You are on page 1of 21

Modelling and Analysis of a Coaxial Tiltrotor

Naman Rawal Abhishek


Graduate Student Assistant Professor
nam.kumar31@gmail.com abhish@iitk.ac.in
Department of Aerospace Engineering
Indian Institute of Technology Kanpur
Kanpur, India

ABSTRACT
This paper proposes a new coaxial tiltrotor design and discusses the development of equations of motion for its analysis
in airplane mode, helicopter mode, and transition mode flights. The proposed design is characterized by a coaxial prop-
rotor system that is capable of converting from a lifter to a thruster between vertical and forward flight conditions. Half
of the wing is also tilted along with the rotor, while the remaining out-board half remains in horizontal position at all
times. The rotor dynamics is modeled using rigid blades with only flap degree of freedom. Inflow is estimated using
Drees’ model. This paper is divided into two parts. First part of the paper compares and highlights the performance
improvement of the proposed vehicle concept over that of the conventional coaxial helicopter. Second part analyses the
problem of transition of the vehicle from helicopter mode to the fixed wing mode in a quasi-steady manner. Proposed
tiltrotor configuration offers significant improvements over helicopter configuration with dramatic improvements in
maximum cruise velocity, range, endurance and rate of climb. The quasi-steady transition analysis shows that the
proposed design can transition from helicopter mode to aircraft mode successfully at wide range of forward speeds.

INTRODUCTION velopment of a real-time flight simulation model for the XV-


15 (Ref. 3). Eight subsequent simulation periods provided
Conventional helicopters continue to be significantly slower, major contributions in the areas of control concepts, cockpit
noisier and suffer from higher vibration than their fixed wing configuration, handling qualities, pilot workload, failure ef-
counterparts. They also have lower range and endurance for fects and recovery procedures, and flight boundary problems
similar all up weights. Fixed wing aircraft seems to be bet- and recovery procedures. The fidelity of the simulation also
ter in every possible aspect, except that they can’t hover or do made it a valuable pilot training aid, as well as a suitable tool
Vertical Take-off and Landing (VTOL). As helicopters pro- for military and civil mission evaluations. Recent simulation
gressed past the World War II, researchers started running periods have provided valuable design data for refinement of
into expected speed limitations inherent to all rotorcraft that automatic flight control systems. Throughout the program, fi-
generate all of its lift and thrust from a rotor in edgewise delity has been a prime issue and has resulted in unique data
flight. In addition to a push for higher cruise speeds there and methods for fidelity evaluation which are presented and
was requirement for greater range. There were several ways to discussed in (Ref. 4).
achieve improved speed and range and they all have their mer-
As discussed in (Ref. 5), even though the XV-3 was gener-
its. Therefore, in 1950s and 1960s there was a significant im-
ally the best behaved, it had several notable issues. During
petus on the development of novel hybrid air vehicles, such as
low speed flight, weak lateral-directional dynamic stability
tiltwings, compound helicopters etc., which could bridge the
and longitudinal and directional controllability were experi-
gap between aircrafts and helicopters. Most of these designs
enced (Ref. 6). These issues were attributed to rotor wash
were complicated and could not be realized with the state of
at low altitudes. During cruise, lateral-directional (dutch-roll)
technology available during 1950s and 60s.
and longitudinal (short-period) damping was reduced as speed
During 1970s work began on the XV-15 (Refs. 1, 2) which was increased (Ref. 7). This was found to be due to large in-
was a 13,000 lb (6000 kg) tiltrotor. Each rotor had three 12.5 plane rotor forces created from blade flapping resulting from
foot radius blades mounted on a gimbaled hub. A large mod- aircraft angular rates. In spite of these issues, the tiltrotor con-
eling effort was undertaken to accompany development and figuration proved to be effective and transition between heli-
flight test of the aircraft. This modeling effort included de- copter and airplane mode was shown to be safe, paving the
way for additional research efforts for this configuration.
Presented at the 4th Asian/Australian Rotorcraft Forum, IISc,
India, November 16–18, 2015. Copyright c 2015 by the An in-depth NASA investigation examined several types
Asian/Australian Rotorcraft Forum. All rights reserved. of rotorcraft for large civil transport applications, and con-
1
cluded that the tiltrotor had the best potential to meet the de- APPROACH
sired technology goals (Ref. 8). Goals were included for hover
and cruise efficiency, empty weight fraction, and noise. The This paper discusses the modelling, performance study and
tiltrotor also presented the lowest developmental risk of the transition analysis of a hybrid tiltrotor/tiltwing vehicle, which
configurations analysed. One of the four highest risk areas consists of a tilting coaxial rotor system (see Fig. 1). In ad-
identified by the investigation was the need for broad spec- dition, the portion of the wing in the downwash of the rotors
trum active control, including flight control systems, rotor also tilt along with the rotor system. This arragnement allows
load limiting, and vibration and noise. Some general infor- the vehicle to take-off and land and transition into a fixed wing
mation on the stability and control of tiltrotors can be found aircraft by tilting the rotor and half-wing combination.
in the Generic Tiltrotor Simulation (GTRS) documentation of
the XV-15 modeling as documented by Sam Ferguson of Sys-
tems Technology, Inc. (Refs. 9, 10).
Even though a lot of research has been done in tiltrotors,
mainly XV-15, a coaxial tiltrotor has never been realised and
is still in experimental stage. Moreover, literature that de-
scribes the basics of coaxial tiltrotor aeromechanics using the
basic Euler equations, flapping equations of motions, and ba-
sic helicopter and airplane theory is difficult to find. A mono
tiltrotor design has been proposed by the Baldwin Technology
Company as an innovative VTOL concept that integrates a tilt- (a) Hover mode (b) Airplane mode
ing coaxial rotor, an aerodynamically deployed folding wing,
Fig. 1. Proposed model for a coaxial tiltrotor
and an efficient cargo handling system (Ref. 11). It was found
that the MTR concept, if practically realized, offers unprece-
The hover performance analysis is carried out using Blade
dented performance capability in terms of payload, range, and
Element Momentum Theory (BEMT). The forward flight
mission versatility.
analysis is performed using a coaxial rotor dynamics analysis
This paper proposes and analyses a new design for coaxial in which the blades are modelled as rigid with hinge offset and
tiltrotor which is different and possibly simpler than that stud- root spring with only flap degree of freedom. Aerodynamic
ied in (Ref. 11). Current design with coaxial rotor promises to loads are estimated using Blade Element Theory (BET) cou-
offer several advantages over conventional tandem rotor tilt- pled to Drees inflow model. The trim analysis in helicopter
rotor systems (Ref. 12): mode and airplane mode is carried out by Newton Raphson
approach.
• lower actuator forces and moments for tilting the rotors The transition analysis is carried out in a quasi-steady man-
as coaxial rotors tend to have significantly lower gyro- ner and the ability of the proposed design to perform success-
scopic loads compared to single rotors during transition. ful transition from helicopter mode to aircraft mode is studied
at different forward speeds. The objective is to develop the
• tilting of the wing portion below the rotor minimizes the methodology for analysis to provide systematic performance
hover losses due to aerodynamic download and interfer- comparison with an equivalent pure helicopter (coaxial con-
ence of the rotor wake and the wing. figuration). Impact of the proposed design changes on the key
performance parameters such as speed, range, endurance etc.
• while a fully tilting wing is ideal to avoid losses, a half are discussed in detail. The detailed derivation of the equa-
tiltwing reduces actuator forces and moments compared tions for hover, forward flight trim and performance analysis
to a full tiltwing. and transition analysis is presented in Appendix A.

