Vous êtes sur la page 1sur 82

UMIST

Department of Mechanical, Aerospace and Manufacturing Engineering


Analysis of thin walled aerospace structures
Dr D.A. Bond
Pariser Bldg. C/21
e-mail: d.bond@umist.ac.uk
Tel: 0161 200 8733

UMIST
2nd YEAR LECTURE NOTES
ANALYSIS OF THIN WALLED
AEROSPACE STRUCTURES
Analysis of thin walled aerospace structures

TABLE OF CONTENTS

TABLE OF CONTENTS
1. INTRODUCTION
2. TERMINOLOGY, SYMBOLS, AND SIGN CONVENTIONS ETC.
2.1 Nomenclature
2.2 Terminology
2.3 Sign Conventions
2.4 Recommended Reading
3. USEFUL APPROXIMATIONS AND FUNCTIONS FOR THIN WALL STRUCTURES
3.1 Product second moment of area IXY
3.2 Thin wall approximations
4. STRUCTURAL IDEALISATION (AFTER THG MEGSON - AIRCRAFT STRUCTURES)
4.1 Introduction
4.2 Idealisation of structures with a linear stress variation in the skin section
5. BENDING OF ARBITRARY CROSS-SECTION PRISMATIC BEAMS
5.1 Planar bending
5.2 Multi-plane bending
5.3 Combined bending of an arbitrary prismatic beam
TUTORIAL SET 1: STRUCTURAL IDEALISATION & AXIAL STRESS DUE TO BENDING
6. TORSION OF MULTICELL SECTIONS
6.1 Torsion of a Single-Cell Closed Section
6.2 Torsional of an open section
6.3 Torsion of Multi-cell sections (Statically Indeterminate)
TUTORIAL SET 2: TORSION OF THIN WALLED STRUCTURES
7. BENDING OF PRISMATICAL BEAM BY TRANSVERSE FORCES
7.1
The Shear Centre
7.2
Bending Shear Stress (Shear stresses induced by bending)
8. CALCULATION OF SHEARING-STRESS DISTRIBUTION AND SHEAR CENTRE POSITION
8.1 Calculation of Shear Centre Co-ordinates:
8.2 Reduction in Work due to Symmetry
8.3 Calculation of Resultant Shearing Stress
9. ANALYSIS OF COMBINED OPEN AND CLOSED SECTIONS
TUTORIAL SET 3: SHEAR STRESS DISTRIBUTION AND SHEAR CENTRES
10. CALCULATION OF DEFLECTIONS
10.1 Beam Deflections due to Pure Bending
10.2 Deflections due to shear loading
10.3 Deflection in thin wall structures by unit load method
11. CONCLUDING REMARKS
TUTORIAL SET 4: DEFLECTIONS

2
Analysis of thin walled aerospace structures

1. INTRODUCTION

Aircraft are not rigid structures. When in operation they deform due to combinations of the aerodynamic
forces that allow them to fly as well as inertial and mass forces (see Figure 1.1). These deflections may be
critical to the way in which an aircraft operates in particular the expected performance of wings and
control surfaces. This course will give a brief introduction to the analysis of thin walled structures that
aircraft are well represented by. On completion of this course a student should be capable of estimating the
deflection and twist of a thin walled aircraft structure under generalised loading conditions. This will
allow an engineering student to assess the effect of aerodynamic loading on the performance of an aircraft
design.

Figure 1.1: Wing deformations due to loading

An aircraft is basically an assembly of stiffened shell structures ranging from the single cell closed section
fuselage to multicellular wings and tail-surfaces each subjected to bending, shear, torsional and axial
loads. Other, smaller portions of the structure consist of thin-walled channel, T-, Z-, 'top hat' or I-sections,
which are used to stiffen the thin skins of the cellular components and provide support for internal loads
from floors, engine mountings etc. Structural members such as these are known as open section beams,
while the cellular components are termed closed section beams; clearly, both types of beam are subjected
to axial, bending, shear and torsional loads.
In this course we shall investigate the stresses and displacements in thin-walled open and single cell closed
section beams produced by bending, shear and torsional loads, and, in addition, we shall examine the
effect on the analysis of idealizing such sections when they have been stiffened by stringers. In general the
structure under consideration shall be a wing however the analysis techniques are valid for all similar thin-
walled aerospace structures.
The value of direct stress at a point in the cross-section of a beam subjected to bending shall be shown to
depend only on the position of the point, the applied loading and the geometric properties of the cross-
section. It follows that it is of no consequence whether or not the cross-section is open or closed. We
therefore derive the theory for a beam of arbitrary cross-section and then discuss its application to thin-
walled open and closed section beams subjected to bending moments.
Strictly speaking, the axial-stress distribution in a beam is statically indeterminate, but the assumption that
plane cross-sections remain plane reduces this problem to a quasi-static problem. Theory yields the well-
known formulae for axial stress [ = My/I] and for the corresponding shearing stress ( = VQ/It) when the
σ

bending moments arise from transverse forces applied to the beam. These formulae are valid for cross-
sections of arbitrary shape provided principal axes are used. However, the position of the principal axes is
usually not known in advance for an asymmetrical section, and must be determined by computing the
section properties with respect to an arbitrary set of axes. It is then necessary to recompute these section
properties with respect to the principal axes, and to resolve the applied bending moments and shearing
forces in the direction of the principal axes.

3
Analysis of thin walled aerospace structures

The use of principal axes for the analysis of an asymmetrical section involves a considerable amount of
tedious computation. The techniques presented in this course formulate the problem using an arbitrarily
chosen set of rectangular axes that pass through the centroid of the cross-section. With this approach the
section properties are calculated once only, that is, with respect to the axes chosen, and the applied
bending moments and shearing forces are referred to these axes. The formulae for axial stress, bending
shear stress (flow) and for bending deflection derived by this approach are directly analogous to the
corresponding principal-axes formulas. In each case, bending moments and shearing forces (with respect
to principal axes) are replaced by quantities called effective bending moments and effective shearing
forces, and section properties (with respect to principal axes) are replaced by section properties with
respect to the. axes chosen. These effective quantities are simple combinations of bending moments,
shearing forces, and section properties with respect to the axes chosen.
The concept of shear centre is defined and methods are given to determine its position. The calculation of
the shear centre co-ordinates is simplified by the use of particular combinations of applied shearing forces
(called shear devices) which make a specified co-ordinate axis a neutral axis. Thus, by applying the shear
device to each axis in turn, the appropriate bending shear stress with respect to each axis is obtained,
together with the two equations for the determination of the shear centre co-ordinates.
In the analysis of the multicell section under torsion, a technique is presented which yields, in one
calculation, the shearing-stress distribution together with the torsion constant of the section. This
technique incorporates redundancies which have a minimum amount of cross-coupling and so leads to
well-conditioned equations for the solution. The combination of this technique with the analysis for
bending shear stress produces a general procedure for the analysis of a multicell section of arbitrary shape,
subjected to any system of forces in the plane of the cross-section. This combined analysis yields the
distribution of resultant shearing stress, the shear centre co-ordinates, and the torsion constant of the
section. If it is required to calculate only the resultant shearing-stress distribution, this can be done more
directly by the alternative procedure presented which does not require the calculation of the shear centre
position.

Trailing Edge
Skin/Plank or
Control Surface

Aft or Closing Spar

Rib
Stringers/Stiffeners

Leading Edge Forward Spar

Figure 1.2: Structure of a wing

Wing sections consist of thin skins stiffened by combinations of stringers, spar webs and caps, and ribs.
The resulting structure frequently comprises of two or more cells and as such is highly redundant (i.e. has
loading that cannot be determined from consideration of statics alone). However, the large number of
closely spaced stringers allow the assumption of a constant shear flow in the skin between adjacent
stringers so that a wing section may be analysed as though it were completely idealised as long as the
direct stress carrying capacity of the skin is allowed for by addition to the existing stringer/boom areas.
We shall investigate the analysis of multi-cellular wing sections subjected to bending, torsional and shear
loads. For this purpose an understanding of the basic structure of a modern aircraft wing is useful. A
summary of the major components is shown in Figure 1.2.

4
Analysis of thin walled aerospace structures

2. TERMINOLOGY, SYMBOLS, AND SIGN CONVENTIONS ETC.

2.1 Nomenclature
A Area enclosed by perimeter of cell.
b Spacing between booms in an idealised section
B Cross-sectional area of axial-force element.
C Centroid of cross-section (axial stress-carrying area).
E Shear centre; modulus of elasticity.
F, FX, FY Applied/Resolved forces
G Shear modulus.
IX, IY, IXY Moment and product of inertia of cross-section, with respect to xy-axes.
J Torsion constant.
Section constant K = 1− I2
K − XY
I X IY
MX, MY Bending moment about x, y-axis.
   
M'X, M'Y Effective bending moments defined by: M ′X = 1  M X + I XY M Y  , M Y′ = 1  M Y + I XY M X 
K   
IY  K IX 
n Number of cells in multicell section.
p Perpendicular distance from tangent at s to the centroid C.
P Axial force.
q Shear flow.
QX, QY Moment about x- and y- axes of axial stress carrying area.
rX, rY Radii of curvature in yz- and xz-planes.
s Distance along perimeter of cross-section.
t Wall thickness.
T Torsion about z-axis.
US Strain energy due to shearing stress.
u, v, w Displacements in x, y, z directions.
VX, VY Shearing force in x, y direction.
   
V'X V'Y Effective shearing forces defined by: V X′ = 1  VX − I XY VY  , VY′ = 1  VY − I XY VX 
K   
IX  K IY 
wX, wY Distributed loading in x, y direction.
x, y, z Co-ordinate axes.
β

Skew load angle


σ

Axial stress.
τ Shearing stress.
Rate of twist or angle between neutral axis and x-axis.
θ

Subscripts:
D Direct stress carrying thickness of skin.
i Typical cell of multicell section.
i,0 Value (of q) at origin (of s).
i-1, i Interior wall between cell (i-1) and cell i.
r Typical axial-force element; wall between axial-force elements Br and Br+l .

5
Analysis of thin walled aerospace structures

2.2 Terminology
It is convenient, for purposes of analysis, to classify thin-wall sections into two classes.
(1) Effective Sections.
In this class the walls of the section are capable of carrying axial stresses (that is, stresses acting normal to
the plane of the cross-section) due to the applied bending moments, in addition to shearing stresses (acting
in the plane of the cross-section) due to the applied shearing forces. The beam may be stiffened or
reinforced by longitudinal members usually called stringers.
Such sections are usually called effective sections, the term “effective” implying that the walls of the
section are fully effective in contributing to the moment of resistance of the section.
(2) Idealised Sections.
In order to simplify the analysis, it is often convenient to formulate an analytical model or idealised
section to represent the behaviour of the actual section. In this case the axial stress-carrying capacity of the
walls is “lumped-in” with the adjacent stringers, so that the idealised section consists of idealised stringers
which carry only axial stresses and idealised walls which carry only shearing stresses. The idealised walls
are called shear fields.
The term axial-force element is used to describe the stringer in an effective section, and the idealised
stringer in an idealised section.
2.3 Sign Conventions

y, v

wy
θ

x, u

Vy
My
wx

Vx
T
P Mx
z, w

Figure 2.1: Sign Convention

(1) The origin of co-ordinates is taken at the centroid C of a cross-section, the z-axis is directed along
the axis of the beam and the x- and y-axes are selected to form a right-handed co-ordinate system as
shown in Figure 2.1
(2) The shearing force components VX, and VY, and the distributed loading components wX, and wY, are
each measured positive in the positive direction of the corresponding co-ordinate axis as shown in
Figure 2.1.
(3) The positive directions of bending moment components MX and MY and of torsion T are in
accordance with the right-hand screw rule as shown in Figure 2.1.
(4) The displacements u, v, w are measured positive in the positive direction of the co-ordinate axes x,
y, z

6
Analysis of thin walled aerospace structures

ds x

s
s=0 dz

Figure 2.2: Sign convention for peripheral variable s

(5) Generally the peripheral co-ordinate s is measured positive in accordance with the right-hand screw
rule as shown in Figure 2.2. Shearing stress (in the plane of the cross-section) is measured positive
τ

in the positive direction of s.


(6) The internal forces (stresses) are taken as the actions on the section which are equivalent to the
forces applied to the section. For example, Figure 2.3(a) shows the applied loading action on a
section and Figure 2.3(b) the equivalent shear-flow distribution. The resultants of the stresses are
taken positive in the positive senses of the reference axes. Thus, all forces, whether applied or stress
resultant, are positive in the positive senses of the reference axes.

(a) (b)
Figure 2.3: Equivalence of applied loads and shear flow

7
Analysis of thin walled aerospace structures

(7) The angles of rotation of cross-sections are measured positive in accordance with the right-hand
screw rule. A radius of curvature is taken positive if the angle of rotation of cross- sections
increases as z increases. With this convention positive curvature is associated with positive bending
moment. Figure 2.4 shows the positive sense of displacement, rotation, radius of curvature and
bending moment in the xz- and yz- planes.

rY
x y

MY
MY MX
MX
v
rX
u

z z
xz- plane yz- plane

Figure 2.4: Sign convention for curatures

2.4 Recommended Reading


Megson: Aircraft Structures for Engineering Students
Niu: Airframe Structural Design
Bruhn: Analysis and Design of Flight Vehicle Structures
Gere and Timeshenko: Mechanics of Materials

8
Analysis of thin walled aerospace structures

3. USEFUL APPROXIMATIONS AND FUNCTIONS FOR THIN WALL STRUCTURES

3.1 Product second moment of area IXY


IXY is defined as:

I XY = ∫ xy dB (1)

IXY = 0 about a set of axes where at least one of the axes is an axis of symmetry. Axes about which IXY = 0
are known as principal axes. Principal axes are the set of orthogonal axes for which the second moments
of area are a maximum and a minimum.
To find IXY about a set of axes other than centroidal axes:

y
y' Area B

C
yc
x

xc x'

Figure 3.1: Parallel axis theorem for product second moment of area

x' = x + x c
y' = y + y c
I X 'Y ' = ∫ x ' y' dB (2)

= ∫ xy dB + ∫ x c y dB + ∫ xy c dB + ∫ x c y c dB
= I XY + x c y c B
This is equivalent of parallel axis theorem for product second moment of area

9
Analysis of thin walled aerospace structures

3.1.1 Example: Product second moment of area for an effective section


Determine the second moments of area about the centroidal axes of the T-section shown below:

y
40mm 80mm

x y
8mm
80mm C x

8mm
Figure 3.2: Cross-section for example 5.3.5

y = 17.6mm
x = 12.0 mm
I X = 1.09 ×10 6 mm 4
IY = 1.31×10 6 mm 4
I XY = 0.34 ×106 mm 4

10
Analysis of thin walled aerospace structures

3.2 Thin wall approximations


3.2.1 Elimination of higher order thickness terms
If thickness t is small in relation to all other dimensions, square and higher terms for t may be ignored
when calculating IX, IY and IXY. If this assumption is used it must be clearly indicated.

y
t y'
yc C

x' x
xc
x

Figure 3.3: Thin wall approximations for t<<x

bt 3
IX = ≈0 I X ' = I X + Bd 2 ≈ y 2c bt
12
tb 3 tb 3
IY = I Y ' = I Y + Bd 2 = + x c2 bt (3)
12 12
I XY = 0 I X 'Y′ = I XY + x c y c B = x c y c bt
3.2.2 Inclined elements

y ds
s
γ

y
x
C x
t
a

Figure 3.4: Thin wall approximations for inclined sections

Thin walled sections often have inclined or curved walls which may also be simplified through use of thin
wall approximations. The second moment of area for such a section may easily be shown to be:

a 3t sin 2 γ
a a
2 2

I X = 2 ∫ ty 2 ds = 2 ∫ t (s sin γ ) ds =
2

0 0
12
a 3t cos 2 γ
a a
2 2

IY = 2 ∫ tx 2 ds = 2 ∫ t (s cos γ ) ds =
2
(4)
0 0
12
a 3t cos γ sin γ
a a

I XY = 2 ∫ txyds = 2 ∫ t (s cos γ )(s sin γ )ds =


2 2

0 0
24

11
Analysis of thin walled aerospace structures

3.2.3 Example: Second moments of area for a Z section stiffener


Using thin wall approximations, determine the second moments of area about the centroidal axes of the Z-
section stiffener shown below:

y h/2
4
3
t

h C
x

2
1 h/2

Figure 3.5: Cross-section of Z section stiffener

th 3
IX =
3

th 3
IY =
12

th 3
I XY =
8

12
Analysis of thin walled aerospace structures

3.2.4 Example: Second moments of area of a C section stiffener


Using thin wall approximations for the c-section channel shown below, determine:
a. the location of the centroid, and
b. the second moments of area about the centroidal axes.

