You are on page 1of 43

O N THE AEROELASTIC

BEHAVIOR OF A
SWEPT- FORWARD WING
The Grumman X-29

Research work in the field of


Aeroelasticity
The author
Piercarlo Mirto is a master’s student at the University of Naples Federico II, as an
aeronautical engineering major. Enthusiast worker, he focuses his attention on Struc-
tural analysis, Aeroelasticity, Aeroacoustics, Impact dynamics, and Aircraft design, try-
ing to find their roots in Aviation history.

Piercarlo Mirto

Personal motivational thought:


“Of course you have to keep your feet on the ground. How
else could a plane take off?”

Profiles: LinkedIn, Academia.


CONTENTS

Contents

Introduction 11

1 Historical background 15
1.1 Development . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
1.2 Challenges . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
1.2.1 Aeroelastic tailoring . . . . . . . . . . . . . . . . . . . . . . . . 16
1.2.2 Instability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.3 Testing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.4 Research flights . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
1.5 New priorities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

2 Static Aeroelasticity 19
2.1 Swept wings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.1.1 Load estimation . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.1.2 Effective angle of attack . . . . . . . . . . . . . . . . . . . . . . 21
2.2 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.2.1 Characteristic ratios . . . . . . . . . . . . . . . . . . . . . . . . 22
2.3 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

3 Numerical methods 25
3.1 Interpolation methods . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.2 Doublet-Lattice Subsonic Lifting Surface Theory . . . . . . . . . . . . 25
3.3 Nastran . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

4 Divergence speed analysis 29


4.1 Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
4.2 Nastran input inizialization . . . . . . . . . . . . . . . . . . . . . . . . 30
4.3 Bulk data input . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
4.3.1 Structural model . . . . . . . . . . . . . . . . . . . . . . . . . . 31
4.3.2 Aerodynamic model . . . . . . . . . . . . . . . . . . . . . . . . 34
4.3.3 Static Aeroelastic Input . . . . . . . . . . . . . . . . . . . . . . 37
4.4 Femap input . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
4.5 Output . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

CONTENTS
CONTENTS

CONTENTS
LIST OF FIGURES

List of Figures

1.1 The Grumman X-29. Credits: NASA . . . . . . . . . . . . . . . . . . . 16


2.1 Geometry of idealized swept wing, showing a chordwise section used
in the computation of aerodynamic loads. . . . . . . . . . . . . . . . . 20
2.2 Static aeroelastic characteristics of a uniform property swept wing, show-
ing spanwise center of pressure change Δ 𝑦★ as a function of the param-
eters 𝑎 and 𝑏/𝑎 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.1 An example of Lattice for a swept tapered wing. . . . . . . . . . . . . . 26
4.1 Idealization of FSW Configuration . . . . . . . . . . . . . . . . . . . . 29
4.2 Structural model overview. . . . . . . . . . . . . . . . . . . . . . . . . 33
4.3 Aerodynamic elements overview. . . . . . . . . . . . . . . . . . . . . . 36
4.4 Spline overview. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
4.5 Three-dimensional views of the FEM model with element thickness high-
lighted. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
4.6 Rotation distribution. . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
4.7 Tranlation distribution. . . . . . . . . . . . . . . . . . . . . . . . . . . 40

LIST OF FIGURES
LIST OF FIGURES

LIST OF FIGURES
GLOSSARY

Glossary

Variables
Physical entity Symbol Unit of measure
Aircraft plunging degree of freedom ℎ ft
Aircraft pitching degree of freedom 𝛼 rad
Aircraft bending stiffness 𝐸𝐼 lb·ft2
Aircraft torsional stiffness 𝐺𝐽 lb·ft2
Aircraft bending-torsion coupling parameter 𝐾 lb·ft
Upward loading about the 𝑦 -axis per unit length 𝑝 lb/ft
Nose-up torque around the 𝑦 -axis per unit lenght 𝐾 lb
Two-dimensional lift coefficient 𝑐𝑟ℓ ★
Two-dimensional lift-curve slope 𝑎0 lb
Aircraft weight 𝑊 lb
Sweep angle Λ deg
Dynamic pressure 𝑞 lb/ft2
Slope of the bending deformation Γ ★
Load factor 𝑛 ★
Wing semispan ℓ ft
Load factor 𝑛 ★
Wing chord at a generic station 𝑐 ft

GLOSSARY
GLOSSARY

GLOSSARY
ABSTRACT

Abstract
The present work aims to introduce the reader to the static divergence problem, a
structural behavior that occurs under certain conditions due to its interaction with the
surrounding moving fluid. For decades this phenomenon made the use of forward-
swept wings almost impossible, resulting in an inability to take advantage of the ben-
efits of stalling that this configuration has over the swept-back configuration.
The work focuses its attention on an ’80s experimental aircraft, the Grumman X-29,
specifically developed to finally overcome this limitation. After a brief historical intro-
duction, it proposes a FEM model to analyze the aircraft response to the divergence
problem. The reader can find the script and all the outputs in the public folder:
https://www.dropbox.com/sh/qkau1lbj4zpbwzi/AAAsAdzoFcnHM0cpyvnIG-kCa?dl=0.

ABSTRACT
ABSTRACT

ABSTRACT
INTRODUCTION

Introduction

Sources: [1], [2]


Aeroelasticity is the study of the effect of aerodynamic forces on elastic bodies, such
as aircraft wings or compressor blades. The generated aerodynamic forces entirely
depend on the deformed shape of the structure in the flow. One of the aeroelastic
problems is the stability — or rather, the instability — of a structure in a flow.
Flexibility is generally associated with lightweight, so aeroelastic problems were dis-
covered and known from the earliest days of flight. Wing divergence, a static aeroelas-
tic problem, has been surmised as the probable cause of S. P. Langley’s failure to control
his machine in its first flight over the Potomac in 1903.
Although numerous other aeroelastic incidents followed in the pre-World War II pe-
riod, problems in aeroelasticity did not attain the prominent role they play now until
the early stages of the war. That is because aircraft speeds were relatively low and their
thickness-to-chord ratio was relatively high, thus giving the structural engineer the re-
quired design flexibility to obtain sufficient bending and torsional rigidities: structures
were sufficiently rigid to prevent most aeroelastic phenomena.
By 1916, problems relating to flutter were prevented by isolating the motions in sev-
eral freedoms. Examples include mass balancing of the lifting surfaces at the expense
of additional weight or the enhancement of the relevant natural frequencies (i.e., stiff-
ness), resulting in an increase of the lowest critical flutter speed safely beyond possible
flight speeds. Thus, the two grassroots solutions for aeroelastic problems were already
well established and, together with damping mechanisms, are still the necessary tools
to prevent aeroelastic instabilities.

The introduction of the swept wing


For most designs developed between the two World Wars, flutter — a dynamic aeroe-
lastic instability that usually involves coupling between an almost pure bending and a
pure torsional mode due to the unswept and more or less constant chord wing plan-
form — would most often occur at a lower airspeed than divergence and, as a result, it
received more attention.
The situation changed in the late ’40s with the first approaches towards transonic flight.
The exploration of this field led to the introduction of the swept wings. The divergence
speed drops dramatically for even slight forward sweep angles due to the wash-in ef-
fect: the spanwise bending of a swept-forward wing induces an increase in the local

INTRODUCTION
INTRODUCTION14

streamwise angle of attack, increasing aerodynamic loads. A swept-back wing experi-


ences the opposite effect, called wash-out.