• the remaining fixed half wings is expected to help dur- PERFORMANCE ANALYSIS IN HOVER
ing the transition by contributing to lift generation at low
forward speeds. process. Numerical studies are done by modifying the XV-15 aircraft
parameters for a coaxial tiltrotor design proposed in this pa-
per. The modified set of parameters were then used as input
The trim analysis for coaxial rotor system is first developed
for a quasi-steady simulation. The complete set of data is pre-
and validated with Harrington rotor’s test data. The trim anal-
sented in Appendix B and are taken from (Refs. 2, 3). Table
ysis for coaxial rotor system is then refined to model coaxial
1 summarises the input parameters used for this simulation.
tiltrotor configuration. The performance benefits in terms of
The following items should be noted prior to this discussion:
range, endurance, rate of climb etc. of proposed coaxial tiltro-
tor design over corresponding coaxial helicopter are system-
atically studied. A quasi-steady analysis of transition of the • The center of gravity is assumed to be constant and does
tiltrotor from helicopter mode to aircraft mode is carried out. not move with the movement of the nacelle.
2
• All the trims were run at 6000 kg gross weight, a 0◦ flight
0.01
path angle and turn rate equal to 0◦ /s.

Power Coefficient (CP)


• Current analysis does not include modelling of stall char- 0.008
acteristics of either rotary wing or fixed wing.
0.006
• The performance analysis provides results for the mini-
mum power consumption of the rotors under a given set 0.004
of conditions, and so provides a datum against which the
efficiency of a real rotor can be measured. 0.002 Present Simulation
Harrington’s Experiment

0
No. of blades per rotor 3 0 0.2 0.4 0.6 0.8 1
−3
Thrust Coefficient (C ) x 10
No. of rotors 2 T

Rotor solidity 0.089


No. of engines 1 Fig. 2. CT vs CP prediction for Harrington Rotor 1 at 450
Max. take-off weight(kg) 6000 RPM and σ = 0.054
Rotor Diameter(m) 8
Wingspan(m) 10 due to fuselage is unavoidable but design adjustments can be
Max. power available(kW) 1800 made such as a tiltwing to avoid vertical drag due to wing.
Fuel weight(kg) 676 In the simplest form the vertical drag can be accounted for
Wing aspect ratio 6.25 by assuming an equivalent drag area fν or a drag coefficient
Specific fuel consumption (kg/Watt-s) 3.8 × 10−4 CDv based on a reference area, say Sre f . This means that the
extra rotor thrust to overcome this drag will be
Table 1. Key design inputs for coaxial tiltrotor simulation