2
x
h

Figure 3.6: Cross section of an idealised C-section stiffener

h
centroid = from centreline of section 2
4
7th 3
IX =
6
5th 3
IY =
12
I XY = 0

13
Analysis of thin walled aerospace structures

4. STRUCTURAL IDEALISATION (AFTER THG MEGSON - AIRCRAFT STRUCTURES)

4.1 Introduction
In general aircraft structures are thin skin sections reinforced by localised stiffeners (often known as spars,
stringers or longerons) used to stiffen more complex structural shapes such as fuselages, wings and tail
surfaces. Thus a two spar wing section could take the form shown in Figure 4.1(a) in which Z-section
stringers are used to stiffen the thin skin while angle sections form the spar flanges.
The exact analysis of such sections is complicated and simplifying assumptions are made to facilitate
solutions. Generally, the number and nature of these simplifying assumptions determine the accuracy and
the degree of complexity of the analysis; the more complex the analysis the greater the accuracy obtained.
The degree of simplification introduced is governed by the particular situation surrounding the problem.
For a preliminary investigation, speed and simplicity are often of greater importance than extreme
accuracy; on the other hand a final solution must be as exact as circumstances allow.

Spar cap/flange boom web


Spar

Skin

Stringers Spar web

(a) Typical wing section (b) Idealised wing section

Figure 4.1: Web boom representation of a complex wing cross section

Complex structural sections may be idealised into simpler forms which behave, under given loading
conditions, in the same, or very nearly the same, way as the actual structure. Different models of the same
structure are required however, to simulate actual behaviour under different systems of loading.
In the wing section of Figure 4.1(a) the stringers and spar flanges have small cross-sectional dimensions
compared with the complete section. Thus, the variation in stress over the cross-section of a stringer due
to, say, bending of the wing would be small. Furthermore, the difference between the distances of the
stringer centroids and the adjacent skin from the wing section axis is small. It would be reasonable to
assume therefore that the direct stress is constant over the stringer cross-sections. Thus we could replace
the stringers and spar flanges by concentrations of area, known as booms, over which the direct stress is
constant and which are located along the mid-line of the skin, as shown in Figure 4.1(b). The stringers and
spar flanges carry most of the direct axial stresses while the skin is mainly effective in resisting shear
stresses although it may also carry some of the direct stresses.
The idealisation shown in Figure 4.1(b) may therefore be taken a stage further by assuming that all direct
stresses are carried by the booms while the skin (web) is effective only in shear. The direct stress carrying
capacity of the skin may be allowed for by increasing each boom area by an area equivalent to the direct
stress carrying capacity of the adjacent skin panels. The calculation of these equivalent areas will
generally depend upon an initial assumption as to the distribution of direct stress in a boom/skin panel.

14
Analysis of thin walled aerospace structures

4.2 Idealisation of structures with a linear stress variation in the skin section

Figure 4.2: Web-boom idealisation

Consider idealising the structure shown in Figure 4.2(a) into a combination of direct stress carrying booms
and shear stress only carrying skin as shown in Figure 4.2(b). In Figure 4.2(a) the direct stress carrying
thickness tD of the skin is equal to its actual thickness t while in Figure 4.2(b) tD = 0. Suppose also that the
direct stress distribution in the actual panel varies linearly from an unknown value 1 to an unknown value
σ

2. Clearly the simplified analysis should predict the extremes of stress 1 and 2 although the distribution
σ σ σ

of direct stress is sacrificed. Since the loading producing the direct stresses in the actual and idealised
panels must be the same we can equate moments to obtain expressions for the boom areas B1 and B2.
Thus, taking moments about the right-hand edge of each panel.
b2 1
+ (σ 1 − σ 2 )t D b
2b
∑M 2 = σ 2t D
2 2 3
= σ 1B1b

...
... (5)
...
t Db  σ 
⇒ B1 =  2 + 2 
6  σ1 
and similarly:
b2 1
+ (σ 1 − σ 2 )t Db = σ 2 B2b
b
∑ M 1 = σ 2t D
2 2 3
(6)
t b σ 
⇒ B2 = D  2 + 1 
6  σ2 
The direct stress distribution in Figure 4.2(a) is caused by a combination of axial load and bending
moment. For axial load only 1/ 2 = 1 and B1 = B2 = tDb/2; for a pure bending moment about the mid
σ σ

plane of the panel l / 2 = -1 and B1 = B2 = tDb/6. Thus, different idealizations of the same structure are
σ σ

required for different loading conditions.


Equations (5) and (6) may be written more generally as:

tabbab  σ 
Ba =  2 + b  (7)
6  σa 
where the a and b subscripts indicate the end of the skin section being idealised.

15
Analysis of thin walled aerospace structures

4.2.1 Example: Idealisation of a wing section (Megson example 9.11)


Part of a wing section is in the form of the two-cell box shown in Figure 4.3(a) in which the vertical spars
are connected to the wing skin through angle sections all having a cross-sectional area of 300mm2. Idealise
the section into an arrangement of direct stress carrying booms and shear stress only carrying panels
suitable for resisting bending moments about the horizontal axis. Position the booms at the spar/skin
junctions.

Figure 4.3: Two cell box-beam idealisation of a wing section

The idealised section is shown in Figure 4.3(b) in which, from symmetry, B1 = B6, B2 = B5, B3 = B4. Since
the section is required to resist bending moments about a horizontal axis the direct stress at any point in
the actual wing section is directly proportional to its distance from the horizontal axis of symmetry as = σ

My/I.
Further, the distribution of direct stress in all the panels will be linear as they are all straight surfaces such
that their distance (y) from the neutral axis (in this case the axis of symmetry) will vary linearly - thus
Equation (7) may be used. Note that in addition to contributions from adjacent panels, the boom areas
include the existing spar flanges. Hence:

t16b16  σ  t b  σ 
B1 = sc1 +  2 + 6  + 12 12  2 + 2 
6  σ1  6  σ1 
where sc1 is the area of the spar cap at point 1. The stress ratios may be calculated from the basic bending
stress equation = My/I. As M and I are constant for the section the stress ratios are simply functions of
σ

the distance of the booms from the neutral axis (remember that material either side of the neutral axis will
have equal but opposite stresses).
3 × 400  − 200  2 × 600  150 
B1 = 300 + 2+ + 2+ 
6  200  6  200 
= 1050mm 2
= B6
Calculate the idealised boom areas B2, B3, B4, B5 (B2 = B5 = 1791.7mm2 and B3 = B4 = 891.7mm2).

16
Analysis of thin walled aerospace structures

4.2.2 Example: Idealisation of a wing section for bending loading about a vertical axis
In the previous example the bending loading is applied about a horizontal axis. If however, the loading is
been applied about the vertical axis, or if the loading had been purely axial, the assumed stress distribution
in the panels of the section would have been different, resulting in different values of boom area.
For the same structure as shown in Figure 4.3 determine the boom areas for bending about a vertical axis.
As structure is symmetric bending will occur about the neutral axis which will be located at the centroid.
First calculate the location of the centroid.
x A = ∑ xab Aab + ∑ xsci sci
x = 524.5mm from skin 16
Then employ Equation (7) as before, eg.:

t32b32  σ  t b  σ 
B3 = sc3 +  2 + 2  + 34 34  2 + 4 
6  σ3  6  σ3 
1.5 × 600  73.5  2 × 200
= 300 + 2+ + (2 + 1)
6  671.5  6
= 816.4mm 2
= B4
Calculate the idealised boom areas B1, B2, B5, B6 (B1 = B6 = 1272.0mm2 and B2 = B5 = 1618.2mm2).
Note how they are different to the values calculated in the previous example. Using the lumped mass/area
model (i.e with boom areas lumped at points and skins having no thickness) re-calculate the position of the
neutral axis – you should find that it is in exactly the same position. That shows that the idealisation
approach maintains the analytical integrity of the structure.

17
Analysis of thin walled aerospace structures

5. BENDING OF ARBITRARY CROSS-SECTION PRISMATIC BEAMS

5.1 Planar bending


5.1.1 Off-axis loading
Consider a beam with a single axis of symmetry (so that the centroid occurs on this axis) in which bending
moments, shear and axial stresses are developed due to an off-axis load F in Figure 5.1. At any section on
the beam the axial and shear forces and bending moment may be determined:

FY F
y y P = FZ
VY = FY
FZ
z C x M X = − FY (L − z )
L
Area B

Figure 5.1: Two dimensional bending induced from off-axis loading

Where P is the axial force, VY the shear force and MX the bending moment about the x axis at any point
along the beam.
From superposition the total axial (z-wise) stress at any position on the beam may then be written as:
P MXy
σ z ( x , y, z ) = + (8)
B IX
This superposition of stress may be shown graphically as in Figure 5.2:

y
y σ

Z = MXyc/IX y y y

σ σ σ
or σ
or σ

Z Z Z Z Z

Z = P/B Z = MXyt/IX
σ σ

P/B < MXyc/IX P/B = MXyc/IX P/B > MXyc/IX


Axial stress due Axial stress due
to axial loading to bending
marks y position of neutral axis

Figure 5.2: Superposition of axial and bending stresses

Note how the position of the neutral axis (axis where axial stress is zero) changes from the centroid (C) to
new position depending on ratio of axial loading stress (P/B) and bending induced stress (MXy/IX).

18
Analysis of thin walled aerospace structures

5.1.2 Eccentric loading


An important case of practical interest occurs when a beam1 is loaded by an eccentric axial load (F) (i.e. a
load not located at the centroid) as shown in Figure 5.3.

y y
F P=F
VY = 0
z C x M X = F.e
e
L
Area B

Figure 5.3: Eccentric loading of a beam

The resultant axial stress distribution is again a superposition of the axial and bending moment induced
stresses:
P M X y F Fey
σz = + = + (9)
B IX B IX
The position of the neutral axis (NA) for a beam subjected to eccentric loading may be found by setting σ

Z
= 0 and is shown in Figure 5.4
F − Fey
=
B IX
(10)
− IX
⇒ y NA =
Be

y
e
F x
C
Neutral
Axis
yNA
Figure 5.4: Position of neutral axis of an eccentric loaded beam

1
How to distinguish between a beam and a column for axial loading: a beam only deflects laterally a small amount under the
influence of an axial load and consequently the bending moment produced due to the eccentricity generated has only a limited
effect on the performance of the beam (short –stocky members). A column deflects a significant amount such that the new
bending moment introduced by the lateral deflection creates significant stresses in the member and leads to column buckling (long
slender members). General rule of thumb is that beams of L/D < 10 don’t have a buckling problem.

19
Analysis of thin walled aerospace structures

5.1.3 Example: Eccentric loading of an engine hanger


A jet engine in a fighter aircraft is often hung from the fuselage on bars known as engine hangers. Such
hangers may have non-uniform cross-sections to accommodate electrical and fuel connections to the
engine. Consider the simplified diagram of the hanger shown in Figure 5.5.
Determine the maximum tensile and compressive stresses in this hanger due to an engine load of P. Also
calculate the position of the neutral axis in the reduced cross-sectional area of the hanger.

b/2 b/2
Equivalent loading of reduced cross sectional area
L

F F
D D
Mx
P=F
VY = 0
z F
Fb
MX =
4
b/2
Mx

F F
b y

x
Section DD

Figure 5.5: Loading of an aircraft engine hanger

b/6
Neutral
2
axis
σ

= -4F/b
y
-ve
compression
+ve
tension
σ

= 8F/b2

x, σ

Figure 5.6: Stress distribution in engine hanger

20
Analysis of thin walled aerospace structures

5.2 Multi-plane bending


5.2.1 Eccentric loading
If the point of application of F is not on one of the principal axes of the cross section there will be
simultaneous bending about both centroidal axes. In this case the axial stress may be obtained from
superposition of the stresses induced by both bending moments and the axial stress.
P M X y MY x
σz = + −
B IX IY
(11)
F Fe y Fe x
= + Y + X
B IX IY
Note that the effect of MY is subtracted as for a positive y and x location a positive MX will produce
tension while a positive MY will produce tension.
The location of the Neutral Axis may be found by setting σ

Z =0
Be X Be
0 = 1+ x+ Y y (12)
IY IX

y
eX
Neutral eY P=F
Axis F x
C VY = 0
M X = F .eY
M Y = − F .eX
Figure 5.7: Position of neutral axis of a doubly eccentric loaded beam

Equation (12) shows that the neutral axis is linear but may or may not intercept the cross-sectional area
(depending on value of loads, eccentricities and second moments of area).

21
Analysis of thin walled aerospace structures

5.2.2 Skew loading


A skewed load may also cause bending about both principal axes. Consider the simplest case of a doubly
symmetric cross-section so that skew load may be resolved into each plane of symmetry. To prevent
twisting about z axis, F must act through a point known as the shear centre of the beam. Methods to locate
the shear centre will be presented later however, for the rectangular cross-section shown in Figure 5.8, the
centroid is also shear centre due to double symmetry.

