A Review of Mathematical Modelling Techniques For Advanced Rotorcraft Configurations

Download as pdf or txt
Download as pdf or txt
You are on page 1of 74

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/346577628

A review of mathematical modelling techniques for advanced rotorcraft


configurations

Article in Progress in Aerospace Sciences · December 2020


DOI: 10.1016/j.paerosci.2020.100681

CITATIONS READS

10 423

3 authors, including:

Renliang Chen Ye Yuan


Nanjing University of Aeronautics & Astronautics Swansea University
40 PUBLICATIONS 286 CITATIONS 27 PUBLICATIONS 121 CITATIONS

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Helicopter View project

Helicopter flight dynamic View project

All content following this page was uploaded by Ye Yuan on 03 December 2020.

The user has requested enhancement of the downloaded file.


1

A Review of Mathematical Modelling Techniques for

Advanced Rotorcraft Configurations

Renliang Chen1, Ye Yuan1, 2, and Douglas Thomson2

1. National Key Laboratory of Rotorcraft Aeromechanics,

Nanjing University of Aeronautics & Astronautics, Nanjing, China

2. Aerospace Sciences Research Division, James Watt School of Engineering,

University of Glasgow, Glasgow, United Kingdom

Abstract: The paper will review the development and application of the mathematical

modelling of the advanced rotorcraft configuration, including compound helicopter

configurations and tilt-rotor vehicles. The mathematical model is the basis for the design

of the flight control system and an essential tool to assess the flying and handling qualities

for helicopters. As the helicopter is a multi-body system, the mathematical modelling of

helicopter should consider the coupling effects among motion, inertia, structure, and

aerodynamics, as well as the unsteady and nonlinear characteristics, to give the physical

principles and mathematical expression of each part. Therefore, the mathematical

modelling of a helicopter is a process of analysing and synthesizing different hypotheses

and subsystem models. Moreover, the advanced helicopter configuration puts forward

higher requirements for the helicopter mathematical modelling in terms of the

aerodynamic interference, blade motion characteristics, and manoeuvre assessment. The

critical issues of helicopter modelling, especially the modelling of the advanced rotorcraft

configurations, will be illustrated in this paper. The emphasis is put on the modelling of

rotor aerodynamics and aerodynamic interaction among the rotor, fuselage, and other parts.

Integrated modelling methods and the manoeuvrability investigation are also the foci of

the paper. Suggestions for future research on helicopter flight dynamics modelling are also

provided.

Keywords: Advanced rotorcraft; Flight dynamics; Wake model; Aerodynamic interference;

Engine/fuel control system; Manoeuvrability analysis


2

Table of Contents

Nomenclature ...................................................................................................................................... 3

1. Introduction – Challenges and Requirements for Rotorcraft Mathematical Modelling ........... 4

2. Overview of helicopter flight dynamics modelling ....................................................................... 8

3. Development in the helicopter flight dynamics modelling ......................................................... 16

3.1 Rotor aerodynamic modelling ................................................................................................ 17

3.1.1 Airfoil aerodynamic model ................................................................................................ 17

3.1.2 Rotor wake model .............................................................................................................. 20

3.1.3 Blade dynamics model ....................................................................................................... 25

3.2 Model of aerodynamic interference among components ...................................................... 30

3.3 Engine/fuel control system modelling .................................................................................... 37

3.4 The integration and calculation of the non-linear systems................................................... 38

3.5 Progress in manoeuvrability analysis .................................................................................... 40

4. Conclusion ..................................................................................................................................... 43

References.......................................................................................................................................... 47
3

Nomenclature

A1C Longitudinal cyclic pitch (Deg)

B1C Lateral cyclic pitch (Deg)

Damping and stiffness matrix

Kre Wake curve parameter

Kβ Equivalent flapping spring stiffness (kg.m2/(rad.s2))

Mβ Static moment of the blade mass (kg.m)

Iβ Inertia moment of the blade (kg.m2)

R Rotor radius (m)

Wtip Flapping amplitude of the original flapping motion (m)

W`0.75 Flapping angle at 0.75R (rad)

V Forward speed (m/s)

a0, a1, a2 Coning angle, lateral and longitudinal flapping angle (rad)

e Flapping hinge offset (m)

Non-dimensional flapping hinge offset

External excitation vector

k Downwash and side-wash factor

q Pitching angular velocity (rad/s)

Dynamic pressure ratio

t Time (s)

vh Induced velocity at hover state (m/s)

vi Induced velocity (m/s)

Ω Rotor rotation speed (rad/s)

β Flapping angle (rad)

ζ Downwash or side-wash effect parameter

Non-dimensional frequency of the first-order flapping


4

1. Introduction – Challenges and Requirements for Rotorcraft Mathematical

Modelling

The aim of helicopter flight dynamics modelling is to construct a correlation among the

helicopter motion, the external forces (moments), and the controllers based on the physical

laws associated with aerodynamic theory and structural dynamics results. The helicopter flight

dynamics model is not only the basis of its control system design but also the primary

measurement to develop and analyse the handling qualities feature of the helicopter. The U.S.

military handling qualities requirement for rotorcraft (ADS-33F-PRF) is explicitly stipulated

that any new rotorcraft should examine the handling qualities with the flight dynamics model
[1]
in each developing phase . With recent advances in the helicopter industry, a range of

advanced rotorcraft configurations have been developed, and the primary types are shown in

Figure 1, namely, the tilt-rotor aircraft, coaxial compound helicopter, and hybrid compound

helicopter. These configurations have the capability to further improve the manoeuvrability

and performance characteristics, such as the maximum flight speed, flight range, and flight

duration. Also, the advanced rotorcraft configurations put forward a higher requirement of the

helicopter flight dynamics modelling technique in order to meet the extension of the flight

range and the manoeuvrability.

(a) Tiltrotor Aircraft (b) Coaxial Compound (c) Hybrid Compound


V22 Osprey Helicopter Helicopter X3 helicopter
SB-1 DEFIANT
Figure 1 The primary advanced rotorcraft configurations

More consideration should be taken in modelling flight dynamics characteristics of

helicopters. Considering the conventional single-rotor helicopter, the rotor system has to
5

provide all the force and moments that the helicopter needs, except the yawing moment is
[2]
provided by the tail rotor . This implies that the lift, the control forces, and the propulsive

force are coupled with each other. Furthermore, it should be mentioned that in most of the

advanced rotorcraft configurations, such as the tiltrotor aircraft and the hybrid compound

helicopter, the yawing moment is provided by a multi-rotor system. Meanwhile, the advanced

rotorcraft is usually equipped with an auxiliary propulsion system or tilts the rotor disc to

provide the propulsive force in high-speed flight, and the elevator and rudder may also be

involved to compensate control power in high-speed flight. There are therefore strong

interactions between the rotor system and other components of the helicopter.

Furthermore, the rotor is composed of flexibility blades with high aspect ratio, and it utilises
[3]
the blade rotational motion to produce lift force . As a result, there are two outstanding

features of the rotor aerodynamics phenomena. Firstly, a blade element may suffer from non-

linear flow phenomena, such as separated flow and shock waves. The resulting stall condition

at the advancing tip brings about great difficulty in performing useful aerodynamic analysis.

Moreover, due to the development of advanced rotorcraft with increased flight speed there is

an increased likelihood of the blade experiencing this stall condition. Secondly, the wake vortex

at the blade trailing edge is rapidly rolled up near the rotor tips and formulates the rotor wake

led by the tip vortex. In hover and low speed forward flight, the tip vortex is trapped near the

rotor disc because of the low flow velocity, leading to severe geometric distortion in the wake.

This induces a significant non-uniform component of the inflow on the rotor disc and

consequently alters the aerodynamic load and motion of the blade, which further influences the

trim characteristics, stability, and control features of the helicopter.

On the other hand, the change of the blade aerodynamic load distribution and motion would,

in turn, affect the strength and geometry of the rotor wake vortex. The interaction among rotor

wake, blade motion, and blade aerodynamic load forms a dynamic system with high-level
6

coupling. Also, the overall motion of the helicopter and the Coriolis force on the blade change

the features of the aerodynamic coupling phenomenon in manoeuvring flight, and this effect

will be more significant for the advanced rotorcraft configurations that equipped with the multi-

rotor system. In multi-rotor systems each rotor can influence the tip vortex motion and the non-

linearity in the associated rotor disc inflow of the other rotors, and therefore alters the

aerodynamic load distribution and the vehicle motion.

The aerodynamic interaction between the rotor system and other parts in the helicopter is

another critical issue that should be considered in the helicopter flight dynamics modelling

process, especially for advanced rotorcraft configurations. As well as the interaction between

the rotor system and the horizontal tails, vertical fins, or fuselage, the unique interference for

advanced rotorcraft should also be taken into consideration, including the aerodynamic

interference between the rotor system and the auxiliary propeller and wing. This aerodynamic

interaction has unique characteristics, determined by the rotor flow field and the helicopter

configuration. Firstly, the rotor flow field contains various types of aerodynamic features, such

as the non-stationary, non-linearity, and a three-dimension effect. These features include the

tip vortex structure in the flow field, the vortex-blade interference, the dynamic separation, the

periodic transonic motion, and the revolution of the trailing vortex and trapped vortex.

Secondly, the rotor system creates a rotational flow, and the multi-rotor system even makes the

rotor flow field further distorted. Effectively, aerodynamic interaction could change the rotor

performance and the pertinent forces and moments provided by other parts of the helicopter to

a large extent. Wind tunnel experiments [4] have demonstrated that the interaction between the

rotor system and wing could occupy around 25 % of the overall vertical force when the tiltrotor

aircraft is in hover state, for example.

Rotor dynamics characteristics are significant aspects that should be carefully considered

during the flight dynamics modelling process. The rotor dynamics feature largely influences
7

the flapping, lagging, and torsion motions of the rotor blade, and these blade motions determine

the direction of the rotor force and drag, which plays a significant role in determining the flight

dynamics characteristics. In addition, advanced rotorcraft has unique rotor dynamics features.

The tiltrotor aircraft usually equips with a gimbal rotor system [5], and the rigid rotor is widely

utilised in the coaxial compound helicopter configurations. Furthermore, the rotor dynamics

have a direct correlation with the rotor rotational speed, and a variable rotor speed strategy is

usually involved in the advanced rotorcraft to reduce the compressibility effect at the advancing

blade tip at higher flight speeds. Meanwhile, the variable rotor speed feature and the flight

dynamics characteristics in large amplitude manoeuvres are coupled with the helicopter

turboshaft engine. Thus, with the aim to enhance the accuracy of the flight dynamics modelling,

the rotor dynamics and turboshaft engine characteristics should also be taken into account.

The flight dynamics mathematical modelling primarily includes the aerodynamic modelling

of rotor, fuselage, horizontal tail, vertical tails, and the potential propulsion device, as well as

the coupled rotor/engine/fuel control system dynamics model. The coupling dynamics and

kinematic model are also needed to capture detailed dynamics features in manoeuvring flight
[2]
. Considering the structure dynamics, aerodynamics, and the multidisciplinary nature of the

interactions present, the mathematical modelling applied should be carefully selected and

improved. Meanwhile, the flight dynamics model should be transformed and generalised based

on the actual situation using a series of integrating methods.

This article discusses the core issues of the flight dynamics modelling of the advanced

helicopter. The foci contain the rotor aerodynamic modelling, the aerodynamic interference

modelling, the coupling dynamics modelling of rotor/engine/fuel control, and the helicopter

flight dynamic modelling integration. Meanwhile, the manoeuvrability analysis methods for

the advanced helicopter will also be discussed in this article. This article puts forward the
8

current status and trends of the flight dynamics modelling development for the advanced

helicopter industry.

2. Overview of helicopter flight dynamics modelling

The helicopter flight dynamics mathematical model has been developed from a simple 6

degree-of-freedom (DOFs) rigid-body model to multi-DOFs model. Also, research objectives

have been extended from steady flight investigation to large amplitude manoeuvring flight

analysis.

The 6 DOFs rigid body flight dynamics model for helicopter derives from the modelling

method of the fixed-wing aircraft. This flight dynamics model is obtained by linearization of

the 6 DOFs rigid body model. This method is convenient for researchers to get the essential

control and stability characteristics of helicopters, and it can be used for the initial design of

the helicopter control system [5-10]. The theory and practice process proves that the linear model

of the flight dynamics is only suitable for the helicopter with lower manoeuvrability

requirements as it fails to take the non-linear effect of the flight dynamics into consideration.

The helicopter flight dynamics characteristics have distinctive non-linearity due to the

aerodynamic coupling among rotor, fuselage, tailplane, and the potential auxiliary propulsion

device. Also, the vehicle’s motion, structural dynamics, and inertia should be considered in
[11-15]
investigating the flight dynamics modelling process . Many researchers have focused on

the modelling of the non-linear characteristics of the helicopter, and a range of helicopter

mathematical models have been developed for theoretical analysis [16-17], numerical simulation
[19-23]
, and real-time simulation [24-28]. Other authors have also contributed to the body of work

on helicopter mathematical models[29-32].

There are two different kinds of helicopter flight dynamics non-linear mathematical model.

The first one is to describe the non-linearity using the differential equations of helicopter

motion. The linearity and non-linearity of the sub-system, such as the rotor model and the
9

dynamics model of other parts, are implicitly expressed in this method. This modelling method

provides a relatively accurate and reliable foundation for the helicopter flight control system

design. The typical example of this model is the ARMCOP model [27, 31]. This model is a low-

order model, which adopts a static inflow model and utilises the blade element theory to

calculate the forces and moments of the rotor system in a periodic average form. In order words,

this method has the feature of time-efficiency and is appropriate for the initial design and

analysis procedure [19-21].

The second type of modelling technique contains not only non-linearity in the helicopter

motion but also includes the non-linearity in every sub-system. The modelling method is

widely used for helicopter flying simulation. The GENHEL simulation package [23] developed

by Sikorsky helicopter company is one of the typical examples using this modelling method. It

still assumes that the fuselage is a rigid body. However, except for the six-rigid-body DOFs of

helicopter motion, this model also has the DOFs of rotor motion, including the flapping,

lagging, and torsion motions, as well as the DOFs of rotor rotational motion. Meanwhile, this

model utilises empirical equations to fit the torsion motion of the blade. Its rotor aerodynamic

model applies the combination of static non-uniform inflow model and the blade element

theory in order to calculate the blade aerodynamics. The rotor downwash on other parts of the

helicopter and other aerodynamic interference effects are also included based on the

experimental and theoretical analysis results. The accuracy of this modelling technique is well

understood, and the method has been widely used for various studies, such as the ground
[24] [25]
numerical simulation , non-linear equation parallel processing investigation , and high-
[26]
order linear model simplification . It is worthy of mention that the GENHEL model also

includes a representation of the engine/fuel control system so that a variable rotor speed

strategy can be simulated with higher accuracy, which is vital for the flight dynamics analysis

of more advanced rotorcraft configurations. According to the relevant flight test results [23, 26],
10

GENHEL model has relatively high precision in mid-speed forward flight range, but its

accuracy in hover and high-speed ranges requires a further improvement due to the lack of the

study on the rotor flow field and its aerodynamic interference on the sub-components at this

flight range.

The University of Maryland developed an improved flight dynamics simulation code,


[33]
HeliUM 2, to overcome the deficiencies in the GENHEL code . It includes flexible rotors

and free-vortex wake models to improve its accuracy across the flight range, and it has been

expanded to include flexible wings and multi-rotor calculation capabilities. This model allows

flight dynamics investigation to be executed for various rotorcraft configurations, from single

rotor helicopter to compound helicopters and tilt-rotor aircraft. Figure 2 indicates the roll rate

response comparison in hover and cruise flight state of XV-15 tilt-rotor aircraft.

