s1) 4 2 − 2 SΩz(s1) 5 a t 26 2 4a t 5 a t

s1SΩz(s2) s1 s2 s2 B B z z x 1 2 x 2 a t SΩy(s2) s3 y s3 y

26 4 2 a2t − 2 5 a t 5 4a t SΩz(s3) SΩy(s3)

Figure B Symmetric U-profile: Distribution of the static moments SΩy(sj ) (left) and SΩz(sj ) (right).

σxs(s1) 3 6 σxs(s1) 32 τz 27 26 τy 160 τz

σxs(s2)

B B z z 3 x τy x σxs(s2) 52 y y

27 3 τz 6 32 τz 160 26 τy σxs(s3) σxs(s3)

Figure C Symmetric U-profile: Distribution of the shear stresses from the shear forces Qz with τz = Qz/(at) (left) and Qy with τy = Qy/(at) (right). (continued)

Elastic Beams in Three Dimensions 82 Chapter 2 – Shear stresses in beams due to torsion and bending

The shear stresses follow from Grashof’s formula (2–74). Figure B (left) shows the distribution of shear stresses from Qz, and Fig. B (right) shows the distribution from Qy. The position of the shear centre is given by Eq. (2–95). Since the shear stresses σxs from Qz are distributed symmetrically around the z-axis it follows that ys = 0, reflecting the fact that the z-axis is a line of symmetry. The resulting shear forces Q1 and Q2 in Fig. A then become

1 6 6 2 3 6 Q1 = 2t 2a τy = Qy,Q2 = t + 2aτy = Qy. (h) · 2 · · 26 13 · 3 · 52 13   Finally, from Eq. (2–95) follows that

1 4 112 zs = Q1 a + a Q2 + Q1a = a. (i) − Qy · 5 · − 65   Hence, for a U-profile the shear centre lies at a considerable distance outside the profile. 

2.3.3 Shear stresses in closed thin-walled sections

Figure 2–14 shows a closed single-cell section of a thin-walled beam. The shear force H(x, s) acting within the wall per unit length of the beam (see Fig. 2–12) in a corresponding open section is uniquely determined by Eq. (2–71) due to the condition H(x, 0) = H0(x)=0 at the boundary s = 0. However, in a closed section H(x, s) is not a priori known in one or more points of the periphery Γ1, for which reason the shear stresses cannot be determined from static equations alone. Hence, a geometrical condition must be introduced, from which the initial value H0(x) at s =0 can be determined. The derivation of this geometric condition will be considered at first.

σxs

s

s n s

n B z Ω(s) S Qz

Γ1 Qy s = L

t(s) H0 s = 0 y

Figure 2–14 Closed single-cell section exposed to bending.

DCE Lecture Notes No. 23 2.3 Shear stresses from bending 83

The arc-length coordinate s is orientated clock-wise with an arbitrarily selected origin O. Then, the local (x,n,s)-coordinate system is orientated with the n-axis orientated toward the interior of the cell. The shear stress σxs(x, s) follows from Eqs. (2–71) and (2–73):

1 H0(x) Qy(x) Qz(x) σxs(x, s)= H(x, s)= − SΩz(x, s) − SΩy(x, s). (2–98a) t(s) t(s) t(s)Iz t(s)Iy where the static moments of the area segment Ω(s) are again given as

s s SΩy(s)= z(s)t(s)ds = zdA, SΩz(s)= y(s)t(s)ds = ydA. (2–98b) Z0 ZΩ Z0 ZΩ Equation (2–98a) is the equivalence of Grashof’s formula (2–74) for a closed thin-walled section. Unlike the case of the open section, H0(x) is generally different from zero and the determination of H0(x) is part of the problem. The shear strain εxs in the local (x,n,s)-coordinate system is given by Eq. (2–76) which is valid for open as well as closed thin-walled sections. This implies the warping u0(x, s) defined by Eq. (2–80), where it is recalled that u0(x, s)=0 in the bending centre B. This is so, because wx(x) has been defined as the total displacement in the x-direction of B. Now, for the closed section, the displacement ux(x, s) must be continuous as a function of s along the periphery Γ1. The axial and bending contributions, i.e. the first three contributions on the right-hand side of Eq. (2–77), are always continuous. A possible discontinuity then stems from the warping. If the profile is open, as illustrated in Fig. 2–12, the ends at s =0 and s = L can move freely relatively to each other. Hence, a warping discontinuitydevelopsbetween these two endpoints. However, in a closed section, the warping of these points must be identical. Hence, the geometrical condition can be formulated as ux0(x, 0) = u(x, L)= u0(x). (2–99) Insertion of Eqs. (2–76) and (2–99) into Eq. (2–80) provides