P=0
VX = F sin β
y VY = − F cos β y
M X = F cos β(L − z ) M
MY
x M Y = F sin β(L − z )
θ

x
β

Fsin
MX
β
z
Neutral Axis
β

F
Fcos

Figure 5.8: Skew loading of a doubly symmetric beam

As there are no axial loads the axial stress in the beam is due only to the two axis bending.
M X y MY x
σZ = − (13)
IX IY
Note second term is negative as where x is positive a positive MY would create a compressive stress.
Again the position of the neutral axis may be found by setting Z = 0
σ

M X y MY x
− =0
IX IY
(14)
y M I I
⇒ tanθ = = Y X = tan β X
x M X IY IY
θ

The neutral axis is found to be a straight line passing through centroid at angle . This also shows that
β θ

unless IX = IY, ≠. That is, the neutral axis is not perpendicular to the plane of the applied load. This
highlights a stability problem with beams of high IX/IY ratio.

22
Analysis of thin walled aerospace structures

5.2.3 Example: Stability problem due to skew loading


The thin I beam shown in Figure 5.9 is to be used as a 4m long cantilever to support a vertical end force of
–40kN.
(1) Determine the maximum stress in the beam.
(2) If the vertical end load is misaligned by -1º recalculate the maximum stress and comment on the
results.

y Typ.0.005m

0.6m
x

0.2m

Figure 5.9: Cantilever I beam cross-section

23
Analysis of thin walled aerospace structures

5.3 Combined bending of an arbitrary prismatic beam


5.3.1 Introduction
A condition of pure bending results when the beam is loaded by only by couples MX, and MY applied at
the end sections (Figure 5.10). The cross-sections of the beam are assumed to remain plane and rotate only
with respect to neutral axes which pass through the centroids of cross-sections.
The development of the general expression for combined bending of a beam is more complex than those
problems previously investigated as the locations of the principal axes are unknown. This would suggest
that the general problem is simply a matter of finding the principal axes and then utilising the simple
theory of bending (Equation (13)). This method however involves a large amount of mathematical
manipulation and a better approach is to derive a general expression about a set of arbitrary axes with their
origin at the centroid of the cross section. Two separate approaches are presented to assist understanding
with how the general equation is derived. The first derives the general expression using a thin walled beam
(section 5.3.2) while the second uses a solid beam (section5.3.3) – both producing the same general
equation.
5.3.2 Derivation of general bending expression from superposition
y, v

MY

MX x, u

MY

C MX
z

Figure 5.10: Pure bending of a beam


Initially assume the beam is a thin-walled section in a condition of pure bending with the x-axis as the
neutral axis. Then the axial-stress distribution across cross-sections is given by:
E
σZ = y (15)
rX
where rX is the radius of curvature in the yz-plane (perpendicular to the x axis).
The couples MX and MY necessary to produce this condition are equal to the components of the stress-
resultant acting on the cross-section. Thus
E E E
M X = ∫ σ Z yt ds = ∫ y t ds = r ∫ y
2 2
dA = IX
rX X rX
(16)
E E E
M Y = − ∫ σ Z xt ds = − ∫ xyt ds = − rX ∫ xy dA = − rX I XY
rX
where IXY is the product second moment of area of the cross section.
The couples MX and MY necessary to cause pure bending with the y-axis as the neutral axis are calculated
in a similar manner. Thus in this case:
E
σZ = − x (17)
rY

24
Analysis of thin walled aerospace structures

and
E E
M X = ∫ σ Z yt ds = − ∫ xyt ds = − I XY
rY rY
(18)
E E
M Y = − ∫ σ Z xt ds = ∫ x 2 t ds = I Y
rY rY
By superposition of these two cases the curvatures (1/rX) and (1/rY), and hence the stresses Z, are σ

calculated for the general case in which the beam is loaded by couples MX and MY acting simultaneously.
Thus:
I I 
M X = E  X − XY 
 rX rY 
(19)
I I 
M Y = E  Y − XY 
 rY rX 
Solving these equations for the curvatures (1/rX) and (1/rY) gives
 I   I 
I X I Y  M X + M Y XY   M X + M Y XY 
1
=  IY = 1  IY  = M ′X
rX (
EI X I X I Y − I XY
2
) EI X K EI X
(20)
 I   I 
I X I Y  M Y + M X XY   M Y + M X XY 
1
=  IX = 1  IX  = M Y′
rY (
EI Y I X I Y − I XY
2
) EI Y K EI Y
where M'X and M'Y are the effective bending moments for the chosen x, y axes, which in general are not
the principal axes of the section. The effective bending moments can be interpreted physically as the
combinations of MX and MY which produce pure bending strains about the x- and y-axis respectively. The
axial-stress distribution is obtained by substituting these expressions for the curvatures into Equations (15)
and (17). Thus:

 y x M′ M′  M X I Y + M Y I XY   M Y I X + M X I XY 
σ Z = E  −  = X y − Y x =   y −   x (21)
 rX rY  I X IY  I X IY − I XY
2
  I X I Y − I XY
2

Equation (21) is the general expression for calculating axial stresses in a cross-section of arbitrary shape,
due to couples MX and MY applied about an arbitrarily-orientated set of rectangular axes x and y which
pass through the centroid of the cross-section.
In the particular case where x and y are principal axes i.e., IXY = 0 (Note: axes of symmetry are principal
axes), which gives K=1, M'X = MX, M'Y = MY and Equation (21) then becomes:
MX M
σz = y− Y x (22)
IX IY
Equation (22) shows that the elementary bending theory can be applied to each principal axis in turn and
superposed to obtain the resultant stress. There is always a set of principal axes for an unsymmetrical
cross-section about which the product second moment of area equals zero. If these axes are used for the x
and y axes then the simplified version of the bending equation may be used. Normally it is simpler to use
the full version of the bending equation and a convenient set of axes than to determine the position of the
principal axes and calculate all dimensions in terms of the co-ordinates from these axes.

25
Analysis of thin walled aerospace structures

5.3.3 Alternate derivation with an arbitrary neutral axis


Consider a solid beam with an arbitrary cross-section as shown in Figure 5.11. The beam supports end
moments MX and MY as shown in Figure 5.10 and bends about the neutral axis (NA) where the direct axial
stress is zero. Assume initially, that the origin of the x and y axes (the centroid of the section) is some
undefined distance from the neutral axis. The stress on an element of area dB at a point (x, y) and
perpendicular distance p from the neutral axis is related to the strain at that point by Hooke’s law i.e., Z =
σ

E Z, where Z is the axial strain at point (x,y).


ε ε

Y dB
A
y p

N C
x X

Figure 5.11: Arbitrary location of the neutral axis for pure bending

If the beam is bent to a curvature rNA about the neutral axis simple bending theory gives:
Mp Ep
σz = = = Eε z (23)
I NA rNA
where INA is the second moment of area about the neutral axis.
As the beam is considered only to support bending moments the sum of the axial stresses over the cross-
sectional face must be zero, therefore:
Ep
0 = ∫ σ z dB = ∫ dB = ∫ pdB (24)
rNA
i.e., the first moment of area of the cross section of the beam about the neutral axis is zero – therefore the
neutral axis must pass through the centroid (this is a definition of the centroid; that the first moment of
area about any axis passing through the centroid is equal to zero – check first year notes). Thus Figure
5.11 may be redrawn with the neutral axis passing through the centroid. Note that this is only valid for
pure bending.
Y dB
A
y p
θ

C
N x X

Figure 5.12: Location of neutral axis for an arbitrary cross-section

26
Analysis of thin walled aerospace structures

From geometry the distance p may be replaced by x and y terms and hence the stress at any point
expressed in terms of its location relative to the centroid (C).
p = y cosθ − x sinθ

σz =
Ep E
= ( y cosθ − x sinθ ) (25)
rNA rNA
The moment about the x and y axes may then be determined by integrating the stress distribution across
the face:

∫ (y ) (I X cosθ − I XY sin θ )
E E
M X = ∫ σ Z y dB = 2
cosθ − xy sin θ dB =
rNA rNA
(26)
∫( )(I XY cosθ − I Y sin θ )
E E
M Y = − ∫ σ Z x dB = − xy cosθ − x sin θ dB = −
2

rNA rNA
θ θ

These two equations may be rearranged to make cos and sin the subjects:

rNA  M X I Y + M Y I XY 
cosθ =  
E  I X I Y − I XY
2

(27)
r  M I + M X I XY 
sinθ = NA  Y X 
E  I X I Y − I XY
2

which may then be substituted into Equation (25) to obtain the general expression for bending of a
prismatic beam:
 M X I Y + M Y I XY  M Y I X + M X I XY 
σ z =   y −   x
 I X I Y − I XY   I X I Y − I XY 
2 2

(28)
 I .y − I XY .x   I .x − I XY .y 
= M X  Y  − M Y  X 
 I X I Y − I XY   I X I Y − I XY 
2 2

which is exactly the same as the expression derived in Equation (21).


5.3.4 Equation of Neutral Axis
The neutral axis is defined by the condition σ

Z = 0. Thus, from Equation (21);


 M X I Y + M Y I XY   M I + M X I XY 
σ Z = 0 =   y NA −  Y X  x NA
 I X I Y − I XY −
2 2
  I I
X Y I XY 
= (M X I Y + M Y I XY ) y NA − (M Y I X + M X I XY )x NA (29)
y NA M I + M X I XY
= tanθ = Y X
x NA M X I Y + M Y I XY
θ θ

Again the neutral axis will be a straight line orientated at some angle to the x axis where is measured
from the positive x-axis and is positive in accordance with the right-hand screw rule.

27
Analysis of thin walled aerospace structures

5.3.5 Example - Effective section maximum stress


The effective section shown below is subject to a bending moment MX =1.5kNm. Determine the
maximum axial stress and the location at which it acts. (Note: thin wall approximations are not valid as
thickness is not significantly less than length.)

y
40mm 80mm

x y
8mm
80mm C x

8mm
Figure 5.13: Maximum stress in a T-section

σ Z = 1.5 y − 0.39 x
maximum occurs at x = −8mm and y = −66.4mm
σ Z = −96MPa (compressive)

28
Analysis of thin walled aerospace structures

5.3.6 Example – Idealised Section axial stress distribution.


Calculate the axial-stress distribution and the position of the neutral axis for the idealised section shown in
Figure 5.14, due to applied bending moments MX = -2MNm and MY = 1MNm.

x
Idealised booms have
0.25m
an area of 0.01m2 F

B
C
1m
y
0.5m

A D

1m 2m

Figure 5.14: Cross-section for example 0

The section characteristics with respect to the xy-axes passing through the centroid C are: IX = 0.0l m4, IY
= 0.0475 m4, IXY = 0.015 m4. Substituting these values into Equation (21) gives:
 M X I Y + M Y I XY
  M Y I X + M X I XY 
σ z =   y −   x
 I X I Y − I XY   I X I Y − I XY 
2 2

 − 2 × 0.0475 + 1× 0.015   1× 0.01 + −2 × 0.015 


= y − x
 0.01× 0.0475 − 0.0152   0.01× 0.0475 − 0.0152 
   
−2
= −320 y + 80 x MNm
The axial stresses at A, B, D, and F are obtained by substituting the co-ordinates of the elements into this
expression for Z.σ

Element x y σ

Z
(m) (m) (MPa)
A -1.25 -0.5 60
B -0.25 0.5 -180
D -0.25 -0.5 140
F 1.75 0.5 -20
The equation of the neutral axis is obtained from Equation (29).
y NA M I + M X I XY
= tanθ = Y X
x NA M X I Y + M Y I XY
1× 0.01 + −2 × 0.015
=
− 2 × 0.0475 + 1× 0.015
= 0.025
θ

Hence = 14º 2', 194º 2'

29
Analysis of thin walled aerospace structures

5.3.7 Example – Effective section axial stress distribution (Example 9.3 in Megson)
Determine the stress distribution for the Z section (typical of a stiffener in an aircraft fuselage) due to a
bending moment MX.

y h/2
4
3
t

h C
x

2
1 h/2

Figure 5.15: Cross-section of Z section stiffener

y -1.72MX/(h3t)
Compression
3
4
+3.43MX/(h3t)
Tension

C
σ
x
Z

-3.43MX/(h3t)
Compression
1
2
+1.72MX/(h3t)
Tension

Figure 5.16: Solution for Example 5.3.7

30
Analysis of thin walled aerospace structures

TUTORIAL SET 1: STRUCTURAL IDEALISATION & AXIAL STRESS DUE TO BENDING

(1) A wing section is shown in Fig Q.1. All dimensions are in mm and L section web caps have a
cross-sectional area of 100mm2 and the T sections caps a cross-sectional area of 200mm2.
(a) Idealise the section for bending about a horizontal axis.
(b) Idealise the section for axial loading.
Answer:

200
300
6mm
200 All outer
walls 2mm 200
150

50
350 500

Figure Q.1 & Q.12

(2) A vertical aluminium pole AB is fixed at the


base and pulled at the top by a cable having a
tensile force T acting at an angle = 25º to
d D
α

the pole. The cable is attached at the outer


surface of the pole. The pole has length
L=2.0m and a hollow circular cross-section A
with outer diameter D=260mm and inner
diameter d=200mm. Determine the allowable
force T in the cable if the allowable α

L
compressive stress in the aluminium pole is
90MPa.
Answer: (99.9kN) B

Figure Q.2

(3) A frame ABC is formed by welding two steel


P
pipes at B. Each pipe has a cross-sectional
area of 10.4×103mm2, second moment of area B
I = 88.2×106mm4 and outside diameter D = D
270mm. Find the maximum tensile and H
compressive stresses in the frame due to a
A C
load P = 14kN if L=2.4m and H=1.8m
Answers: (25.3MPa, -26.1MPa)

L L

Figure Q.3

31
Analysis of thin walled aerospace structures

(4) A circular, cylindrical tower having height h,


inside diameter d and outside diameter D
begins to lean slightly. What is the maximum d D
permissible angle of inclination from the
α

vertical if there is to be no tension in the α

tower? (Consider that the only load is the


uniformly distributed weight of the tower
itself and assume that the angle is very
α

small.) h
Answer: ( = sin-1 [(D2 + d2)/(4hD)])
α

Figure Q.4

(5) A solid bar of circular cross-section is subjected to an axial tensile force T = 26kN and a bending
moment M=3.2kNm. Based upon an allowable stress in tension of 120MPa, determine the
required diameter D of the bar (disregard the weight of the bar).
Answer: (66.2mm)
(6) The cantilever beam shown has a rectangular cross-section of height h. When the loads P have
the directions as shown, the neutral axis is always located above the centroid of the cross-section.
Let s represent the distance from the x axis to the neutral axis. (a) Obtain an expression for s as a
function of the distance z from the fixed support, disregarding the weight of the beam and (b) plot
a graph showing s versus z for the case where L=1m and h = 120mm.
Answer: (s = 1.2/(1-z))

P
y y s
P N A
h z x
C

Figure Q.6

(7) A cantilever beam of rectangular cross-section


supports an inclined load P at its free end. y
Calculate the maximum tensile stress due to the
load with b = 75mm, h=150mm, beam length =
θ

1.4m, P = 800N and = 30º.