1
∆T = Dv = ρ ν̄ 2 fnu (1)
Validation of Hover Analysis 2

The modelling process involved developing a trim routine for where ν̄ is the average velocity of the rotor slipstream.
a generic coaxial rotor and then modifying the math model Using the Simple Momentum Theory, an expression for
to incorporate the effects of a tilting mast. The hover per- calculating the net rotor power requirement is given by
formance prediction for a regular coaxial rotorcraft using the (Ref. 14):
BEMT is validated against the experimental results obtained
by Harrington for nominally full scale rotor systems (Ref. 13).  
The results comparing the current predictions with experi- v
W  Vc u
u W
mental data for Harrington Rotor 1 with 25ft diameter and P=  +t    + P0 (2)
fν 2
solidity of 0.054 for coaxial rotor with untwisted blades and a 1− A 2ρ A f ν A
taper ratio of approximately 3:1 are shown in Fig. 2. Predicted
results show excellent correlation.
Using the above equation, the download penalty is calcu-
The coupled trim code for coaxial rotor has also been de- lated for three different configurations based on the area of
veloped and validated and the combination of the two analysis wing that is tilted along with the rotor system and shown in
would be used to carry out performance evaluation and tran- Table 2. These three configurations are:
sition analysis of the novel coaxial tiltrotor/tiltwing vehicle.

1. No wing area under tilt


Hovering Performance Comparison
2. Half wing area under tilt
An important feature of the proposed coaxial tiltrotor is the
half tiltwing design. It is normally assumed that the total
3. Full wing area under tilt
thrust, T, required by the rotor system is equal to the weight
of the helicopter, W. However, there is usuaiiy an extra incre-
ment in power required because of the download or vertical It can be intuitively stated that the actuator force required
drag, Dv , on the helicopter fuselage that results from the ac- to tilt a wing is directly proportional to the wing area under
tion of the rotor slipstream velocity. Typically, the vertical tilt. On the other hand Table 2 shows an inverse relation in
download on the fuselage can be up to 5% of the gross takeoff power required to hover with area of wing under tilt. Thus,
weight, but it can be much higher for rotorcraft designs such qualitatively, it can be stated that a half tiltwing would provide
as compounds or tilt-rotors that have large wings situated in a trade-off between actuator forces and download penalty and
the downwash field below the rotor. The download penalty is therefore incorporated in the current hybrid tiltrotor design.
3
Vehicle Configuration Power Required in Hover
0.07
No Tiltwing 2245 kW
Half Tiltwing 1546 kW
0.06 λ
Full Tiltwing 1165 kW u
λl

Table 2. Comparison of power required in hover for dif- 0.05

Inflow, λ
ferent tiltwing configurations
0.04
COAXIAL TILTROTOR SIMULATION
0.03
The simulation was designed to incorporate the elements of
basic rotary wing dynamics along with reasonable assump- 0.02
tions while maintaining the accuracy of the model. The per-
formance analysis involves running the simulation for oper- 0.01
0 20 40 60 80 100
ational extremities of the proposed vehicle and obtaining the True Airspeed (in m/s)
trim results. A qualitative and quantitative analysis is then
carried out to predict the performance advantage of one oper-
Fig. 3. Non-dimensional inflow in helicopter mode( w/o
ational mode over another.
wings)
The most important factor in the development of vehicles
that can operate in multiple configurations/orientations is the 1
feasibility of transition. To investigate the proposed design’s λ
0.9 u
transition capabilities, a quasi-static analysis is carried out for λ
l
a prescribed transition corridor. The results of this analysis
0.8
then throws light on the feasibility and nature of transition.
Inflow, λ

0.7
Trim Analysis
0.6

The two operational extremities of the proposed design are


0.5
helicopter mode (w/o wings) and airplane mode (with wings)
for which the performance analysis is carried out. The vehicle 0.4
configuration transitions from helicopter mode (w/o wings)
(βM = 0◦ ) to airplane mode (with wings) (βM = 90◦ ) where 80 100 120 140 160 180 200
True Airspeed (in m/s)
βM is the angle between the coaxial rotor mast with the verti-
cal axis of the aircraft. The analysis was carried out for speeds
from 0 m/s to 105 m/s (378 km/hr) and from 90 m/s (324 Fig. 4. Non-dimensional inflow in airplane mode (with
km/hr) to 200 m/s (720 km/hr). wings)

