Sei sulla pagina 1di 84

Aeroengine Fan Blade Design

Accounting for Bird Strike

Arthur Blair

A thesis submitted in partial fulfilment


of the requirements for the degree of

BACHELLOR OF APPLIED SCIENCE

Supervisor: S.A. Meguid

Department of Mechanical and Industrial Engineering


The University of Toronto

March 2008
Abstract

Bird strikes are a serious and growing problem, the risk of which must be

minimized by airports and aerospace engineers. Birds undergoing high speed

impact may be represented by a hydrodynamic material impinging on an ob-

ject, and the finite element method is a powerful tool for modelling this event.

In this thesis project, a finite element model of a bird striking a rigid target

is initially developed. Investigations are performed into various aspects of the

rigid target model, and it is found that a single hexahedral brick element is

sufficient to model a rigid plate. Investigations into the bird model show that a

hemispherically-ended cylinder made up of 13824 brick elements with a polyno-

mial equation of state and a density of 950 kg m−3 is a suitable model of a bird.

The model is validated using experimental data from impacts on rigid plates

of gelatin projectiles, which have been found to be suitable bird substitutes.

The model is found to fit the data well, particularly at lower impact velocities.

A generic, deformable fan blade model is then produced, and perpendicular

impacts near the blade tip of 1.8 kg birds are carried out for a range of blade

geometries. It is found that blades with larger chord widths suffer greater plas-

tic strain, while increasing a blade’s length decreses the maximum plastic strain

experienced in the impact area.


Acknowledgements

The author would like to thank Professor Shaker Meguid for supervising this thesis
project, and also Professor Pinhas Ben-Tzvi and Amy Franklin for providing useful
advice.

i
Contents

List of figures vi

List of tables vii

Nomenclature ix

1 Introduction 1
1.1 Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 The finite element method . . . . . . . . . . . . . . . . . . . . . . . . 2

2 Literature review 4
2.1 Bird strikes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.1.1 History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.1.2 Prevention . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.1.3 Regulations and testing . . . . . . . . . . . . . . . . . . . . . 11
2.2 Impact theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.2.1 Development . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.2.2 Description . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.3 Finite element modelling of bird strikes . . . . . . . . . . . . . . . . . 18
2.3.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.3.2 Formulations . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.3.3 Bird model . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

3 Finite element modelling 24


3.1 Approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

ii
3.1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.1.2 Software . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.1.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.2 Initial modelling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.2.1 Development . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.2.2 Evaluation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.2.3 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.3 Rigid target model investigation . . . . . . . . . . . . . . . . . . . . . 35
3.3.1 Geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.3.2 Material model . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.3.3 Material properties . . . . . . . . . . . . . . . . . . . . . . . . 40
3.3.4 Element selection . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.3.5 Discretization . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.3.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.4 Bird model investigation . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.4.1 Geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.4.2 Material model . . . . . . . . . . . . . . . . . . . . . . . . . . 45
3.4.3 Material properties . . . . . . . . . . . . . . . . . . . . . . . . 45
3.4.4 Element selection . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.4.5 Discretization . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.5 Simulation options . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
3.5.1 Velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
3.5.2 Contact definition . . . . . . . . . . . . . . . . . . . . . . . . . 53

4 Results and discussion 56

iii
4.1 Validation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
4.1.1 Experimental data . . . . . . . . . . . . . . . . . . . . . . . . 56
4.1.2 Comparison . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
4.2 Application to flexible fan blades . . . . . . . . . . . . . . . . . . . . 60
4.2.1 Fan blade model . . . . . . . . . . . . . . . . . . . . . . . . . 60
4.2.2 Geometry investigations . . . . . . . . . . . . . . . . . . . . . 61

5 Conclusions 64

References 66

iv
List of Figures

2.1 The number of bird strikes reported per year in the USA, 1990-2006 [1] 6
2.2 The proportion of bird strikes reported at different phases of flight . . 7
2.3 A passive bird deterrent measure at Auckland International Airport . 8
2.4 Simplified bird impact phases . . . . . . . . . . . . . . . . . . . . . . 16
3.1 Initial state and longitudinal stress contour plot of cylindrical alu-
minium impactor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.2 Hemispherically-ended bird model with key dimensions . . . . . . . . 27
3.3 Construction process of hemispherically-ended bird model . . . . . . . 28
3.4 Expected pressure output for a hydrodynamic impactor . . . . . . . . 29
3.5 Deformation over time of a bird striking a rigid target . . . . . . . . . 32
3.6 Pressure vs. time plot for initial bird model . . . . . . . . . . . . . . 32
3.7 Change in total, kinetic and hourglass energy over time . . . . . . . . 33
3.8 Change in solution timestep over time . . . . . . . . . . . . . . . . . 33
3.9 Four differently shaped models of a rigid target . . . . . . . . . . . . 36
3.10 Pressure vs. time plot for circular and square 3D rigid plates . . . . . 37
3.11 Pressure vs. time plot for circular and square 2D rigid plates . . . . . 37
3.12 Pressure vs. time plot for 3D square plates of differing thickness . . . 38
3.13 Deformation of a bird striking a very thin rigid plate . . . . . . . . . 39
3.14 Rigid target meshes of different densities . . . . . . . . . . . . . . . . 42
3.15 Pressure vs. time plot for rigid targets of different mesh densities . . 42
3.16 Two bird model shapes . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.17 Pressure vs. time plot for two different bird shapes . . . . . . . . . . 44
3.18 Pressure vs. time plot for different hydrodynamic polynomial coefficients 45

v
3.19 Pressure vs. time plot for different bird densities . . . . . . . . . . . . 46
3.20 Comparison of bird model deformation with and without merged nodes. 47
3.21 Four different bird mesh densities . . . . . . . . . . . . . . . . . . . . 49
3.22 Pressures vs. time for four levels of mesh refinement . . . . . . . . . . 50
3.23 Hugoniot pressure vs. number of elements used in bird mesh . . . . . 50
3.24 Hourglass energies for four levels of mesh refinement, as a proportion
of the largest . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
3.25 Pressure vs. time plot for three different bird velocities . . . . . . . . 52
3.26 Pressure vs. time plot for four different contact definitions . . . . . . 54
3.27 Hugoniot pressures for four different contact definitions . . . . . . . . 55
4.1 Hugoniot pressure vs. impact velocity from experimental data . . . . 57
4.2 Example of data extraction from printed graph . . . . . . . . . . . . 57
4.3 Comparison with experimental data for velocity of 117 m s−1 . . . . . 58
4.4 Comparison with experimental data for velocity of 158 m s−1 . . . . . 59
4.5 Comparison with experimental data for velocity of 171 m s−1 . . . . . 59
4.6 Basic fan blade geometry . . . . . . . . . . . . . . . . . . . . . . . . . 61
4.7 Deformation during collision with fan blade . . . . . . . . . . . . . . 62
4.8 Correlation between maximum chord width and peak plastic strain . 62
4.9 Correlation between blade height and peak plastic strain . . . . . . . 63

vi
List of Tables

3.1 Basic bird model definition . . . . . . . . . . . . . . . . . . . . . . . . 34


3.2 Basic rigid target model definition . . . . . . . . . . . . . . . . . . . . 34
3.3 Basic simulation options . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.4 Improved rigid target definition . . . . . . . . . . . . . . . . . . . . . 41
4.1 Basic fan blade model definition . . . . . . . . . . . . . . . . . . . . . 60

vii
Nomenclature

β Fluid compressibility

∆ρ Density change across shock wave

∆p Pressure change across shock wave

δ0 Kronecker delta function

γ Ratio of specific heats at constant pressure and volume

ρ
µ Mass density changing ratio µ = ρ0
−1

ρ Bird density

ρ Instantaneous bird density

ρ0 Initial bird density

ρ1 , ρ2 Fluid density either side of a shock wave

σ, σ 0 , σm Total, deviatoric and hydrostatic stress components

A Bird aspect ratio

a Speed of sound

Cn Linear polynomial equation of state coefficients

d0 Deviatoric strain rate tensor

E Young’s modulus

L Bird length

viii
Lc Length of cylindrical portion of hemispherically-ended bird

m Bird mass

P Pressure

p1 , p2 Fluid pressure either side of a shock wave

pH Hugoniot pressure

ps Stagnation pressure

pwh Water hammer pressure

R Bird radius

T Impact duration

u0 Bird impact velocity

us Shock wave speed

ix
1 Introduction

Every airborne traveller faces the dangers inherent in flying. The risks to man-made
aviation machines are well within acceptable limits, allowing flight to be among the
safest modes of transport. Events frequently transpire that cause intersection of the
dangers faced by birds and humans, although the consequences for humans are, fortu-
nately, most often much less severe than for their avian counterparts. But bird strikes,
as such events are known, are an increasing problem for an increasingly mobile species.
With populations of wild birds flourishing all over the world, particularly those wont
to gather in large flocks, like geese, and the growth of low-cost airlines colluding with
the expanding transport demands of emerging economies and general globalization
to push up global air traffic density, bird strikes are perhaps more significant an issue
now than ever before.
The mitigation of this threat may take many forms, and requires efforts from a
multitude of organizations and disciplines. Among the most crucial of these is the
design of all aircraft components likely to face exposure to bird impacts. Turbofan
engines are of particular interest, since they are the primary powerplant for the ma-
jority of large planes, and of all the recorded bird strike-related incidents involving
airliners or executive jets, 77% were caused directly by engine ingestion [2].
Finite element analysis (FEA) is an invaluable, computational tool in the design
process, which allows any variety of events to be simulated for minor cost outlays
when compared with performing a test on real, manufactured components. In this
thesis project, FEA is employed to simulate bird strikes on aeroengine fan blades,
with a view to understanding the ways in which blade design may be used to develop
engines that are more robust in their response to bird strike.

1
1.1 Objectives

• Develop a finite element model of a bird strike event that displays the correct
behaviour.

• Investigate every aspect of the model to understand fully the important param-
eters.

• Validate the model against available experimental data.

• Use the validated model to investigate the effect of fan blade design on the
response to bird strike.