Answer: (7.43MPa)
h C
x

b P

Figure Q.7

32
Analysis of thin walled aerospace structures

(8) An I beam with flanges 12in. wide and an overall


depth of 18in. is used as the main spar for a wing.
At one section of the spar the resultant line of
action of the load is at a point one inch from either
axis of symmetry. Find the position relative (to P
the centroid of the cross-section) of the line of
1”
18”
zero stress on the section given that the principal C
1”
second moments of area for the spar cross-section
are 1933 and 262in.4 and the cross-sectional area
is 34in.2.
Answer: (56.85in below C at an angle of -82.28º)
12”
Figure Q.8

(9) For the simplified section shown calculate the


position of the principal axes (about which Ixy = 0)
and find the values of the principal second
t = 0.5”
moments of area.
Answers: (34.607º, 28.61in4, 3.34in4) 5” (constant)
Not straight forward – some thinking required.

(10) Calculate the maximum axial stress in the section 6.5”


and state the point at which it acts when a bending
moment of 2000lbf.ft is applied about the
Figure Q.9 & Q.10
horizontal axis.
Answer:

(11) A 3000Nm bending moment is applied to a beam


with the right angle cross section shown.
Calculate the maximum axial stress in the beam 100mm
and state the point at which it acts.
Answer: (-63.3MPa at bottom LH corner)

125mm
C 30º

3000Nm

Thickness = 10mm

Figure Q.11

(12) Calculate the maximum axial stress in the wing section shown in Fig Q.1 and state the point at
which it acts when combination of a 2kNm bending moment about the horizontal and an axial
load of 5kN are applied.
Answer:

33
Analysis of thin walled aerospace structures

6. TORSION OF MULTICELL SECTIONS

6.1 Torsion of a Single-Cell Closed Section


6.1.1 Batho Equation
The torsion of a single-cell section is statically determinate and the structural analysis of such a section is
relatively easy provided:
(1) The cross-section does not vary along its length (i.e. along the z axis).
(2) The cross-section is closed.
(3) The wall thickness is small compared to the cross-sectional dimensions.
(4) The member is subjected to end torques only.
(5) The ends are not restrained from warping.
The analysis of closed sections may also be extended to consider the torsion and warping of open sections.

y
ds

s T
s=0 dz

T
z

Figure 6.1: Single closed cell under pure torsion (about its shear centre)

Consider a typical closed, thin walled torsion member. The thickness (t) may vary with circumferential
position but may not vary along the length (z direction). As the wall thickness is thin compared to other
dimensions, the shear stress is assumed to be parallel to the median line of the thickness and, provided that
the wall remains in the elastic region, constant across the thickness.
Introduce Shear Flow (q) = Shear Force per unit length of surface.
q =τ t (30)

Figure 6.2: Equivalence of shear stress and shear flow

34
Analysis of thin walled aerospace structures

The use of shear flow eliminates the need to consider variation of thickness in the s direction. Shear flow
is considered positive in the direction of increasing s.
∂q
q+ ds q
∂s ds

dz
q
s
∂q
q+ dz
∂z
z

Figure 6.3: Closed cell skin element under pure torsion loading

Consider the small element of the torsion member as shown in Figure 6.3. As there are no axial stresses
for the element to be in equilibrium:
 ∂q  ∂q
∑F z =  q + ds  dz − qdz = 0 ⇒
 ∂s  ∂s
=0 (31)

 ∂q  ∂q
∑F s =  q + dz  ds − qds = 0 ⇒
 ∂z  ∂z
=0 (32)

From Equation (31) the change in shear flow in the s direction is zero i.e., the shear flow must be constant
around a closed section → q = .t = constant. Now, consider the cross-section of a thin-walled tube as
τ

shown in Figure 6.4.


s
ds

q.ds
t

p dA = ½p.ds
O
A = Total area enclosed
by median line

Figure 6.4: Derivation of Batho equation

The elemental torque (dT) resisted by a small element of the tube (ds) about any point O is:
dT = pq.ds (33)
where p is the perpendicular distance from point O to the tangent of the median line of ds. Total torque
therefore is:

T = ∫ pq .ds = q ∫ p .ds = 2q ∫ dA = 2qA


(34)
= 2τ t A
This is known as the Batho equation.

35
Analysis of thin walled aerospace structures

6.1.2 Bredt- Batho expression for rate of twist


The theory associated with the torsion of closed section is known as the Bredt-Batho theory. Closed
arbitrary cross-section tubes will, in general, both warp and twist under torsion. Consider the deformation
of the element of Figure 6.5 when subjected to torsion.
dw
γ

ds
r.d φ
dz γ

d φ

r Undistorted
surface element

Distorted surface element

O
Figure 6.5: Derivation of Bredt-Batho expression for rate of twist
Twist (rotation about the z axis) produces twist shear strain:
r .dϕ
γt ≈ (35)
dz
Warping (translation parallel to the z axis) produces warping shear strain:
dw
γw ≈ (36)
ds
Total shear strain in a surface element:
r .dϕ dw τ q T
γ =γt +γ w = + = = =
dz ds G Gt 2 AGt
(37)
T r .dϕ
⇒ ds = ds + dw
2 AGt dz
where d /dz is the rate of twist of the tube. Integrating Equation (37) around the section gives:
φ

1 Tds dϕ dϕ

2 A Gt
=
dz ∫ r .ds + ∫ dw =
dz
2A

dϕ 1 Tds
dz 4 A 2 ∫ Gt
=

1 qds
2 A ∫ Gt
= (38)

=
T
{ if G is constant around the section }
GJ
where J is known as the torsional constant (equal to polar second moment of area for circular cross
sections) and is equal to:
4 A2
J= (39)
ds
∫t

36
Analysis of thin walled aerospace structures

6.1.3 Example: Torsion of a closed section


Calculate the shear stress and rate of twist for the thin-walled steel tube (G = 76GPa) shown below when it
is loaded by a 10kNm torque.

8mm

50mm

100mm

Figure 6.6: Thin walled steel tube

max = 35MPa
θ

= 0.38º/m

37
Analysis of thin walled aerospace structures

6.2 Torsional of an open section


6.2.1 Torsion constant (J)
A thin wall open cross-section may be considered to consist of a series of small rectangular elements.
Each of these elements can be considered to consist of a series of thin-walled sections. The enclosed area
within these thin walled sections is dAn:
dAn = 2n(b − t + 2n ) (40)

dAn = (b-t+2n).(2n)

b
dn

t/2 - n

n
t/2 - n

Figure 6.7: Torsional constant for an open section

The inverse thickness integral of the generalised thin walled section is:
ds 2(b − t + 2n ) 2n
∫ t
=
dn
+
dn
(41)

Therefore the torsional constant J may be determined by substituting this expression into Equation (39):

4(dAn ) 16n 2 (b − t + 2n )
t t
2 2
2 2
bt 3
J=∫ =∫ dn ≈ when b>>t (42)
ds 2(b − t + 2n ) + 2n 3
0
∫t 0

For open thin walled structures with a number of ‘lengths’:


bt 3
J ≈∑ (43)
3

38
Analysis of thin walled aerospace structures

6.2.2 Maximum shear stress in open sections


The solution of the problem of torsion of bars with a uniform, general shaped cross section was given by
Saint-Venant in 1855. He used the so-called semi-inverse method. That is, at the start he made certain
assumptions as to the deformation of the twisted bar and showed that with these assumptions he could
satisfy the equations of equilibrium and the boundary conditions. Then by applying the uniqueness of
solutions of the elasticity equations he showed that the assumptions made at the start were correct and the
solution obtained was the exact solution of the torsion problem, provided that the torques on the ends are
applied as shear stresses in exactly the manner required by the solution itself (For further details, read
Timoshenko & Goodier, Theory of Elasticity, Chapter 10, or Megson, Aircraft Structures for Engineering
Students Chapter 3).
A sketch of the shear stress distribution for rectangular sections subject to torsion is shown in Figure 6.8.
Note the difference with the circular section also shown in Figure 6.8. The magnitude of the shear stress
on the circular cross section varies proportionally with the distance from the centre and reaches maximum
at the outer surface. In contrast, the shear stress at the corner of the rectangular torsion member (usually
farthest from the centre) approaches zero. The maximum shear stress on a rectangular cross section
actually occurs at the middle of the longer edge, which is the point on the periphery of the cross section
that is nearest the centre. The maximum shear stress in a rectangular prismatic bar subjected to torsion
may be expressed (according to Saint-Venant's method) in the form:
T
τ max = (44)
αbt 2
where is a dimensionless constant obtained by a theory of elasticity solution, and b/t 1. If b >> t (i.e.

the rectangular bar becomes a thin walled open section) so that the maximum stress in a thin walled
α ≈ ⅓

open section may be approximated by:


T Tt
τ max = = (45)
bt 2 J
3
For values of b/t = 10 or greater, the error in using this approximation is less than 6%.

T T

max
t

Figure 6.8: Shear flow/stress for rectangular members

39
Analysis of thin walled aerospace structures

6.2.3 Example: Maximum shear stress in an open section (I beam) due to torsion.
An I-beam has a width of 100mm and depth of 150mm as shown in Figure 6.9. Calculate the maximum
torque that could be applied to this beam if the yield stress of the material the beam is made from is
240MPa. Assume a safety factor of 3.

6mm

138mm

6mm
T

6mm

100mm

Figure 6.9: Torsion of an I-beam

Tmax = 324.5Nm

40
Analysis of thin walled aerospace structures

6.3 Torsion of Multi-cell sections (Statically Indeterminate)


6.3.1 Order of Redundancy and Choice of Variables
Since the single-cell section is statically determinate, it follows that the n-cell section has (n-1) degrees of
redundancy. The redundant forces can be taken either as
(1) the shear flows in the interior walls, Figure 6.10(1), or
(2) the shear flows in the outer walls Figure 6.10(2).

qi,0 (1)

(2)

qi,0

Figure 6.10: Redundancy choices in multi cell sections

In the case of choice (1), the influence of each redundancy extends over the whole cross-section and the
computations become laborious when the number of cells is large. In general, it is better to use choice (2)
because in this case each redundancy stresses only the cell in which it is applied. Consequently there is a
minimum cross-coupling between the redundancies. With this choice, regardless of the number of cells, at
the most only three unknowns are involved in each of the equations expressing the solution and these
unknowns are banded about the principal diagonal. This gives well-conditioned equations as shown by
Equation (50) and illustrated by the four-cell section of Example 6.3.6.
Also, although the section has (n-1) static redundancies, the analysis can be made using n variables, that
is, by treating the shear flow in the outer wall of each of the n cells as an initially unknown quantity
("redundancy"). This approach permits both the stress distribution and the torsion constant J to be
evaluated in one calculation in a simple manner and is illustrated in Example 6.3.6.
6.3.2 Solution of multicell section -solution involving (n - 1) redundancies
The shear flow in the n-cell section (Figure 6.11) is expressed in terms of the n constant value shear flows
q1, ... , qn representing the shear flows in the outer walls of the section. These shear flows are the
independent variables of the problem. The dependent variables are the shear flows in the interior walls,
which are determined by the equilibrium equations at the junctions of the interior walls with the outer
walls of the section. Thus, the shear flow in the interior wall between cell (i - 1) and cell i is qi-1,i = qi-1 - qi
(i = 2, .., n). The positive sense of qi-1,i is determined by considering the interior wall between cell (i-1) and
cell i to be part of cell (i-1), following the general rule that anticlockwise shears are positive. In the left
outer wall (cell 1) a positive shear is down, but in the right outer wall (cell n) a positive shear is up.

41
Analysis of thin walled aerospace structures

Figure 6.11: Nomenclature for a multi cell section

The equation of equilibrium gives:


n
T = 2∑ qi Ai (46)
1

where Ai is the area enclosed by cell i.


Since the stress distribution is expressed in terms of n independent variables, but only one equilibrium
equation is available (i.e. the net effect of all the shear flows has to be equivalent to the applied torque),
the section is statically indeterminate to the (n-1)th degree. The additional (n - 1) equations required for the
solution are obtained from the compatibility equations that state that the rate of twist in all of the cells
must be the same:
θ1 = θ 2 = ,...,= θ n (47)

where
dϕ 1 qds
θi = = ∫ (48)
dz 2 Ai G i t
is the rate of twist of cell i.
Equations (46) and (47) may be solved for q1, …, qn. The shear flows in the interior walls are obtained by
substituting these values into the equilibrium condition for these walls i.e., qi-1,i = qi-1 - qi.
θ

To calculate the torsion constant J of the section, firstly the rate of twist is calculated by substituting the
θ

values of qi thus obtained into any one of the i of Equation (47). J is then obtained by substituting this
value of into Equation (38).
Note: for sections where G is not constant only an effective torsional rigidity may be determined i.e., GJ.

42
Analysis of thin walled aerospace structures

6.3.3 Example: Shear flow in an idealised section


Calculate the shear-flow distribution in the idealised section of Figure 5.14 due to a torsion T = l MNm.
Also calculate the torsion constant J of the section. Assume G is constant for the section.

x
Idealised booms have
an area of 0.01m2 0.25m
F
B q2 F
B
C A2
1m q1
y q1,2
0.5m A1 q2
A D
A
1m 2m q1 D

Figure 6.12: Shear flow in an idealised section subject to a torque

Denote the cells ADB and BDF by the subscripts 1 and 2 respectively.
The independent variables are the shear flows q1 and q2 in the outer walls of cells 1 and 2.
√ √

A1 = 0. 5m2, A2 = 1m2, AB = 2m, DF = 5m


n
T = 2∑ qi Ai
1

10 = q1 + 2 q2
6

θ1 = θ 2
from Equation (48)
1 qds 1 qds

2 A1G 1 t
= ∫
2 A1G 2 t
1
1
(( ) ) ((
1 + 2 q1 + (q1 − q 2 ) = 2 + 5 q 2 − (q1 − q 2 )
1
2
) )
Solving these two simultaneous equations for q1 and q2 gives
q1 = 316 kN/m, q2 = 342 kN/m
Hence the shear flow in the interior wall, q1,2 = -26 kN/m, the minus sign indicating that it acts in the
negative sense, that is, in the sense BD.
θ

The rate of twist due to T is obtained by substituting these values of q into either cell. Using cell 1,

θ1 =
1

2 Ai G t
i
qds
=
1
2 × 0.5 × G × 0.005
( (
316 1 + 2 − 26 × 1 ) )
147.36 × 10 6
=
G
which may then be used to estimate the torsion constant for the section:
T
J= = 0.006786m 4

43
Analysis of thin walled aerospace structures

6.3.4 Solutions involving n cells


The shear flow throughout the section is expressed in terms of the n redundancies q1, …, qi, …, qn , where
qi is the shear flow in the outer walls of cell i. The independent variables are qi and the shear-flow
distribution is as shown in Figure 6.11.
The rate of twist of cell i is given by
1 qds
θi = ∫
2 Ai G t
i
(49)
1  ds ds ds 
=  − qi −1 ∫i−1,i + qi ∫i − qi +1 ∫i ,i +1 
2 Ai G  t t t 
where the first and last integrals are only over the shared edges of cells i-1, i and i+1. This may be re-
arranged into a format amenable to solution by matrix methods
ds ds ds
− qi −1 ∫ + qi ∫ − qi +1 ∫ = 2GθAi
i −1,i t i t i ,i +1 t
 ∫1 dst − ∫2 ,1 dst   q1   A1 
   A 
 − ∫1,2 t
ds
∫2 t
ds
− ∫2 ,3 tds
  q2  (50)
 2
 ... ... ...   ...  = 2Gθ  ... 
 ds     
∫i −1,i t ∫i t ∫i+1,1 t   qi 
ds ds
  Ai 
 ∫n−1,n t
ds ds  
∫n t   qn   An 

This shows that, regardless of the number of cells, at the most only three unknowns are involved in each
equation. Also, these unknowns are banded about the principal diagonal and the coefficients of the
principal diagonal are predominant. Thus the equations are well conditioned and direct solution by
elimination is possible even for a large number of redundancies. These characteristics of the solution
equation are a consequence of the choice of redundancies: this choice minimises cross- coupling.
6.3.5 Effect of warping
In the preceding sections it has been assumed that the torque is applied at the ends of the member and that
all sections are free to warp (i.e. there is no axial restraint). Owing to the variation in transverse shear
stress, for example in the flanges of a channel or I-section, there is also a variation in longitudinal
complementary shear stresses, which results in the axial movement of one flange with respect to the other.
Therefore, cross-sections that were initially plane, do not remain so during torsion and there is warping of
any cross section. In practice, however, there are often cases where one or more sections of a member are
constrained in some way so that cross sections remain plane, i.e. warping is prevented.
The warping constraint has most significance near the ends of the torsion member. It decays with distance
from the constraint so that after several section depths the influence of warping is negligible. Non-circular
closed and solid-section members also exhibit warping, but the effect is much smaller in comparison with
the open section.
Resisting torque in a section is thus supplied in two ways, by simple torsion and also by torque set up
through the restraint of warping. In short, when warping is restrained, angles of twist are generally
reduced and hence torsional stiffness increased.