Figures 3 and 4 shows trim results for non dimensional In airplane mode (with wings), the cyclic pitch require-
rotor inflow on upper and lower rotor. In helicopter mode (w/o ments are very small compared to the collective pitch require-
wings), the upper rotor experiences a lesser inflow compared ments. This result seem intuitive as the vehicle is moving
to the lower rotor. This can be attributed to the fact that the in the axial direction. The collective power requirement for
lower rotor operates in the vena contracta of the upper rotor. both rotors are almost equal. The difference between the two
In airplane mode (with wings), the inflow velocities seen by collectives is negligible compared to the high requirements in
both the rotors are almost equal for a given flight condition as forward flight.
the difference due to interference is very small compared to The collective pitch requirement shown in the results may
the high inflow velocities as it operates at higher speeds. appear absurd as they range from about 0◦ to 45◦ which is
Figures 5 and 6 the control angle requirements to trim impossible to achieve without encountering stall. A further
the vehicle in a given flight condition. In helicopter mode investigation of local blade element reveals the effective an-
(w/o wings), the collective pitch requirement for lower rotor gle of attack at the blade element. The results are shown in
is slightly higher than the upper rotor which is again a con- Figures 8, 7 for extremities i.e. helicopter mode (w/o wings)
sequence of the lower rotor operating under the influence of and airplane mode (with wings). It is clear from the plots that
the wake of the upper rotor. At hover, the cyclic pitch re- the blade root might be under stall but for most part, the ef-
quirements are zero but as the vehicle gains forward speed, fective angle of attack is within reasonable limits.
the rotors tilts backwards as the longitudinal cyclic pitch re- The attitude of the vehicle in helicopter mode (w/o wings)
quirement increases in the negative direction. Lateral cyclic is shown in Fig. 9. Pitch requirements in helicopter mode
pitch requirement variation is very small as the rotor mast tilts (w/o wings) shows a decrease which means a nose down
to counter the unbalanced side slip forces and roll moments. movement which when compared to the longitudinal cyclic
4
Control Angles (in degrees) 20 20

10 15

α (in degrees)
0 10

θ β = 0°, V = 0 m/s
−10 0
u
5 M

e
θ
0
l

−20 θ1c 0
θ1s

−30 −5
0 20 40 60 80 100 0 0.2 0.4 0.6 0.8 1
True Airspeed (in m/s) Non−dimensional radial position r/R

Fig. 5. Control angles in helicopter mode (w/o wings) Fig. 7. Effective angle of attack at the blade for βM = 0◦ at
V = 0 m/s
90
θ 20
80 0
u
Control Angles (in degrees)

70 θ0
l
18
60 θ1c
α (in degrees)
50 θ
1s
16
40
30
14 °
β = 90 , V = 180 m/s
e

20 M

10 12
0
80 100 120 140 160 180 200 10
True Airspeed (in m/s) 0 0.2 0.4 0.6 0.8 1
Non−dimensional radial position r/R
Fig. 6. Control angles in airplane mode (with wings)
Fig. 8. Effective angle of attack at the blade for βM = 90◦
pitch requirement in helicopter mode depicts that the rotor at V = 180 m/s
mast tilts backward as the vehicle pitches in nose down di-
The induced power is generated due to the drag induced
rection. The roll requirement is very negligible in helicopter
as a consequence of development of lift at the blade element.
mode (w/o wings). In airplane mode (with wings), the roll
Climb Power is the power required to climb. Parasite Power
and pitch requirements are shown in Fig. 10.
is required to overcome the drag of the helicopter. Profile
power is needed to turn the rotor in air. Since the vehicle
Performance Prediction is in forward flight, the rotor disk is always under climb and
thus climb power contributes significantly to the total power
To analyse the performance of the coaxial tiltrotor, several requirement. Parasite power is accounted for using the para-
performance parameters are evaluated and compared for he- site drag which is calculated in terms of equivalent flat plate
licopter mode (w/o wings) against the airplane mode (with area. For a tiltwing configuration, this flat plate area will be
wings). These performance parameters include calculation of a function of the mast angle but is assumed to be constant for
Power Requirements, Rate of Climb, Endurance and Range. this analysis.
Further, the advantages of a half tiltwing design are investi-
gated using Simple Momentum Theory. The power curves for both helicopter mode (w/o wings)
and airplane mode (with wings) are shown in Fig. 11. The
With all the power transmitted to the main rotor system following inferences are drawn from the results obtained.
through the shaft, we have P = ΩQ. It can be shown that the
nondimensional power coefficient is equal to the nondimen- • The power requirements for helicopter mode (w/o wings)
sional torque coefficient. The total power can be written as a reaches the power available for much lower airspeed than
combination of four components as: when the vehicle is in the airplane mode (with wings).
This is a reflection of the clear airspeed advantage that
P = Pi + Pc + Pp + P0 (3) the airplane possesses over the conventional helicopter.
5
5 2500
pitch Helicopter Mode
Attitude/ Euler Angles (in degrees)

roll Airplane Mode


2000 Power Available

Power (in kW)


1500

1000
−5

500

−10 0
0 20 40 60 80 100 0 50 100 150 200 250
True Airspeed (in m/s) True Airspeed (in m/s)

