Tesi Ri Riorganiz
Tesi Ri Riorganiz
Tesi Ri Riorganiz
Candidate
Giovanni D’Onofrio
ID number 1697308
Abstract
The aim of this thesis is the preliminary design of a sandwich panel, conceived
to support a rotating mechanism and connect it to an Earth observation
satellite through aluminium brackets. The rotating mechanism includes a
flywheel whose purspoe is the balancing of vibrations induced by a rotating
antenna. The design of the panel is crucial for the satellite, considering the
high flywheel mass, the reached peak accelerations and the complex vibrational
environment during launch. In order to find an optimal configuration of
the panel, different core and skins thicknesses, core densities, number of
laminate plies and laminate stacking sequences were investigated. A carbon
fibre reinforced polymer (CFRP) and aluminium 7075 T6 were the evaluated
material for the skins, while aluminium was used for the honeycomb core. The
chosen design criteria are related to the modal decoupling between the system
and the satellite, and to weight minimization. The design is therefore based
on the value of the bending stiffness of the plate. Structural resistance to
the quasi-static loads exerted by the Vega C launcher was verified. Since the
system is excited by disturbs coming from the rotating mechanism imbalances,
a fatigue life assessment was needed on the brackets. The stress cycles caused
by the disturbances were identified on appropriate nodes by means of modal
frequency and transient analyses. The ESACRACK software was used to
calculate the damage accumulated during the entire load history of the system.
The optimal sandwich configuration involved CFRP symmetrical and balanced 8
ply laminates with stacking sequence [0/90/45/-45]s, and a 50.8 mm honeycomb.
A solution involving the same core and 0.5 mm aluminium skins can be adopted
in case of heat dissipation issues. It was verified that the critical point in
terms of fatigue of the brackets is a corner point, requiring the definition of a
proper fillet in a more detailed model. It also was determined that the flywheel
imbalances do not cause any fatigue damage to the structure.
v
Contents
Introduction 1
2 Composite theory 17
2.1 Micromechanics . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.1.1 Constitutive relation . . . . . . . . . . . . . . . . . . . 18
2.1.2 Elastic constants of a composite ply . . . . . . . . . . . 19
2.2 Macromechanics . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.2.1 Off-axis stiffness of a lamina . . . . . . . . . . . . . . . 20
2.2.2 Strain of a laminate . . . . . . . . . . . . . . . . . . . . 22
2.2.3 Constitutive equations of a laminate . . . . . . . . . . 23
2.3 Relevant stacking sequences . . . . . . . . . . . . . . . . . . . 25
2.3.1 Symmetric laminate . . . . . . . . . . . . . . . . . . . 26
2.3.2 Balanced laminate . . . . . . . . . . . . . . . . . . . . 26
2.3.3 Stacking sequence notation . . . . . . . . . . . . . . . . 27
2.4 Strength and failure criteria . . . . . . . . . . . . . . . . . . . 27
2.4.1 Strength of a unidirectional composite ply . . . . . . . 28
2.4.2 Failure criteria . . . . . . . . . . . . . . . . . . . . . . 28
Conclusions 85
1
Introduction
Before, during and after launch satellite structures are subjected to many
different types of excitation, such as static, sine, random, shock and acoustic
loads. The induced vibrations are transmitted from the launcher to the satel-
lite through their interface, or via the air inside the fairing. The satellite is
composed of a main structure and many lighter subsystems, each characterized
by its own dynamic behaviour, depending on its mass and stiffness. One
of the most important goal in the design of space structures is therefore to
avoid coupling between dynamics of the satellite and that of its subsystems,
since the resulting resonances can be destructive. During launch the load
environment is the most critical, because of the high static accelerations and
the simultaneous existence of multiple loads of different types. A second
design goal is therefore to achieve structural resistance to these loads, ensuring
that structures are stressed in their elastic field. In space applications high
mechanical properties must always be combined with low density, since weight
is associated with the power consumption necessary to put the satellite in orbit.
structures. CFRPs are produced in very thin plies constituted of carbon fibre
filaments embedded in a polymeric matrix. Plies are stacked with various
orientations to form a laminate, and laminates can be attached at the ends of
a honeycomb to form a sandwich structure. In a sandwich laminates are called
”skins” and the central honeycomb is called ”core”. The sandwich structure
is designed to withstand in-plane bending loads. The honeycomb is a very
light modular structure, generally made of aluminium hexagonal prisms, the
purpose of which is to allow an efficient use of the skin material.
The present work is part of the project for the development of an Earth obser-
vation satellite by Thales Alenia Space. The satellite is equipped with a large
rotating antenna, which can generate a nutation motion of the spacecraft and
vibrations which can disturb and degrade the functioning of sensitive payloads.
In order to balance its motion, an electrically driven rotating mechanism has
been designed to be mounted on the satellite. The preliminary design of the
sandwich panel on which the rotating mechanism is to be mounted is the
subject of this study. The panel is equipped with brackets for its mounting on
the satellite, and must accommodate the interfaces of the rotating mechanism.
Since this is a preliminary study, the aim is to determine the project feasibility
and its main features, taking in consideration the dynamic, static and fatigue
behaviour.
In the first two chapters the theoretical background of the study is presented.
The general architecture, the types of load, the typical disturbances and the
most common materials for satellite applications are the topics of the first
chapter, together with the definitions of laminate and sandwich. The second
chapter covers the main arguments of composite theory, with the determina-
tion of the constitutive equations of composite lamina and laminate, their
specification for symmetric and balanced laminates and the description of the
main composite failure criteria.
In the following sections the work is presented, starting with the description of
the finite element model. A model of the panel has been built with Patran
software, while the brackets have been imported from an existing model. The
flywheel has been modelled as a point mass, having already been designed to be
decoupled from the panel. The design of the panel has been carried out follow-
ing a dynamic criterion, together with mass minimization and static resistance.
The dynamic criterion deals with dynamic decoupling, traduced in the need
to have the frequencies of the first bending mode of the panel above a certain
threshold. To do this, the bending stiffness of the panel has to be sufficiently
increased with respect to its mass. Structural resistance deals with quasi-static
loads, which are the worst combinations of static and dynamic loads exerted
at any phase of the mission. Static and modal analyses were performed with
Nastran finite element solver and Patran pre- and post-processor. The use
Contents 3
of CFRP for the sandwich skins has been investigated and compared with a
high strength aluminium alloy. Despite having lower mechanical properties
and higher density with respect to CFRP, aluminium is still an alternative for
its superior thermal conductivity. In fact, the heat generated by the electric
motor driving the rotating mechanism could cause heat dissipation issues. In
addition to the different skin materials and thicknesses, cores of different thick-
nesses and densities have been considered. An optimization of the laminate
stacking sequence was necessary in the case of CFRP skins. The laminate
stacking sequence has been determined with a logic of minimization of the
maximum failure index on the most loaded ply. The most satisfying solutions
have been found, and an optimum region has been identified. The maximum
loads exchanged at the panel-bracket and panel-flywheel interfaces have been
collected for a future design of the joints.
Chapter 1
Background on satellite
structures
In this chapter some general notions, useful for the understanding of the subject
covered in this text, are given. After a brief introduction on the main satellite
subsystems, the focus is moved to the structural domain, which is central for
a mechanical design. The main load sources on a spacecraft are described
together with their main examples, with particular attention to the flight phase
of the satellite-launcher system. The centrifugal load caused by imbalances are
then explained in detail, since they are the reason for the fatigue verification
carried out on the brackets. Towards the end of the chapter the most common
materials in the airspace industry are briefly explained. A deeper view is given
for composites and the laminate and sandwich configurations in which they
find relevant applications.
The bus contains one or several payloads, namely the instruments or equipment
which performs the user mission [7]. An illustrative example of a satellite
exploded view is given in Figure 1.1 [8].
panels, which do not carry significant loads but provide support for smaller
components [5] [24].
The structure is the backbone of the satellite, providing:
• the required strength in order to bear high loads during the launch phase
and throughout the whole satellite life;
The shape of a satellite structure can depend on the type of attitude control.
The two main types of attitude control are satellite spin stabilization or three-
axis stabilization. The first one applies when the satellite axis has to be fixed
due to some mission requirements. The satellite attitude is be controlled with
a spin around this axis without additional effort. The structure in this case
has a cylindrical symmetry in order to balance forces and moments around
the spin axis. In the three-axis stabilization active attitude control is required.
