DURANVENEGAS 486qrja279dale963lsei649xitu TH
DURANVENEGAS 486qrja279dale963lsei649xitu TH
THÈSE
par
1 Introduction 5
1.1 Context . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.2 Description of the wake of a rotor . . . . . . . . . . . . . . . . . . . . . . . 8
1.2.1 Rotor regimes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.2.2 Momentum theory . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.2.3 Vortex methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.3 Aerodynamic loads . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
1.3.1 An introduction to airfoil aerodynamics . . . . . . . . . . . . . . . . 16
1.3.2 Blade element theory . . . . . . . . . . . . . . . . . . . . . . . . . . 18
1.4 Blade deformation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
1.4.1 Mechanical properties of deformable materials . . . . . . . . . . . . 20
1.4.2 Beam model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
1.5 A note on fluid-structure interactions . . . . . . . . . . . . . . . . . . . . . 23
1.6 Helical vortices in a rotor wake . . . . . . . . . . . . . . . . . . . . . . . . 23
1.7 Stability of wake solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
1.8 General overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2 Wake model 29
2.1 Chapter overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
2.2 Generalized Joukowski model . . . . . . . . . . . . . . . . . . . . . . . . . 30
2.3 Far-wake solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2.3.1 Vortex filament framework . . . . . . . . . . . . . . . . . . . . . . . 31
2.3.2 Deformed helical vortex pairs . . . . . . . . . . . . . . . . . . . . . 39
2.3.3 Analysis of the effect of a varying core size . . . . . . . . . . . . . . 50
2.3.4 Discussion in the context of rotors . . . . . . . . . . . . . . . . . . . 50
2.4 Near-wake solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
2.4.1 Framework . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
2.5 Rb∗ = 0 case (Standard Joukowski model) . . . . . . . . . . . . . . . . . . . 55
2.5.1 Framework . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
2.5.2 Wake solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
2.5.3 Induced velocities . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
2.5.4 Energy extracted by a wind turbine . . . . . . . . . . . . . . . . . . 65
2.6 Rb∗ 6= 0 case (Generalized Joukowski model) . . . . . . . . . . . . . . . . . 67
2.6.1 Topology of the solutions . . . . . . . . . . . . . . . . . . . . . . . . 67
2.6.2 Parametrical analysis . . . . . . . . . . . . . . . . . . . . . . . . . . 68
2.6.3 Comparison with the standard model . . . . . . . . . . . . . . . . . 74
3
2.7 Chapter conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
4 Rigid blade 87
4.1 Chapter overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
4.2 Rigid blade framework . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
4.2.1 Induced velocity on the rotor . . . . . . . . . . . . . . . . . . . . . 88
4.2.2 Circulation law and equivalent Joukowski profile . . . . . . . . . . . 89
4.2.3 Aerodynamic loads . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
4.2.4 Tip effect correction . . . . . . . . . . . . . . . . . . . . . . . . . . 91
4.2.5 Model parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
4.3 Comparison with numerical and experimental results . . . . . . . . . . . . 92
4.3.1 MEXICO rotor project . . . . . . . . . . . . . . . . . . . . . . . . . 92
4.3.2 The Unsteady Aerodynamic Experiment (UAE) . . . . . . . . . . . 96
4.3.3 A hovering case: NASA Technical report TM81232 . . . . . . . . . 98
4.4 Chapter conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
6 Conclusions 113
6.1 Achievements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
6.2 Future work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
Appendices 117
4
Chapter 1
Introduction
Contents
1.1 Context . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.2 Description of the wake of a rotor . . . . . . . . . . . . . . . . 8
1.2.1 Rotor regimes . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.2.2 Momentum theory . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.2.3 Vortex methods . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.3 Aerodynamic loads . . . . . . . . . . . . . . . . . . . . . . . . . 16
1.3.1 An introduction to airfoil aerodynamics . . . . . . . . . . . . . 16
1.3.2 Blade element theory . . . . . . . . . . . . . . . . . . . . . . . . 18
1.4 Blade deformation . . . . . . . . . . . . . . . . . . . . . . . . . . 20
1.4.1 Mechanical properties of deformable materials . . . . . . . . . . 20
1.4.2 Beam model . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
1.5 A note on fluid-structure interactions . . . . . . . . . . . . . . 23
1.6 Helical vortices in a rotor wake . . . . . . . . . . . . . . . . . . 23
1.7 Stability of wake solutions . . . . . . . . . . . . . . . . . . . . . 24
1.8 General overview . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
5
Figure 1.1: Left: V164 offshore wind turbine model from MHI Vestas Offshore Wind, with
a diameter of 164m and a total power of 9.5MW (source: www.mhivestasoffshore.com).
Right: Detail of the blades of a Kaman HH-43B Huskie helicopter (source:
www.nationalmuseum.af.mil )
1.1 Context
The term rotor is used in many technical domains, from mechanical to electrical engi-
neering, to define rotating parts of a machine or a mechanical assembly. In this thesis, we
are interested in rotors as defined in aeronautics: a system of rotating airfoils. This kind
of rotor can transform the movement of a fluid into mechanical rotation or vice-versa. It
has been used for centuries and they are very common in many daily applications. In
particular, we are interested in two of its main uses: harvesting and propulsion.
Harvesting devices are probably the most ancient rotor applications. The principle is
to transform the kinetic energy of a flow into a mechanical or electrical energy through
the rotation of the system. The use of watermills and windmills to obtain useful mechan-
ical energy has been documented for more than a thousand years. Nowadays, one of the
main rotor applications in harvesting is wind turbine. The total installed power in wind
farms grows every year, introducing more and bigger devices that generate new technical
challenges. In the last years, the tendency has been to design more powerful devices by
increasing the diameter of the rotor (Figure 1.1a shows a turbine Vesta V164, with 164m
of diameter), however the great increase in the length of the blades introduces important
effects on their deformation that can affect the performance of the rotor. Furthermore,
the construction of populated wind farms presents some technical problems caused by the
interaction between devices. The wake generated behind a rotor can interfere in the fol-
lowing row, altering the incoming flow and reducing the efficiency of the turbines. Having
a good comprehension of the dynamic mechanisms of the wake or blade deformation is
fundamental to give a solution to these problems.
Propulsion devices follow a principle inverse to that of wind turbines: a rotation is
forced mechanically in order to generate a motion of the fluid and induce a thrust force
perpendicular to the rotor disc. This kind of rotors can be found in boat propellers or
helicopters. The main difficulty in the analysis and design of helicopter rotors lies in
6
Figure 1.2: Rotor with very flexible blades. Left: at rest; right: rotating. From Sicard
(2014).
the many different flow configurations they can adopt, depending on the external wind
conditions and if the vehicle is ascending, descending or in horizontal flight. In this sense,
several sorts of transitions and instabilities can occur on the wake while helicopter is
operating. These instabilities may lead to dangerous flight situations, as the so called
Vortex Ring State, where a vortex ring system is generated around the rotor, producing a
problematic loss of lift. Again, understanding the behaviour of the wake is fundamental to
prevent this kind of issues. The effect of deformation of the blades is also very important
in helicopters because of the slenderness of the blades and the great aerodynamic forces
exerted on them. In figure 1.1b, a helicopter Kaman HH-43B is shown at rest where the
deformation of the blades can be perfectly noticed.
Another particular case of propulsion rotors is found in drones, whose use in society
has grown noticeably in the last decade. These devices are quite recent and have not
involved as much research as helicopters. Frequently, their rotors are designed as copies
of aircraft propellers at a smaller scale, however, drones have particular characteristics
that should be taken into account:
- For different reasons, as reducing manufacturing cost, blades are often made of
plastic. The use of polymeric materials for the design of the blades further increases
the importance of the deformation in the study of their performance. In figure 1.2
is displayed a case of a very flexible drone rotor, where the blades are extremely
deformed at rest and get straight under the effect of the centrifugal forces when
rotating.
7
Figure 1.3: NASA GL-10 convertible drone. Source: www.nasa.gov
inclination can be modified from vertical to horizontal position (figure 1.3). For
this kind of drone, rotors have to be specifically designed to operate in these two
different configurations, consequently, a good comprehension of blades deformation
or rotors wakes interaction is required.
In conclusion, in spite of their long history, rotors are a constantly developing tech-
nology. Understanding the dynamic of the wake, its interaction with the rotor and the
mechanisms of deformation of the blades is important to address new challenges in rotors
design. Understanding the coupling mechanisms between the flow and the flexibility of
the rotor would be very useful from the point of view of optimization. Introducing the
elasticity of the blades on rotor design could help, for example, to improve the performance
of drones or enlarge the operational range of wind turbines. Having a precise description
of the wake may also help to predict the transition to dangerous helicopter flight regimes
or avoid interference between turbines in wind farms. However, these problems are very
complex and heavy from the point of view of computation. The fluid-structure interac-
tion problem resulting from the effect of the flow on the blade deformation and vice-versa
becomes very heavy to compute using Direct Numerical Simulation methods. This com-
plexity makes impossible any optimization procedure or any extensive parametric analysis
with current computation means. In this work, different simplifications are proposed at
every level to reduce the complexity and computational cost of the problem resolution,
resulting in a light model with low computation times allowing to address optimization
analysis. This model would provide a versatile tool to address new challenges in rotor
design.
8
(a) (b)
(c) (d)
(e)
Figure 1.4: Scheme of the flow crossing the rotor plane for different regimes under a
external axial flow. The black arrows represent the streamlines of the flow, while the
dashed line represent the projection of the curve enveloping the wake. The red arrow
shows the direction and the relative magnitude of the total velocity crossing the rotor
plane. (a) Climbing flight. (b) Hover. (c) Slow descent. (d) Vortex-Ring State. (e)
Windmill-brake / Wind turbine.
9
dependency with the external wind. In this case, rotation of the blades is imposed,
generating a lift force that balance the weight of the aircraft. There are three main ways
to control the flight: collective pitch, cyclic pitch and anti-torque. The collective pitch
changes the angle of attack of all the blades at the same time, modifying the lift and
allowing a vertical climbing or descending. The cyclic pitch implies a local variation of
the angle of attack, causing an horizontal displacement of the helicopter. The anti-torque
is a system to avoid uncontrolled rotation of the cabin produced by the torque of the rotor.
Several solutions exist to compensate the torque, the most common is with a small tail
rotor. The regulation of this system, over- or under-compensating the torque, introduces a
controlled rotation on the cabin. In drones, as it was explained in the previous section, size
restriction makes it impossible to include complex pitch control systems. Flight control
is made by modifying the speed of the different rotors. In conclusion, many different flow
regimes can be reached depending on the operational conditions of the device.
In this work, we will focus on uniform axial flow regimes, that is, when the external
flow is homogeneous and aligned with the rotor axis. This corresponds to the normal
operational conditions of wind turbines and vertical flight regimes of helicopters and
drones. In particular, the non-homogeneous character of the wind owing to the presence
of the ground is not taken into account for the wind turbines. For helicopters, the forward
flight regimes are not considered in the present work. The external flow is always assumed
to be perpendicular to the rotor plane. Under this hypothesis, the analysis of different
operating states can be reduced to a few rotor regimes (figure 1.4). The rotation of the
blades induces a flow perpendicular to the rotor plane. This induced flow can have the
same direction as the external flow or be opposed to it. Depending on the regime, a rotor
can accelerate or decelerate the external flow. Figure 1.4 shows a sketch of the mean
flow (azimuthally averaged) streamlines crossing the rotor plane for the different rotor
regimes. The streamlines are the family of curves that are parallel to the flow direction,
so they never cross each others. In a three-dimensional flow, an ensemble of streamlines
that form a closed surface is called streamtube. In figure 1.4, as axisymmetric flow is
assumed, streamlines also correspond to streamtubes of circular cross section.
In the case of helicopters or drones in ascending flight (figure 1.4a) the external flow
and the induced flow have the same direction, so the flow is accelerated. In the case of
hovering flight (figure 1.4b), there is no external flow, so all the displacement of the fluid is
generated by the rotor. The most interesting cases is probably when the external and the
induced flows are in opposite directions; in that situation, different regimes can appear.
The simplest regime to analyze is probably associated with wind turbines or windmill
brake descending helicopter (figure 1.4e), where the external flow is strong and the total
flow has always the same direction across the whole rotor plane. However, there are
regimes associated with helicopters in descending flight where the induced flow is strong
compared to the external flow and the total flow changes its direction across the rotor
plane. In soft descending flight regime (figure 1.4c), the flow crossing the rotor plane
goes upwards out of the rotor disc and downwards inside of it. If the descending speed
is increased, the Vortex Ring State regime (figure 1.4d) can be reached, where the flow
direction changes inside the rotor disc, generating lift problems with serious consequences
to the security of the helicopter.
To study the flow generated by the rotor in these different axial regimes, various
10
(a) (b)
Figure 1.5: (a) Scheme of the streamtube around a rotor disc in a wind turbine configura-
tion. The arrows represent the 1D momentum theory velocities. In solid grey, the circular
areas of the stream tube upstream, downstream and on the rotor position. (b) Velocity
and pressure distributions along the axial direction.
models are commonly used. In this chapter, the momentum theory and the vortex theory
will be introduced. Concerning the wake description, in addition to uniform axial flow
hypothesis, no time dependance will be considered in this work. Viscous effects, as shear
of dissipation phenomena, will be also neglected.
11
Particular assumptions are made for pressure distribution. Far upstream and down-
stream, it is supposed to be equal to the undisturbed ambient pressure (figure 1.5b).
The net action of the lateral pressure acting on the stream tube is zero. Then, applying
Bernouilli’s equation at both sides of the rotor, the total pressure drop is given by
∆p = 12 ρ(u20 − u21 ). (1.2)
This pressure drop is supposed to be uniform all over the rotor area, so the thrust and
power extraction are simply obtained as:
T = ∆pAR = 12 ρAR (u20 − u21 ), (1.3)
P = uR T = 21 ρAR uR (u20 − u21 ). (1.4)
According to assumptions imposed for the pressure and assuming an inviscid, incom-
pressible and steady flow, one can deduce the velocity across the rotor as the mean value
between upstream and downstream velocities:
1
uR = (u0 + u1 ). (1.5)
2
An important characteristic of a rotor that can be deduced from momentum theory is
the axial interference factor, which can be defined as the ratio of the undisturbed external
flow velocity to the total flow velocity behind the rotor:
u0 + uR
a= , (1.6)
u0
so velocities in the far wake and in the rotor plane can be reformulated as u1 = (1 − 2a)u0
and uR = (1 − a)u0 . This permits to give an expression of power extraction in terms of
the axial interference factor:
P = 2ρAR u30 a(1 − a)2 . (1.7)
It is also very common to use of the so-called power coefficient, which gives a non-
dimensional value for the power extracted by the rotor:
P
CP = 1 3
= 4a(1 − a)2 . (1.8)
2
ρA R u0
Then, differentiating with respect to a, a maximum theoretical obtainable power CPmax =
16/27 = 0.593 for a value of a = 1/3 is deduced. This is known as the Betz limit, it implies
that only the 59.3% of the kinetic energy of the flow inside a stream tube of the same
diameter of the disk can be transformed to useful work by the rotor. In addition, this is
just a conservative limit that does not take into account the losses related to dissipation
processes. Moreover, rotation has not been considered here. This can be included in
Momentum theory as explained for instance in Sørensen (2016).
Momentum theory gives very useful results to predict the behaviour of the flow in
wind turbines or climbing flight regimes, however, it cannot predict many practical effects.
Although it has progressively been modified to consider the presence of multiple blades,
the influence of the vortical wake or three-dimensional effects (see Sørensen, 2016), it
assumes a particular topology of the flow streamlines that is not always satisfied. The
observed regimes where the flow is not uniform at each cross section or when it changes
direction, as it occurs in descending flight and Vortex Ring State, cannot be described by
the Momentum theory. For these situations, different methods have to be used to describe
the flow — one of them is presented in the next section.
12
(a) (b)
Figure 1.6: (a) Wake generated by a rotating blade during hovering regime with a devel-
oping vorticity sheet and an helical tip vortex (Gray, 1956). (b) Circulation profile along
a blade.Tip and hub vortices are rolled-up behind the ends of the blade. In the middle, a
vorticity sheet is shed.
13
(a) (b)
Figure 1.7: Rotor wake models proposed by (a) Betz (1926) and (b) Joukowski (1929).
(b)
(a) (c)
Figure 1.8: Examples of visualization of the tip vortex in (a) a wind turbine (from Vermeer
et al., 2003) (b) a boat propeller (Cavitation Research Laboratory/AMC) and (c) a Bell
AH-1 Cobra helicopter
14
Figure 1.9: (a) Schematic of the velocity induced by a vortex filament of circulation Γ
at a distance r. (b) Schematic of the local self-induced velocity of a vortex filament and
the cut-off length δa. In dotted line, the circle osculating the filament at the point where
self-induction is being calculated.
In the free-vortex methods, the geometry of the vortex elements forming the wake is
calculated in the time domain, providing a much more precise structure than that of the
prescribed shape models. In many practical problems, viscous effects can be neglected
and an inviscid approach can be used. In this framework, vortex elements move cohesively
as material elements with the local fluid velocity, giving the fundamental equation for the
free-vortex method:
dr
= V (t, r), (1.9)
dt
where r(t) is the position of a point of the vortex element in a time.
The local velocity V is the sum of the external velocity field and the velocity induced
by the vortical elements. This contribution is given for a vortex filament of circulation Γj
by the Biot-Savart law (figure 1.9):
r × dl
Z
Γj
VjInd = , (1.10)
4π ||r||3
where dl is the differential element of the vorticity filament. The total induced velocity
is obtained as the sum of the induction of all the vortical elements. The main difficulty
of the Biot-Savart formula is the presence of a singularity at the position of the vortex
element. To resolve this issue, different techniques have been developed in the literature
(this will be treated in detail in section 2.1). One of the most commonly used is the cut-off
method (Thomson, 1883; Crow, 1970). The idea of this method is to exclude a portion
of the filament on both sides of the singularity in the Biot-Savart integral. The length
of each portion is proportional to the vortex core size a and it is called cut-off length
δa. The proportional factor δ depends on the characteristics of internal structure of the
vortex (for details about the calculation of δ, see Saffman, 1992).
By applying the cut-off method, a simple asymptotic expression can be derived for the
15
self-induction of a curved filament of vorticity due to the effect of its local curvature:
Γ L
Vloc = κb log , (1.11)
4π δa
where κ is the local curvature of the filament and L is the length of integration from both
sides of the considered point. This expression is the so-called Local Induction Approach
(LIA). It is the leading order approximation in the limit a/L → 0. It clearly demonstrates
that the self-induced velocity of a curved filament diverges as its core size goes to zero.
It is also mainly directed along the binormal direction b of the filament (figure 1.9).
The velocity surrounding a vortex element, coming from an external flow or self-
induction, in general, displaces and deforms vortical structures. Only a few vortex struc-
tures, like vortex rings or helices, have the particularity to move without changing shape
(this property will be discussed in section 2.3). In this work, we will be interested in this
kind of structures, looking for structures that are stationary in an adequate frame. In the
context of rotors, the frame will be moving with the blades.
Free and constrained wake approaches (Leishman, 2006; Bhagwat & Leishman, 2000;
Bliss, 1993) are examples of numerical studies using the Biot-Savart law. The main
advantages of vortex methods are that they give accurate approximations for the wake
while they are easy to compute and in general much less computationally costly than
direct numerical simulations. However, they suffer from stability problems and they can
also become very expensive if a high number of elements is used. A good way to have
a simple and relatively efficient description of the wake is to use the free-vortex method
procedure with a reduced vortex model involving only a few vortices as the Joukowsky
model (Gupta & Leishman, 2005b).
16
(a) (b)
Figure 1.10: (a) 2D airfoil under a external flow u. The dashed line circle C represents
a closed line surrounding the airfoil with a differential of length ds. The integration of
the velocity field u along the closed line C gives the circulation Γ around the profile.
(b) Scheme of the lifting line theory. The circulation Γ(r) along the wing is computed
applying Kutta-Joukowski law at every 2D element.
of the flow. This force can be decomposed into lift FL , acting perpendicularly to the
flow direction, and drag FD , in the same direction as the flow. For a 2D airfoil, Kutta-
Joukowski law links the lift force with the circulation Γ around the airfoil, the density of
the fluid ρ and the speed of the airfoil U through:
FL = ΓρU. (1.12)
This law is conceived for two-dimensional potential flows, that is, inviscid, incompressible
and irrotational, however, it is still a good approximation for real viscous flows in typical
aerodynamic applications. To apply it in real flows, it is specially important to have a
smooth merging between the fluid coming from upper and lower surfaces, without fluid
circulation near the trailing edge of the airfoil. This is known as the Kutta condition
and it is normally guarantied with a sharp trailing edge and moderated angles of attack
(Kuethe & Chow, 1976).
Using Kelvin’s circulation theorem, the circulation can be mathematically defined as
the integral of the velocity field u around a closed contour C surrounding the airfoil (figure
1.10a): I
Γ= u · ds. (1.13)
C
In real flows, even when viscous effects are very weak, there always exist a narrow viscous
boundary layer around the airfoil. To apply Kutta-Joukowski theorem, the closed line
surrounding the airfoil must be located outside the boundary layer.
For a straight airfoil of length l and chord c, lift and drag forces are related to the
properties of the flow felt by the blade through the so-called lift and drag coefficients,
which depend exclusively on the geometry of the airfoil:
2FL 2FD
CL = ; CD = . (1.14)
ρU 2 cl ρU 2 cl
17
These two coefficients are crucial in applied aerodynamics. They are used to predict the
aerodynamic forces on a specific airfoil under particular flow conditions. In figure 1.11b, it
is illustrated a typical distribution of CL and CD as a function of the angle of attack α for
a NACA0009 profile (from Abbott & Von Doenhoff, 1960). This distribution changes with
Reynolds numbers, so several tables are normally given for an specific airfoil geometry.
Historically, the coefficients have been determined experimentally in wind tunnel tests for
different airfoil geometries (Jacobs et al., 1935). Numerous compilations can be found
in the literature with lift and drag coefficients data of the most commonly used airfoils
(Abbott & Von Doenhoff, 1960).
The relation between circulation and lift extracted from Kutta-Joukowski formula can
be finally reformulated to obtain a simple expression for the circulation that includes the
lift coefficient:
1
Γ = cU CL . (1.15)
2
This formula is a priori valid just for 2D airfoils. A very useful and simple model to
compute the circulation of a 3D airfoil is based on treating it as a series of 2D elements (see
figure 1.10). This is the case of the lifting-line theory (Prandtl, 1921), which assumes that
any section of a finite wing experiences the same lift as an equivalent section of a infinite
wing with the same uniform circulation. The circulation profile is then calculated by
applying Kutta-Joukowski formula (1.15) to each section of the wing (figure 1.10b). This
model takes into account the variation of the blade geometry and the flow properties along
the wing so some 3D effects are taken into account in the circulation and aerodynamic
loads distribution. However, the method is not powerful enough to capture more complex
phenomena as tip and root losses (Leishman, 2006). Nevertheless, the lifting-line theory
approach has been widely used and improved, for example, to adapt it to wings with
arbitrary camber or sweep angle (Phillips & Snyder, 2000).