Composite materials
The introduction of composite material in the area of aircraft design in the early 70s
has led to new airframe design concepts as well as the re-evaluation of older concepts.
Today almost every aerospace company develops products made with fiber-reinforced
composite materials. Consequently, the successful application of laminated compos-
ite materials in aircraft structures, coupled with their anisotropic property, has gener-
ated a renewed interest in aeroelasticity.
The technology to design for a predetermined aeroelastic response of a lifting sur-
face using composite materials has been named aeroelastic tailoring, defined by ref-
erence [2] as
“ the embodiment of directional stiffness into an aircraft structural design
to control aeroelastic deformation, static or dynamic, in such a fashion as to
affect the aerodynamic and structural performance of that aircraft in a ben-
eficial way. ”
As a result, a great deal of research has been devoted to the improvement of the aeroe-
lastic stability of wings using composites.
A significant example of this mindset is the experimental aircraft Grumman X-29,
where the anisotropic nature of the fiber composite is used to minimize the torsional
divergence problem. Thus, the directional properties of laminated composite mate-
rials can be oriented to alter static and dynamic characteristics of composite aircraft
wings, leading to aeroelastic tailoring and thus to possible optimum design.

INTRODUCTION Piercarlo Mirto – Corso di Studi in Ingegneria Aeronautica, Università degli Studi di Napoli Federico II
Chapter
1
Historical background
Sources: [3]

In the ’80s, the forward-swept wing was not a new concept. The technology had been
previously implemented on the German plane Junkers Ju-287 in 1944. The Germans
worked extensively with this type of aircraft but abandoned the project during the Sec-
ond World War partly due to the war’s outcome.
A second example appeared in 1964: the HBF-320 Hansa Jet, a ten-seat business jet
that was designed and produced by German aircraft manufacturer Hamburger Flugzeug-
bau.
The Grumman’s experimental aircraft X-29 — figure 1.1 — was the third plane to
fly using a forward-swept wing and the first American aircraft. However, it achieved
what no forward-swept wing plane achieved before: supersonic speed. The configu-
ration sought to test new canard control surfaces and advanced materials to improve
maneuverability response at a high angle of attack.
Although promising and reviewed favorably by test pilots, the aircraft’s new tech-
nologies made it one of the most aerodynamically unstable aircraft in aviation history.
An incredibly complex system was needed to continually calibrate the controls to keep
the X-29 in the air. Still, the promise of the potential new technologies was so enticing
that even the Russians couldn’t help themselves from trying to replicate them.

1.1 Development
The X-29 project started at the height of the Cold War. It involved a collaboration
between NASA , the US Air Force, and the Defense Advanced Research Project
Agency (DARPA ). They hoped the unique characteristics would outperform other
modern aircraft.
In 1981 DARPA awarded a fixed-price contract to the aircraft company Grumman to
create this ultimate fighter aircraft. As part of the deal, two planes were constructed.
Grumman would carry out four test flights after successful testing, to then hand the
aircraft over to the US military.
Design work commenced in 1984. Before this project, the Unites States prioritized
aft-swept-wings due to structural limitations that prohibited different wing types. The
primary design challenge The company faced laying how to prevent the wings from
twisting due to their direction. If that wasn’t challenging enough, it was incumbent
16 Chapter 1 Historical background

upon the team to keep the aircraft light. They needed to develop stiffer wings, that
required new materials.
The two airplanes were built by Grumman modifying existing airframes of a F-5A Free-
dom Fighter (Northrop corporation).

1.2 Challenges
The new shape of the X-29 aircraft presented several challenges for American engi-
neers. Among these was the matter of weight. The forward-facing wing designs would
be subject to extreme twisting force by the headwind. This threatened to break them
where they weren’t reinforced: the aerodynamic lift would twist the wing’s leading edge
upward, which would lead to structural failure.

1.2.1 Aeroelastic tailoring


For what said, making the wings out of metal was out of the question because the
added weight would kill prospective performance. As a solution, Grumman decided
to use a new mix of composite materials which have since become standard for both
commercial and military planes.
Anisotropic elastic coupling solved the problem by bending and twisting the car-
bon fiber composite material without compromising weight. Therefore, the X-29 in-
troduced a laminate that could produce coupling between torsion and bending. This
resulted in an increased lift. The bending loads would force the wingtips upward. Fur-
thermore, the coupling resisted torsion loads, which would otherwise twist the wing

Figure 1.1 The Grumman X-29. Credits: NASA

Piercarlo Mirto – Corso di Studi in Ingegneria Aeronautica, Università degli Studi di Napoli Federico II
1.3 Testing 17

to higher angles of attack. The leading edge would be turned downwards and there-
fore solve this problem. By reducing lift, the loads decreased and prevented aeroelastic
divergence in the material.

1.2.2 Instability
The aircraft architecture with the center of gravity behind the aerodynamic center was
intrinsically unstable. Knowing that without a ballast equal in weight to a car on the
nose of the airplane they would not be able to stabilize the aircraft, Grumman decided
to approach the issue from a novel angle: the design choice referred to as a three-
surface aircraft. It employed three longitudinal controls through the canard, forward-
swept wing, and aft-straight control surfaces. As a result, the canards and wings reduce
both wave drag and trim drag. The strikes provided trim wherever the center was off
and therefore reduced drag.
That didn’t eliminate all the problems. Throughout development, the new struc-
ture produced an unprecedented challenge. NASA ’s Armstrong flight research center
historian Christian Gelzer stated:
“ it was unflyable, literally, without a digital flight computer on board which
made connections to the flight path 40 times a second. ”
Determining the problem early allowed the team sufficient time to tackle it. The en-
gineers decided to control forward canards and rear flaps with a trifecta of redundant
computers as a fail-safe system. Stability was achieved through the flight control sys-
tem, capable of up to forty corrections per second. The computer system, backed by
three equally redundant analog computers rechecked and corrected its own assess-
ments, drastically reducing the probability of errors.
While any of the three machines could theoretically assist in allowing the aircraft to
fly, redundancy made the system nearly infallible. Each computer sent measurements
and made quick assessments, weighted against the two others. Any malfunctions were
immediately detected. At the time, calculations by experts concluded that the three-
computer system was as unlikely to fail as the regular mechanics on a conventionally
arranged plane. Still, the unlikely possibility of system failure was terrifying. Always
Christian Gelzer said:
“ the engineers concluded that if all three flight computers had failed to-
gether, the airplane would have broken up around the pilot before the pilot
had a chance to eject. ”

1.3 Testing
Finally, the aircraft was completed, and the first X-29 was ready for testing. Its maiden
flight took off out of Edwards air force base on December 13th, 1984 by chief test pilot
Grumman’s Chuck Sewell.
Pilots who flew the X-29 had nothing but positive things to say about the experi-
mental test bed. Due to its fantastic thrust-to-weight ratio during takeoff and minimal
turbulence, pilots agreed that other aircraft paled in comparison.
After pilots successfully performed five consecutive flights in a single day, the plane
was approved.

Aeroelasticity
18 Chapter 1 Historical background

1.4 Research flights


Following its maiden flight, the X-29 commenced a four-month NASA test program.
Despite complications and initial challenges regarding the instability of the aircraft,
the three-computer system proved invaluable. The plane was reliable and even more
critically deployable. By August 1986, the plane flew multiple times for research mis-
sions with durations all over three hours. However, the first model of the aircraft had
no spin recovery parachute since all test flights avoided any maneuver which would
take the plane out of a controlled flight. Seeing this drawback, the second model was
equipped with a parachute and tested for high angles of attack. It proved to be maneu-
verable up to 25°, but reaching a maximum angle of 67°.
Between 1984 and 1991 the two X-29 planes flew 242 times. The Dryden flight re-
search center published an internal report on the new technologies and techniques
shown by the X-29. New applications of previous technologies, such as aeroelastic
tailoring, improved the plane’s control of structural divergence, aircraft control, and
handling. Its successes mostly relied on the three-surface longitudinal control and the
advancements made with the second model regarding the high angle of attack.
Regardless of the program’s future, the X-29 had proven essential data for military air-
craft production.

1.5 New priorities


The X-29 broke more flight records than any other X-label experimental program. Still,
it was scrapped soon after in 1992. The general consensus among aviation historians
and enthusiasts is that the cancellation was due to the Department of Defense pri-
oritizing newly researched stealth technologies for aircraft. The forward-swept wing
design of the X-29 was incompatible with this new focus. Moreover, others believe the
seemingly superior Russian Sukhoi Su-47 intimidated the United States and led the
Department of Defense to gravitate toward more ambitious projects.