Fig. 9. Attitude/Euler angles in helicopter mode (w/o Fig. 11. Predicted rotor power vs. true airspeed
wings) Z WT O −W f
V∞ dW
R=− (4)
6 WT O cP
Attitude/ Euler Angles (in degrees)

pitch The total distance covered during a range flight is the in-
4
roll
tegral of this expression and is shown in Fig. 12 for both he-
2 licopter (w/o wings) and airplane mode (with wings). The
predictions made from the results are itemized below.
0
• The maximum range of the vehicle increases by 100% in
−2 airplane mode (with wings) than in helicopter mode (w/o
wings). The maximum range is predicted to exceed 1600
−4 kms in airplane mode (with wings) compared to 800 kms
in helicopter mode.
−6
80 100 120 140 160 180 200
• The vehicle is capable of flying longer distances at higher
True Airspeed (in m/s)
speeds in airplane mode (with wings) than in helicopter
mode (w/o wings).
Fig. 10. Attitude/Euler angles in airplane mode (with
wings) • Upto 400% increase is predicted in range when flying in
• The maximum cruise speed in helicopter mode is pre- airplane mode (with wings) at 100 m/s (360 km/hr) as
dicted to exceed 100 m/s (360 km/hr) whereas in airplane compared to flying in helicopter mode (w/o wings) at the
mode (with wings), the vehicle can theoretically achieve same speed.
speeds exceeding 200 m/s (720 km/hr).

• Upto 75% reduction in power requirements is predicted Endurance


while flying in airplane mode (with wings) as compared
to helicopter mode (w/o wings). The maximum reduc- Total time an airplane stays in the air on a full tank of fuel for
tion in power requirements can be achieved while flying a given flight condition is defined as the endurance of an air-
at airspeeds close to 100 m/s (360 km/hr) than in heli- craft. The generalised endurance parameters are specific fuel
copter mode (w/o wings) at the same speed. consumption, aircraft weight, engine power, and fuel weight.
Endurance can be calculated using the following equation.

Z WT O −W f
Range of Flight dW
E =− (5)
WT O cP
Range of flight is defined as the total ground distance tra-
versed on a full tank of fuel. The weight of fuel consumed The total time an airplane can fly is the integral of this
per unit power per unit time or Specific Fuel Consumption(c) expression and is shown in Fig. 13 for both helicopter and
dictates the range of an aircraft. The generalised range equa- airplane mode (with wings). The inferences drawn from the
tion for propeller-driven airplane is given by: results are presented below.
6
The rate of climb(ROC) capability vs. airspeed is shown
2000 in Fig.14 for both flight modes. As a result of much lower
Helicopter Mode
Airplane Mode power requirements in airplane cruise, the rate of climb capa-
bility far exceeds that of in helicopter mode (w/o wings) . The
1500 power available can be judged by comparing it with the en-
Range (in km)

gine power installed in XV-15 which consists of two engines


of about 1250 kW whereas for a coaxial rotor the power avail-
1000 able is assumed to be 1800 kW. Following inferences can be
drawn from the results for climb capabilities:

500 • The maximum rate of climb of the vehicle is predicted to


exceed 20 m/s (72 km/hr).
• The vehicle can climb faster as higher airspeeds in air-
0
0 50 100 150 200 250 plane mode (with wings) as compared to helicopter mode
True Airspeed (in m/s) (w/o wings) .

Fig. 12. Predicted range vs. true airspeed • Upto 500% increase is predicted in rate of climb when
the vehicle is flying in airplane mode (with wings) at an
• The maximum endurance of the vehicle is predicted to airspeed of 100 m/s (360 km/hr) than in helicopter mode
exceed 4.5 hours. (w/o wings) .
• Upto 450% increase in endurance is predicted in airplane
mode (with wings) than in helicopter mode (w/o wings). 30
The maximum improvement can be achieved by flying at Helicopter Mode
100 m/s (360 km/hr). Airplane Mode
25
Rate of Climb (in m/s)

• For the same endurance, the vehicle is now capable of


flying two times faster in airplane mode (with wings) 20
compared to helicopter mode (w/o wings).
15

10
5

5
4 Helicopter Mode
Airplane Mode
Endurance (in hrs)

0
0 50 100 150 200 250
3 True Airspeed (in m/s)