(a) Hover state


11

(b) Cruise state


Figure 2 XV-15 cruise pitch rate response comparison [33]
where HeliUM curves represent the results obtained from HeliUM 2 code; the curve marked

“ID Model” comes from a state-space model derived from “Flight Data” using system

identification method; the “GTRSIM” represents the results from a state-space model derived

from the GTRSIM code, which is constructed based on the wind tunnel experiments of the XV-

15 tiltrotor aircraft [34]. The cruise state means that the tilt-rotor aircraft is flying in an aeroplane

mode with the forward speed of 180 knots, and the hover state indicates the tilt-rotor aircraft is

in hover state with the helicopter mode. According to Figure 2, the HeliUM 2 curves follow

the ID Model curves with good agreement, suggesting good accuracy of the HeliUM 2

simulation code for the flight dynamics modelling of tilt-rotor aircraft.

In order to validate the accuracy of the HeliUM 2 model in dealing with the coaxial

compound helicopter, the trim results of X2TD helicopter (x2 technology demonstrator, a

coaxial compound helicopter) in different forward speeds are shown in Figure 3. The results

derived from the GENHEL code are also provided in Figure 3, where A1C and B1C represent
12

the longitudinal and lateral cyclic pitches, respectively. It should be mentioned that due to

differences in the propeller modelling method, HeliUM 2 data are not presented in the propeller

collective results. Based on the results, the HeliUM 2 software has better accuracy in the

longitudinal control comparison, indicating the increased precision in the aerodynamic

interference calculation between the rotor wake and other parts. However, the accuracy of the

collective pitch is lower than the GENHEL model. This is because the aerodynamic

interference inside the coaxial rotor system is significant so that the rotor wake calculation

method used in HeliUM 2 (free wake model) may lead to additional errors.
12
Flight
9 GenHel
HeliUM
6

3
Upper Rotor
A1C (deg)

-3

-6

-9

-12
-50 0 50 100 150 200 250 300
Airspeed (ktas)
(a) Upper rotor A1C
12
Flight
9 GenHel
HeliUM
6

3
Upper Rotor
B1C (deg)

-3

-6

-9

-12
-60 0 60 120 180 240 300
Airspeed (ktas)
(b) Upper rotor B1C
13

12
Flight
9 GenHel
HeliUM
6

Lower Rotor
A1C (deg)
0

-3

-6

-9

-12
-60 0 60 120 180 240 300
Airspeed (ktas)
(c) Lower rotor A1C
12
Flight
9 GenHel
HeliUM
6

3
Lower Rotor
B1C (deg)

-3

-6

-9

-12
-60 0 60 120 180 240 300
Airspeed (ktas)
(d) Lower rotor B1C
50
Flight
GenHel
40
Collective Pitch at 75% R(deg)

30
Propeller

20

10

-10
-60 0 60 120 180 240 300
Airspeed (ktas)
(e) Propeller collective pitch at 75% R
14

100
Flight
90
GenHel
80 HeliUM

70

Stick Position (%)


60

Collective
50

40

30

20

10

0
-60 0 60 120 180 240 300
Airspeed (ktas)
(f) Collective stick position
Figure 3 Trim comparison of X2TD coaxial compound helicopter [35]
It is possible to tune the flight dynamics model using empirical factors to improve the

accuracy in steady and small amplitude manoeuvring flight. However, the aerodynamic

characteristics are significant in large amplitude manoeuvres, especially for the advanced

rotorcraft due to its multi-rotor system and potential auxiliary propulsion device. Using

empirical factors can be difficult to establish an accurate model in such scenarios. Ferguson [36,
37] [38-40]
and Yuan have investigated the manoeuvrability characteristics of the compound

helicopter, but the accuracy of the simulation results is still questionable due to the lack of the

relevant flight test data. The verified simulation research on the large amplitude manoeuvre is

mainly related to the conventional helicopter. The off-axis response simulation results of UH-

60 helicopter from different researchers are shown in Figure 4. In these figures, t is the time,

and q represents the pitching angular velocity.


15

(a) Comparison with Zhao and Curtiss’s model, GENHEL’s result with AEFRA flight test data

(b) Comparison of Takahashi model with AEFA flight test data

Figure 4 Large amplitude manoeuvre comparison with different researchers


According to these results, a significant error can be found in different modelling methods,

indicating that the modelling technique for the large-amplitude manoeuvre still does not

replicate the real vehicle. This inaccuracy exists in the different types of helicopter flight

dynamics models. The comparison results for BO-105 and AH-64 helicopters also show similar

phenomena in large amplitude manoeuvring flight [43, 44].

In order to improve the precision of the helicopter flight dynamics model, researchers have

investigated this coupling effect with different aspects, including the rotor aerodynamic

characteristics, the blade dynamics, the unsteady rotor wake feature, and the 2-D airfoil
[45-55]
unsteady aerodynamics characteristics . In their rotor blade aerodynamic models, the
16

influence of airfoil aerodynamic unsteady and dynamic stall characteristics are both involved.

The blade dynamics model can include elastic blade representations based on the finite element

method (FEM), and an improved rotor free wake model can also be used in the helicopter flight

dynamic modelling. The relevant results demonstrate that a significant improvement can be

seen when these methods are combined [48, 54].

With the development of the advanced rotorcraft industry, the importance of developing a
[56, 57]
universal manoeuvrability assessment method is growing , particularly for use in

assessing vehicle subjective handling qualities features and to improve its flight dynamics

characteristics. Researchers have done a great deal of work, trying to apply different types of

flight dynamics models for manoeuvrability investigation. In theoretical research, initially

energy methods were the primary approach to study the helicopter manoeuvring. This method

is based on the theory of energy conservation to calculate the trajectory of helicopter

manoeuvring flight. When considering real conditions and specific circumstances, the accuracy
[58-62]
of the energy method is relatively low . In the late 1980s, the inverse simulation method

was developed to analyse the manoeuvrability of the helicopter [63-65]. The inverse simulation

method utilises the mathematical description to give the flight trajectory and then obtains the

control input during the manoeuvre by inversely calculating the flight dynamics equations.

Although this method is successful in the research of helicopter manoeuvring flight and can

acquire meaningful results, the inverse simulation method still needs improvement in practical

application due to the increasing complexity of the helicopter flight dynamics model. Also, the

non-linear optimisation method and other manoeuvrability analysis methods have their

inefficiency in practice. Therefore, how to investigate the helicopter manoeuvring flight

characteristics is still a key issue in the research of the helicopter flight dynamics.

3. Development in the helicopter flight dynamics modelling


17

Helicopter flight dynamics mathematical modelling involves the rotor aerodynamic

modelling, helicopter aerodynamic interference modelling, the engine/fuel control system

modelling, and the manoeuvring flight analysis method. This section would illustrate the status

and development of the above modelling and analysis methods.

3.1 Rotor aerodynamic modelling

The rotor provides most of the lift, the control power, and the propulsion that the helicopter

needs all of which are dependent on aerodynamic characteristics. Meanwhile, the advanced

rotorcraft usually adopts the auxiliary propeller to provide the propulsion in the high-speed

flight, and its aerodynamic modelling process is similar to the rotor modelling methods.

In modelling the helicopter flight dynamics, the critical aspects of rotor aerodynamic

modelling can be concluded in three aspects, which are the airfoil aerodynamic model, the rotor

wake model, and the blade dynamics model. These three aspects have mutual effects on each

other.

3.1.1 Airfoil aerodynamic model

The airfoil aerodynamic model is the basis of the rotor aerodynamics model, determining

the lift and drag of each blade element (airfoil) in different flight ranges. The most simplified

airfoil aerodynamic modelling method is to use the 2D lift-curve slope theory to calculate the

lift coefficients at the corresponding angle of attack, in conjunction with an empirical equation

to calculate the airfoil drag coefficients. This method has been widely utilised to the initial

flight dynamics estimation and performance calculation. Its weakness is that it fails to consider

unique features of the helicopter airfoil aerodynamics, such as reverse flow and dynamic stall
[66-73]
characteristics . Therefore, the accuracy of this method is relatively low and cannot be

utilised to investigate the flight dynamics characteristics of the advanced helicopter

comprehensively.
18

The airfoil aerodynamic characteristics of the helicopter are quite different from the fixed-

wing aircraft due to phenomena of dynamic stall, compressibility effects, reverse flow, and the

radial flow effect. These features are summarised in Figure 5. The main challenge to improve

the accuracy of the airfoil aerodynamic model is to determine the appropriate modelling

method to simulate these features.

Figure 5 The features of the airfoil aerodynamic characteristics


Firstly, the dynamic stall phenomenon is due to the leading-edge vortex, which provides

additional suction over the upper airfoil surface as it converts downstream. This increased

suction leads to performance gains in lift and stall delay, but it quickly becomes unstable and

detaches from the airfoil, leading the lift to decrease rapidly [74-76]. The helicopter rotor usually

occurs this phenomenon at the blade tip of the retreating side, and it is more significant in the

advanced rotorcraft as the increase of the forward speed exaggerates the flapping motion. It is

hard to calculate the dynamic stall effect on a blade section even with a detailed CFD method
[77-79]
, and therefore, a series of empirical or semi-empirical methods were developed based on

experimental results, including an unsteady/dynamic stall aerodynamic model constructed by


[80-81]
Leishman and Beddoes . The calculation process is composed of three parts, the bound

flow aerodynamics calculation, the trailing separation estimation, and the dynamic stall

(advancing separation) aerodynamics calculation. In the bound flow aerodynamics calculation,


19

an exponential response method is adopted to calculate the airfoil aerodynamics. The pressure

on the leading edge is used to estimate airflow separation occurrence based on the correlation

between the vertical force coefficient and the critical vertical force coefficient of the airflow

separation. The critical coefficient is a function of Mach number, which can be empirically

determined according to steady airfoil aerodynamic test data. Also, The French Aerospace Lab

(ONERA) developed a 2-D dynamic stall model for airfoil aerodynamics model based on the
[83-85]
Hopf bifurcation . The lift and moment coefficients are calculated on the basis of non-

linear ordinary differential equations, and the relevant parameters in those equations are also

based on wind tunnel experiments.

The compressibility effect and reverse flow also influence the airfoil aerodynamics to a large

extent. As forward speed increases, the local Mach number at the blade tip of the advancing

side is close to the local speed of sound, reducing the aerodynamic efficiency and sharply

increasing the drag. Meanwhile, the area of the reverse flow spreads with forward speed. The

angle of attack in the reverse flow area is much higher than the stalling incidence, which means

that any simple approach to calculating the lift and drag characteristics of the blade section

would be inaccurate. The maximum forward speed of the advanced helicopter is much greater

than the conventional helicopter, and consequently, the effect of the compressibility and

reverse flow have a more significant influence on its flight dynamics characteristics. Therefore,

NASA provided the aerodynamic characteristics tables of various airfoil types according to a

range of wind tunnel experiments, which demonstrates the aerodynamic lift and drag with

different angle of attack (-180 degrees to 180 degrees) and Mach number (0.0 to 1.0) [23]. Thus,

the results obtained can be utilised to determine the airfoil aerodynamics in the reverse flow

area and the influence of the compressibility effect. Other researchers have combined the airfoil

aerodynamics model with a CFD technique to determine these effects and calculate the
[86-89]
aerodynamic properties of the airfoil in different flight ranges , however, the additional
20

time-cost of incorporating a CFD approach would deteriorate the computing efficiency of the

flight dynamics model.

Radial flow can influence the airfoil aerodynamics characteristics, especially in the high-

speed flight range, suggesting that it should be considered in the advanced high-speed rotorcraft.

According to the relevant experiments [90], the radial flow can induce an additional normal force

on inboard blade sections due to centrifugal and Coriolis forces and therefore alter the airfoil

aerodynamics and flight dynamics characteristics. Thus, the radial flow could delay the stall of

the rotor disc to some extent. A series of correction methods have been put forward to simulate

the radial flow [90-91], but the parameters in the correction equations have to be determined by
[91]
the relevant experiments or appropriate CFD technique. Breton utilised the lifting-line-

prescribed wake vortex scheme to calculate this effect and used a wind tunnel experiment to

verify this method. The calculation results share a similar trend with the experimental results,

indicating that it can be adopted in flight dynamics models to improve their accuracy.

In conclusion, different methods have been developed to enhance the precision of the airfoil

aerodynamics model. Nevertheless, most of these methods have to rely on the relevant

experiments or CFD calculations.

3.1.2 Rotor wake model

The induced velocity is determined by the rotor wake model, which alters the rotor

aerodynamics and the flight dynamics characteristics. In addition, advanced rotorcraft usually

utilises a multi-rotor system, and therefore, the aerodynamic interaction between rotor discs

plays a significant effect on the induced velocity on each disc and consequently changes the

flight dynamics features of the advanced rotorcraft. Therefore, with the aim to accurately

simulate the flight dynamics characteristics of the advanced rotorcraft, the interaction in the

multi-rotor system should also be considered in the rotor wake model.


21

There are two important requirements for the rotor wake model: accuracy and computing

efficiency. The accuracy of the rotor wake model strongly influences the validity of the flight

dynamics model, and the computing efficiency defines the capability of the flight dynamics

model for analysis in practical circumstances. High time cost limits the ability of the flight

dynamics model to investigate the control response and handling qualities.

Different types of wake models have been developed [92], such as the most straightforward
[93-97]
rotor disc uniform induced velocity model , the finite-state inflow [98-101], the fixed wake

method [102-108], the free wake model [109-113], and higher-resolution wake models developed [114-
121]
recently.

The uniform induced velocity model is a simplified rotor wake model with high computing

efficiency, which is based on the relationship between the induced velocity and the

aerodynamic loading on the rotor disc [92, 95]. This model can be regarded as a particular wake

model derived from the momentum theory. However, the feasibility of this method is limited.

When the helicopter is in forward flight, the wake will tilt backwards, and the induced velocity

distribution becomes non-uniform, reducing the accuracy of the uniform induced velocity

model. Thus, Coleman at [93] built a linear induced velocity model with the wake correction,

enabling the method to simulate the induced velocity distribution in the mid to high speed

forward flight. However, the accuracy of this model is reduced in low-speed forward flight due

to the wake distortion. In addition, these induced velocity models could only be utilised for the

single rotor system, and cannot consider the aerodynamic interaction in the multi-rotor system,

which means it cannot be directly used for the most of the advanced helicopter configurations.

Carpenter and Friedovich [94] expanded the momentum theory to the dynamic inflow model,

in which the additional inertia effect produced by the disturbance on the rotor disc is considered.

This dynamic inflow model has the capability to simulate the dynamic change of the induced

velocity, and this induced velocity is given in the form of first-order ordinary differential
22

[95]
equations. Pitt, Peters, and He constructed the Pitt-Peters dynamic inflow model and its
[96, 97]
generalised form (Pitt-He finite-state (high order) inflow model) according to the

acceleration potential theory. This wake model has been widely used for the flight dynamics

analysis and control response calculation. Nevertheless, the prerequisite of these models is that

the inflow velocity should be much more than the induced velocity on the rotor disc. Thus, the

dynamic inflow model only can be used for rotors with low loading or when the helicopter is

in mid to high speed forward flight. In addition, Ferguson [37] utilised a revised dynamic inflow

model to investigate the flight dynamics characteristics of the coaxial compound helicopter.