Piercarlo Mirto – Corso di Studi in Ingegneria Aeronautica, Università degli Studi di Napoli Federico II
Chapter
2
Static Aeroelasticity
Sources: [4], [5], [6], [7], [8]

Several simplifying features characterize static aeroelastic problems. By definition,


time does not appear as an independent variable, so vibratory inertial forces are elimi-
nated from equilibrium equations. In addition, aerodynamic forces can be based upon
the assumption of steady flow rather than the more complex unsteady theories.
The majority of static aeroelastic problems may be divided into two main classes:
• Load distribution problems.
• Stability problems.
This chapter will focus on the first category. Specifically, we are interested in the static
instability known as divergence, that in Ref. [4], is defined as follows:
“ Divergence is the static instability of an airfoil in torsion which occurs when
the torsional rigidity of the structure is exceeded by aerodynamic twisting
moments. ”

2.1 Swept wings


The differential equations of equilibrium, in terms of bending and torsional deforma-
tions, of a slender swept wing are provided in Ref. [7]. It is formulated by considering
the aerodynamic loads to act upon chordwise segments of the wing, perpendicular to
the wing reference axis, as shown in Figure 2.1.
The reference axis, labeled as the 𝑦 -axis, is located equidistant between the front and
rear edges of the box beam and lies in the geometric midplane of the box, halfway
between the upper and lower surfaces of the cover sheets.
The governing equations of equilibrium for such a wing are found to be:
 
𝑑2 𝑑2ℎ 𝑑𝛼
𝐸𝐼 −𝐾 = 𝑝( 𝑦) (2.1)
𝑑 𝑦2 𝑑 𝑦2 𝑑𝑦
 
𝑑 𝑑2ℎ 𝑑𝛼
−𝐾 2 − 𝐺 𝐽 = −𝑡 ( 𝑦 ) . (2.2)
𝑑𝑦 𝑑𝑦 𝑑𝑦

The structural parameters 𝐸𝐼 , 𝐺 𝐽 , and 𝐾 are respectively the beam bending stiffness,
torsional stiffness, and bending-torsion coupling parameter. The variables ℎ ( 𝑦 ) and
20 Chapter 2 Static Aeroelasticity

Figure 2.1 Geometry of idealized swept wing, showing a chordwise section used in the computa-
tion of aerodynamic loads.

𝛼 ( 𝑦 ) represent respectively the upward bending displacement of the wing reference


axis and the nose-up twisting rotation of the wing sections. The term 𝑝 ( 𝑦 ) is the up-
ward load per unit length, measured along the 𝑦 -axis, while 𝑡 ( 𝑦 ) is the nose-up torque
per unit length, also measured about the 𝑦 -axis.

2.1.1 Load estimation


Reference [6] describes one method that may be used to compute the wing loads 𝑝 ( 𝑦 )
and 𝑡 ( 𝑦 ) for this type of slender wing model. For this calculation, aerodynamic strip
theory is used to develop the following expressions:

𝑝 ( 𝑦 ) = 𝑐 𝑐𝑟ℓ 𝑞 cos2 Λ + 𝑞 𝑐 𝑎0 cos2 Λ · ( 𝛼 − Γ tg Λ) − 𝑛𝑊


𝑡 ( 𝑦 ) = 𝑞 𝑐 𝑒 𝑐𝑟ℓ cos2 Λ + 𝑞 𝑐2 𝑐𝑚.𝑎.𝑐. cos2 Λ − 𝑛𝑊 𝑑 + 𝑞 𝑐 𝑒 𝑎0 · ( 𝛼 − Γ tg Λ) cos2 Λ .

the term 𝑐𝑟ℓ corresponds to the two-dimensional lift coefficient for wing sections per-
pendicular to the reference axis for a wing with no flexibility, while 𝑎0 refers to the two-
dimensional lift-curve slope of the same section.
Γ is the slope of the bending deformation, defined as Γ = 𝑑𝑑ℎ𝑦 , while 𝑛𝑊 represents the
distributed inertia load per unit length for a wing of weight 𝑊 pounds per unit length
operating at a load factor 𝑛.
The cited report proceeds in the analysis introducing the nondimensional coordinate
𝑦
𝜂 =1− .

Piercarlo Mirto – Corso di Studi in Ingegneria Aeronautica, Università degli Studi di Napoli Federico II
2.2 Results 21

2.1.2 Effective angle of attack


The equations 2.1 are coupled together by the dependence of the aerodynamic load
upon the term ( 𝛼 − Γ tg Λ) and by the structural coupling parameter 𝐾 . Once defined
the first term as 𝛼𝑒 , it is possible to combine them to form a single equation in terms
of a new variable. It results:
𝑑 3 𝛼𝑒 𝑑𝛼𝑒 𝑑 𝑓1
3
+𝑎 − 𝑏 𝛼𝑒 = − 𝑓2 . (2.3)
𝑑𝑦 𝑑𝑦 𝑑𝑦

The involved quantities are:

𝑎 = 𝑞 ( Ω𝑇 𝑎0 ) Θ
𝑏 = 𝑞 ( Ω 𝐵 𝑎0 ) Ξ
   2

𝑐 𝑐𝑚.𝑎.𝑐. 𝑛𝑊 𝑑 ℓ
𝑓1 = Θ − 1 + 𝑞 Ω𝑇 𝑐𝑟ℓ +

𝑒 𝑐𝑟ℓ 𝐺𝐽
 
 𝑛𝑊 ℓ3
𝑓2 = −Ξ 𝑞 Ω 𝐵 𝑐𝑟ℓ − ,
𝐸𝐼

with
𝑐 𝑒 ℓ2 cos2 Λ 𝑐 ℓ3 cos2 Λ
Ω𝑇 = Ω𝐵 =
𝐺𝐽 𝐸𝐼 
1 − 𝑘 tg Λ tg Λ − 𝑔
  
Θ= Ξ=
1−𝑔𝑘 1−𝑔𝑘

Note that appear the structural coupling parameter nondimensional counterparts 𝑘 =


𝐾 / 𝐸𝐼 and 𝑔 = 𝐾 /𝐺 𝐽 .

Boundary conditions
The boundary conditions, in terms of the variable 𝛼𝑒 , are found by first noting that
both Γ (the bending slope) and 𝛼 are zero at the effective root of the wing (𝜂 = 1). This
leads to the boundary condition
𝛼𝑒 =0.
𝜂 =1

The fact that the bending moment and the twist are zero at the effective wing tip leads
to the equation
𝑑𝛼𝑒
=0.
𝑑𝑦 𝜂 =0

The wing tip zero shear condition, when combined with the torsion equation, evalu-
ated at 𝜂 = 0, leads to the following result:

𝑑 2 𝛼𝑒
+ 𝑎 𝛼𝑒 = 𝑓1
𝑑 𝑦2 𝜂 =0 𝜂=0 𝜂=0

2.2 Results
Ref. [6] and [8] solve the wing deformation problem defined by Equation and its associ-
ated boundary conditions, restricting to an untwisted, uniform planform wing whose

Aeroelasticity
22 Chapter 2 Static Aeroelasticity

pitch attitude is specified. In this case, both 𝑓1 ( 𝜂) and 𝑓2 ( 𝜂) are constants.


Firstly, it is necessary to find the roots of the following characteristic equation, ob-
tained from Equation 2.2:
𝑟3 + 𝑎 𝑟 − 𝑏 = 0 .