2
Fig. 14. Predicted rate of climb capability vs. true airspeed

1 TRANSITION ANALYSIS
During transition, the mast angle changes the aircraft config-
0 uration as it varies from 0◦ to 90◦ where βM = 0◦ corresponds
0 50 100 150 200 250
to the helicopter mode (w/o wings) when the rotor plane is
True Airspeed (in m/s)
parallel to the horizontal axis. The mast tilts forward with a
step size of 2.5◦ until the rotor plane is perpendicular to the
Fig. 13. Predicted endurance vs. true airspeed
horizontal axis. The operating speed range for βM = 0◦ is 0
m/s to 50 m/s with a step size of 2.5 m/s. With every sub-
Rate of Climb sequent step increment in mast angle, each velocity step gets
incremented by 2.5 m/s. i.e., at βM = 2.5◦ , the velocities span
Another important parameter in the assessment of aircraft per- from 2.5 m/s to 52.5 m/s. Likewise, for βM = 90◦, the velocity
formance is climb capability, particularly the rate of climb. range is 90 m/s to 140 m/s. This is the prescribed transition
The maximum rate of climb is characterized by the ratio of corridor for a quasi-steady transition analysis.
the excess power to the aircraft gross weight as given by For better visualization, the transition results are plotted
with a colour gradient. A range of colours are generated cor-
Pavail − Preq responding to trim results for every value of βM . The colour
Vcmax = (6) field is indicated using a colour bar.
WT O
7
Inflow Velocities
0.7 90
Figures 15 and 16 shows the variation of inflow on lower and 80
0.6
upper rotor with respect to forward speed. Note that ’inflow’
70

Inflow upper rotor (λu)


here refers to the flow velocity perpendicular to the rotor disk. 0.5
As the rotor disk tilts forward, more and more of the resultant 60
flow becomes perpendicular to the rotor disks or parallel to 0.4 50
the mast.
0.3 40
For βM close to zero, the inflow ratio dips with increase 30
in airspeed before rising again. This is consistent with con- 0.2
ventional rotor system results. For mast angles greater than 20
20◦ , the inflow ratio increases with increase in airspeed. This 0.1
10
seems intuitive as the rotor disk is now more aligned with the
0 0
oncoming flow. 0 50 100 150
True Airspeed (in m/s)
Figure 17 shows the variation of flow parallel to the rotor
disks. For a conventional helicopter, flow parallel to the rotor Fig. 16. Non-dimensional inflow in lower rotor
disks is a measure of the forward velocity of the vehicle which
is why this quantity is referred to as advance ratio, denoted by 0.35 90
µ . However in this analysis, advance ratio is the flow velocity

Advance ratio w.r.t. rotor disc (µ)


parallel to the rotor disk in longitudinal direction. The results 80
0.3
for flow parallel to rotor disk shows an increase for small mast 70
angles. As the mast tilts further forward, the flow becomes 0.25
60
more and more perpendicular to the rotor disk. From the Fig.
0.2 50
17, it can be seen that the flow lines starts decreasing for mast
angles greater than 45◦ . 0.15 40

30
0.1
0.7 90 20
80 0.05
0.6 10
70 0 0
Inflow lower rotor (λl)

0.5 0 50 100 150


60 True Airspeed (in m/s)
0.4 50
40
Fig. 17. Non-dimensional advance ratio on both rotors
0.3
30 control sticks can be assigned to the collective pitch require-
0.2 ment and would influence the throttle. Similarly the cyclic
20
pitch controls can be assigned to the two pedals and would
0.1
10 influence the orientation of the vehicle.
0 0 Figures 18 and 19 shows the collective pitch requirement
0 50 100 150
True Airspeed (in m/s)
for both the rotors. The collective pitch requirement shows an
initial decrease which would mean a pull back of the collec-
tive stick control at lower mast angles for increasing speed.
Fig. 15. Non-dimensional inflow on upper rotor
As the mast tilts forward, the control stick is pushed to gain
airspeed. For higher mast angles, the collective pitch require-
ment decreases relatively.
Trim Control Angles
Fig. 20, depicts the lateral cyclic pitch control angle. The
The proposed vehicle design has four control parameters, two lateral cyclic pitch required to trim the aircraft as the nacelle
collective pitch control, one for each rotor and two cyclic pitch tilts forward decreases initially, followed by a sharp before
control, common for both rotors. These four control param- evening out to nearly shows. This follows intuition as the mast
eters are depicted in Figures 18, 19, 20, and 21. To control would not require to tilt laterally to maintain flight in airplane
the vehicle, the pilot must be able to influence all the control mode (with wings). The lateral cyclic pitch control shows
parameters at the same time. Since, there are four control pa- huge variations during transition to balance the roll moments
rameters, four control columns are required. Let us assume and side forces generated.
that the vehicle has two control sticks located in front of the Fig. 21, depicts the longitudinal cyclic pitch control an-
pilot and two pedal controls activated by the pilot’s feet. Two gle. The longitudinal cyclic pitch required to trim the aircraft
8
as the nacelle tilts forward tends to be negative, which would
imply a tilt opposite to the direction of motion. This results in 4 90
apparent negative speed stability while converting (aft stick 80
requirement for increasing airspeed). There is a sharp de- 2 70
crease in the rate of change of longitudinal cyclic pitch. The

(in degrees)
longitudinal cyclic pitch eventually tends to even out to nearly 60
zero as the rotor mast tilts close to 90◦ . This follows intuition 0
50
as the mast would not require to tilt laterally to maintain flight
40
in airplane mode (with wings). −2