L 1 L H(x, s) u (x)= u (x)+2 ε (x, s)ds = u (x)+ ds ⇒ 0 0 xs 0 G t(s) Z0 Z0 L H(s) H(s) ds = ds =0. (2–100) t(s) t(s) Z0 IΓ1 Insertion of Eq. (2–98a) gives the following formulation of the geometrical condition from which the initial condition H0 can be determined: ds Q S Q S H (x) − z Ωy ds − y Ωz ds =0. (2–101) 0 t(s) I t(s) I t(s) IΓ1 y IΓ1 z IΓ1

With H0(x) determined, H(x, s) and the shear stress σxs(x, s) can next be determined from Eq. (2–98a).

Example 2.11 Shear stresses in a symmetric thin-walled single-cell section Figure A shows a double-symmetric thin-walled single-cell profile. The thickness is everywhere t. We want to determine the shear stresses from a shear force Qy acting at the shear centre. (continued)

Elastic Beams in Three Dimensions 84 Chapter 2 – Shear stresses in beams due to torsion and bending

The cross-sectional area and moment of inertia around the z-axis become 5 A = 8at, I = a3t. (a) z 3 The origin of the arc length coordinate is chosen at the lower left corner. The corresponding distribution s of the static moment SΩz(s)= 0 z(s)t(s)ds has been shown in Fig. B. R

t(s) B = S z a x

Qy

2 a 2 a 3 y 3 Figure A Geometry of the symmetric rectangular thin-walled single-cell section.

3 2 2 a t

B = S z

1 2 x 8 a t 1 2 8 a t s y

3 2 2 a t

Figure B Distribution of the static moment SΩz(s) in the symmetric rectangular thin-walled single-cell section.

Next, the following line integrals are calculated:

ds 1 8a = (a + 3a + a + 3a)= , (b) t t t IΓ2

S (s) Ωz ds t IΓ1 1 2 1 1 3 3 2 1 1 3 = a2t a 3a a2t a2t a a2t a 3a a2t = 6a3. (c) t 3 · 8 · − 2 · · 2 − 2 · − 3 · 8 · − 2 · · 2 −   (continued)

DCE Lecture Notes No. 23 2.3 Shear stresses from bending 85

From Eq. (2–101) then follows that

8a Qy 3 9 Qy H0 ( 6a ) = 0 H0 = . (d) 5 3 · t − 3 a t − ⇒ − 20 at

−18τ 18τ

B = S z x 3τ 3τ Qy

−18τ 18τ y

Figure C Shear stresses in the symmetric rectangular thin-walled single-cell section with τ = Qy/(40at).

The distribution of shear stresses σxs(x,s) follows from Eq. (2–98a) and has been indicated in Fig. C. The arrows indicate the positive direction of the shear stresses. These will be referred to as the shear flow in what follows, due to the analogous behaviour of water flowing down through a system of pipes. 