However, an empirical correction has to be applied to modify the parameters in the dynamic

inflow model, hindering its further utilisation.

Peters further developed the finite state inflow method based on the Galerkin treatment of

the potential flow equations, allowing this method to compute induced flow everywhere in the

flow field [98-101], suggesting that it could be used to determine the aerodynamic interaction in

the multi-rotor system. According to the comparison against flight tests, the obtained result

gives acceptable precision for flight dynamics analysis so that this rotor wake model has the

potential to be adopted into the flight dynamics modelling of the advanced rotorcraft, such as

the coaxial compound helicopter and the tilt-rotor aircraft. However, this method is only
[99]
verified in the hover state of the coaxial helicopter configuration , and the numerical

convergence of this method is also a significant impediment for its further development.

In order to develop a generalised rotor wake model to accurately simulate the non-linearity

of the induced velocity on the rotor disc throughout the flight range, as well as calculate the

aerodynamic interaction in the multi-rotor system, Barocela [102], Krothapalli [103], Zhao [104],

Rosen and Isser [105, 106], and Keller [107, 108] put forward different wake distortion models using

the pre-scheduled curvature method. They defined a parameter Kre for the wake curve to reflect

the proportional relationship between the rotor induced velocity gradient and the wake
23

curvature. However, the parameter of Kre must change along with the forward speed to ensure
[109]
its accuracy at different flight ranges. Bhagwat constructed the correlation between Kre

and the forward speed, rotor angular acceleration, and rotor thrust, enhancing the feasibility of

the method. However, the pre-scheduled wake distortion method still fails to fully reflect the

distortion of the rotor wake geometry and cannot take the aerodynamic interaction of the multi-

rotor system into account. Thus, it cannot be used to precisely simulate the flight dynamics

characteristics of the advanced rotorcraft.

The free-wake model is another approach to calculate the rotor wake of the helicopter, which

was developed based on the rotor vortex theory [110], and this model solves for the rotor wake

geometry directly, and in principle do not require experimental data for formulation purposes.

In this method, the wake system was usually decomposed into two main parts. Firstly, a near

wake of trailed and shed vorticity behind each blade and second, a far wake comprising the

rolled-up tip vortices from the blade. Then, the numerical solution to the free-wake problem

can be described by the integration of a system of ordinary differential equations. These are

obtained after the spatial discretisation of a series of partial differential equations that govern

the positions of the tip vortices. A set of collocation points are specified on the trailed vortex

filaments, and these points are numerically converted through the flow field at the local velocity.

The curved tip vortices generated by the blades are usually divided into a number of smaller

straight-line segments. The local velocities at each collocation point on the vortex filament are

then calculated by the application of the Biot-Savart law. Thus, this method allows the vortex

element to move with the local airflow velocity and can automatically simulate the self-

induction and distortion of the wake. The induced velocity vector at anywhere in the flow field

can be obtained using this method. The free-wake method can be used for advanced rotorcraft

modelling as it can not only calculate the aerodynamic interaction in the multi-rotor system but

also be able of capturing the effect of wake distortion during flight. In its early development,
24

the explicit Euler time marching method was widely utilised, however, its numerical stability

is relatively weak. To solve this numerical instability, two numerical methods have been put

forward. The first method is to introduce a constraint of the periodic condition during the renew

process in each time step, referred to as the classical relaxation free wake method. The second

method is to combine the forecast-correction method with the high order time marching format

to reduce the numerical oscillation, referred to as the time-accuracy free wake method.

The line vortex discrete embedded free wake model is a relatively mature method for the

flight dynamics modelling of helicopter. This method could not only guarantee the precision

but also reduce the overall time cost. Meanwhile, this method shows a significant efficiency

advantage in computing the aerodynamic interference between rotors, which is essential for the

advanced rotorcraft modelling process. However, the line vortex discrete embedded free wake

model is based on the potential flow theory, excluding the viscous effects. In order to improve

the accuracy of this method, researchers applied the empirical coefficients into the vortex core

model, and the position of the tip vortex distortion are used to include the effect of viscosity.

However, these empirical coefficients impede its applicability for the manoeuvrability and

control response analysis. Lee and Na [111, 112] constructed a new rotor wake method with vortex

blob method, successfully solving the problem of the numerical convergence. However, its low

calculation efficiency in tackling the self-induced velocity limits its further development and

utilisation for the flight dynamics modelling of the advanced rotorcraft.

Yuan, Chen, and Li developed a multi-transmutable-vortex-ring (MTVR) wake model based


[113]
on the rotor disc assumption and fixed wake theory . The induced velocity everywhere in

the flow field is attainable with this rotor wake model, and its computing efficiency is much

enhanced, making it able to achieve the real-time requirement. The validation results indicate

that this model can accurately simulate the induced velocity distribution of the coaxial
25

compound helicopter in different forward speeds. However, the wake variables of this method

are extensive and that leads to difficulty in the convergence during the calculation.
[114]
In recent years, with the development of the fast multipole method (FMM) , many

researchers tried to combine this method with high-resolution general vortex method to

construct the rotor wake model. Brown [115, 116] firstly built the vortex transport method (VTM)

method for the high accuracy rotor wake calculation according to the finite volume method.

He and Zhao [117, 118] constructed the viscous vortex particle method (VVPM) for high precision

rotor wake estimation. These methods not only inherit the advantage of non-viscous free wake

method but also consider the effect of the viscous dissipation and the wake geometry alteration

on the rotor aerodynamic characteristics. However, these methods usually utilise the lift-line

or lift-surface model due to the convergence requirement, reducing the accuracy in calculating

the airflow characteristics around the blade. The rotor CFD method is widely utilised to

accurately simulate the rotor wake influence, and the effects of the airflow separation, dynamic
[119]
stalling, and the shock wave on the rotor wake are all considered in this method . This

method usually suffers from numerical dissipation problems, leading to over fast attenuation

on the vorticity. For this purpose, some researchers combined the wake method with a CFD
[120, 121]
method to develop a high precision wake calculation approach . A CFD method is

utilised to capture the flow field detail characteristics, and the rotor wake model is used to

calculate the wake viscous dissipation and topological structural change. Therefore, this

method can obtain the induced velocity in the flow field and be used to accurately construct

the flight dynamics model for advanced rotorcraft, especially the simulation of the aerodynamic

interaction in multi-rotor systems. Nevertheless, the time cost of this method is extremely high,

decreasing the computing efficiency of the flight dynamics model.

3.1.3 Blade dynamics model


26

The blade motion consists of three parts: flapping motion, lagging motion, and torsion

motion (pitching motion). The flapping motion is the most important component to helicopter

flight dynamics modelling as it decides the control power and propulsion of the rotor system.

The blade motion of the advanced rotorcraft is usually different from the conventional

helicopter in order to improve its performance. Coaxial compound helicopters, such as X2TD

and SB-1 helicopter, utilise what is referred to as a rigid rotor to delay the dynamic stall
[122-126]
phenomenon in high-speed flight range . The tilt-rotor aircraft adopts gimbal rotors to

deal with the aeroelastic instability problem in high-speed aeroplane mode [127]. These features

significantly change the blade motion characteristics and consequently alters the flight

dynamics of the advanced helicopter.

In steady flight, the periodic characteristics of the blades’ aerodynamics are similar as their

motion trajectories are same, and the tip planes of different rotor blades maintain in the same

shape. Thus, the coning angle, longitudinal flapping angle, and the lateral flapping angle can

be used to fully describe the flapping motion of the blades. This description and modelling

method are regarded as rotor plane method. The orientation of the aerodynamic forces and the

effect on the vehicle motion can be easily determined using this method. However, the

aerodynamic and inertia force on each blade will be different in large amplitude manoeuvring

flight. Meanwhile, the turbulent environment could also make the trajectories of rotor blades
[50].
located in different planes Consequently, the trajectories of different rotor blades are no

longer kept in the same plane so that the accuracy of the rotor plane method is reduced. In order

to improve the accuracy of the blade motion calculation for manoeuvring flight, two different

approaches have been developed.

The first method is to assume that the rotor blades still have the same dynamic trajectory,

and it can be described as [17]:


27

(1)

where: a0, a1, and a2 represent the coning angle, lateral flapping angle, and longitudinal

flapping angle; , , and denote the damping matrix, stiffness matrix, and external

excitation vector, respectively. This approach only takes the dynamic change of the rotor disc

into account, indicating that it is only suitable for the small-to-moderate amplitude

manoeuvring flight.

The second method abandons the hypothesis that each blade should be kept in the same plane
[23]
and separately investigates the flapping motion of each blade in rotational coordinates .

Compared with the first method, this approach can sufficiently capture the Coriolis force

derived from the vehicle angular motion and the inertia force from manoeuvring flight. It

should be mentioned that this method is also appropriate for the blade motion modelling in the

steady flight. The path of each blade is similar to the others, and the trajectory would be

therefore back to the same rotor plane.

The blade motion has a direct relationship with the design of the rotor hub. In terms of a

rotor modelled by a centrally located flapping hinge, the first order flapping frequency of the

blade is the same as the rotor rotational speed, which makes it easy to model. The design of the

rotor hub becomes unique in some helicopter configurations, such as the gimbal and high-

rigidity rotor hub design. These types of rotor hub alter the flapping frequency and blade motion

characteristics and consequently change the control and stability characteristics of the

helicopter. With the aim of taking these effects into consideration and maintaining the

computing efficiency, the blade motion can be simplified by modelling the flapping dynamics

by the simplification shown in Figure 6, in which e is the flapping hinge offset, Kβ is the

stiffness of equivalent flapping spring, and β is the flapping angle.


28

Rot tion is
l de

Hu l ne
Fl ppin hin e ith sprin onstr int
Figure 6 Equivalent articulated or hingeless rotor [17]
In this simplified method, the equivalent flapping spring is used to adjust the flapping

frequency in order to ensure it matches that of the real rotor. This simplified method is

appropriate for rotors with the flapping frequency less than 1.1 Ω, where Ω is the rotor

rotational speed.

The gimbal rotor hub of the tilt-rotor aircraft can be simulated using this method [2], however,

there are some other advanced rotorcraft configurations equipped with highly rigid rotor blades

where the flapping frequency can be more than 1.4 Ω. In order to simulate the blade motion of

the rigid rotor, there are two requirements that should be met to guarantee accuracy: the first

one is to ensure the flapping frequency remains the same before and after the equivalent system

is defined; secondly, the flapping mode after the equivalence should be as similar as possible

compared with the original flapping mode. The equivalent model of the rigid blade flapping

motion is shown in Figure 7, which can satisfy the conditions discussed above.

Figure 7 Equivalent method for rigid blade flapping [128]


29

The non-dimensional equivalent flapping offset is calculated using the equation shown

below to guarantee the similar flapping mode

(2)

where: Wtip is the flapping amplitude of the original flapping motion; R is the rotor radius;

W`0.75 denotes the flapping angle at 0.75 R.

To keep the flapping frequency, the additional flapping constraint spring is needed, and its

stiffness can be expressed as [3]

(3)

where: is the non-dimensional flapping frequency of the first-order flapping; Mβ is the

static moment of the blade mass; Iβ is the blade inertia moment.

This method assumes the blade is rigid in the flapping motion, which is the mainstream

approach in the helicopter flight dynamics modelling as it would simplify the calculation

process and improve the computing efficiency. With the development of the advanced

rotorcraft, there are higher requirements for rotor motion modelling. The combination of an

elastic blade model with a helicopter flight dynamics model has drawn growing attention due
[129-133]
to the improvement in precision it provides . The Finite-Element-Method (FEM)

embedded elastic model can represent the elastic deformation of the coupling between blade

flapping, lagging, and torsion motion, which further improves the precision in the blade model,

especially for advanced rotorcraft equipped with the rigid rotors. Duval, He, and Turnour and
[129128-131]
Celi constructed different flight dynamics models with this elastic blade motion

approach, and they pointed out that the elastic blade motion model could efficiently improve

the precision of the helicopter in off-axis control response. References [132, 133] combined the

FEM method with advanced wake model and airfoil unsteady/dynamic stall model, suggesting
30

that the accuracy is further improved when calculating the rotor loading in different flight

ranges. The FEM incorporated an elastic blade model is a powerful approach for the advanced

helicopter flight dynamics modelling to enhance the calculation accuracy in the high speed and

manoeuvring flight. However, the additional time cost brought by the FEM method and other

elastic blade motion models reduces the time efficiency of the flight dynamics model.

3.2 Model of aerodynamic interference among components

The aerodynamic interference among helicopter components is the most challenging feature

to capture in a flight dynamics model. The interference among the rotor system, fuselage,

horizontal and vertical tails alters the flow field and pressure distribution on each component,

influencing the resultant force and moment. In addition, the aerodynamic interference becomes

more extensive for advanced rotorcraft. As well as the additional force and moment due to the

interference, advanced rotorcraft configurations usually experience other resulting problems.

For example, the rotor wake changes the aerodynamic characteristics of the horizontal and

vertical tails, and consequently control power of the elevator and rudder are altered. Further,

for tilt-rotor aircraft aerodynamic interference influences the lift-to-drag ratio of the wing and

affects the performance characteristics. Coaxial compound helicopters have different wake

features due to the interference between the coaxial rotor system, which affects the

aerodynamic characteristics of other parts of the helicopter significantly. The hybrid compound

helicopter has two auxiliary propellers situated at each side of the wing, and their wakes will

couple with the rotor wake, which leads to a significant wake effect on the vertical and

horizontal tail. Also, poor propeller inefficiency may occur in the hybrid compound helicopter

when the forward speed equals to the induced velocity of the propeller. Therefore, it is worth

exploring aerodynamic interference among the helicopter components in more detail and

detailing modelling methods to capture their effects.


31

Figure 8 is a schematic diagram of the aerodynamic interference between rotor and


[134]
horizontal tail when the helicopter is flying from hover to forward flight . The rotor wake

sweeps backwards and impacts the horizontal tails to produce the nose-up moment as the

forward speed increases. Then, as the forward speed further increases, the rotor wake sweeps

upward missing the horizontal tails area, which then returns to providing a nose-down moment.

This phenomenon alters the trim characteristics of the helicopter and may induce the helicopter

dynamic instability [27]. Also, the wake interference also changes the inter-axis coupling effect

and consequently damages the handling qualities of the helicopter.

This interference of the rotor wake on the vehicle not only degrades the handling qualities

of the helicopter but could also damage the helicopter during flight. During the development

of the AH-64 helicopter, there were a number of horizontal tail redesigns to avoid rotor wake

effect on the horizontal tail, and this was the reason for an accident during the flight test

programme. Finally, manufacturers have had to change the location of the horizontal tail and

adopt an all-moving horizontal tail to meet the handing qualities requirement [135], as shown in

Figure 9. Interference induced vibration occurred in the YUH-61 helicopter owing to the lower
[136]
distance between the rotor and fuselage . This led the Boeing Company to start a

programme called UTTAS for seven years to research the internal mechanism between the
[58, 137]
aerodynamic interference. The programme included a large number of experiments ,

which produced a large volume of test data. This programme has pushed forward the research

on the aerodynamic interference for helicopter development to a large extent.