Those solutions can be expressed as −2 𝛽 and 𝛽 ± i𝛾 . Consequently, the complete so-


lution is given by:  
𝑓2 𝑓3 ( 𝜂 )
𝛼𝑒 ( 𝜂 ) = 1− . (2.4)
𝑏 𝑓3 ( 1)
The function 𝑓3 ( 𝜂) is defined as
   2    
4 𝛽2 𝛽𝜂 −2 𝛽 𝜂 5 𝛽 + 𝛾2 − 𝛽𝛾 23 𝛽3
𝑓3 ( 𝜂 ) = 𝑒 + 𝑒 cos ( 𝛾 𝜂) + sin ( 𝛾 𝜂)
9 𝛽2 + 𝛾2 9 𝛽2 + 𝛾2 9 𝛽2 𝛾 + 𝛾3

2.2.1 Characteristic ratios


Equation 2.4 provides information necessary to compute the ratio between the total
lift developed by a flexible composite wing and the lift developed by a similar, but in-
flexible wing. Under the strip theory restriction, the lift of the rigid wing is found to
be
𝐿𝑟 = 𝑞 𝑐 𝑐𝑟ℓ ℓ cos2 Λ − 𝑛𝑊 ℓ .

The lift on the flexible wing is


∫ ℓ
𝑞 𝑐 𝑎0 cos2 Λ 𝛼𝑒 𝑑 𝑦 .

𝐿 = 𝐿𝑟 +
0

Performing the necessary integration yields an expression for the flexible-to-rigid wing
lift ratio:
𝐿 −2 𝛽𝛾 𝑒−2 𝛽 + 2 𝛽𝛾 𝑒 𝛽 cos 𝛾 + ( 3 𝛽 2 + 𝛾 2 ) 𝑒 𝛽 sin 𝛾
= 2 −2 𝛽 .
𝐿𝑟 4𝛽 𝑒 + 𝑒 𝛽 [( 5 𝛽 2 + 𝛾 2 ) cos 𝛾 + ( 3 𝛽 2 − 𝛽𝛾 2 ) sin 𝛾 ]
This equation is an expression of the lift effectiveness of the wing. Note that when the
denominator becomes zero, then the wing will diverge.
The ratio of the bending moment at the wing root for the flexible wing to that com-
puted for the rigid wing is found to be

𝛾 𝑒 − 2 𝛽 + 𝑒 𝛽 (−𝛾 cos 𝛾 + 3 𝛽 sin 𝛾 )


 
𝑀
=2
𝑀𝑟 4 𝛽 2 𝛾 𝑒 − 2 𝛽 + 𝑒 𝛽 [( 5 𝛽 2 𝛾 + 𝛾 3 ) cos 𝛾 + ( 3 𝛽 3 − 𝛽𝛾 2 ) sin 𝛾 ]

The bending moment is about the effective root of the wing as shown in Figure 1.
The twisting moment ratio expression is identical to that calculated for the ratio
L/Lr. The spanwise center of pressure, 𝑌𝐶𝑃 , measured along the reference axis, reads
as follows:
ℓ 𝑀 𝐿𝑟
𝑌𝐶𝑃 = .
2 𝑀𝑟 𝐿

2.3 Discussion
The change in the spanwise position of the center of pressure, due to wing flexibility,
affects both the overall directional stability of the aircraft and the wing root stresses.

Piercarlo Mirto – Corso di Studi in Ingegneria Aeronautica, Università degli Studi di Napoli Federico II
2.3 Discussion 23

An outboard shift of the center of pressure results in an increase in stress levels near
the wing root. The equations obtained previously may be used to compute the change
in the spanwise location of the center of pressure position along the swept reference
axis.
Reference [8] presents a study of this change for uniform planform metallic wings, as
a function of two parameters, related to dynamic pressure and the sweep angle. On
the other hand, Ref. [7] presents similar results by suitably redefining parameters to
conform to the composite wing analysis. The two parameters are coefficient 𝑎, 𝑎 is
referred to as the dynamic pressure parameter, and the ratio of 𝑏/𝑎, which is the sweep
parameter given as:
𝑏 ℓ tg Λ − 𝑔
  
𝐺𝐽
= .
𝑎 𝑒 1 − 𝑘 tg Λ 𝐸𝐼
Figure 2.2 shows the influence of these parameters on the movement of the center of
pressure (CP) along the swept reference axis. It introduces the parameter
Δ 𝑦𝐶𝑃
Δ 𝑦★ = .
ℓ/2

Positive Δ 𝑦★ corresponds to an outward CP movement while, conversely, inboard CP


movement is denoted by negative values of Δ 𝑦★.
Labeling the upper right quadrant of Figure 2.2 as Quadrant 1, and numbering the
other quadrants consecutively, counterclockwise, we see that values of 𝑎 and 𝑏/𝑎 that
lie in Quadrant 1 correspond to wings for which divergence is very unlikely to occur.
In Quadrant 2, divergence is very likely, except for wings falling into a very narrow
range of 𝑎 and 𝑏/𝑎. In addition, shifts in the CP locations are outboard so unfavorable
effects of aeroelasticity upon wing stresses may be expected in this region.
In Quadrant 3, divergence is impossible, while, in Quadrant 4, divergence may be pos-
sible for certain combinations of 𝑎 and 𝑏/𝑎.
Reference [8] notes that subsonic, metallic sweptback wings (𝑘 = 𝑔 = 0) are de-
scribed by values of 𝑎 and 𝑏/𝑎 lying in the first quadrant, while subsonic, metallic
swept-forward wings fall into Quadrant 2. For a to be negative for a metallic wing, the
chordwise aerodynamic center must lie behind the wing elastic axis. This situation is
usually associated with supersonic flight. Quadrant 3 illustrates swept-forward metal-
lic wing behavior in the supersonic flight regime, while Quadrant 4 displays metallic
sweptback wing behavior at supersonic speeds. When the wing is constructed of com-
posite materials, the likely effect of sweep on aeroelastic behavior is not as clear as it
is for metallic wings. The elastic coupling between bending and torsion introduces a
new parameter into the definitions of a and b/a. For a particular value of wing sweep
angle and dynamic pressure, the values of a and b/a may be modified by tailoring the
composite laminate.

Aeroelasticity
24 Chapter 2 Static Aeroelasticity

Figure 2.2 Static aeroelastic characteristics of a uniform property swept wing, showing spanwise
center of pressure change Δ 𝑦★ as a function of the parameters 𝑎 and 𝑏/𝑎

Piercarlo Mirto – Corso di Studi in Ingegneria Aeronautica, Università degli Studi di Napoli Federico II
Chapter
3
Numerical methods
Sources: [9], [10]

3.1 Interpolation methods


In the previous work of the undersigned (Ref. [11]), the interpolation problem has
already been treated, even if not explicitly described. The structural and aerodynamic
models generally obey different requirements, which result in different geometrical
schemes. A typical example is the use of a classical beam-like model for the wing (one-
dimensional), parallel to the use of a simple panel mesh as an aerodynamic model
(two-dimensional).
One of the most critical subjects when dealing with aeroelastic calculations is the inter-
connection between the structural modal behavior and the aerodynamic mesh. This
problem is solved by a mathematical algorithm named spline, which allows the inter-
polation between structural (independent) and aerodynamic (dependent) degrees of
freedom.
 
Mathematically speaking, the splining methods lead to an interpolation
 matrix 𝐺 𝑘𝑔
that relates the components of the structural grid point deflections 𝑢𝑔 to the deflec-
tions of the aerodynamic grid points {𝑢𝑘 }:
 h i
𝑢𝑘 = 𝐺 𝑘𝑔 𝑢𝑔 .