As shown in Example 2.11, a symmetric thin-walled single-celled profile exposed to a shear force acting along the line of symmetry will have a symmetric distribution of shear stresses. Especially, the shear stress at the line of symmetry vanishes. Hence, if the origin of arc-length coordinate is placed at the line of symmetry we must have H0(x)=0, entailing that a single- celled symmetric profile can be analysed from the static equations alone. For a thin-walled profile with N cells, arc-length parameters sj are introduced for all cells, orientated clock-wise as shown in Fig. 2–15. The origins O1, O2,...,ON for the arc-length coordinates are chosen arbitrarily. The shear forces per unit length H0j (x) at the origins are unknown and must be determined from geometrical conditions in addition to the static shear flow equations. Similarly to Eq. (2–99) these are determined by the conditions that the warping must be continuous along all the peripheries Γj of the cells. At first, the static moments SΩy and SΩz for the open profile in Fig. 2–15 are determined. Especially, the variation of SΩy(x, sj ) and SΩz(x, sj ) along Γj as a function of sj are registered. Next, the calculation of the shear force per unit length Hj (x, sj ) within each cell can be arranged as illustrated in Fig. 2–16. Firstly, the shear force per unit length Hsj (sj ) of the open, branched profile is calculated, orientated in the direction of the arc-length coordinate sj . This is given by Grashof’s formula (2–74), i.e.

Qz Qy Hsj (sj )= − SΩy(sj ) − SΩz(sj ). (2–103) Iy Iz

Elastic Beams in Three Dimensions 86 Chapter 2 – Shear stresses in beams due to torsion and bending

σ xs H0N Γ0 S Qz

ON

Qy H01 N Γj s s O1 N x j B j z

1 O s1 j

sk k H0j Ok

H0k y

Figure 2–15 Closed multi-cell section exposed to bending.

N HN j

H1 1 Hk Hj k

N HsN j

Hs1 1 Hsk Hsj k

N H0N j

H01 1 H0k H0j k

Figure 2–16 Arrangement of calculations of shear forces per unit length within cells.

DCE Lecture Notes No. 23 2.3 Shear stresses from bending 87

In each cell, the initial conditions induce a constant shear force per unit length H0j . Due to the chosen sign convention, the net shear force on the periphery segment Γjk between cell j and cell k becomes Hsj − H0k. Hence, on this segment the total shear force per unit length becomes

Hj (sj )= Hsj (sj )+ H0j − H0k. (2–104) Continuity requires that Eq. (2–101) is fulfilled for all N cells, i.e. H (s ) j j ds =0, j =1, 2,...,N. (2–105) t(s ) j IΓj j Insertion of Eq. (2–104) leads to the following N linear equations for the determination of H01, H02,...,H0N :

N dsj Hsj (sj ) (H0j − H0k) = − dsj , j =1, 2,...,N. (2–106) Γjk t(sj ) Γj t(sj ) kX=1 Z I The coefficient matrix in Eq. (2–106) is identical to this for the determination of Prandtl’s stress function Sj in the cells in the corresponding St. Venant torsion problem. With H0j determined from Eq. (2–106), Hj (sj ) can be calculated from Eq. (2–104). Finally, the shear stresses σxs(sj ) follow from Eq. (2–98a) with positive sign when acting in the direction of sj .

Example 2.12 Shear stresses in a single-symmetric double-cell section The profile shown in Fig. A is identical to that considered in Example 2.11. However, a partition with the wall thickness t has been included as shown in the figure. The bending centre B is placed as indicated. Because the section is symmetric around the z-axis the indicated coordinate system with origin in B is a principal axis coordinate system. The section is still exposed to a shear force Qy acting through the shear centre S. Due to the symmetry the shear centre is placed on the z-axis (yS = 0). The zs will be determined as a part of the solution. The cross-sectional area and inertial around the z-axis become 7 A = 9at, I = a3t. (a) z 4

13 9 a

a t(s) t(s) t(s) 2 S B z a x Qy a 2

y a 2a

Figure A Geometry of the rectangular thin-walled double-cell section. (continued)

Elastic Beams in Three Dimensions 88 Chapter 2 – Shear stresses in beams due to torsion and bending

O1 O2

1 s1 s2 2 H1 H2

Γ1 Γ2

Figure B Definition of the origins of the arc-length coordinates in the rectangular thin-walled double-cell section.