32

Nose up pit hin moment

utside o the rotor


e ound r In re se the do nlo d
on hori ont l t il

Rotor e
The e ound r mo es rd

Figure 8 Aerodynamic interference of rotor wake with horizontal tail

Figure 9 The horizontal tail before and after the redesign in AH-64 helicopter [12]
The downwash and side wash effects of the rotor on other parts of the vehicle mainly

influences the dynamic pressure, angle of attack and sideslip angle at the helicopter sub-

components. Issues caused by rotor wake instability and time-varying characteristics can be

very difficult to describe analytically within a model, and so data from wind tunnel experiments

are widely used to predict this influence. Figure 10 shows wind tunnel experimental results for

dynamic pressure on the tailplane of YUH-61A helicopter [8]; where is the dynamic pressure

ratio (the difference between the local pressure and the free flow pressure divided by the free-

flowing pressure), and vh is the induced velocity in hover state. As shown in Figure 10, the

horizontal tail is affected by the rotor wake, increasing its dynamic pressure above the pressure

in the free-flow.
33

1.4
Flight Test
1.2 Wind tunnel test

Dynamic pressure ratio


1.0

0.8

0.6

0.4

0.2

0.0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0
V/vh

Figure 10 Dynamic pressure ratio at the horizontal tail [8]


The downwash and side-wash caused by the rotor wake can be expressed as follow

(4)

where vi represents the induced velocity; k denotes downwash or side-wash factor caused by

rotor wake, which is determined by the rotor wake skewing angle and its relative position from

the rotor hub. In practice, the factor k is obtained by the relevant wind tunnel experiments or

numerical calculation techniques, such as CFD.

The correlation between the downwash factor k and the rotor wake skewing angle is

illustrated in Figure 11, demonstrating that the downwash factor increases with the rotor wake

skewing angle. It also indicates that the effect of the rotor wake on the horizontal tail diminishes

as forward speed increases. On the other hand, the downwash factor is close to 2.0 when the

rotor wake skewing angle is 90 degrees. In other words, the downwash velocity is twice that

on the rotor disc in hover state, which is in line with the result derived from momentum theory.
34

2.0

1.5

k
1.0

0.5

0.0
0 20 40 60 80 100
c/(Deg)

Figure 11 Relationship between rotor downwash factor and wake angle [9]
Based on the experimental results mentioned above, the rotor wake effect is dependent on

the flight status, the configuration, and the rotor design. During the modelling process, it is a

significant challenge to determine which parts of the helicopter are affected by the rotor wake

and what additional influence there is due to this interference. Moreover, the dynamic motion

of the rotor wake also occurs during manoeuvring flight, enlarging the effects of aerodynamic

interference. Both features should be considered during the modelling process.

In the 1980s to 1990s, researchers started to implement theoretical studies to simulate


[138-145]
aerodynamic interference with the developmental free wake research . A theoretical

prediction model of the rotor/fuselage aerodynamics was developed to determine the unsteady

aerodynamic loading on the rotor and fuselage, including the effects of aerodynamic

interference.

In order to validate the accuracy of this model, Georgia Institute of Technology [146-149] and
[150, 150]
the University of Maryland separately built rotor/fuselage combined experimental

model from 1989 to 1991. In these experiments, the flow field and the pressure distribution

around the fuselage in different conditions were measured. These results have become the

validation baseline for subsequent theoretical analysis methods [54, 104].


35

[109,
With the breakthrough in the numerical stability of unsteady rotor free wake methods
152]
, the theoretical prediction model is relatively straightforward to use in helicopter flight

dynamics modelling. Horn [46] and Ribera [48] combined the time-precision free wake analysis

method with an existing flight dynamics model. Wachspress [153] introduced the time-precision

free wake model into the rotorcraft simulation process, and validations illustrate that the free

wake model could improve the accuracy of the predictions of helicopter aerodynamic

characteristics, especially for the aerodynamic interference derived from the rotor wake. D'

Andrea [49] adopted the time-precision free wake model with unstructured surface element grid

method and developed the ADPANEL method for the aerodynamic interference analysis.

Recently, with the development of the advanced rotorcraft, a set of research has been put

forward towards the aerodynamic interference of the tiltrotor aircraft and compound helicopter.

The main focus in the aerodynamic interference of tiltrotor aircraft is the interference

between the rotor system and the wing. This interference determines the flight dynamics

characteristics and performance characteristics of the aircraft, and could be changed with the
[154]
nacelle incidence angle, flight speed, and other flight states. Yeo calculated the

aerodynamic interference effect on the performance characteristics of the tiltrotor aircraft with

CFD/CSD coupled method on the CAMRAD II platform. Based on the analysis results, the

aerodynamic interference effect improves the aircraft lift-to-drag ratio, and the interference

velocities reduce the total induced velocity along the wingspan and, thus, reduce wing induced

power. Jung [155] investigated the aerodynamic interference between the rotor system the wing

with different sideslip angles and nacelle incidence angles based on a CFD flow solver. The

results indicated that aerodynamic interference magnifies the fluctuating amplitudes of the yaw

and roll moments with the increase of the sideslip angle, and the aerodynamic interference is

more significant when the nacelle is tilted forward. These results provide a deeper insight into

the aerodynamic effect of the tiltrotor aircraft; however, the time cost of these calculation
36

process is extremely high. Thus, the empirical factors are widely used in the current flight

dynamics model [156-158], especially the model that needs to achieve the real-time requirement,

to deal with the aerodynamic interference in the tiltrotor aircraft.

There are different types of aerodynamic interference inside the compound helicopter that

play a major effect on its flight dynamics characteristics. Apart from the rotor-tail surface

interference, the rotor-wing interaction and the rotor wake effect on the propeller also
[159]
contribute to the flight dynamics features and performance characteristics. Yeo

investigated the performance characteristics of different compound helicopter based on

CAMARD II. The results indicated that rotor/wing interference effects are examined for a

compound helicopter at a high-speed cruise flight condition where the rotor carries about 7%

of the gross weight and the wing carries about 93% of the gross weight. The interference

velocity on the rotor is relatively large due to the large wing lift compared to small rotor
[160]
induced velocity. However, interference power is very small in cruise. Stokkermans

utilised the unsteady CFD simulation technique to investigate the installation effects of the

latera rotors for a hybrid compound helicopter featured a box-wing design. The results

suggested that the main interaction in cruise was between the wing and lateral rotors, resulting

in a propulsive efficiency increase up to 10.6% due to wingtip vortex energy recovery. In hover

the main rotor slipstream resulted in a near perpendicular inflow to the lateral rotors, with a

disturbance from the wings due to the deflection of the main rotor slipstream. Although the

high-accuracy results can be obtained from these CFD solvers, their time cost is still significant

and consequently cannot be directly adopted into the flight dynamics modelling. On the other

hand, Ye applied the MTVR wake model into an aerodynamic interference calculation and

utilised the flight test data to verify its accuracy [113]. The comparison indicates that this method

could capture the primary influence of the rotor wake on the other parts of the helicopter, and

the trim calculations follow the flight test results with good accordance. Additionally, the
37

computing efficiency of the MTVR model is better than the free wake model and can be used

to the control response simulation and manoeuvrability analysis.

3.3 Engine/fuel control system modelling

The engine has a significant coupling effect on the flight dynamics characteristics of the
[161, 162]
helicopter . During the steady flight, the flight state of the helicopter is relatively

constant so that the power requirement and the power output of the engine roughly remain

constant. However, when the helicopter is in manoeuvring flight or experiencing atmospheric

disturbances, the pilot needs to keep changing the control input of the helicopter and the airflow

around the helicopter is therefore influenced. Thus, helicopter flight dynamics characteristics

and the associated engine power output also vary to a large extent. Therefore, the dynamic

characteristics of the engine must be taken into consideration as additional lag or overshoot

effect may occur due to the engine characteristics, and consequently influence the flight
[162]
dynamics and handling qualities of the helicopter . Moreover, advanced rotorcraft

configurations usually employ multi-rotor systems or rotor-propeller combination systems,

further complicating the formation of the power output and engine/fuel control system

modelling process.

There is significant research into engine dynamics characteristics independent from

helicopter flight dynamics studies. Research relating to the rotorcraft engine dynamics usually

assumes that the power output of the engine and the rotorspeed are invariable and fails to

consider the coupling effect between the rotor dynamics and engine. Ballin built a real-time

simulation model of the T-700 turboshaft engine taking into account the aerodynamic and
[163]
thermodynamic characteristics of the engine . Then, Ahmet utilised an identification

method to construct a simplified linear model of the T-700 turboshaft engine on the basis of
[164]
the Ballin model . As mentioned above, the dynamics characteristics of the engine are

excluded in many helicopter flight dynamics studies, and it is often assumed that the power
38

required can be met instantly by the engine. For example, the ARMCOP model, widely used

for rotor system design and analysis, ignores the dynamic influence of the engine. [165]

Many researchers have tried to improve vehicle modelling accuracy by incorporating the

dynamic effect of the turboshaft engine. Talbot put forward a simplified engine/fuel control
[166]
system model for helicopter flight dynamics investigation . This model utilised a second-

order transfer function to represent the engine dynamics effect, including the compressor,

throttle control, power turbine, and fuel control. This method is simple in structure and easily

adopted in flight dynamics models. However, the detailed response of the engine is neglected

in the simplified transfer function. The GENHEL helicopter simulation package included an

engine model of the T-700 turboshaft engine and its fuel control system. This helicopter

rotor/engine integrated model enhanced the accuracy of the flight dynamics analysis in

manoeuvring flight. The engine model in the GENHEL package ignores the inlet pressure of

the engine turbine, which reduces its precision during the large amplitude manoeuvre [167, 168].

A specific schematic modelling method was developed for the T-53 turboshaft engine used
[169]
in the XV-15 tilt-rotor aircraft . The model is composed of equations to calculate engine

horsepower during transient and steady-state based on the operating characteristics of the

combined engine-fuel control system. The model includes the dynamic effect of the engine

with more detail and ensures the real-time requirement for the associated flight dynamics can

be met. However, the parameters in this model are determined by many experiments and can

only be suitable for a specific combination of engine and helicopter type. In addition, Cranfield

University developed a gas turbine performance simulation code called TURBOMATCH [170-
172]
. It is a long-standing and validated tool of the engine suitable for both steady and transient

conditions. Researchers have used it for flight dynamics modelling and analysis of the

advanced coaxial rotor configuration, such as the X2TD helicopter [172].

3.4 The integration and calculation of the non-linear systems


39

Helicopters experience a wide range of nonlinear effects which produces unique dynamic

characteristics. The calculation and integration of this non-linear system is a great challenge in

the flight dynamics modelling process. Also, the additional components present in advanced

rotorcraft put forward the higher requirement for this integration and calculation procedure.

First, the vehicle body motion of the helicopter features low-frequency characteristics and

strong coupling, and therefore, the governing differential equations describing the helicopter

motion are non-linear. Furthermore, aerodynamic interference plays a significant effect on

these non-linear characteristics. The downwash or side wash of the rotor wake causes the

aerodynamics on the fuselage, horizontal tail, and vertical tail to be discontinuous, bringing

about further problems in the calculation process.

Second, the flight dynamics model needs to incorporate a rotor wake model to capture the

aerodynamic interference. However, current rotor wake models still suffer from the problems

of numerical instability and calculation inefficiency. The combination of the flight dynamics

model and the discrete rotor wake model exaggerates this effect. In order to improve the

numerical stability, Theodore [173] and Ribera [48] adopt the simplified free wake model of Bagai
[174] [109] [45] [47]
and Bhagwat into the FLEXUM model. Spoldi and Horn utilised a similar

simplification when combining the CHARM free wake model of the CDI company with the

GENHEL model. A loose coupling method is needed for these applications in order to reduce

the numerical instability and improve calculation efficiency. According to the published

research mentioned above, the integrated method between wake model and flight dynamics

model indicates the potential of the rotor wake model in enhancing the accuracy and computing

efficiency of the flight dynamics model.

Third, the unsteady and dynamic stall characteristics of the airfoil aerodynamics influence

the air loading of the rotor blade. The Leishman-Beddoes model [80], ONERA model [175, 176],

or the Johnson model [3] can be used in blade loading calculation, to consider those effects. The
40

unsteady blade aerodynamic characteristics mainly focus on the effect of shed vortex in the

wake. When the rotor wake model and the airfoil unsteady aerodynamics model are both

utilised to model the helicopter flight dynamics characteristics, care should be taken to avoid

the repeated inclusion of the unsteady effect from the shed vortex.

Fourth, the governing equations of the helicopter flight dynamics model need to be expressed

with a specific format due to the coupling effect of the blade dynamics feature and the

interaction between the rotor and fuselage, impeding the utilisation of traditional solving
[131]
methods. Tornour and Celi utilised analytical solutions to separate the inertia coupling

related to the vehicle acceleration from the rotor/fuselage coupling dynamic functions, and this

part was rescheduled to a first-order ordinary differential equation. This method has been

widely used for flight dynamics analysis and simulation.

Finally, the objective to introduce the engine/fuel control system is to make the helicopter

flight dynamics model suitable for manoeuvrability investigation. Compared with the steady

flight, the non-linearity significantly increases in manoeuvring flight, and this non-linearity

will couple with the engine/fuel system, which will put forward higher requirements in the

system integration and solution procedures.

Predictably, with the increasing complexity of the rotor aerodynamics model and

aerodynamic interference model, a growing effort will be needed to develop its integration and

calculation methods. The critical challenge for the high-confidence helicopter flight dynamics

modelling method is to strike a balance of the modelling accuracy and computing time cost to

ensure both the accuracy and efficiency of the flight dynamics model can be satisfied at all

flight conditions.

3.5 Progress in manoeuvrability analysis

Flight test and pilot-in-the-loop simulation have been widely used for manoeuvrability

analysis for the helicopter. Many the flight tests have focussed on different Mission-Task-
41

Elements (MTEs) using various helicopter configurations [177-181], providing valuable material

for the helicopter flight dynamics assessment. Flight testing is expensive, potentially dangerous

and can only take place once a prototype aircraft is available, and not in the early design phase

where potential benefits could be identified.

Pilot-in-the-loop simulation is widely used due to its economic efficiency and capability to

be adopted in the early design process. There are a number of helicopter simulators being used
[182-185]
across the world to assess the manoeuvrability of different helicopter configurations .

Nevertheless, the method puts forward a higher requirement for the flight dynamics modelling

technique. The flight dynamics model has to meet the real-time requirement when adopted into

the flight simulator. In other words, the application of the high-precision method, including the

rotor free wake method and FEM rotor dynamics model, narrows the feasibility of this method.

Effectively, the handling qualities assessment from the pilot-in-the-loop simulation still has a

significant error compared with the flight test results, and it cannot be used to replace the flight

test entirely at present.

On the other hand, with the increase of the flight dynamics modelling accuracy, a range of

novel methods for manoeuvrability analysis have been put forward, including the nonlinear

optimal control (NOC) method and the inverse simulation method.

The NOC method is based on the collocation and the numerical optimization method and

has been adopted into the manoeuvrability investigation of the conventional helicopter [186-189]
[190, 191]
and tilt-rotor aircraft . This method utilises a human operator model to take the pilot

biometric lag into the simulation process and improve its accuracy. However, this method has

internal numerical instability and can suffer from convergence problems once the discrete

modelling technique, such as the rotor free wake model, is adopted. Therefore, the NOC

approach is only suitable for the task profile (e.g. flight range and flight duration calculation)

investigation and the small-to-moderate amplitude manoeuvrability analysis.


42

The inverse simulation has been created and steadily developed in recent years and has been

employed to investigate the manoeuvrability and subjective handling qualities of various


[64, 65, 192-201]
helicopter configurations . Plenty of progress has been made for the inverse
[64-65]
simulation method to improve its accuracy and efficiency. Rutherford has adopted the

individual-blade-motion into the inverse simulation approach, making it possible to consider

the blade motion effect in more detail. Cameron introduced a pilot model into the inverse

simulation method [197]. Thus, the pilot-induced oscillation is included in the calculation results.