The same principle of the lifting-line theory to obtain load distributions on finite wings
is specifically applied to rotors in the Blade Element Theory.
18
(a) (b)
Figure 1.11: (a) Scheme of the Blade Element Theory. The aerodynamic loads are calcu-
lated at each element of the blade from the effective velocity U received. (b) Example of a
typical distribution of lift and drag coefficients (Here for NACA0009 airfoil, from Abbott
& Von Doenhoff (1960)).
Each blade element is studied independently, calculating the amplitude U and orienta-
tion (angle of attack α) of the local velocity (figure 1.11). This local velocity is composed
of two main contributions: the external flow and the flow induced by the wake. In free
vortex methods, the induced flow is computed through Biot-Savart law.
Now, the contributions to lift and drag per unit of span associated with each element
can simply be derived from Kutta-Joukowski as:
1 1
FL (r) = ρU 2 cCL , FD (r) = ρU 2 cCD . (1.16)
2 2
These formulas give a very simple way to calculate aerodynamic forces. In practice, it
is just necessary to compute the local velocity and the angle of attack, then CL and
CD are directly obtained from tabulated data. In rotor blades, local velocity is mainly
governed by the rotation of the blades, so contrary to airplane wings, Reynolds numbers
may change from tip to hub. In this work, this effect will not be considered, and we shall
use the Reynolds number at the tip to choose lift and drag coefficients distributions.
Once lift and drag forces distributions are computed, they can be integrated along the
blades to obtain the total thurst, torque and power delivered by the rotor.
The main difficulty in BET is to estimate the local velocity U along the blades. While
the external flow and the rotor angular speed are known, the induced velocity has to be
calculated, so a model for the wake has to be used together with BET. Glauert (1935),
de Bothezat (1919) and Reissner (1937, 1940) linked BET with the momentum theory
to obtain the induced velocity distribution for the airplane propeller problem. Later,
Gustafson & Gessow (1946) and Gessow & Gustafson (1948) developed a similar approach
for hovering helicopter rotors. This combined theory is known as the Blade Element
Momentum theory and is widely used for the design of wind turbines (Hansen et al.,
2006). BET has also been linked with vortical descriptions of the blades. These theories
19
(a) (b)
Figure 1.12: (a) Prismatic bar under an axial tension. (b) Typical stress-strain diagram
for steel where most common transitions can be seen from linear elastic deformation to
fracture.
have been developed from the beginning of the twentieth century by Joukowski (1929),
Glauert (1922) or Lock et al. (1925). Betz (1920) and Goldstein (1929) developed a
prescribed vortex wake theory for propellers which was later extended by Theodorsen
(1948). The combination of BET with prescribed vortex wake models was also early
applied to the helicopter rotor problem (Knight & Hefner, 1937).
P
σ= , (1.17)
A
20
that we suppose uniformly distributed over the cross section area.
On the other hand, the strain, , is a non-dimensional measure of the deformation of
the material. Under an axial load, a straight bar of length L will be shortened or lengthen
a distance δ, depending on whether the load is compression or traction. Then, strain is
defined as
δ
= . (1.18)
L
Strain can also be described laterally, as a lateral contraction occurs when the bar is
elongated or an expansion when compressed.
To understand the behavior of a particular material and employ it in a design problem,
it is important to know the intrinsic relation between stress and strain. This relation is
illustrated in stress-strain diagrams. A typical strain-stress diagram of steel is shown in
figure 1.12b. Initially, when the applied load is further increased, there is a linear relation
between strain and stress and the material has an elastic behaviour, where deformation
is completely reverted if load is removed. However, if load is still increased, the material
becomes plastic, the deformation continues at constant stress and it remains even if load
is removed. After that, if loading is increased, plastic deformation continues with a non-
linear variation of σ before a great slimming of the material (necking) and, finally, its
fracture. Depending on the material, the regions of the diagram can completely change,
being enlarged or reduced. In this work, we are interested in the linear region and an
elastic behaviour will always be supposed for rotor blades.
Staying in the elastic region, the description of the mechanic behaviour of the material
is greatly simplified and it can be described by just two parameters: Young’s modulus E
and Poisson ratio ν.
Young’s modulus describes the linear relationship between stress and strain in the
elastic region (figure 1.12b). It is defined by the well known Hooke’s law:
σ
E= , (1.19)
Poisson ratio describes how much a material is laterally contracted when axially elongated:
lateral strain
ν=− (1.20)
axial strain
These two properties are easy to measure by applying a pure axial traction on a pris-
matic bar. Once the elastic properties of the material are determined from a experimental
test, they can be used to calculate the deformation of more complex geometries. For that,
stress and strain distributions have to be computed for the whole structure.
21
Figure 1.13: Upper: Internal and external forces and moments exerting on a chunk of
beam. In the classical Euler-Bernouilli beam theory the length of the beam has to be
much bigger than the rest of dimensions. Lower: asymptotic limits of rod (a ∼ b << L)
and ribbon (a << b << L) models.
important simplification in the resolution. This theory describes the deformations of the
blade through 1D equations along the centerline, bringing a great reduction of the com-
putation cost. In the seventieth century, Jacob Bernouilli (1654-1705) first postulated the
principles of the theory, later, Daniel Bernouilli (1700-1782) and Leonhard Euler (1707-
1783) extended the theory by formulating the differential equation for the motion of a
vibrating beam and analyzing the deformation of the beam under different loading con-
ditions (see Han et al., 1999). Later, Timoshenko (1922) improved the theory including
shear deformation and rotational bending effects.
The theory is applicable to linear elastic problems and is based on several assumptions
concerning the cross section of the beam: it is infinitely rigid in its own plane, it remains
plane after deformation and it is always normal to the deformed axis of the beam (Bauchau
& Craig, 2009). Under these hypotheses, a force and momentum balance is made at every
section of the beam. In figure 1.13, the forces and moments involved in the balance of
a small section of beam are represented. T and M are the internal forces and moments
and f and m are the external forces and moments per unit length. Applied to a flexible
rotor, the blades can be modeled as cantilevered beams where the external forces are the
aerodynamic loads, the weight of the blades and centrifugal force. One of the advantages
of beam theory is that it allows the computation of both static and dynamic problems.
In our case, we will use it to obtain static blade deformation equilibrium.
The link between forces and moments on the structure with the elastic properties
of the material is given by a constitutive relation, relating E and ν with the internal
moments of the beam. For that, in addition to the hypothesis of slenderness (the span
is bigger than the other lengths of the beam) some other geometrical assumption has to
22
be made. Different asymptotic limits can be considered (see Carrera et al., 1926), each
one with its own constitutive relation. Two of the most common models are rod and
ribbon (figure 1.13). The combination of force and moment balance with the constitutive
equation permits to calculate beam bending and torsion.
The rod model (Dias & Audoly, 2014) is the simplest one, it assumes that the length
of the blade is very large compared to the other dimensions involved in the problem. The
constitutive law gives a linear expression for the internal moment in terms of the local
torsion and bending. In this case, torsion and bending are not intrinsically coupled, which
gives an important simplification to the problem.
For the ribbon model (Volovoi et al., 2001; Dias & Audoly, 2015), the height of the
cross section of the beam is supposed to be small compared to its width. In this case,
the constitutive law becomes more complicated, yielding an intrinsic non-linear coupling
between bending and torsion.
Independently of the intrinsic characteristics of the model, in flexible rotors there is
always a coupling between bending and torsion because of the effect of aerodynamic loads:
when a blade twists or bends, it changes its position with respect to the flow, this produces
a variation of the aerodynamic loads and, in turn, of twisting and bending. A number
of authors have shown that it is possible to take advantage of this coupling and improve
rotors adaptability to non-optimal operational conditions. Two possible strategies can be
used to exploit the coupling: active (i.e. forcing blade twist to obtain a bending response)
or passive (using just the natural coupling given by the blade geometry), which is the most
common due to its simplicity and economy. The effect of coupling has been analyzed to
improve the performance of different kinds of rotor devices, from wind turbines (Maheri
& Isikveren, 2010) to drones (Lv et al., 2013). As an example, load mitigation in wind
turbines is one of the possible applications of the bending-twisting coupling, which can
be useful to reduce fatigue damage on the blades (Lobitz & Veers, 2003; Bottasso et al.,
2013).
23
early works by Kelvin (1880), Da Rios (1916), Levy & Forsdyke (1928) and Joukowski
(1929).
For instance, Betchov (1965) and Kida (1981) showed using the local induction ap-
proximation that these structures rotate and translate without changing form. When the
vortex is (infinitely) thin, Hardin (1982) and Kawada (1936) (see Fukumoto et al., 2015)
provided an exact expression for the induced velocity field inside and outside the cylin-
der containing the helix. Ricca (1994), Kuibin & Okulov (1998) and Boersma & Wood
(1999) showed how the singularity of Hardin expression can be extracted to compute the
self-induced motion of the helix. Velasco Fuentes (2018) recently applied these results
to compute the motion of vortex elements on a helical vortex, emphasizing the role of
tangential velocities. In these works, the vortex core size is implicitely assumed to be
a small parameter. Lucas & Dritschel (2009) and Selçuk et al. (2017) have shown how
helical vortices with a thick core can be obtained numerically by enforcing the helical
symmetry in the governing equations.
Solutions with more complex geometries are scarce in the literature. Walther et al.
(2007) looked at equilibrium solutions composed of undeformed helical vortex pairs of
same pitch but different radius. They demonstrated that undeformed helical vortex pairs
of identical pitch and opposite circulation were possible for a particular radius ratio.
This analysis was further pursued by Okulov (2016) for helical vortices with same sign
circulation. Reducing the framework to nearly parallel filaments (Klein et al., 1995),
Kwiecinski & Van Gorder (2018) were recently able to provide more exotic solutions.
Pairing instabilities
Among the various instabilities occurring in the context of vortex dynamics, the pairing
instability, that is the spontaneous grouping by pair of vortices, is probably the most
important one. This instability was first described in the context of 2D point vortices.
Crow instability (Crow, 1970) is a 3D version of the pairing instability that appears in
the wake of airplanes. It results from the 3D interactions of the counter-rotating vortex
pair that is generated by the airplane. It leads to the deformation of the vortices that
ultimately touch, reconnect and form an array of vortex rings (figure 1.14).
24
(a) (b)
Figure 1.14: (a) Scheme of two parallel counter-rotative vortices developing a symmetric
Crow instability mode. (b) Chain of vortex rings of a Crow instability
Figure 1.15: (a) Long wave instability results for a single helical vortex obtained by
Widnall (1972). Displacement prediction for several wavelength perturbations. (b) Three
dimensional view on the left and developed plan view on the right of the unstable mode
associated with k = 1 for a pair of helical vortices (from Quaranta, 2017). (c) Growth
rate of symmetric modes as a function of the wavenumber k for a pair of helical vortices.
Theoretical distribution from Gupta & Loewy (1974) expressions.
25
(a) (b)
Figure 1.16: Spatio-temporal evolution of a perturbation (a) convectively and (b) abso-
lutely unstable.
Independently of the analysis of the geometry and growth rate of the different unstable
modes, a spatio-temporal stability analysis can be performed to study its propagation
along the wake (see Huerre & Monkewitz, 1985, 1990). In our case, the objective of this
26
kind of analysis is to determine if, for a given wake solution, the instability propagates
downstream or spreads upstream and affect the flow in the rotor plane.
To study the nature of the instability, the impulse response is analyzed in the (z, t)
plane (figure 1.16), where z is the direction of the flow. A particular velocity of propagation
v = z/t is represented in the plane as a straight line radiated from the initial location of
the perturbation. In a unstable system, the perturbation is amplified but spread while it
is advected downstream. Of particular interest are the rear and front velocities v− and
v+ of propagation of the wavepacket. If there is a single wavepacket, in any frame moving
at a velocity, v, such that v− < v < v+ , the perturbation grows at any fixed position.
Then, two different situations can occur depending on whether v+ and v− have the same
sign or not. If they have the same sign (figure 1.16a), the instability is advected away
from its source. In this case, the flow is said convectively unstable. In the other case, if
v+ and v− have different signs (figure 1.16b), the perturbation propagates both upstream
and downstream as it grows. In this case, the flow is absolutely unstable.
In the context of helicopter, it has been suggested that the transition to the VRS could
be associated with a change of the local pairing instability from convective to absolute
(Leishman et al., 2004; Bolnot et al., 2014). For this reason, we shall also consider the
spatial-temporal development of the perturbations in this work.
Wake
− Hypothesis: The flow is assumed inviscid and incompressible. The external flow will
be uniform and aligned with the rotor axis.
Aerodynamic loads
− Model: BEM theory is used with the Kutta-Joukowski law to compute the circulation
and aerodynamic forces distribution along the blades.
27
− Hypothesis: As for the wake, inviscid and incompressible flow is assumed. For the
computation of lift and drag coefficients, the dependence on Reynolds number along
the blades is not taken into account, the value at the tip is used.
Blades deformation
− Model: Beam theory is applied combined with a rod model to compute the deformation
of blades.
− Hypothesis: Linear elastic properties are assumed, as well as all beam theory conditions
concerning cross section.
At all levels steadiness is imposed, looking for stationary solutions for the whole coupled
problem.
In chapter 2, the model for the wake is analyzed. Two different models are presented
and compared: a classical Joukowski model, where two vortices are detached per blade,
one at the tip and one at the axis, and a generalized Joukowski model where the central
axis is not located on the axis but displaced on the blade. Both models are compared with
experimental and numerical results. Then, in chapter 3, a stability analysis is presented
for Joukowski model solutions in several rotor regimes, comparing it with analytical results
from the bibliography. An analysis of the convectively/absolutely unstable solutions is
also presented, identifying the regimes where this transition occurs.
In chapter 4, Kutta-Joukowski law and blade element theory are applied to obtain
the circulation profile and the aerodynamic forces exerted along the blade. A particular
methodology is presented to calculate the circulation of the wake vortices from the cir-
culation profile on the blade. The solutions obtained for the wake in chapter 2 are used
to calculate the flow in the rotor plane. The numerical results obtained from the present
model are compared to numerical and experimental results found in the literature
Finally, in chapter 5, the models for the blade deformation are developed and coupled
with the induction effect from the wake. A strong fluid-structure coupling is implemented,
looking for a convergence between the stationary solutions for the wake structure and the
blade deformation.
28
Chapter 2
Wake model
Contents
2.1 Chapter overview . . . . . . . . . . . . . . . . . . . . . . . . . . 30
2.2 Generalized Joukowski model . . . . . . . . . . . . . . . . . . . 30
2.3 Far-wake solutions . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2.3.1 Vortex filament framework . . . . . . . . . . . . . . . . . . . . 31
2.3.2 Deformed helical vortex pairs . . . . . . . . . . . . . . . . . . . 39
2.3.3 Analysis of the effect of a varying core size . . . . . . . . . . . 50
2.3.4 Discussion in the context of rotors . . . . . . . . . . . . . . . . 50
2.4 Near-wake solutions . . . . . . . . . . . . . . . . . . . . . . . . . 53
2.4.1 Framework . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
2.5 Rb∗ = 0 case (Standard Joukowski model) . . . . . . . . . . . . 55
2.5.1 Framework . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
2.5.2 Wake solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
2.5.3 Induced velocities . . . . . . . . . . . . . . . . . . . . . . . . . 60
2.5.4 Energy extracted by a wind turbine . . . . . . . . . . . . . . . 65
2.6 Rb∗ 6= 0 case (Generalized Joukowski model) . . . . . . . . . . . 67
2.6.1 Topology of the solutions . . . . . . . . . . . . . . . . . . . . . 67
2.6.2 Parametrical analysis . . . . . . . . . . . . . . . . . . . . . . . 68
2.6.3 Comparison with the standard model . . . . . . . . . . . . . . 74
2.7 Chapter conclusion . . . . . . . . . . . . . . . . . . . . . . . . . 76
29
(a)
(b) (c)
Figure 2.1: (a) PIV visualization of the wake generated by a rotor in wind turbine regime
(Quaranta et al., 2015). In yellow, it can be observed the hub helical vortex. (b) Numer-
ical calculation of a three-bladed wind turbine wake from Ivanell et al. (2010) using the
Actuator Line Method. (c) Scheme of the generalized Joukowski rotor model. To simplify
the figure, only the vortices detached from a single blade are represented.
30
and two free vortices of opposite circulation detached from the axis and the tip of the
blade. For a N blade rotor, these vortices form in the far-wake a uniform helical braid
composed of N vortices of circulation Γ plus a central vortex of circulation −N Γ (figure
1.7b). However, in many actual rotors, it can be observed that the internal vortex is not
necessarily placed on the axis, but displaced on the radial direction, generating another
helical vortex. In figure 2.1, experimental and numerical examples are shown where this
configuration appears. In this section, a generalized Joukowski model taking into account
this characteristic of the wake is presented. From each blade, the wake is composed of a
bound vortex on the blade, and two free helical vortices, one detached from the tip with
a circulation Γ and one detached near the axis with a circulation −Γ (figure 2.1c). These
two helices will not necessarily have the same pitch.
In the standard Joukowski model, the far-wake is spatially homogeneous. It is known
that helical vortices with an uniform external axial flow keep undeformed under their own
induction (Kida, 1981). With a straight hub vortex on the rotation axis, the angular
rotation of the helix is modified but not its shape. However, if the hub vortex is not on
the axis but has the form of a helix, the mutual induction between tip and hub vortices
will not be homogeneous any more: a radial induced velocity will appear on the helices
and they will be deformed. In addition, because of this mutual induction, the axial and
azimuthal flow on each vortex will not be the same and the helices will normally have
different pitches.
Finding the far-wake solution when the hub vortex is not on the axis is therefore
already an issue –this problem is treated in the next section–. These far-wake solutions
will be obtained calculating the structures resulting of the interaction of infinite helical
vortex pairs. The details of the calculation of the near wake and its coupling with far-wake
solution will be presented in section 2.4.
31
(b)
(a)
Figure 2.2: Discretization procedure of the vortex filaments. (a) Discretization in seg-
ments of two filaments of circulation Γi and Γj . (b) Arc of circle formed by three consec-
utive points of a discretized filament for the computation of the local contribution.
special cases: first to single helices to validate the method, then to non-interacting helical
pairs to introduce the geometrical parameters that are used to define the solutions.
Lagrangian description
We consider small core size vortices which can be described as vortex filaments. In this
framework, all the vorticity is concentrated along lines which move as material lines in
the fluid according to
dξ
= U (ξ) = U ∞ + U ind (ξ), (2.1)
dt
where ξ is the position vector of the vortex filament, U the velocity field, composed of an
external field U ∞ and a field U ind (ξ) induced by the vortex filaments. When there are
N vortices, this induced velocity is given by the Biot-Savart law
N
(ξj − ξ) × dτj
Z
ind
X Γj
U (ξ) = 2 , (2.2)
j=1
4π |ξ j − ξ|
where the integrals cover each vortex filament defined by its circulation Γj , its position
vector ξj and its tangent vector τj .
On the vortex line, the Biot-Savart integral is singular, and the self-induced velocity
diverges. To avoid this singularity, one has to assume a small but finite core size a.
The self-induced motion is then obtained by an integral of the same form but without
considering the interval [−δa, δa] around the singular point. This so-called cut-off method
is explained in length in textbooks (see Saffman, 1992). The value of δ depends on the
vortex core model. Here, we shall assume a Gaussian vorticity profile for which δ ≈ 0.8736.
Vortex discretization
We follow the vortex method approach described for instance in Leishman (2006). Each
vortex filament is discretized in small segments in order to compute the velocity field and
follow its displacement (see figure 2.2a).
32
The velocity field induced by a given segment [ξin , ξin+1 ] of the ith vortex at a point
ξjm can be calculated explicitly as
pn
N X
X
U ind
(ξjm ) = loc
Uj,m + Useg m
i,n (ξj ), (2.5)
n=1 i=1
where pn is the number of points discretizing the nth vortex, and assuming implicitly that
Useg m seg m
j,m (ξj ) = Uj,m−1 (ξj ) = 0 .
This formula is tested against direct calculation of the cut-off integral in figure 2.3 for a
single helix (see also Gupta & Leishman, 2005a). We observe that a good approximation
is obtained as soon as the helix is divided in 25 or more segments by turn when the
local contribution is included. When the local contribution is not taken into account, a
much larger number of segments by turn of order O(2πρ/a) is needed to obtain a good
approximation. In practice, we use pn = 30 in most calculations.
33
0.4
0.2
0
Ub
−0.2
−0.4
−0.6
0 20 40 60 80 100
pn
Figure 2.3: Comparison of the cut-off formula (dashed red line) with the approximate
formula (2.5) (solid black line). Binormal component of the induced velocity versus the
number pn of segments by turn for an helix of circulation Γ = 1, pitch h = 1, radius
R = 1, and core size a = 0.05. The two contributions Uloc and Useg are also indicated in
dash-dot and dotted lines, respectively.
0.1 10 2
0.05
10 1
0
10 0
ΩR2 /Γ
W R/Γ
-0.05
10 -1
-0.1
10 -2
-0.15
-0.2 10 -3 -1
10 -1 10 0 10 1 10 2 10 10 0 10 1 10 2
h/R h/R
(a) (b)
Figure 2.4: Non-dimensionalized rotation rate ΩR2 /Γ (a) and axial speed W R/Γ (b) of a
right-handed helix of circulation Γ, radius R, pitch h and core size a as a function of h/R
for a/R = 0.03. The solid black curves correspond to the numerical results obtained in
this work. The solid grey curves correspond to the theoretical expressions (2.6-2.7), the
dashed grey curves correspond to the results obtained by Velasco Fuentes (2018), that is
the same expressions without the 1/4 terms. Both theoretical results are for an equivalent
Rankine vortex of core size a/R = 0.0408.
34
local curvature (Ricca, 1994; Kuibin & Okulov, 1998; Boersma & Wood, 1999):
p
ΩR2 ln(2/) + 2(1 + p2 ) − ln( 1 + p2 ) − (1 + p2 )3/2 [2/p − W(p)] − 1/4
= , (2.6)
Γ 4π(1 + p2 )3/2
p
WR ln(2/) − ln( 1 + p2 ) + (1 + p2 )3/2 W(p) − 1/4
= , (2.7)
Γ 4π(1 + p2 )3/2
where p = h/(2πR), = a/[R(1 + p2 )] and W(p) is the function defined in Boersma &
Wood (1999) by
∞
sin2 t
H(1/2 − t)
Z
1
W(p) = 2 2 2 3/2 − 2 3/2
dt. (2.8)
0 (p t + sin t) (p + 1) t
This correction term permits to take into account the deformation of the vortex core in-
duced by curvature. In these theoretical works, a Rankine vortex model (uniform vorticity
in the core) is used, while we use a Gaussian vorticity profile. We have thus applied the
correction factor aRankine ≈ 1.36 aGaussian to the core size to account for the different vortex
models (Widnall, 1972; Saffman, 1992). As it can be seen on this figure, the agreement
between the numerical results and Velasco Fuentes (2018) is good and almost perfect
for both Ω and W when the correction term is included. This comparison is a strong
validation of our numerical approach.