44
Analysis of thin walled aerospace structures

6.3.6 Example: Shear flow distribution in a multi cell section


Calculate the shear-flow distribution in the four-cell section of Figure 6.13, due to a torsion T = 0.5 MNm.
Also calculate the torsion constant J of the section. Assume that G is constant.

q1 q2 q3 q4

q1 q4
q1,2 q2,3 q3,4

q1 q2 q3 q4

q1,2 q1 q2

Figure 6.13: Shear flow in multi cell section

 ( 20×.005
0.6
+ 0.005
1
+ 0.003
1
) ( 0.−0031 )   q1  0.6
 ( 0.−0031 ) ( 02×.005
1.0
+ 0.003
1
+ 0.002
1
) ( 0.−0051 )    
   q2  = 2Gθ 1.0 
 ( 0.−0051 ) ( 20×.005
0. 4
+ 0.002
1
+ 0.003
1
) ( 0.−0031 )   q3  0.4
     
 ( 0.−0031 ) ( 0.005 + 0.005
2×0.8 1
+ 0.003
1
)  q4  0.8
 773 − 333   q1  0.6
 − 333 1233 − 500  q   
   2  = 2Gθ 1.0 
 − 500 993 − 333  q3  0.4
    
 − 333 853   q4  0.8
 q1  0.001676
q   
 2  = 2Gθ 0.002090
 q3  0.002036
   
 q4  0.001733
n 0.001676
2∑ qi Ai 0.002090
= 4[0.6 1.0 0.4 0.8] 
T 
J= = 1
Gθ Gθ 0.002036
 
 0.001733
= 0.02118m 4
θ

therefore 2G = 2T/J = 47.2MPa so that the shear flow distribution can be calculated to be:
 q1  79
 q  99  q1,2   − 20
 2  =   kN / m and    
 q3  96 q2 ,3  =  3  kN / m
    q3 ,4   14 
 q4  82

45
Analysis of thin walled aerospace structures

TUTORIAL SET 2: TORSION OF THIN WALLED STRUCTURES

(1) The thin walled open section shown is to be loaded by


a 10Nm Torque. Using G = 50GPa, estimate: 28mm
t = 2mm
(a) the torsional constant (J) of the section,
(b) the rate of twist of the section and
(b) the maximum shear stress in the section.
Answers: (1184mm4 , 0.169, ± 33.78MPa) t = 4mm
50mm
t = 2mm

60mm

Figure Q.1

(2) The thin walled open section shown is to be loaded by


a Torque T = 10a. Determine: a
(a) the torsional constant (J) of the section,
(b) the rate of twist of the section and a/2
(b) the maximum shear stress in the section. t
Answers: (5at3/3, 6/(Gt3), ± 6/t2) 2a
Constant
Thickness a/2

Figure Q.2

(3) A simplified wing section is shown in Figure Q.3. the


semi-circular nose section is made from a different
material than the remainder of the wing box. The nose All walls 1.6mm 2mm
100mm thick except as
material has a shear modulus of 22GPa while the
indicated
remainder of the section has a modulus of 27.5GPa. If
the section can be considered to be subjected to a
uniform torque of 4.5kNm determine:
(a) the maximum shear stress in the structure, and 200mm 50mm
(b) the rate of twist for such a loading/section.
Answers: (58.8MPa, 29.3×10-6 rad/mm) Figure Q.3

(4) A simplified wing section is shown in Figure Q.4. If


the wing span is 3m and the wing can be considered to
be subjected to a uniform torque of 11kNm All walls 5mm 4mm
determine: 120mm thick except as
indicated
(a) the maximum shear stress in the structure, and
(b) the twist at the wing tip.
Answers: (85.54MPa, 6.3º)
200mm 80mm

Figure Q.4

46
Analysis of thin walled aerospace structures

(5) A wing is simplified to a thin walled cantilever box


Fuselage
section with the dimensions shown in Figure Q.4. The Uniformly distributed
top and bottom skins are made a different thickness Torque 20Nm/mm
and from different material than the forward and aft
skins. The shear modulus of the upper and lower skins 1.2mm
is 18GPa while for the vertical skins it is 26GPa. The z 2.1mm 2.1mm 250mm
aerodynamic loads may be assumed to produce a
1.2mm
uniformly distributed torque of 20Nm/mm along the 2500mm
length of the wing. Calculate:
1000mm
(a) The maximum shear stress in the wing due to
torsion, and
(b) determine an expression for the distribution of
Figure Q.5
twist along the span of the wing.
Answers: (83.3MPa, =8.14×10-9(2500z - z2/2) rad)
φ

(6) The simple wing box section shown in Figure Q.6 is


used in wing of span 2.5m. If the shear stress must not
exceed 140MPa, find:
(a) the maximum torque that may be applied to the
section.

If that torque is then assumed to be constant along the


span, find:
Figure Q.6
(b) the twist of the wing tip relative to the wing
root and
(c) the torsional rigidity of the box.
Answers (102.417kNm, 1.46º, GJ=10MNm2/rad)
Shaded Areas: A34 = 6450mm2, A16 = 7750mm2
Wall lengths: s34 = 250mm, s16 = 300mm
Wall thickness: t12 = 1.63mm, t34 = 0.56mm, t23 =
t45 = t56 = 0.92mm, t61 = 2.03mm
and t25 = 2.54mm
(7) Determine the shear stress distribution in the walls of
the three cell wing section shown in Figure Q.7 when
subjected to a anticlockwise torque of 11300Nm.
Answer: shown in Figure Q.7(b)
Areas: AI = 0.258m2,
AII = 0.355m2,
AIII = 0.161m2
Wall lengths: s13 = s24 = 0.775m,
s35 = s46 = 0.508m, Figure Q.7(a)
s12(outer) = 1.650m,
s12(inner) = 0.508m,
s34 = 0.380m,
s56 = 0.254m.
Wall thickness: t35 = t46 = t56 = 0.92mm,
t13 = t24 = t12(outer) = 1.220mm,
t12(inner) = 2.03mm,
t34 = 1.63mm.
Figure Q.7(b)

47
Analysis of thin walled aerospace structures

(8) Determine the torsional stiffness of the four cell wing section
shown in Figure Q.7 (Hint: use Mathcad or Matlab to help
with matrix manipulation).
(Answer: 522.5×106G MPa/rad)
Areas: AI = 0.1615m2,
AII = AIII = 0.291m2,
AIV = 0.226m2
Wall lengths: s12 = s78 = 0.762m,
s23 = s67 = s34 = s56 = 0.812m, Figure Q.8
s45(outer) = 1.525m,
s45(inner) = 0.356m,
s36 = 0.406m,
s18 = 0.254m,
s27 = 0.356m.
Wall thickness: t12 = t78 = t23 = t67 = t34 = t56 = t18 = 0.915mm,
t45(outer) = 0.711mm,
t45(inner) = t27 = 1.220mm,
t36 = 1.625mm.

48
Analysis of thin walled aerospace structures

7. BENDING OF PRISMATICAL BEAM BY TRANSVERSE FORCES

7.1 The Shear Centre


Equation (21) gives the axial-stress distribution due to couples MX and MY applied to a general section.
When these couples arise from transverse forces applied to the beam, shearing forces VX and VY are
generated in the sections, and bending together with twisting of the beam occurs in the general case. That
is, the deformation of a cross-section consists of rotations about the x- and the y-axes (bending
deformations) together with a rotation about an axis parallel to the z-axis (twisting).
However, there is a unique point in the section through which the action of the applied shearing forces VX
and VY causes the section to bend but not to twist. This point is called the shear centre (E) of the section,
and in the general case is located by two co-ordinates (xE, yE). A method to determine these co-ordinates
will be discussed in a later lecture.
The reciprocal of this statement is that the shear centre is also the point about which the section will twist
(i.e. centre of twist). Normally shear loads do not act through the shear centre so for analysis purposes
they have to be replaced by a shear load at the shear centre plus a torsional moment. The displacement of
the section may then be found by a superposition of the shear induced deformation and the torsional
induced twist.
The ability to locate the shear-centre is therefore important. Where a cross-section has an axis of
symmetry the shear-centre must lie on that axis so that for cruciform or angle sections the shear-centre is
located at the intersection as
(1) all shear flows must pass through this point and thus
(2) taking moments about this point gives a net torque of zero (i.e. no twist)

E E

Figure 7.1 Shear centres of open sections

7.2 Bending Shear Stress (Shear stresses induced by bending)


The shearing stresses corresponding to VX and VY applied at E generate the bending stresses Z, and for
σ

this reason they are called bending shear stresses. The resultant of the bending shear stresses is statically
equivalent to the shearing force resultant applied at E.
The bending shear stresses are derived from Equation (21) and expressed in terms of the shearing force
components by using the following two linking equations:
(1) The equations of equilibrium relating shearing force and bending moment, and
(2) The differential equations of equilibrium relating shearing stress and axial stress.

49
Analysis of thin walled aerospace structures

7.2.2 Equilibrium Equation relating Shearing Force and Bending Moment


Consider the yz plane of a beam under generalised loading. At some point z along the beam consider the
equilibrium of a small elemental length (dz):
The distributed load wY(z) may be considered constant over the small element length dz and the variation
of VY and MX may be well approximated by a linear function.

y wy(z)

∂M x
Mx + dz
Mx ∂z
A
Vy
z
z dz ∂V y
Vy + dz
∂z
Figure 7.2: yz-plane element

Other equilibrium equations are obtainable from Figure 7.2 and although these are not required for the
analysis of bending shear stress, they are necessary for the calculation of deflections (covered in Section
9) and are recorded here for reference purposes. Thus, in the yz-plane:

 ∂V 
∑F
Y ⇒ − VY + wY dz +  VY + VY Y dz  = 0
 ∂z 
∂VY
wY = −
∂z (51)
dz  ∂M X 
∑M A ⇒ − M X − wY dz . +  M X +
2  ∂z
dz  − VY dz = 0

∂M X
{
as dz 2 is negligible VY = } ∂z
which can be extended to the xz-plane to give
∂VY ∂ 2M X
wY = − =−
∂z ∂z 2
(52)
∂V ∂2MY
wX = − X =
∂z ∂z 2

50
Analysis of thin walled aerospace structures

7.2.3 Equilibrium Equation relating Shearing Stress and Axial Stress


Introduce ‘s’ a parameter to measure distance around a surface. The origin for ‘s’ can be at any point
although there are some locations that simplify analysis. The choice of a starting point for s will be
discussed in later lectures. To derive relationships between shear forces and axial forces consider a small
element of the surface of dimensions ds by dz.

∂σ s σz
σs + ds
y ∂s
∂τ τ
τ+ ds ds
ds x
∂s
dz
s
τ
s
dz ∂τ
s=0
τ+ dz σs
∂z
∂σ z
z σz + dz
∂z z t

Figure 7.3: Generalised stress element

On the small element of the surface both q and the axial and hoop stresses σ

Z and σ

S may be assumed to
vary linearly so that the following general stress element may be drawn.

∂σ s σz
σs + ds
∂s
∂q q
q+ ds ds
∂s
dz
s q
∂q σs
q+ dz
∂z
∂σ z
σz + dz
∂z z t

Figure 7.4: Generalised shear flow element

S is normally zero unless the structure is a pressure vessel (eg. a/c fuselage). Use force equilibrium to
gain relationship between shear flow and stresses. The wall thickness t may vary with s, but is considered
independent of z, so that the equation of equilibrium of forces in the z-direction is:
 ∂q   ∂σ Z 
∑F Z =  q + ds dz − q .dz +  σ Z +
∂s  ∂z
dz .t .ds − σ Z .t .ds = 0
   (53)
∂ q ∂σ Z
⇒ +t =0
∂s ∂z

 ∂q   ∂σ S 
∑F S =  q + dz  ds − q .ds +  σ S +
∂z 
ds .t .dz − σ S .t .dz = 0
∂ s 
  (54)
∂ q ∂σ S
⇒ +t =0
∂z ∂s

51
Analysis of thin walled aerospace structures

The general expression for the calculation of bending shear stress is obtained from Equation (21), using
the equilibrium equations above. Thus, differentiating Equation (21) with respect to z gives:

∂σ Z ∂ M X  I Y .y − I XY .x  ∂ M Y  I X .x − I XY .y 
=   −  
∂z ∂z  I X IY − I XY  ∂ z  I X I Y − I XY 
2 2

 I .y − I XY .x   I .x − I XY .y 
= VY  Y  + VX  X  (55)
I
 X Y I − I 2
XY  I
 X YI − I 2
XY 

 V I − VY I XY   VY I Y − VX I XY 
=  X X  x +   y
 I X I Y − I XY   I X IY − I XY
2 2

Therefore the partial derivative of q with respect to s may be related to the shear loading and structural
geometry by:

∂q  V I − VY I XY   V I − V X I XY 
= −t . X X  x − t . Y Y  y
∂s  I X I Y − I XY 
2
 I X I Y − I XY 
2

−t  V I  t  V I 
=  V X − Y XY  x −  VY − X XY  y (56)
KI Y  IX  KI X  IY 
− tV X′ tV ′
= x− Y y
IY IX
where V'X and V'Y are the effective shearing forces for the chosen axes. V'Y bears the same relationship to
M'X as VY bears to MX. Similarly, in the xz-plane, V'X is related to M'Y in the same manner as VX is related
to MY. Thus the effective shearing forces V'X and V'Y, can be interpreted physically as the combinations of
VX and VY which, acting through the shear centre E, produce bending strains about the x- and the y-axes
respectively. These relationships provide a useful method for the calculation of the shear centre position of
an asymmetrical section.
Finally, the general expression for calculating the bending shear flow at any point s on the cross-section is
obtained by integrating Equation (56) with respect to s. Thus:

∂q
s
q (s ) = ∫ ds
0
∂s
 V X I X − VY I XY  s  V I − V X I XY s
= qi ,0 −   ∫ t x.ds −  Y Y  ∫ t y .ds
 I X I Y − I XY  0  I X I Y − I XY
2 2
0 (57)
V X′ VY′
s s

I Y ∫0 I X ∫0
= qi ,0 − t x.ds − t y .ds

V X′ V′
= qi ,0 − QY − Y Q X
IY IX
where QX, QY are the first moments of area for all axial stress-carrying material lying between the origins
of s and the point s under consideration, about the x- and the y-axes respectively. qi,0 is the constant of
integration, that is, the values of q at the origin of s in cell i.
Detailed methods for determining the constants qi,0 are given later. However for open sections the shear
flow on a free edge must be zero, therefore choosing the origin of s to be at the free edge sets qi,0 = 0.
Similarly if, the section is doubly symmetrical, a suitable choice of origin will also make the constants qi,0
= 0, and corresponding reductions in work are also possible for sections with one axis of symmetry.