Reaction wheels can be used to modify the orientation, each controlling the
rotation around one axis. For this reason the structure doesn’t need to have a
specific shape, and is often box-like.
The most critical environment takes place that during launch, due to the high
accelerations in lateral as well as in axial direction. Those excitation sources
are transmitted from the launcher to the spacecraft. The transmission can
8 1. Background on satellite structures
be structural via their interface, or acoustic via the air which surrounds the
satellite inside the fairing of the launcher. Once on orbit the acceleration levels
reduce drastically and, due to the decreasing air density, loads become purely
structural. Computing the micro-vibration environment which takes place on
orbit can be still useful in terms of performance requirements of high precision
pointing devices. Each launch phase causes a certain flight environment, which
is a source of load for the spacecraft. The intensity of the environments is
measured in terms of acceleration, and load distributions can be analysed
in the time or frequency domain, based on the convenience. The possible
environments generated during a spaceflight are:
• shock events (above 500 Hz) expressed in terms of the shock response
spectrum function (SRS)3 , usually of the acceleration.
These environments can occur simultaneously especially during the flight, and
their occurrence is variable depending on the flight conditions. P. Berlin [5] gives
a detailed description of the circumstances in which these load environments
take place during launch.
Quasi-static loads are expressed in terms of load factors, which are dimensionless
multiple of the gravitational acceleration representing the inertial force acting
on a structure. Those factors are provided in the launcher vehicle user’s
manual, and are employed in the preliminary design of a spacecraft structure
[19]. A launcher is subject to both lateral and longitudinal static accelerations
during launch, reaching values above 4g (39 m/s2 ). The loads associated to
1
QSL are considered equivalent to static loads, but they take into account both static
loads and dynamic loads with frequencies sufficiently below the first natural frequency of
the structure
2
Power Spectral Density expresses the power distribution of a signal as a function of
frequency
3
Shock Response Spectrum is defined as the maximum response of a singe degree of
freedom system as a function of its natural frequency and for a given damping ratio
1.2 Structural loads 9
peaks in figure are associated to the time intervals between ignition and cut-off
of each solid propellant stage. The acceleration time history is not constant
in these three phases because of propellant consumption, which reduces the
system mass. When the first three stages have been jettisoned and the vehicle
is already out in space, the AVUM liquid propellant stage allows to place the
payload in the prescribed orbit with very small acceleration.
An example of sinusoidal low frequency vibration is the POGO effect4 , which
takes place during launch. The vibrations due to this effect derive from
resonances in the rocket fuel lines and can induce critical stresses on the
satellite. Acoustic noise is experienced in the first moments of lift-off and when
the system reaches transonic speed. In the first case pressure waves bounce
back from the ground in the launcher, while in the second case the acoustic
noise is generated in the boundary layer on the external fairing surfaces. In
addition to the random vibrations generated by acoustic phenomena mechanical
random vibrations can take place, such as that those coming from propellant
combustion in the engines. Shock loads result from very short and violent
transients, such as stage ignition and cut-off, and fairing jettison. A situation
similar to a pressure vessel can happen in the fairing as the external pressure
decreases, requiring the presence of venting to prevent high pressure differences.
4
PrOpulsion Generated Oscillations
10 1. Background on satellite structures
Out in space structural loads can be induced for example by thermal gradients
caused by solar radiation. This phenomenon has to be taken into account also
since in space no heat exchange through convection exists. Small meteorites
and orbital debris are abundant in space and can hit the satellite surface. Even
very minute solids can cause significant damage due to their high velocity (up
to 7.5 m/s). Moving parts can cause shock loads on the structure, like solar
panels during their unfolding. Other components like reaction wheels and
thrusters are sources of micro-vibrations which have to be taken into account
because they are severe disturbs for high precision instruments.
The frequency range of excitation is of primary importance in that it affects the
applicability of the analysis methodologies. The term ”low frequency” doesn’t
refer to a single and fixed range, but rather to the interval between 0 Hz (static
excitation) and the first natural frequencies of the analysed structure. The high
frequency range contains a great number of modes thus it is characterized by
high modal density. In the high frequency domain the contribution of a mode
to the total response is drastically reduced. For low frequencies of excitation,
applying a modal approach calculating each mode in the frequency interval
of interest is an efficient way to predict the structural response accurately.
That’s why in this cases Finite Element Analyses (FEA) are widely used.
At higher frequencies the use of FEA becomes progressively prohibitive in
terms of computational cost needed to reach sufficient accuracy. To study the
high frequencies other models such as the Statistical Energy Analysis (SEA)
stochastic model have been developed.
the fact that a static imbalance can be detected without putting the wheel in
rotation, while a dynamic imbalance shows up only on a rotating system. They
are both due to constructive asymmetries with respect to the axis of rotation,
and usually they can be minimized but not avoided during manufacturing.
The existing models for these disturbances are showed in Figure 1.3 and are
briefly explained [19]. An imbalance is said to be static when the rotor CoG is
eccentric of a quantity rs with respect to the spin axis [1]. This situation is
modelled with a mass ms representing the imbalance, placed at a distance rs
from the rotation axis. The rest of the rotor mass is perfectly balanced. This
mass put under rotation generates a centrifugal force rotating at the same
frequency as the flywheel. The force magnitude is:
Fs = ms rs ω 2 = Us ω 2 (1.1)
exists, the rotational axis of is no more a principal axis of inertia, and the
inertia tensor I is no longer diagonal. Its non diagonal terms are equal to Ud
[2].
The rotating centrifugal force and moment generally occur together, in the
plane orthogonal to the rotation axis. The moment vector is in quadrature
with respect to the force because of the vector product. The magnitudes are
both proportional to the square of the frequency, so in high speed devices
the centrifugal forces may become dangerous even with a small value of the
imbalances Us and Ud .
1.4 Materials
To meet the requirements specified in Section 1.1, several types of materials
have been developed, which in principle can be classified as metal alloys and
composites. Common metal alloys today used in spacecraft structures are [23]:
• aluminium alloys, widely employed for their moderate cost, light weight,
ductility, high specific stiffness and specific strength, good fatigue re-
sistance and high thermal conductivity. They are however sensitive to
corrosion and moisture, and their temperature range of applicability is
low, below 150°C;
• titanium alloys, with moderate density, good resistance to corrosion, good
mechanical properties which hold at high temperatures (up to 500°C)
and long fatigue life. The disadvantages are related to their high cost
and low ductility;
• steels, characterized by good mechanical properties and employed in
heavily loaded structure, where the main requirement is strength. Steels
are not widely used in aerospace applications due to their high density,
which is up to 2.5 times higher than aluminium alloys.
A variety of composite materials has been developed in the past decades to
overcome the mechanical behaviour of metal alloys. Composite materials are
systems usually composed of a filler imbedded in a matrix. These two main
elements together form a ply, which is generally very thin, below 1 mm. The
filler is constituted of fibres disposed along one direction. The consequent
anisotropic mechanical behaviour allows to obtain the best performances when
fibres are oriented in the load direction. Fibres provide strength, stiffness and
fatigue resistance and are extremely light. The most employed fibre material is
carbon (graphite), but also fibreglass or aramid (Kevlar) find applications. The
matrix is an homogeneous and ductile material. Its function is to hold fibres
together ensuring their simultaneous work under loading. Matrices can be
made of metal and ceramic, but polymeric materials are by far the most used.
Epoxy, a thermosetting polymer, is the most affirmed material for composite
1.4 Materials 13
Table 1.1. Tensile properties of the main fibres used in aerospace applications
cracks. These imperfections are created when the thin filaments are joined into
bundles before being imbedded in the matrix. It is very difficult to prevent
this process.