Hess [198, 199] has tried to use a pilot model to conduct the inverse simulation of aggressive

mission tasks, demonstrating that this analysis allows well-established compensatory models

of human pilot behaviour to produce realistic pilot responses in discrete manoeuvres. Lee [200]

utilised the pilot model to conduct the inverse simulation of helicopter shipboard operations.

Results show that the unsteadiness of the ship airwake has a significant impact on pilot

workload when the helicopter is operating near the deck and superstructure of the ship.
[201]
Meanwhile, to enhance the efficiency and precision of the results, Ye has adopted the

Automatic Differentiation (AD) method into the flight dynamics model, accelerating the

computing speed during the inverse simulation process. The AD method is based on the chain

rule of the differentiation process. The calculation results indicate that the AD method

embedded inverse simulation method could satisfy the real-time requirement. In other words,

the proper control inputs can be calculated in advance of the real-time period using this inverse

simulation method, which widens the application of the inverse simulation approach in the area

of automatic control, control system design, and flight simulator assessment.

However, the inverse simulation method is still under developing. Firstly, it utilises a pre-

determined trajectory, a mathematical description of the manoeuvring task needs to be given

in advance. However, some manoeuvres, including some of the mission-task-elements (MTEs)


[1]
defined in helicopter handling qualities requirements (ADS-33F-PRF) do not have a fixed
43

trajectory. In other words, it is hard to describe these MTEs mathematically. Secondly, inverse

simulation obtains the control input by inversely solving the helicopter flight dynamics

equations, which may have more than one solution in some flight states. Thus, additional

constraint conditions are needed in order to obtain an optimized solution, and these constraints

may not be entirely realistic in the context of actual vehicle piloting strategies. Although the

inverse simulation approach can obtain the cockpit input that satisfies the performance

requirement according to the handling qualities, the obtaining control strategy is one of the

many control methods that could meet the requirement for given MTEs. Moreover, the

advanced rotorcraft usually has redundant control inputs, including the rudder and the elevator,

which also influences the inverse simulation results. In some research, additional boundary
[36, 38]
conditions need to be added to investigate its manoeuvrability . In short, there are still

many challenges for the further development of the inverse simulation method.

The main focus in manoeuvre studies recently has been the assessment of handling qualities

ratings based on the control input results obtained from NOC or inverse simulation method.

Wavelet analysis methods have been widely utilised due to its excellent time-frequency

resolution [202-205]. In wavelet analysis, the finite-length bandpass filter is introduced to illustrate

the signal energy in the frequency spectrum and time histories. With the wavelet analysis

method, the main frequency components during the control input can be identified. According

to the research of Thits hler nd ’ Conner [205]


, the pilot workload and pertinent handling

qualities rating are dependent on the main frequency components. Additionally, the numerical

correlation between the main frequency range and handling qualities ratings can be constructed.

Thus, the pertinent handling qualities rating can be calculated in a straightforward manner once

the control input is obtained.

4. Conclusion
44

Significant advances in helicopter mathematical modelling techniques are detailed in this

paper. However, there is still a need for further improvements to investigate flight performance,

flight dynamics, and handling qualities characteristics especially for advanced rotorcraft

configurations. The specific research topics that will require future attention are as follows:

1) Improvement in the accuracy of the rotor aerodynamics model is still the most critical

aspect of the helicopter flight dynamic modelling. Rotor flow field models have

advanced from initial slipstream theory to the high-resolution rotor wake model able to

calculate the aerodynamic characteristics found in advanced rotorcraft configurations.

However, the high-resolution rotor wake model usually uses the lift-line or lift-surface

models and excludes the detailed flow features around the blade, which limits its overall

precision. The accuracy of conventional aerodynamic modelling methods is reduced to

a large extent in high-speed flight due to dynamic stall and reverse area flow effects,

making them of limited use in the simulation of advanced high-speed configurations.

With multi-rotor systems used in the advanced helicopters, the distortion of the rotor

wake put forward higher requirements for the rotor aerodynamics calculation process,

and advances in this area are certainly a priority for improved predictions.

2) Although much progress has been achieved in the aerodynamic interference calculation,

most methods are still dependant on the use of wind tunnel experiments or CFD

simulation, and so are only valid for the vehicle tested (or class of vehicle at best). This

lack of generality is an aspect of rotorcraft modelling that needs further attention.

Methods such as the free wake model can numerically calculate the wake induced

velocity in the flow field, and obtain aerodynamic interference, but their extreme time

cost hinders their utilisation for the flight dynamics modelling. This problem is amplified

for vehicles with multiple rotors or auxiliary propulsion devices where the time cost of
45

free wake and other vortex-based methods can incur unmanageable computational

overheads for flight dynamics research.

3) Engine/fuel system modelling has a significant influence on the flight dynamics

characteristics predicted, especially for large-amplitude manoeuvring flight. The change

of the required power leads to additional lag or overshooting effect produced by the

engine/fuel control system. However, this effect is usually neglected in the flight

dynamics modelling. At the moment, simplified transfer functions of the engine can be

utilised in flight dynamics analyses of the helicopter. Nevertheless, relevant experiments

are still required in these modelling methods in order to construct the engine/fuel system

models, and in particular, a generalised modelling method to simulate the engine effect

for various rotorcraft configurations is in urgent demand.

4) Rotorcraft dynamics can be considered as a set of highly coupled sub-systems. The

vehicle motion, the rotational motion of the rotor, and the wake motion are coupled with

the unsteady aerodynamic and inertia loading to ensure the precision. This structure puts

forward a higher requirement for the approach taken to solve the flight dynamics model.

The associated approach should not only consider the coupling effect among the

helicopter’s s stems, but also have the capability to allow each sub-component to

exchange data efficiently and with a relatively low time cost. Only with these

considerations, the flight dynamics model could strike a balance between precision and

time efficiency.

5) The manoeuvrability of the advanced helicopter has drawn a range of research interest

in recent years, and the NOC method and inverse simulation approach could be utilised

to investigate the manoeuvrability of various rotorcraft configurations with relatively low

time and financial cost. Also, the wavelet method can be used to obtain the handling

qualities rating from the calculated control inputs. However, due to the deficiency of the
46

NOC and inverse simulation methods, further improvements on the accuracy of the

manoeuvrability analysis method and the resultant handling qualities rating method are

still needed.
47

References

1. Blanken, Chris L., et al. Proposed Revisions to Aeronautical Design Standard-33E

(ADS-33E-PRF) Toward ADS-33F-PRF. CCDC AvMC Redstone Arsenal United States,

2019. Available at: https://apps.dtic.mil/sti/pdfs/AD1080657.pdf

2. Padfield, Gareth D. Helicopter Flight Dynamics: Including a Treatment of Tiltrotor

Aircraft. John Wiley & Sons, 2018, chapters 3, 4, and 10.

https://doi.org/10.1002/9781119401087

3. Johnson, Wayne. Helicopter theory. Courier Corporation, 2012, chapters 4 and 7.

Available at:

https://books.google.co.uk/books/about/Helicopter_Theory.html?id=SgZheyNeXJIC

4. Ghiringhelli, Gianluca, et al. "Multi-Body Analysis of the 1/5 Scale Wind Tunnel Model

of the V-22 Tiltrotor." 55th International Annual Forum of the American Helicopter

Society (AHS). 1999. Available at: https://vtol.org/store/product/multibody-analysis-of-

the-15-scale-wind-tunnel-model-of-the-v22-tiltrotor-4990.cfm

5. Manimala, Binoy, Gareth D. Padfield, and Daniel Walker. "Load alleviation for a tilt-

rotor aircraft in airplane mode." J. Aircr. 43 (1) (2006): 147-156.

https://doi.org/10.2514/1.13565

6. Keller, Jeffrey D., et al. "A Free Wake Linear Inflow Model Extraction Procedure for

Rotorcraft Analysis." American Helicopter Society 73rd Annual Forum, Fort Worth, TX.

2017. Available at: https://vtol.org/store/product/a-free-wake-linear-inflow-model-

extraction-procedure-for-rotorcraft-analysis-12111.cfm

7. Yanguo, Song, and Wang Huanjin. "Design of flight control system for a small

unmanned tilt rotor aircraft." Chin. J. Aeronaut. 22 (3) (2009): 250-256.

https://doi.org/10.1016/S1000-9361(08)60095-3

8. Blake, Bruce B., and Irvin B. Alansky. "Stability and Control of the YUH‐61A." J.
48

Am. Helicopter Soc 22 (1) (1977): 2-10. https://doi.org/10.4050/JAHS.22.1.2

9. Kisielowski, E., A. A. Perlmutter, and J. Tang. Stability and Control Handbook for

Helicopters. No. DCR-186. DYNASCIENCES CORP BLUE BELL PA, 1967. Available

at:

https://pra.org/publicdl/Engineering%20Design%20Papers/stability%20and%20control

%20handbook%20for%20helicopters%2067-63.pdf

10. Hess, Ronald A. "Analytical assessment of performance, handling qualities, and added

dynamics in rotorcraft flight control." IEEE Trans. Syst. Man Cybern. Paart A-Syst. Hum.

39 (1) (2008): 262-271. https://doi.org/10.1109/TSMCA.2008.2007943

11. Dowell, Earl H., and Deman Tang. "Nonlinear aeroelasticity and unsteady

aerodynamics." AIAA J. 40 (9) (2002): 1697-1707. https://doi.org/10.2514/2.1853

12. Sturisky, Selwyn H., et al. "Development and validation of a comprehensive real time

AH-64 Apache simulation model." Proceedings of the 48th Annual Forum of the

American Helicopter Society. Fairfax, Virgina: AHS, 1992: 1267-1280. Available at:

https://vtol.org/store/product/development-and-validation-of-a-comprehensive-real-

time-ah64-apache-simulation-model-850.cfm

13. Anderson, W. D., et al. REXOR Rotorcraft Simulation Model. Volume I. Engineering

Documentation. No. LR-27463-VOL-1. Lockheed-california co burbank, 1976.

Available at: https://apps.dtic.mil/sti/citations/ADA028314

14. Johnson, Wayne. "Assessment of aerodynamic and dynamic models in a comprehensive

analysis for rotorcraft." Comput. Math. Appl. 12 (1) (1986): 11-28.

https://doi.org/10.1016/0898-1221(86)90086-6

15. Kathryn, B. H. "A mathematical model of the UH-60 helicopter." NASA TM-85890

(1984). Available at: https://apps.dtic.mil/dtic/tr/fulltext/u2/a145899.pdf

16. Phillips, James D. "Mathematical model of the SH-3G helicopter." NASA-TM-84316


49

(1982). Available at: https://www.semanticscholar.org/paper/Mathematical-model-of-

the-SH-3G-helicopter-Phillips/b6ca1e9745c34f92684071a84c0ff36402726b5f

17. Weber, Jeanine M., Tung Y. Liu, and William Chung. "A mathematical simulation model

of the CH-47B helicopter, volume 2." NASA-TM-84351-VOL-2 (1984). Available at:

https://ntrs.nasa.gov/citations/19840024310

18. Chen, Robert TN. "Effects of primary rotor parameters on flapping dynamics." NASA-

TM-78575. (1980). Available at: https://ntrs.nasa.gov/citations/19800006879

19. Markley, F. Landis, et al. "UH-60 flight data replay and refly system state estimator

analysis." 28th Aerospace Sciences Meeting, 08 January 1990 - 11 January 1990, Reno,

NV, U.S.A. https://doi.org/10.2514/6.1990-181

20. Johnson, Wayne. "CAMRAD-a Comprehensive Analytical Model of Rotorcraft

Aerodynamics and Dynamics." AD-A0900513, (1994). Available at:

https://www.semanticscholar.org/paper/CAMRAD-A-COMPREHENSIVE-

ANALYTICAL-MODEL-OF-AND-

Johnson/7dd2caaae96d817b8dc02319dfdd607a2b498a5e

21. Lee, Bochan, and Moble Benedict. "Development and Validation of a Comprehensive

Helicopter Flight Dynamics Code." AIAA Scitech 2020 Forum, AIAA 2020-1644. 2020.

https://doi.org/10.2514/6.2020-1644

22. Sheridan, P., et al. "Mathematical modeling for helicopter simulation of low speed, low

altitude, and steeply descending flight." NASA contractor report, NASA-CR-166385

(1982). Available at: https://ntrs.nasa.gov/citations/19820024498

23. Howlett, John James. "UH-60A Black Hawk engineering simulation program. Volume

1: Mathematical model." NASA-CR-166309, (1981). Available at:

https://ntrl.ntis.gov/NTRL/dashboard/searchResults/titleDetail/N8428806.xhtml
50

24. Ballin, Mark G. "Validation of a real-time engineering simulation of the UH-60A

helicopter." NASA-TM-8360, (1987). Available at:

https://ntrs.nasa.gov/citations/19870008283

25. Sarathy, S., and Vadrevu Murthy. "An advanced rotorcraft flight simulation model-

Parallel implementation and performance analysis." AIAA-1993-3550, Flight

Simulation and Technologies, 09 August 1993 - 11 August 1993. Monterey, CA, U.S.A.

https://doi.org/10.2514/6.1993-3550

26. Kim, Frederick D., Roberto Celi, and Mark B. Tischler. "High ‐Order State Space

Simulation Models of Helicopter Flight Mechanics." J. Am. Helicopter Soc 38 (4 (1993):

16-27. https://doi.org/10.4050/JAHS.38.16

27. Bailey, James Earl, Ravivarma K. Prasanth, and Kalmanje Krishnakumar. ARMCOP

Helicopter Flight and Engine Model for the UH-1 TRS Simulator. University of Alabama,

College of Engineering, Bureau of Engineering Research, 1991. Available at:

https://www.worldcat.org/title/armcop-helicopter-flight-and-engine-model-for-the-uh-

1-trs-simulator/oclc/25517551

28. Heffley, Robert K., and Marc A. Mnich. "Minimum-complexity helicopter simulation

math model." NASA-CR-177476, (1988). Available at:

https://ntrs.nasa.gov/citations/19880020435

29. Talbot, Peter D. A mathematical force and moment model of a uh-1h helicopter for flight

dynamics simulations; NASA-TM-73254, 1977. Available at:

https://ntrs.nasa.gov/citations/19770024231

30. He, Chengjian, and WILLIAMD LEWIS. "A parametric study of real time mathematical

modeling incorporating dynamic wake and elastic blades." AHS, Annual Forum, 48 th,

Washington. 1992. Available at:

https://vtol.org/store/product/a-parametric-study-of-real-time-mathematical-modeling-
51

incorporating-dynamic-wake-and-elastic-blades-844.cfm

31. Lewis, Michael S. "A Piloted Simulation of One‐on‐One Helicopter Air Combat in

Low Level Flight." J. Am. Helicopter Soc 31 (2) (1986): 19-26.

https://doi.org/10.4050/JAHS.31.2.19

32. Chen, Robert TN, et al. "Helicopter mathematical models and control law development

for handling qualities research." NASA-CR-249, (1988). Available at:

https://ntrs.nasa.gov/citations/19880007259

33. Celi, R. "HeliUM 2 Flight Dynamic Simulation Model: Development, Technical

Concepts, and Applications." Proceedings of the 71st Annual Forum of the American

Helicopter Society. 2015. Available at:

https://vtol.org/store/product/helium-2-flight-dynamic-simulation-model-development-

technical-concepts-and-applications-10213.cfm

34. Ferguson, Samuel W. "A mathematical model for real time flight simulation of a generic

tilt-rotor aircraft." NASA CR-166536 (1988). Available at:

https://rotorcraft.arc.nasa.gov/Publications/files/CR-166536_882.pdf

35. Fegely, Cody, et al. "Flight dynamics and control modeling with system identification

lid tion o the Si ors X2 Te hnolo ™ Demonstr tor." Ameri n Heli opter

Society 72nd Annual Forum, West Palm Beach, FL. 2016. Available at:

https://vtol.org/store/product/flight-dynamics-and-control-modeling-with-system-

identification-validation-of-the-sikorsky-x2-technology-demonstrator-11500.cfm

36. Ferguson, Kevin, and Douglas Thomson. "Maneuverability Assessment of a Compound

Helicopter Configuration." J. Am. Helicopter Soc 61 (1) (2016): 1-15.

https://doi.org/10.4050/JAHS.61.012008

37. Ferguson, Kevin, and Douglas Thomson. "Examining the stability derivatives of a

compound helicopter." Aeronaut. J. 121 (1235) (2017): 1-20.