1.2 The finite element method

The finite element method is a powerful technique for engineering applications as


it allows complex geometries and loading situations to be analysed computationally.
With modern computers the process can be carried out in high detail at relatively
low cost.
The method may be described briefly in the following way: Take a real-world
object that is subject to known loads, and represent it as a mesh of simple geometrical
elements, which are each defined by a set of nodes. Describe the known loads in
terms of discrete nodal forces, and apply a constitutive law to each element in turn
to calculate nodal displacements. Derive further results such as stress and strain for
each element, and recombine all the elements to produce a new representation of the
whole object.
A dynamic analysis, such as that which must be applied to bird strike, is essentially
a frequent repetition of these steps in order to analyze how the results change over

2
time.
In most practical applications, the implementation of the method described above
is handled entirely by a contained piece of commercial software, and so the bulk of
any investigation, such as the present one, is concerned with the other two parts of
any FE-based analysis: pre- and post-processing.
Pre-processing is the preparation of all input data that are required by the solver
program in order to produce a solution. For each object being modelled, the process
can be divided into several parts:

• Geometry - Create an idealized continuum model of the real-world object.

• Material model - Specify the constitutive law that governs the object’s material.

• Material properties - Define the material’s properties.

• Element selection - Select a suitable element type.

• Discretization - Discretize the continuous model into a mesh of finite elements.

These steps must be taken for every entity to be modelled, hence the headings in
sections 3.4 and 3.3 correspond to the above bullet points. Once all the entities are
discretized, boundary and intial conditions must be specified.
Post-processing covers the range of results and analyses that are produced and
performed once the solution has been completed. For the present investigation, some
of the key things to investigate will be the maximum pressures produced within the
bird upon impact, and the extent of plastic deformation in the fan blade.

3
2 Literature review

2.1 Bird strikes

Bird strikes are a serious and complicated issue. This section aims to give an overall
view of the problem and the measures that are taken to minimize its risk.

2.1.1 History

Bird strikes have occurred throughout the history of aviation, with their frequency
and consequence mirroring the industry’s growth and expansion. An early aviator
was Calbraith P. Rodgers, who trained with the Wright brothers and in 1911 became
the first person to fly across the United States [3]. The following year however, and
just nine years after the brothers’ own ground-breaking flight, he became the first bird
strike fatality. Rodgers was performing in an air show at Long Beach in California
when a seagull became entangled in the control wires of his Wright Flyer, which
subsequently crashed into the sea [4].
Birds began to pose a greater threat as aircraft grew faster, but the piston engines
ubiquitous in the 20th century’s first half were relatively robust and resistant to the
avian threat when compared with the gas turbine powerplants that would dominate
its second [5].
On 4th October 1960 three of the four turboprop engines on a Lockheed Electra
each ingested several starlings as it took off from Boston Logan International Airport.
Two engines lost power and one shut down entirely, causing the plane to stall and
crash into the sea. 62 of the 72 people on board lost their lives, making it the most
serious incident ever caused by bird strike [2]. Although no single accident would again

4
be as catastrophic, the total number of people killed in bird strike-related incidents
would continue to rise over the next 55 years to four times the fatality count in
Boston. Bird strikes became a more common occurance as the commercial airline
industry expanded and air traffic densities increased. In February 1973 a Learjet 24
crashed shortly after take-off from De Kalb, Georgia due to the ingestion into both
engines of a flock of cowbirds, resulting in the deaths of all 7 on board [2].
One factor that has increased the risk posed by bird strikes is the general trend
towards having fewer, more powerful engines per plane, rather than several smaller
ones. It is estimated that in 2008 only 10% of aircraft in the USA will have three
or four engines, as opposed to 75% 40 years ago [6]. This decreased redundancy has
made the ingestion response of individual engines more important, since the common
factor in the worst bird strike-related disasters is the loss of thrust in a significant
proportion of the available engines. On 15th September 1988 an Ethiopian Airlines
Boeing 737 hit a flock of speckled pigeons during take-off. The aircraft, which was
carrying 104 people, attempted to circle back to the airport, but both engines failed
before the circuit was completed. It crash-landed in the countryside, resulting in 35
fatalities [2].
Awareness of the risk of bird strikes has increased over time, and measures have
been taken to reduce the threat. It is evident however, that in recent times the risk
has only increased, and perhaps to a significant extent. The FAA has found that
almost four times as many bird strikes to civilian aircraft in the USA were reported
in 2006 than in 1990 (see figure 2.1) [1]. Increased awareness and better reporting
may account for some of the increase, although it is estimated that around 80% of
bird strikes still go unreported, so there are likely other, more serious causes.

5
8000

7000

6000
Number of reported bird strikes

5000

4000

3000

2000

1000

0
1990 1991 1992 1993 1994 1995 1996 1997 1998 1999 2000 2001 2002 2003 2004 2005 2006
Year

Figure 2.1: The number of bird strikes reported per year in the USA, 1990-2006 [1]

As shown by the trend towards fewer engines per plane, developments in engine
technology often have unintended negative consequenses in terms of vunerability to
bird strikes. A second example is engine noise, which is being steadily decreased with
each new design iteration in an effort to maximise efficiencies and minimize noise
pollution. It has been observed that quieter aircraft are more likely to suffer bird
strikes, simply due to the reduced warning given to birds in their path [7]. This
can also explain why more birds are hit during landing and approach phases than
during take-off (figure 2.2), since the high thrust needed to become airborne results
in more noise. A further example of unintended negative consequences is the growth
of low-cost airlines and their tendency (particularly in continental Europe) to use
smaller, more rural airports. While local bird populations may have grown used to
the infrequent flights of light aircraft, the rapid influx of twin-engined, jet-powered

6
planes can expose serious shortcomings in the minor airports’ abilities to deal with
the avian threat, and this is believed to increase the risk to passengers [8].
Developments that are perceived positively in other fields can also negatively
impact the bird strike problem, most notably the size of wild bird populations. Since
1980, the number of Canada Geese, a particularly threatening species from the point
of view of aircraft, have grown in Canada and the USA by 8% per year. Indeed, all but
one of the 14 common species with masses above 8 lb, the majority of which feature
noticably in bird strike statistics, are known to have enjoyed significant population
increases in recent decades [9]. Similar trends have been noticed in Europe, and
concern has been expressed over the suitability of current regulations given the rapid
increase in the number of “large” flocking birds [10].
The financial costs suffered by airlines due to bird strikes are very significant.
Determining accurate figures is difficult, but it has been conservatively estimated
that the total yearly cost to worldwide, commercial aviation is US$1.28 billion [11].
The International Bird Strike Committee believes that the total number of civilians
that have been killed in accidents directly caused by bird strikes is at least 242 [12].

Take-off Landing

Descent

Climb

En route Landing approach

Figure 2.2: The proportion of bird strikes reported at different phases of flight

7
2.1.2 Prevention

Prevention is a crucial part of minimizing the overall risk that bird strikes pose to
aircraft. Pilots rarely report visually identifying a bird or flock prior to a strike occur-
ring [13], and therefore are extremely unlikely to be able to take effective measures to
prevent them. Given this fact together with the observation that the overwhleming
majority of strikes take place during the takeoff and landing phases of flight [1] (see
figure 2.2), it is clear that airports have a significant responsibility to reduce the fre-
quency which with strikes occur. The Civil Aviation department of Transport Canada
has produced informative guides to the overall approach that airports should take to
wildlife [14, 15, 16]. A range of measures specific to bird prevention are possible [5],
and a range of the most common issues and techniques are summarized here.

Figure 2.3: A passive bird deterrent measure at Auckland International Airport

8
The “passive” approach is to remove any features of the airport or surrounding
area that attract birds, or to modify them in some way such that they are no longer
attractive or accessible. These features are generally categorised into those that pro-
vide a source of food, water or shelter.
Food sources are often man-made, such as the food waste produced by on-site
catering facilities. Large garbage disposal centres are sometimes found by the slip-
roads and intersections that lead to airports, and cause birds to fly across runways
in order to reach them. Other edible attractions are provided by the vegetation and
wildlife that is present both in the airport’s land and the surrounding area. Food
sources are difficult to control. The variety of bird species in any given region can
mean that attempts to deter a particular type result simply in the attraction of
another.
Water sources are more easily manipulated, in general it is desirable to minimize
the presence of open water at or near to an airport. Where sources can’t be removed,
birds can be discouraged from landing by employing covering wires, or steep banks
that would inhibit the detection of approaching predators.
Shelter is commonly provided by architectural features of airport buildings, which
are unlikely to have been designed with the minimization of potential bird nesting
sites in mind. A serious bird-management strategy can therefore include the retro-
active modification of certain features so as to prevent access to holes, ledges, vents
and crevices that would otherwise be attractive.
The “active” approach includes a large array of activities designed to disperse birds
or otherwise exclude or remove them from an airport. Pyrotechnics are a particularly
effective dispersal technique. Modified shotguns and starter pistols, or dedicated

9
launchers, can be used to project a variety of ammunition into the air, which then
emit loud bangs, screeches and flashes of light. Gas cannons are automated devices
that allow loud bangs to be produced in a particular direction at a set interval. Such
devices need to be moved around and their timings altered fairly regularly however,
as birds can quickly become used to the pattern of their output.
Birds use distress calls to warn other members of the species to disperse when they
have encountered some cause of serious danger. Playing the amplified recordings of
such calls is therefore an effective dispersal technique. The use of multiple sources
yields better results, and the calls must be specific to the particular birds in a given
species and region in order to be effective. Visual dispersive techniques include scare-
crows (see figure 2.3), radio-controlled model aircraft, and models or real carcasses of
gulls arranged in an “agony posture”. These can all have a limited effect depending
on the situation.
In cartain situations in which many of the above methods have not been effective,
and which involve non-endangered species that are thought unlikely to replace them-
selves quickly, the removal through lethal means of a section of a bird population
may be considered. A final, more widely applicable approach is to analyze the mi-
gration patterns of the most common local species and to try and adjust flight paths
and patterns accordingly. This method can prove very effective, since the migratory
movements of large flocking birds provide one of the biggest bird strike threats.
The cost of implementing a comprehensive bird management program can be of
the order of US$100, 000 per year [11], however this is typically significantly lower
than the cost to airlines of the bird strikes that, as a result of the program, did not
occur. Most commercial airports run such programs [17].