1c
30

θ
50 90 −4 20

80 10
40
70 −6 0
0 50 100 150
θ0 (in degrees)

30 60 True Airspeed (in m/s)

50
20 Fig. 20. Lateral cyclic pitch of both rotors vs. true airspeed
40
0 90
u

10 30

20 −5 80
0
10 −10 70
θ1s (in degrees)
−10 0 −15 60
0 50 100 150
−20 50
True Airspeed (in m/s)
−25 40
Fig. 18. Collective pitch of upper rotor vs. true airspeed −30 30

−35 20

50 90 −40 10

80 −45 0
40 0 50 100 150
70 True Airspeed (in m/s)
θ0 (in degrees)

30 60
Fig. 21. Longitudinal cyclic pitch of both rotors vs. true
50
20 airspeed
40
speed and the mast tilts further forward, the wing starts con-
l

10 30 tributing to the lift and the rotors start acting as propellers to


20 provide thrust in forward flight. Also, the pitch requirement
0 decreases as the wing generates higher lift for smaller pitch
10 angles at higher speeds. Thus, the sharp decrease in pitch.
−10 0 The roll requirement, shown in Fig. 23.
0 50 100 150
True Airspeed (in m/s)
Power in Transition
Fig. 19. Collective pitch of lower rotor vs. true airspeed
The power consumption over the transition range is shown in
Fig. 24. It can be noticed that the power consumption de-
Attitude Parameters creases as the rotor mast starts tilting forward upto 15◦ de-
noted by the green region. The transition from green to cyan
The pitch requirement, shown in Fig. 22, shows an increasing to denotes an increase in the rate of power consumption. This
pitch requirement in the early stages of the transition which rate increases upto 40◦ followed by a decrease in the rate of
later evens out. At all points of transition, the pitch attitude power consumption denoted by the region in blue to black.
is decreasing with increasing airspeed. The increasing pitch
requirements with increasing mast angle for smaller mast an- CONCLUSIONS
gles can be attributed to lesser contribution of lift in providing
lift at lesser speeds. Thus, the vehicle pitches up to allow the This paper discusses the development of the equations of mo-
rotors to continue providing lift. As the vehicle gains forward tion for coaxial tiltrotor aircraft covering airplane mode, heli-
9
25 90 3000 90
80 80
20 2500
70 70
Pitch (in degrees)

15 60 2000 60

Power (in kW)


50 50
10 1500
40 40
5 30 1000 30
20 20
0 500
10 10
−5 0 0 0
0 50 100 150 0 50 100 150
True Airspeed (in m/s) True Airspeed (in m/s)

Fig. 22. Pitch attitude vs. true airspeed Fig. 24. Power consumption vs. true airspeed
2. The maximum range of the vehicle increases by 100%
0.5 90
in airplane mode than in helicopter mode (w/o wings).
80 Up to 400% increase is predicted in range when flying
70 in airplane mode at 100 m/s (360 km/hr) as compared to
0
flying in helicopter mode (w/o wings) at the same speed.
Roll (in degrees)

60

50 3. The maximum endurance of the vehicle is predicted to


−0.5 exceed 4.5 hours and the vehicle can fly up to two times
40
faster for the same endurance. A maximum possible
30 increase of 450% in endurance is predicted in airplane
−1 20 mode compared to the highest possible speed in heli-
copter mode (w/o wings).
10

−1.5 0 4. A maximum of 500% increase is predicted rate of climb


0 50 100 150 when the vehicle is flying in airplane mode at an airspeed
True Airspeed (in m/s) of 100 m/s (360 km/hr) as compared to flying in heli-
copter mode (w/o wings) at the same speed.
Fig. 23. Roll attitude vs. true airspeed
5. The analysis demonstrates the capability of the coaxial
copter mode, and conversion mode flight. Coaxial rotor trim tiltrotor of transitioning from helicopter mode to aircraft
and performance analysis is developed and refined for tiltrotor mode at wide range of speeds. The simulated vehicle
analysis. Subsequent analysis addresses the performance and could transition from helicopter mode to aircraft mode
transition aspects of the coaxial tiltrotor aircraft from the per- for all the airspeeds between 0 m/s to 90 m/s.
spective of a quasi-steady trim. Qualitative analysis demon-
strated substantial benefit of coaxial tiltrotor design if it were 6. The proposed half tiltwing design is expecter to require
to be technically realized. The coaxial tilt rotor was modelled smaller actuators with reduced power requirements when
using rigid blades with root spring and hinge offset. Rotor in- compared to a full tiltwing while giving advantage of re-
flow was modelled using Drees’ model. Coupled rotor trim duced download penalty.
analysis was carried out using Newton-Raphson to perform
performance and transition analysis. In general, the proposed
coaxial tiltrotor design offers significantly improved perfor- APPENDIX A
mance in the aircraft mode when compared with a coaxial he-
Governing Equations
licopter of same disc loading. The following specific conclu-
sions can be drawn from this conceptual study:
First, the equations of motion are developed. These equations
are then modified for an identical counter-rotating coaxial ro-
1. The maximum cruise speed is predicted to be 100% more tor system. The two set of rotor equations provided the coax-
in airplane mode than in helicopter mode. Up to 75% ial rotor loads which are then added to the airframe loads for
reduction in power requirements is anticipated while fly- trim analysis. The transition analysis is carried out in quasi-
ing in airplane mode compared to helicopter mode (w/o steady manner to demonstrate the feasibility of the vehicle to
wings) at higher flight speeds. complete successful transition.
10
Axes Systems