1 2 2 2 a t a t

SΩz(s1) SΩz(s2)

1 2 1 2 1 2 8 a t 8 a t 8 a t 1 2 8 a t

1 2 2 2 a t a t

Figure C Distribution of the static moments SΩz(s1) and SΩz(s2) in the rectangular thin-walled double-cell section.

The origin of the arc-length coordinates are chosen as shown in Fig. B. The static moments SΩz(s1) and SΩz (s2) are next calculated with the sign and magnitude as indicated in Fig. C. Especially, it is seen that SΩz(s1)= SΩz(s2). The shear force per unit length in the open profile is given by Eq. (2–103): − Qy Hsj (sj )= SΩz(sj ), j = 1, 2. (b) − Iz Next, Eq. (2–106) provides

a a a a Hsj (s1) (H10 H20) +(H10 0) +(H10 0) +(H10 0) = ds1, (c) − t − t − t − t − t(s1) IΓ1

2a a 2a a Hsj (s2) (H20 0) +(H20 0) +(H20 0) +(H20 H10) = ds2. (d) − t − t − t − t − t(s2) IΓ2 The right-hand sides are calculated by insertion of Eq. (b) with the distribution of SΩz(sj ) shown in Fig. C. The results become

H (S ) 4 Q 2 a2t a a2t a2t 2 a2t s a2t 4 Q sj 1 y y (e) ds1 = 3 2 a a a = , t(s1) 7 a t 3 · 8 − 2 · 2 − 2 − 3 8 − 2 2 − 7 t IΓ1   H (s ) 4 Q 2a 2 2a 2 a2t 12 Q sj 2 y 2 2 2 2 y (f) ds2 = 3 2 a t + a a t + a a t + a t a = . t(s2) − 7 a t 2 · · 3 · 2 − 3 8 7 t IΓ2   (continued)

DCE Lecture Notes No. 23 2.3 Shear stresses from bending 89

Then, the following solutions are obtained for the initial values of the shear forces per unit length:

4 Qy 12 Qy 4H 10 H20 = , H10 + 6H20 = , (g) − − − 7 a − − 7 a which implies that

24 Qy 88 Qy H10 = ,H20 = . (h) 322 a − 322 a

Next, Hj (sj ) is calculated from Eqs. (2–103) and (2–104) along both peripheries Γ1 and Γ2. Finally, the shear stresses σxs(sj ) are computed from Eq. (2–98a). Due to the constant wall thickness this reduces to 1 σ (s )= H (s ), j = 1, 2. (i) xs j t j j

The flow of the shear stresses σxs has been shown in Fig. E.

68τ 24τ 96τ

88τ 88τ 88τ 88τ 23τ 23τ 23τ

68τ 24τ 96τ

Figure D Distribution of the shear stresses in the rectangular thin-walled double-cell section with τ = Qy/(322at).

13 9

Q4 Q1 a 2 S B z a Q5 Q3 Q2 x Qy y a 2 Q4 Q1

80 1449 Figure E Shear forces in the wall segments of the rectangular thin-walled double-cell section. (continued)

Elastic Beams in Three Dimensions 90 Chapter 2 – Shear stresses in beams due to torsion and bending

Next, the shear forces Q1,Q2,Q3,Q4 and Q5 in the wall segments shown in Fig. E are calculated. These become:

1 1 Qy 12 Q1 = (96 88) 2a t = Qy, (j) 2 − · · · 322 at 483

2 1 Qy 167 Q2 = (96+ 23) a t = Qy, (k) 3 · · · · 322 at 483 2 1 Qy 191 Q3 = (112 + 23) a t = Qy, (l) 3 · · · · 322 at 483 1 1 Qy 33 Q4 = (68 24) a t = Qy, (m) 2 − · · · 322 at 483 2 1 Qy 125 Q5 = (68+ 23) a t = Qy. (n) 3 · · · · 1288 at 483 It is seen that Q2 + Q3 + Q5 = Qy. The static equivalence (2–83) of the shear stresses is then fulfilled. The position zs of the shear centre follows from Eq. (2–95):

1 14 4 13 80 zs = (Q4 Q1)a Q2 a + Q3 a + Q5 a zs = a. (o) − Qy − − · 9 · 9 9 ⇒ − 1449   The position of the shear centre has been illustrated in Fig. E. 