10
2.1.3 Regulations and testing

All major civial aviation authorities enforce strict regulations on the required safety
characteristics of new engine designs, which include specifications for bird ingestion
[18]. The only method by which these airworthiness certification requirements are
met is full-scale testing. Full-scale tests in this context mean attaching an engine to a
fixed rig and running it under various conditions, such as extremes of temperature or
impact situations. For bird ingestion requirements, bird carcasses are literally fired
into the running engine and various criteria must be met depending on the kind of
impact being simulated.
The requirements of the USA’s Federal Aviation Administration (FAA) are the
most commonly referred to in the literature, and they are broadly similar to other
large bodies’ recommendations (such as the european Joint Aviation Authority), so
they will be described here.
The FAA distinguishes between impacts of “large”, “medium” or “small” birds.
Medium and small birds are assumed likely to form flocks, so multiple impacts are
considered. An engine must be able to withstand such impacts without suffering a
sustained loss of power of more than 25%, and without significant deterioration of
its handling characteristics. Large birds are assumed not to be encountered in flocks,
so only single impacts are considered. An engine must be able to withstand such an
impact without hazardous fragments penetrating its casing, and without catching fire
or losing the ability to shut down.
A small bird is assumed to have a mass of 0.085 kg. A small bird flock impact must
involve one bird per 0.032 m2 of engine inlet area, up to a maximum of 16 birds. The
definition of a large or medium bird and the quantity involved in a flock encounter all

11
depend on the size of the engine involved. For an engine with an inlet area greater
than 3.90 m2 , a large bird has a mass of 3.65 kg. For the same engine, impacts of
three medium birds of mass 1.15 kg must be withstood. The largest engines (> 4.50
m2 ) must be able to ingest four such medium birds [18].
There has been criticism of the relation between these regulations and the real-
world situations that engines are likely to face, particularly with respect to the rapid
growth in populations of large flocking birds, which are thought to pose a threat
greater than that anticipated by the tests [9, 19, 20]. This has led to FAA to increase
the mass of the “large” bird to its current value of 3.65 kg, although the majority of
the literature still refers to the older, 1.85 kg bird size.

12
2.2 Impact theory

2.2.1 Development

The earliest researchers of bird strike assumed that the strengths of a bird’s flesh and
muscle fibres would be properties of significant consequence in the event of impact [21].
However, Barber et al. showed in 1975 [22] and in subsequent work [23, 24, 25] that
a bird could be more accurately thought of as behaving like a fluid when undergoing
high speed impact. This placed bird strikes firmly in the realm of soft-body impacts,
which are characterized by resultant stresses sufficiently in excess of the strength of
the impactor so as to allow the strength to be negelected. In short, it allowed birds to
be thought of as cylinders governed by a hydrodynamic material model, analagously
to the approach taken with other impacting liquids. Although bird strikes had not
been extensively studied before this time, a large body of work already existed on the
subject of water droplets and jets impinging on rigid surfaces. Indeed, in 1928 Cook
had observed that a cylinder of fluid colliding with a rigid object was, in the initial
phase of impact, equivalent to the well-known water hammer problem [26].
The water hammer problem concerns a flow in a cylindrical pipe that is suddenly
blocked, resulting in a pressure wave travelling back along the pipe [27]. Cook equated
the initial kinetic energy of the cylinder of fluid to the potential energy stored after
impact due its compressibility to find equation 2.1, which gives the peak pressure:

r
ρ0
pwh = u0 (2.1)
β

β is the compressibility of the fluid, ρ its density and u0 its impact velocity. Cook
further noted that the pressure due to this “first encounter” would usually be sig-

13
nificantly larger than that generated when the impacting fluid was flowing steadily
onto the surface of the target [26]. In 1961 Bowden and Brunton expressed the water
hammer pressure using the more general water hammer equation [28]:

pwh = ρu0 a (2.2)

Where a is the speed of sound in the hydrodynamic material. They explained that
the pressure in the fluid behind the shock wave – the shocked region – would be given
by equation 2.2 for a finite period of time, and that it would be the overall peak
pressure. The duration of this peak period is dependent on the release waves, which
are generated at the impacting end of the cylinder’s free, outer surface. The release
waves travel radially, towards the centre of the cylinder, and reduce the strength of
the initial shock wave as they go [29]. Upon converging at the central axis, the period
of peak pressure would end. After this period the fluid would flow radially, with the
pressure having the reduced magnitude of the “steady impact” phase identified by
Cook [26, 28].
The speed of any pressure wave through a material is given by equation 2.3 [30].

 
∆p ∆ρ
u2s = 1+ (2.3)
∆ρ ρ

Sound waves are a special case of the general pressure wave phenomenon for which
the pressure change is assumed to be infinitesimally small, resulting in the expression
for speed of sound shown in equation 2.4 [30]:

∂p
a2 = (2.4)
∂ρ

14
In a typical water hammer problem, it may be reasonably assumed that the shock
wave is equivalent to a sound wave, but a high energy impact event can easily produce
waves with supersonic velocities [30]. This led Bowden and Field to suggest in 1964
that the general shock wave speed be used in place of the speed of sound in the water
hammer equation [31]. The pressure and density changes across a non-isentropic
shock wave are governed by the Rankine-Hugoniot relations [32]:

γ+1 ρ1
p1 γ−1 ρ2
−1
= γ+1 (2.5)
p2 γ−1
− ρρ12

So Bowden and Field’s expression for the transient peak pressure experienced by a
cylinder of fluid impacting on a rigid target is now known as the Hugoniot pressure:

p H = ρ0 u 0 u s (2.6)

2.2.2 Description

In 1978 Wilbeck combined the above analytical work relating to water impacts with
evidence from the experimental studies of bird impacts, and produced the first com-
prehensive fluidic theory of bird strikes [29]. Since this work has been the basis of
most subsequent bird strike investigations [33, 34, 35], it is useful to describe the
physics of the impact event as understood by Wilbeck.
Figure 2.4 depicts the four phases of the impact of a cylindrical body of fluid on a
rigid target at an oblique angle. More complex geometries, such as hemispherically-
ended cylinders, complicate the theory but behave in the same broad manner. The

15
four phases can be described as follows:

(a) The cylinder of fluid approaches the flat target at an oblique angle and with a
purely axial velocity. Its internal pressure is equal to that of the ambient atmosphere.

(b) The leading face of the cylinder impacts the target and its constituent particles
are instantaneously stopped, creating a violent shock wave – the Hugoniot shock –
that travels back along the length of the cylinder as adjacent particles are brought to
rest. The shocked region behind the wave is subject to a transient Hugoniot pressure
of high magnitude.

(c) The huge pressure gradient between the radially unconfined free surface of the
shocked region and its interior causes the generation of release waves, which trigger
rapid radial expansion at the shocked end. The release waves travel at the speed of
sound back towards the centre of the cylinder, reducing the strength of the Hugoniot
shock wave as they progress.

(a) Before impact (b) Initial shock (c) Release waves (d) Steady flow

Figure 2.4: Simplified bird impact phases

16
(d) The transient period ends when the release waves meet at the cylinder’s central
axis [28]. Steady flow is established as the cylinder flows onto the surface of the target.
A steady stagnation pressure is reached, which is maximum at the centre of impact.
The impact forces decay completely once the fluid’s velocity is entirely radial.

Wilbeck analysed the one-dimensional flow across a shock wave in order to find
the Hugoniot pressure, and thus derived equation 2.6, as suggested by Bowden and
fields [29, 31].
The stagnation pressure experienced by the fluid during the steady flow stage that
follows the initial shock can be easily derived using Bernoulli’s equation, and is given
by equation 2.7:
1
ps = ρ0 u20 (2.7)
2

The total duration of the impact event can be estimated by assuming that the
trailing edge of the bird continues at the original impact velocity until all the fluid
has begun to flow radially. This results in equation 2.8:

L
T = (2.8)
u0

17
2.3 Finite element modelling of bird strikes

2.3.1 Overview

The finite element method was developed in the middle of the 20th century as an
adaptation of the Matrix Structural Analysis technique that was tailored for use with
programmable, as opposed to human, computers [36]. Finite element modelling of
bird strikes began in the late 70s and early 80s. These early simulations were largely
connected with military aircraft, and dealt with canopy and windshield impacts [37,
38]. At this time the computing power available to most institutions was minimal
when compared with the situation today, hence the inital predominance of military
involvement. Indeed, the researchers behind one of the earliest bird simulations to be
performed at a university found that, having calculated the degree of mesh refinement
necessary to produce results of an accurate order of magnitude, they did not have
sufficient computing power to run the simulation [39]. A limited amount of work
was done throughout the 80s, again mostly relating to transparency impacts. Early
simulations of damage to engine fan blades took place in 1982 [40] and 1988 [41], but it
was not until the late 90s and the turn of the millenium that computing power became
sufficiently inexpensive for highly complex simulations to be performed with relative
ease. The majority of investigations into fan blade damage have taken place in the
past five years. Three different formulations have been used frequently, as described
in section 2.3.2. The bulk of the research has centered around the modelling of the
bird (see section 2.3.3).

18
2.3.2 Formulations

An important observation made by Barber et al. in 1978 was that the loads exerted
on a non-rigid target by an impacting bird are inextricably linked to the physical
response of the target itself [24]. Three implementations of the finite element method
have been successfully employed to simulate this crucial coupling. These are the
Lagrangian [35, 42, 43], Abtrary Lagrangian-Eularian (ALE) [44, 45, 46] and Smooth
Particle Hydrodynamics (SPH) [34, 47, 48] formulations.
The three formulations differ mainly in their method of discretization [33]. The
Lagrangian formulation is based on Lagrangian mechanics, and was the earliest ap-
proach to be used in bird strike analysis. In Lagrangian analyses, the bird is divided
into elements and the element mesh is bounded to the material. Birds undergo ex-
treme deformation during a strike, and this can cause problems in Lagrangian meshes
as some elements may experience negative volumes. Solving dynamic transient non-
linear problems requires a time step, which is often a function of the aspect ratio of
the smallest element in the mesh [49]. This causes further problems during high dis-
tortion in lagrangian analysis because the time step may become unacceptably small.
These problems have been overcome by using an elimination procedure whereby ma-
terial is removed from any element in which the strain rate or time step go beyond
acceptable values [49, 35, 43].
The ALE formulation allows elements to contain more than one material (e.g.
water and air), and uses arbitrary reference coordinates that allow material to flow
through the mesh (Eularian behaviour) rather than the nodes moving with the mate-
rial (Lagrangian) [46]. It does not experience a time step problem, but high distortion
can still cause negative volume issues.