The sandwich structure is built attaching two skins to the sides a central core,
through two adhesive films. The core is often an honeycomb, a low-density
modular structure composed of hexagonal prisms, with axes perpendicular
to the skins. The materials used for the core are mainly aluminium and
Nomex6 , but also fibreglass and CFRP are used. Aluminium is preferred when
low weight is a primary constraint and the component is not heavily loaded,
or when good thermal conduction is required between the two faces of the
sandwich. The skins can be anisotropic laminates or single plies of isotropic
metal alloys (mainly aluminium), while for adhesives epoxy or other polymers
are used. The peculiar properties of sandwich structures are the low weight
and buckling resistance provided by the honeycomb, and a good stiffness to
weight ratio. Designing it properly, a sandwich structure can ensure higher
6
material composed of aramid fibres in a phenolic resin matrix
16 1. Background on satellite structures
Chapter 2
Composite theory
A large part of this chapter is taken from Prof. Gaudenzi’s lectures [11],
and from R. M. Jones’ book [12]. In this chapter only the most important
equations will be derived, and the fundamental path which allows to derive the
remaining ones will be explained. The mechanics of composites includes two
complementary theories, micromechanics and macromechanics. The first one
is the study of the properties of a single ply through those of its component
materials, while the second is about the stacking of plies to form a laminate and
the influence of the properties and disposition of the plies with the laminate
structural response.
2.1 Micromechanics
The composite ply is a continuous system composed of matrix and fibres, so
the mechanical properties of the lamina are combinations of those of these two
phases. The realistic hypotheses that each phase is isotropic, and homogeneous
and perfect bonding exists between the two phases are made. Assuming this,
to characterize the behaviour of the constituents, only Young’s modulus, shear
modulus and Poisson’s ratio are needed.
Ef
Ef , ν f , G f =
1 + νf
Em
Em , νm , Gm =
1 + νm
stress become:
− νE121
1
0
E
ν112 1
F = − E1 E2
0
(2.5)
1
0 0 G12
E1 ν12 E2 0
1
Q= ν12 E2
E2 0
(2.6)
1 − ν12 ν21
0 0 G12 (1 − ν12 ν21 )
where only the first, second and sixth rows have been extracted from Equation
2.3. The remaining out-of-plane strains γ23 and γ31 are zero, while ε3 can be
found from the third row of Equation 2.3 applying the plane stress condition.
The direction of fibres x is be called longitudinal direction L, and y is the
transverse direction T . The ply lies on the LT plane, and the rewritten
constitutive relation are:
1
− νETTL 0
εL EνLLT
σ
L
1
− EL
εT = 0 σT (2.7)
ET
1
γLT 0 0 G
τLT
LT
Ef Em
ET = (2.8)
Ef vm + Em vf
1
A compatibility equation is written on the displacements, and displacements are related
to strains through the definition of strain. The compatibility equation expresses the total
displacement in a certain direction as a function of the displacements of fibres an matrix
20 2. Composite theory
Gf Gm
GLT =
Gf vm + Gm vf
Note that in the first two equations the term containing Em is much smaller
than the one with Ef . In the first equation it means that EL is fibre dominated,
while in the second it means that ET is matrix dominated. In the third relation
GL T is matrix dominated, because Gm is much smaller than Gf . This means
that an orthotropic ply behaves mainly as its fibres with respect to longitudinal
stresses, and as its matrix with respect to transverse and shear stresses. For
the Poisson’s coefficients, their definitions and the compatibility equation on
the transverse direction allow to obtain the poisson equivalent coefficient of
the ply.
νLT = νf vf + νm vm (2.9)
With these four relations one can derive the elastic constants of the ply knowing
those of the constituents, and then write the stiffness and flexibility matrices
under the made assumptions.
2.2 Macromechanics
In this section the fundamental equations which govern the elastic behaviour
of a laminate will be derived under some simplifying assumptions on stresses
and strains. The hypothesis allow to reduce a three-dimensional problem to a
simpler two-dimensional one, with good confidence on the obtained model.
The basic assumption is of perfectly bonding in the laminate, so that the
laminae work perfectly together when stacked on the laminate. This can
be true if the adhesives are strong enough. A second hypothesis is that the
displacement distribution is continuous at the plies interfaces, so that no
relative slip is allowed. Under these assumptions the laminate acts as a single
plate.
The aim is to derive the stress-strain relation for the single layers for a general
in-plane orientation, and define the forces and moments on a laminate through
those on the individual plies. Combining together these 2 systems of equations
we will be able to derive the stiffnesses of a laminate.
cos2 θ sin2 θ
2cosθsinθ
2
T = sin θ
cos2 θ −2cosθsinθ (2.10)
−cosθsinθ cosθsinθ cos2 θ − sin2 θ
τ12 τxy
ε1 εx
ε2 = T εy (2.12)
ε12 εxy
Recall the definition of stiffness matrix in plane stress, which is Equation 2.6.
σ1 Q11 Q12 0 ε1
σ 12 = σ2 = Q12 Q22 0 ε2 = Q ε12 (2.13)
12
τ12 0 0 Q66 γ12
Shear strains are defined as γij = 2εij . The matrix R allows to pass from
the shear vector containing ε12 (or εxy ) to that containing γ12 (or γxy ). This
matrix allows to write Equation 2.14.
ε1 ε1 εx
−1 −1
ε12 ε2 = R ε2 = RT R εy = RT R ε xy
= (2.14)
Equation 2.11 is inverted, and Equations 2.13 and 2.14 are substituted into it.
σ xy = Q εxy (2.16)
xy
We are dealing with lamina whose principal coordinates are rotated of an angle
θ with respect to a chosen reference (x, y, z). The matrix Q contains the
xy
22 2. Composite theory
transformed reduced stiffnesses, and allows to write Equation 2.16, which links
stresses and strains of the plate in the (x, y, z) coordinate system. The matrix
is in general full.
Q11 Q12 Q13
Q = Q21 Q22 Q26 (2.17)
xy
Q61 Q62 Q66
The Qij are made explicit in Jones’ book [12], and they have the properties:
Knowing Q for any ply, it is possible to calculate the stiffness matrix of the
xy
laminate. Omitting for simplicity the (xy) subscripts, the stress-strain relation
for the k th layer of a composite laminate, oriented of an angle θk , is given by
Equation 2.20.
σk = Q εk (2.20)
k
It has to be noticed that while for the hypothesis of perfect bonding the strain
distribution through the laminate thickness is a line, the stress distribution is
not, since the Qij are in general different for each lamina. This means that
in the xz or yz plane the strain distribution will be a line, while the stress
distribution will be a broken line.
Let (u, v, w) be the components of the laminate displacement along the (x, y, z)
directions, and mark with the 0 subscript the quantities referring to the middle
plane. As a result of the Kirchhoff’s hypotheses, the displacements at any
point of the plate is a linear function of the z coordinate.
∂w0
u(x, y, z) = u0 (x, y) − z
∂x
2.2 Macromechanics 23
∂w0
v(x, y, z) = v0 (x, y) − z (2.22)
∂y
w(x, y, z) = w0 (x, y)
By derivation of the displacements, strains are obtained.
∂u0 ∂ 2 w0
εx = −z
∂x ∂x2
∂v0 ∂ 2 w0
εy = −z (2.23)
∂y ∂y 2
∂u0 ∂v0 ∂ 2 w0
γxy = + − 2z
∂y ∂x ∂x∂y
ε0x
εx χx
0
εy = εy + z χ y (2.24)
0
γxy γxy χxy
Figure 2.1. Forces and moments on the mid plane of a lamina, general case
∂uj 2 2
2
εi = ∂ui
∂xi ; γij = ∂ui
∂xj + ∂xi ; χi = − ∂∂xu2k ; χij = −2 ∂x
∂ uk
i ∂xj
i
24 2. Composite theory
These loads are forces and moments per unit of width. The integrals of the
induced stresses along thickness, that is between the positions zk−1 and zk ,
must be equal to the resultant forces and moments. Recalling Equations 2.20
and 2.24 we can write:
Nx Z zk σx Z zk Q11 Q12 Q16 εx
Ny = σy dz = Q12 Q22 Q26 εy dz = (2.25)
zk−1 zk−1
Nxy k τxy k Q16 Q26 Q66 k γxy
Mx Z zk σx Z zk Q11 Q12 Q16 εx
M
y
= σ
y zdz =
Q
12 Q22 Q26 εy zdz =
(2.26)
zk−1 zk−1
Mxy k τxy k Q16 Q26 Q66 k γxy
where it has been assumed that the ply stiffness matrix is constant within the
ply itself. This is true unless a temperature or moisture gradient exists along
its thickness.
For the entire laminate of total thickness h, composed of N plies of thickness
hk = zk − zk−1 , forces and moments are expressed in Equations 2.27 and 2.28.
The middle plane strains and curvatures are constant with z, so they can be
taken outside the integral. The product of Q and the integrals gives the
k
stiffness matrices of the laminate.