For both rings and helices, there exist infinitely many other moving frames where the
vortex structure is steady. The displacement associated with this frame just has to remain
tangent to the structure. The condition of steadiness for the frame velocity VF can then
be written as
(VF (ξjm ) + Uind (ξjm )) × τjm = 0. (2.9)
In this frame, the vortex elements are moving along the vortex structure. For a ring, any
rotation around the ring axis can for instance be added. For an helix of pitch h, any
rotation and translation along the helix axis can also be added if the rotation rate Ωa and
axial speed Wa of this additional movement satisfy
where the sign is + for right-handed helices, and − for left-handed helices.
An helical braid composed of N identical vortices of same axis, separated with each
other by an azimuthal angle 2π/N also forms a steady solution in an adequate frame.
When N 6= 1, a straight hub vortex placed on the central axis can be added without
introducing any deformation on the helices. This is not possible when N = 1. The
external helix indeed generates an horizontal velocity on the axis that induces a radial
displacement of any straight structure place at this position. In that case, we expect the
hub vortex to deform and to move out from the rotation axis. Our objective is to describe
such a structure, as well as the structures composed of N vortex pairs with root vortices
not on the rotation axis. We shall see that there still exist steady solutions in those cases.
35
Figure 2.5: Configuration of two undeformed co-axial helical vortices. Here both helices
have the same orientation: κ = 1.
1 1 κ
= − , (2.11a)
L hint hext
L L
φ̃ = 2π −E , (2.11b)
min(hint , hext ) min(hint , hext )
where E(x) denotes the integer part of x. When κ = −1, L is smaller than hext and hint :
both helices have performed less than a complete rotation in one axial period L. When
κ = 1, L is always larger than the smaller pitch; the helix with the smaller pitch has then
performed more than a complete rotation in a period. The other helix has performed just
one complete revolution less.
When hext 6= hint , the undeformed helical pair corresponds to a steady solution only
if the mutual induction of one helix on the other is negligible. This occurs when the core
size becomes sufficiently small. In this limit, the locally induced velocity is the dominant
contribution to the induced velocity which then becomes constant and oriented along the
local binormal vector. Each helix then behaves as if the other helix was not present. A
priori, they rotate and translate at different speeds. But, owing to the possibility to add
any displacement along the helical line, it is possible to find a frame where both helical
structures become steady. The frame selection is graphically explained in figure 2.6b.
In that figure, we plot each helix in the (φ, z) plane at t = 0 (solid lines) and t = 1
36
(a)
(b)
Figure 2.6: (a) Self-induced velocities of a pair of coaxial vortices when the mutual induc-
(0)
tion is negligible. (b) Schematic diagram showing how the moving frame velocities ΩF
(0)
and WF are obtained. Solid and dashed lines represent the two helices at t = 0 and t = 1
respectively. The frame velocity is given by the vector connecting the crossing points of
solid lines with dashed lines. The axial and azimuthal components of this vector provides
(0) (0)
WF and ΩF respectively.
(dashed lines) using two different colors. Each helix corresponds to a straight line with
a slope equal to the helix pitch. The self-induced velocity of each helix, together with
their decomposition on the axial and azimuthal direction is also indicated. Any vector
that connects any two points from the lines at the two distinct times provides a possible
frame velocity vector that keeps the considered helix steady. The vector that keeps both
helices steady is the one that connects the crossing points associated with each instant.
Such a vector exists as soon as the helix lines are not parallel in the (φ, z) plane, that is
if hext 6= hint or κ = −1.
(0) (0)
The angular velocity ΩF and axial velocity WF of the frame are given from the
self-induced velocity of each helix by
SI SI
(0) 2π(Wint − Wext ) + (ΩSI SI
ext hext − Ωint κhint )
ΩF = , (2.12a)
hext − κhint
(0) SI hext (0)
WF = Wext + (Ω − ΩSI
ext ). (2.12b)
2π F
The second equation shows that condition (2.10) is satisfied by the external vortex. It is
immediate to obtain
(0) SI κhint (0)
WF = Wint + (ΩF − ΩSI int ), (2.13)
2π
37
Parameters defining the deformed helical structures
The two-helix structure obtained above is no longer a solution if the two helices interact.
Indeed, the velocity field of one helix on the other contains a radial component that
moves the structure radially. Each helix is therefore expected to be deformed by the field
induced by the other helices. Being inspired by the non-interacting solutions, we shall
search for steady solutions that still exhibit a spatial periodicity. We consider solutions
composed of N pairs of counter-rotating vortices (the external vortex having a positive
circulation Γ, the internal a negative circulation −Γ), with a 2π/N azimuthal symmetry.
By construction, we then assume that the solutions are invariant by the transform φ →
φ + 2π/N . We also assume that there exist an axial distance L > 0 and an angle φ̃
satisfying 0 ≤ φ̃ < 2π/N , such that the solutions are invariant by the double operation
z → z + L and φ → φ + φ̃. We do not want the solution to repeat several times in a spatial
period, so we further assume that there is a single location in an axial period L where
internal and external vortices are at the same azimut. We shall choose this particular
azimut to define from their radial positions the radius Rint and Rext of the internal and
external vortex (Rint < Rext ). We also define the mean pitch hint and hext for each vortex
from the azimuthal angle covered in an axial period. For the external vortex, if this angle
is φext , we have hext = φ2πL
ext
. If we add the vortex core size a that we assume identical and
constant for all the vortices, we obtain 5 spatial length scales from which we can form 4
independent non-dimensional parameters:
Rint hext hint a
R∗ = , h∗ = , α= , ε= . (2.14)
Rext Rext hext Rext
To these 4 real positive parameters, we should add the number N of vortex pairs and the
index κ = ±1 that defines the relative orientation of internal and external vortices. In
most cases, we shall keep κ = 1 and N = 1. The parameter R∗ will be varied between
0 and 0.75, h∗ between 0.1 and 2, α between 0.5 and 2. The parameter ε will always be
considered small, and typically equal to 0.03. Even for this small value of ε, finite core
size effects can become important if h and α are too small. This provides a limitation on
the values of the parameters that we can consider. Here, only the extreme cases (h ≈ 0.1
and α ≈ 0.5) are expected to give rise to consequent finite core size effets.
In the paper, the vortex core size is also assumed to be constant. This approximation
is discussed in section 2.3.3.
Note that both the normalized period L/Rext and the angle φ̃ can be obtained from
the above geometrical parameters:
h∗
L 2π 1 1
= , φ̃ = −E . (2.15)
Rext N |1/α − κ| N |1/α − κ| |1/α − κ|
With each solution is associated a moving frame where the solution is steady. From
the angular velocity ΩF and the axial velocity WF of the frame, we can construct two
other dimensionless parameters using the external radius Rext and the circulation Γ of the
vortices:
R 2 ΩF Rext WF
Ω = ext , W = . (2.16)
NΓ NΓ
These two parameters characterizing the frame velocity are functions of the 6 geometrical
parameters R∗ , h∗ , α, ε, N and κ.
38
2.3.2 Deformed helical vortex pairs
In this section, we describe the deformed helical structures as the geometrical parame-
ters vary. After having introduced an approximated solution, we successively study the
geometrical characteristics, the frame velocity and the induced velocity.
dr Vrind dφ Ωind − ΩF
= ind , = ind . (2.17)
dz Vz − WF dz V z − WF
For the mutual induction, we use the formula given by Hardin (1982) for a perfect helix.
(1) (1)
The corrected frame velocities ΩF and WF are obtained by using the definition of L for
the internal and external vortex:
Z L I (0) (0) (1)
(1) (1) 2πL ΩM SI
int (rint , φint (z), z) + Ωint − ΩF
φint (L) − φint (0) = = (0) (0) (1)
dz , (2.19a)
hint MI SI
0 Vz,int (rint , φint (z), z) + Vz,int − WF
Z L I (0) (0) (1)
(1) (1) 2πL ΩM SI
ext (rext , φext (z), z) + Ωext − ΩF
φext (L) − φext (0) = = (0) (0) (1)
dz . (2.19b)
hext MI SI
0 Vz,ext (rext , φext (z), z) + Vz,ext − WF
Note that equations (2.19a,b) give (2.12) if we neglect the mutual induction. The condition
of periodicity of the radial deformation does not give an additional constraint because it
is automatically satisfied for each vortex.
This simple first order approximation for the helix deformation is compared to nu-
merical results for two typical examples in figure 2.7. We clearly see that the agreement
strongly depends on the pitch ratio α. This approximation tends to underestimate the
deformation of the external vortex but to overestimate that of the internal vortex. The
39
1.02
1 External
1
vortex
0.98
0.98 External
0.96 vortex
r/R
r/R
0.94
0.96
Internal 0.92 Internal
vortex vortex
0.94 0.9
0.88
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
z /L z /L
(a) (b)
Figure 2.7: Radial position of the internal and external vortices for a single counter-
rotating helical pair for (a) R∗ = 0.5, κ = 1, h∗ = 1, α = 0.9 and ε = 0.03 (b) R∗ = 0.5,
κ = 1, h∗ = 1, α = 1.4 and ε = 0.03. Solid line: numerical solution. Dashed line: first
order approximation.
* * *
40
(a) R∗ = 0.3 (b) R∗ = 0.45 (c) R∗ = 0.60 (d) R∗ = 0.75
error is always larger for the internal vortex. The maximum error can then be quantified
using
(1)
(1) max(|rint − rint |)
Er,int = (2.20)
Rint
which measures the maximum deviation between numerical and first order solutions. This
quantity is plotted in figure 2.8 as a function of h∗ for N = κ = 1 and various α, R∗ and
ε. This figure shows that except for very large R∗ (R∗ = 0.7), the error increases with
h∗ . The error also increases with ε, R∗ and with the distance of α to 1. It becomes
particularly large (superior to 30%) for large α and large h∗ . Note however that the error
remains small for α ≈ 1, small R∗ and small ε.
41
0
10
−2
, ∆r max
10
ext
∆r max
int
−4
10
−6
10
0.1 0.2 0.3 0.4 0.5 0.6
R∗
(a) (b)
(c)
Figure 2.10: Representation of 2 vortex pairs (a) and 3 vortex pairs (b) for h∗ = 1,
R∗ = 0.4, ε = 0.05, α = 2, κ = 1. (c): Maximum deformation ∆rmax int ext
(red) and ∆rmax
(black) versus R∗ for 1 vortex pair (solid lines), 2 vortex pairs (dashed lines), 3 vortex
pairs (dash-dot lines). h∗ = 1, α = 1.5, κ = 1 and ε = 0.05.
* *
ext
(a) ∆rmax int
(b) ∆rmax
Figure 2.11: Maximum deformation contours in the (α, h) plane for R∗ = 0.5, ε = 0.03,
int ext
κ = 1, N = 1. The thick solid line corresponds to the line ∆rmax = ∆rmax .
42
α = 0.95 2.2
1 α = 1.05
2
α=1
1.8
r/R ex t
0.995 1.6
L/W F
1.4
0.99 1.2
(a) (b)
with the respect to z axis. They therefore mainly induce a velocity field in the radial and
azimuthal directions. The ratio Vr /Vz that defines the slope of the deformation thus gets
int
large, implying large ∆rmax ext
and ∆rmax . In the opposite, when h∗ goes to zero, Vr /Vz goes
to zero as well: the helical vortices are thus no longer deformed.
Concerning the effect of α, the increase of the deformation of the external helix with
α can be understood by the same argument. It is associated with a decrease of hint , and
int
therefore an increase of the ratio Vr /Vz . The variation of ∆rmax with respect to α is less
∗ int
simple. For h < 1, ∆rmax is maximum for α close to 1. This value α = 1 is special
for κ = 1 because the axial period L and the frame axial velocity WF get infinite [see
equation (2.15)]. It therefore corresponds to a singular limit in our description. However,
the radial positions rint and rext of internal and external vortices can still be plotted as a
function of z/L and we obtain a well-defined curve as α → 1. Similarly, WF /L converges
to a non-zero constant as α → 1, which means that a finite time T = L/WF is needed to
advect a perturbation on the period L at the velocity WF . The singular case α = 1 could
therefore be described by an alternative way by considering the solution in the frame
moving at the velocity WF . In this frame it should correspond to the temporal dynamics
of perfect helices with the same R∗ , h∗ and ε. This is indeed what we have checked in
figure 2.12. In figure 2.12(a), we show that the variation of rint as a function of z/L for α
close to one, is well described by the variation of rint as a function of t/T in the temporal
problem. We also check in figure 2.12(b) that L/WF converges to the temporal period
obtained in the temporal problem as α → 1.
Structure velocities
As explained above (§2.3.2), we can obtain different approximations of the frame veloc-
ity ΩF and WF by neglecting the helix deformations and/or the mutual induction. The
43
* *
(a) (b)
Figure 2.13: Frame velocity versus h∗ for R∗ = 0.5, ε = 0.03, α = 1.4, κ = 1, N = 1. (a)
2
Angular velocity Ω = Rext ΩF /Γ, (b) axial velocity W = Rext WF /Γ. Solid line: numerical
(1) (1)
result. Dashed line: first order approximation (ΩF , WF ). Dash-dotted line: leading
(0) (0)
order approximation (ΩF , WF ).
(0) (0)
leading order approximation (ΩF , WF ) neglects both the mutual induction and the de-
(1) (1)
formation. The first order approximation (ΩF , WF ) neglects the vortex deformation
but takes into account the mutual induction. This approximation is obtained by solv-
(1) (1)
ing the equations (2.19a,b) for ΩF and WF . In figure 2.13, we have compared both
approximations to numerical results for a typical case. We clearly see that the leading
order approximation does not capture the variations of ΩF and WF with h∗ , while the
first order approximation follows both qualitatively and quantitatively these variations.
By comparing other configurations, we have observed that the first order approximation
provides a good approximation of ΩF and WF as soon as R∗ ≤ 0.5 and ε ≤ 0.1. In prac-
tice, these approximations have been used as guess values for the numerical calculation.
The variations of the frame velocities Ω and W with respect to h∗ and α are shown in
figure 2.14. Both contour plots exhibit similar features: a same singularity line α = 1 and
a single contour where Ω and W vanish. These zero level contours are different for Ω and
W but both cross the singular line α = 1 at the same value (here h∗ ≈ 0.175). This special
point on the singular line α = 1 corresponds to the particular solution obtained by Walther
et al. (2007). For these parameters, the vortices are undeformed helices of same pitch.
There therefore exist infinitely many frames where the helices are stationary as any values
of axial speed Wa and angular velocity Ωa can be added provided (2.10) is satisfied. This
explains the degeneracy observed at this point in figure 2.14. These qualitative features
do not depend on ε and R∗ . It is interesting to note that a contour of Ω may cross
twice a contour of W (look at the saddle point region close to (α, h) ≈ (1.7, 0.4)). The
coordinates of the two crossing points then correspond to couples of parameters (α, h)
having the same frame velocities (Ω, W ).
By contruction, the vortex elements are advected along the stationary vortex structure.
This tangential velocity is different for the internal and the external vortex and varies
44
* *
(a) (b)
Figure 2.14: Contour values of the frame velocity in the (α, h) plane for R∗ = 0.5, ε = 0.03,
2
N = 1 and κ = 1. (a) Angular velocity Ω = ΩF Rext /Γ, (b) Axial velocity W = WF Rext /Γ.
The dashed line (α = 1) indicates a line where Ω and W are not defined. The thick solid
curve corresponds to the level zero.
along the vortex structure, as illustrated in figure 2.15a. This variation is associated with
the deformation of the helices. It is then important when the level of deformation is high.
For each vortex, we define a mean tangential velocity V̄tan and a measure ∆Vtan of the
fluctuation around this mean using
1 z0 +L max |Vtan − V̄tan |
Z
V̄tan = Vtan (z)dz , ∆Vtan = . (2.22)
L z0 |V̄tan |
ext int
The measures ∆Vtan and ∆Vtan for the external and internal vortex are plotted as a
∗
function of h in figure 2.15b for a few cases. We do observe an increase of the tangen-
tial velocity fluctuation with h∗ and R∗ , in agreement with the increase of the vortex
deformation (see figure 2.11).
The mean tangential velocity of each vortex is shown in the (α, h) plane in figure 2.16.
A positive value corresponds to an advection in the positive axial direction, a negative
value to an advection in the opposite direction. Not surprisingly, the mean tangential
velocity blows up as α → 1 like ΩF and WF . It is also interesting to note that the contour
2
curves are similar (in shape) for both vortices and close to those obtained for ΩF Rext /Γ
in figure 2.14(a).
In figure 2.17a, we have displayed on the same plot the parameters for which mean
tangential velocities and ΩF vanish. We clearly see that mean tangential velocities and
ΩF vanish for almost the same parameters. This means that there is a very small region
of parameters around the line ΩF = 0 where internal and external vortices propagate
in different directions. This region is delimited by the solid lines shown in figure 2.17a.
Everywhere else, both vortices propagate in the same direction. We shall see in section
2.3.4 that this condition on the direction of propagation of the vortices is necessary for
the solution to describe the flow generated by a rotor.
It is interesting to compare more precisely the mean tangential velocity with the veloc-
ity associated with the moving frame. Assuming that each structure is approximatively
45
0.35
-2
0.3
-2.5
0.25 R∗ = 0.8
int
, ∆Vtan
-3
Vtan Rext /Γ
0.2
-3.5
ext
∆Vtan
0.15
-4 R∗ = 0.6
0.1
-4.5
0.05 R∗ = 0.3
-5
0 0.2 0.4 0.6 0.8 1 0
1 1.5 2
z/L
h∗
(a)
(b)
Figure 2.15: (a) Variation of the tangential velocity on a period in the internal vortex
(dashed line) and in the external vortex (solid line) for a typical case (R∗ = 0.6, h∗ = 1.5,
α = 1.4 and ε = 0.03, N = 1, κ = 1). (b) Maximum tangential velocity fluctuation in
the internal vortex (dashed lines) and external vortex (solid lines) as a function of h∗ for
different values of R∗ and α = 1.4 , ε = 0.03, N = 1, κ = 1 .
* *
(a) (b)
Figure 2.16: Contour values of the mean tangential velocities in the (α, h) plane for
R∗ = 0.5, ε = 0.03, N = 1, κ = 1. (a) External vortex: V̄tan ext
Rext /Γ. (b) Internal vortex:
int ext int
V̄tan Rext /Γ. The dashed line (α = 1) indicates a line where V̄tan and V̄tan are not defined.
The thick solid curve corresponds to the level zero.
46
*
(a) (b)
Figure 2.17: (a) Value of h∗ versus α for which ΩF = 0 (black dashed line), V̄tan ext
= 0 (blue
int ∗
line), V̄tan = 0 (red line) for R = 0.5, ε = 0.03, N = 1, κ = 1. (b) Comparison of mean
tangential velocity V̄tan (solid lines) and tangential frame velocity VF tan (dashed lines) for
external and internal vortices (normalized by Γ/Rext ). Parameters are R∗ = 0.5, α = 1.2,
ε = 0.03, N = 1, κ = 1.
for the internal and external vortex, respectively. The difference between VF tan and V̄tan
is associated with the vortex induction. In figure 2.17b, we can observe the similar values
of V̄tan and VF tan in the whole range of h∗ between 0.6 and 1.4 for a typical case. This
means that the most important part of the tangential velocity is associated with the frame
velocity and the vortex induction contribution remains in general small.
Induced flow
From the point of view of applications, it is useful to know the velocity field induced by
the vortex structure. An illustration of the axial and angular components of such a field
in a plane perpendicular to the structure axis is shown in figure 2.18. In these contour
plots, the axial velocity WF and angular velocity ΩF have been subtracted such that the
velocity field vanishes far from the center. The vortex cores where the velocity field is
smoothed have also been indicated. We clearly see that the induced velocity field exhibits
strong inhomogeneities.
In figure 2.19, we show the azimuthally averaged flow versus r at different axial loca-
tions. In the core region of each vortex (at a distance smaller that a for the vortex center),
each velocity profile has been replaced by a linear fit. We do see small fluctuations of the
profiles between different locations but the profiles remain close to the profiles generated
by N pairs of perfect helices. These ideal profiles are given (for infinitely thin vortices)
47
1.5 1.5 3
1 1 2.5
0.5 0.5 2
0 1.5
y
0
y
−0.5 −0.5 1
−1 −1 0.5
−1.5 −1.5 0
−1.5 −1 −0.5 0 0.5 1 1.5 −1.5 −1 −0.5 0 0.5 1 1.5
x x
q
(a) Vzind Rext /Γ (b) (Vxind )2 + (Vyind )2 Rext /Γ
1.2
z=0
0 1 z = L/4
z = L/2
0.8
V̄zind hext /(N Γ)
/(N Γ)
Hardin
-0.1
0.6
Ω̄ind Rext
2
0.4
-0.2
z=0 0.2
z = L/4
-0.3 z = L/2 0
Hardin
-0.2
0 0.5 1 1.5 2 0 0.5 1 1.5 2
r/Rext r/Rext
(a) (b)
Figure 2.19: Azimuthally averaged induced velocity profile for the same parameters as
in figure 2.18. Solid lines: numerical results at different axial locations. Dashed line:
theoretical prediction for perfect helices using Hardin model. (a): Angular velocity
Ω̄ind Rext
2
/(N Γ). (b) Axial velocity V̄zind hext /(N Γ).
48
*
Figure 2.20: Contours of mass axial flow rate M (normalized by ρN ΓRext ) in the (α, h)
plane for R∗ = 0.5, ε = 0.03, N = 1 and κ = 1. Contours correspond from top to bottom
to {1, 1.25, 1.5, 2, 2.5, 3, 4, 6, 10}. Solid lines: M ; red dashed lines: M H .
by Hardin (1982)
1
1 − if r < Rint ,
NΓ
α
V̄zH = 1 if Rint < r < Rext , (2.24)
hext
0 if r > Rext ,
2
Rext
N Γ − if Rint < r < Rext ,
Ω̄H = 2 2πr2 (2.25)
Rext
0 elsewhere .
The fluctuations are mainly associated with the radial displacement of the vortices. When
the helices are less deformed, the fluctuations are smaller. It is interesting to note that
the azimuthally averaged axial flow changes sign close to the axis when hint < hext , that
is α < 1.