19
The SPH fomulation is a more recently developed Discrete Element Method
(DEM) that uses particles of mass rather than an element mesh to represent the
bird [50]. It was originally intended for use in hypervelocity impact problems in as-
trophysics [51]. SPH avoids problems caused by mesh distortion, but still requires
further development to become reliably stable and consistent [50].

2.3.3 Bird model

Finite element simulations of bird strikes have primarily modelled the bird as either
a flat- or hemispherically-ended cylinder. Flat cylinders were more common in early
simulations due to their simpler geometry, and the fact that with a more coarse mesh
the difference between the two geometries is likely to be less anyway [41, 52, 42].
Barber et al. had also recommended a cylinder as being a suitable representation
of a bird in 1978 [24]. The International Birdstrike Research Group developed a set
of standardized artificial bird designs to be used in real-world tests, based on the
biometrics of the thirty most commonly struck bird species, and concluded that flat-
ended cylinders, hemispherically-ended cylinders and ellipsoids may all be suitable
to represent a bird, depending on the situation and the species being modelled [53].
More recent numerical sumulations have tended to use a hemispherically-ended shape
[49, 34, 48, 35, 54, 43, 33, 55], although flat cylinders are still occasionally used [46],
particularly when bird slicing is investigated [56]. An investigation into bird model
geometry found that the highest pressures are produced with flat-ended cylinders,
followed by hemispherically-ended cylinders and ellipsoids [57]. The same investiga-
tion also concluded that the aspect ratio of a bird model – the ratio of its total length
to maximum diameter – does not have a significant effect on the results. An aspect

20
ratio of 2:1 [35, 34, 44] is most commonly used.
A material model for the bird is required so that the deformations and stresses it
experiences may be calculated. In general the stress on any object has two compo-
nents: one hydrostatic (normal) and one deviatoric (shear) [58]:

σ = σ 0 + σm (2.9)

σ is the overall stress, while σ 0 and σm are the deviatoric and hydrostatic components
respectively. Different relations are used to represent each component in material
models that are expected to experience large strains. The deviatoric component for
a bird material has been modelled in several ways. A viscous, Newtonian fluid model
has been used [52, 42] such that:
σ 0 = 2µd0 (2.10)

µ is the dynamic viscosity and d0 is the deviatoric part of the strain rate tensor. The
most commonly cited material model is elastic-plastic hydrodynamic [42, 49, 48],
where the deviatoric response is governed by an isotropic, hypoelastic theory [52].
As explained in section 2.2 however, a bird’s behvaiour during a high-speed col-
lision is little different to that of a fluid [29], so in practice most investigations have
taken a purely hydrodynamic approach and ignored all deviatoric components. With-
out deviatoric stresses, the overall stress tensor can be expressed as simply:

σ = σm = −P δ 0 (2.11)

where δ 0 is the Kronecker delta function, i.e. the overall stress tensor is a diagonal
matrix (therefore with no shear components) whose principal components are all

21
equal to the hydrostatic pressure, P . A range of models have been used to model the
hydrodynamic behaviour [49, 48], but by far the most commonly applied has been
the linear polynomial equation of state [52, 42, 49, 44, 34, 48, 54, 45, 46, 33, 55]:

P = C0 + C1 µ + C2 µ2 + C3 µ3 + C4 + C5 + C6 µ2 E

(2.12)

In this equation µ is the mass density changing ratio given by

ρ
µ= −1
ρ0

where ρ and ρ0 are the instantaneous and initial densities respetively. One of the
earliest studies to use this model utilized the following values of the coefficients, Cn
[52]:
C0 = C4 = C5 = C6 = 0

C1 = 2324 MPa
(2.13)
C2 = 5026 MPa

C3 = 13927 MPa

The same values have later been used [42, 57], although when Langrand et al. com-
pared several material models [49] they suggested setting C1 = 2250 MPa and all
other coefficients to zero (therefore making C1 essentially the bulk modulus [34]),
which has since become a popular approach [54].
The most important material property of the bird is density. Barber et al. con-
cluded that a density equivalent to gelatin with 10% porosity was a suitable repre-
sentation of bird tissue [24], and Wilbeck went on to specify a density of 950 kg m−3
[29], which has been commonly used in numerical simulations [48, 33, 45]. The In-

22
ternational Birdstrike Research Group’s effort to design a standard artificial bird
also investigated densities, and subsequently derived the following relationship as a
suitable correlation [53]:
ρ0 = −0.063 log m + 1.148 (2.14)

ρ0 (g cm−3 ) and m (g) are the initial density and mass of a theoretical bird respectively.
The most common bird mass used is 1.82 kg, for which the above relationship would
suggest a density of 943 kg m−3 . A more commonly used density is 938.5 kg m−3 .

23
3 Finite element modelling

3.1 Approach

3.1.1 Introduction

There are two events to be modelled: a bird striking a rigid target, and a bird striking
a flexible fan blade. The purpose of the rigid target test is to decouple the results
from the response of the target structure, and therefore to validate just the model
of the bird against available experimental data. The flexible fan blade test will use
the validated bird model to investigate how design of fan blades affects the observed
loads and deformations.
Based on information from the literature review, and much experimentation, a
basic model of each event is created in section 3.2 that broadly exhibits the expected
characteristics. The Lagrangian formulation is used throughout.
The following two sections, 3.3 and 3.4, look in detail at the two objects that must
be successfully modelled for validation purposes: a rigid target and a bird. The basic
models from section 3.2 are used to conduct various investigations into the effect that
each part of the modelling process has on the results. The headings within these
sections correspond to those parts of the modelling process identified in section 1.2
that must be performed for every object being modelled.
The results from these investigations are used to validate the bird model in sec-
tion 4.1. The model is then discussed in terms of application to fan blade design in
section 4.2

24
3.1.2 Software

The modelling of bird strike is a dynamic, transient, non-linear problem. Suitable


software must therefore be used. The present investigation utilizes the static analysis
program ANSYS in conjunction with the explicit dynamic module, LS-DYNA. All
pre- and post-processing is performed in ANSYS, which then generates a *.k keyword
input file used by LS-DYNA to generate a solution.

3.1.3 Results

A variety of results can be produced from any bird strike analysis. The most common
output is a plot of pressure vs. time, as it is the large pressure at the contact interface
between the bird and the fan blade that causes deformation.
Pressures have been measured in different ways in the literature. They can be
found by dividing the known force on an individual element by the average element
area, or by dividing the entire reaction force perpendicular to the plate by the average
area of the entire bird. When a purely hydrodynamic material model is used, however,
the direct “stresses” on a given element are equivalent to a measure of pressure in
a fluid. This is corroborated by the fact that in all hydrodynamic simulations the
direct stresses on an element are observed to be equal in each direction. The maximum
pressure is expected at the centre of impact, so the stresses on the element at the
centre of the leading face of the bird model are used to generate pressure plots.
It is useful to normalize results, so that different simulations can be easily com-
pared. Pressures are normalized by dividing by the theoretical stagnation pressure,
given in equation 2.7. Time values are normalized by dividing by the total impact
time.

25
3.2 Initial modelling

3.2.1 Development

The first step taken towards modelling a bird strike was to simulate a simple impact
event. A cylindrical impactor was used, with a linear, elastic, isotropic material model
and the material properties of aluminium. It was discretized using a coarse mesh of
solid brick elements. The target was constructed from single, 2D rigid shell element,
and the “automatic surface-to-surface” contact regime was defined between the two.

(a) Initial model (b) Stress contour during impact

Figure 3.1: Initial state and longitudinal stress contour plot of cylindrical aluminium
impactor

Figures 3.1a and 3.1b show the initial state of an early model and the contours
of longitudinal direct stress produced within the projectile upon impact respectively.
These early simulations showed that an impact event could be successfully modelled
with ANSYS/LS-DYNA. The next step was to create a more bird-like impactor.
It was decided that a hemispherically-ended cylinder should be used to model the

26
bird initially, as this is more commonly used in the literature. See section 3.4.1 for a
comparison of bird shapes.

LC

Figure 3.2: Hemispherically-ended bird model with key dimensions

The dimensions of the bird model are not set to a single set of values, since they are
dependent on its other properties, which may be varied between simulations. It was
therefore deemed useful to find two equations based on the geometry of the selected
bird shape that would provide the two dimensions necessary to define it, namely its
radius and the length of its purely cylindrical portion. Equations 3.1 and 3.2 were
subsequently derived, which allow correctly sized geometries to be generated for any
combination of mass, density and aspect ratio. Figure 3.2 shows an outline of the
bird model with the relevent dimensions labelled.

m
r
R= 3 4  (3.1)
πρ 3
+ 2(A − 1)

Lc = 2r(A − 1) (3.2)

27
Difficulties were encountered meshing the hemispherically-ended bird model in
ANSYS. Any spherical or hemispherical (or similar) volume that had been generated
directly would not allow itself to be meshed with brick elements due to “incorrect
topology”. This issue was solved by revolving the volumes from areas, rather than
creating the volumes directly, as shown in figure 3.3.

(a) Base area (b) Revolved volume (c) Discretized mesh

Figure 3.3: Construction process of hemispherically-ended bird model

The main criterion for bird-like impact behaviour is conformation to the hydro-
dynamic impact theory as described in section 2.2.2. The early simulations showed
that solid aluminium impactors experienced a single peak stress value, which quickly
fell to negligible levels as the impactor rebounded off the target. A material behaving
hydrodynamically should not rebound, but rather flow continously onto the target
following the initial impact event (see figure 2.4).
Section 2.2.2 identified that the temporal impact pressure profile of a hydrodynam-
ically modelled bird exhibits two clear phases: a peak, transient Hugoniot pressure
followed by a lower, steady stagnation pressure. Only once all the fluid has begun to
travel radially do the pressures die away. Figure 3.4 shows the expected form of a
pressure vs. time plot.