ε0x
Nx A11 A12 A16 B11 B12 B16 χx
0
Ny = A12 A22 A26 εy + B12 B22 B26 χy (2.29)
0
Nxy A16 A26 A66 γxy B16 B26 B66 χxy
2.3 Relevant stacking sequences 25
ε0x
Mx B11 B12 B16 D11 D12 D16 χx
0
My = B12 B22 B26 εy + D12 D22 D26 χy (2.30)
0
Mxy B16 B26 B66 γxy D16 D26 D66 χxy
The Aij terms are the extensional stiffnesses of the laminate. A11 and A22 are
related to extension and A66 to shear. The term A12 is the extension-extension
coupling term, while A16 and A26 represent the shear-extension coupling. The
Bij coefficients couple extension and bending, and they constitute the principal
difference in the behaviour of a laminate with respect to the lamina. Their
presence means that loading a laminate with an extensional or shear force
can result in bending (B11 ,B12 ,B22 ) and twisting (B16 , B26 ,B66 ), and also that
a laminate subjected to moment will deform on its plane by extension and
shearing strain. The Dij are bending stiffnesses. The bending terms are D11
and D22 , while D66 is related to twisting. D12 is the bending-bending coupling,
while D16 and D26 represent the bending-twisting coupling.
Equations 2.29 and 2.30 are the constitutive equations for a composite laminate.
They express its mechanical behaviour, which appears in general to be very
complex due to the presence of many coupling terms. That’s why in practice
engineers search for laminate configurations which allow to delete the coupling
terms, reducing the number of stiffness coefficients and giving some practical
advantages in the use of laminate and sandwich structures.
general case where all the coupling terms exist, but in space applications some
particular configurations are employed, which simplify the design.
ε0x
Nx A11 A12 A16
0
Ny = A12 A22 A26 εy (2.32)
0
Nxy A16 A26 A66 γxy
Mx D11 D12 D16 χx
My = D12 D22 D26 χy (2.33)
Xε,c < ε1 < Xε,t Yε,c < ε1 < Yε,t |γ12 | < Sε (2.37)
σ12 σ1 σ2 σ22 2
τ12
− + + <1 (2.38)
X2 X2 Y2 S2
σ12 σ1 σ2 σ2 Xc + Xt Yc + Yt τ2
− + − 2 + σ1 + σ2 + 122 < 1 (2.39)
Xc X t Xc Xt Yc Yt Xc Xt Yc Yt S
The presence in the equation of different strengths for tension and com-
pression implicate a difference in the stress space representation with
respect to Tsai-Hill. The limit surfaces into which the criteria are satisfied
are both ellipsoids, but Hoffman is represented by a single ellipsoid while
in Tsai-Hill 4 different portions of ellipsoid are traced in each of the
quadrants of the σ1 − σ2 plane. Those portions together form a unique
surface, but they are different in slope. For equal strengths in tension
and compression the two criteria become equivalent.
Chapter 3
This chapter contains a description of how the finite element model of the system
was built, of the employed materials and of the selected element properties.
The finite element solver and the pre-and post- processor used for the design
are respectively MSC Nastran and Patran. The panel is a circular plate meant
to be mounted on the bottom side of a satellite, around its longitudinal axis. A
number of brackets have to be positioned circumferentially on the panel. The
brackets have the function to constrain the plate to the satellite on their upper
surface. The mounting of the system on the satellite is guaranteed on the top
faces of each bracket. A heavy rotating mechanism has to be mounted on the
panel through 3 equally spaced rectangular interfaces of known dimensions.
The FEM model has been built starting from an already existing one, which
has been modified to be consistent with the present application. Figure 3.1
shows this initial model. The panel lies on the xy plane and has a radius of
a distance of 0.45 m from the plate centre, are connected to the panel. Each
connection is modelled with 2 Multi-Point Constraints (MPC) of type RBE2
and a BAR2 element. The BAR2 element connects two coincident nodes,
placed between the lower surface of the bracket and the plate, and a scalar
bush property is assigned to it. The RBE2 is a rigid multi-point constraint,
which in this case binds all the 6 DoFs of the dependent nodes to those of an
independent node. The dependent nodes are placed on the plate and on the
bracket. The bar element with the bush property is in practice a spring with
very high stiffness, used to extract the forces and moments exchanged between
panel and the bracket. Knowing this loads it is possible to dimension the bolts.
Another similar system is used to represent the interfaces between the upper
surface of each bracket and the satellite, where the constraints are placed.
In each of the upper springs, the node belonging to the satellite has been
linked with a bar element and a bush property to the correspondent node of
the bracket. The system is constrained to the satellite in many nodes with
springs of proper rigidity, to ensure the solidity of the mounting.
In the original model the plate was constituted of both QUAD4 and HEX8
elements. The elements belonging to the plate have been deleted, and a new
mesh was created for the plate using QUAD4 and TRIA3 elements. The mesh
has been built setting the dependent nodes of the RBE2 constraints as hard
points. The dimension of the element have been reduced to 6.5 mm, while that
of the QUAD4 elements used for the brackets is 4 mm.
Shell properties are defined on the elements of both plate and brackets. The
thicknesses of the bracket surfaces vary from 3 mm to 10 mm. Given the low
thickness, it is possible to model the elements as shells. This is possible also
for the panel, whose thickness is much smaller with respect to the diameter.
A non structural mass has been added to the panel to take into account non
modelled items, which are mainly the 2 sandwich adhesives and the inserts,
33
necessary to join the panel with the other components. The value for the non
structural mass is estimated with the inverse of the plate surface, and is equal
to 1.273 kg.
The rotating mechanism has already been designed, so the all relevant data
for its modelling are known. Its most relevant characteristics is the 99 kg
mass and its moments of inertia, which are are Ix = Iy = 4.77 kg · m2 and
Iz = 9.54 kg · m2 . The moments of inertia are calculated with respect to
the principal axes, centred on its CoG. To model the rotating mechanism, a
lumped mass property has been added to a point element, placed at a vertical
distance of 10 cm from the panel. This simple modelling is possible because
it is known that the flywheel is dynamically decoupled from the panel. In
fact the natural frequencies of the rotating mechanism fall into a frequency
range above 140 Hz, while the main modes of our system are surely below 100
Hz. The rotating mechanism has been designed to have this modal behaviour
since the vibrational decoupling between mechanically interacting systems is
crucial in the design of space structures. This phenomenon will be explained
further in detail in the next chapter. The radius of its flywheel is 0.44 m, so
it can fit between the brackets. The wheel is connected to the plate through
3 equispaced radial interfaces, placed at a distance of 19 mm from the edge
of the panel. Each interface models a bolted joint with 4 bolts placed at the
corners of a rectangle. The interface dimensions are 97 mm radially and 85
mm circumferentially. Each bolt is modelled with a BAR2 element with a
bush property and an RBE2 constraint as it was done for the interfaces of the
brackets. It is important to notice that each of the springs at the rotating
mechanism interface model 1 bolt, while those at the bracket interfaces model
the entire bolted joint. This mean that the loads at exchanged at the panel-
bracket interfaces have to be divided by the number of bolts constituting the
joint.
To place the wheel on the panel, it was necessary to eliminate 2 brackets. The
two brackets, symmetrical with respect to the y axis, were deleted in order to
make room for the interfaces of the rotating mechanism. The obtained model
is showed in Figure 3.2.
For the core of the sandwich panel, 2 types of aluminium honeycombs are
available. The materials used is aluminium 7075 T6 for both, but one is denser,
having smaller hexagonal cells. Honeycomb materials have been defined in
Nastran as orthotropic materials. To distinguish among them they have been
named Honeycomb 50 and Honeycomb 70, referring to their density values. The
relevant elastic properties of the cores are listed in Table 3.1. Since honeycombs
are not designed to bear in-plane normal stresses, the in-plane elastic moduli
and Poisson’s ratio have not been reported and very small values have been
set for them. The possible core thicknesses are 25.4 mm (1 inch) and 50.8
mm (2 inches). For the reasons cited in Section 1.5 a thicker honeycomb will
guarantee better flexural and vibrational performances, at the price of a very
small weight increase. The skins are designed in two different configurations:
34 3. Finite element model
Honeycomb 50 Honeycomb 70
• CFRP laminates.
allows a weight saving around 40%. The aluminium choice is still valid due
to lower cost and excellent thermal conductivity. This last aspect is relevant
since the electric actuator which drives the rotating mechanism generates heat.