Example 2.13 Distribution of shear stresses in open and closed sections due to bending and St. Venant torsion

Q Q Q

Q Q

Mx Mx

Figure A Shear flow in open and closed sections due to transverse shear forces and torsional moments.

Fig. A shows the shear flow in open and closed thin-walled sections exposed to a transverse shear force Q or a torsional moment M. For closed sections the shear stresses σxs(x,s) are uniformly distributed over the wall thickness for both loadings. For open section this is only the case in the case for the shear force loading, whereas the shear stresses for St. Venant torsion is linearly varying in the thickness direction around the mid-line of the wall. 

DCE Lecture Notes No. 23 2.4 Summary 91

2.4 Summary

The computation of shear stresses in beams has been the focus of this chapter with detailed explanation of the theory for thin-walled sections subjected to torsion and/or bending. A brief summary of the main findings is given in the following. Uncoupling of bending and torsion requiresthat the shear force acts throughthe so-called shear centre. The shear centre always lies on a line of symmetry within a symmetric cross-section. Hence, the position of the shear centre coincides with that of the bending centre for double- symmetric sections. St. Venant torsion is characterised by the fact that the beam is allows to warp freely, leading to a homogeneous twist of the cross-section along the beam axis. The homogenous torsion problem can be defined in terms of the warping functionor, alternatively, in terms of Prandtl’s stress function. Shear stresses due to homogeneous torsion vary linearly over the thickness of the wall in open thin-walled profiles. However, in a closed thin-walled section with one or more cells, the shear stresses due homogeneous torsion is homogeneous over the thickness. Homogeneous torsion induces warping in a beam. The warping increases linearly with the tor- sional moment. However, in a circular profile (a cylinder as well as a tube) there is no warping. The torsional stiffnesses of open and closed cross-sections are different . Thus, aclosed cross- section has a much higher torsional stiffness than an open cross-section with a similar cross- sectional area. Likewise, the ultimate strength of a closed section is higher than that of a similar open section. Shear stresses from bending can generally be analysed by means of the so-called stress func- tion. However, this requires the use of a numerical scheme, e.g. the finite-element method. Grashof’s formula defines the shear stresses from bending in open thin-walled sections. For closed thin-walled sections, Grashof’s formula must be adjusted be an additional term that Warping takes place due to bending since the shear stresses are accompanied by shear strains. For a rectangular cross-section, the beam warps into an S-shape. Shear flow is a graphical interpretation of the direction in which the shear stresses are acting on the cross-section of a beam. In the case of bending, this forms an analogy to water streaming down a system of pipes. Internal walls in a section will not increase the torsional stiffness and strength significantly. However, the inclusion of an internal wall oriented in the direction of the shear force will provide an increase of the shear strength and stiffness.

The theory presented in this chapter can be used for the determination of the shear stresses and warping deformations in beams subjected to any combination of bending and homogeneous torsion. However, if torsion is prevented, e.g. at one end of the beam, another theory must be applied as described in the next chapter.

Elastic Beams in Three Dimensions 92 Chapter 2 – Shear stresses in beams due to torsion and bending

DCE Lecture Notes No. 23 References

Krenk, S. (2001). Mechanics and analysis of beams, columns and cables (2nd ed.). Springer Verlag. SaintVenant, B. (1855). Memoire sur la torsion des prismes. Mem. Acad. Sci. Savants Etrangers 14, 233– 560. Timoshenko, SP (1921). On the correction for shear of the differential equation for transverse vibrations of static bars. Philosophical Magazine, Series 6 41, 744–746. Vlasov, V.Z. (1961). Thin-walled elastic beams. Israel Program for Scientific Translations.

— 93 — 94 References

DCE Lecture Notes No. 23

ISSN 1901-7278 DCE Lecture Notes No. 23