28
Figure 3.4: Expected pressure output for a hydrodynamic impactor

The polynomial equation of state given previously in equation 2.12 was selected,
and the assumption was made that deviatoric stresses could be ignored. The poly-
nomial coefficients for the hydrodynamic model suggested by Langrand et al. were
used. These coefficients are investigated further in section 3.4.2.
In order to utilize the polynomial equation of state with no deviatoric stresses in
ANSYS, the Null material model was used. In addition to the polynomial coefficients,
it requires three quantities to be given: pressure cutoff, relative volume for erosion
in tension and relative volume for erosion in compression. These inputs do not affect
the results directly, but they must be set to suitable values in order to prevent the
simulation from failing. After some experimentation, it was found that the pressure
cutoff must be set to any negative value of large magnitude, and the relative vol-
umes for erosion in tension and compression should be of the order of 10 and 0.001
respectively.
Wilbeck’s suggested density of 950 kg m−3 [29] was used. The Null material
model also requires a Young’s modulus of elasticity and Poisson’s ratio to be defined,

29
although these do not greatly affect the outcome and need only be set to sensible
orders of magnitude. Density is investigated in further in section 3.4.3.

3.2.2 Evaluation

A simulation of a bird hitting a rigid plate was carried out using the basic model.
Figure 3.5 shows the resulting deformation of the bird over time, which corresponds
very well with the behaviour that is expected based on the findings in section 2.2, as
well as the observations of real impacts made in the literature [24]. The leading face
of the bird deforms immediately upon impact (3.5a), and begins to expand radially
(3.5b). The rest of the bird then continues flowing onto the plate (3.5c) until almost
the entire bird is expanding radially (3.5d), at which point the maximum radius of
the bird has been increasd by a factor of 5.
The non-dimensionalized pressure, as measured at the bird’s central, leading node,
is shown over time in figure 3.6. This resembles very clearly the desired pressure profile
described previously in figure 3.4. There is a very brief initial peak, followed by a
more steady period during which the normalized pressure tends towards a value of 1,
i.e. the absolute pressure is nearing the theoretical stagnation pressure.
ANSYS SOLID164 elements use reduced, one-point integration, which helps to
reduce computation time considerably, but makes them vunerable to zero-energy or
“hourglass” modes. These modes can cause a simulation to be invalid if the hour-
glassing energy becomes too high. Figure 3.7 shows the increase in hourglass energy
during the simulation. The values shown are normalized by dividing by the total
energy. Clearly the hourglass energy increases steadily throughout the simulation,
ending at around 20% of the total. It is recommended that hourglass energy be kept

30
below 10% of the total energy, so this is unaccaeptably high. The hourglass energy
does not rise above 10% until half way through the simulation however, when the
peak pressure has occured and the stagnation value has already been reached, i.e.
the main period of interest is over. It should not therefore preclude this model from
being used in further investigations, but for validation purposes the mesh should be
refined, as this is the primary method of reducing hourglassing.
Figure 3.8 shows the timestep used during the simulation. The timestep is based
on the size of the smallest element in the mesh, hence it decreases significantly as the
bird deforms. The reduction of the timestep to unmanagably small values has caused
problems in past investigations in the literature, however with modern computers
timesteps of the order of even 10−10 can be accomodated, and timestep was not an
issue in this simulation.

3.2.3 Summary

Tables 3.1 and 3.2 summarize the definitions of the bird and the rigid target used in
this initial model, while table 3.3 summarizes the simulation options used.

31
(a) t = 0.00 (b) t = 0.33 (c) t = 0.66 (d) t = 1.00

Figure 3.5: Deformation over time of a bird striking a rigid target

7
Non-dimensionalized pressure

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Non-dimensionalized time

Figure 3.6: Pressure vs. time plot for initial bird model

32
1.2

0.8
Non-dimensionalized energy

0.6 Total energy


Kinetic energy
Hourglass energy
0.4

0.2

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Non-dimensionalized time

Figure 3.7: Change in total, kinetic and hourglass energy over time

2.00E-06

1.80E-06

1.60E-06

1.40E-06

1.20E-06
Timestep

1.00E-06

8.00E-07

6.00E-07

4.00E-07

2.00E-07

0.00E+00
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Non-dimensionalized time

Figure 3.8: Change in solution timestep over time

33
Parameter Description
Geometry Hemispherically-ended cylinder
m = 1 kg
LC = 0.0938 m, R = 0.0469 m
Material model Hydrodynamic with polynomial EoS
C1 = 2250 MPa, C0,2,3,4,5,6 = 0
Pressure cutoff = −1 × 10−12 Pa
Relative volume for erosion in tension = 10
Relative volume for erosion in compression = 0.001
Material properties ρ0 = 950 kg m−3
E = 1 GPa, ν = 0.3
Element selection SOLID164
Discretization 3968 elements, average size = 0.00938 m

Table 3.1: Basic bird model definition

Parameter Description
Geometry Circular 3D plate
Diameter = 0.4690 m
Thickness = 0.0469 m
Material model Rigid
All nodes fully constrained
Material properties ρ = 1 × 105 kg m−3
E = 100 GPa
ν = 0.3
Element selection SHELL163
Discretization 1536 elements, average size = 0.0234 m

Table 3.2: Basic rigid target model definition

Parameter Description
Initial velocity 100 m s−1
Contact definition Automatic Surface to Surface
Time step multiple 0.9
Simulation time 0.0015 s

Table 3.3: Basic simulation options

34
3.3 Rigid target model investigation

The purpose of the rigid target impact test is to validate only the model of the bird, so
the model of the rigid target itself should not be a significant factor. However, it was
observed that the pressures generated do vary to a certain extent, depending on the
exact approach taken to model the rigid plate. The following investigations therefore
aim to give an understanding of the significance of each aspect of the rigid model,
with a view to producing a rigid target that allows the bird model to be reliably and
accurately validated.

3.3.1 Geometry

There are two broad methods of modelling the geometry of a flat, rigid target, assum-
ing that shell elements will be used (see section 3.3.4). Either a three-dimensional
volume can be created, or a two-dimensional area can be used in conjunction with a
specified perpendicular thickness value. The differences between 3D and 2D shapes
will first be examined, followed by an investigation into the effect of thickness.

Shape

The oblique impact of a homogenous projectile with a circular cross-section on a flat


target should be an axisymmetric event, so it is natural to give the target a round
shape. Given that the target’s constituent elements will have quadrilateral faces
however, it would seem sensible to use a square shape, so as to minimize element
distortion. The four configurations shown in figure 3.9 were therefore investigated,
consisting of three- and two-dimensional versions of a round and a square plate.
Impact simulations were carried out with the standard bird model described in

35
(a) 3D circular plate (b) 3D square plate (c) 2D circular plate (d) 2D square plate

Figure 3.9: Four differently shaped models of a rigid target

section 3.2, resulting in the nondimentionalized pressures over time shown in figures
3.10 and 3.11.
The results do not show significant variation, as expected. The Hugoniot pressures
produced using both 3D shapes are identical, where as in the 2D case the square plate
experienced a slightly higher peak pressure. The steady stagnation period showed
more variation. The pressures exerted against the 3D circular plate dropped away
before the end of the impact event, while the square plate produced a more constant
and extended stagnation measurement, in addition to having less variation in pres-
sure immediately after the transient peak. In general the 2D shapes produced more
variable readings, with neither case giving a particularly steady stagnation period. As
in the 3D case, the circular plate saw a larger secondary peak in pressure immediately
after the Hugoniot peak.
Further simulations were carried out using a higher mesh density in the bird model,
and broadly similar trends to those described were observed. It can therefore be
concluded from these results that, while the outputs corresponding to each shape do

36
9

7
Non-dimensionalized pressure

5
3D circular plate
4 3D square plate

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Non-dimensionalized time

Figure 3.10: Pressure vs. time plot for circular and square 3D rigid plates

7
Non-dimensionalized pressure

5
2D circular plate
4 2D square plate

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Non-dimensionalized time

Figure 3.11: Pressure vs. time plot for circular and square 2D rigid plates

37
not differ greatly, a three-dimensional plate with a square shape is the most desirable
configuration to be used in a rigid target impact test, due to the steadiness of the
impact pressure it experiences after the initial peak.

Thickness

The second geometrical aspect of interest is the rigid plate’s thickness. Following
on from the recommendation of the investigation into the shape of the rigid plate,
simulations were carried out using a three-dimensional square plate (as shown in
figure 3.9b) with three different levels of thickness. Thicknesses were defined relative
to the average element size in the plate. The “medium” thickness level was the same as
that used in the basic model – twice the average element width. “Large” and “small”
1
thicknesses corresponded to 5 and 3
times the average element size respectively.

7
Non-dimensionalized pressure

5
Large thickness
4 Medium thickness
Small thickness
3

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Non-dimensionalized time

Figure 3.12: Pressure vs. time plot for 3D square plates of differing thickness

38
The temporal pressure profiles that were produced for the three levels of thickess
are shown in figure 3.12. The plots for the large and medium thicknesses overlap
exactly, but the pressures experienced by the thinnest plate show significant distur-
bances. In this third simulation it was noticed that significant penetration of the
bird through the plate had occured, with the leading, central elements of the bird
becoming extremely distorted. Figure 3.13 shows a plot of the deformation of the
bird near to the end of the impact period.
This penetration event did not occur with the larger thickness values, and is clearly
the cause of the disturbance in the pressure plot for the thin plate. From this it may
be inferred that thicknesses below the order of one element width can hamper the
accuracy of a simulation, and that they should be kept larger than this amount, in
which case they should have no effect on the outcome.

Figure 3.13: Deformation of a bird striking a very thin rigid plate

3.3.2 Material model

In ANSYS/LS-DYNA an object is made rigid by applying the Rigid material model.


No inputs are required other than basic material properties. All nodal degrees of
freedom can be fully constrained, so no constitutive law is necessary.

39
3.3.3 Material properties

The density, Young’s modulus of elasticity and Poisson’s ratio of the rigid material
must be specified, but they are only required to calculate the stiffness of the contact
interface between the target and the bird. They have no effect on the results when
set to realistic values.