3.2 Doublet-Lattice Subsonic Lifting Surface Theory


There is no precise definition of the Doublet Lattice Method (DLM) and the associated
formulae. Ref. [12] provides a possible mathematical formulation behind it, here not
reproduced. What is important to say here, is that formulating the problem of upwash
in the continuum, its kernel function turns out to suffer singularity. The doublet lattice
method is an empirical device that simplifies the integration of this singularity.
Primarily, the advantage gained by the doublet lattice method lies in the relative sim-
plicity of the resulting computer program, especially for complex configurations.
With the doublet lattice method, the continuous pressure doublet sheet is replaced
by a set of pressure doublet lines with finite length. In Figure 3.1 there is a pattern of
nine doublet lines for a swept and tapered wing. Each line (bolded) is contained in
26 Chapter 3 Numerical methods

Figure 3.1 An example of Lattice for a swept tapered wing.

its own box. The simplicity is that all boxes are treated identically, regardless of their
proximity to the wing boundary (i.e. leading edge, trailing edge, or wing tip).
The doublet line is placed at the quarter chord of each box1 . The upwash 𝑤 ( 𝑥, 𝑦, 0)
is evaluated at the 3/4 chord midspan of each box. The empirical nature of the doublet
lattice method arises in the choice of the 3/4 chord and 1/4 chord: by no means there
is any mathematical nor aerodynamical proof that this is the correct location.
As a matter of fact, for nonrectangular wings with swept and tapered boxes, the pro-
grammer is left to his own devices to invent meaningful 1/4 chords and 3/4 chords.
One must be on guard and realize that the flow field generated by this lattice of dou-
blets will not be smooth, especially near the wing surface. What is important is that
the upwash at the 3/4 chord is approximately the same whether one has a constant
strength doublet line at the 1/4 chord or has a continuous doublet sheet with the cor-
rect strength. It is not important the truthfulness of the generated velocity field, but
only the correctness of the downwash values.
The doublet lattice user needs to understand the approximations incurred in dis-
cretizing a doublet sheet into trapezoidal boxes. It should be immediately obvious that
since we have assumed the pressure to be constant within each box, a sufficient num-
ber of boxes is required to capture the steady state (zero frequency) pressure function
accurately.
On the other hand, it is not obvious that we need to increase the number of boxes as we
increase the frequency of oscillation. For instance, the pressure field over a rigid wing,
plunging at high frequency, is not trivial and requires a significant number of boxes to
resolve the standing (pressure) waves. The required box density depends on a combi-
nation of wing deformation and the frequency of motion. The box density should be
increased as the deformation becomes more. spatially wavy and as the temporal fre-
quency of motion increases The doublet lattice user must perform convergence stud-
ies to determine the appropriate box density for their application.
1 Note that to call this a "doublet lattice" is a misnomer. If one views the doublet line segments alone,
no lattice is formed. The name "doublet lattice" arises from the correctly named "vortex lattice" methods
applicable to unsteady incompressible or steady compressible flow over planar wings.

Piercarlo Mirto – Corso di Studi in Ingegneria Aeronautica, Università degli Studi di Napoli Federico II
3.3 Nastran 27

3.3 Nastran
Ref. [9] briefly synthesizes the possibilities of the software Nastran about aeroelastic
analysis. The four solution sequences are illustrated in table 3.1. Specifically, SOL 144
addresses static aeroelasticity and, as such, is useful for making a preliminary assess-
ment of the aircraft design loads and provides estimates for rigid and elastic stability
and control derivatives.

Solution Description
SOL 144 Static Aeroelasticity
SOL 145 Aerodynamic Flutter
SOL 146 Dynamic Aeroelastic Response
SOL 200 Design Sensitivity and Optimization

Table 3.1 Solution Sequences Related to Aeroelasticity.

Aeroelasticity
Chapter
4
Divergence speed analysis
Sources: [13], [14]

The consistent unit system adopted for the analysis is the English Gravitational unit
system [ slug]-[ ft]-[s]. More about that in Ref. [15].

4.1 Model
The (extremely idealized) model — Figure 4.1 — is that of from Ref. [13], which has
been implemented in Nastran in Ref. [14]. The aircraft is assumed perfectly symmet-
ric; only the right-hand side is modeled. The Figure shows the aerodynamical and
structural elements separately for convenience.

Figure 4.1 Idealization of FSW Configuration

The wing has an aspect ratio of 4.0, no taper, twist, or camber, but an incidence of
0.1 deg relative to the fuselage, and a forward sweep angle of 30 deg. The canard has an
30 Chapter 4 Divergence speed analysis

aspect ratio of 1.0, no taper, twist, camber, incidence, or sweep, and is hinged about its
quarter-chord.
The chords of both the wing and canard are 10.0 ft, the reference chord is chosen as the
same length, and the reference area is 𝑆 = 200 ft2 for the half-span model. In Ref. [14]
both subsonic ( 𝑀 = 0.9) and supersonic ( 𝑀 = 1.3) speeds are considered. Here, due to
the limits of the Nastran| Student edition only the subsonic will be developed.
The left part shows the aerodynamic modeling. For the Doublet-Lattice methods of
aerodynamic analysis, the half-span model of the wing is divided into 32 equal aero-
dynamic boxes, and the half-canard is divided into eight equal boxes. Aerodynamic
forces on the fuselage are neglected.
The right side of the Figure shows the structural idealization. Four weights are located
at the one-quarter and three-quarter span and chord positions of the wing and are
assumed to be connected to the 50% chord elastic axis by rigid streamwise bars. The
weights are 600 lb forward, and 400 lb aft, giving a wing centroid at 45% of the wing chord.
The wing is assumed to be uniform with equal bending ( 𝐸𝐼 𝑦 ) and torsion (𝐺 𝐽 ) stiff-
nesses of 25 · 107 lb·ft2 and is connected to the fuselage at its root. The right-side fuse-
lage is assumed to have the same bending stiffness as the wing and is shown with four
equal and equidistant weights (1500 lb each per side).
The fuselage length is 30.0 ft. The total weight per side is 8000 lb, the center of gravity
is 12.82 ft forward of the intersection of the fuselage and wing elastic axis, and the cen-
troidal moment of inertia in pitch per side is 𝐼 𝑦 = 892 900 lb·ft2 . For the subsonic case, the
airplane is assumed to be flying at a Mach number 𝑀 = 0.9 at sea level (𝑞 = 1200 lb/ ft2 ).
The low-speed characteristics (but at 𝑀 = 0.9) are obtained by assuming a low value
of dynamic pressure, 𝑞 = 40 lb/ft2 , to illustrate the behavior of the quasi-rigid vehicle.

4.2 Nastran input inizialization


Listing 4.1 provides the ID and CEND entries.

Listing 4.1
1 I D NXN , H A 1 4 4 A
2 TIME 5 $ CPU TIME IN MINUTES
3 SOL 144 $ STATIC AERO
4 $
5 $
6 $\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\
7 CEND
8 $///////////////////////////////////////////////////////////////////////////////
9 $
10 $
11 TITLE = EXAMPLE HA144A: 30 DEG FWD SWEPT WING WITH CANARD
12 S U B T I = S Y M M E T R I C F L I G H T CONDITIONS , DOUBLET - L A T T I C E A E R O
13 L A B E L = HALF - S P A N MODEL , S T A T I C S Y M M E T R I C L O A D I N G
14 ECHO= BOTH
15 SPC = 1 $ SYMMETRIC CONSTRAINTS
16 DISP = ALL $ PRINT ALL DISPLACEMENTS
17 STRESS = ALL $ PRINT ALL STRESSES
18 FORCE = ALL $ PRINT ALL FORCES
19 AEROF = ALL $ PRINT ALL AERODYNAMIC FORCES
20 APRES = ALL $ PRINT ALL AERODYNAMIC PRESSURES
21 $
22 SUBCASE 1

Piercarlo Mirto – Corso di Studi in Ingegneria Aeronautica, Università degli Studi di Napoli Federico II
4.3 Bulk data input 31

23 TRIM = 1 $ 1 G LEVEL FLIGHT (LOW SPEED)


24 SUBCASE 2
25 TRIM = 2 $ 1 G LEVEL FLIGHT (HIGH SUBSONIC SPEED)
26 OUTPUT(PLOT)
27 PLOTTER = NASTRAN
28 SET 1 = ALL
29 F I N D SCALE , O R I G I N 1 , S E T 1
30 PLOT SET 1
31 PLOT STATIC DEFORMATION 0 , ORIGIN 1 , SET 1 , OUTLINE

4.3 Bulk data input


The bulk data have been organized into structural and aerodynamical data, presented
in listings 4.2 (rows 39-126), 4.3 (rows 128-202) and 4.4 (rows 203-221) respectively.
This section will describe all the related inputs, indicating the code rows where the
reader can find them.