A part of this heat may have to be dissipated by the panel, so the aluminium
option is taken into account for future design.
Besides constitutive models, strength values have been set for orthotropic
materials to be able to apply a failure theory. For the two orthotropic core
materials only the shear stress limit has a physic value, which is of 1.0 MPa
35
for Honeycomb 50, and 2.2 MPa for Honeycomb 70. The values of other stress
limits have been set very high, since the correspondent failure modes are not
allowed for the core. The only relevant failure property of the 7000 series
aluminium is the yield strength of 430 MPa. No plastic behaviour is possible
for honeycomb and skins. Stress limits of CFRP are reported in Table 3.4. For
Xt Yt Xc Yc S
1776 30 715 95 56
honeycombs and CFRP, a value of 15 MPa has been set for the bonding shear
stress limit, given by the adhesive plies which bind the skins to the core. The
material of the bracket is the same aluminium used for the panel skins. The
elastic properties of brackets are the same as those listed in Table 3.2.
Material properties are used build a laminated composite material. It contains
informations about the number, thickness, material and orientation of skin plies
and core. The laminated composite material is the main input given to the shell
property of the panel. The core is modelled into the laminated composite table
as a ply of appropriate thickness. The operative part of the design is to find
laminated composite configurations which optimally satisfies the design criteria.
37
Chapter 4
In this chapter informations about applied loads, criteria of the design and
relevant parameters are given, in order to justify the choices made in the choice
of the optimal sandwich configuration, and to explain the settings used for
the modal and static analyses. After an explanation of the applied dynamic
criterion, fundamental in the design of space structures, a description of the
quasi-static loads used for the static verification is given, together with the
chosen thresholds. For what concerns the results, the bending modes of the
panel are showed, and then the actual comparative study for the derivation of
an optimal panel configuration is explained. On the optimal configurations
further static analyses, necessary for the complete static characterization of
the panel, have been carried on. At the end of the chapter some further
considerations on the interface loads and on the stresses on the brackets are
made.
It is important to point out that the main design criterion which has to be
satisfied is related to the dynamics of the system. A comparison between the
dynamics of the system which includes panel, brackets and flywheel with those
of the the rest of the satellite is crucial for a good design of a space structure.
This is why this design is based on the fulfillment of a dynamic requirement,
verified through modal analyses. The following static analyses are just needed
for a necessary verification of the material resistance to the most severe loads
applied before and during launch.
The main interest in the study of the panel vibrational behaviour is on the
frequency of the principal out-of-plane bending mode. In fact, due to its geom-
etry and constraints, the most relevant mode for the panel is the one which
causes the displacement of the majority of the panel mass in the z direction,
that is the out-of-plane direction. Regardless of the particular configuration of
the skins and core of the sandwich, this mode is always the first main mode of
the system, and it is characterized by a high value of the modal effective mass
fraction associated to the z translation.
have to be multiplied again by a design factor, which for satellites is the product
of three different factors: qualification factor KQ; model factor KM and project
factor KP . The qualification factor KQ is used to determine the qualification
loads, which are loads used for the dimensioning or test qualification of a
structure. The project factor KP is related to the maturity of the program,
and its value is progressively reduced during the development of the project.
The model factor KM takes into account the uncertainties of the mathematical
model. Design limit loads are obtained multiplying the limit load factors by a
total coefficient 9.81 · 1.25 · 1.1 · 1.2. In this preliminary design loads are being
strongly increased, since we want to give a very conservative design solution
to compensate the uncertainties present at this level of analysis. In future
designs the model can be refined and the safety factors decreased. The obtained
loads are tabled in Table 4.1. Those design loads are inertial accelerations,
applied at the system CoG. Note that some load events are much more critical
than others, and that longitudinal compression accelerations are higher than
longitudinal tension accelerations. This means that compression load cases are
more dangerous, and among them the worst cases are lift-off, 1st stage flight
with maximum acceleration and tail off, and most of all 2nd stage ignition and
flight, 3rd stage ignition.
Lateral loads can have any direction on the xy plane. It is common practice
to verify 8 in-plane directions at regular intervals of 45°, covering the entire
circumference from 0° to 315°. The 0° direction is the x and the 90° direction
is y. For each of the 6 flight phases 2 load cases are examined, coupling the
4.3 Conditions for the static verification 41
QSL [m/s2 ]
Load Event Longitudinal Lateral
Compress. Tension
1st stage flight with max. acc. and tail off -80.9 16.2 ±11.3
2nd stage ign. and flight, 3rd stage ign. -80.9 48.6 ±21
of the Hoffman failure surface, while the actual stress has to be sufficiently
inside it. The software calculates the principal stresses on each element of each
ply, and uses them to find failure indexes. If F I > 1 ply failure has occurred.
In this work we will consider laminate failure to take place when just one ply
has failed, adopting a first ply failure criterion. This is not always true, since
a redistribution of stresses could lead to a non-critical stress state after the
failure of a single ply. The approximation is justified by the fact that it is in
favour of safety and leads to a simple elaboration of the results.
A Factor of Safety (FoS) is considered to define the Margin of Safety (MoS),
which has the expression [23]:
σallowable
M oS = −1 (4.2)
σactual · F oS
The static criterion used to assess structural integrity is simply:
M oS > 0 (4.3)
where the FoS value is 1.2 for aluminium and 2 for CFRP. Very high factors of
safety are being used especially for CFRP skins, because we are preliminarily
ensuring the feasibility and robustness of the project. For aluminium, since
σallowable = 300MPa, from Equations 4.2 and 4.3 an equivalent expression for
the static criterion can be found.
For CFRP the equivalent static criterion is found recalling also the definition
of failure index.
F I < 0.5 (4.5)
The inequalities of Equations 4.4 and 4.5 are the conditions checked in the
static analyses under several combinations of QSL, in order to asses structural
integrity.
laminates.
A common 8 ply balanced and symmetric stacking sequence used in sandwich
structures is [0\90\45\ − 45]s, where the 0◦ external ply is the farthest from
the bending neutral axis of the sandwich, while the −45◦ one is the closest.
The θ = 0◦ direction coincides with the first principal direction of the laminate,
and with the x axis of the model. The θ = 0◦ and θ = 90◦ plies carry most of
the in-plane normal stresses, while the θ = ±45◦ layers are inserted to sustain
in-plane shear stresses. For this reason, a CFRP laminate has to include at
least one ply oriented in each of these 4 directions (0◦ , 90◦ , +45◦ , −45◦ ). It is
good practice to not put ±45◦ plies externally in a laminate.
An example of sandwich setup in Patran with [0\90\45\ − 45]s stacking
sequence is given in Figure 4.2. Skins and core have been grouped into black
rectangles. CFRP and honeycomb materials are respectively mat.2000001 and
the margin of safety is negative, some design changes are necessary. Those
modifications include:
• increasing aluminium skin thickness, starting from a value of 0.3 mm at
intervals of at least 0.2 mm;
• increasing the number of CFRP plies from a minimum of 8, adding 2
plies per skin if their orientation is 0◦ or 90◦ , and 4 plies per skin if it is
±45◦ to maintain a balanced and symmetric configuration;
• changing the core thickness from 25.4 mm to 50.8 mm, which allows to
consistently increase the natural frequencies and the flexural behaviour
of the skins, at the cost of a consistent mass increase;
• changing the core material, since Honeycomb 70 is denser and more
performing than Honeycomb 50, leading to better results both in dynamic
and static terms with an increase in mass.
With each of these design changes the sandwich stiffness rises. Also the
system mass rises. Considering that the main contribution to the mass of
the entire system is not given by the panel but by the 99 kg of the rotating
mechanism, and that the stiffness is only given by the panel, it’s clear how
these modifications increase the natural frequencies of the system.
s
k
fnat ∝ (4.6)
m
Some of the extracted outputs depend on whether skin material is isotropic or
orthotropic. Von Mises stresses are plotted in the first case, failure indexes in
the second. The margin of safety positivity is then assessed. This is equivalent
to consider the inequalities of Equations 4.4 and 4.5.
Forces and moments on the bush elements are extracted to have an estimate
of the loads exchanged with brackets and rotating mechanism. We need these
loads for a future design of the bolted joints connecting the panel to brackets
and wheel. The equilibrium of forces and moments for the brackets leads to
the equality of the loads exchanged from a bracket to the panel and to the
satellite.