3.3.4 Element selection

The ANSYS shell element SHELL163 has most commonly been used to model a rigid
target. However, experimentation showed that identical results are produced with
SOLID164 elements. This is to be expected, since all the nodes of every element are
fully constrained. The main practical difference between the two is that shell elements
can be used to mesh a simple two-dimensional area, where as solid brick elements such
as SOLID164 must always have a three-dimensional form. Overall, SOLID164 is the
better choice, since in section 3.3.1 it was shown that a 3D rigid target is probably
more desirable anyway, and furthermore, the SOLID164 element is also used to mesh
the bird, so using it for the target means that only one element definition is required
for the whole simulation.

3.3.5 Discretization

The rigid target may be discretized using a range of mesh densities, from a very fine
mesh all the way down to a single element. Simulations were carried out with four
different levels of mesh refinement, as depicted in figure 3.14.
The pressure profiles that resulted from the simulations, shown in figure 3.15,
again exhibit variation only in the stagnation period, while the Hugoniot pressures

40
are indistinguishable. The three least-fine densities show the least variation, where
as the finest mesh displays a more significant unsteadiness.
These results indicate that a single element can be used with no less accuracy
than a medium mesh. It is unclear based on this sample alone however, whether the
results given by the fine mesh are more desirable. As part of the validation process,
a single element rigid target should be compared with one that has a fine mesh in
order to choose between them.

3.3.6 Summary

Based on the investigations performed in this section, an updated definition of the


rigid plate is summarized in table 3.4. This model definition was used for the bird
model investigations in the next section (3.4).

Parameter Description
Geometry Square 3D plate
Width = 0.4690 m
Thickness = 0.0156 m
Material model Rigid
All nodes fully constrained
Material properties ρ = 1 × 105 kg m−3
E = 100 GPa
ν = 0.3
Element selection SOLID164
Discretization Single element

Table 3.4: Improved rigid target definition

41
(a) Fine (b) Medium (c) Coarse (d) Single element

Figure 3.14: Rigid target meshes of different densities

7
Non-dimensionalized pressure

5
Fine mesh
4 Medium mesh
Coarse mesh
3 Single element

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Non-dimensionalized time

Figure 3.15: Pressure vs. time plot for rigid targets of different mesh densities

42
3.4 Bird model investigation

3.4.1 Geometry

The geometry of a bird model for impact simulation purposes does not need to be at
all complex, or indeed to obviously resemble a bird. This is because the effect of a
real bird colliding with a rigid plate or compliant fan blade has long been recognized
as not significantly different from that of a homogenous lump of fluid. Variation is
still possible among the simplified geometries available however, and these will be
investigated.
The two shapes most often used to represent a bird are cylinders with either hemi-
spherical or flat ends. While it has previously been concluded that hemispherically-
ended cylinders are the more accurate of the two [57], other investigations have suc-
cesfully used flat-ended ones [50]. From the earlier review of literature it is clear
that different conclusions about bird geometry may be drawn depending on the exact
modelling approach taken, so it is of use to investigate the two cases using the model
developed in the present project.
Simulations were performed using the updated rigid target model defined in sec-
tion 3.3.6, with the two shapes shown in figure 3.16.
The resulting pressure profiles over time, shown in figure 3.17, differ greatly. The
flat-ended cylindrical bird experienced a Hugoniot pressure nearly twice as high as did
the hemispherically-ended model, while its stagnation pressure immediately dropped
to negligible levels until a second peak near the end of the impact period. These
results very clearly suggest that a cylinder with hemispherical ends is more suitable
to be used with the model that has been developed.

43
(a) Round-ended cylinder (b) Flat-ended cylinder

Figure 3.16: Two bird model shapes

16

14

12
Non-dimensionalized pressure

10

8 Round-ended cylinder
Flat-ended cylinder
6

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Non-dimensionalized time

Figure 3.17: Pressure vs. time plot for two different bird shapes

44
3.4.2 Material model

Tests were carried out using the polynomial coefficients suggested by Langrand et al.
[49] and Brockman [52]. The main difference between them is that Langrand et al. set
all but C1 to zero. As figure 3.18 shows, this simplification is entirely justified, since
the change in Hugniot and stagnation pressure due to specifying the extra values is
negligible.

7
Non-dimensionalized pressure

5
Langrand et al.
4 Brockman et al.

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Non-dimensionalized time

Figure 3.18: Pressure vs. time plot for different hydrodynamic polynomial coefficients

3.4.3 Material properties

The hydrodynamic impact theory suggests that the peak pressure in an impacting
fluid is directly proportional to its density (equation 2.6). Simulations were carried
out with bird models of four different densities, with the resulting pressures shown in

45
figure 3.19.

40

35

30

25 950 kg/m3
Pressure (MPa)

938 kg/m3
920 kg/m3
20
800 kg/m3

15

10

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Non-dimensionalized time

Figure 3.19: Pressure vs. time plot for different bird densities

3.4.4 Element selection

Any bird mesh impacting on a target will clearly experience nodal displacements
in all three cartesian directions, so a three-dimensional brick element is suitable.
The ANSYS/LS-DYNA elements SOLID168 and SOLID164 are both brick elements,
however SOLID168 uses a tetrahedral shape that is more suited to complicated, non-
uniform geometries. SOLID164 has a hexahedral shape ideal for simpler geometries,
and is generally the more accurate of the two [59]. SOLID164 was therefore selected
to mesh the bird, and it was used with the default 1-point integration setting, which
saves CPU time and copes better with large distortions.

46
3.4.5 Discretization

Node merging

Since the bird model is made from several separate volumes, the nodes associated with
each volume have to be merged so that the discretized bird acts as one object. It was
observed that when nodes are not merged, the bird instead behaves as two separate
hemispheres and a cylinder. Figure 3.20 shows a split image of the deformation of
a bird model with and without merged nodes. Note that the in the left half of the
image the cylindrical portion of the bird is still completely straight even though the
front hemisphere has deformed considerably. In the right half, with nodes properly
merged, the cylinder deforms with the hemisphere, thus more correctly modelling the
impact of a soft-body hemispherically-ended projectile.

(a) With merged nodes (b) Without merged nodes

Figure 3.20: Comparison of bird model deformation with and without merged nodes.

47
Mesh refinement

The number of elements used in the bird model can have a large effect on the results, as
well as a very significant effect on the computation time required to run the simulation.
In order to decide what level of mesh refinement is suitable in this project, rigid target
impact simulations were performed using the four levels of mesh density shown in
figure 3.21.
The resulting pressure plots over time are shown in figure 3.22. The stagnation
period of each curve does not differ drastically, although the timing of the secondary
peak after the Hugoniot pressure is different in each case. The Hugoniot pressure itself
does vary. Figure 3.23 shows how the maximum pressure varied with the number of
elements used. There is an appreciable difference between the Hugoniot pressures
recorded using coarse and fine meshes. However, the change in moving from a fine to
a very fine mesh is not as significant. Taking into account the fact that the very fine
mesh involves more than three times as many elements as the fine one, and that the
computation time for the finest mesh was several times the amount required for all the
other levels of mesh refinement combined, it can be concluded that the fine mesh, with
13824 elements, should provides a reasonable level of accuracy in further simulations.
The results also show that even the most coarse mesh, for which simulations take
very little time to complete, can be used for gauging the general level of pressure to
be expected from a given scenario.
In section 3.2 it was noted that the level of hourglass energy in the basic model was
too high. Mesh refinement is one method of reducing hourglass energy. Figure 3.24
shows the maximum hourglass energy for the first three meshes as a proportion of the
maxmimum hourglass energy produced in the basic model (equivalent to the coarse

48
(a) Coarse mesh (3968 elements) (b) Medium mesh (7900 elements)

(c) Fine mesh (13824 elements) (d) Very fine mesh (45684 elements)

Figure 3.21: Four different bird mesh densities

mesh). It is clear that refining the mesh helps to reduce hourglass energy, although
the effect is not very strong, so any saving may outweighed by the disadvantage of
the extra computing time.

49
10

7
Non-dimensionalized pressure

5 3968 elements
7900 elements
4 13824 elements
45684 elements
3

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Non-dimensionalized time

Figure 3.22: Pressures vs. time for four levels of mesh refinement

50

45

40

35
Hugoniot pressure (MPa)

30

25

20

15

10

0
0 5000 10000 15000 20000 25000 30000 35000 40000 45000 50000

Number of elements

Figure 3.23: Hugoniot pressure vs. number of elements used in bird mesh

50
100%
Non-dimensionalized maximum hourglass energy

80%

60%

40%

20%

0%
2000 4000 6000 8000 10000 12000 14000 16000

Number of elements

Figure 3.24: Hourglass energies for four levels of mesh refinement, as a proportion of
the largest

51
3.5 Simulation options

The investigations within this section look at the parameters that must be defined
for every simulation, but which are not related to a particular object being modelled.

3.5.1 Velocity

The hydrodynamic impact theory suggests that velocity is a very important determi-
nant of the magnitude of the Hugoniot pressure. To confirm this, simulations were
carried out at three velocities, resulting in the pressure profiles shown in figure 3.25.
Note that the pressures were not normalized, since the theoretical stagnation pressure
is also dependent on velocity, so dividing by it would to some extent reduce the effect
of the velocity being changed.

100

90

80

70

60
Pressure (MPa)

50 100 m/s
150 m/s
40 250 m/s

30

20

10

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Non-dimensionalized time

Figure 3.25: Pressure vs. time plot for three different bird velocities

52
The results show the expected behaviour in that the both the Hugoniot and stag-
nation pressures increases significantly with velocity.