52

https://doi.org/10.1017/aer.2016.101

38. Yuan, Ye., et al. "Heading control strategy assessment for coaxial compound

helicopters." Chin. J. Aeronaut. 32 (9) (2019): 2037-2046.

https://doi.org/10.1016/j.cja.2019.04.008

39. Yuan, Ye, Douglas Thomson, and Renliang Chen. "Investigation of Lift Offset on Flight

Dynamics Characteristics for Coaxial Compound Helicopters." J. Aircr. 56 (6) (2019):

2210-2222. https://doi.org/10.1016/j.cja.2019.04.008

40. Yuan, Ye., D. Thomson, and R. Chen. "Variable rotor speed strategy for coaxial

compound helicopters with lift–offset rotors." Aeronaut. J. 124 (1271) (2020): 96-120.

https://doi.org/10.1017/aer.2019.113

41. Zhao, Xin, and H. C. Curtiss Jr. "A study of helicopter stability and control including

blade dynamics." NASA-CR-183245, (1988). Available at:

https://ntrs.nasa.gov/citations/19890001524

42. Takahashi, Marc D. "Rotor‐State Feedback in the Design of Flight Control Laws for a

Hovering Helicopter." J. Am. Helicopter Soc 39 (1) (1994): 50-62.

https://doi.org/10.4050/JAHS.39.50

43. Von Grunhagen, W. "Dynamic inflow modeling for helicopter rotors and its influence

on the prediction of cross-couplings." Proceedings of the AHS Aeromechanics

Specialists Conference, Bridgeport, CT. 1995. Available at:

https://vtol.org/store/product/dynamic-inflow-modelling-for-helicopter-rotors-and-its-

influence-on-the-prediction-of-crosscoupling-13272.cfm

44. Chaimovich, M., et al. "Investigation of the flight mechanics simulation of a hovering

helicopter." Proceedings of the 49th Annual Forum of the American Helicopter Society,

Fairfax Virginia: AHS, 1992: 1237-1256. Available at:

https://vtol.org/store/product/investigation-of-the-flight-mechanics-simulation-of-a-
53

hovering-helicopter-848.cfm

45. Spoldi, S., and P. Ruckel. "High Fidelity Helicopter Simulation using Free Wake, Lifting

Line Tail and Blade Element Tail Rotor Models." Annual Forum Proceedings-American

Helicopter Society. Vol. 59. No. 2. American Helicopter Society, INC, 2003. Available

at: https://vtol.org/store/product/high-fidelity-helicopter-simulation-using-free-wake-

lifting-line-tail-and-blade-element-tail-rotor-models-4186.cfm

46. Ji, Honglei, Renliang Chen, and Pan Li. "Real-time simulation model for helicopter flight

task analysis in turbulent atmospheric environment." Aero. Sci. Technol. 92 (2019): 289-

299. https://doi.org/10.1016/j.ast.2019.05.066

47. Horn, Joseph F., et al. "Implementation of a free-vortex wake model in real-time

simulation of rotorcraft." J. Aerosp. Comput. Inf. Commun. 3 (3) (2006): 93-107.

https://doi.org/10.2514/1.18273

48. Ribera, Maria. Helicopter flight dynamics simulation with a time-accurate free-vortex

wake model. Diss. University of Maryland, 2007. Available at:

https://drum.lib.umd.edu/handle/1903/6876

49. D Andrea, A. "Development of a multi-processor unstructured panel code coupled with

a CVC free wake model for advanced analyses of rotorcrafts and tiltrotors." Annual

Forum Proceedings-American Helicopter Society. Vol. 64. No. 1. American Helicopter

Society, INC, 2008. Available at: https://vtol.org/store/product/development-of-a-

multiprocessor-unstructured-panel-code-coupled-with-a-cvc-free-wake-model-for-

advanced-analyses-of-rotorcrafts-and-tiltrotors-3213.cfm

50. Ji, Honglei, Renliang Chen, and Pan Li. "Distributed Turbulence Model with Rigorous

Spatial Cross-Correlation for Simulation of Helicopter Flight in Atmospheric

Turbulence." J. Am. Helicopter Soc 64 (4) (2019): 1-13.

https://doi.org/10.4050/JAHS.64.042011
54

51. Pulla, Devi Prasad, and Albert Conlisk. "A lifting surface study of helicopter

aerodynamics in ground effect." AIAA 2007-1279, 45th AIAA aerospace sciences

meeting and exhibit, Reno, Nevada, 2007. https://doi.org/10.2514/6.2007-1279

52. Barra, F. "Development of a tilt-rotor model for real-time flight simulation." Proc. 15th

PEGASUS Student Conference (Glasgow, UK, April 2019). 2019. Available at:

https://www.pegasus-europe.org/wp-

content/uploads/Student_Conference/papers/2019/Paper_Barra.pdf

53. Gao, Han, and Ramesh K. Agarwal. "Numerical Study of a Hovering Helicopter Rotor

Blade in Ground Effect." AIAA 2019-1099, AIAA Scitech 2019 Forum. 2019, San Diego,

California, https://doi.org/10.2514/6.2019-1099

54. Lu, Yang, et al. "A method for optimizing the aerodynamic layout of a helicopter that

reduces the effects of aerodynamic interaction." Aero. Sci. Technol. 88 (2019): 73-83.

https://doi.org/10.1016/j.ast.2019.03.005

55. Aydemir, Hakan, and Ugur Zengin. "Real-time Simulation Infrastructure for Model-

based Design of Helicopter Flight Control System.", AIAA 2018-0124, 2018 AIAA

Modeling and Simulation Technologies Conference, Kissimmee, Florida, 2018.

https://doi.org/10.2514/6.2018-0124

56. Lewis, Richard B. "Hueycobra Maneuvering Investigations." 26th American Helicopter

Society, Annual National Forum, Washington, DC. 1970. Available at:

https://vtol.org/store/product/hucycobra-maneuvering-investigations-2982.cfm

57. Lewis, I. I., et al. Engineering Flight Test AH-1G Helicopter (HUEYCOBRA).

Maneuvering Limitations. No. USAASTA-69-11. Army Aviation Systems Test Activity

Edwards AFB CA, 1971. Available at:

https://apps.dtic.mil/dtic/tr/fulltext/u2/855629.pdf

58. Yamakawa, George M., Donald G. Broadhurst, and John R. Smith. Utility tactical
55

transport aircraft system (UTTAS) maneuver criteria. No. USAASTA-71-32. Army

Aviation Systems Test Activity Edwards AFB CA, 1972. Available at:

https://apps.dtic.mil/dtic/tr/fulltext/u2/902767.pdf

59. Wood, T. L., and C. L. Livingston. An energy method for prediction of helicopter

maneuverability. No. BHC-TR-299-099-557. Bell Helicopter Textron INC FORT

WORTH TX, 1971. Available at: https://apps.dtic.mil/dtic/tr/fulltext/u2/a021266.pdf

60. Wells, C. D., and T. L. Wood. "Maneuverability ‐Theory and Application." J. Am.

Helicopter Soc 18 (1) (1973): 10-22. https://doi.org/10.4050/JAHS.18.1.10

61. Wood, T. L., D. G. Ford, and G. H. Brigman. Maneuver Criteria Evaluation Program.

BELL HELICOPTER TEXTRON INC FORT WORTH TX, AD-782207, 1974.

Available at: https://apps.dtic.mil/dtic/tr/fulltext/u2/782209.pdf

62. Wood, T., and T. Waak. Improved Maneuver Criteria Evaluation Program. Bell

Helicopter Textron INC FORT WORTH TX, 1979. Available at:

https://apps.dtic.mil/dtic/tr/fulltext/u2/a080408.pdf

63. Thomson, Douglas, and Roy Bradley. "Inverse simulation as a tool for flight dynamics

research—Principles and applications." Prog. Aeosp. Sci. 42 (3) (2006): 174-210.

https://doi.org/10.1016/j.paerosci.2006.07.002

64. Rutherford, Stephen, and Douglas G. Thomson. "Improved methodology for inverse

simulation." Aeronaut. J. 100. (993) (1996): 79-86.

https://doi.org/10.1017/S0001924000067348

65. Rutherford, Stephen, and Douglas G. Thomson. "Helicopter inverse simulation

incorporating an individual blade rotor model." J. Aircr. 34 (5) (1997): 627-634.

https://doi.org/10.2514/2.2239

66. Eversman, Walter. "A reduced cost rational-function approximation for unsteady

aerodynamics." AIAA-90-1155-CP, 31st Structures, Structural Dynamics and Materials


56

Conference, Long Beach, CA, U.S.A. 1990. https://doi.org/10.2514/6.1990-1155

67. Tyler, Joseph C., and J. Gordon Leishman. "Analysis of pitch and plunge effects on

unsteady airfoil behavior." J. Am. Helicopter Soc 37 (3) (1992): 69-82.

https://doi.org/10.4050/JAHS.37.69

68. Yen, Jing G., and Mithat Yuce. "Correlation of Pitch‐Link Loads in Deep Stall on

Bearingless Rotors." J. Am. Helicopter Soc 37 (4) (1992): 4-15.

https://doi.org/10.4050/JAHS.37.4

69. Mello, Olympio AF, and Omri Rand. "Unsteady, Frequency ‐ Domain Analysis of

Helicopter Non‐Rotating Lifting Surfaces." J. Am. Helicopter Soc 36 (2) (1991): 70-81.

https://doi.org/10.4050/JAHS.36.70

70. Shih, C., et al. "Unsteady flow past an airfoil pitching at a constant rate." AIAA J. 30 (5)

(1992): 1153-1161. https://doi.org/10.2514/3.11045

71. Beddoes, T. S. "A synthesis of unsteady aerodynamic effects including stall hysteresis."

Paper 17, 1st European Rotorcraft Forum, Southampton, UK. Available at:

https://dspace-erf.nlr.nl/xmlui/handle/20.500.11881/2079

72. AZUMA, AKIRA, and AKIRA OBATA. "Induced flow variation of the helicopter rotor

operating in the vortex ring state." J. Aircr. 5(4) (1968): 381-386.

https://doi.org/10.2514/3.43954

73. Bragg, Michael B., Douglas C. Heinrich, and Abdollah Khodadoust. "Low-frequency

flow oscillation over airfoils near stall." AIAA J. 31 (7) (1993): 1341-1343.

https://doi.org/10.2514/3.49069

74. Rosti, Marco E., Mohammad Omidyeganeh, and Alfredo Pinelli. "Numerical Simulation

of a Passive Control of the Flow Around an Aerofoil Using a Flexible, Self Adaptive

Flaplet." Flow Turbul. Combust. 100 (4) (2018): 1111-1143.

https://doi.org/10.1007/s10494-018-9914-6
57

75. Mai, Holger, et al. "Dynamic stall control by leading edge vortex generators." J. Am.

Helicopter Soc 53 (1) (2008): 26-36. https://doi.org/10.4050/JAHS.53.26

76. Le Pape, Arnaud, et al. "Dynamic stall control using deployable leading-edge vortex

generators." AIAA J. 50 (10) (2012): 2135-2145. https://doi.org/10.2514/1.J051452

77. Spentzos, Agis, et al. "Investigation of three-dimensional dynamic stall using

computational fluid dynamics." AIAA J. 43 (5) (2005): 1023-1033.

https://doi.org/10.2514/1.8830

78. Spentzos, Agis, et al. "Computational fluid dynamics study of three-dimensional

dynamic stall of various planform shapes." J. Aircr. 44 (4) (2007): 1118-1128.

https://doi.org/10.2514/1.24331

79. Wang, Shengyi, et al. "Numerical investigations on dynamic stall of low Reynolds

number flow around oscillating airfoils." Comput. Fluids. 39 (9) (2010): 1529-1541.

https://doi.org/10.1016/j.compfluid.2010.05.004

80. Leishman, J. G., and T. S. Beddoes. "A generalised model for airfoil unsteady

aerodynamic behaviour and dynamic stall using the indicial method." Proceedings of the

42nd Annual forum of the American Helicopter Society. Washington DC, 1986.

Available at: https://vtol.org/store/product/a-generalised-model-for-airfoil-unsteady-

aerodynamic-behaviour-and-dynamic-stall-using-the-indicial-method-1381.cfm

81. Leishman, J. G. "Modeling sweep effects on dynamic stall." J. Am. Helicopter Soc 34 (3)

(1989): 18-29. https://doi.org/10.4050/JAHS.34.3.18

82. Leishman, J. Gordon, and T. S. Beddoes. "A Semi‐Empirical model for dynamic stall."

J. Am. Helicopter Soc 34 (3) (1989): 3-17. https://doi.org/10.4050/JAHS.34.3.3

83. Truong, V. K. "A 2-d dynamic stall model based on a hopf bifurcation." Proceedings of

the 19th European Rotorcraft Forum. Cernobbio: ERF, 1993: 23. (1993). Available at:

https://dspace-erf.nlr.nl/xmlui/handle/20.500.11881/2281
58

84. Ortun, Biel, et al. "Rotor loads prediction on the ONERA 7A rotor using loose

fluid/structure coupling." J. Am. Helicopter Soc 62 (3) (2017): 1-13.

https://doi.org/10.4050/JAHS.62.032005

85. Truong, Khiem-Van, Hyeonsoo Yeo, and Robert A. Ormiston. "Structural dynamics

modeling of rectangular rotor blades." Aero. Sci. Technol. 30 (1) (2013): 293-305.

https://doi.org/10.1016/j.ast.2013.08.014

86. Yeo, Hyeonsoo, Khiem-Van Truong, and Robert A. Ormiston. "Comparison of one-

dimensional and three-dimensional structural dynamics modeling of advanced geometry

blades." J. Aircr. 51 (1) (2014): 226-235. https://doi.org/10.2514/1.C032304

87. Massaro, Andrea, and Ernesto Benini. "Multi-objective optimization of helicopter

airfoils using surrogate-assisted memetic algorithms." J. Aircr. 49 (2) (2012): 375-383.

https://doi.org/10.2514/1.C001017

88. Won, Y. S., et al. "Aerodynamic performance evaluation of basic airfoils for an

agricultural unmanned helicopter using wind tunnel test and CFD simulation." J. Mech.