4.3.1 Structural model

Listing 4.2 Structural data input


39 $===============================================================================
40 $ * * * STRUCTURAL DATA * * *
41 $===============================================================================
42 $
43 $ -------------------------------------------------------------------------------
44 $ * * GRID GEOMETRY * *
45 $ -------------------------------------------------------------------------------
46 $
47 $ FUSELAGE GRID
48 $
49 GRID ,90 , ,15. ,0. ,0.
50 $
51 GRID ,97 , , 0. ,0. ,0.
52 GRID ,98 , ,10. ,0. ,0.
53 GRID ,99 , ,20. ,0. ,0.
54 GRID ,100 , ,30. ,0. ,0.
55 $
56 $ WING GRID
57 $
58 GRID ,110 , ,27.11325 ,5. ,0.
59 GRID ,120 , ,21.33975 ,15. ,0.
60 $
61 GRID ,111 , ,24.61325 ,5. ,0.
62 GRID ,112 , ,29.61325 ,5. ,0.
63 GRID ,121 , ,18.83975 ,15. ,0.
64 GRID ,122 , ,23.83975 ,15. ,0.
65 $ -------------------------------------------------------------------------------
66 $ * * STRUCTURAL STIFFNESS PROPERTIES * *
67 $ -------------------------------------------------------------------------------
68 $
69 $ FUSELAGE STRUCTURE
70 $
71 CBAR ,101 ,100 ,97 ,98 ,0. ,0. ,1.
72 CBAR ,102 ,100 ,98 ,90 ,0. ,0. ,1.
73 CBAR ,100 ,100 ,90 ,99 ,0. ,0. ,1.
74 CBAR ,103 ,100 ,99 ,100 ,0. ,0. ,1.

Aeroelasticity
32 Chapter 4 Divergence speed analysis

75 $
76 PBAR ,100 ,1 ,2.0 ,.173611 ,0.15 ,0.5 , , ,+P B 1
77 + PB1 ,1.0 ,1.0 ,1.0 , -1.0 , -1.0 ,1.0 , -1.0 , -1.0 ,+P B 2
78 + PB2 , , ,0.0
79 $
80 $ WING STRUCTURE
81 $
82 CBAR ,110 ,101 ,100 ,110 ,0. ,0. ,1.
83 CBAR ,120 ,101 ,110 ,120 ,0. ,0. ,1.
84 $
85 PBAR ,101 ,1 ,1.5 ,0.173611 ,2.0 ,0.462963 , , ,+P B 3
86 + PB3 ,0.5 ,3.0 ,0.5 , -3.0 , -0.5 ,3.0 , -0.5 , -3.0 ,+P B 4
87 + PB4 , , ,0.0
88 $
89 MAT1 ,1 ,1.44E +9 ,5.40E + 8
90 $
91 RBAR ,111 ,110 ,111 ,123456
92 RBAR ,112 ,110 ,112 ,123456
93 RBAR ,121 ,120 ,121 ,123456
94 RBAR ,122 ,120 ,122 ,123456
95 $ -------------------------------------------------------------------------------
96 $ * * MASS AND INERTIA PROPERTIES * *
97 $ -------------------------------------------------------------------------------
98 $
99 $ FUSELAGE MASSES
100 $
101 CONM2 ,97 ,97 ,0 ,1500.0
102 CONM2 ,98 ,98 ,0 ,1500.0
103 CONM2 ,99 ,99 ,0 ,1500.0
104 CONM2 ,100 ,100 ,0 ,1500.0
105 $
106 $ WING MASSES
107 $
108 CONM2 ,111 ,111 ,0 ,600.0
109 CONM2 ,112 ,112 ,0 ,400.0
110 CONM2 ,121 ,121 ,0 ,600.0
111 CONM2 ,122 ,122 ,0 ,400.0
112 $ -------------------------------------------------------------------------------
113 $ * * STRUCTURAL PARAMETERS * *
114 $ -------------------------------------------------------------------------------
115 PARAM , GRDPNT , 9 0
116 PARAM , WTMASS , . 0 3 1 0 8 1
117 PARAM , AUNITS , . 0 3 1 0 8 1
118 $ -------------------------------------------------------------------------------
119 $ * * STRUCTURAL CONSTRAINS * *
120 $ -------------------------------------------------------------------------------
121 SPC1 ,1 ,1246 ,90
122 SPC1 ,1 ,246 ,97 ,98 ,99 ,100
123 $
124 SUPORT ,90 ,35
125 $
126 OMIT1 ,4 ,110 ,120

Grid geometry

The first thing the code defines is of course the grid geometry, illustrated in Figure
4.2. The fuselage is made of five points, GRID 97, GRID 98, GRID 99, GRID 100 and GRID 90.
Specifically, all those grid points except GRID 90 will host a lumped mass of 1500 lb.
The wing’s elastic axis is made of GRID 100, GRID 110, and GRID 120. Grid points 111, 112,

Piercarlo Mirto – Corso di Studi in Ingegneria Aeronautica, Università degli Studi di Napoli Federico II
4.3 Bulk data input 33

121, and 122 will be home to concentrated masses.

Figure 4.2 Structural model overview.

Stiffness properties
Rows from 63 to 92 define the connection elements and their associated properties.
BAR elements are used between grid points. The material (row 89) is the same for all
of them, with 𝐸 = 1.44 · 109 lb/ft2 and 𝐺 = 5.40 · 108 lb/ft2 .
The wing stiffnesses were assumed to be equal in bending and torsion, 𝐸𝐼 𝑦 = 𝐺 𝐽 =
25.0 · 107 lb·ft2 ; that leads to 𝐼 𝑦 = 0.173611 ft4 and 𝐽 = 0.462963 ft4 , respectively. Values of
cross-sectional area, 𝐴 = 1.5 ft2 , and chordwise inertia, 𝐼 𝑧 = 20 ft4 , are chosen arbitrarily.
A nominal symmetrical rectangular cross-section with a 6.0 ft chord and 1.0 ft depth
is also assumed for the wing structural box for stress recovery purposes at the four
corners.
The half-fuselage material properties are assumed to be the same as in the wing with
the same vertical cross-sectional moment of inertia, 𝐼 𝑦 = 0.173611 ft4 . The remaining
fuselage cross-sectional area properties are selected arbitrarily for stiffness and stress
recovery, specifically, 𝐴 = 2.0 ft2 , 𝐼 𝑧 = 0.15 ft4 , 𝐽 = 0.5 ft4 , and the points selected for
stress recovery are at 𝑦, 𝑧 = ±1.0, ±1.0.
Grid points 111, 112, 121, and 122 are connected to the wing’s elastic axis by rigid bars.
The wing forward CONM2 weights are 600 lb, and the aft weights are 400 lb.

Structural parameters
PARAM entries select GRID 90 as the inertial property reference point and convert the input
weights to masses in slugs. PARAM,WTMASS,1/G provides the conversion of weight to mass;
PARAM,AUNITS,1/G allows for the input of the accelerations using load factors (Gs).

Structural contraints
There are two rigid body motions in this model: vertical translation and rotation in
pitch. A SUPORT Bulk Data entry defines a reference point for these rigid body modes on
GRID 90, DOFs 3 and 5. Component 4 (roll) of wing grid points 110 and 120 are omitted
from the calculation to illustrate this means of reducing the problem size and thus has

Aeroelasticity
34 Chapter 4 Divergence speed analysis

no effect on the results. GRID 90 is constrained longitudinally, and all of the fuselage
grid points are constrained for symmetry using SPC1 entries (rows 119-120).