3.5.2 Contact definition

Defining a suitable contact regime is an important part of any dynamic impact anal-
ysis, since it is this definition that allows objects to interact within the model. LS-
DYNA provides a very wide range of contact definitions to choose from, although
the majority of these are designed for a small set of specific situations. In general,
four contact types are recommended for most simulations [59], and these are analysed
here.
The first two types are Automatic Node-to-Surface (ANTS) and Automatic Surface-
to-Surface (ASTS), which both require Contact and Target entities to be specified.
They work by detecting penetration of the nodes or surfaces respectively of the Con-
tact entity through a surface of the Target entity. The second two are the Automatic
(ASSC) and General (AG) definitions of the Single Surface contact type. For both
of these definitions, ANSYS/LS-DYNA automatically determines which are the con-
tacting surfaces, and no further inputs are required. The difference between the two
is that the General definition requires the orientation of shell elements to be specified,
where as the Automatic type determines this itself.
Simulations were performed with the four contact types, in addition to another two
types (General Node-to-Surface and General Surface-to-Surface), which resulted in er-
rors. The four working types produced the pressure profiles shown in figure 3.26. The
Automatic Surface-to-Surface and Automatic Single Surface types produced identical
profiles, but the profiles and Hugoniot pressures varied among the rest. The Gen-

53
eral Single Surface type produced the highest Hugoniot pressure (see figure 3.27), but
the instantaneous pressure fluctuated considerably, suggesting the simulation was less
stable.
These results suggest that either the Automatic Surface-to-Surface or Automatic
Node-to-Surface contact definitions could be suitable. Both result in the expected
pressure profile without significant fluctuations. ANTS is commonly used for contacts
between a small and a large surface [59], which might suggest that its low Hugoniot
pressure is the most accurate. However, both types should be investigated during the
validation stage (section 4.1).

12

10
Non-dimensionalized pressure

6 Auto. surface-to-surface
Auto. node-to-surface
Automatic general
4 Automatic single surface

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Non-dimensionalized time

Figure 3.26: Pressure vs. time plot for four different contact definitions

54
50

45

40

35
Hugoniot pressure (MPa)

30

25

20

15

10

0
Auto. node-to-surface Automatic single surface Auto. surface-to-surface Automatic general

Contact definition

Figure 3.27: Hugoniot pressures for four different contact definitions

55
4 Results and discussion

4.1 Validation

4.1.1 Experimental data

Two sets of experimental data have been used. Barber et al. provided data for
over one hundred rigid target impact tests carried out using real birds with masses
between 50 g and 130 g [22]. Some pressure transducer temporal outputs were also
given, but poor reproduction of the original report means that these can not be
read accurately. The tables of data that provide the bird mass, impact velocity and
corresponding Hugoniot pressure for each test are still accessible, however. These
have been extracted and processed to produce figure 4.1. The full data include tests
where pressures were measured at various radial locations on the target plate, but
figure 4.1 shows only the Hugoniot pressures that were measured at the plate’s centre,
i.e. the stagnation point. These show that the hydrodynamic theory’s prediction that
the Hugoniot pressure should increase with the square of velocity is broadly correct,
although typically a theoretical Hugoniot pressure is much higher than that measured
experimentally.
More detailed pressure profile data is available from Wilbeck [29]. Wilbeck inves-
tigated several different artificial projectile materials in order to find a suitable bird
substitute, with the conclusion that porous gelatin was the most ideal. An example of
one of the pressure profiles produced using a gelatin impactor is shown in figure 4.2a.
Profiles were provided for several different velocities. It was necessary to extract the
data so that a meaningful comparison could be made with the simulated results. A
simple Matlab program was created that allowed any raster image of a graph to be

56
manually traced and calibrated, resulting in a set of coordinates that could then be
plotted against the output of ANSYS/LS-DYNA. Figure 4.2b shows the extracted
data from figure 4.2a

80

70

60
Hugoniot pressure (MPa)

50

40

30

20

10

0
0 25 50 75 100 125 150 175 200 225 250

Impact velocity (m/s)

Figure 4.1: Hugoniot pressure vs. impact velocity from experimental data

5
Non-dimensionalized pressure

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Non-dimensionalized time

(a) Printed graph (b) Extracted data

Figure 4.2: Example of data extraction from printed graph

57
4.1.2 Comparison

Figures 4.3, 4.4 and 4.5 show the resulting comparisons that were made after simula-
tions were performed to replicate three of Wilbeck’s tests involving gelatin impactors.
Very good agreement can be seen between the Hugoniot pressures produced in
the simulation and those measured in the first two experimental data sets. The third
set, in which the impact velocity is 171 m s−1 , does not show such good agreement.
However, Wilbeck gives data for another test that was performed at 258 m s−1 and
which gives a Hungoniot pressure lower than that measured for the 171 m s−1 case,
suggesting that the significant jump in Hugoniot measurements between 158 m s−1
and 171 m s−1 may be due to overshoot in the original pressure transducer rather
than inaccuracy on the part of the finite element model.

5
Non-dimensionalized pressure

Simulation
3 Wilbeck

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Non-dimensionalized time

Figure 4.3: Comparison with experimental data for velocity of 117 m s−1

58
6

5
Non-dimensionalized pressure

3 Simulation
Wilbeck

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Non-dimensionalized time

Figure 4.4: Comparison with experimental data for velocity of 158 m s−1

10

7
Non-dimensionalized pressure

5 Simulation
Wilbeck
4

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Non-dimensionalized time

Figure 4.5: Comparison with experimental data for velocity of 171 m s−1

59
4.2 Application to flexible fan blades

The following study is not intended to apply to a specific fan blade design, since there
are a wide range of designs available. Rather, a generic geometry will be used, and
the aim will be to improve understanding of the effect of changing particular aspects
of that geometry.

4.2.1 Fan blade model

Figure 4.6 shows the basic fan blade shape than was created in ANSYS. The nodes
at the bottom edge are fully constrained, so as to represent the fixed connection to
the hub. The common Piecewise Linear Plasticity material model is used, with the
Cowper-Symonds method of calculating the yield stress based on the effective strain
rate. The blade is meshed with around 1000 SHELL163 elements.

Parameter Description
Geometry Generic blade shape
Length = 1 m
Maximum chord width = 0.5 m
Material model Piecewise Linear Plasticity
Bottom edge fully constrained
Material properties ρ = 4 × 103 kg m−3
E = 1 × 101 1 Pa
ν = 0.3
σy = 1 × 108 Pa
Etan = 1 × 105 Pa
F = 40
C = 5, P = 10
Element selection SHELL163
Discretization ≈ 1000 elements, average size = 0.02 m

Table 4.1: Basic fan blade model definition

60
(a) Front view (b) Top view

Figure 4.6: Basic fan blade geometry

4.2.2 Geometry investigations

Many simulations were carried out, first varying the maximum width (chord) of the
blade and then the height. Identical birds travelling at the same velocity were used,
and impact always occured near the leading edge at the tip of the blade (as shown
in 4.7). The maximum plastic strain sustained at the impact location is used as
an indicator of the blade’s robustness and ability to withstand strikes. Figures 4.8
and 4.8 show the resulting correlations between the two geometrical parameters and
plastic strain.

61
(a) (b) (c)

Figure 4.7: Deformation during collision with fan blade

0.006

0.005

0.004
Maximum plastic strain

0.003

0.002

0.001

0.000
0.35 0.4 0.45 0.5 0.55 0.6 0.65 0.7 0.75

Maximum chord width (m)

Figure 4.8: Correlation between maximum chord width and peak plastic strain

62
0.004500

0.004000

0.003500

0.003000
Maximum plastic strain

0.002500

0.002000

0.001500

0.001000

0.000500

0.000000
0.6 0.7 0.8 0.9 1 1.1 1.2 1.3

Blade height (m)

Figure 4.9: Correlation between blade height and peak plastic strain

The results for maximum chord width show that wider blades experienced more
plastic strain when stuck near their leading edge. For blade height the results followed
an opposite trend, with longer blades suffering less plastic deformation.

63
5 Conclusions

An initial model of a bird striking a rigid plate was developed that displayed the
expected behaviour, and that produced a pressure profile exhibiting the characteristics
of a transient, peak Hugoniot pressure followed by a more steady stagnation period,
as predicted by the hydrodynamic impact theory.
Investigations were performed into the model of the rigid plate, and it was found
that a three-dimensional square shape is the most desirable configuration, and fur-
thermore that a single element could be used with accpetable results.
Investigations into the model of the bird confirmed the view that a hemispherically-
ended cylinder with an length to diamter ratio of two is the best simplified represen-
tation of a bird’s torso. Node merging was found to be important when using such a
geometry. Investigations into mesh refinement showed that little extra benefit was to
be found in using a very dense mesh, and that 13824 elements is sufficient to produce
good results while not requiring excessive computation time. It was also found that
even a more coarse mesh of 3968 elements produces reasonable results. Experimen-
tal data from Barber et al. was extracted, and it was found that the experiments
broadly showed the trend, as implied by hydrodynamic impact theory, that Hugoniot
pressure increases with the square of impact velocity. Simulations were carried out
to compare with pressure profiles from Wilbeck, and the model proved very accurate
at low velocities. The discrepancy observed at a higher velocity may have been due
to overshoot in the original pressure measurements.
A generic, deformable fan blade model was developed, and impact tests were
carried out with blades of different chord widths and lengths. Results comparing
the maximum plastic strain experienced by each blade suggest that wider blades are

64
susceptible to greater damage from bird impacts, while longer ones are more robust.

65
References

[1] Associate Administrator for Airports. Wildlife strikes to civil aircraft in the
united states 1990-2006. Serial Report 13, Federal Aviation Administration Na-
tional Wildlife Strike Database, Washington, DC, July 2007.

[2] John Thorpe. Fatalities and destroyed civil aircraft due to bird strikes 1912-
2002. Technical Report IBSC26/WP-SA1, International Bird Strike Committee,
Warsaw, May 2003.

[3] Roger E. Bilstein. Flight in America: from the Wrights to the Astronauts. JHU
Press, 3rd edition, 2001.

[4] The Washington Post. Fall kills aviator, April 1912.

[5] V. E. F. Solman. Birds and aircraft. Biological Conservation, 5(2):79–86, 1973.

[6] Edward C. Cleary and Richard A. Dolbeer. Wildlife hazard management at


airports: a manual for airport personnel. Staff publications, USDA National
Wildlife Research Center, July 2005.

[7] Joanna Burger. Jet aircraft noise and bird strikes: Why more birds are being
hit. Environmental Pollution (Series A), 30:143–152, 1983.

[8] John Thorpe. The bird strike implications of low cost airlines using smaller
airports. Technical report, International Bird Strike Committee, Warsaw, May
2003.

66
[9] Richard A. Dolbeer and Paul Eschenfelder. Amplified bird-strike risks related
to population increases of large birds in north america. Technical Report
IBSC26/WP-OS4, International Bird Strike Committee, May 2003.

[10] Safety Regulation Group. Large flocking birds. Technical report, UK Civil
Aviation Authority, Aviation House, Gatwick Airport South, May 2002.