For the applications, it is also useful to evaluate the mass flow rate induced by the
structure. The mass axial flow rate is defined by
ZZ
M= ρVzind rdrdφ. (2.26)
that is
MH (R∗ )2
π
= ∗ 1− . (2.28)
ρN ΓRext h α
This expression provides a good approximation of the mass flow rate of the deformed
structure, as observed in figure 2.20 for a typical example.
49
Note that M H changes sign when α < (R∗ )2 . In this regime, the induced axial flow
is sufficiently negative close to the axis to inverse the positive mass flow rate occurring
between the internal and external vortices. For the parameters of figure 2.20, this is
expected for very small α (α < 0.25).
As soon as a changes, the self-induced velocity is modified and therefore a different equi-
librium solution is obtained.
In figure 2.21, we have analysed the effect of a varying core size in an extreme case
(R∗ = 0.8, h∗ = 1.4, ε = 0.03, α = 1.4, N = 1, κ = 1). As seen in figure 2.15(b), for this
value of R∗ , the variations of Vtan are the largest: they reach 32% for the internal vortex,
and 23% for the external one. In figure 2.21(a), we have plotted the variation of the core
size in a period for the converged solution in both vortices. The variations of the core size
are weaker than of the tangential velocity as expected from (2.29). They are around 10%
for the internal vortex, and 7% for the external vortex, with respect to the mean core size
a = 0.03. In figure 2.21(b), the radial position of the vortices is shown. The solution with
a varying core size is compared to the solution with a constant core size. We observe that
the difference between both solutions is very small. The largest gaps between both radial
positions are 0.78% and 0.66% for internal and external vortices respectively. In terms
of moving frame velocities, the differences are also extremely small: we obtain the values
W = 1.826 and Ω = 3.697 for a varying core size, while we had W = 1.809 and Ω = 3.610
for a constant core size.
This comparison garantees that the effects of a varying core size is negligible for all
the cases that we have considered. It a posteriori justifies the use of the approximation
of a constant core size in our work.
50
0.034 1
0.9
0.032
0.8
0.03 External
r/Rext
0.7
a/Rext
vortex
0.028 0.6
0.5 Internal
0.026 vortex
0.4
0.024 0.3
0 0.5 1 0 0.5 1
z/L z/L
(a) (b)
Figure 2.21: (a) Variation of the core size a/Rext for the external (solid line) and internal
(dashed line) vortex. The mean core size value in each vortex is a/Rext = 0.03. (b) Radial
position of the internal and external vortices for a constant core size (dashed line) and
for a variable core size obtained using (2.29) (solid line). The parameters in both figures
are R∗ = 0.8, h∗ = 1.4, α = 1.4, N = 1, κ = 1 and a mean core size a/Rext = 0.03.
propagation of the vortices, which is given by the sign of the mean tangential velocity, we
can then deduce the side where the rotor should be. For example, if V̄tan > 0, the vortices
move in the positive direction, the rotor should then be located on the negative side.
By analysing the directions of propagation of the vortices and of the external wind,
one can built the diagram shown in figure 2.22. The different domains are limited by
the curves WF = 0 and ΩF = 0 and the line α = 1, which correspond to changes of
signs of the direction of propagation of the vortices or of the external wind. In this
figure, we have also indicated the typical azimuthally averaged axial flow corresponding
to each regime. For example, in the white region on the right of the line α = 1, WF > 0,
ext int
V̄tan < 0, V̄tan < 0 and M > 0: both vortices moves in the same negative direction as
the external wind, which is opposite to the direction of the mass flow rate. Assuming a
positive axis downwards, this situation corresponds to the so-called windmill brake regime
of an helicopter: the helicopter is going downwards, while the flow and the vortices are
going upwards with respect to the helicopter rotor. This regime also corresponds to the
wind turbine regime. The difference with the other windmill brake regime obtained for
α < 1 is in the azimuthally averaged axial flow which is stronger than the external wind
close to the axis in that case.
In the light gray regions, the vortex velocities are positive while WF < 0. This regime
corresponds to a climbing regime: the vortices moves downwards as the external flow and
the mass flow rate. In the dark gray regions, WF and the vortex velocities are positive as
the mass flow rate. This situation corresponds to a slow descending regime: the vortices
moves downwards while the external flow goes upwards. It does not tell us anything on
the behavior of the vortices close to the rotor. It is not excluded that the vortices exhibit
a complex pattern near the rotor as observed in the so-called vortex ring state (Drees &
Hendal, 1951; Quaranta, 2017; Durán Venegas & Le Dizès, 2018).
Note that the curve ΩF = 0 is not exactly the limit between the windmill brake regime
51
Figure 2.22: Diagram of the different rotor flow regimes. White regions: windmill brake
regime or wind turbine regime (both vortices are going upwards as the external wind).
Light gray regions: ascending regime (both vortices are going downwards as the external
wind). Dark gray regions: slow descending regime and vortex ring state (both vortices
are going downwards while the external wind is going upwards). Close to the line ΩF = 0,
there exist a small region where one vortex is going upwards while the other is going
downwards: such a solution cannot describe the (far) wake of a rotor.
52
and the slow descending regime. There is actually a small region close to this curve where
the solution cannot correspond to any helicopter flight regime. This region has been
displayed in figure 2.17a for a particular set of parameters. In this region, internal and
external vortices move in opposite directions. For this reason, they cannot be created by
a single rotor located far away.
2.4.1 Framework
To obtain the solutions for the near wake, a similar procedure as for the far-wake (section
2.3.1) is followed. Just a few modifications have to be made concerning induced velocity
contributions and boundary conditions.
Under the same assumptions of small core size vortices and inviscid framework, the
concentrated filaments of vorticity move in the fluid as material lines advected by the
velocity field:
dξ
= U (ξ) = U ∞ + U ind (ξ), (2.30)
dt
where ξ is the position vector of the vortex filament, U the velocity field, composed of the
external velocity U ∞ and the induced velocity U ind (ξ) generated by the vortex filaments.
This induced velocity can be in turn decomposed in four velocity contributions:
U ind (ξ) = U tip (ξ) + U hub (ξ) + U blade (ξ) + U FW (ξ), (2.31)
where U tip is the velocity induced by the tip helical vortices, U hub is the contribution
from the hub vortex, U blade is the contribution from the bound vortex on the blade and
U FW is the induction effect from the far-wake. The velocity induced by the vortices is
calculated with Biot-Savart law. The discretization and the application of Biot-Savart
law at each point is done as in section 2.3.
Far-wake contribution
To model the effect of the far wake, it is imposed that, after a certain distance from the
rotor (figure 2.23), the wake adopts the geometry of the far-wake solutions previously
calculated in section 2.3. The imposed far-wake geometry is extended from the end of
the calculation domain for a given number of turns. The far-wake velocity contribution
U FW is calculated as the velocity induced by the far-wake structure on the vortices of the
calculation domain.
53
Figure 2.23: Schematic of the complete vortical structure generated at the wake (to
simplify the figure, only tip vortex from a single blade is represented). In solid line, the
calculation domain. In dashed line, the prescribed far-wake structure.
To obtain the geometrical parameters (2.14) of the far-wake structure, we first calculate
the wake geometry assuming that the vortices in the far wake are perfect helices. So, in
the first iteration, the far-wake contribution is obtained by calculating the induction of
N pairs of undeformed helical vortices. After obtaining this first approximation, the
actual far-wake structure is calculated as in section 2.3 with the following geometrical
parameters:
R∗ = Rhub
FW FW
/Rtip , h∗ = hFW FW FW FW ∗ FW
tip /Rtip , α = hhub /htip , ε = a/Rtip . (2.32)
The value of the parameters is adjusted iteratively until convergence of the solution for
the whole wake.
54
Figure 2.24: Local reference frame used by Leishman et al. (2002)
Problem parameters
Rtip ΩR Γ Rhub a
λ= , η= 2 , Rb∗ = , ε= , (2.36)
V∞ Rtip ΩR Rtip Rtip
where λ is known as the tip-speed-ratio and η and ε represent the vortex strength and the
vortex core size. Note that λ can be either positive or negative; positive values correspond
to descending flight and wind turbine regimes and negative values to ascending flight. The
parameter Rb∗ is the radii ratio of the vortex emission points at the rotor blades. The
parameter ε, which measures the vortex core size, is assumed small as before.
The details for the computation of far-wake and near-wake solutions are presented in
the appendix.
2.5.1 Framework
Stationary solutions for the particular case Rb∗ = 0 can be obtained with the same proce-
dure as presented in section 2.4.1. The only difference is the hub vortex that is considered
as a straight filament of circulation −N Γ instead of a braid of helical vortices. Fixing
the parameter Rb∗ to 0, the problem is just described by four parameters: λ, η, ε and N .
Contrarily to the generalized Joukowski model, N cannot be fixed to 1 for the classical
model. Indeed, the velocity induced by a single helix exhibits on the helix axis a radial
component: a straight vortex placed on the axis cannot remain there. A value of N = 2
will be fixed for standard Joukowski model solutions.
55
Figure 2.25: Map of the rotor regimes as a function of λ and η. The striped band
represents the region where solutions cannot be obtained with the presented model.
In this section, solutions are obtained by a slightly different procedure. The free-
vortex discretization scheme presented by Leishman et al. (2002) will be used, where
vortex position is parametrized through two angular coordinates ψ and ζ:
dξ(ψ, ζ) ∂ξ(ψ, ζ) ∂ξ(ψ, ζ)
= ΩR + = U (ξ) (2.37)
dt ∂ψ ∂ζ
where ψ(t) is the angular position of the blade at a time t and ζ(t) is the total azimuth
of the blade from its position when the element ξ was created (figure 2.24). In a frame
rotating with the blade at the angular velocity ΩR , the steady problem can be expressed
as:
dr 1 dφ 1 dz 1
= Ur , = (Ω − ΩR ), = (Uz + U∞ ) . (2.38)
dζ ΩR dζ ΩR dζ ΩR
where r, φ and z are the radial, angular and axial position of the vortex at ζ and Ur , Ω
and Uz are the respective components of the induced velocity.
This formulation gives geometrically equivalent solutions to those obtained from equa-
tions (2.34). The only difference is a different angular discretization of the vortices (A
comparison of both formulations will be addressed in the appendix). Now, three variables
have to be calculated instead of two, increasing the computational cost; this formulation
will however facilitate the stability analysis of the solutions that is performed in chapter
3.
56
(a) (b) (c) (d)
Figure 2.26: Wake structure for a two bladed rotor in (a) λ = −10 ascending flight, (b)
λ = ∞ hovering flight, (c) λ = 19 descending flight and (d) λ = 11 VRS - descending
flight. For all cases η = 0.05 and ε = 0.01.
Figure 2.27: Wake structure for a two bladed rotor in windmill brake flight configuration
(a) λ = 3.3, (b) λ = 4.3, (c) λ = 5.1 and (d) λ = 6.1. For all cases η = 0.05 and ε = 0.01.
57
1 5
λ =3.3
4 λ =4.3
0 λ =5.1
λ =6.1
3
-1
z/Rb
z/Rb
2
-2
1
λ =-10
-3 λ =-Inf
λ =19 0
λ =11
-4 -1
0.6 0.8 1 1.2 1 1.5 2
r/Rb r/Rb
(a) (b)
Figure 2.28: Radial position of the wake in the axial direction for different tip-speed-ratio
values. With η = 0.05 and ε = 0.01.
1 1.5
0.5 1
h/Rb
h/Rb
0.5
0 λ = -10.0 λ = 3.3
λ=∞ λ = 4.3
λ = 19.0 λ = 5.1
λ = 11.0 0 λ = 6.1
-0.5
0 5 10 0 5 10
φ/2π φ/2π
(a) (b)
Figure 2.29: Evolution of the local pitch along the wake for different tip-speed-ratio values.
With η = 0.05 and ε = 0.01.
goes upwards, the downwards flow induced by the vortices is stronger. If the descending
speed is increased, the vortices can be emitted upwards and then change direction to finally
go downwards and cross the rotor plane. Owing to the displacement of the vortices above
and below the rotor, we denote this regime as the VRS regime. With this definition,
the transition to the VRS regime corresponds to the situation where the vortices are
emitted towards the center tangentially to the rotor plane. The VRS regime is reached at
a different value of λ depending on η. When η is small, we observe that the VRS regime
appears for a negative λ (see figure 2.25, where the VRS frontier crosses the hovering
flight line 1/λ = 0 at η = 0.052, for ε = 0.05).
If 1/λ increases, there is a critical value of 1/λ above which a downwards wake ceases
to exist. It is above a second critical value of 1/λ that a new solution with a wake
going upwards is obtained. For these large values of 1/λ, the wake expands radially. It
corresponds to the windmill brake regime or wind turbine regime. In figure 2.27, some
58
2
1.5
0.5
0 0.5 1 1.5 2
1.5
0.5
0 0.5 1 1.5 2
Figure 2.30: Time-averaged dye visualization of the wake behind half of the rotor plane in
a wind turbine configuration approaching to VRS at λ = 7.23 (upper) and λ = 11.2 (lower)
(from Quaranta, 2017). Right plots: qualitative comparison of the wake geometry with
the numerical model for wind turbine configuration at high tip-speed-ratio (up: λ = 8,
down: λ = 10. η = 0.02, ε = 0.01). The rotor disc is represented in black dashed line. In
black solid line, the projection of the tip vortices and, in red, their radial evolution.
vortical structures of wind turbine regime are plotted for different values of λ and, in
figure 2.28b, their radial trajectories. As observed for the helicopter flight regimes, the
stronger the external flow is (that is, the smaller λ), the less deformed the structure is.
For slow external flow in the wind turbine regime, there is a limit value of λ –just before
the interruption of the solutions– from which the vortices are emitted downwards and
then cross the rotor plane going upwards (see the case of λ = 6.1 in figure 2.28b). This
kind of structure was found experimentally by Quaranta (2017) just before the transition
to VRS from a windmill brake regime (figure 2.30). Information on vortex circulation and
core size have not been measured in the experiment, so a quantitative comparison cannot
be performed. However a very good qualitative agreement with the wake geometry is
found using the numerical model
The intervals of λ of existence of downward and upward structure do not seem to
overlap. For all the considered parameters, it was found an interval of λ where both
59
Figure 2.31: Contour maps of (a) radius and (b) pitch of the helix in the far wake,
non-dimensionalized by Rtip . With ε = 0.01.
60
Figure 2.32: Induced velocity contours in a cross section (z=0 plane) for η = 0.05, ε = 0.01
and λ = −15 (left plots) λ = ∞ (center), λ = 11 (right). η = 0.05, ε = 0.01. Upper plots:
Vzind Rtip /Γ, lower plots: Vφind Rtip /Γ.
6 5
λ = -20 λ = -20
λ =∞ λ =∞
4 λ = 11 λ = 11
λ = 4.3 λ = 4.3
V̄φ Rb /Γ
V̄z Rb /Γ
2 0
-2 -5
0 0.5 1 1.5 0 0.5 1 1.5
r/Rb r/Rb
(a) (b)
Figure 2.33: Azimuthally averaged induced velocity in axial and azimuthal directions in
the rotor plane for different values of λ with η = 0.05 and ε = 0.01.
61
4 5
4
3
3
V̄φ Rb /Γ
V̄z Rb /Γ
2 2
1
1
0
0 -1
0 0.5 1 1.5 0 0.5 1 1.5
r/Rb r/Rb
(a) (b)
Figure 2.34: Azimuthally averaged induced velocity in axial and azimuthal directions in
the rotor plane (solid black) and in the far-wake (dashed red) for λ = −20, η = 0.05 and
ε = 0.01. For the axial velocity, the external flow is also included.
ture. Figure 2.32 shows the axial and angular components of the velocity field induced
at the rotor plane for different helicopter regimes. In these contour plots, the external
velocity V∞ and the angular velocity ΩR have been subtracted such that the velocity
field vanishes far from the center. For a clearer visualization, the induction of the bound
vortex has not been plotted. We can see that the homogeneous character of the induced
velocity field varies with λ. For climbing and hovering configurations (figure 2.32, left
and center), the velocity field is approximatively homogeneous inside the rotor disc, with
strong variations near the vortices. However, for VRS regimes, the vortices cross the rotor
disc producing a very heterogeneous velocity field (figure 2.32, right). For all cases, as
for infinite helices (Hardin, 1982; Saffman, 1992), the induced flow is mainly concentrated
inside the rotor disc, and vanishes rapidly outside of it.
Figure 2.33 shows the azimuthally averaged flow versus r in the rotor plane for different
rotor regimes. In the core region of each vortex (at a distance smaller that a from the
center of the vortex), each velocity profile has been replaced by a linear fit. For the
axial flow, we obtain similar profiles in almost all the regimes, with a constant upwards
flow inside the rotor disc and a varying downwards flow outside that vanishes far from
the rotor disc. The total flow grows when the climbing velocity is reduced and also
in soft descending flight and VRS regimes. In VRS, two different regions are observed
inside the rotor disc, that are delimited by the location where the vortices cross the rotor
plane. For the induced azimuthal velocity, the flow profiles are very similar for all the
regimes, in clockwise direction for wind turbine regime and counterclockwise for the rest.
Outside the rotor disc, the azimuthal flow is inexistent; inside, it is not zero but it vanishes
asymptotically as V̄φ ∼ 1/r. For λ = −20 and λ = ∞, both profiles are almost equivalent.
In VRS, it is also equivalent in the inner region, inside the position where the vortices
cross the rotor disc; from this point to the border of the rotor disc, it changes direction
and overlaps with the wind turbine curve.
In figure 2.34, the azimuthally averaged flow is compared in the rotor plane and
62
Figure 2.35: Induced velocity variation as a function of climb and descent velocity. Com-
parison with momentum theory and experimental data from Washizu et al. (1966). In
coloured background, the regions where momentum theory applies.
the far-wake for a climbing flight case (λ = −20). For both cases, similar profiles are
found. The induced axial flow is constant inside the wake and the rotor disc and vanishes
outside. As expected for a climbing flight regime, the flow is accelerated and the induced
velocity is higher in the far-wake. In this case, the total velocity in the rotor disc has
approximately the mean value between the external flow and the velocity in the far-wake,
so the description of momentum theory for the flow on the rotor (1.5) is verified. The
induced azimuthal flow is also similar for both cases. It is zero outside the vortices and
it vanishes as V̄φ ∼ 1/r inside. In non-zero regions, the azimuthal induced velocity is the
double at the far-wake compared to the rotor plane: V̄φFW r/Γ ' 0.3182 and V̄φR r/Γ '
0.1591
It is common in the literature (see Leishman, 2006) to use the averaged induced velocity
Vi crossing the rotor disc to estimate the performance of a rotor. More specifically, one
considers the induced velocity ratio Vi /Vh for each climb velocity ratio V∞ /Vh , where Vh
is the induced velocity Vi in the hovering regime. In our case, we calculate the induced
velocity as: Z Rb
Vi = rV̄z (r)dr/Rb2 . (2.39)
0
As it was observed in figure 2.33, the averaged velocity inside the rotor disc is almost
constant, so the induced velocity Vi will be very close to the averaged velocity observed
on the axis. This does not happen for VRS regimes, where there are different flow regions
inside the rotor disc.
The induced velocity ratio can be approximated with the Momentum theory (Leish-
man, 2006), given by the Froude-Rankine limit. However, this approximation is not valid
for soft descending flight and VRS regimes. For these particular cases, the performance
of the rotor is normally measured experimentally. Figure 2.35 shows the induced velocity
ratio against the climb velocity ratio and compares it with the Froude-Rankine limit and
some experimental results from Washizu et al. (1966). The agreement with the momen-
tum theory is very good for climbing flight and wind turbine regimes, that is, for climb
63
(a) (b) (c) (d)
Figure 2.36: Streamlines of the averaged azimuthal flow for (a) λ = −10 (b) λ = ∞ (c)
λ = 20 (d) λ = 4 with η = 0.05 and ε = 0.01. Black solid line: induced flow. Black
dashed line: total flow. Red solid line: radial position of the tip vortex. Black dotted
line: rotor disc.
velocity ratios greater than 0 and smaller than -2. For soft descending and VRS regimes,
that is, for −2 < V∞ /Vh < 0, the calculated induced velocity ratio is in good agreement
with the experimental measurements.
In figure 2.36, the streamlines of the azimuthally averaged flow are displayed in the
(r, z) plane for several flight regimes. In solid lines are plotted the streamlines of the
induced flow while the total flow is represented with dashed lines. In fast climbing flight
and wind turbine regimes (figures 2.36a and 2.36d), the external flow dominates, so the
total flow remains unidirectional. When the external flow is reduced, the induced flow
becomes more important, until the limit case of the hovering flight regime (figure 2.36b),
where the external flow is zero. In VRS, the flow becomes bidirectional (figure 2.36c):
outside of the vortical tube the flow goes upwards, while it goes downwards inside of it.
This change of direction is associated with a loop of the streamlines around the border
of the rotor disc. It is important to notice that these streamlines are obtained under an
axisymmetric flow assumption. However, the real flow induced by an helical vortex is not
that homogeneous (2.32). For this reason, although the helical vortex correspond with a
streamline of the complete 3D flow, the projection of the helical vortex in the (r, z) plane
does not really correspond with an streamline of the azimuthally averaged flow.
The total mass rate crossing the rotor disc can be easily calculated by integrating the
azimuthally averaged flow:
Z Rb
MT = 2πρ (V∞ + V̄zInd (r))rdr, (2.40)
0
where ρ is the density of the fluid. In figure 2.37 a contour map of the non-dimensional
mass flow rate MT /πρRb3 Ω is shown in the (1/λ,η) plane. When external flow speed is
high or vortex circulation is small, the influence of η on the mass flow rate vanishes and
it mainly depends on the tip speed ratio. A simple analytic expression can be obtained
64
3
Figure 2.37: Contour map of the non-dimensional mass flow rate MT /πρRtip Ω. In a
vertical rotor, negative values represent a flow from up to down.
65
(a) (b)
Figure 2.38: Contour maps of (a) axial interference factor a∗ and (b) power coefficient
CP . The dashed line corresponds to the contour of a∗ = 1/3.
In figure 2.38a, the contour map of the interference factor is shown in the (1/λ, η) plane
(in the wind turbine regime). As expected, the interference factor of the rotor is weaker
for low values of λ and of η, that is, for high external wind speed and low vortex induction.
On the other hand, the axial interference factor grows near the frontier of the wind turbine
regime, where the induction effect of the wake is very important compared to the external
flow. This feature is directly related with the mass flow rate variation across the rotor
(figure 2.37): a high interference factor implies a weak mass flow rate.The limit case a∗ = 1
matches with the contour line of autorotation at MT = 0.