4.3.2 Aerodynamic model

Listing 4.3 Aerodynamic data input


128 $===============================================================================
129 $ * * * AERODYNAMIC DATA * * *
130 $===============================================================================
131 $
132 $ -------------------------------------------------------------------------------
133 $ * * ELEMENT GEOMETRY * *
134 $ -------------------------------------------------------------------------------
135 AEROS ,1 ,100 ,10.0 ,40.0 ,200.0 ,1
136 $
137 CORD2R ,1 ,0 ,12.5 ,0. ,0. ,12.5 ,0. ,10. ,+C R D 1
138 + CRD1 ,20. ,0. ,0.
139 CORD2R ,100 ,0 ,15.0 ,0.0 ,0.0 ,15.0 ,0.0 , -10.0 ,+C R D 1 0 0
140 + CRD100 ,0.0 ,0.0 ,0.0
141 $ -------------------------------------------------------------------------------
142 $ * * SPLINE FIT ON THE LIFTING SURFACES * *
143 $ -------------------------------------------------------------------------------
144 $
145 $ BEAM SPLINE FIT ON THE WING
146 $
147 SPLINE2 ,1601 ,1100 ,1100 ,1131 ,1100 ,0. ,1. ,2 ,+S P W
148 + SPW , -1. , -1.
149 $
150 CAERO1 ,1100 ,1000 , ,8 ,4 , , ,1 ,+C A W
151 + CAW ,25. ,0. ,0. ,10. ,13.45299 ,20. ,0. ,10.
152 $
153 PAERO1 , 1 0 0 0
154 $
155 SET1 ,1100 ,99 ,100 ,111 ,112 ,121 ,122
156 $
157 $
158 $
159 CORD2R ,2 ,0 ,30. ,0. ,0. ,30. ,0. ,10. ,+C R D 2
160 + CRD2 ,38.66025 ,5.0 ,0.
161 $
162 $ CONTROL SURFACE DEFINITION
163 $
164 AESURF ,505 ,ELEV ,1 ,1000
165 AELIST ,1000 ,1000 ,THRU , 1 0 0 7
166 $
167 $ BEAM SPLINE FIT ON CANARD
168 $
169 SPLINE2 ,1501 ,1000 ,1000 ,1007 ,1000 ,0. ,1. ,1 ,+S P C
170 + SPC ,1. , -1.
171 $
172 CAERO1 ,1000 ,1000 , ,2 ,4 , , ,1 ,+C A C
173 + CAC ,10. ,0. ,0. ,10. ,10. ,5. ,0. ,10.
174 $
175 SET1 ,1000 ,98 ,99
176 $
177 $
178 $ -------------------------------------------------------------------------------
179 $ * * USER SUPPLIED INPUT DATA * *

Piercarlo Mirto – Corso di Studi in Ingegneria Aeronautica, Università degli Studi di Napoli Federico II
4.3 Bulk data input 35

180 $ -------------------------------------------------------------------------------
181 $
182 $ PRESSURE MODIFIERS (WEIGHTING MATRIX)
183 $
184 DMI , WKK ,0 ,3 ,1 ,0 , ,80 ,1
185 $^$
186 DMI , WKK ,1 ,1 ,1.0 ,THRU , 8 0
187 $
188 $ I N I T I A L D O W N W A S H E S ( E . G . , D U E T O INCIDENCE , T W I S T O R C A M B E R )
189 $
190 DMI , W2GJ ,0 ,2 ,1 ,0 , ,40 ,3
191 $^$
192 DMI , W2GJ ,1 ,9 ,0.0017453 ,THRU , 4 0
193 DMI , W2GJ ,2 ,9 ,0.0017453 ,THRU , 4 0
194 DMI , W2GJ ,3 ,9 ,0.0017453 ,THRU , 4 0
195 $
196 $ PRESSURES (E.G. , AT ZERO ANGLE OF ATTACK)
197 $
198 DMI , FA2J ,0 ,2 ,1 ,0 , ,40 ,3
199 $^$
200 DMI , FA2J ,1 ,1 ,0.0 ,THRU , 4 0
201 DMI , FA2J ,2 ,1 ,0.0 ,THRU , 4 0
202 DMI , FA2J ,3 ,1 ,0.0 ,THRU , 4 0

Element geometry
The reference geometry is specified on the AEROS entry. This entry specifies the aerody-
namic coordinate system CORD2R 1, the aerodynamic reference coordinate system for
rigid body motions CORD2R 100, a reference chord of REFC = 𝑐 = 10.0 ft, a reference span of
REFB = 𝑏 = 40.0 ft (the full span), a reference area of REFS = 𝑆 = 200.0 ft2 (half-model), and
symmetric aerodynamic loading (SYMXZ = 1).
CORD2R 100 provides the NACA reference axes for the stability derivatives. The trim
angle of attack is the angle of attack of the structural axis at the SUPORT point.

Spline fit on the lifting surfaces


The Doublet-Lattice method for surfaces is specified on the CAERO1 entries. CAERO1 1000
specifies the canard with a 2 × 4 division into boxes. CAERO1 1100 specifies the wing with
an 8 × 4 division into boxes. Figure 4.3 shows that input. The PAERO1 entry is required
even though the fuselage modeling is being neglected.
The SPLINE2 1501 (row 167) and SET1 1000 (row 173) entries specify a linear spline
on the canard to interconnect the structural and aerodynamic grids. The Cartesian
coordinate system CORD2R 1 is for the spline on the canard through the one-quarter
chord hinge line. The simplicity of the canard splining makes it uncharacteristic of
what can be expected in practice, but it also provides an illustrative example of the use
of the smoothing factors DTHX and DTHY.
It is seen that DTHX has been set to 1.0, implying that the canard is restrained in a roll
about the centerline by a rotational spring. If DTHX was set to 0.0, the spline would be
overdetermined since the two grid points that lie on the same line perpendicular to
the spline axis each have a value for the bending slope when only one is allowed. On
the other hand, if DTHX were set to -1.0, the splining would become singular because
displacements at two points are not sufficient to define a plane, i.e., rotation of the
canard about the centerline would not be precluded. Setting DTHY to -1.0 ensures that
only the transverse displacements of the two grid points are used in determining the

Aeroelasticity
36 Chapter 4 Divergence speed analysis

Figure 4.3 Aerodynamic elements overview.

displacement and rotation of the spline.


Next, the trim surface is defined at row 162 by an AESURF entry as the elevator (canard)
ELEV using coordinate system CORD2R 1 for its hinge line and defining the aerodynamic
boxes using AELIST 1000, which specifies aerodynamic box numbers 1000 through 1007.
The SPLINE2 1601 and SET1 1100 entries specify a linear spline on the wing, Figure 4.4.
SET1 1100 includes GRIDs 99 and 100 for a good spline fit for the wing in the root region.
CORD2R 2 is the Cartesian coordinate system through the wing elastic axis.

Figure 4.4 Spline overview.

CORD2R 100 is for the rigid body motions of the aerodynamic reference point; specifi-
cally, the pitch and moment axis is at the canard mid chord at GRID 90. This coordinate
system is the standard NACA body axis system with the x-axis forward and the z-axis
downward. The stability derivatives are output using this coordinate system.

Piercarlo Mirto – Corso di Studi in Ingegneria Aeronautica, Università degli Studi di Napoli Federico II
4.3 Bulk data input 37

User supplied data

Additional aerodynamic data are included in DMI entries to account for the differences
between test and theory1 , experimental pressure data at some reference condition,
e.g., zero angle of attack2 , and any initial downwash distribution arising, e.g., from
incidence, camber, or twist3 . In the cited example, 𝑊𝑘𝑘 = 1.0 and 𝑓 𝑒 = FA2J = 0.0 for all
of the wing and canard aerodynamic boxes, and 𝑤𝑔𝑗 = 2WGJ = 0.1 deg = 0.001745 rad for the
wing boxes and W2GJ = 0.0 for the canard boxes.