The point mass representing the rotating flywheel is by far the heaviest part
of the model, therefore it is the main source inertial forces. In the model loads
are transmitted from the wheel to the panel through an RBE2 multi-point
constraint. This constraint rigidly transfers the displacements to the panel,
as if the rotating mechanism supports were infinitely rigid. The highest loads
are therefore on the sandwich regions around the RBE2 dependent nodes. We
expect these interfaces to be the most critical regions. Von Mises stresses on
the brackets are monitored in this analysis, but the aluminium brackets are not
expected to be critical since they are not in direct contact with the rotating
mechanism.
4.5 Modes of the panel 45
(a)
(b)
(c)
Figure 4.3. Aluminium skins, shapes of the panel modes: 1st (a), 2nd (b)
and 3rd (c)
of the panel is always the first in terms of natural frequency. In practice, the
situations with the frequency of the main bending mode above 40 Hz and
4.6 Comparative study of admissible configurations 47
satisfying the static verification, have been considered good. Only the force
components of the interface loads have been reported, to avoid the presence
of too many outputs. The interface moments are in any case 2 orders of
magnitude lower than forces, since the applied inertial loads are translational
accelerations.
The results have been tabulated in Excel, and the most significant and ex-
planatory data are here reported. It is important to remark that the panel
mass values include 1.27 kg of non structural mass, representing non-modelled
items.
For the aluminium skins, each of the 4 possible honeycomb combinations was
examined, which hare obtained changing the material and thickness of the
core. For every core configuration, the minimum thickness of the skins which
allows to satisfy the design criteria was found, with the objective of weight
minimization. The results are showed in Figure 4.4. The first 3 columns define
the sandwich, the 4th and 6th are related to the design criteria, the 5th concerns
the static verification, and the last 3 are needed for optimization. Each row
defines a tested configuration, and some adjacent rows containing the same val-
ues have been joined together to give a clear view. The aluminium sandwiches
with the 25.4 mm core need very thick skins, thus their mass is excessive with
respect to other good solutions. That’s why they have been discarded without
further analyses. With a 50.8 mm core of the lighter material (Honeycomb 50)
we obtain the best performances for the aluminium case. Skins are only 0.5
mm thick, and maximum Von Mises stresses are not critical (MoS=1.45). This
is a first relevant result of this study: in general the use of a thicker honeycomb
raises the natural frequencies allowing to use thinner skins, with a good saving
in mass. The Von Mises stress distribution on the critical skin, which is the
lower one, is given in Figure 4.5. The components of the inertial load are (14.9,
-14.9, -80.9) m/s2 , where the x and y components correspond to a in-plane
load of magnitude 21 m/s2 . The three interfaces with the rotating mechanisms
48 4. Design of the panel
Figure 4.5. Von Mises stress distribution, aluminium skin, lower skin, inertial
load (14.9, -14.9, -80.9) m/s2
are visible, since in those points the stresses reach the highest values. The
most critical region corresponds an interface which lies in the direction of the
lateral load. The rest of the plate is substantially unloaded. This situation,
with the plate unloaded except for the regions around some interface nodes,
happens with every plate shape and load case.
In the CFRP skin case a further degree of freedom is given by the choice of
the stacking sequence. The stacking sequence affects the stress distribution in
each ply, then it changes the value of the maximum failure index. The other
parameters are not affected, since no mass and stiffness are added or subtracted
to the system. Figure 4.6 shows the relevant cases for a CFRP skin with 8 plies.
The analyses with a 25.4 mm core have not been reported, because in that case
the modal design requirement can not be satisfied. With the thicker core good
configurations are found, with a lower mass with respect to the aluminium
skin, and with almost the same MoS. The principal out-of-plane mode has a
natural frequency of 55.4 Hz, which is 15.4 Hz higher than the threshold. This
4.6 Comparative study of admissible configurations 49
is a very good dynamic result and it is combined with low weight. The total
skin thickness is 0.6 mm. Figure 4.7 shows the failure index distribution in the
most critical load case and on the critical ply, which is the 45◦ external ply of
the upper skin. The anisotropic behaviour of the carbon fibre composite is
Figure 4.7. Failure index distribution, CFRP 8 ply laminate, upper skin,
inertial load (0, -21, -80.9) m/s2
visible in figure: although the total load vector lies on the yz plane, the failure
index distribution is asymmetric with respect to the y axis, since the fibres of
the concerned ply are oriented at θ = 45◦ . Knowing the most loaded ply, an
optimization of the stacking sequence can be performed. The most loaded ply
is moved towards the centre of the symmetric laminate, that is towards the
right in the stacking sequence notation. With this change the distance from the
bending neutral axis is reduced by small amounts, multiples of the thickness of
the lamina. Employing the thicker honeycomb this optimization is not much
visible, and very little gain on the maximum failure index is obtained.
To have a better comprehension of the behaviour of the sandwich, a 10 ply
laminate is analysed. The best 8 ply configuration just found will probably
be the overall optimum choice. The principal attempted configurations for a
10 ply CFRP laminate are visible in Figure 4.8. A group of 2 layers can be
inserted symmetrically in the laminate, with orientations of 0◦ or 90◦ . The
±45◦ plies instead can be added only in groups of four to maintain the laminate
balanced and symmetric. Adding 0◦ plies is not useful, since those plies are
always the least stressed. Adding 90◦ plies good results are obtained, with the
use of the thinner and denser honeycomb. The dynamic requirement is hardly
satisfied. With a stacking sequence optimization, the maximum failure index
can be brought under the 0.5 threshold of the static criterion, but remains very
close to this limit. The logic for stacking optimization is always to put the
most loaded plies in at the centre of the laminate, lowering their distance from
50 4. Design of the panel
Figure 4.9. Failure index distribution, CFRP 10 ply laminate, upper skin,
inertial load (0, -21, -80.9) m/s2
the neutral axis. The optimum 10 ply configurations allow a small mass saving
of 0.14 kg with respect to the 8 ply situation. Every stacking sequence in
which the 45◦ layers are in the middle and the 0◦ plies are external is optimal.
The critical layer in the most severe load case is visible in Figure 4.9. This is
again the external 45◦ ply of the upper skin. The 10 ply laminate is a very
border-line solution for both the dynamic and static design requirements. To
complete the possible cases, some 12 ply configurations have been tested. The
reason is that with a 12 ply skin, 45◦ and -45◦ plies can be added maintaining
a balanced and symmetric laminate. Adding layers of the same orientation as
the critical one the stresses will be conveniently redistributed, reducing the
maximum failure index. The results are showed in 4.10. The first 3 rows of
the table show that adding 0◦ and 90◦ plies is not useful, as it was expected.
4.6 Comparative study of admissible configurations 51
The fourth row shows that employing the lighter honeycomb doesn’t lead to
a good configuration from the vibrational point of view. With the denser
honeycomb good configurations are found. Optimizing the stacking sequence
some satisfying modal and static parameters are obtained (MoS=0.24). Weight
naturally increases, but its still lower than that of the aluminium sandwich.
The stresses on the oblique plies are redistributed, so the most loaded plies
are now the 90◦ ones. Figure 4.11 shows the failure index distribution on the
critical ply. In this plot both loads and composite fibres are symmetrical with
Figure 4.11. Failure index distribution, CFRP 10 ply laminate, upper skin,
inertial load (0, -21, -80.9) m/s2
in Figures 4.12 and 4.13. The points lying in the top-left region of both
(270◦ ). With this load, the failure index reaches the highest value, highlighted
with a red box. The worst case happens when the lateral load points towards
the negative y direction because two brackets have been deleted on this side
of the panel to make room for the flywheel interfaces. Not only this, but all
the load events become more severe if lateral load points towards this region
(225◦ , 270◦ , 315◦ ). This is testified in the reported tables, where in the last 3
tables (225◦ , 270◦ , 315◦ ) each failure index is higher than the corresponding
value of the first 5 tables.