From the point of view of wind turbine applications, it is specially interesting to es-
timate the power extracted from the flow by the rotor. In Sørensen (2016), the author
proposes an expression derived from the momentum theory to calculate the power ex-
tracted by a wind turbine. This expression is obtained from a balance of the energy flux
computed upstream and downstream of the rotor:
R1
u2θ1
Z Z Z
P = 1/2ρV∞3 A0 − 1/2ρu31 dA1 − 1/2u2θ1 − dr1 ρu1 dA1 . (2.43)
A1 A1 r1 r1
The two first terms correspond to the axial momentum theory, the last one correspond
to the losses due to rotation. A1 is the cross-section area of the stream tube enveloping
the rotor disc in the far-wake, A0 is the area of the same streamtube upstreams. In our
case, u1 and uθ1 are the azimuthally averaged axial and azimuthal velocities inside this
streamtube in the far-wake. The crossareas A0 and A1 of the streamtube enveloping the
rotor disc are calculated with the relation uA = Cst. In the literature, it is usual to use
the power coefficient CP , defined as:
P
CP = (2.44)
1/2 ρV∞3 AR
66
Figure 2.39: Region map in 1/λ and η of the different solution topologies and rotor
regimes. I) Tip and hub vortices go downwards. This kind of solutions are found in
climbing flight regime. II) Tip and hub vortices go downwards and upwards respectively.
For negative values of λ, climbing flight regime is found, for positive values, soft descending
and VRS regimes. Hover flight is found at 1/λ = 0 (dashed line). III) Both vortices go
upwards. This kind of solution is found in wind turbine (windmill brake flight) regimes.
In figure 2.38 we show a contour map of the power coefficient of the rotor for different
values of λ and η. For the wind turbine regime, we find a maximum value of CP = 0.53,
just on the contour line of a∗ = 1/3. This is in good agreement with the momentum
theory, that establishes that the maximum theoretical power coefficient extracted by a
wind turbine is CPmax = 0.59 for an axial interference factor of a∗ = 1/3 (Sørensen, 2016).
67
2.6.2 Parametrical analysis
In this section, the effect of the tip-speed ratio λ, the vortex strength η, the radii ratio R∗
and the number of blades n on the geometry of the wake is analyzed for the three types
of solutions.
In figure 2.40, the radial trajectories of the vortices in the wake are shown for the three
type of solutions when λ, η and n are varied. For climbing type I solutions, the wake is
contracted downwards for tip and hub vortices. This contraction grows when the external
flow is reduced or the circulation is increased. Not a big variation is noticed when the
number of blades is increased, just a slightly higher contraction for the tip vortex, while
the hub vortex remains practically unchanged. A particular modification of the radial
trajectory of the internal vortex is observed for high values of λ or η. A small expansion
is observed near the rotor, followed by an important contraction to the far-wake state.
This occurs when the transition limit to region II is approached (a particular analysis will
be addressed later about this point). For type II solutions, similar geometries of the wake
are observed independently of the rotor regime. Climbing (λ = −100), hovering (λ = ∞)
and soft descending (λ = 200) configurations are very similar. The same contraction
of type I solutions is observed for tip vortices. This contraction grows from climbing
flight to hovering and it continues to grow in the soft descending regime (just like for
Rb = 0 case). In hover, the radial trajectory of the tip vortex is almost unchanged when
vortex circulation or the number of blades is modified. The hub vortex trajectory is
however highly deformed for this type of solution. Firstly, it gets contracted while it goes
downwards, then, it goes upwards in the form of a perfect helix. The radius of the far-
wake hub vortex is not very sensitive to the variation of λ, however, it noticeably increases
with η. For type III solutions, typical wind turbine wake geometries are obtained. A
radial expansion is observed for both vortices, especially the external one. This expansion
increases with the tip speed ratio, the vortex circulation or the number of blades.
In figure 2.41, the local pitch distribution of the helical vortices is shown for the same
cases as figure 2.40. The local pitch is defined for each vortex as
where z(φ) is the axial location of the helix at its azimuthal position φ. For climbing type
I solutions, the pitch remains almost constant for both vortices, with a soft increase just
after the emission from the rotor. For a single bladed rotor, the pitch is always bigger for
the tip vortex, however, an opposed behaviour is observed for 2 or 3 blades. When the
external flow is decreased, the wake is decelerated and the pitch is reduced, in the other
hand, when vortex circulation is increased, the wake is accelerated and the pitch grows.
The inverse behaviour of climbing flight is observed for wind turbine type III solutions.
The pitch is softly reduced after the emission from the rotor and it is larger for the hub
vortex in single bladed rotors. When λ is increased, the pitch is reduced, as observed
in climbing flight, however, in wind turbine configuration, when η is increased the wake
is decelerated and the pitch is reduced. For type II solutions, the same tendency as in
climbing flight is observed for the tip vortex when λ, η or n are modified. For the hub
vortex, a completely different evolution is observed. A strong variation is observed when
the vortex changes direction and goes upwards, then, the pitch becomes constant.
68
2 2 2
λ = -10 η = 0.01 n=1
λ = -20 η = 0.02 n=2
λ = -50 η = 0.04 n=3
0 0 0
z/Rtip
z/Rtip
z/Rtip
-2 -2 -2
-4 -4 -4
-6 -6 -6
0 0.5 1 0 0.5 1 0 0.5 1
r/Rtip r/Rtip r/Rtip
2 2 2
λ = -100 η = 0.01 n=1
λ=∞ η = 0.02 n=2
1 λ = 200 1 η = 0.04 1 n=3
0 0 0
z/Rtip
z/Rtip
z/Rtip
-1 -1 -1
-2 -2 -2
-3 -3 -3
0 0.5 1 0 0.5 1 0 0.5 1
r/Rtip r/Rtip r/Rtip
6 6 6
4 4 4
z/Rtip
z/Rtip
z/Rtip
2 2 2
0 0 0
λ=9 η = 0.01 n=1
λ = 12 η = 0.02 n=2
-2 λ = 18 -2 η = 0.04 -2 n=3
Figure 2.40: Radial trajectories of tip (black) and hub (red) vortices along the wake for
different values of λ, η and n. The default parameters are η = 0.01, n = 1 and λ = −20
for region I solutions (upper line), λ = ∞ for region II (middle line) and λ = 12 for region
III (lower line). For all the cases Rb∗ = 0.3 and ε = 0.05. The horizontal thick grey line
represents the rotor disc.
69
1 0.8
λ = -10 η = 0.01 1 n=1
λ = -20 0.7 η = 0.02 n=2
0.8 λ = -50 η = 0.04 n=3
0.6 0.8
h/Rtip
h/Rtip
h/Rtip
0.6
0.5
0.6
0.4 0.4
0.4
0.3
0.2
0.2 0.2
0 2 4 0 2 4 0 2 4
φ/2π φ/2π φ/2π
0.4
0.2
0.2
0.2
0.1 0.1
h/Rtip
h/Rtip
h/Rtip
h/Rtip
Figure 2.41: Local pitch distribution of tip (black) and hub (red) vortices in the wake
along the azimuthal position φ of each vortex for different values of λ, η and n. The
distributions of the figures and the chosen parameters are the same as in figure 2.40.
70
1
0
0.5 20
0
-5 15
z/Rtip
z/Rtip
z/Rtip
-0.5
-10 -1 10
-1.5 5
-15
-2
0
-20 -2.5
0 0.5 1 0 0.5 1 0 0.5 1
r/Rtip r/Rtip r/Rtip
0.42 0.2 0.52
0.4
0.1 0.5
0.38
h/Rtip
h/Rtip
h/Rtip
0 0.48
0.36
-0.1 0.46
0.34
Figure 2.42: Radial trajectories (upper plots) and local pitch distribution (lower plots)
of tip (black) and hub (red) vortices in the wake for two different values of radii ratio:
Rb∗ = 0.3 (solid line) and Rb∗ = 0.7 (dashed line). The tip-speed-ratio is λ = −20 for left
plots (climbing flight, region I), λ = ∞ for middle plots (hover, region II) and λ = 12
for right plots (windmill brake flight, region III). For all the cases η = 0.01, n = 1 and
ε = 0.05.
71
0.5
0
-0.5
z/Rtip
-1
-1.5
-2 λ=50
λ=55
-2.5
0 0.5 1 1.5
r/Rtip
(a) (b)
Figure 2.43: Transition in climbing flight regime from type I solutions (λ = 50) to type
II solutions (λ = 55), with η = 0.01, Rb∗ = 0.3 and ε = 0.05. (a) Radial trajectories of
tip (black) and hub (red) vortices. For the tip vortex, the transition is very smooth, in
both cases the vortex goes downwards and the geometry is conserved. The hub vortex
goes downwards for λ = 50 and upwards for λ = 55; a overlap of both geometries can
however be observed near the rotor disc. This similar behaviour can also be noticed in the
perspective view in figure (b). For a better visualization, just the tip vortex for λ = 50 is
shown in light grey line.
In figure 2.42, the effect of the parameter Rb∗ is illustrated. Radial trajectories and
local pitch distribution for the three different kinds of solutions are shown for Rb∗ = 0.3
and Rb∗ = 0.7. In climbing flight and wind turbine configurations, when Rb∗ is increased,
tip and hub vortices approach and the mutual interaction becomes more important. In
figure 2.42 (left and right plots) it can be perfectly noticed in the far-wake the periodical
deformation studied in section 2.3. For Rb∗ = 0.3, the radial interaction of both vortices
is very weak and the helices remain undeformed. For the type II solutions, the effect of
Rb∗ is clearly visible near the rotor. When the hub vortex goes downwards, it influences
the contraction of the tip vortex structure: this contraction is weaker close to the rotor
but larger in the far-wake. An opposite behaviour is observed for the hub vortex: the
far-wake contraction is reduced. The pitch of the external helix remains unchanged. For
the hub vortex, the pitch distribution is altered near the rotor when Rb∗ is modified, but
it reaches a similar value in the far-wake.
In figure 2.40, it is interesting to note the similar behaviour of tip vortex solutions
in regions I and II. In both cases, the tip vortex moves downwards with similar radial
trajectories and pitch distribution. In figure 2.43, the transition between both regions is
illustrated for a climbing flight regime example. For the tip vortex, there is no difference
on the geometry when the frontier of the existence domain is reached from both sides and
solutions seem to match perfectly. For the hub vortex, solutions are completely different
in the far wake, as it goes downwards for region I and upwards for region II. However,
near the rotor, a long as both solutions go downwards, they overlap.
In figure 2.44, the contour maps of the far-wake parameters h∗ and α are shown in the
72
(a) h∗ (b) α
Figure 2.44: Contour maps of the far-wake parameters h∗ and α in the (1/λ, η) with
Rb∗ = 0.3 and ε = 0.05. In grey, regions where no solution has been obtained.
(1/λ, η) plane. From the contours of h∗ , one can notice the continuity of the tip vortex
solution across the frontier between regions I and II. However, for wind turbine regimes,
the geometry completely changes. Generally, h∗ grows for small values of λ, that is, when
the external flow is strong, and decreases for slow windmill brake descending and ”fast”
soft descending, where VRS are expected. Concerning α, we first observe that its range
of variation is limited in region I and III. As it was observed in figure 2.40, α is always
smaller than 1 for climbing flight regime (region I), that is, the hub vortex has a smaller
pitch than the tip vortex. For the wind turbine regime (region III), α is always bigger
than one. For both cases, α approaches 1 when η is reduced. For type II solutions, the
definition of h∗ and α is not strictly the same as presented in section 2.3, in this case,
each vortex goes in a different direction and they never interact in the far wake. However,
the same logic of the rest of regions is conserved: α < 1 for climbing flight and α > 1 for
descending flight, obtaining α ' 1 at hovering flight independently of the value of η.
73
2.6.3 Comparison with the standard model
Almost no difference is found between tip vortex solutions obtained with standard and
generalized Joukowski models. In figure 2.45a, a 3D view of a wake is shown for a hovering
flight regime obtained with both models. Both solutions are almost identical, with a slight
radii difference at the far wake. In figure 2.45b, a contour map of the non-dimensional
pitch h∗ is plotted in the (1/λ, η) plane. A very good overlap of the models is noticed for
all the regimes. Just a small variation is observed in the VRS region. In this figure, it
can also be noticed that more extreme solutions can be found with the standard model.
It can be expected that the agreement between both models disappears if high values
of Rb∗ parameter are considered, especially for VRS regimes, where vortices are greatly
deformed due to their mutual interaction (figure 2.42). In figure 2.46, radius and pitch
of the external helix at the far-wake are compared for different rotor regimes with several
values of parameter Rb∗ . One can note that the pitch is the same for both standard and
generalized models independently of the regime or the radii ratio Rb∗ . However, a visible
variation of the tip vortex radius in the far-wake is observed when high values of Rb∗ are
chosen, specially in VRS cases. For Rb∗ < 0.5, the difference of the far-wake radius remains
small, even in VRS. Only for Rb∗ = 0.7, a deviation of about 25% is found in VRS. Similar
tendencies are observed for different values of vortex strength η.
(a) (b)
Figure 2.45: (a) Perspective of the wake for λ = ∞ and η = 0.04 (To simplify the
visualization, just the vortices of a single blade are plotted). (b) Contour map of h∗ . In
black solid line: standart Joukowski model (Rb∗ = 0). In red dashed line: the generalized
Joukowski model (Rb∗ = 0.3). For both cases ε = 0.01 and N = 2.
74
2 1
0.9
1.5
h∞ /Rtip
r∞ /Rtip
0.8
1
0.7
0.5
0.6
0 0.5
-0.3 -0.2 -0.1 0 0.1 -0.3 -0.2 -0.1 0 0.1
1/λ 1/λ
2 2.4
2.2
1.5 2
h∞ /Rtip
r∞ /Rtip
1.8
1
1.6
0.5 1.4
1.2
0 1
0.1 0.15 0.2 0.25 0.3 0.35 0.1 0.15 0.2 0.25 0.3 0.35
1/λ 1/λ
Figure 2.46: Pitch and radius of the tip vortex at the far-wake for climbing flight and
VRS (upper plots) and wind turbine regimes (lower plots). Comparison between standart
Joukowski solutions (solid line) and generalized model solutions for Rb∗ = 0.3 (◦), Rb∗ = 0.5
() and Rb∗ = 0.7 (4). η = 0.02 and ε = 0.01.
75
2.7 Chapter conclusion
In this chapter, the geometry of the wake behind a rotor has been studied. A Joukowski
model has been used for the wake, composed of a bound vortex on the blade and two
free vortices emitted from the tip and hub. Solutions are presented for a standard model,
where the hub vortex is straight and placed on the axis, and a generalized model, where
hub vortex is out of axis, generating a helical vortex.
Near- and far-wake structures have been studied separately. Far-wake solutions have
been modeled as infinite structures of coaxial helical vortex pairs. These solutions have
been proved to be spatially periodic and stationary in an appropriate moving reference
frame. An extensive parametric analysis has been made for this kind of solutions. For
the standard case, far-wake solutions are uniform helical vortices with a straight vortex
on the axis.
Near-wake solutions have been obtained for any kind of flight configuration. For the
standard case, the induced flow across the rotor has been computed and compared with
Momentum theory and experimental measurements; a good agreement has been found for
all rotor configurations. For the generalized model, three different topologies of solutions
have been found, with two small regions in the climbing flight and VRS regimes, where
no solutions have been obtained. This region without solutions at the transition to VRS
is also found in the standard configuration. A VRS region without solution is also found
with the standard model.
76
Chapter 3
Contents
3.1 Chapter overview . . . . . . . . . . . . . . . . . . . . . . . . . . 78
3.2 Perturbation model . . . . . . . . . . . . . . . . . . . . . . . . . 79
3.3 Modal decomposition . . . . . . . . . . . . . . . . . . . . . . . . 81
3.4 Spatio-temporal evolution . . . . . . . . . . . . . . . . . . . . . 84
3.5 Chapter conclusion . . . . . . . . . . . . . . . . . . . . . . . . . 86
77
3.1 Chapter overview
In this chapter, the stability of the solutions obtained for the vortical wake is analyzed,
in particular, solutions obtained with the standard Joukowski model.
The stability of a perfect helical vortex was studied by Betchov (1965) and Kida
(1982) taking into account the local induction approximation where the velocity on a
point of the vortex just depends on the local curvature. However, this simplification is
only valid for large pitch helices, where the effect of the successive turns of the helix is
small compared to local curvature effects. A stability analysis of the problem taking into
account the induction effect of the whole structure was made by Widnall (1972). Widnall
calculated the self-induced velocities of a single helix using the Biot-Savart law with the
cut-off method. She then calculated analytically the displacement of the perturbations
and identifying unstable modes at different wavelengths. The modes of instability found
by Betchov (1965) or Kida (1982) were completed with new modes for small pitch, where
the effect of interaction of successive turns of the helix is important. It was also found
how the most unstable perturbation appears when the displacement of the neighboring
loops are out of phase. Some years later, Fukumoto & Miyazaki (1991) included an axial
flow along the vortex core, finding that the instability may be suppressed for some cases
with large pitch. Gupta & Loewy (1974) extended the results of Widnall for multiple
interdigitated helices following the same principle of Biot-Savart integration and cut-off
formula and obtaining the growth rates of the different unstable modes. In the case of
multiple helical vortices, different instability modes appear from the mutual interaction
of the different helices. Similar to the single helix case, the most unstable modes for two
interlaced helices exhibit an out-of-phase symmetry pairing. Okulov (2004) reconsidered
the case of multiple helices using not the Biot-Savart integration but analytic expressions
for the induced velocity introduced by Hardin (1982). Later, Okulov & Sorensen (2007)
added the central vortex present in Joukowski model. Concurrently, similar stability
analysis have been made for an infinite array of vortex rings (Levy & Forsdyke, 1927;
Bolnot et al., 2014), which presents important similarities with helices of small pitch.
A number of authors have worked on these stability problems using numerical simu-
lations and experimental studies of rotor wakes. Ivanell et al. (2010) and Sarmast et al.
(2014) studied numerically the instability of the wake in wind turbines, using the ac-
tuator line method (Sorensen & Shen, 2002) to model the rotor. In these works, the
authors calculated the spatial growth rate of the individual modes through a spectral
analysis of the non-linear flow. A temporal linear stability analysis was recently per-
formed by Brynjell-Rahkola & Henningson (2017) and Selçuk et al. (2018), obtaining the
growth rate of different unstable modes using time-stepping methods on the linearised
Navier-Stokes equations. On the experimental side, Quaranta et al. (2015) did a com-
plete comparison with Widnall (1972) results, reproducing the main unstable modes with
their growth rates. They also explained why the instability mechanism is associated with
local pairing. In a more recent work, Quaranta et al. (2018) presented a similar analysis
for two interlaced helical vortices and captured the unstable pairing modes predicted by
Gupta & Loewy (1974).
In this chapter, a temporal linear stability analysis of our solutions will be performed.
In section 3.2, the methodology to perturb the solutions and analyze the evolution of the
perturbation in time will be presented. In section 3.3, the evolution of the perturbation
78
will be processed to obtain the growth rate as a function of wave number. In section
3.4, spatio-temporal analysis of the perturbation will be performed, studying how they
propagate in the wake. Thanks to this analysis, the rotor regimes for which the solution
is convectively or absolutely unstable will be identified.
where s0 = (r0 (ζ), φ0 (ζ), z0 (ζ)) is the base solution and s0 = (r0 , φ0 , z 0 ) is the perturba-
tion. Introducing the complete perturbed solution s on the problem equation (2.37) and
linearizing, we obtain a first order system of equations for the temporal evolution (on ψ)
of the perturbation:
∂r0 ∂r0
1 ∂Ur 0 ∂Ur 0 ∂Ur 0
=− + r + φ + z , (3.2)
∂ψ ∂ζ ΩR ∂r ∂φ ∂z
∂φ0 ∂φ0
1 ∂Ω 0 ∂Ω 0 ∂Ω 0
=− + r + φ + z , (3.3)
∂ψ ∂ζ ΩR ∂r ∂φ ∂z
∂z 0 ∂z 0
1 ∂Uz 0 ∂Uz 0 ∂Uz 0
=− + r + φ + z . (3.4)
∂ψ ∂ζ ΩR ∂r ∂φ ∂z
The system is solved numerically. For that, a finite difference discretization scheme
has been applied in ψ and ζ:
s0j+1 − s0j
1
= −Dζ + ∇U (s0 ) s0j (3.5)
∆ψ ΩR
where the index j represents the discretized position in ψ, Dζ is the discretization matrix
on ζ and ∇U (s0 ) is the Jacobian matrix of the complete induced velocity field on the
base solution. To avoid numerical stability problems, a robust discretization scheme has
to be chosen for the matrix Dζ . In our case, a five point symmetric discretization scheme
has been used.
The perturbation is introduced as an impulsion displacement in the axial direction at
the azimuthal position ζp :
0
s0j=0 (ζ) = 0 , (3.6)
Ap δ(ζ − ζp )
where Ap is the amplitude of the initial perturbation and δ is the Dirac delta function.
The perturbation is applied for each vortex in the same axial position, so symmetry is
forced.
79
×10-3 ×10-3 ×10-3
2 2 2
1 1 1
0 0 0
z′
z′
z′
-1 -1 -1
-2 -2 -2
0 2 4 6 0 2 4 6 0 2 4 6
ζ/2π ζ/2π ζ/2π
Figure 3.1: Axial coordinate of the perturbation at three different instants. Parameters
of the base solutions are λ = −10, η = 0.04, ε = 0.05. Perturbation amplitude and initial
position Ap = 10−4 Rtip and ζp = 4π.
0.8
0.6
φ/2π
0.4
0.2
0
10.5 11 11.5 12 12.5 13 13.5
z/R
Figure 3.2: View of the perturbed helical vortex in the axial-azimuthal plane. Solid line:
perturbed solution. Dashed line: base solution. The same parameters of figure 3.1 are
used for the base solution. The system is dominated by the pairing instability.
In figure 3.1, the evolution of an axial perturbation on a climbing flight wake is shown.
In this case, the perturbation is introduced at ζp = 4π, however, for climbing flight, similar
propagations are normally found independently of the initial position of impulse forcing.
As soon as it is created, the perturbation propagates downstream and starts to interact
with the surrounding turns of the helices. This interaction produces new peaks in the
perturbed signal. Unstable solutions are found for all rotor regimes. The perturbation is
always amplified in time and the system is dominated by a pairing instability mode. In
figure 3.2, a view of the perturbed solution in the axial-azimuthal axis is shown where the
pairing pattern can be perfectly seen.
80
(a) (b)
Figure 3.3: Schematic of the deviation d of the perturbed solution. (a) Unperturbed and
(b) perturbed system.
-11 2
σ2h2∞ /Γ
-13 k=3
1
k = 0.5 σ/Ω
-14
k=2 0.5
-15
-16 0
0 1 2 3 0 2 4
ψ/2π k
(a) (b)
Figure 3.4: (a) Temporal evolution of the perturbation decomposed for different wavenum-
bers. In dashed line, the position in time where the slope is measured to obtain the growth
rate. (b) Growth rate as a function of the wave number. Solid line: distribution for a
wake solution with λ = −10, η = 0.04, ε = 0.05 and ζp = 4π. Dashed line: theoretical
prediction for h/R = 1.03 and a/R = 0.055, obtained from the geometrical parameters of
the solution in the far-wake.