4.3.3 Static Aeroelastic Input

Listing 4.4 Aerodynamic data input


203 $===============================================================================
204 $ * * * SOLUTION SPECIFICATIONS * * *
205 $===============================================================================
206 $
207 $ -------------------------------------------------------------------------------
208 $ * * AERODYNAMIC DOFS * *
209 $ -------------------------------------------------------------------------------
210 AESTAT ,501 ,A N G L E A
211 AESTAT ,502 ,P I T C H
212 AESTAT ,503 ,U R D D 3
213 AESTAT ,504 ,U R D D 5
214 $
215 $ -------------------------------------------------------------------------------
216 $ * * TRIM CONDITIONS * *
217 $ -------------------------------------------------------------------------------
218 TRIM ,1 ,0.9 ,40.0 ,PITCH ,0.0 ,URDD3 , -1.0 , ,+T R 1
219 + TR1 , URDD5 , 0 . 0
220 TRIM ,2 ,0.9 ,1200.0 ,PITCH ,0.0 ,URDD3 , -1.0 , ,+T R 2
221 + TR2 , URDD5 , 0 . 0

Aerodynamic DOFs

The foregoing input is typical for any aeroelastic analysis. The entries in the Bulk Data
specifically for static aeroelastic analysis begin with the AESTAT entries, which specify
the trim parameters. The parameters are angle of attack, 𝛼 = ANGLEA; pitch rate, 2𝑞𝑐𝑉 =
¥ /𝑔 = URDD5.
PITCH; normal load factor 𝑧¥ /𝑔 = URDD3; and pitch acceleration, 𝜃

Trim conditions

The two TRIM entries specify the flight condition at Mach number, 𝑀 = 0.9 and level
flight with no pitch rate, 2𝑞𝑐𝑉 = PITCH = 0, a one-g load factor, 𝑧¥/𝑔 = URDD3 = −1.0, and no
pitching acceleration, 𝜃¥ /𝑔 = URDD5 = 0.0.
The first entry, TRIM 1, specifies the low-speed condition with dynamic pressure 𝑞 =
𝑄 = 40 lb/ft2 , and the second entry, TRIM 2, specifies 𝑞 = 𝑄 = 1200 lb/ft2 , both at 𝑀 = 0.9
at sea level.
1 see for correction factors 𝑊 𝑘 of Eq. 1-21 in Ref. [7]
𝑘
2 the additive coefficients 𝑓 𝑒 also of Eq. 1-21 in Ref. [7]
3 the additional downwash 𝑤𝑔 of Eq. 1-2 in Ref. [7]
𝑗

Aeroelasticity
38 Chapter 4 Divergence speed analysis

4.4 Femap input


Figures 4.5a and 4.5b show the complete Analysis Model, highlighting the element
thicknesses.

(a)

(b)

Figure 4.5 Three-dimensional views of the FEM model with element thickness highlighted.

4.5 Output
Figure 4.6 outlines the rotation distributions for Case 2. Figure 4.7 does the same for
the displacement distribution.

Piercarlo Mirto – Corso di Studi in Ingegneria Aeronautica, Università degli Studi di Napoli Federico II
4.5 Output 39

(a) Around the 𝑥 -axis

(b) Around the 𝑦 -axis

(c) Around the 𝑧 -axis

Figure 4.6 Rotation distribution.

Aeroelasticity
40 Chapter 4 Divergence speed analysis

(a) Along the 𝑥 -axis

(b) Along the 𝑦 -axis

(c) Along the 𝑧-axis

Figure 4.7 Tranlation distribution.

Piercarlo Mirto – Corso di Studi in Ingegneria Aeronautica, Università degli Studi di Napoli Federico II
BIBLIOGRAPHY

Bibliography

References
[1] G. M. Ashawesh, “Flutter behaviour of composite aircraft wings,” Ph.D. disserta-
tion, CRANFIELD UNIVERSITY, CRANFIELD, BEDFORD, UK, 1999. [Online]. Avail-
able: https://www.cond-mat.de/events/correl11/manuscript/Koch.pdf.
[2] M. H. Shirk, T. J. Hertz, and T. A. Weisshaar, “Aeroelastic tailoring - Theory, prac-
tice, and promise,” JOURNAL OF AIRCRAFT, vol. 23, no. 1, pp. 6–18, 1986. [Online].
Available: https://arc.aiaa.org/doi/pdf/10.2514/3.45260.
[3] Dark_Skies. “X-29 - The Most Unstable Fighter Jet Ever Built,” YOUTUBE. (2021),
[Online]. Available: https://youtu.be/9GEB-7Gw-_0.
[4] R. Rosenbaum, “Simplified Flutter Prevention Criteria for Personal Type Aircraft,”
FEDERAL AVIATION ADMINISTRATION, WASHINGTON DC, Tech. Rep. ADA955270,
Jan. 1955, Distribution Statement: APPROVED FOR PUBLIC RELEASE. [Online].
Available: https://apps.dtic.mil/sti/citations/ADA955270.
[5] F. Marulo, “Static aeroelastic phenomena,” Teaching material from the course
"Aeroelasticity", 2023.
[6] R. L. Bisplinghoff, H. Ashley, and R. L. Halfman, Aeroelasticity.: ADDISON-WESLEY,
1955, pp. 474–489.
[7] T. A. Weisshaar, “Forward Swept Wing Static Aeroelasticity,” VIRGINIA POLYTECH-
NIC INSTITUTE, Tech. Rep. ADB042815, Apr. 1979, Distribution Statement: AP-
PROVED FOR PUBLIC RELEASE. [Online]. Available: https : / / apps . dtic . mil / sti /
citations/ADB042815.

[10] F. Marulo, “Splines - Interfacing the structure with aerodynamics,” Teaching ma-
terial from the course "Aeroelasticity", 2023.
[11] P. Mirto, Finite Element Analysis of a metal Wing-Box, Published via Dropbox:
https://www.dropbox.com/s/2azdqtdsbxc8q8s/Finite_Element_Analysis_of_a_metal_wingbox.pdf?dl=
0, Jan. 2023.
[12] M. Blair, “A COMPILATION OF THE MATHEMATICS LEADING TO THE DOU-
BLET LATTICE METHOD,” WRIGHT LABORATORY, Tech. Rep., Mar. 1992, Distri-
bution Statement: APPROVED FOR PUBLIC RELEASE. [Online]. Available: https:
//apps.dtic.mil/sti/citations/ADA256304.

BIBLIOGRAPHY
INSIGHTS42 Chapter 4 Divergence speed analysis

[13] W. P. Rodden and J. R. Love, “Equations of Motion of a Quasisteady Flight Vehicle


Utilizing Restrained Static Aeroelastic Characteristics,” JOURNAL OF AIRCRAFT,
vol. 22, no. 9, pp. 802–809, 1985. [Online]. Available: https://arc.aiaa.org/doi/10.2514/
3.45205.

[14] Aeroelastic Analysis User’s Guide, SIEMENS, GRANITE PARK ONE, USA, 2014. [On-
line]. Available: https://apps.dtic.mil/sti/tr/pdf/ADB042815.pdf.
[15] P. Mirto, Consistent Physical Units for structural engineering, May 2023. [Online].
Available: https : / / www . dropbox . com / s / 9vxb6ah0zsugxwj / Consistent _ Physical _ Units _ for _
structural_engineering.pdf?dl=0.

Insights
[8] F. W. Diederich and K. A. Foss, “Charts and Approximate Formulas for the Estima-
tion of Aeroelastic Effects on the Loading of Swept and Unswept Wings,” NATIONAL
ADVISORY COMMITTEE FOR AERONAUTICS, LANGLEY FIELD, UNITED STATES, Tech.
Rep. 93R20225, Sep. 1951, Distribution Statement: APPROVED FOR PUBLIC RE-
LEASE. [Online]. Available: https://ntrs.nasa.gov/citations/19930090935.
[9] E. H. Johnson, “MSC Developments in Aeroelasticity,” THE MACNEAL-SCHWENDLER
CORPORATION, LOS ANGELES, CALIFORNIA 90041, Tech. Rep., Oct. 2011, [On-
line]. Available: https : / / www . researchgate . net / publication / 242155910 _ MSC _ Developments _
in_Aeroelasticity.

INSIGHTS Piercarlo Mirto – Corso di Studi in Ingegneria Aeronautica, Università degli Studi di Napoli Federico II
Updated on: July 18, 2023

You might also like