Sci. Technol. 31 (12) (2017): 5829-5838. https://doi.org/10.1007/s12206-017-1125-x

89. Forrest, J. S., C. H. Kaaria, and I. Owen. "Evaluating ship superstructure aerodynamics

for maritime helicopter operations through CFD and flight simulation." Aeronaut. J. 120

(1232) (2016): 1578-1603. https://doi.org/10.1017/aer.2016.76

90. Guntur, S.; Sørensen, N.N.; Schreck, S.; Bergami, L. Modeling dynamic stall on wind

turbine blades under rotationally augmented flow fields. Wind Energy 19, 2016, 383–

397. https://doi.org/10.1002/we.1839

91. Breton, S.P.; Coton, F.N.; Moe, G. A Study on Rotational Effects and Different Stall

Delay Models Using a Prescribed Wake Vortex Scheme and NREL Phase VI Experiment

Data. Wind Energy 11, 2008, 459–482. https://doi.org/10.1002/we.269


59

92. Chen, Robert TN. "A survey of nonuniform inflow models for rotorcraft flight dynamics

and control applications." NASA-TM-102219, (1989). Available at:

https://core.ac.uk/download/pdf/42825131.pdf

93. Coleman, Robert P., Arnold M. Feingold, and Carl W. Stempin. Evaluation of the

induced-velocity field of an idealized helicopter rotor. No. NACA-WR-L-126. NASA,

1945. Available at: https://apps.dtic.mil/dtic/tr/fulltext/u2/a801123.pdf

94. Carpenter, Paul J., and Bernard Fridovich. "Effect of a rapid blade-pitch increase on the

thrust and induced-velocity response of a full-scale helicopter rotor." NACA-TN-3044.

(1953). Available at: https://ntrs.nasa.gov/citations/19930083686

95. Pitt, Dale M., and David A. Peters. "Theoretical prediction of dynamic-inflow

derivatives." (1980) 16th European Rotorcraft and Powered Lift Aircraft Forum, Bristol,

England, Sep 16-19, 1980. Available at: https://dspace-

erf.nlr.nl/xmlui/handle/20.500.11881/1796

96. Peters, David A., and Cheng Jian He. "Finite state induced flow models. II-Three-

dimensional rotor disk." J. Aircr. 32 (2) (1995): 323-333.

https://doi.org/10.2514/3.46719

97. Peters, David A. "How dynamic inflow survives in the competitive world of rotorcraft

aerodynamics." J. Am. Helicopter Soc 54 (1) (2009): 11001-11001.

https://doi.org/10.4050/JAHS.54.011001

98. Fei, Zhongyang, and David A. Peters. "Applications and data of generalised dynamic

wake theory of the flow in a rotor wake." IET Contr. Theory Appl.9 (7) (2015): 1051-

1057. https://doi.org/10.1049/iet-cta.2014.0710

99. Huang, Jianzhe, and David Peters. "Real-time solution of nonlinear potential flow

equations for lifting rotors." Chin. J. Aeronaut. 30 (3) (2017): 871-880.

https://doi.org/10.1016/j.cja.2017.02.007
60

100. Peters, David A. "Two-dimensional incompressible unsteady airfoil theory—an

overview." J. Fluids Struct. 24 (3) (2008): 295-312.

https://doi.org/10.1016/j.jfluidstructs.2007.09.001

101. Hong, JunSoo, David A. Peters, and Robert A. Ormiston. "A dynamic-inflow-based

induced power model for general and optimal rotor performance." J. Am. Helicopter Soc

63 (2) (2018): 1-11. https://doi.org/10.4050/JAHS.63.022008

102. Barocela, Ed, et al. "The effect of wake distortion on rotor inflow gradients and off-

axis coupling." AIAA-97-3579, 22nd Atmospheric Flight Mechanics Conference, New

Orleans, LA, U.S.A. 1997. https://doi.org/10.2514/6.1997-3579

103. Krothapalli, Krishnamohan R., J. V. R. Prasad, and David A. Peters. "Helicopter rotor

dynamic inflow modeling for maneuvering flight." J. Am. Helicopter Soc 46 (2) (2001):

129-139. https://doi.org/10.4050/JAHS.46.129

104. Zhao, Jinggen. Dynamic wake distortion model for helicopter maneuvering flight. Diss.

Georgia Institute of Technology, 2005. Available at:

https://pdfs.semanticscholar.org/1796/cdf2d690b36af9e52da91a5a89b56b9ad493.pdf

105. Raz, Reuben, Aviv Rosen, and Tuvia Ronen. "Active aerodynamic stabilization of a

helicopter/sling-load system." J. Aircr. 26 (9) (1989): 822-828.

https://doi.org/10.2514/3.45847

106. Rosen, Aviv, and Aharon Isser. "A new model of rotor dynamics during pitch and roll

of a hovering helicopter." J. Am. Helicopter Soc 40 (3) (1995): 17-28.

https://doi.org/10.4050/JAHS.40.17

107. Keller, J. D., and H. C. Curtiss. "A critical examination of the methods to improve the

off-axis response prediction of helicopters." Annual Forum Proceedings-American

Helicopter Society. Vol. 54. American Helicopter Society, 1998. Available at:
61

https://vtol.org/store/product/a-critical-examination-of-the-methods-to-improve-the-

offaxis-response-prediction-of-helicopters-4650.cfm

108. Keller, Jeffrey D. "An investigation of helicopter dynamic coupling using an analytical

model." J. Am. Helicopter Soc, 41 (4) (1996): 322-330.

https://doi.org/10.4050/JAHS.41.322

109. Bhagwat, M. J. "Mathematical modeling of the transient dynamics of helicopter rotor

wakes using a time-accurate free-vortex method." PhD dissertaion, University of

Maryland at College Park, MD (2001). Available at:

https://drum.lib.umd.edu/handle/1903/25270

110. i, Ashish, nd J. Gordon Leishm n. "Rotor Free‐W e Modelin Usin seudo‐

Implicit Technique—Including Comparisons with Experimental Data." J. Am.

Helicopter Soc. 40 (3) (1995): 29-41. https://doi.org/10.4050/JAHS.40.29

111. Chung, K. H., et al. "A Study on Rotor Tip-Vortex Pairing Phenomena by using Time-

Marching Free-Wake Method." annual forum proceedings-american helicopter society.

Vol. 56. No. 1. American helicopter society, inc, 2000. Available at:

https://vtol.org/store/product/a-study-on-rotor-tipvortex-pairing-phenomena-by-using-

timemarching-freewake-method-4771.cfm

112. Lee, Duck Joo, and Seon Uk Na. "Numerical simulations of wake structure generated

by rotating blades using a time marching, free vortex blob method." Eur. J. Mech. B-

Fluids 18 (1) (1999): 147-159. https://doi.org/10.1016/S0997-7546(99)80011-9

113. Yuan, Ye, Renliang Chen, and Pan Li. "Trim investigation for coaxial rigid rotor

helicopters using an improved aerodynamic interference model." Aero. Sci. Technol. 85

(2019): 293-304. https://doi.org/10.1016/j.ast.2018.11.044


62

114. Greengard, Leslie, and Vladimir Rokhlin. "A fast algorithm for particle simulations."

J. Comput. Phys. 135 (2) (1997): 280-292. https://doi.org/10.1016/0021-9991(87)90140-

115. Brown, Richard E. "Rotor wake modeling for flight dynamic simulation of helicopters."

AIAA J. 38 (1) (2000): 57-63. https://doi.org/10.2514/2.922

116. Brown, Richard E., and Andrew J. Line. "Efficient high-resolution wake modeling

using the vorticity transport equation." AIAA J. 43 (7) (2005): 1434-1443.

https://doi.org/10.2514/1.13679

117. He, Chengjian, and Jinggen Zhao. "Modeling rotor wake dynamics with viscous vortex

particle method." AIAA J. 47 (4) (2009): 902-915. https://doi.org/10.2514/1.36466

118. Zhao, Jinggen, and Chengjian He. "A viscous vortex particle model for rotor wake and

interference analysis." J. Am. Helicopter Soc. 55 (1) (2010): 12007-12007.

https://doi.org/10.4050/JAHS.55.012007

119. Guntur, Srinivas, and Niels N. Sørensen. "A study on rotational augmentation using

CFD analysis of flow in the inboard region of the MEXICO rotor blades." Wind

Energy 18 (4) (2015): 745-756. https://doi.org/10.1002/we.1726

120. CAO, Yi-hua, et al. "Combined free wake/CFD methodology for predicting transonic

rotor flow in hover." Chin. J. Aeronaut. 15 (2) (2002): 65-71.

https://doi.org/10.1016/S1000-9361(11)60132-5

121. Yongjie, Shi, et al. "A new single-blade based hybrid CFD method for hovering and

forward-flight rotor computation." Chin. J. Aeronaut. 24 (2) (2011): 127-135.

https://doi.org/10.1016/S1000-9361(11)60016-2

122. Kim, Hyo Won, et al. "Interactional aerodynamics and acoustics of a hingeless coaxial

helicopter with an auxiliary propeller in forward flight." Aeronaut. J. 113 (1140) (2009):

65-78. https://doi.org/10.1017/S0001924000002797
63

123. Feil, Roland, et al. "Aeromechanics analysis of a high-advance-ratio lift-offset coaxial

rotor system." J. Aircr. 56 (1) (2019): 166-178.

https://doi.org/10.2514/1.C034748

124. Schmaus, Joseph H., and Inderjit Chopra. "Aeromechanics of rigid coaxial rotor models

for wind-tunnel testing." J. Aircr. 54 (4) (2017): 1486-1497.

https://doi.org/10.2514/1.C034157

125. Ruddell, Andre J. "Ad n in l de on ept (A C™) de elopment." J. Am.

Helicopter Soc 22 (1) (1977): 13-23. https://doi.org/10.4050/JAHS.22.1.13

126. Walsh, D., et al. "High airspeed testing of the sikorsky x2 technologytm demonstrator."

American Helicopter Society 67th Annual Forum, Virginia Beach, VA. 2011. Available

at: https://vtol.org/store/product/high-airspeed-testing-of-the-sikorsky-x2-technology-

tm-demonstrator-5325.cfm

127. Padfield, Gareth D., Victoria Brookes, and Michael A. Meyer. "Progress in Civil Tilt‐

Rotor Handling Qualities." J. Am. Helicopter Soc 51 (1) (2006): 80-91.

https://doi.org/10.4050/1.3092880

128. Leishman, Gordon J. Principles of helicopter aerodynamics with CD extra. Cambridge

university press, 2006. Chapters 2 and 3. Available at:

https://books.google.co.uk/books/about/Principles_of_Helicopter_Aerodynamics_wi.ht

ml?id=nMV-TkaX-9cC&redir_esc=y

129. Du Val, R. "A real-time blade element helicopter simulation for handling." ERF-1989-

59, Proceedings of the 15th European Rotorcraft Forum, Amsterdam: ERF, 1989: 766-

785. Available at:

https://dspace-erf.nlr.nl/xmlui/bitstream/handle/20.500.11881/2653/ERF%201989-

59.pdf?sequence=1
64

130. He, Cheng-Jian, and Ronald Du Val. "An unsteady airload model with dynamic stall

for rotorcraft simulation." AHS, Annual Forum, 50 th, Washington, DC. 1994. Available

at: https://vtol.org/store/product/an-unsteady-airload-model-with-dynamic-stall-for-

rotorcraft-simulation-561.cfm

131. Turnour, Stephen R., and Roberto Celi. "Modeling of flexible rotor blades for helicopter

flight dynamics applications." J. Am. Helicopter Soc 41 (1) (1996): 52-66.

https://doi.org/10.4050/JAHS.41.52

132. Zhao, J., and C. He. "Rotor Blade Structural Loads Analysis Using Coupled

CSD/CFD/VVPM." Proceedings of the American Helicopter Society 69th Annual

Forum. 2013. Available at: https://vtol.org/store/product/rotor-blade-structural-loads-

analysis-using-coupled-csdcfdvvpm-8813.cfm

133. Pan, Li, and Chen Renliang. "A mathematical model for helicopter comprehensive

analysis." Chin. J. Aeronaut. 23 (3) (2010): 320-326.

https://doi.org/10.1016/S1000-9361(09)60222-3

134. Yeo, Hyeonsoo, William G. Bousman, and Wayne Johnson. "Performance analysis of

a utility helicopter with standard and advanced rotors." J. Am. Helicopter Soc 49 (3)

(2004): 250-270. https://doi.org/10.4050/JAHS.49.250

135. Curtiss, H. C., and T. R. Quackenbush. "The influence of the rotor wake on rotorcraft

stability and control." ERF-1989-70, Fifteenth European Rotorcraft Forum, Sep 12-15,

Amsterdam. (1989).

https://dspace-erf.nlr.nl/xmlui/bitstream/handle/20.500.11881/2643/ERF%201989-

70.pdf?sequence=1

136. Kenyon, Adam R., and Richard E. Brown. "Wake Dynamics and Rotor ‐Fuselage

Aerodynamic Interactions." J. Am. Helicopter Soc 54 (1) (2009): 12003-12003.

https://doi.org/10.4050/JAHS.54.012003
65

137. SCARPATI, T., R. FEENAN, and W. STRATTON. "The results of fabrication and

testing of the prototype composite rotor blades for HLH and UTTAS." Aircraft

Systems and Technology Meeting, Los Angeles, CA, U.S.A. 1975.

https://doi.org/10.2514/6.1975-1010

138. Smith, Charles A., and Mark D. Betzina. "Aerodynamic loads induced by a rotor on a

body of revolution." J. Am. Helicopter Soc 31 (1) (1986): 29-36.

https://doi.org/10.4050/JAHS.31.1.29

139. Lorber, Peter F., and T. Alan Egolf. "An Unsteady Helicopter Rotor ‐ Fuselage

Aerodynamic Interaction Analysis." J. Am. Helicopter Soc 35 (3) (1990): 32-42.

https://doi.org/10.4050/JAHS.35.32

140. Crouse, Gilbert L., J. Gordon Leishman, and Naipei Bi. "Theoretical and experimental

study of unsteady rotor/body aerodynamic interactions." J J. Am. Helicopter Soc 37 (1)

(1992): 55-65. https://doi.org/10.4050/JAHS.37.55

141. Mavris, Dimitris N., Narayanan M. Komerath, and Howard M. McMahon. "Prediction

of Aerodynamic Rotor‐Airframe Interactions in Forward Flight." J. Am. Helicopter Soc

34 (4) (1989): 37-46. https://doi.org/10.4050/JAHS.34.37

142. Komerath, N. M., D. M. Mavris, and S. G. Liou. "Prediction of unsteady pressure and

velocity over a rotorcraft in forward flight." J. Aircr. 28 (8) (1991): 509-516.

https://doi.org/10.2514/3.46056

143. Quackenbush, T. R., C ‐ MG Lam, and D. B. Bliss. "Vortex methods for the

computational analysis of rotor/body interaction." J. Am. Helicopter Soc 39 (4) (1994):

14-24. https://doi.org/10.4050/JAHS.39.14

144. Affes, H., et al. "The three-dimensional boundary layer flow due to a rotor-tip vortex."

23rd Fluid Dynamics, Plasmadynamics, and Lasers Conference, Orlando, FL, U.S.A.