The magnitudes of forces and moment vectors have been collected at the bracket
and wheel interfaces. The components (x,y,z) are showed only for the highest
loads (red boxes in the tables). Forces and moments at the panel-bracket
interface have to be divided by the unknown number of bolt constituting the
joint while, since the rotating mechanism interfaces are known, the loads are
here referred to the single bolt. Forces and moments at the interfaces are
useful values for the dimensioning of the joints. Moments are much smaller
than forces. For this reasons they have not been considered in the comparative
tables. To how the values of the interface load components, the vector plots
showing the components of the maximum interface forces are given in Figures
4.20 and 4.21. The colors in figure green for the x component, yellow for the y
and red for the z. From Figure 4.20 the single node connecting the bottom of
the bracket with the plate is visible. Another node on the top of the bracket
connects it to the satellite. The force components have to be distributed over
a unknown number of bolts. Figure 4.21 shows the interface force components
on the four nodes of the panel-wheel interface. The most loaded node in figure
is the one on the left. The maximum values of the interface forces are resumed
The moment vectors are not reported since as their values are extremely small.
The values of the force components suggest that normal and shear forces on
the bolted joints allow to design the bolted joints with standard diameters and
materials.
Figure 4.22 shows the Von Mises stress distribution over the brackets, for the
optimum 8 ply configuration and the worst load case. The load case is the
same that causes the maximum forces exchanged between plate and brackets.
The maximum Von Mises stress value is of 74.9 MPa, much lower than the
limit of 250 MPa (MoS=2.3). This confirms that the brackets are not critical
in terms of quasi-static loads exerted during launch. The thicknesses of the
brackets are 3 mm for the vertical faces, 5mm for the upper base and 10 mm
for the lower base. The design of the brackets is not the aim of this work, but
from this analysis it is clear that an reduction of the thicknesses is possible, in
order to optimize the material usage.
4.7 Analysis on the optimum configuration 55
Chapter 5
where mi , ki , bi and pi (ω) are respectively the mass, stiffness, damping and
force associated to the ith mode. Solving the equation for ξi (ω), each individual
modal response can be computed. The values for the n analysed modes are
then summed up to find the physical solution x.
The damping has not been defined directly as a modal damping bi , but through
the quality factor Qi .
1 mi ωi
Qi = = (5.3)
2ζi bi
This quantity is defined from the more known damping ratio ζi , and it is
inversely proportional to the actual damping. Its distribution is defined in the
frequency range of interest for the modal extraction, which in our case is up to
800 Hz. In this range a constant value Q = 100 has been set.
The force and moment vectors have been defined through their x and y
components, assigning magnitude distributions and phases. The application
region is the point mass modelling the rotating mechanism. The the solution is
computed in the interval [0,30] Hz, every 0.1 Hz. Since the in-plane load vectors
rotate at the same angular speed, they are defined in the same frequency range.
The loads defined in this analysis have the form of Equation 5.4 [14].
Ud = 50 g · cm2 = 5 · 10−6 kg · m2
The imbalances are very small, and the range of wheel rotating frequencies is
not high too. For these reasons the frequency distributions of the centrifugal
force and moment modelling the disturbances have low values, as showed in
Figures 5.1 and 5.2. The functions are parabolic, and their maximum values
verify at 30 Hz and are 0.53 N and 0.18 Nm. The fatigue analysis is pursued
anyway, since for a service life of 12 years and a frequency of 30 Hz, the number
of stress cycles on the bracket is very high, in the order of 1010 cycles. This
high number of cycles can be dangerous for the brackets, even with small loads.
In addition, the real disturbances could be much higher than the typical values
employed in this analysis.
In defining the load phases, the quadrature of the moment with respect to
the force has been respected. The initial direction of the force is along the
positive x axis, and that of the moment is the positive y axis. Phase values
are θF x = θM y = 0 and θF y = θM x = π4 . A scale factor of -1 is set for Mx , since
in the first 180° of the rotation it must assume negative values. To avoid the
computation of the rigid body modes and their contribution to the solution,
the lowest frequency of interest for the mode extraction has been set to 0.1 Hz.
64 5. Fatigue verification of the brackets
the stress distribution on the critical brackets are equal, the top-left one in
Figure 5.3 is taken as a reference.
Near the bases of each bracket the 3 parallel vertical faces form as many angles.
In those regions the fatigue life of the component is reduced by the presence
of a fatigue stress concentration factor, since geometrical discontinuities are
present. For this reason the most loaded node on an angle point of the bracket
is the second point to verify in terms of fatigue. The node with the highest
Von Mises stress is called Node 1, while the corner node with highest stresses
is called Node 2. Node 1 and Node 2 are located in the same bracket. The
positions of the 2 critical regions are visible in Figure 5.4, where Von Mises
stress values are displayed. Note that the stress at Node 2 is nearly a half of
that at Node 1. The fact that the overall most loaded region of the bracket is
the lower base, is coherent with its thickness (10 mm) that is the highest. The
most external positions along the shell thickness, where the most tense and
compressed fibres lie, are called Z1 and Z2. Tables 5.1 and 5.2 resume the stress
components on Node 1 and Node 2 in these two positions. Only the non-zero
stress components are reported, in the global coordinate system. The stress
components are extremely small as expected, since load magnitudes are small.
On Node 1 normal stresses are slightly higher on the lower element face and
they change sign through the thickness, going from the most compressed fibres
to the most tense. This is coherent with the typical linear distribution of the
bending stresses along the thickness. There will be a neutral bending xy plane
in the shell elements close to their middle plane. All the stress components are
different from zero for Node 2, because the shell elements to which it belongs
are on a generically oriented plane. The values are nearly constant between
the lower and upper element face. Stresses on Node 2 are slightly higher at
the upper element face. A more detailed view of some stress components on
the critical bracket is given in Figures 5.5, 5.6, 5.8 and5.7, where some of the
previously tabled values on Node 1 and Node 2 are evidenced. The plots refer
only to the lower element face (Z1) and are viewed in the global coordinate
66 5. Fatigue verification of the brackets
Figure 5.4. Von Mises stress on the critical bracket, values on Node 1 (92.7
Pa) and Node 2 (50.9 Pa) highlighted
σx σy τxy
Lower (at Z1) -90.8 -32.8 24.3
Upper (at Z2) 84.5 29.8 -21.5
Table 5.1. Stress components [Pa] on lower and upper element face, Node 1
system. The global coordinate system coincides with the local element system
for the element to which Node 1 belongs, but this is not valid for Node 2.
This is testified that the shell element containing Node 1 has a null σz value.
The distributions of τxz and τyz are omitted. The labels on the right refer to
all brackets, and the figures are just a zoom on the examined bracket. The
vertical faces are almost unloaded, with exception of the regions where the free
5.2 Results of the frequency analysis 67
The mean values and amplitudes are listed in Table 5.3. The number of cycles
is equal to the number of wheel rotations. Considering an uninterrupted service
life of 12 years and 30 rotations per second N = 11353 · 106 cycles.
5.4 ESALOAD and ESAFATIG 71
the real load history of the system has been assumed, defining an event file which
contains a series of load generating events. Each event must be associated to a
load curve, which contains informations about the number of cycles versus mean
and alternating accelerations (but also pressure or temperature), expressed
as multiples of the gravitational acceleration. Load curves can be taken from
an existing database containing standard pre-calculated curves, which are
modifiable. ESALOAD also gives the possibility to define load curves, giving
informations about the load input at the base of the structure and about the
dynamics of the system through a transmissibility function. Putting together
the informations from all the events and the load curves associated to each
event, the software creates the load spectrum. This file contains load cycles
versus number of cycles for every defined load event. The software is then able
to calculate the stress spectrum, which is the same as the load spectrum but
with stresses in place of accelerations, knowing the unit stresses. Unit stresses
are the stress components induced by a unitary load (1g acceleration) in the
point of interest for the fatigue analysis. Knowing the stress components due
to a unitary load on a specific point, and the complete history of the load
cycles, ESALOAD gives in output the complete set of stress cycles (maximum
and minimum stress values) to which the structure is subject during its service
life. This output is the stress spectrum file.
The stress spectrum is the input for ESAFATIG, together with the S-N curve
of the material chosen from the software material database [20]. From stress
spectrum file, the minimum and maximum stresses Smin and Smax are derived
by combining the stress components into equivalent uniaxial stresses. The
stress ratio R is found dividing Smin by Smax . The S-N curves are represented
analytically by Equation 5.6.