The particularity of introducing a Dirac perturbation is that all the wavenumbers are
excited and they grow together in a wave packet. A spectral analysis is made to separate
and find the growth rate of each wavenumber. Instead of analyzing the perturbation in a
particular frame, we first consider the total deviation d of the vortex from its unperturbed
position (figure 3.3). To have a equispaced discretization of the vortices in the azimuthal
direction, original solution in ζ is interpolated and parametrized in terms of the azimuthal
coordinate φ.
At each time step, spatial Fourier transform is applied to the deviation function d(φ)
in order to obtain the amplitude of different modes. For each mode, associated with a
wavenumber k, the temporal evolution of the amplitude Ak (ψ) is obtained (figure 3.4a).
After a transient, the amplitude evolution can be fitted by an exponential expression
Ak (ψ) ∼ exp(σψ), where σ is the growth rate of the instability mode of wavenumber k.
81
Figure 3.5: Growth rate distribution for different rotor regimes with ε = 0.01 and ζp = 20π
for all cases. Solid line: numerical analysis of the perturbed solutions. Dashed line:
theoretical prediction from Gupta & Loewy (1974).
82
Then, it can be calculated as:
d
σ(k) = log [Ak (ψ)] (3.7)
dψ
In figure 3.4a, the temporal evolution of the perturbation associated with different wave
numbers is plotted. To obtain the growth rate, the slope of the perturbation is measured
after 2 turns of rotor (ψ = 4π), when the exponential growth is established.
In figure 3.4b we show a typical distribution of the growth rate as a function of the
wavenumber. We plot the non-dimensional growth rate σ ∗ , defined as:
σ ∗ = Re(σ)2h2∞ /Γ (3.8)
It can be observed the more unstable mode at k = 1 with a value of σ ∗ ' π/2, which
corresponds to the local pairing mode shown in figure 1.15b. Successive highly unstable
modes are found for odd wavenumbers. A comparison is made with the analytical results
obtained by Gupta & Loewy (1974) for infinite multiple helical vortices. The theoretical
curve for σ(k) is calculated by taking h/R = hF W /RF W and a/R = a/RF W as parameters
for the infinite helix.
In their work, Gupta & Loewy (1974) used the same procedure as that of Widnall
(1972). On a perfect helix that rotates and axially translates without changing shape due
to its self-induction, they consider perturbations of the form:
δrh = δ r̂h exp(σt + ikθh ), (3.9)
where δrh is the perturbation in a cylindrical frame, δ r̂h is the amplitude of the perturba-
tion in the three directions and θh is the azimuthal coordinate of the helix. The evolution
of the filament shape is obtained by computing its self-induced velocity from the Biot-
Savart law. To desingularize the integral, contrary to Widnall (1972), who used a cut-off
method, Gupta & Loewy (1974) used a Rosenhead formula (Saffman, 1992), that replaces
3
the denominator of Biot-Savart integral by |(r − r0 )2 + µ2 | , where µ is a parameter pro-
portional to the core size. This procedure is equivalent to cut-off for µ = a exp(−3/4)
(Saffman, 1992). However, Gupta & Loewy (1974) erroneously used µ = a, so a correc-
tion has to be made to compare with their results. Independently of the desingularization
procedure, the main difference with Widnall (1972) is that they extended the analysis to
multiple helical vortices.
In the theory, different growth rate distributions are obtained when pitch or core size
are modified, consequently, the growth rate will be also modified when the rotor regime
is changed and the wake structure varies. In figure 3.5, examples of the growth rate
distribution are presented for different rotor regimes. For all cases, the same distribution
can be observed, with most unstable modes located at k = 1 and k = 3. A good agreement
is found with the theoretical prediction independently of the regime.
It is important to note that, for this modal analysis, the perturbation has to be
introduced where the wake structure is already uniform. Near the rotor, the spatial
distribution of the vortex is not uniform and the dynamic of the perturbation is not
clear, so exponential growth is not necessarily expected. Here, we have therefore mainly
computed the stability properties of the far-wake, which is well described by perfect
helices. This explains the very good agreement with Gupta & Loewy (1974) that we have
observed even for the non-conventional VRS regime.
83
(a) (b) (c)
(d) (e)
Figure 3.6: Spatio-temporal evolution of the perturbation for (a) a convectively unstable
case (λ = −50), (c) an absolutely unstable case (λ = 20) and (b) a case on the convec-
tive/absolute limit (λ = −600). For the three cases, the perturbation is introduced at 6
turns of the helix from the rotor. (d) Typical temporal evolution of the amplitude of the
perturbation at a frame velocity v ∗ . (e) Growth rate versus frame velocity for the three
cases of upper figures: λ = −50 (solid), λ = −600 (dashed), λ = 20 (dash-dotted).
84
axial direction z, which is related to coordinate ζ by:
dz 1
= (Uzind + V∞ ). (3.10)
dζ ΩR
In figure 3.6a-c, temporal evolution of the wavepacket is plotted. The axial direction and
time are normalized as:
ΩR /2π ΩR
z∗ = z, t∗ = t. (3.11)
N Γ/2h∞ + V∞ 2π
Note that normalized time is equivalent to rotor azimuth ψ. If we move in the wake
with the wavepacket, we will see how the amplitude of the perturbed solution grows in
time. Then, following the same procedure as for the modal analysis, the growth rate
can be calculated by measuring the slope of the amplitude evolution (figure 3.4a). This
procedure can be used for any frame velocity v = z/t, obtaining a particular amplitude
evolution A(t, v) and its related growth rate σ(v) (figure 3.6d). Then, downstream v+ and
upstream v− propagation velocities are calculated as the values of v where the growth rate
is zero, that is, where the perturbation is no longer amplified. Transition from convective
to absolutely unstable is found when v− = 0 (figure 3.6b).
In figure 3.6e, σ ∗ (v ∗ ) distribution is plotted for the three regimes of the figure. v ∗ is
the frame velocity normalized as
v
v∗ = . (3.12)
N Γ/2h∞ + V∞
Under this normalization, the maximum growth rate values are found at v ∗ = 1.103,
v ∗ = 1.111 and v ∗ = 1.145 for λ = −50, λ = −600 and λ = 20 respectively, which
correspond to the velocity of displacement of the vortices. This velocity does not coincide
with v ∗ = 1 because the axial induced velocity of the helices is not exactly Γ/2h∞ , as
in the two-dimensional case of an array of vortices. For the three cases, the maximum
growth rate is closed to σ ∗ ' π/2, and it corresponds to the contribution of mode k = 1,
as it was seen in the previous section. A narrower peak of σ ∗ (v ∗ ) corresponds to a slower
expansion of the wavepacket. If the expansion is slower than the advection, the system
remains convectively unstable, otherwise the system becomes absolutely unstable. In
other words, if there are negative values of v ∗ for positive values of σ ∗ , the system is
absolutely unstable.
Concerning rotor regimes, σ ∗ (v ∗ ) has a narrower peak for high external velocity regimes,
and it gets wider when approaching hover and soft descending flights. In figure 3.6, the
velocity of expansion of the perturbation is minimal for the climbing flight and it grows
when approaching VRS. In figure 3.7, the regions of stability for in the (1/λ,η) plane are
shown. Climbing and windmill brake flight regimes are generally convectively unstable.
The transition to absolutely unstable regimes occurs near hovering flight, which is also
the region where the transition to VRS is observed. Another transition to absolutely un-
stable solutions is observed in wind turbine regimes, just before the region where solutions
cannot be obtained. This transition occurs around the limit when the wake overtakes the
rotor plane and the mass flow across the rotor disc falls to zero. However, for this kind of
solution, the wake structure requires a great distance from the rotor to become uniform,
so it is not excluded that the impulse has been positioned too close to the rotor in this
case.
85
Figure 3.7: Convective (light grey) and absolute (dark grey) instability regions (ε = 0.01).
86
Chapter 4
Rigid blade
Contents
4.1 Chapter overview . . . . . . . . . . . . . . . . . . . . . . . . . . 88
4.2 Rigid blade framework . . . . . . . . . . . . . . . . . . . . . . . 88
4.2.1 Induced velocity on the rotor . . . . . . . . . . . . . . . . . . . 88
4.2.2 Circulation law and equivalent Joukowski profile . . . . . . . . 89
4.2.3 Aerodynamic loads . . . . . . . . . . . . . . . . . . . . . . . . . 91
4.2.4 Tip effect correction . . . . . . . . . . . . . . . . . . . . . . . . 91
4.2.5 Model parameters . . . . . . . . . . . . . . . . . . . . . . . . . 92
4.3 Comparison with numerical and experimental results . . . . 92
4.3.1 MEXICO rotor project . . . . . . . . . . . . . . . . . . . . . . 92
4.3.2 The Unsteady Aerodynamic Experiment (UAE) . . . . . . . . 96
4.3.3 A hovering case: NASA Technical report TM81232 . . . . . . . 98
4.4 Chapter conclusion . . . . . . . . . . . . . . . . . . . . . . . . . 100
87
Figure 4.1: Schematic of the airfoil and the different contributions to the local velocity U
for a climbing helicopter flight configuration
so the modulus of the local velocity and the angle of attack (figure 4.1) at each radial
position of the blade are given by :
q
U (r) = (V∞ + V̄zind (r))2 + r2 (−Ω + Ω̄ind (r))2 , (4.3)
88
1.2 1.2 1.2
1 1 1
U/RΩ
U/RΩ
0.6 0.6 0.6
φ[◦ ]
φ[◦ ]
30
15 10
20 10 5
5 0
10 0 -5
0.2 0.4 0.6 0.8 1 0.2 0.4 0.6 0.8 1 0.2 0.4 0.6 0.8 1
r/R r/R r/R
Figure 4.2: Comparison of the local velocity (upper) and the angle of the flow with
the rotor plane (lower), considering just the external flow and the rotation of the blades
(dashed line) and considering also the velocity induced by the wake (solid line). The wake
is calculated with a circulation parameter of η = 0.05, a normalized core size of ε = 0.01
and a tip-speed-ratio of λ = −5 (left), λ = −15 (center) and λ = ∞ (right).
V∞ + V̄zind (r)
α(r) = β(r) − φ(r) = β(r) − arctan , (4.4)
r(−Ω + Ω̄ind (r))
where β is the pitch angle of the blade and φ is the angle of the flow with the rotor plane.
In figure 4.2, the effect of the wake induction on the rotor is illustrated. The local
velocity and its angle with the rotor plane is shown for different tip-speed-ratio and
compared with the external velocity observed by the rotor without considering the induced
velocity. For the local velocity U , there is not a significant difference for any tip-speed-
ratio, just a small difference that appears near the axis, where the contribution from the
rotation of the blades is reduced. However, the angle φ is crucially affected when the
tip-speed-ratio grows. For λ = −5, a small gap is observed in the effective angle when
the induction is considered or not. For λ = −15, the difference grows, especially near the
axis, where the azimuthal velocity is reduced with the radial position. Finally, in hovering
flight, the external axial flow disappears and the angle φ falls to 0 if the induced velocity
is not considered.
1
Γ̄(r) = c(r)U (r)CL (α; r), (4.5)
2
89
Figure 4.3: Scheme of the circulation profile Γ̄ along a blade and the equivalent Joukowski
profile of constant circulation Γ between the vortex emission points Ri and Re .
where c is the chord of the blade and CL is the lift coefficient as a function of the angle
of attack for a given airfoil geometry.
As it was presented in section 1.2.3, the variations of the circulation profile along the
blade are responsible of the emission of a vorticity sheet on the wake. In the Joukowski
model analyzed in chapter 2, all the vorticity is concentrated in two vortices, at the tip
and the hub. To obtain this wake configuration, the circulation profile along the blade
has to be constant, so all the variation of circulation is produced at the ends of the blade
and concentrated just in two vortices. In real cases, there are always variations on the
circulation profile. To obtain wake solutions such as those studied in the previous chapter,
the variable circulation profile is transformed into an equivalent Joukowski profile.
From a circulation profile Γ̄(r), calculated from Kutta-Joukowski law (figure 4.3), the
circulation Γ of the equivalent Joukowski profile is obtained as the maximum value on the
calculated profile:
Γ = max(Γ̄(r)). (4.6)
For the generalized Joukowski model, the radial position of the points of the blade where
the vortices are emitted are calculated as the centroid of the gradient of circulation at
both sides of the maximum value of Γ̄(r):
Z r(Γ) Z R
dΓ̄ dΓ̄
r dr r dr
dr r(Γ) dr
Ri = Z0 r(Γ) ; Re = Z R , (4.7)
dΓ̄ dΓ̄
dr dr
0 dr r(Γ) dr
90
positive and negative circulation (1.6b). In the simplest case where the circulation profile
has a single maximum, each blade is expected to create two concentrated vortices of op-
posite circulation at the positions of the center of mass of ∂r Γ distribution at both sides
of the maximum. In that case, after the roll-up phase, the flow is then composed of N
pairs of vortices of opposite circulation for a N blade rotor. In our model, we assume that
this rolling-up is produced instantly after the rotor. In the standard Joukowski model,
it is assumed that all hub vortices merge together in a single straight vortex on the axis
of circulation −N Γ. In the case of several local maximum values in the circulation along
the blade, the vorticity sheet will be expected to roll-up in several vortices of alternative
circulation sign.
In figure 4.4, the effect of Prandtl’s tip loss factor on the circulation and the aerody-
namic loads is illustrated. In the three cases a similar effect is observed. Before applying
the correction, circulation and aerodynamic forces grow along the blade (which is a typi-
cal distribution on the straight rotor blades) until the tip, where they fall suddenly to 0.
When the correction factor is applied, the pressure drop near the tip is modeled and the
fall of the circulation and the forces is smoother.
91
×10-3
0.03 0.03 3
0.02 0.02 2
ft /ρR3 Ω2
fn /ρR3 Ω2
Γ/ΩR2
0.01 0.01 1
0 0 0
-0.01 -0.01 -1
0.2 0.4 0.6 0.8 1 0.2 0.4 0.6 0.8 1 0.2 0.4 0.6 0.8 1
r/R r/R r/R
(a) (b) (c)
Figure 4.4: (a) Circulation, (b) normal force and (c) tangential force profiles along the
blade with (solid line) and without (dashed line) tip-loss correction. Wind turbine con-
figuration at λ = 7 for the UAE blade (section 4.3.2).
92
(b)
Figure 4.5: (a) MEXICO wind turbine installation (from Schepers & Snel (2007)). (b)
Chord distribution of the blade (c) Pitch angle (d) Lift coefficient of the three airfoil
geometries.
and the rotor loads. The experiment consists of a three bladed wind turbine model, 4.5
m in diameter, installed in a wind tunnel of squared section (9.5 × 9.5 m2 ) (figure 4.5).
The airfoil profile is not constant along the span but it changes between three different
geometries: DU91W2-25, RISO121 and NACA64418 (in figure 4.5 are shown the lift
coefficients for the three airfoil geometries). Chord and pitch distributions are shown in
figure 4.5b and figure 4.5c. All the measurements were made for three different tip-speed-
ratio values, modifying the external flow at V∞ = 10 m s−1 (λ = 10), V∞ = 15 m s−1
(λ = 6.67) and V∞ = 24 m s−1 (λ = 4.17) for a fixed rotation of 424.5rpm.
In figure 4.6 the distributions of the local velocity and of the angle of attack are shown
for the three tip-speed-ratio configurations analyzed in MEXICO. The local velocity is
very similar for the three cases, especially at the tip; however, the flow is slightly stronger
near the axis, when the wind speed is higher (λ = 4.17). Concerning the angle of attack,
the differences are more important. For λ = 10 the angle of attack does not vary very
much: the variation of the incident velocity direction is balanced by the twist of the
blade. When the wind speed increases, the blade twist is not sufficient, so the variable
distribution due to the rotation of the blades (see figure 4.2) becomes more evident. A
maximum variation of 9◦ and 18◦ are observed for the angle of attack for λ = 6.67 and
λ = 4.17 respectively. A comparison with numerical results extracted from Schepers &
Snel (2007) is presented. A perfect agreement is observed for the local velocity (it mainly
depends on the external velocity, as was analyzed in section 4.2.1). For the angle of attack,
a very good agreement is found with the published numerical results, both qualitatively
and quantitatively.
In figure 4.7, the circulation profile along the blades is shown for the three rotor
configurations. A quite constant distribution is observed, especially for the higher tip-
93
120 120 120
This work This work This work
100 DTU1 100 DTU1 100 DTU1
DTU1 DTU1 DTU1
CEN CEN CEN
U [m/s]
U [m/s]
U [m/s]
80 80 80
60 60 60
40 40 40
20 20 20
0.5 1 1.5 2 2.5 0.5 1 1.5 2 2.5 0.5 1 1.5 2 2.5
r[m] r[m] r[m]
9 18 35
8 This work 16 This work This work
DTU1 DTU1 30 DTU1
7 DTU1 14 DTU1 DTU1
CEN CEN CEN
6 12 25
α[◦ ]
α[◦ ]
α[◦ ]
5 10 20
4 8
15
3 6
2 4 10
0.5 1 1.5 2 2.5 0.5 1 1.5 2 2.5 0.5 1 1.5 2 2.5
r[m] r[m] r[m]
Figure 4.6: Distribution of the local velocity (upper) and the angle of attack (lower)
for the three different axial velocities: λ = 10 (left), λ = 6.67 (center) and λ = 4.17
(right). The results obtained with our model (solid line) are compared to three different
numerical results extracted from Schepers & Snel (2007): two different contributions from
DTU (Delft University of Technology) and one from CENER (National Renewable Energy
Center of Spain).
Figure 4.7: Distribution of the circulation along the blades for the three different axial
velocities: λ = 10 (left), λ = 6.67 (center) and λ = 4.17 (right). In red line, the equivalent
Joukowski constant circulation profile.
94
20 45 120
40 100
15 80
35
ft [N/m]
ft [N/m]
ft [N/m]
60
10 30
40
25
5 20
20 0
0 15 -20
0.5 1 1.5 2 2.5 0.5 1 1.5 2 2.5 0.5 1 1.5 2 2.5
r[m] r[m] r[m]
Figure 4.8: Distribution of the normal (upper) and tangential (lower) forces for λ = 10
(left), λ = 6.67 (center) and λ = 4.17 (right). The markers correspond to the experimental
measurements from Schepers & Snel (2007).
speed ratio cases. For the low tip-speed ratio case, the circulation is softened near the
tip, so the vortex emission point is displaced inside the blade. These distributions are in
good agreement with the hypothesis of Joukowski’s constant circulation profile. It can be
observed a linear increase of the maximum circulation value with the external velocity
In figure 4.8, distribution of the normal and tangential aerodynamic forces is shown.
The similar growing tendency is observed for both forces for every tip-speed-ratio. The
strength of the forces grows with the local velocity, that also increases with the external
wind. The tangential force is approximately one order of magnitude weaker than the
normal force. Compared to the experimental measurements, a very good agreement is
found, especially for the normal force at V∞ = 10 m s−1 and V∞ = 15 m s−1 . For the
tangential force, a quantitative deviation is noticed, although the agreement is still qual-
itatively good. A discrepancy between theory and experimental data is observed for the
tangential force for V∞ = 24 m s−1 , however, the same observation was also made with all
the numerical results presented in Schepers & Snel (2007) for the same case.
In figure 4.9, the geometry of the wake is shown for λ = 6.67 (V∞ = 15 m s−1 )
and λ = 10 (V∞ = 10 m s−1 ). As it was analyzed in chapter 2, the wake gets more
expanded for high tip-speed-ratios. In this case, the solutions are compared to CFD
results from Carrion et al. (2015). The agreement is specially good for V∞ = 15 m s−1 .
For V∞ = 10 m s−1 , some differences are found for the pitch of the hub vortices. This
deviation can be explained by the dissipation of the vortices in the CFD results. In
Carrion et al. (2015), the external vortices start to dissipate at a distance of 1.2R from
the rotor, this probably affects the induced flow on the hub helices: they are not slowed
down by the external vortices so they get axially expanded.
95
1.2 1.2
1 1
0.8 0.8
r/Rtip
r/Rtip
0.6 0.6
Carrion et al Carrion et al
0.4 0.4 This work
This work
0.2 0.2
0 0
0 0.5 1 1.5 2 2.5 0 0.2 0.4 0.6 0.8 1
z/Rtip z/Rtip
(a) (b)
Figure 4.9: Geometry of the wake for (a) λ = 6.67 and (b) λ = 10. The markers
correspond to the position of the tip (black) and hub (red) vortices of the wake crossing
the r − z plane. Comparison with numerical CFD results from Carrion et al. (2015).
96
(b)
Figure 4.10: (a) UAE Phase VI wind turbine installation (from Sicklinger et al. (2015)).
(b) Chord distribution of the blade (c) Pitch angle (d) Lift coefficient of the airfoil ge-
ometriy.
Figure 4.11: Distribution of the circulation along the blades for the three different axial
velocities: λ = 7.58 (left), λ = 5.41 (center) and λ = 3.79 (right). In red line, the
equivalent Joukowski constant circulation profile.
97
140 250 300
120 250
200
100
200
80 150
fn [N ]
fn [N ]
fn [N ]
150
60 100
100
40
50 50
20
0 0 0
0.2 0.4 0.6 0.8 1 0.2 0.4 0.6 0.8 1 0.2 0.4 0.6 0.8 1
r/R r/R r/R
8 25 50
6 20 40
4 15 30
ft [N ]
ft [N ]
ft [N ]
2 10 20
0 5 10
-2 0 0
0.2 0.4 0.6 0.8 1 0.2 0.4 0.6 0.8 1 0.2 0.4 0.6 0.8 1
r/R r/R r/R
Figure 4.12: Distribution of the normal (upper) and tangential (lower) forces for λ =
7.58 (left), λ = 5.41 (center) and λ = 3.79 (right). The markers correspond with the
experimental measurements from Hand et al. (2001).
98
(a) (b)
Figure 4.13: (a) Scheme of the rotor installation for NASA technical report TM81232.
(b) Lift-coefficient profile of the NACA0012 airfoil.
0.06 0.6
0.4
Γ̄/ΩR2
0.02
0
0.2
-0.02
-0.04 0
0.2 0.4 0.6 0.8 1 0.5 0.6 0.7 0.8 0.9 1
r/R r/R
(a) (b)
(c)
Figure 4.14: (a) Circulation and (b) lift force along the blade. (c) Radial trajectory of
the wake. Comparison between our numerical results (-) and the experimental data from
NASA technical rapport TM81232 (◦).
99
inaccuracy can be expected.