52
Analysis of thin walled aerospace structures

7.2.4 Example – Shear flow in an effective open section (Example 9.4 from Megson)
Determine the variation of shear flow in the thin walled Z section shown due to a vertical shear force VY
applied through the centroid (no twisting will occur due to this loading as centroid is also the shear
centre).

q12 =
VY
h3
(
5 .16 s12 − 1 .74 hs1 )
q 23
V
(
= Y3 − 3 . 42 s 22 + 3 .42 hs 2 + 0 . 42 h 2
h
)
q 34
VY
(
= 3 5 .16 s 32 − 3 . 42 hs 1 + 0 . 42 h 2
h
)
y

Vy
h/2
3 4 0.42Vy/h
s3

t
h C
x

s1 s2
2
1 h/2 0.42Vy/h

Figure 7.5: Cross-section and solution for example 7.2.4

53
Analysis of thin walled aerospace structures

7.2.5 The effect of boom idealisation on shear flow induced by shear stresses
Currently Equation (57) makes no obvious provision for the effects of booms (particularly important in
structural idealisation) which cause discontinuities in the skin and therefore interrupt the shear flow.
Consider the equilibrium of the rth boom in the elemental length of the beam shown in Figure 7.6(a) which
carries shear loads VX and VY acting through its shear centre E. These shear loads produce direct stresses
due to bending in the booms and skin and bending shear stresses in the skin. Suppose that the shear flows
in the skin adjacent to the rth boom of cross-sectional are Br are q1 and q2.

VY

VX

Figure 7.6: (a) Element of shear loaded open section beam with booms and (b) equilibrium element

Then equilibrium of the linear force element (boom) in Figure 7.6(b) and Equation (55) give:
∂σ Z
q 2 − q1 = − Br
∂z
 V I − VY I XY   V I − V X I XY 
= − X X  Br x −  Y Y  Br y (58)
 I X I Y − I XY   I X I Y − I XY
2 2

V′ V′
= − X Br xr − Y Br y r
IY IX
This gives the change in shear flow induced by a boom which itself is subjected to a direct load ( ZBr). σ

Each time a boom is encountered the shear flow is incremented by this amount so that if, at any distance s
around the profile of the section, k booms have been passed, the QX and QY at the point is given by:
(1) For an idealised section:
k
QX = ∑ Br yr
0
k
(59)
QY = ∑ Br xr
0

(2) For an effective section including axial-force elements:


k s
QX = ∑ Br yr + ∫ yt .ds
0 0
(60)
k s
QY = ∑ Br xr + ∫ xt .ds
0 0

54
Analysis of thin walled aerospace structures

7.2.6 Example – Shear flow in an idealised section


Determine the variation of shear flow in the idealised C section shown due to a vertical shear force VY
applied through the shear centre.

Y
200mm
2 1

s
4.8kN

400mm
C
E X

3 4

200mm

Figure 7.7: Comparison of shear force induced shear flow in effective and idealised sections

As the effective direct stress carrying area of the walls is zero the section properties refer to the boom only
so:
IX = 48×106mm4 and IXY = 0
Equations (57) and (59) combine to give:
k
q (s ) = qi ,0 −
VY
IX
∑B y
0
r r

Observing the shear flow at a free edge must be zero (i.e. q4,0 = 0) this simplifies further to:
k
q (s ) = −
VY
IX
∑B y
0
r r

so that the following shear flows may be calculated:


q12 = -6N/mm
q23 = -12N/mm
q34 = -6N/mm

55
Analysis of thin walled aerospace structures

8. CALCULATION OF SHEARING-STRESS DISTRIBUTION AND SHEAR CENTRE


POSITION

8.1 Calculation of Shear Centre Co-ordinates:


8.1.1 Fundamental approach
To determine the co-ordinates (xE, yE) of the shear centre E, Equation (57) is used to calculate the
following bending shear-flow distributions:
(1) due to VY acting alone at the shear centre (VX = 0) for the co-ordinate xE, and
(2) due to VX acting alone at the shear centre (VY = 0) for the co-ordinate yE.

Figure 8.1: Loading cases used to locate the shear centre

The co-ordinates of E are then obtained by equating the moments of these bending shear flows to the
moment of the corresponding shearing force component as shown in Figure 8.1. Thus

VY x E = ∫ pqds
(61)
− V X y E = ∫ pqds

where p is the perpendicular from the tangent at s to the centroid C. To apply this procedure, however,
both terms in Equation (57) for q must be evaluated for each element of the cross-section, because, in each
case, V'X as well as V'Y may be present. This is the method that Megson uses and can involve some
significant calculations.
Equation (57) shows that the shear flow induced by bending is obtained throughout the section by a
process of integration when the constants at the origins of s are known. These constants can be determined
directly from the boundary conditions in the case of the open section (where at free edges q = 0), or by
making use of symmetry conditions, if applicable (Section 8.1.3). In all other cases the constants (qi,0) may
be determined by applying the condition that the section does not twist when shearing forces are applied
through the shear centre. From Equation (38):
dϕ T q
=∫ 2
ds = ∫ ds = 0
dz 4 A Gt 2 AGt
⇒∫ i
(q + qi ,0 ) ds = qi ds + q ds = 0
t ∫t i ,0 ∫
t
(62)

qi
∫ ds
⇒ qi ,0 = − t
ds
∫t

56
Analysis of thin walled aerospace structures

Equation (62) must be satisfied for each cell of an n-cell section, giving n equations for the determination
of the constants q1,0, q2,0,…. qn,0 at the origins of s.
The calculations of this method may be simplified by eliminating one term at a time from Equation (57).
This is done by applying a suitable ratio of forces VX:VY as follows.
8.1.2 Alternative Approach using Shear Devices
An alternative and generally more useful approach for determining (xE, yE) is provided by the use of shear
devices derived from the effective shearing forces V'X, and V'Y. The condition for the x-axis to be a neutral
axis is obtained by setting V'X = 0. This gives:
I XY
VX = VY (63)
IX
This gives V'Y = VY. Hence the shear-flow distribution, becomes (from Equation (57)):
VY
q = qi ,0 − QX (64)
IX
Thus the application of shearing force components VY together with VX = VY.IXY/IX causes the x-axis to be
a neutral axis, and the shear-flow distribution is given by Equation (64).

Figure 8.2: Shear device loading cases used to locate the shear centre

Taking moments about the centroid C gives the equation:


I XY
VY xE − VY y E = ∫ pqds (65)
IX
Thus the shear centre E lies on a line lX, (Figure 8.3) which has the equation:
IX I
y=
I XY
x− X
I XYVY ∫ pqds (66)

57
Analysis of thin walled aerospace structures

The slope of lX is the ratio VY:VX and the intercept of lX, on the y-axis is determined by the moment
equilibrium Equation (65). Similarly, the y-axis is a neutral axis when V'Y = 0, which, following a similar
argument to that above, may be used to determine that shear centre E lies on a line lY which has the
equation
I XY 1
y=
IY
x−
VX ∫ pqds (67)

Figure 8.3: Location of shear centre

The intersection of lX and lY gives the co-ordinates (xE, yE) of the shear centre E.
The shearing stresses calculated by Equation (64) and its equivalent when V'Y = 0 cause the section to
bend about a specified axis without twisting, but the resultants of these stresses, when superposed, give the
following force components:
I XY
(1) V X + VY along the x-axis, together with
IX
I XY
(2) VY + V X along the y-axis
IY
In order that the stresses calculated in the process of determining the shear centre E can be superposed
directly in the calculation of the resultant stresses due to specified applied forces VX and VY, it is
necessary that the shear devices used are as shown in Figure 8.4.
The corresponding shear-flow distributions are:
VY′
q = qi ,0 − QX
IX
(68)
V′
q = qi ,0 − X QY
IY
for x and then y as the neutral axis respectively.

58
Analysis of thin walled aerospace structures

Figure 8.4: Shear devices to be used for simultaneous calculation of shear flow and shear centre.

The superposition of the shear devices of Figure 8.4(a) and (b) gives the following force components:
I XY
(1) VX′ + VY′ = VX along the x-axis, together with
IX
I XY
(2) VY′ + VX′ = VY along the y-axis
IY
That is, the use of V'X and V'Y in the shear devices gives stress distributions that satisfy:
(1) the deformation requirement that the section bends about a specified axis without twisting, and
(2) the equilibrium requirement that superposition of the stresses calculated in step (1) gives a force
resultant equal to the specified applied force system.
The application of this technique is shown in the following examples for asymmetrical open, single-cell,
and multicell sections respectively. The following examples also show how the work involved in
calculating the shear centre co-ordinates by the use of the shear device can be used directly in the solution
of the shearing-stress distribution due to arbitrarily-located shearing forces.

59
Analysis of thin walled aerospace structures

8.1.3 Example: Asymmetrical Open Section


The channel section shown in Figure 8.5 is of uniform thickness 5 mm, and is loaded by a 2 kN force
applied at the shear centre E and acting parallel to the web BD. Calculate the shearing-stress distribution
and the position of the shear centre. The centroid C of the section is located as shown in Figure 8.5(a).

Figure 8.5: Asymmetrical Open Section

Solution methodology:
• Determine section properties.
IX = 15.55×10-6m4, IY = 4.358×10-6m4, IXY = 2.223×10-6m4
• Use the shear device of Section 8.1.2 to determine the combination of shearing forces required to
make the x-axis the neutral axis.
2.157kN in the y direction and 0.308kN in the x direction
• Determine shear flow/stress in shear webs and then take moments about B to determine an equation
for the location of the shear centre (lX).
2.157x - 0.308y = 0.07709 where x and y co-ordinates are from point B
• Similarly, determine the combination of shearing forces required to make the x-axis the neutral axis.
-0.157kN in the y direction and -0.308kN in the x direction
• Determine shear flow/stress in shear webs and then take moments about B to determine a second
equation for the location of the shear centre (lY).
-0.157x + 0.308y = 0.0108 where x and y co-ordinates are from point B
• Solve the two equations lX and lY for the co-ordinates of the shear centre (xE, yE)
xE = 0.0439 m, yE=0.0574 m where xE and yE co-ordinates are from centroid.
• The resultant shearing-stress distribution shown in Figure 8.5(b) is obtained by adding the
calculated stress distributions from the two cases when x- and y-axes were made neutral axes.

60
Analysis of thin walled aerospace structures

8.1.4 Asymmetrical Closed Section.


In the idealised section shown in Figure 8.6, the axial-force elements at A, B, and D each have a cross-
sectional area 0.002 m2. The elements at F and G each have a cross-sectional area of 0.001 m2 and all
shear fields are of equal thickness. Calculate the position of the shear centre of the section, and the shear-
flow distribution due to the 100 kN force applied in the location as shown. The centroid C of the section is
located as shown in Figure 8.6(a).

q1,0

Figure 8.6: Asymmetrical Closed Section

Solution methodology:
• Determine section properties.
IX = 320×10-6m4, IY = 1200×10-6m4, IXY = 320×10-6m4
• Determine the combination of shearing forces required to make the x-axis the neutral axis, the shear
flow in the shear webs (using Equation (62) to gain initial values) and then take moments about D
equating resultant shear flow and applied shear forces to determine an equation for the location of
the shear centre (lX).
136.4kN in the y direction and 136.4kN in the x direction
q1,0 = -8.19kN/m
x - y = -0.5337 where x and y co-ordinates are from point D
• Similarly, for the y-axis as the neutral axis.
-136.4kN in the y direction and –36.4kN in the x direction
q1,0 = -68.53kN/m
-x + 3.75y = -0.05538 where x and y co-ordinates are from point D
• Solve the two equations lX and lY for the co-ordinates of the shear centre (xE, yE)
xE = -0.15 m, yE = -0.014 m where xE and yE co-ordinates are from centroid.
• Determine the shear flow due to torsion (as shear load does not pass through shear centre)
• The resultant shearing-stress distribution shown in Figure 8.6(e) is obtained by adding the
calculated stress distributions from the two cases when x- and y-axes were made neutral axes and
the torsion induced shear loading.

61
Analysis of thin walled aerospace structures

8.1.5 Example: Asymmetrical multi-cell section


The two-cell section shown in Figure 5.14 is subjected to shearing forces VY = 200 kN and VX = 100 kN,
acting through the element A. Calculate the position of the shear centre of the section and the shear-flow
distribution. The origins for the peripheral co-ordinates s1 and s2 are taken in the fields AD and DF
respectively.

q2,0

q1,0 Fig (a)

Figure 8.7: Asymmetrical Multi-Cell Section


Solution methodology:
• Determine section properties.
IX = 10-2m4, IY = 4.75×10-2m4, IXY = 1.5×10-2m4
• Determine the combination of shearing forces required to make the x-axis the neutral axis, the shear
flow in the shear webs (using Equation (62) to gain initial values) and then take moments about the
centroid equating resultant shear flow and applied shear forces to determine an equation for the
location of the shear centre (lX).
320kN in the y direction and 480kN in the x direction
q1,0 = 49kN/m and q2,0 = 101kN/m
xE – yE = -0.4313 where x and y co-ordinates are from centroid
• Similarly, for the y-axis as the neutral axis.
-120kN in the y direction and -380kN in the x direction
q1,0 = -55.4kN/m and q2,0 = -67.9kN/m
-0.316xE + yE = -0.0758 where x and y co-ordinates are from centroid
• Solve the two equations lX and lY for the co-ordinates of the shear centre (xE, yE)
xE = -0.194 m, yE = 0.0146 m where xE and yE co-ordinates are from centroid.
• Determine the shear flow due to torsion (as shear load does not pass through shear centre)
• The resultant shearing-stress distribution shown in Figure 8.7(e) is obtained by adding the
calculated stress distributions from the two cases when x- and y-axes were made neutral axes and
the torsion induced shear loading.