S-N curve of the material. The final output of ESALOAD is total damage
accumulated Φ. To find the actual fatigue damage f , the value of Φ has to be
multiplied by a factor of safety, equal to 4. The design is said to be acceptable
if the fatigue damage is lower than 1.
f = 4Φ < 1 (5.9)
The ESAFATIG program flow chart is showed in Figure 5.12. The term PFCI
in figure stands for potential fracture critical item. The PFCI is a part of the
74 5. Fatigue verification of the brackets
system which for some reason is considered critical in terms of fatigue. In the
present analysis the PCFI is the most loaded bracket, in particular the areas
of Node1 and Node 2.
cycles due to the production and assembly phases are not considered. The load
curves related to the horizontal, vertical and tilting handling are taken from
an existing load curve database, together with the road and air transportation
curves. The curves are defined in ESALOAD for careful handling and transport.
The transport curves are available per km travelled. The handling is defined
between each phase of Figure 5.13. The assumed transport events are:
load input values, for a total of 6 qualification tests. These loads are applied
76 5. Fatigue verification of the brackets
modal parameters needed to define the transmissibility function are the natural
frequency ωk , the damping ratio ζk and the MEMF values. The first 4 modes
have been taken as representative of the whole system. For the damping ratio
ζk a small value 0.01 is set for the selected modes, to consider a very critical
situation in terms of amplification of the load at the junction.
Given a load input and a transmissibility function, the program calculates the
response ẍ of the system and an equivalent number of cycles. The different
load types require different mathematical treatments, being defined each with
its proper type of input. A detailed view on the calculation of response and
number of cycles for every type of load is given in [19]. For sine (harmonic)
loads, expressed as acceleration amplitudes q̈r specified at each frequency, the
response at the point of interest is found simply applying Equation 5.11.
ẍ = T (ω)q̈r (5.11)
For the bracket a different transfer function exists for each spatial direction,
while the defined sine test inputs are the same in every direction. The cal-
culation of the number of cycles changes with the load type as well. For a
sinusoidal load the frequency range is divided in small ranges of amplitude
fi − fi−1 in which the response ẍ is assumed constant. The cycles Ncyc are
calculated as
60(fi − fi−1 )
Ncyc = (5.12)
voct ln(2)
where voct is the sweep rate, measured in octaves per minute, that is the
derivative in time of the harmonic load frequency.
The random load environment refers for example to the acoustic pressure
generated inside the fairing during lift-off. Due to the probabilistic approach
needed to characterize the random excitation, the input must be given through
its power spectral density function Sr (ω). The PSD of the forcing signal q̈r is a
78 5. Fatigue verification of the brackets
real and positive function which represents the distribution in frequency of the
root of the mean square of the signal q̈rms . The PSD contains informations of
the energy (power) content of a signal in frequency. Knowing the PSD of the
input and the transmissibility T (ω), the PSD of the response in the direction
p is calculated as:
S p (ω) = T (ω)T ∗ (ω)Sr (ω) (5.13)
where T ∗ (ω) is the complex conjugate of T (ω). The knowledge of S p (ω) allows
to reconstruct the root mean square of the response in each direction ẍp . If
the mean of the signal is zero, the definition of root mean square (or standard
deviation) is given by Equation 5.14.
sZ
∞
ẍprms = S p (f )df (5.14)
0
Since any random process has the tendency to follow a Gaussian distribution,
taking a value of the maximum response equal to 3 times its root mean square
value allows to have a very good confidence that this maximum acceleration is
not exceeded. The number of cycles for random acoustic inputs is calculate
under the additional assumption that the excitation is stationary. The analo-
gous to the number of cycles for a random signal is the rate of upcrossings of a
certain acceleration level. The expected rate of upcrossings of an acceleration
level b is:
−b2
p
vbp = v0p e 2M0 (5.15)
where v0p is the number of zero crossing per second of an excitation in the
direction p. M0p is the variance of the motion in the same direction, when the
mean value is assumed to be null. The relations for the calculation of v0p and
the generic Mip are showed respectively in Equation 5.16 and 5.17.
v
u p
u M2
v0p = t (5.16)
M0p
sZ
∞
M0p = S p (f )df (5.17)
0
The last load curve has defined from the already existing curve related to
the ascent phase of the Ariane 5 launcher. It is based on low frequency
measurements at the launcher-satellite interface with many different payloads.
This load curve, together with the sine and random flight load curves already
defined, allows a description of the most important loads occurring during flight.
Two corrections are applied, scaling acceleration amplitudes with appropriate
coefficients. This is done under the realistic hypothesis that, changing the
structure to which ascent loads are applied, the only changing parameter are
acceleration amplitudes. The first correction is related to the fact that the
5.5 Definition of the load history 79
known from the result of the frequency analysis. The selected nodes are the
one with the highest peak stress value (Node 1) and the most stressed node
between those where the faces of the bracket form an angle (Node 2). The
unit stresses of Node 1 and Node 2 are obtained through a static analysis on
the bracket. The unit stress values are defined for each DoF, so an inertial
acceleration of magnitude 1g has been applied along each individual DoF.
The in-plane stress components have been extracted, and are tabled in Table
5.7. ESALOAD requests the stress components in the local coordinate system
of the element, therefore only the 3 in-plane stress components are different
from zero. The unit stresses have been associated to every load curve defined
in this section. The outputs of this association are the stress spectra of the
two potentially critical points. The stress spectrum contains each existing
stress cycle, defined by the maximum and minimum stress components and
80 5. Fatigue verification of the brackets
Node 1 Node 2
DoF σx σy τxy σx σy τxy
Trasl. X 7.8 2.5 2 3 4.3 0.6
Trasl. Y 1.7 1.0 0.1 2.1 2.9 0.4
Trasl. Z 0.6 0.5 0.7 1.9 2.6 0.3
Rot. X 4.5 2.2 0.3 3.7 5.2 0.7
Rot. Y 11.8 4.4 2.2 2.0 2.9 0.4
Rot. Z 1.5 0.7 0.2 0.8 1.0 0.2
Table 5.7. Unit stresses of the potentially critical nodes [MPa]
the associated number of cycles. The informations in Table 5.3 have been
expressed as maximum and minimum values, and added to the stress spectrum
file of the respective node. This complete stress spectrum is an input of the
ESAFATIG package, together with the S-N curve of the 7075 aluminium and
the value of the stress concentration factor. The final output is the linear
damage accumulated on Node 1 and Node 2.
For ductile material under dynamic loading a fatigue stress concentration factor
has to be calculated. It is surely smaller than Kt = 1, since plasticization in
5.6 Fatigue damage assessment 81
ductile materials makes less critical the stress concentration around a notch. It
will be conservatively assumed that Kt = Kf , since to know the exact Kf the
effective fillet radius is needed. Von Mises stresses due to the imbalances of
Figure 5.4 may suggest that the stress concentration on Node 2 results in an
equivalent stress around 3 times the reference value. This would be misleading,
since in the model does not include any fillet in correspondence of its corner
points. A fillet radius should be designed in order to lower the concentration
of stresses on Node 2, employing a specialistic text [17].
The ESAFATIG output files are generated. Their heading is showed in Figure
5.17. The file contains a series of rows representing the maximum, minimum
and equivalent stresses, filtered according to a minimum threshold. This
threshold is inversely proportional to the number of cycles. A section extracted
from the ESAFATIG output file, regarding Node 1, is showed in Figure 5.18.
For every row in figure, the program checks if the equivalent stress Seq is below
the yield stress Syld and the experimental stress Sexp . Sexp is the value used in
the experiments needed for the definition of the S-N curve. The last column
is the Miner linear damage accumulated in consequence of the corresponding
stress cycles. From line number 28 on the damages are the highest, in the order
of 10− 2. This is true for both Node 1 and Node 2. It has been verified that
the correspondent lines are due to the random qualification test and random
flight loads. The random events are the only causing wide band loads, in the
[20,2000] Hz range. As was visible in Figure 5.14, the first natural frequency
of the bracket is relatively high. The transmissibility functions Tp (ω) have
been set according to the modal parameters of the bracket, therefore they
5.6 Fatigue damage assessment 83
Conclusions
Bibliography
[2] J. Alcorn, C. Allard, and H. Schaub. Fully coupled reaction wheel static
and dynamic imbalance for spacecraft jitter modeling. Journal of Guidance,
Control, and Dynamics, 41(6):1380–1388, 2018.
[25] X.-S. Yi, S. Du, and L. Zhang. Composite materials engineering, volume
2. 2018.