The discrepancy observed in the wake geometry (figure 4.14c) may not be associated
with the model. Indeed, the experimental rotor is fixed on a big shaft aligned with the
wake axis. We suspect that the shaft limits the contraction of the wake. It is not excluded
that it could also influence the lift forces. For these reasons, it is difficult to have a definite
opinion on the errors associated with the model for this case.
100
Chapter 5
Flexible blade
Contents
5.1 Chapter overview . . . . . . . . . . . . . . . . . . . . . . . . . . 102
5.2 Deformation model . . . . . . . . . . . . . . . . . . . . . . . . . 102
5.3 Fluid-structure interaction . . . . . . . . . . . . . . . . . . . . . 105
5.4 Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
5.4.1 Description of the rotors . . . . . . . . . . . . . . . . . . . . . . 105
5.4.2 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
5.5 Chapter conclusion . . . . . . . . . . . . . . . . . . . . . . . . . 109
101
5.1 Chapter overview
In this chapter, we implement the last part of the model: the deformation of the rotor
blades. The model for the rigid rotor presented in the previous chapter will be still
used to compute aerodynamics loads. However, these loads will deform the blades. We
introduce a elastic beam model to obtain the deformation of the blades under different
conditions. Most of the content of the chapter is extracted from the paper A strongly-
coupled model for flexible rotors (Durán Venegas et al., 2019), published in the Journal of
Fluids and Structures. In this paper, all the coupled model has been developed for a rod
model, including a detailed framework and a parametric analysis for two different blade
geometries.
102
Moreover, in the following, we assume that the airfoil profile is the same all along
the blade, meaning that we can write I = I ∗ c4 and J = J ∗ c4 , where I ∗ and J ∗ are
dimensionless quantities. The cross section area of the airfoil can also be written as
A = A∗ c2 where A∗ is a dimensionless quantity.
The blade deformation problem is then defined by the parameters linked to the blade
profile (CL , CD , I ∗ , J ∗ and A∗ ), one parameter characterizing the blade aspect ratio
c∗ = c/Rb and three parameters associated with the blade material
E ρb
ν, E ∗ = , ρ∗b = , (5.7)
ρb gRb ρ
where Rb is the length of the blade, ρb and ρ the density of the blade and of the fluid
respectively, and g the gravitational acceleration. The parameter E ∗ compares the elastic
forces with gravity. It is also useful to introduce the Froude number Fr that compares
gravitational forces with centrifugal forces:
Rb Ω2
Fr2 = . (5.8)
g
∂ 2 (M · d1 ) ∂T
+ · d2 = 0, (5.10)
∂s2 ∂s
where κ2 = 0 has been used again to obtain the second term and neglect the term coming
from the external moments.
From the constitutive law (5.6) and using the definition (5.5), the first term of equation
(5.10) can be written as
∂ 2 (M · d1 )
3
∂I ∂ 2 θ ∂ 2 I ∂θ
∂ θ
=E I 3 +2 + . (5.11)
∂s2 ∂s ∂s ∂s2 ∂s2 ∂s
The second term of (5.10) is the projection on the direction d2 of the forces exerted on
the blade (see equation (5.1)). They include the aerodynamic force perpendicular to the
blade fn , calculated using equation (4.8), the centrifugal force and the weight of the blade
per unit length. In the following, the rotor axis will be assumed to be aligned with the
direction of gravity, as for a helicopter in vertical flight. The projection on d2 of the
centrifugal force and weight can then be written:
Z s
fcent · d2 = Aρb rb (s)er · d2 = Aρb Ω rb (s) sin θ = Aρb Ω Rb cos θds sin θ, (5.12)
0
103
(a) (b)
Figure 5.1: Blade schemes showing the local orthonormal frame (a) and bending and
torsion angles (b).
Developing equation (5.10) with (4.8), (5.11)-(5.13) we obtain the final equation for
θ. The equation for γ is obtained by projecting equation (5.2) on d3 with (5.5) and (5.6).
Finally, the two dimensionless differential equations for the bending and torsion angles θ
and γ are:
ρ∗b E ∗ I ∗ ∗ 4 000 ∗ 0 ∗ 3 00 ∗0 ∗2 ∗ 00 ∗ 3 0
1 ∗ ∗2
c θ + 8c c θ + 4(3c c + c c )θ − c U (CL (α) cos α + CD (α) sin α)
Fr2 Z s 2 ∗ ∗ ∗2 (5.14)
∗ ∗ ∗2 ρAc
+ρb A c cos θds sin θ + b 2 cos θ sin(β + γ) = 0,
0 Fr
∗ ∗ ∗
ρb E J0 ∗ 4 00 ∗3 ∗0 0
1 ∗2 ∗2 ac
2 c γ + 4c c γ − c U [(CL (α) cos α + CD (α) sin α)δcm + Cm,ac (α)] = (5.15)
0,
2(1 + ν)Fr 2
where the prime denotes differentiation with respect to s. The last two terms of the
bending equation (5.14) correspond to the centrifugal force and the weight of the blade.
The second term of the torsion equation (5.15) corresponds to the aerodynamic moment.
ac
The quantity δcm is the distance (non-dimensionalized by c) from the center of mass to
the aerodynamic center of the airfoil, typically located at a distance of c/4 from the
leading edge for subsonic flows. The coefficient Cm,ac is the pitching moment coefficient
of the airfoil, which is in general constant for small angles of attack and equal to zero for
symmetric airfoils. The angle of attack α(s) and the normalized incident velocity U ∗ (s) =
U (s)/(ΩRb ) are obtained from the wake model discussed in the next two subsections.
These equations have to be integrated with the following boundary conditions
that correspond to a “clamped” condition at the blade root, and a “free” condition at the
blade tip.
In order to analyze the blade deformations in a more intuitive way, we also introduce
the bending deflection function fθ (s) that measures the distance of each blade element
from its original undeformed position.
104
Figure 5.2: Fluid-structure loop scheme.
5.4 Applications
5.4.1 Description of the rotors
In this section, we apply the model to two different two-bladed rotors, named here rotor
A and rotor B. For both rotors, we consider a same NACA0012 profile for the blade
cross-section. The characteristic constants for this profile are I ∗ = 0.0033, J ∗ = 3.28 and
A∗ = 0.0822. For the aerodynamic coefficients CL and CD , we assume the functions given
in Chritzos et al. (1955) and shown in figure 5.3(b) and, for the aerodynamic moment,
ac
we use Cm,ac = 0 and δcm = 0.15. The blades of rotor A have a simple geometry with a
uniform dimensionless chord c∗ = 0.1 and a uniform twist angle β. We shall vary this angle
β. Rotor B is inspired by the rotor used in the experiments by Quaranta et al. (2015),
which was designed to have a constant circulation profile. The original geometry has been
slightly modified to operate in a larger range of tip-speed ratio λ. The distributions of
105
2
1.5
0.5
CL , CD
0
-0.5
-1
-1.5
0 50 100 150
α[◦ ]
Figure 5.3: (a) Blade geometry of rotor A (upper plot) and rotor B (lower plot). (b) Lift
coefficient CL (dashed line) and drag coefficient CD (solid line) with respect to the angle
of attack α for a NACA0012 profile, from Chritzos et al. (1955). (c) Chord c∗ and twist
angle β (in degree) along the rotor B blade. Adapted from Quaranta et al. (2015).
chord and twist angles are shown in figure 5.3(c). The geometry of the blades of each
rotor is illustrated in figure 5.3(a).
In the following section, we shall vary the operational conditions of the rotors (i.e. the
angular rotation Ω and the external wind speed V∞ ) that affect the tip speed ratio λ and
the Froude number F r. The material properties of the blade will also be varied such that
the effect of the dimensionless parameters ρ∗b and E ∗ will be considered. However, the
Poisson ratio will be kept fixed and equal to ν = 0.5.
Except for the twist angle of the rotor A blade, the geometry of the two rotors without
external forces will not be varied. We shall also not analyse the effect of the vortex core
size that is fixed to ε = 0.01.
5.4.2 Results
In this section, we first analyse the effect of the tip-speed ratio λ on the wake character-
istics and blade shape of rotor A. We fix the Young modulus E ∗ , the density ratio ρ∗b , the
Froude number Fr, and vary λ by changing the external fluid velocity.
In figure 5.4, we have illustrated the radial trajectories of the vortices in the wake
when λ is varied. In the rotor frame, the external wind is going downwards when λ is
negative and upwards when it is positive. For a helicopter, this means that negative
values of λ correspond to climbing flight, and positive values of λ to descending flight.
Hover is associated with an infinite value of λ. Normal flight situations of a helicopter
are shown in figure 5.4(a). For this case, the wake is contracting and moves downwards.
The contraction increases when λ becomes more negative, that is when the climbing
speed decreases. This contraction process continues when the climbing speed vanishes
and changes sign, i.e. when we move to a slow descent flight regime corresponding to
large positive values of λ. However, for this regime, the vortices are emitted upwards
and therefore cross the rotor plane before going downwards. For smaller values of λ, a
downward wake ceases to exist, and we jump to another type of solution shown in figure
106
1 8
7
0
6
-1
5
-2 4
z/Rb
z/Rb
-3 3
2
-4
λ = -10 1
λ = -20 λ=3
-5 λ = -∞ λ = 4.2
0
λ = 15 λ = 5.7
-6 -1
0.6 0.7 0.8 0.9 1 0.9 1 1.1 1.2 1.3
r/Rb r/Rb
(a) (b)
Figure 5.4: Radial position of the wake in the axial direction for different tip-speed-ratio
values of rotor A. (a) Helicopter regimes for β = 40◦ . Climbing flight (λ = −10, λ = −20),
hover (λ = ∞), and weakly descending flight (λ = 15). (b) Wind turbine (or windmill
brake) regimes for β = 15◦ . Other parameters are E ∗ = 107 , ρ∗b = 100, Fr2 = 1000.
5.4(b): the wake expands and goes upwards. This situation corresponds to the so-called
windmill brake regime of helicopters. If gravity was not aligned with the rotor axis, it
would correspond to the wind turbine regime. As expected, the stronger the external
wind (i.e. the smaller λ), the less expanding is the wake.
The intervals of λ where downward and upward wakes exist do not seem to overlap.
For all the parameters that we have considered, we have found an interval of λ where
both solutions cease to exist. The limits of this interval are different for each case. We
suspect that this is reminiscent of the Vortex Ring State (VRS) (Drees & Hendal, 1951).
In figure 5.5, we have plotted the circulation profile obtained on the blade for the
configurations shown in figure 5.4. For normal flight conditions (figure 5.5(a)), the circu-
lation profile is found not to vary much. It is mainly associated with the blade rotation
that prescribes a linear dependence of the circulation on the radial coordinate. The small
bump observed for small values of s/Rb is associated with the lift coefficient crisis ob-
tained for small angles of attack [see figure 5.3(b)]. The small overshoot of circulation
observed when λ = 15 corresponds to an induction effect of the vortices that are above
the rotor plane for s/Rb > 0.7. In the windmill brake regimes (figure 5.5(b)), a larger
effect of variation is observed, with an increase of 40 % of the total circulation when λ
changes from 3 to 5.7. This can be understood by the larger relative contribution of the
axial wind in the total velocity for these cases.
The effect of λ on the blade deformation is shown in figure 5.6. The material chosen for
the blade is weakly flexible so the deformation in terms of torsion angles (left plots) and
bending angles (central plots) remains small. For normal flight conditions (upper plots),
the deformation increases as the climbing velocity decreases. The largest deformation
is reached for the slow descending regime. In the windmill brake regime (lower plots),
the opposite behavior is observed : the deformation diminishes when the descent speed
decreases. Both behaviors are in agreement with the variation of circulation with respect
107
0.06 0.06
0.05
0.05
0.04
0.04
Γ̄/ΩRb2
Γ̄/ΩRb2
0.03
0.03
0.02
λ = -10
λ = -20 0.02 λ=3
0.01 λ = -∞ λ = 4.2
λ = 15 λ = 5.7
0 0.01
0.2 0.4 0.6 0.8 1 0.2 0.4 0.6 0.8 1
s/Rb s/Rb
(a) (b)
Figure 5.5: Circulation profile obtained along the blade of rotor A for different tip-speed-
ratio values. (a) Helicopter regimes (β = 40◦ ); (b) Wind turbine regimes ( β = 15◦ ).
Other parameters are the same as in figure 5.4.
-6 -3
×10
2 ×10 8
6
1.5
4
γ[rad]
fθ
1
2
2 6
1.5
4
γ[rad]
fθ
1
2
0.5
λ=3 0 λ=3
0 λ = 4.2 λ = 4.2
λ = 5.7 λ = 5.7
-0.5 -2
0.2 0.4 0.6 0.8 1 0.2 0.4 0.6 0.8 1
s/Rb r/Rb
Figure 5.6: Blade deformation of rotor A for the helicopter regimes with β = 40◦ (upper
plots) and for the wind turbine regimes with β = 15◦ (lower plots). Left: Torsion angle
γ; Center: Bending angle θ; Right: Bending deflection function fθ . The other parameters
are the same as in figure 5.4.
108
to λ. Larger circulation leads to larger deformation. Note however that the variations
of γ and θ along the blade are much smoother and do not exhibit the bumps and jumps
observed in the circulation profile.
The effect of the other parameters on blade deformation for a typical climbing regime
of rotor A (λ = −20) is analyzed in figure 5.7. The torsion angle γ remains always much
smaller than the deflection angle θ. However, both are similarly affected by variations of
E ∗ (upper plots), ρ∗b (central plots) and Fr (lower plots): twisting and bending increase
when E ∗ or ρ∗b decreases or when Fr increases. These similarities can be understood by
going back to equation (5.14). For the parameters of figure 5.7, weight and centrifugal
forces remains small. If the corresponding terms (third and fourth terms in equation
(5.14)) are neglected, the equations for θ and γ become dependent on the parameters E ∗ ,
ρ∗b and Fr2 through the combination ρ∗b E ∗ /Fr2 only. It is then obvious that decreasing ρb
by a factor of 10 is equivalent to decreasing E ∗ by the same factor, or to increasing Fr2
by a factor of 10.
In figure 5.7, the deformation is relatively small. Highly deformed cases can never-
theless be calculated in the same way. In figure 5.8 we illustrate such a configuration by
considering a very compliant material for rotor A. We observe in figure 5.8(c), that the
deflection reaches 50% of the blade length when we consider the flow conditions λ = −10,
ρ∗b = 1 and Fr2 = 100. The same rotor at rest is however almost undeformed as seen
on this figure. We see also that, even for this very deformed case, the torsion remains
very small compared to the deflection. This agrees with the assumption made in equation
(5.9), where the terms involving τ 2 were neglected.
So far, we have considered the simple geometry of rotor A. There is no difficulty to
consider the more complex geometry of rotor B. In figure 5.9, we compare the circulation
profile and blade deformation of both rotors for the same flow conditions, and the same
material. We have chosen a value of λ close to conditions for which the rotor B has been
designed. In figure 5.9(a), we observe that the circulation profile associated with rotor
B is almost constant for s/Rb > 0.4 in agreement with the design properties (Quaranta
et al., 2015). The circulation profile obtained by rotor A is completely different. It gives
a smaller vortex circulation but it gets more deformed by the flow (see figure 5.9(b-c)).
109
×10 -5 0.04
1.2
E * = 10 6 E * = 10 5
0.035 * 6
1 E * = 10 7 E = 10
* 7
E * = 10 8 0.03 E = 10
0.8
0.025
γ[rad]
0.02
fθ
0.6
0.015
0.4
0.01
0.2
0.005
0 0
0.2 0.4 0.6 0.8 1 0.2 0.4 0.6 0.8 1
s/Rb r/Rb
×10 -5
1.2 0.05
*
ρ b = 10 ρ *b = 10
1 ρ *b = 10 2 0.04 ρ *b = 10 2
ρ *b = 10 3 ρ *b = 10 3
0.8 0.03
γ[rad]
0.6 0.02
fθ
0.4 0.01
0.2 0
0 -0.01
0.2 0.4 0.6 0.8 1 0.2 0.4 0.6 0.8 1
s/Rb r/Rb
-4
×10 -6 5
×10
1.2
*
Fr = 1 Fr 2 = 1
1 Fr * = 10 4 Fr 2 = 10
Fr * = 100 Fr 2 = 100
0.8 3
γ[rad]
2
fθ
0.6
0.4 1
0.2 0
0 -1
0.2 0.4 0.6 0.8 1 0.2 0.4 0.6 0.8 1
s/Rb r/Rb
Figure 5.7: Effect of the other parameters on blade deformation of rotor A for β = 30◦ and
λ = −20. The default parameters are E ∗ = 106 , ρ∗b = 100 and Fr2 = 100 Left column:
Twist angle γ; Central column: Bending angle θ; Right column: Bending deflection
function fθ . Upper line: Young Modulus effect. Middle line: Blade density effect. Lower
line: Froude number effect.
110
-5
×10
14
0.5
12
0.4
10
8 0.3
γ[rad]
6 0.2
fθ
4
0.1
2
0
0
-2 -0.1
0.2 0.4 0.6 0.8 1 0.2 0.4 0.6 0.8 1
s/Rb s/Rb
(d)
Figure 5.8: Highly deformable case. Rotor A with a twist angle β = 30◦ for E ∗ = 106 ,
ρ∗b = 1 and λ = −10. Solid line: Fr2 = 100; Dashed line: Fr = 0 (rotor at rest). (a) Torsion
angle γ; (b) Bending angle θ; (c) Bending deflection function fθ . (d) Three-dimensional
visualization of the wake and the deformed rotor.
-7 ×10 -4
0.08 2.5 ×10 8
0.07
2 6
0.06
1.5
0.05 4
Γ̄/ΩRb2
γ[rad]
0.04
fθ
1
0.03 2
0.5
0.02
0
0
0.01
0 -0.5 -2
0.2 0.4 0.6 0.8 1 0.2 0.4 0.6 0.8 1 0.2 0.4 0.6 0.8 1
r/Rb s/Rb r/Rb
Figure 5.9: Effect of the blade geometry. Solid lines: rotor A with β = 20◦ ; Dashed lines:
rotor B. (a) Circulation Γ̄ along the blade; (b) Torsion angle γ; (c) Bending deflection
function fθ . The other parameters are λ = 3.5, E ∗ = 107 , ρb = 100, Fr2 = 100.
111
112
Chapter 6
Conclusions
6.1 Achievements
The purpose of this work was to develop a model simple enough to allow an extensive
parametric analysis but still physically relevant. The model provides the equilibrium state
of a flexible rotor in vertical flight conditions. The solution is assumed stationary (in the
frame rotating at the rotor angular speed), but takes into account the strong coupling
between the rotor geometry and its wake. A simple but complete description of the blade
deformation and of the wake generated by the rotor has been provided. We have shown
that the model is able to describe rigid as well as very flexible rotors. The present model
also works in operational conditions where classical momentum theory does not apply.
In particular, it has been shown that it can describe rotors in slow descending regimes
of helicopter flight, where the vortices created by the rotor move above the rotor plane
before being advected downwards. This regime is often referred to as the Vortex Ring
State in the helicopter community.
The achievements of this work are divided in two main parts: the analysis of the wake
and its coupling with the blades, rigid or flexible.
To model the wake, a Lagrangian free-vortex model has been implemented on a pre-
scribed Joukowski model. A generalization of the original Joukowski model has been
developed where hub vortices are not necessarily placed on the axis but emitted from
the blade out of axis. In that situation, each blade creates two counter-rotating vortices
emitted at two different non-vanishing radii. This configuration is close to the actual wake
generated by the rotors used for wind turbines. A complete analysis of far-wake structure
for this generalized Joukowski model has been addressed. New numerical solutions have
been obtained that extend the uniform helices that are usually used to describe the far
wake generated by a rotor in axial wind. These solutions are spatially periodic and steady
in a rotating and translating frame. We have shown that these solutions can be considered
as deformed helices. They exhibit radial variations that increase as the internal vortex get
closer to the external vortex. Simple numerical approximations have been provided that
capture the main features of these solutions. We have also considered the variation of the
vortex core size associated with the deformation, proving that it has a very weak effect
on the main characteristics of the solutions. However, in this study, the inner structure
of the vortices has not been considered. The phenomena occurring in the vortex cores,
113
such as short-wavelength instabilities, have therefore not been analyzed.
After having looked at the properties of the far-wake, the near wake has been analyzed.
Solutions have been obtained for a wide range of rotor regimes. Classical and general-
ized Joukowski models have been used to describe the wake. Both models have shown
to provide very similar solutions for tip vortices. A discontinuity on the topology of the
solutions has been found at the transition from wind turbine to VRS regimes. For gener-
alized model, another discontinuity is also found for climbing flight rotors. The total flow
induced on the rotor disc has been calculated for the wide range of regimes and compared
with momentum theory prediction. For VRS regimes, for which momentum theory does
not apply, our induced flow predictions are in good agreement with experimental data.
An analysis of the maximum power extracted by a wind turbine has also been performed.
A very good agreement has been also found with momentum theory prediction.
The stability of classical Joukowski model solutions has been analyzed, showing that
all the regimes are unstable. The instability growth rate has been extracted as a function
of the wavenumber for different rotor regimes. We have shown that it is very close to
the analytical results obtained by Gupta & Loewy (1974) for uniform helices. Finally,
the spatio-temporal evolution of the perturbation has also been analyzed in order to
identify the nature convective or absolute of the instability. Climbing flight and wind
turbine regimes have been found to be convectively unstable, while VRS regimes produce
absolutely unstable solutions. This naturally questions whether the solutions we have
obtained in the VRS regime can be observed. It nevertheless agrees with the idea that in
this regime we expect a complex temporal dynamics.
Concerning the coupling with the blades, a parametric study has been performed with
a flexible rotor model. It has been shown that we are capable to describe rigid as well as
very flexible rotors. The rigid version of the model has been compared to experimental
data for several rotors. A good agreement has been found for the lift distribution and
the wake geometry for a wind turbine rotor exhibiting an almost constant circulation
distribution along its blades. However, further comparisons are needed to fully assess the
validity of the model in hover or for other rotors.
It is worth emphasizing that only steady solutions have been considered. The dynamic
of strongly unsteady regimes as experimentally observed in VRS regimes (Stack et al.,
2005; Quaranta, 2017) are outside the scope of the present work. Moreover, we have
no information on the stability of the coupled solutions. We have shown that the wake
is unstable with respect to long-wavelength instability (Widnall, 1972; Gupta & Loewy,
1974; Quaranta et al., 2015). But other instabilities associated with the blade flexibility,
such as flutter, may also be present (Eloy et al., 2007; Shelley & Zhang, 2011).