1993. https://doi.org/10.2514/6.1993-3081
66

145. Berry, John D., and Susan L. Althoff. "Inflow velocity perturbations due to fuselage

effects in the presence of a fully interactive wake." 46th AHS, Annual Forum; May 21,

1990 - May 23, 1990; Washington, DC; United State, (1990). Available at:

https://vtol.org/store/product/inflow-velocity-perturbations-due-to-fuselage-effects-in-

the-presence-of-a-fully-interactive-wake-995.cfm

146. Norman, Thomas R., and Gloria K. Yamauchi. "Full-scale investigation of

aerodynamic interactions between a rotor and fuselage." 47th AHS Annual Forum; May

06, 1991 - May 08, 1991; Phoenix, AZ; United States (1991). Available at:

https://vtol.org/store/product/fullscale-investigation-of-aerodynamic-interactions-

between-a-rotor-and-fuselage-866.cfm

147. Liou, S. G., N. M. Komerath, and H. M. McMahon. "Velocity measurements of

airframe effects on a rotor in a low-speed forward flight." J. Aircr. 26 (4) (1989): 340-

348. https://doi.org/10.2514/3.45766

148. Liou, S. G., N. M. Komerath, and H. M. McMahon. "Measurement of the interaction

between a rotor tip vortex and a cylinder." AIAA J. 28 (6) (1990): 975-981.

https://doi.org/10.2514/3.25153

149. Brand, A. G., H. M. McMahon, and N. M. Komerath. "Surface pressure measurements

on a body subject to vortex wake interaction." AIAA J. 27 (5) (1989): 569-574.

https://doi.org/10.2514/3.10147

150. Leishman, J. G., and N. P. Bi. "Measurements of a rotor flowfield and the effects on a

fuselage in forward flight." 16th European Rotorcraft Forum, Glasgow, Scotland, 18-21

Sept. 1990. Available at:

https://dspace-erf.nlr.nl/xmlui/bitstream/handle/20.500.11881/2572/ERF1990-Vol2-II-

11-1.pdf?sequence=1
67

151. Bi, Nai-pei, J. Gordon Leishman, and Gilbert L. Crouse Jr. "Investigation of rotor tip

vortex interactions with a body." J. Aircr. 30 (6) (1993): 879-888.

https://doi.org/10.2514/3.46430

152. Bhagwat, Mahendra J., and J. Gordon Leishman. "Correlation of helicopter rotor tip

vortex measurements." AIAA J. 38 (2) (2000): 301-308.

https://doi.org/10.2514/2.957

153. Wachspress, Daniel A., Todd R. Quackenbush, and Alexander H. Boschitsch.

"Rotorcraft interactional aerodynamics with fast vortex/fast panel methods." J. Am.

Helicopter Soc. 48 (4) (2003): 223-235. https://doi.org/10.4050/JAHS.48.223

154. Yeo, Hyeonsoo, and Wayne Johnson. "Performance and design investigation of heavy

lift tilt-rotor with aerodynamic interference effects." J. Aircr. 46 (4) (2009): 1231-1239.

https://doi.org/10.2514/1.40102

155. Jung, Y. S., You, J. Y., & Kwon, O. J. (2014). Numerical investigation of prop-rotor

and tail-wing aerodynamic interference for a tilt-rotor UAV configuration. J. Mech. Sci.

Technol., 28(7), 2609–2617. https://doi.org/10.1007/s12206-014-0617-1

156. Di Francesco, G., & Mattei, M. (2016). Modeling and Incremental Nonlinear Dynamic

Inversion Control of a Novel Unmanned Tiltrotor. J. Aircr., 53(1), 73–86.

https://doi.org/10.2514/1.C033183

157. Lu, K., Liu, C., Li, C., & Chen, R. (2019). Flight Dynamics Modeling and Dynamic

Stability Analysis of Tilt-Rotor Aircraft. Int. J. Aerosp. Eng., 2019, 1–15.

https://doi.org/10.1155/2019/5737212

158. YAN, X., & CHEN, R. (2019). Augmented flight dynamics model for pilot workload

evaluation in tilt-rotor aircraft optimal landing procedure after one engine failure. Chin.

J. Aeronaut., 32(1), 92–103. https://doi.org/10.1016/j.cja.2018.06.010

159. Yeo, H eonsoo. 2019. “Desi n nd Aerome h ni s In esti tion o Compound


68

Heli opters.” Aerosp. Sci. Technol. 88, 2019: 158–73.

https://doi.org/10.1016/j.ast.2019.03.010.

160. Stokkermans, T., Veldhuis, L., Soemarwoto, B., Fukari, R., & Eglin, P. (2020).

Breakdown of aerodynamic interactions for the lateral rotors on a compound helicopter.

Aerosp. Sci. Technol, 101, 105845. https://doi.org/10.1016/j.ast.2020.105845

161. Neighbors III, W. Kenneth, and Stephen M. Rock. "Integrated flight/propulsion

control-subsystem specifications for performance." J. Guid. Control Dyn.18 (3) (1995):

572-578. https://doi.org/10.2514/3.21425

162. Rock, Stephen M., and Ken Neighbors. "Integrated flight/propulsion control for

helicopters." J. Am. Helicopter Soc 39 (3) (1994): 34-42.

https://doi.org/10.4050/JAHS.39.34

163. Ballin, Mark G. "A high fidelity real-time simulation of a small turboshaft engine."

NASA-TM-100991, (1988). Available at: https://ntrs.nasa.gov/citations/19880016994

164. Duyar, Ahmet, Zhen Gu, and Jonathan S. Litt. "A simplified dynamic model of the

T700 turboshaft engine." J. Am. Helicopter Soc 40 (4) (1995): 62-70.

https://doi.org/10.4050/JAHS.40.62

165. Chen, Robert TN. "A simplified rotor system mathematical model for piloted flight

dynamics simulation." NASA-TM-78575, (1979). Available at:

https://ntrs.nasa.gov/citations/19790015806

166. Talbot, Peter D., et al. "A mathematical model of a single main rotor helicopter for

piloted simulation." NASA-TM-84281, (1982). Available at:

https://ntrs.nasa.gov/citations/19830001781

167. Kaplita, Thaddeus T. "UH-60 Black Hawk engineering simulation model validation

and proposed modifications." NASA-CR-177360, (1986). Available at:

https://ntrs.nasa.gov/citations/19870008277
69

168. KIM, FREDERICK. "Analysis of propulsion system dynamics in the validation of a

high-order state space model of the UH-60." Flight Simulation Technologies Conference,

Hilton Head Island, SC, U.S.A. 1992. https://doi.org/10.2514/6.1992-4150

169. Conner, David A., et al. "Xv-15 tiltrotor low noise terminal area operations."

20040110255, NASA (1998). Available at: https://ntrs.nasa.gov/citations/20040110255

170. Aydin, Hakan, et al. "Component–based exergetic measures of an experimental

turboprop/turboshaft engine for propeller aircrafts and helicopters." Int. J. Exergy 11 (3)

(2012): 322-348. https://doi.org/10.1504/IJEX.2012.050228

171. Enconniere, Julien, Jesús Ortiz-Carretero, and Vassilios Pachidis. "Mission

optimisation for a conceptual coaxial rotorcraft for taxi applications." Aero. Sci. Technol.

72 (2018): 14-24. https://doi.org/10.1016/j.ast.2017.10.031

172. Enconniere, Julien, Jesus Ortiz-Carretero, and Vassilios Pachidis. "Mission

performance analysis of a conceptual coaxial rotorcraft for air taxi applications." Aero.

Sci. Technol. 69 (2017): 1-14. https://doi.org/10.1016/j.ast.2017.06.015

173. Theodore, Colin, and Roberto Celi. "Helicopter flight dynamic simulation with refined

aerodynamics and flexible blade modeling." J. Aircr. 39 (4) (2002): 577-586.

https://doi.org/10.2514/2.2995

174. Bagai, Ashish, and J. Gordon Leishman. "Rotor free-wake modeling using a

pseudoimplicit relaxation algorithm." J. Aircr. 32 (6) (1995): 1276-1285.

https://doi.org/10.2514/3.46875

175. Truong, V. K. "An analytical model for airfoil aerodynamic characteristics over the

entire 360° angle of attack range." J. Renew. Sustain. Energy 12 (3) (2020): 033303.

https://doi.org/10.1063/1.5126055

176. Truong, Khiem Van. "Modeling aerodynamics, including dynamic stall, for

comprehensive analysis of helicopter rotors." Aerospace 4.2 (2017): 21.


70

https://doi.org/10.3390/aerospace4020021

177. Ivler, Christina M., et al. "Design and flight test of a cable angle feedback flight control

system for the RASCAL JUH-60 helicopter." J. Am. Helicopter Soc 59 (4) (2014): 1-15.

https://doi.org/10.4050/JAHS.59.042008

178. Blanken, Chris L., and Heinz‐Ju Pausder. "Investigation of the Effects of Bandwidth

and Time Delay on Helicopter Roll‐Axis Handling Qualities." J. Am. Helicopter Soc 39

(3) (1994): 24-33. https://doi.org/10.4050/JAHS.39.3.24

179. Welsh, W., et al. "Flight test of an active vibration control system on the UH-60 black

hawk helicopter." AHS, Annual Forum, 51 st, Fort Worth, TX. 1995. Available at:

https://vtol.org/store/product/flight-test-of-an-active-vibration-control-sytem-on-the-

uh60-black-hawk-helicopter-407.cfm

180. Hutto, A. J. "Flight‐Test Report on the Heavy‐Lift Helicopter Flight‐Control System."

J. Am. Helicopter Soc 21 (1) (1976): 32-40. https://doi.org/10.4050/JAHS.21.1.32

181. Kääriä, Christopher H., et al. "An experimental technique for evaluating the

aerodynamic impact of ship superstructures on helicopter operations." Ocean Eng. 61

(2013): 97-108.4 (2013): 663-686. https://doi.org/10.1016/j.oceaneng.2012.12.052

182. Klyde, David H., et al. "Piloted simulation evaluation of tracking MTEs for the

assessment of high-speed handling qualities." 74th Annual Forum Proceedings-AHS

International. Phoenix, AZ, May 14–17, 2018. Available at:

https://vtol.org/store/product/piloted-simulation-evaluation-of-tracking-mtes-for-the-

assessment-of-highspeed-handling-qualities-12779.cfm

183. Ferguson, Kevin M. Towards a better understanding of the flight mechanics of

compound helicopter configurations. Diss. University of Glasgow, 2015. Available at:

http://theses.gla.ac.uk/6859/
71

184. Padfield, Gareth D., and Mark D. White. "Flight simulation in academia HELIFLIGHT

in its first year of operation at the University of Liverpool." Aeronaut. J. 107 (1075)

(2003): 529-538. https://doi.org/10.1017/S0001924000013415

185. White, Mark D., et al. "Acceptance testing and commissioning of a flight simulator for

rotorcraft simulation fidelity research." Proc. Inst. Mech. Eng. Part G-J. Aerosp. Eng.

227 (4) (2013): 663-686. https://doi.org/10.1177/0954410012439816

186. Celi, Roberto. "Optimization-based inverse simulation of a helicopter slalom

maneuver." J. Guid. Control Dyn. 23 (2) (2000): 289-297. https://doi.org/10.2514/2.4521

187. Johnson, Eric N., and Suresh K. Kannan. "Adaptive trajectory control for autonomous

helicopters." J. Guid. Control Dyn. 28 (3) (2005): 524-538.

https://doi.org/10.1007/s11768-015-4062-1

188. Chi, Cheng, et al. "Analysis of low-speed height-velocity diagram of a variable-speed-

rotor helicopter in one-engine-failure." Aero. Sci. Technol. 91 (2019): 310-320.

https://doi.org/10.1016/j.ast.2019.05.003

189. Liu, Hao, et al. "Robust optimal attitude control of hexarotor robotic vehicles."

Nonlinear Dyn. 74 (4) (2013): 1155-1168. https://doi.org/10.1007/s11071-013-1031-4

190. Carlson, Eric Bernard. Optimal tiltrotor aircraft operations during power failure.

Minnesota: University of Minnesota, 1999. Available at:

https://ui.adsabs.harvard.edu/abs/1999PhDT.......225C/abstract

191. Muro, D., et al. "An optimal control approach for alleviation of tiltrotor gust response."

Aeronaut. J. 116. (1180) (2012): 651-666. https://doi.org/10.1017/S0001924000007119

192. Thomson, D. G., and R. Bradley. "An investigation of the stability of flight path

constrained helicopter manoeuvres by inverse simulation." 13th European Rotorcraft

Forum. Arles: ERF, 1987: 122-142 (1987). Available at: http://eprints.gla.ac.uk/139153/


72

193. Hess, R. A., C. Gao, and S. H. Wang. "Generalized technique for inverse simulation

applied to aircraft maneuvers." J. Guid. Control Dyn.14 (5) (1991): 920-926.

https://doi.org/10.2514/3.20732

194. Thomson, Douglas G., and Roy Bradley. "The principles and practical application of

helicopter inverse simulation." Simul. Pract. Theory. 6 (1) (1998): 47-70.

https://doi.org/10.1016/S0928-4869(97)00012-8

195. Thomson, Douglas G., and Roy Bradley. "Mathematical definition of helicopter

maneuvers." J. Am. Helicopter Soc. 42 (4) (1997): 307-309.

https://doi.org/10.4050/JAHS.42.307

196. Leacock, Gary, and Douglas Thomson. "Helicopter handling qualities studies using

pilot modeling and inverse simulation." Annual Forum Proceedings-American

Helicopter Society. Vol. 54. American Helicopter Society, 1998. Available at:

https://vtol.org/store/product/helicopter-handling-qualities-studies-using-pilot-

modelling-and-inverse-simulation-7252.cfm

197. Cameron, N., D. G. Thomson, and D. J. Murray-Smith. "Pilot modelling and inverse

simulation for initial handling qualities assessment." Aeronaut. J. 107. (1074) (2003):

511-520. https://doi.org/10.1017/S0001924000134013

198. Hess, Ronald A., Yasser Zeyada, and Robert K. Heffley. "Modeling and simulation for

helicopter task analysis." J. Am. Helicopter Soc. 47 (4) (2002): 243-252.

https://doi.org/10.4050/JAHS.47.243

199. Hess, Ronald A. "Simplified approach for modelling pilot pursuit control behaviour in

multi-loop flight control tasks." Proc. Inst. Mech. Eng. Part G-J. Aerosp. Eng. 220 (2)

(2006): 85-102. https://doi.org/10.1243/09544100JAERO33

200. Lee, Dooyong, et al. "Simulation of helicopter shipboard launch and recovery with

time-accurate airwakes." J. Aircr.42 (2) (2005): 448-461. https://doi.org/10.2514/1.6786


73

201. Yuan, Ye, Douglas Thomson, and David Anderson. "Application of Automatic

Differentiation for Tilt-Rotor Aircraft Flight Dynamics Analysis." J. Aircr. (2020): 1-6.

https://doi.org/10.2514/1.C035811 (Published online)

202. Kl de D H, S hul e C, Thompson M, et l. “Use o W elet S lo r ms to

Characterize Rotorcraft Pilot-Vehi le S stem Inter tions.” Ameri n Heli opter

Society International 66th Annual Forum Proceedings, Phoenix, AZ: AHS, 2010: 1-11.

Available at: https://vtol.org/store/product/use-of-wavelet-scalograms-to-characterize-

rotorcraft-pilotvehicle-system-interactions-1660.cfm

203. Lu, Linghai, Michael Jump, and Michael Jones. "Tau coupling investigation using

positive wavelet analysis." J. Guid. Control Dyn, 36. (4) (2013): 920-934.

https://doi.org/10.2514/1.60015

204. Thomson, D. G., and Roy Bradley. "The use of inverse simulation for preliminary

assessment of helicopter handling qualities." Aeronaut. J. 101. (1007) (1997): 287-294.

https://doi.org/10.1017/S0001924000066148

205. Trits hler J K, ’Connor J C. “Use o Time-Frequency Representations for Interpreting

H ndlin Qu lities Fli ht Test D t .” J. Guid. Control Dyn. 39. (12) (2016): 2768-2775.

https://doi.org/10.2514/1.G000401

View publication stats

You might also like