62
Analysis of thin walled aerospace structures

8.2 Reduction in Work due to Symmetry


If the section has one axis of symmetry it is advantageous to select the co-ordinate axes so that one of
these axes coincides with the axis of symmetry. With this choice, the co-ordinate axes are the principal
axes of the section and IXY = 0. Hence, K = 1, V'X = VX, V'Y = VY, and Equation (57) is simplified. The
stress distribution is symmetrical about the axis of symmetry for loading applied along this axis and anti-
symmetrical for loading applied perpendicular to the axis of symmetry. Hence the constants qi,0 due to this
loading are zero at the intersections of the axis of symmetry with the section (see Figure 8.8).
The value of qi,0 has only to be determined for loading by shearing forces acting perpendicular to the axis
of symmetry. The complete solution for loading by both shearing force components VX and VY is obtained
by super-position.

Figure 8.8: Symmetric shear flow distribution in symmetric sections.

The shear centre lies on the axis of symmetry and so requires the calculation of only one co-ordinate to
locate its position. Also, in the analysis due to forces acting perpendicular to the axis of symmetry, the
origin of the co-ordinate axes can be taken at any convenient point on the axis of symmetry and need not
necessarily be taken at the centroid C of the cross-sectional area. This often simplifies the algebra
considerably and is illustrated by the analysis of Example 8.2.1.
Similarly, if the section has double symmetry, that is, symmetry about two perpendicular axes, the co-
ordinate axes should be selected to coincide with the axes of symmetry. In this case the constants qi,0
involved in each solution are zero when the origins for the peripheral co-ordinate s are taken at the
appropriate intersections of the axes of symmetry with the section. The shear centre of a double-
symmetrical section coincides with the centroid of the cross-sectional area.

63
Analysis of thin walled aerospace structures

8.2.1 Example: Section with One Axis of Symmetry.


The section shown in Figure 8.9is loaded by a shearing force 10-2MN applied at the shear centre E and
perpendicular to the axis of symmetry. Calculate the shearing-stress distribution and the shear centre
position. Select the co-ordinate axes with origin at F and the x-axis along the axis of symmetry. Take the
origin for the peripheral co-ordinate s at the intersection A of the section with the x-axis.

Figure 8.9: Analysis of a singularly-symmetric section.

Solution methodology:
• Determine relevant section properties.
IX = 3.21×10-6m4, IXY = 0
• By symmetry, the stress distribution is antisymmetrical about the x-axis and the shear centre lies, on
the x-axis.
• Determine the shear flow in the shear webs using Equation (62) to gain the initial value at A, and
then take moments about F equating resultant shear flow and the applied vertical shear force to
determine the location of the shear centre (xE).
Shear stress distribution is as shown above
qA,0 = 32.89kN/m →
τ

A = 32.89MPa
xE = 0.1075m from origin at F.

64
Analysis of thin walled aerospace structures

8.3 Calculation of Resultant Shearing Stress


Two approaches are possible to calculate the resultant shearing-stress distribution.
8.3.1 General Method
The general method is the approach previous examples have employed. The shear centre position is
calculated using the methods of section 8.1.2. The applied forces are then transferred to the shear centre
producing, in general, bending shear stresses due to bending about each co-ordinate axis, together with
shearing stresses due to torsion about the shear centre. The resultant shearing stress is the sum of these
three components. It may be noted that each application of the section 8.1.2 shear devices produces
directly the bending shear distribution about the relevant co-ordinate axis, together with one equation to
determine the shear centre position. This is the general approach to the problem and, in addition to the
shearing-stress distribution, the calculations produce the shear centre co-ordinates and also the torsion
constant J of the section. This approach is illustrated in Example 8.3.2.
8.3.2 Example: General Method (including torsion constant) for a section with One Axis of Symmetry
The axial-force elements at A, H, D, and F of the idealised section shown in Figure 8.10 each have a
cross-sectional area 0.005m2. The elements at B and G each have a cross-sectional area 0.02m2. Each
shear field is 5mm thick. Calculate the shear-flow distribution, the shear centre position, and the torsion
constant J of the section. Select the coordinate axes with origin at the mid-depth of AH and the x-axis
along the axis of symmetry. The origins for the peripheral coordinates s1 and s2 are taken in the fields AB
and BD respectively

q1,0 q2,0

Figure 8.10: Section with one axis of symmetry

65
Analysis of thin walled aerospace structures

Solution methodology:
• Determine necessary section properties.
IX = 0.375×10-2m4, IXY = 0m4
• By symmetry, the shear-flow distribution is anti-symmetrical about the x-axis and the shear centre
lies on the x-axis.
• Determine the change in shear flow at force elements A, B and D.
50kN/m, 200kN/m and 50kN/m
• Use Equation (62) to gain initial values of shear flow and hence the shear flow distribution in the
structure and then take moments about the origin to determine x co-ordinate of the shear centre (xE).
q1,0 = -19.6kN/m and q2,0 = 32.6kN/m
xE = 0.895 m
• Determine the torsion constant (see Example 6.3.6) and shear flow due to torsion (as shear load
does not pass through shear centre).
J=2.826×10-3 m4
• The resultant shearing-stress distribution is obtained by adding the calculated stress distributions
from the bending and torsion induced shear loading.

66
Analysis of thin walled aerospace structures

8.3.3 Alternate/Reduced method


Alternatively, if it is required to calculate only the resultant shearing-stress distribution, this can be done
more directly by the following procedure, which does not require the calculation of the shear centre
position.
The n initially unknown shear flows qi,0 (i = 1,… n) are defined as resultant shear flows, one in an outer
wall of each cell, as before (section 7.2). Equation (57) then expresses the resultant shear-flow distribution
throughout the section in terms of these n unknowns qi,0. The n equations necessary for the determination
of the n unknowns qi,0 are provided by:
(1) the moment equilibrium equation, and
(2) the (n - 1) equations of compatibility (Equation (47)) which express the conditions for each cell to
twist at the same rate under the specified applied forces.
The solution of these n equations gives qi,0 which are substituted into the expressions for shear flow
derived from Equation (57), to give the resultant shear-flow distribution through-out the section.
This alternative procedure illustrates clearly that the section is statically determinate in bending and the
only redundancies involved are those due to torsion. The alternative procedure is more direct and involves
less computation that that of the General Method when only the shear-flow distribution is required.
However, in most practical analyses, it is usually also necessary to determine the shear centre position and
the torsion constant J.
8.3.4 Example: Alternative solution for Example 8.3.2.
Using the problem shown in Figure 8.10, q1,0 and q2,0 are designated as the intially-unknown shear flows in
the fields AB and BD, due to the 150 kN force acting in its specified position.
• Applying Equation (57) gives the shear-flow distribution throughout the section as before, but in
terms of the unknowns q1,0 and q2,0.
qBG = q1,0 – q2,0 +200
qDF = q2,0 +50
qFG = q2,0
qGH = q1,0
qHA = q1,0 – 50
• Determine a relationship between q1,0 and q2,0 using the moment equilibrium equation about A
-77.5 = q1,0 + 0.5q2,0

θ θ

Determine a relationship between q1,0 and q2,0 using the compatibility equation ( 1 = 2).
-450 = 8q1,0 - 9q2,0
• Solve for q1,0 and q2,0 and substitute back into the expressions for shear flow above to gain the
solution for shear flow in the section.

67
Analysis of thin walled aerospace structures

9. ANALYSIS OF COMBINED OPEN AND CLOSED SECTIONS

In some cases the cross-section of a beam is formed by a combination of open and closed components. For
example. a wing section in the region of an undercarriage bay could take the form shown in Figure 9.1 in
which the nose portion is a single cell closed section and the cut-out forms an open channel section. Such
composite sections may be analysed using, where appropriate, a combination of the methods presented in
the earlier lectures.

Figure 9.1: Combined open and closed section

(1) Bending: the direct stresses due to bending are given by Equation (21).
(2) Shear loading: Where shear loads do not act through the shear centre its position must be found and
the loading system replaced by shear loads acting through the shear centre together with a torque;
the two loading cases are then analysed separately. Again assume that the cross-section of the beam
remains undistorted by the loading.
(3) Torsion: Generally, in the torsion of composite sections, the closed portion is dominant since its
torsional stiffness is far greater than that of the attached open section portion which may therefore
be frequently ignored in the calculation of torsional stiffness. It may be assumed, in simplified
analyses, that the torsional rigidity of the complete section is the sum of the torsional rigidities of
the open and closed sections.

68
Analysis of thin walled aerospace structures

TUTORIAL SET 3: SHEAR STRESS DISTRIBUTION AND SHEAR CENTRES

(1) Calculate the shearing-stress distribution in the thin-wall beam of Figure Q.1 due to a shearing force F
applied along the centre-line of the web thickness.
Select the co-ordinate axes coincident with the axes of symmetry and the origin for the peripheral co-
ordinate s at the free end A of the flange AB.
Since the section is doubly symmetrical the shear centre coincides with the centroid C, and the stress
distribution is symmetrical about the y-axis and anti-symmetrical about the x-axis.
Answer: as shown in Figure Q.1(b).

-Fdt1b/IX

-2Fdt1b/IX

-F(2dt1b + ½d2 t2 )/IX

Figure Q.1

(2) Calculate the shearing-stress distribution and the position of the shear centre of the section shown in
Figure Q.2. The wall thickness t is constant around the periphery and the shearing force F is applied at
the shear centre E, perpendicular to the axis of symmetry.
Select the co-ordinate axes having origin at the centre 0 of the circular arc and the x-axis along the axis
of symmetry.
By symmetry, the stress distribution is anti-symmetrical about the x-axis and the shear centre lies on
the x-axis.
Answer: as shown in Figure Q.2(b).

Figure Q.2

69
Analysis of thin walled aerospace structures

(3) Locate the shear centre of the following thin walled open sections.
Answers:
(a) x = -4r/ , y=0 π

(b) x = -3tf /(htw+6tf (b1+b2), y = 0


(c) x = 3d2(1- )/(h(1+12d/h)), y = 0
β

(d) x= -45a/97, y = 46a/97


3b2 (h12 + h22 )
(e) x = h2 + 6b(h1 + h2 ) , y = 0
3 2 2

3bh 2 (b + 2a ) − 8ba 3
x = h (h + 6b + 6a ) + 4a (2a − 3h) , y = 0
2 2
(f)

y y
tf y
d β

r tw t t/
β

t
h
O x
x h C x
tf
β

t t/
Constant Thickness = t
b1 b2

Figure Q.3(a) Figure Q.3(b) Figure Q.3(c)

y y y
a b Constant b
thickness
t
2t (both) a
Constant
thickness
x t x
2a h2 h
h1
t
x
a

2a
Figure Q.3(e) Figure Q.3(f)
Figure Q.3(d)

70
Analysis of thin walled aerospace structures

(4) Determine the shear flow distribution in the singly


symmetrical three-cell wing section shown in Figure
Q.4(a) when it carries a shear load of 100 kN applied
through its shear centre and hence find the distance
of the shear centre from the spar web 34. Assume E
that all direct stresses are resisted by the booms while
the skin is effective only in shear. The shear modulus xE
G is constant throughout.
Answer: as shown in Figure Q.4(b)
Boom Areas: B1 = B6 = 2500mm2,
B2 = B5 = 3800mm2,
B3 = B4 = 3200mm2, Figure Q.4(a)
Cell Areas: AI = 0.265m2, AII = 0.580m2,
AIII = 0.410m2
Wall lengths: s12 = s56 = 1.025m,
s23 = s45 = 1.275m,
s34(outer) = 2.200m, s34(inner) = 0.400m,
s16 = 0.330m,
s25 = 0.460m.
Thickness: t12 = t56 = 1.25mm, Figure Q.4(b)
t23 = t45 = t16 = 1.65mm,
t34(outer) = 2.250mm,
t34(inner) = t25 = 2.65mm.

(5) Calculate the shear flow distribution for the idealised Y 150mm 150mm
open csection shown in Figure Q.5 when a shear
force of 200kN is applied vertically through the shear
centre. Also determine the location of the shear
3 2 1
centre.
Boom Areas: B1 = B6 = 250mm2, X
100mm
B2 = B5 = 500mm2,
B3 = B4 = 250mm2,
Answers: q12 = q56 = 250N/mm 4 5 6
q23 = q45 = 750N/mm
q34 = 1000N/mm
Figure Q.5
xE = -307mm

(6) Calculate the shear flow distribution for the idealised Y 150mm 150mm
closed section shown in Figure Q.6 when a shear
force of -200kN is applied vertically through linear
force element 4. Also determine the location of the
3 2 1
shear centre.
Boom Areas: B1 = B6 = 250mm2, X
100mm
B2 = B5 = 500mm2,
B3 = B4 = 250mm2,
Answers: q12 = q56 = -155.6N/mm 4 5 6
q23 = q45 = 344.4N/mm
q34 = 594.4N/mm
Figure Q.6
xE = 110.24mm

71
Analysis of thin walled aerospace structures

(7) The axial-force elements at the corners A,


G, H, and M of the idealised section
shown in Figure Q.7 each have a cross-
sectional area 0.01m2 . The elements at B,
D, F, J, K, and L each have a cross-
sectional area 0.005m2, and all shear
fields are of equal thickness. Calculate the
shear-flow distribution due to the 500 kN
force applied as shown.
q1,0 q2,0
Answers: as shown in Figure Q.7(b)

Figure Q.7

(8) A singly symmetric wing section consists


of two closed cells and one open cell as
shown in Figure Q.8. The webs 25, 34
and the walls 12, 56 are straight, while all
other walls are curved. All walls of the
section are assumed to be effective in
carrying shear stresses only, direct
stresses being carried by booms 1 to 6. Figure Q.8
Calculate the distance of the shear centre Boom Areas: B1 = B6 = 645mm2,
aft of the web 34. The shear modulus is B2 = B5 = 1290mm2,
constant for all walls. B3 = B4 = 1935mm2,
Answer: 241.4mm Cell Areas: AI = 0.093m2,
(9) Calculate the shear flow distribution for AII = 0.258m2
the section shown in Figure Q.8 if a Wall lengths: s12 = s56 = 0.510m,
vertical shear load of 200kN acts along s23 = s45 = 0.765m,
web 34. s34(outer) = 1.015m,
Answer: s34(inner) = 0.304m,
s25 = 0.304m.
Thickness: t12 = t56 = 0.559mm,
t23 = t45 = 0.915mm,
t34(outer) = 0.559mm,
t34(inner) = 2.030mm
t25 = 1.625mm

72
Analysis of thin walled aerospace structures

10. CALCULATION OF DEFLECTIONS

10.1 Beam Deflections due to Pure Bending


When the beam is subjected to transverse forces, these should be resolved along the co-ordinate axes and
replaced by parallel forces acting through the shear centre axis. These forces produce bending moments
MX and MY about the coordinate axes together with a torsion T about the shear centre axis. The torsional θ

deflection about the shear centre axis is calculated using the expression for rate of twist, = T/GJ, where
the torsion constant J is determined using Equations (38) and (39).
The deflections due to the bending moments MX and MY are calculated using the curvatures (1/rX,) and
(1/rY) given by Equation (20). Using the sign rules given in Figure 2.4, it follows that, for small
deflections, the curvatures may also be related to deflections to give:
1 d 2v M ′
=− 2 = X
rX dz EI X
(69)
1 d 2u M Y′
= =
rY dz 2 EI Y
where u and v are the deflections along the x- and y-axis respectively. These are the equations for the
deflection of the shear centre axis. The deflection components of the shear centre axis due to bending, u
(in the xz-plane) and v (in the yz-plane) are obtained by integrating Equation (69). The resultant deflection
is the vector sum of u and v.
10.1.1 Example: Deflection due to pure bending
The section of a uniform cantilever beam, 10 m long, is shown in Figure 10.1. Calculate the deflection
equations of the shear centre axis of this beam due to bending moments MX = -2 MNm and MY = 1 MNm
applied to the tip section. Use a modulus of elasticity E = 70 GPa.
Y

X
Y
Idealised booms have
0.25m
an area of 0.01m2
1MNm

C
1m
2MNm X
0.5m
10m

1m 2m
Z

Figure 10.1: Idealised cross-section subject to pure bending

The section characteristics with respect to the xy-axes passing through the centroid C are: IX = 0.0l m4, IY
= 0.0475 m4, IXY = 0.015 m4. These give values for K and the effective bending moments of K = 0.5263,
M'X = -3.2MNm, M'Y = -3.8MNm. Integrating Equation (69) twice using the boundary for a cantilever
beam that at z = 0, du/dz = dv/dz = u = v = 0 gives:
u = -0.05715×10-2 z2 m and v = 0.2286×10-2 z2 m
The resultant deflection is the vector sum of u and v. For example, the deflection components at the tip
section (z = 10 m), are vT = 0.2286 m, uT = -0.05715 m and the resultant tip deflections T, is: δ

δ T = 0.22862 + 0.057152 = 0.2356m at an angle tan-1(v/u) = 104º 2', 284º 2' to the x-axis.
θ

Thus δ

T is perpendicular to the neutral axis for this loading, that is, = 14º 2', 194º 2' (See example 5.3.6).