Concerning the description of the wake, it would be interesting to include the internal
structure of the vortices in the description of the helical wake. Fukumoto & Okulov (2005)
and Blanco-Rodrı́guez et al. (2015) among others have shown that the core of helical vor-
tices is deformed due to the effects of curvature, torsion and strain. In this thesis, we
114
have seen that the helical vortices generated by a rotor are subject to important deforma-
tions. These deformations are known to be responsible of short-wavelength instabilities
that develop in vortex cores. Local curvature induces the curvature instability (Hattori
& Fukumoto, 2014; Blanco-Rodrı́guez & Le Dizès, 2017) while strain causes the ellip-
tic instability (Blanco-Rodrı́guez & Le Dizès, 2016). Both instabilities are expected to
be present in the inner core of solutions. Numerical and experimental analysis of these
instabilities in rotor wakes could be an interesting future work.
Optimization
One of the objectives of this work was to develop a simple, computationally low-cost model
to perform extensive parametric studies. The next step is to use the model together with
an optimization procedure. For a rigid configuration, one could imagine modifying the
blade geometry in order to optimize, for example, the performance of the rotor under
particular flow conditions. For a deformable rotor configuration, same optimization pro-
cedure could also be performed by adding elastic properties of the material. Blades can
be designed to take advantage of the bending-twisting coupling, for example, for load mit-
igation in wind turbines (Lobitz & Veers, 2003; Bottasso et al., 2013). Passive strategies
just depending on blade geometry and material properties could a priori be addressed in
future with the present method.
115
116
Appendices
117
Appendix A
Numerical methodology
In this appendix, the details of calculation of the different steps of the model are presented.
I) Far-wake solutions
− Parameters: R∗ , h∗ , α, ε
∗ h∗
L = (A.1)
N |1/α − κ|
3. Definition of the initial guess structure as two perfect helices along the discretized
axial coordinate:
Typically, the period is repeated at least 7 times from each side, giving a extended
structure of nT = 7(n − 1) + 1 elements.
119
5. Computation of the induced velocity Ukind at each point k as explained in section
2.3.1. First, the induction of the discretized vortices is obtained by adding the
contribution of every segment of all the discretized extended structures, including
external and internal vortices. The induction of a single segment between points
{m, m + 1} of vortex i on the point k of vortex j is computed through the discretized
Biot-Savat formula:
j,k j,k j,k j,k j,k j,k j,k j,k
seg j Γi ((1 − |ri,m |2 )|ri,m | + (1 − |ri,m |2 )|ri,m+1 |)(ri,m × ri,m+1 )ri,m · ri,m+1
Ui,m (ξk ) = j,k j,k j,m 2 j,k j,k j,k
,
4π ((ri,m · ri,m+1 )2 − |ri,m | |ri,m+1 |2 )|ri,m ||ri,m+1 |
(A.3)
j,k
where ri,m is the vector from point m of vortex i to point k of vortex j.
Then, the contribution of the curved section of vortex between the two adjacent
points k − 1 and k + 1 is computed:
!
k k
loc Γ j ∆φj ρj
Uj,k = ln bkj . (A.4)
4πρkj δa
where Ṽzj and Ω̃j are the induced axial velocity and rotation normalized by N Γ/Rext
2
and N Γ/Rext respectively; the same normalization is used for W̃F and Ω̃F . These two
equations are solved using a Newton-Raphson method. To integrate, a trapezoidal
rule is employed. Initial guess values for the numerical resolution are obtained by:
Z L∗
1 h
∗
i
Ω̃0F = ∗ int ext ext int
2π(Ṽz − Ṽz ) + h (Ω̃ − αΩ̃ ) dz , (A.7a)
h (1 − α) 0
Z L∗
h∗ 0
0 ext ext
W̃F = Ṽz + (Ω̃ − Ω̃ ) dz . (A.7b)
0 2π F
7. Radial and azimuthal coordinates of the deformed helical structure are calculated
from equations
dr Ṽr dφ Ω̃ − Ω̃F
= , = . (A.8)
dz Ṽz − W̃F dz Ṽz − W̃F
120
Which are discretized by a finite difference scheme:
rkext − rk−1
ext
(Dz r ext )k = , (A.10)
∆z
where ∆z = L∗ /n. Dirichlet conditions are applied on the first element of the
position vectors:
r0ext = 1, r0int = R∗ , φext int
0 = 0, φ0 = 0. (A.11)
As induced and frame velocities depend on the geometry of the vortices, the system
is coupled. It is solved with Matlab software. Particularly, function fsolve is used.
When several blades are considered (N > 1), the resolution process does not change.
The 2π/N azimuthal periodicity is imposed. It is just necessary to include the
induction effect of the discretized segments of the other vortices by applying Biot-
Savart formula (A.3).
3. Definition of the initial guess structure as two perfect helical vortices advected by
the external flow:
121
4. Computation of the induced velocity as explained before in section A.I.5, including
the contribution of the discretized segments and the local curved vortex section.
To complete the computation of the induced velocity U ind , the contribution of the
far-wake has to be added. For that, a far-wake structure is calculated as in section
A.I., taking as geometric parameters:
tip
R∗ = rnhub /rntip , h∗ = (zntip − zn−n )/rntip ,
hub hub tip tip ∗ tip tip
t
(A.13)
α = (zn − zn−nt )/(zn − zn−nt ), ε = εR /rn .
Then, the induced velocity of the far-wake structure is computed on the point k using
discretized Biot-Savart formula (A.3) and including the effect of every segment of
internal and external vortices of the far-wake structure:
n
XFW n
XFW
ind
UFW (ξkj ) = seg
UextFW,m (ξkj ) + seg
UintFW,m (ξkj ) (A.14)
m=1 m=1
A simpler way to estimate the effect of the far-wake is to consider the far-wake
structure as perfect helices of constant pitch and radius instead of computing the
ind
deformed structures. The difference on the induced flow contribution UFW is small
and it does not have an important effect on the near-wake geometry, on the contrary,
computation time is noticeably reduced.
5. The coordinates of the position of the vortices are computed as in section A.I through
the following discretized system:
For standard Joukowski model, the same process is followed. Parameters Rb∗ and α
disappear and system (A.15) is reduced to r tip and z tip variables. The induction of the
straight hub vortex is computed with Biot-Savart formula (2.3). As occurs for generalized
far-wake solutions, when several blades are considered (N > 1), the same process is
followed, just the induction effects of the other vortices have to be considered to compute
induced velocities.
1. Guess initial values of Γ, Rtip and Rhub . Normally, a good guess for tip vortex is
Rtip = 1. For Γ and Rhub , it depends on the blade geometry.
122
2. Compute the wake geometry as in section A.II taking
3. With the wake structure obtained in previous point, compute the induced velocity
on the blades: Z 2π
ind 1
V̄ (r) = URind (r, θ)dθ, (A.18)
2π 0
where URind (r, θ) is the induced velocity on the rotor plane, computed from the
contribution of all the segments of the discretized wake with Biot-Savart formula
A.3. Velocities are normalized by N Γ/R. Then, local velocity U (r) and angle of
attack α(r) are computed
q
U (r) = (1/λ + η V̄zind (r))2 + r2 (η Ω̄ind (r) − 1)2 , (A.19)
1
Γ̄(r) = c(r)U (r)CL (α; r)F, (A.21)
2
2 N (1 − r)
F = arccos exp − . (A.22)
π 2r sin φ
Γ = max(Γ̄(r)), (A.23)
r(Γ)
Z R
dΓ̄
Z
dΓ̄
r dr r dr
dr r(Γ) dr
Rhub = Z0 r(Γ) , Rtip = Z R
, (A.24)
dΓ̄ dΓ̄
dr dr
0 dr r(Γ) dr
6. Iterate from points 2 to 5 until the relative variation of Γ, Rtip and Rhub is lower
than 10−4 :
|Γq+1 − Γq |
< 10−4 , (A.25)
|Γq |
where q is the iteration number.
123
IV) Deformed blades
− Parameters: c(r), c0 (r), c00 (r), β(r), A∗ , I ∗ , J ∗ , CL (α, r), CD (α, r); λ, ε; ρ∗b , E, ν
1. Discretize the blade in the local span coordinate s in nb elements: sk = k/nb for
k = {0, ..., nb }. Initially, the blade is assumed undeformed, so bending and twisting
angles are θk = 0 and γk = 0.
The deformed blade position is described by ξb (s) ≡ {rb (s), φb (s), zb (s)}.
2. Compute the wake structure and its induction on the rotor assuming rigid blades.
Follow the algorithm of section A.III.
3. Compute circulation distribution Γ̄(s) and calculate the vortex strength η = Γ/ΩR Rb2
and the position of the vortices on the blade stip and shub . This is computed with
(A.21-24) replacing r by blade span coordinate s.
5. Obtain bending θ(s) and twist γ(s) angles by solving the beam model equations
(5.14) and (5.15) discretized with a finite difference scheme:
∗ 4 θk+2 − 3θk+1 + 3θk − θk−1 ∗ 0 ∗ 3 θk+1 − 2θk + θk−1
ck + 8ck ck +
∆s3 ∆s2
(A.28)
Fr2
∗0 ∗2 ∗ 00 ∗ 3 θk+1 − θk
+4(3ck ck + ck ck ) = ∗ ∗ Fθ (sk ),
∆s ρb EI
2(1 + ν)Fr2
∗ 4 γk+1 − 2γk + γk−1 ∗ 3 ∗ 0 γk+1 − γk−1
ck + 4c k kc = Fγ (sk ). (A.29)
∆s2 2∆s ρ∗b EJ ∗
Note that Fθ and Fγ are computed from θ and γ distributions from the previous
iteration, so the system is linear and completely uncoupled. Just the boundary
conditions have to be included:
θnb +1 − θnb θnb +2 − θnb+1 + θnb
θ0 = 0, = 0, = 0,
∆s ∆s2 (A.30)
γnb +1 − γnb −1
γ0 = 0, = 0,
∆s
124
6. Compute a new wake geometry using as boundary conditions the emission points of
the vortices in the deformed blade:
where rb (s) and zb (s) are computed from θ(s) and γ(s) as
Z s Z s
rb (s) = cos(θ(s̃))ds̃, zb (s) = cos(β(s̃) − γ(s̃)) sin(θ(s̃))ds̃ (A.32)
0 0
The vortices are assumed to shed from the same azimuthal position. This simplifies
the model and, as the induced velocity on the blades is azimuthally averaged, it
does not not imply a big difference on the solution.
7. Compute the induced velocity on the blades along the span coordinate s
Z 2π
ind 1
V̄ (s) = U ind (rb (s), zb (s); φ)dφ, (A.33)
2π 0
where velocities are normalized by N Γ/Rb . Then, local velocity U (s) and angle of
attack α(s) are computed:
q
U (s) = (1/λ + η V̄zind (s))2 + rb2 (s)(η Ω̄ind (s) − 1)2 , (A.34)
With this distribution of α, we obtain lift and drag coefficients distribution CL (s)
and CD (s) by interpolating the airfoil data CL (α) and CD (α).
8. Compute a new circulation distribution Γ̄(s) and calculate vortex strength Γ and
the new position of the vortices on the blade stip and shub .
125
126
Bibliography
Abbott, I. & Von Doenhoff, A. 1960 Theory of wing sections. Dover Publications.
Bauchau, O. & Craig, J. 2009 Structural Analysis. Solid Mechanics and Its Applica-
tions. Springer.
Bazilevs, Y., Hsu, M.-C., Akkerman, I., Wright, S., Takizawa, K., Henicke,
B., Spielman, T. & Tezduyar, T. E. 2011 3D simulation of wind turbine rotors at
full scale. Part I: Geometry modeling and aerodynamics. Int. J. Numer. Meth. Fluids
65, 207–235.
Betchov, R. 1965 On the curvature and torsion of an isolated vortex filament. J. Fluid
Mech. 22, 471–479.
Betz, A. 1920 Das maximum der theoretisch möglichen ausnützung des windes durch
windmotoren. Zeitschrift das gesamte Turbinewessen.
Betz, A. 1926 Windenergie und ihre Ausnützung durch Windmühlen. Göttingen: Van-
denhoeck und Ruprecht.
Bliss, D. B. 1993 New vortex dynamics methods for rotor free wake analysis. technical
report ARO-25747. Tech. Rep.. Durham, USA.
127
Bolnot, H., Le Dizès, S. & Leweke, T. 2014 Spatio-temporal development of the
pairing instability in an infinite array of vortex rings. Fluid Dyn. Research 46, 061405.
de Bothezat, L. 1919 The general theory of blade screws. tr 29. Tech. Rep.. NACA.
Carr, L. W. 1988 Progress in analysis and prediction of dynamic stall. J. Aircraft 25,
6–17.
Carrera, E., Giunta, G. & Petrolo, M. 1926 Beam structure: classical and ad-
vanced theories. Wiley & Sons.
Crow, S. C. 1970 Stability theory for a pair of trailing vortices. AIAA J. 8 (12), 2172–
2179.
Dias, M. A. & Audoly, B. 2014 A non-linear rod model for folded elastic strips.
Journal of the Mechanics and Physics of Solids 62, 57–80.
Dias, M. A. & Audoly, B. 2015 Wunderlich, meet Kirchhoff: A general and unified
description of elastic ribbons and thin rods. J. Elast 119, 49–66.
Drees, J. M. & Hendal, W. P. 1951 The field of flow through a helicopter rotor
obtained from wind tunnel smoke tests. J. Aircraft Eng. 23, 107–111.
128
Durán Venegas, E. & Le Dizès, S. 2018 Structure and stability of rotor generated
vortices. In European Fluid Mechanics Conference 12 .
Durán Venegas, E. & Le Dizès, S. 2019 Generalized helical vortex pairs. J. Fluid
Mech. 865, 523 – 545.
Durán Venegas, E., Le Dizès, S. & Eloy, C. 2019 A strongly-coupled model for
flexible rotors. J. Fluids Struct. .
Froude, R. 1889 On the part played in propulsion by difference of fluid pressure. Trans.
Inst. Naval Arch. 30, 390–405.
Froude, W. 1878 On the elementary relation between pitch, slip and propulsive effi-
ciency. Trans. Inst. Naval Arch. 19, 22–33.
Fukumoto, Y. & Okulov, V. L. 2005 The velocity induced by a helical vortex tube.
Phys. Fluids 17, 107101.
Gessow, A. & Gustafson, F. B. 1948 Effect of rotor-blade twist and plan-form taper
on helicopter hovering performance. tn 1542. Tech. Rep.. NACA.
Goldstein, M. A. 1929 On the vortex theory of screw propellers. Proc. R. Soc. Lond.
A 123, 440–465.
129
Gupta, B. P. & Loewy, R. G. 1974 Theoretical analysis of the aerodynamic stability
of multiple, interdigitated helical vortices. AIAA J. 12, 1381–1387.
Gupta, S. & Leishman, J. G. 2005a Accuracy of the induced velocity from helical
vortices using straight-line segmentation. AIAA J. 43, 29–40.
Gustafson, F. B. & Gessow, A. 1946 Effect of rotor tip speed on helicopter rotor
performance and maximum forward speed. naca arr no. l6a16. Tech. Rep.. NACA.
Han, S., Benaroya, H. & Wei, T. 1999 Dynamics of transversely vibrating beams
using four engineering theories. Journal of Sound and Vibration 225, 935–988.
Hand, H., Simms, D., Fingersh, L., Jager, D., Cotrell, J., Schreck, S. &
Larwood, S. 2001 Unsteady Aerodynamics Experiment Phase VI: Wind tunnel test
configurations and available data campaingns. Tech. Rep.. NREL/TP-500- 29955.
Hardin, J. C. 1982 The velocity field induced by a helical vortex filament. Phys. Fluids
25, 1949–52.
Hattori, Y. & Fukumoto, Y. 2014 Modal stability analysis of a helical vortex tube
with axial flow. J. Fluid Mech. 738, 222–249.
Ivanell, S., Mikkelsen, R., Sorensen, J. & Henningson, D. 2010 Stability anal-
ysis of the tip vortices of a wind turbine. Wind Energ. 13, 705–715.
130
Katz, J. & Plotkin, A. 1991 Low-Speed Aerodynamics: from wing theory to panel
methods. Mc Graw-Hill.
Kawada, S. 1936 Induced velocity by helical vortices. J. Aero. Sci. 3 (3), 86–87.
Kerswell, R. R. 2002 Elliptical instability. Annu. Rev. Fluid Mech. 34, 83–113.
Kida, S. 1981 A vortex filament moving without change of form. J. Fluid Mech. 112,
397–409.
Kida, S. 1982 Stability of a steady vortex filament. J. Phys. Soc. Japan 51, 1655–1662.
Klein, R., Majda, A. J. & Damodaran, K. 1995 Simplified equations for the inter-
action of nearly parallel vortex filaments. J. Fluid Mech. 288, 201–248.
Knight, M. & Hefner, R. 1937 Static thrust of the lifting airscrew. Tech. Rep.. NACA
TN 626.
Leishman, J. G., Bhagwat, M. J. & Ananthan, S. 2004 The vortex ring state
as a spatially and temporally developing wake instability. J. Am. Helicopter Soc. 49,
160–175.
Levy, H. & Forsdyke, A. G. 1928 The steady motion and stability of a helical vortex.
Proc. R. Soc. Lond. A 120, 670–690.
131
Lobitz, D. & Veers, P. S. 2003 Load mitigation with bending/twist coupled blades
on rotors using modern control strategies. Wind Energy 6.2, 105–117.
Lock, C., Bateman, H. & Townend, H. 1925 An extension of the vortex theory
of airscrews with applications to airscrews of small pitch and including experimental
results. Aeronautical Research Committee. Reports and Memoranda 1014.
Lv, P., Mohd-Zawawi, F., Benard, E., Prothin, S., Moschetta, J. M. & Mor-
lier, J. 2013 An application of adaptive blades on convertible mavs. International
Journal of Micro Air Vehicles 5 (4), 229 – 243.
Mullerners, K. & Raffel, M. 2012 The onset of dynamic stall revisited. Exp. Fluids
52, 779–793.
Munk, M. 1924 Elements of the wing section theory and of the wing theory. rept no 191.
Tech. Rep.. NACA.
Okulov, V. L. 2004 On the stability of multiple helical vortices. J. Fluid Mech. 521,
319–342.
Okulov, V. L. & Sorensen, J. N. 2007 Stability of helical tip vortices in a rotor far
wake. J. Fluid Mech. 576, 1–25.
Okulov, V. N. 2016 An acentric rotation of two helical vortices of the same circulations.
Regular and Chaotic Dyn. 21, 267–273.
Payne, P. R. 1959 Helicopter Dynamics and Aerodynamics. Pitman & Sons, London.
Prandtl, L. & Betz, A. 1927 Vier abhandlungen zur hydrodynamik und aerodynamik.
Gottingen Nachr. pp. 88–92.
Prouty, R. W. 1986 Helicopter Performance, Stability and Control . Prindle, Weber &
Schmidt, PWS Engineering, Boston.
132
Quaranta, H. U., Brynjell-Rahkola, M., Leweke, T. & D.S., H. 2018 Local
and global pairing instabilities of two interlaced helical vortices. J. Fluid Mech. 863,
927 – 955.
Quaranta, U. 2017 Instabilities in a swirling rotor wake. PhD thesis, Aix Marseille
Université.
Rankine, W. J. M. 1865 On the mechanical principles of the action of propellers. Trans.
Inst. Naval Arch. 6, 13–39.
Reissner, H. 1937 On the vortex theory of the screw propeller. J. Aeronaut. Sciences 5
(1), 1–6.
Reissner, H. 1940 A generalized vortex theory of the screw propeller and its application.
TN 750. Tech. Rep.. NACA.
Ricca, R. L. 1994 The effect of torsion on the motion of a helical vortex filament. J.
Fluid Mech. 273, 241–259.
Saffman, P. G. 1992 Vortex dynamics. Cambridge University Press.
Sarmast, R., Dadfar, R., Mikkelsen, R. F., Schlatter, P., Ivanell, S.,
Sorensen, J. & Henningson, D. 2014 Mutual inductance instability of the tip
vortices behind a wind turbine. J. Fluid Mech. 755, 705–731.
Schepers, J. G., Boorsma, K., Cho, T., Gomez-Iradi, S., Schaffarczyk, P.,
Jeromin, A. & Sørensen, N. N. 2012 Analysis of mexico wind tunnel measurements:
Final report of IEA task 29, Mexnext (phase 1). Tech. Rep.. Research Centre of the
Netherlands (ECN).
Schepers, J. G. & Snel, H. 2007 Mexico, model experiments in controlled conditions.
ECN-E-07-042. Tech. Rep.. Research Centre of the Netherlands (ECN).
Selçuk, C., Delbende, I. & Rossi, M. 2017 Helical vortices: Quasi-equilibrium states
and their time evolution. Phys. Rev. Fluids 2, 084701.
Selçuk, C., Delbende, I. & Rossi, M. 2018 Helical vortices: linear stability analysis
and nonlinear dynamics. Fluid Dyn. Res. 50, 011411.
Shelley, M. J. & Zhang, J. 2011 Flapping and bending bodies interacting with fluid
flows. Annu. Rev. Fluid Mech. 43, 449–465.
Shen, W. Z., Mikkelsen, R., Sørensen, J. N. & Bak, C. 2005 Tip loss corrections
for wind turbine computations. Wind Energy 8 (4), 457–475.
Sicard, U. 2014 Development of an extremely flexible, variable-diameter rotor for a
micro-helicopter. PhD thesis, The University of Texas at Austin.
Sicklinger, S., Kürkchübasche, A., Belsky, V., Wüchner, R. & Bletzinger,
K.-U. 2015 Fluid-structure interaction with controls forwind turbine simulation using
a novel co-simulation approach. Journal of Wind Engineering and Industrial Aerody-
namics 144, 135–145.
133
Snel, H., Schepers, J. & Montgomerie, B. 2007 The mexico project (model exper-
iments in controlled conditions): The database and first results of data processing and
interpretation. Journal of Physics: Conference Series 75.
Sorensen, J. & Shen, W. 2002 Numerical modeling of wind turbine wakes. J. Fluids
Eng. 124, 393–399.
Sørensen, J. N. 2016 General momentum theory for horizontal axis wind turbines,
Springer series: Research topics in wind energy, vol. 4. Springer.
Stack, J., Caradonna, F. & Savas, O. 2005 Flow visualizations and extended thrust
time histories of rotor vortex wakes in descent. J. Am. Helicopter Soc 50, 279–288.
Velasco Fuentes, O. 2018 Motion of a helical vortex. J. Fluid Mech. 836, R1.
Vermeer, L. J., Sørensen, J. N. & Crespo, A. 2003 Wind turbine wake aerody-
namics. Prog. Aero. Scie. 39, 467–510.
Washizu, K., Azuma, A., Koo, J. & Oka, T. 1966 Experiments on a model helicopter
rotor operating in the vortex ring state. Journal of Aircraft 3(3), 225–230.
Widnall, S. E. 1972 The stability of a helical vortex filament. J. Fluid Mech. 54,
641–663.
134