[11] John R. Allan. The costs of bird strikes and bird strike prevention. In Human
Conflicts with Wildlife: Economic Considerations, pages 147–153, Lincoln, NE,
2000. USDA National Wildlife Research Center Symposia.

[12] John Thorpe. Fatalities and destroyed civil aircraft due to bird strikes 2002-
2004. Technical Report IBSC27/WP II-3, International Bird Strike Committee,
Athens, May 2005.

[13] V. E. Jacobi and M. Beklova. Can the pilot of an aircraft prevent a collison with
birds? Technical Report IBSC15 WP10, International Bird Strike Committee,
1981.

[14] Transport Canada. Sharing the skies (TP 13549). Ottawa, 2nd edition, March
2004.

[15] Transport Canada. Wildlife control procedures manual (TP 11500). Ottawa, 3rd
edition, March 2002.

[16] Transport Canada. Land Use in the Vicinity of Airports (TP 1247). Ottawa,
8th edition, May 2005.

[17] Paul A. Gallaher. The airport environment and its effect on wildlife damage to
aircraft. Master’s thesis, Embry-Riddle Aeronautical University, May 2003.

67
[18] Federal Aviation Administration. US Code of Federal Regulations, Title 14, Part
33, Subpart E, Section 33.76.

[19] Paul Eschenfelder. Jet engine certification standards. Technical Report


IBSC25/WP-IE1, International Bird Strike Committee, Amsterdam, April 2000.

[20] John Downer. When the chick hits the fan: Representativeness and reproducibil-
ity in technological tests. Social Studies of Science, 37(1):7–26, 2007.

[21] A. W. R. Allcock and D.M. Collin. The development of a dummy bird for use
in bird strik research. Technical Report C.P. No. 1071, Ministry of Technology
Aeronautical Research Council, London, 1969.

[22] John P. Barber and Henry R. Taylor. Characterization of bird impacts on a


rigid plate: part 1. Technical Report AFF111-TE-75-15, University of Dayton
Research Institute, January 1975.

[23] John P. Barber, John M. Klyce, Philip F. Fry, and Henry R. Taylor. Impact of
soft bodies on jet engine fan blades. Technical Report AFML-TR-77-29, Univer-
sity of Dayton Research Institute, April 1977.

[24] John P. Barber, Henry R. Taylor, and James S. Willbeck. Bird impact forces
and pressures on rigid and compliant targets. Technical Report AFFDL-TR-77-
60, University of Dayton Research Institute, 300 College Park Avenue, Dayton,
Ohio 45459, May 1978.

[25] James S. Wilbeck and John P. Barber. Bird impact loading. The Shock and
Vibration Bulletin, 48(2):115–120, September 1978.

68
[26] Stanley S. Cook. Erosion by water-hammer. In Proceedings of the Royal Society
of London, volume 119, pages 481–488, July 1928.

[27] W. P. Graebel. Engineering Fluid Mechanics. Taylor & Francis, first edition,
2001.

[28] F. P. Bowden and J. H. Brunton. The deformation of solids by liquid impact at


supersonic speeds. In Proceedings of the Royal Society of London, volume 263,
pages 433–450, October 1961.

[29] James S. Wilbeck. Impact behavious of low strength projectiles. Technical


Report AFML-TR-77-134, Air Force Materials Laboratories, Air Force Wright
Aeronautical Laboratories, Wright Patterson Air Force Base, Ohio 45433, July
1978.

[30] Frank M. White. Fluid Mechanics. McGraw-Hill, sixth edition, 2008.

[31] F. P. Bowden and J. E. Field. The brittle fracture of solids by liquid impact,
by solid impact, and by shock. In Proceedings of the Royal Society of London,
volume 282, pages 331–352, November 1964.

[32] S. M. Yahya. Fundamentals of compressible flow. New Age Publishers, 2003.

[33] M. A. Lavoie, A. Gakwaya, M. Nejad Ensan, and D. G. Zimcik. Review of


existing numerical methods and validation procedure available for bird strike
modelling. ICCES, 2(4):111–118, 2007.

[34] Alastair F. Johnson and Martin Holzapfel. Modelling soft body impact on com-
posite structures. Composite Structures, 61:103–113, 2003.

69
[35] A. Airoldi and B. Cacchione. Modelling of impact forces and pressures in la-
grangian bird strike analyses. International Journal of Impact Engineering,
32:1651–1677, 2006.

[36] C. A. Felippa. A historical outline of matrix structural analysis: a play in three


acts. Computers & Structures, 79(14):1313–1324, June 2001.

[37] R. E. McCarty. Computer analysis of bird-resistant aircraft transparencies. In


Survival and Flight Equipment Association, Annual Symposium, number 17th,
pages 93–97, 1980.

[38] R. E. McCarty. Finite element analysis of F-16 aircraft canopy dynamic re-
sponse to bird impact loading. In Structures, Structural Dynamics, and Ma-
terials Conference, number AIAA-1980-804, Seattle, Wash., May 1980. USAF,
Flight Dynamics Laboratory, Wright-Patterson AFB, Ohio.

[39] T. Belytschko, E. Privitzer, W. Mindle, and T. Wicks. Computer simulation of


canopy-pilot response to bird-strike. Technical Report ADA080122, Northwest-
ern University Department of Civil Engineering, October 1979.

[40] M. S. Hirschbein. Bird impact analysis package for turbine engine fan blades. In
Structures, Structural Dynamics and Materials Conference, Collection of Tech-
nical Papers, Part 2, pages 326–334, May 1982.

[41] N. F. Martin. Nonlinear finite element analysis to predict fan blade impact
damage. In ASME, SAE, and ASEE, Joint Propulsion Conference, number
AIAA-1988-3163, July 1988.

70
[42] Frederick Stoll and Robert A. Brockman. Finite element simulation of high-
speed soft-body impacts. Technical Report AIAA-97-1093, American Institute
of Aeronautics and Astronautics, University of Dayton Research Institute, 300
College Park Avenue, Dayton, OH 45469-0110, 1997.

[43] R. H. Mao, S. A. Meguid, and T. Y. Ng. Finite element modeling of a bird


striking an engine fan blade. Journal of Aircraft, 44:583–596, March-April 2007.

[44] Rajeev Jain and K. Ramachandra. Bird impact analysis of pre-stressed fan
blades using explicit finite element code. In Proceedings of the International Gas
Turbine Congress, number IGTC2003TokyoTS-009, Tokyo, November 2003.

[45] A. G. Hanssen, Y. Girard, L. Olovsson, T. Berstad, and M. Langseth. A numeri-


cal model for bird strike of aluminium foam-based sandwich panels. International
Journal of Impact Engineering, 32:1127–1144, 2006.

[46] Vijay K. Goyal, Carlos A. Huertas, Jose R. Borrero, and Tomas R. Leutwiler.
Robust bird-strike modeling based on ALE formulation using LS-DYNA. In 47th
AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materi-
als Conference, number AIAA 2006-1759. American Institute of Aeronautics and
Astronautics, 2006.

[47] S. Audic, M. Berthillier, J. Bonini, H. Bung, and A. Combescure. Prediction of


bird impact in hollow fan blades. In 36 AIAA/ASME/SAE/ASEE Joint Propul-
sion Conference and Exhibit, number AIAA 2000-3201, Huntsville, Alabama,
July 2000. American Institute of Aeronautics and Astronautics.

71
[48] M. A. McCarthy, C. T. McCarthy, J. R. Xiao, A. Kamoulakos, J. Ramos, J. P.
Gallard, and V. Melito. Modelling of bird strike on an aircraft wing leading edge
made from fibre metal laminates part 2: Modelling of impact with SPH bird
model. Applied Composite Materials, 11:317–340, 2004.

[49] B. Langrand, A. S. Bayart, Y. Chauveau, and E. Deletombe. Assessment of


multi-physics FE methods for bird strike modelling - Application to a metallic
riveted airframe. International Journal of Crashworthiness, 7(4):415–428, 2002.

[50] Vijay K. Goyal, Carlos A. Huertas, Tomas R. Leutwiler, and Jose R. Borrero.
Robust bird-strike modeling based on SPH formulation using LS-DYNA. In 47th
AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materi-
als Conference, number AIAA 2006-1878. American Institute of Aeronautics and
Astronautics, 2006.

[51] Rade Vignjevic. Review of development of the smooth particle hydrodynamics


(SPH) method. Crashworthiness, Impact and Structural Mechanics (CISM),
Cranfield University, UK.

[52] R. A. Brockman and T. W. Held. Explicit finite element method for transparency
impact analysis. Final technical report Sep 1988-Sep 1990 WL-TR-91-3006, Uni-
versity of Dayton Research Institute, 300 College Park Avenue, Dayton, OH
45469-0110, June 1991.

[53] Richard Budgey. The development of a substitute artificial bird by the interna-
tional birdstrike research group for use in aircraft component testing. Techni-
cal Report IBSC25/WP-IE3, International Bird Strike Committee, Amsterdam,
April 2000.

72
[54] Koh Chee Chuan. Finite element analysis of bird strikes on composite and glass
panels. Master’s thesis, National University of Singapore, 2006.

[55] S. T. Jenq, F. B. Hsiao, I. C. Lin, D. G. Zimcik, and M. Nejad Ensan. Simula-


tion of a rigid plate hit by a cylindrical hemi-spherical tip-ended soft impactor.
Computational Materials Science, 39:518–526, 2007.

[56] D. Chevrolet, S. Audic, and J. Bonini. Bird impact analysis on a bladed disk. In
RTO AVT Symposium on ”Reduction of Military Vehicle Acquisition Time and
Cost through Advanced Modelling and Virtual Simulation”. RTO AVT, April
2002.

[57] S. A. Meguid, R. H. Mao, and T. Y. Ng. FE analysis of geometry effects of an


artificial bird striking an aeroengine fan blade. International Journal of Impact
Engineering, 2007.

[58] Abdel-Rahman A. Ragab and Salah Eldin Ahm Bayoumi. Engineering Solid
Mechanics. CRC Press, 1999.

[59] ANSYS, Inc., Southpoint, 275 Technology Drive, Canonsburg, PA 15317. ANSYS
LS-DYNA User’s Guide, January 2007.

73

Potrebbero piacerti anche