73
Analysis of thin walled aerospace structures

10.2 Deflections due to shear loading


The previous section considered the deflection of a beam due to pure bending (i.e the application of only
bending moments). If the beam is loaded by shear forces so that the beam deflects due to a combination of
shear loading and bending moments, calculation of deflection may need to include the shear effects. In
most engineering structures the effect of shear stresses on deflections is small so that equation (69) is
sufficiently accurate. However, in thin walled structures this may not be the case and shear effects should
be considered.
Firstly consider the effect of shear loading on a rectangular cross-section element where the shear force
(VY) and bending moment (MX) at any location along the beam are given by VY = P and MX = -P(L-z).
The bending moment distribution along the beam may be used to gain a first order approximation for the
deflection of the neutral axis (from equation (69)):
− M' X
v' ' = =
P
(L − z ) = 1 (PL − Pz )
EI X EI X EI X
1  P 2 
v' =  PLz − z + C1 
EI X  2 
1  PL 2 P 3 
v=  z − z + C1 z + C2 
EI X  2 6 
The constants C1 and C2 would normally be determined by considering the slope and deflection at z = 0 to
equal 0. Therefore the deflection at any location along the beam would be predicted to equal:

v=
P
6 EI X
(
3 Lz 2 − z 3 )
Y

P
X
γ

P τ

L
τ

h
τ

Z τ

Figure 10.2: The effect of shear loads on beam displacements

Figure 10.2 shows the deflection of the neutral axis caused by shear. Assuming an even distribution of
shear across the cross-section of the beam, Hooke’s law predicts the shear strain (along the neutral axis) in
the beam to equal ( =) /G = P/bhG = Ph2/12IXG. Thus the actual slope of the neutral axis at z = 0 is equal
γ
τ

to and the C1 constant now equals Ph2/12IXG so that the deflection of the beam neural axis with both the
γ

shear and bending effects included equals:

v=
P
6 EI X
(3 Lz 2 − z 3 ) +
Ph 2 z
12 I X G
Considering the shear loading to be uniformly distributed is however, overly simplified as the shear
distribution is not uniform across a cross-section. This is obvious as Figure 10.2 shows the upper and
lower free edges of the beam having a non-zero shear stress. This is obviously not possible and beam

74
Analysis of thin walled aerospace structures

theory (recall lectures from semester 1) may be used to show that the variation in transverse shear stress
across a solid (or thick wall) prismatic beam cross section may be predicted using the semi-inverse method
of St Venant to give:
VY Q( y )
τ ZY ( y ) = (70)
I X b( y )
where VY is the y direction shear force on the section, IX is the second moment of area of the section about
its centroidal X axis, b(y) the width of the cross-section at height y from the neutral (X) axis and Q is the
first moment of area about the neutral (X) axis of the cross sectional material bounded by the edge of the
beam and the height y.
Equation (70) predicts the shear stress at height y = 0 (i.e at the height of the neutral axis) for a rectangular
cross-section beam to equal ZY = Ph2/8IX. Therefore the rotation of the neutral axis at z = 0 is equal to
τ

Ph2/8IXG and the actual deflection of the beam including shear effects is given by:

v=
P
6 EI X
(3 Lz 2 − z 3 ) +
Ph 2 z
8I X G
10.2.1 Example: The effect of shear loading on I-beam displacement
Determine the deflection of a 10m long I beam loaded by a vertical end load of 250kN. Use the results of
Tutorial sheet 3 Q.1 to assist in considering the effects of shear. Consider the beam to have the following
dimensions: t1 = t2 = 10mm, b = d = 50mm.

-Fdt1b/IX

-2Fdt1b/IX

-F(2dt1b + ½d2 t2 )/IX

Figure 10.3: The effect of shear loads on I beam displacement

75
Analysis of thin walled aerospace structures

10.3 Deflection in thin wall structures by unit load method


The deflection of thin-walled structures are also affected by bending and shear (and torsion should the
loads not be applied through the shear centre). The calculation of such deflections is easily obtained by use
of the unit-load method (recall semester 1 lectures).
10.3.1 Deflection due to Bending
Considering the effects of bending first, the increment in total complementary energy due to the
application of a virtual unit load may be shown to be:
 
∆C = − ∫  ∫ σ z ,1ε z ,0 dA dz + ∆M (71)
L A 
where L is the length of beam under consideration, A is the cross-section of the beam, z,1 is the stress in σ


the z-direction due to the virtual unit load, z,0 is the strain in the z direction due to the actual load case and
ε

M is the deflection of the beam due to bending at the point of application and in the direction of the of


the unit load. Since the system is in equilibrium, C = 0 and


 
∆M = ∫  ∫ σ z ,1ε z ,0 dA dz (72)
L A 
Equation (21) may be used to express this bending induced deflection in terms of applied moments:
M ′X ,1 M′
σ Z ,0 = y − Y ,1 x
IX IY
(73)
σ M′ M′
ε Z ,0 = Z ,0 = X ,0 y − Y ,0 x
E EI X EIY
Equation (72) may then be written as:

1   M X′ ,1 M Y′ ,1  M ′ M′  
∆M = ∫ ∫  y − x  X ,0 y − Y ,0 x  dA dz
E L  A  I X IY  I X IY  
1   M ′X ,1 M ′X ,0 2  M Y′ ,1 M ′X ,0 M ′X ,1 M Y′ ,0  M′ M′  
 xy + Y ,1 2 Y ,0 x 2 dA dz
E ∫L  ∫A 
= 2
y −  +   (74)
IX  I X IY I X IY  IY  
1  M ′X ,1M ′X ,0  M Y′ ,1M ′X ,0 M ′X ,1 M Y′ ,0  M′ M′ 
= ∫ −  +  I XY + Y ,1 Y ,0 dz
E L IX 
 I X IY I X IY  IY 
This simplifies dramatically if the section has at least one axis of symmetry so the IXY = 0

1  M ′X ,1M ′X ,0 M Y′ ,1 M Y′ ,0 
∆M =
E ∫L 
 + dz (75)
IX IY 
10.3.2 Deflection due to shear stress/shear flow
The shear stress induced deflection of a beam introduced by application of a virtual unit load may be
shown to be (in a similar way to bending above):
 
∆S = ∫  ∫ τ 1γ 0 dA dz (76)
L A 

76
Analysis of thin walled aerospace structures

where L is the length of beam under consideration, A is the cross-section of the beam, 1 is the shear stress τ

in the s direction due to the virtual unit load, 0 is the strain in the s direction due to the actual load case

γ

and S is the deflection of the beam due to shear at the point of application and in the direction of the of
the unit load. This may be re-arranged into an expression of shear flow as 1 = q1/t, 0 = 0/G = q0/tG and
τ

γ
τ

dA = tds to give:
 q1q0 
∆S = ∫  ∫ ds dz (77)
LA
Gt 
Equation (57) may be used to express the shear flows in terms of applied shear forces:
M ′X ,1 M′
σ Z ,0 = y − Y ,1 x
IX IY
(78)
σ M′ M′
ε Z ,0 = Z ,0 = X ,0 y − Y ,0 x
E EI X EI Y
Equation (72) may then be written as:

1   M X′ ,1 M′  M ′ M′  
∆M = ∫  ∫  y − Y ,1 x  X ,0 y − Y ,0 x  dA dz
E L  A IX IY  I X IY  
1   M ′X ,1 M ′X ,0 2  M Y′ ,1 M ′X ,0 M ′X ,1 M Y′ ,0  M′ M′  
 xy + Y ,1 2 Y ,0 x 2 dA dz
E ∫L  ∫A 
= 2
y −  +   (79)
IX  I X IY I X IY  IY  
1  M ′X ,1M ′X ,0  M Y′ ,1M ′X ,0 M ′X ,1 M Y′ ,0  M′ M′ 
= ∫ −  +  I XY + Y ,1 Y ,0 dz
E L  IX  I X IY I X IY  IY 

This simplifies dramatically if the section has at least one axis of symmetry so the IXY = 0

1  M ′X ,1M ′X ,0 M Y′ ,1 M Y′ ,0 
∆M =
E ∫L 
 + dz (80)
IX IY 

Determine the vertical and horizontal tip deflection for the unsymmetrical cantilever beam load by a
vertical end load F passing through its shear centre.

Y, v
L

X, u

Z
F

Figure 10.4: Idealised cross-section subject to pure bending

77
Analysis of thin walled aerospace structures

Bending moment and effective bending moments at any point z along the length of the beam are.
M X = F (L - z ), MY = 0
M X F (L - z ) I XY F (L - z )
M ′X = = , M Y′ =
K K KI X
Integrating Equation (69) twice using the boundary for a cantilever beam that at z = 0, du/dz = dv/dz = u =
v = 0 gives:

FI XY  z2 z3  FL2 I XY
u=  L -  ⇒ uT =
KEI X IY  2 6  3EKI X I Y

F  z2 z3  FL3
v=  L -  ⇒ v =
KEI X  2 6  3KEI X

78
Analysis of thin walled aerospace structures

Add section on unit load method for deflections (Megson section 4.8, 9.10 and 10.3.8)
Add a bit on why obtaining shear flow for a web is important (maybe deflection example or buckling from
Niu)
Add a bit on tapered sections (Megson 10.3.5)
Static Divergence?

79
Analysis of thin walled aerospace structures

11. CONCLUDING REMARKS

The scope of the present work is confined to the calculation of stress distributions and deflections of thin
wall sections according to the assumptions of the engineering theory of bending and the Batho theory of
torsion.
Whilst this may be considered adequate for many engineering purposes, additional analysis is necessary if
account is to be taken of the axial constraint problem. This problem arises if the warping displacements of
cross-sections, which occur as a consequence of shear deformations, are inhibited by the existence of
constraints. Such constraints may occur, for example, if the root section if forced to remain plane, or if the
torsion varies along the length of the beam. The effect of warping constraint is to induce in the beam an
additional system of axial stresses and corresponding shearing stresses.

80
Analysis of thin walled aerospace structures

TUTORIAL SET 4: DEFLECTIONS

(1) A uniform thin walled beam with cross section as shown carries a downward load P at its free end.
Derive expressions for the vertical and horizontal components of the deflection of the beam midway
between the supports B and D. The constant modulus of the beam is E. Ignore the effect of shear
stresses.
Answers: (u = 0.186PL3/Ea3t, v = 0.177PL3/Ea3t)
a y, v
2 3 L
a/2
t L x, u
1
2a
6 L
a/2 A
4 5
a B

Typical Cross-Section
C

z, w
D

P
Figure Q.1
(2) A thin walled cantilever with walls of constant thickness (t) has a cross-section as shown. The beam is
loaded by a vertical force P at the fee end and a horizontal force 2P at the mid-section. Both forces act
through the shear centre of the beam and thus produce no twist. Determine the stress in the beam at
point 1 and 2 on the cross-section at the built-in end and mid-section of the beam. Ignore the effect of
shear stresses.
Answers: (Built-in end: -1.85PL/td2, 0.1PL/td2 , Mid Span: -0.05PL/td2, -0.63PL/ td2)

Figure Q.2

81
Analysis of thin walled aerospace structures

(3) A uniform cantilever of arbitrary cross-section and length L has section properties IX, IY, and IXY. The
beam weighs w N/m. Determine magnitude of a horizontal force applied at the free end of the beam
that is sufficient to keep the deflection of the beam (due to its own weight) in the vertical plane (i.e.
deflection in the x direction should equal zero). Also determine the deflection of the beam at the free
end id the shear stress effect may be assumed negligible.
Answers: (3wLIXY/8IX, wL4/8EIX)
(4) A thin-walled cantilever has a constant cross-section of uniform thickness with the dimensions shown
in Figure Q.4. It is subjected to a uniform distributed load (-w) acting in the xz-plane. Ignoring the
effects of shear stresses:
(a) Derive expressions for the vertical and horizontal components of the tip deflection of the beam in
terms of W, L, E, a and t.
(b) Calculate the direct stresses at the points 1 and 2 of the cross-section at the built-in end when a =
30mm, w = -10N/m and L = 1000mm.
Answers: (u = 3wL/44Ea3t, v = 18wL4/44Ea3t, σ

z1 =-757.2MPa, σ

z2 =-253.7MPa)

Figure Q.4
(5) A uniform thin-walled beam of open cross-section with length L and thickness t, shown in Figure Q.5,
is simply supported at its end points with its web vertical. It carries a downward vertical force F at its
middle. Show that for this thin-walled cross-section with the uniform thickness t, the second moments
of area (IX, IY) and product moment of area (IXY) are; IX = 3.25at3, IY = 1.67a3t and IXY = -1.75a3 t.
If a = 20mm, t = 2mm, F = 500N, L = 2m and elastic modulus E = 200GPa, determine the vertical
component of the deflection of the beam midway between the support (Ignore the shear stresses).
Answer: (18.375mm)

Figure Q.5

82

Vous aimerez peut-être aussi