In order to define and transfer forces, moments, and motions


for all the moving parts of the vehicle, several sets of coor-
dinate systems need to be defined. The coordinate systems
and the corresponding transformation matrices are discussed
in the following subsections.

Gravity Coordinate System

The gravity axes system is shown in Fig. 25. It is fixed


to the aircraft centre of gravity when it is stationary on the
ground. The z-axis points vertically downwards in the direc-
tion of gravity or the center of the Earth. The x-axis points
North and the y-axis points East. Fig. 26. Body-fixed coordinate system

The nacelle axes system shown in Fig. 27 is defined


with respect to the body axes system with the help of mast
angle(βM ). When βM = 0◦ , the vehicle is in helicopter mode.
In this configuration, the x-axis runs parallel to the x-axis of
the body frame, the y-axis runs out the right wing, and the z-
axis is perpendicular to the x and y axes directed downward.
This system is centered at the nacelle axis of rotation. It is
fixed to the nacelle and rotates with it about the y-axis or na-
celle axis of rotation).

Fig. 25. Gravity coordinate system

Body Coordinate System

The body axes system is defined with respect to the grav-


ity axes system with the help of euler angles which determine
the attitude of the aircraft. This system is fixed to the air-
craft at the CG and rotates with it.As shown in Fig. 26 the
x-axis is directed out of the nose of the aircraft and is paral-
lel to the longitudinal axis, the y-axis is directed out of the
right wing and the z-axis, perpendicular to the x and y axes, Fig. 27. Nacelle-fixed coordinate system
directed downward. The roll, pitch and yaw then refers to the
rotation about x-axis, y-axis and z-axis, respectively. TG→B
denotes the rotation matrix defined by the Euler Angles.
    
 îB  cos βM 0 − sin βM  îNAC 
    jˆB =  0 1 0  jˆNAC (8)
 îB     îG  k̂B
 
sin βM 0 cos βM

k̂NAC

jˆB = TG→B jˆG (7)
k̂B k̂G
   
Non-rotating Hub Coordinate System
where
The non-rotating hub axes system is centered at the rotor

cθ cψ cθ sψ −sθ
 hub as shown in Fig. 28. In helicopter mode, i.e. βM = 0◦ , the
TG→B = sφ sθ cψ − cφ sψ sφ sθ sψ + cφ cψ sφ cθ  x-axis runs parallel to the x-axis of the body frame, however
cφ sθ cψ + sφ sψ cφ sθ sψ − sφ cψ cφ cθ it is directed aft. The y-axis runs parallel to the body y-axis
and the z-axis is directed upwards. The non-rotating hub axis
system is fixed to the hub, therefore when the nacelle rotates,
Nacelle Coordinate System the axis rotates with it.
11
Fig. 28. Non-rotating hub-fixed coordinate system Fig. 29. Rotating blade-fixed coordinate system
    
 îNAC  −1 0 0  îNR  parallel to the blade, directed out of the blade. When βk = 0,
jˆNAC =  0 1 0  jˆNR (9) the z-axis is directed downward and the y-axis is then perpen-
k̂NAC 0 0 −1 k̂NR
   
dicular to the x and z axes, directed opposite to the direction
of motion.
    
 îB  − cos βM 0 sin βM  îNR 
jˆB =  0 1 0  jˆNR (10)
 iˆ 
    ′
 îROT  cos βk − sin βk 0
k̂B − sin βM 0 − cos βM k̂NR
   
jˆROT =  0 1 0  ĵ′ (13)
 ˆ′ 
k̂ROT sin βk 0 cos βk
 
k
Rotating Hub Coordinate System
Momentum Theory for Coaxial Rotors
The rotating hub axes system, shown in Fig. 29 is fixed to
the center of the rotor and rotates with it. It is defined with The Momentum Theory or Simple Momentum Theory pro-
respect to the non-rotating hub axes system with the help of vides the most parsimonious flow model to study the hovering
the azimuthal angle of rotation of the kth blade (ψk ) . When performance of a coaxial rotor system. The main advantage
ψk = 0, the rotor blade passes over the tail of the aircraft. of the Simple Momentum Theory, however, is that it provides
At this point, the x-axis is parallel to the x-axis of the body results for the minimum power consumption of the rotors un-
however, it is directed aft. The y-axis runs parallel to the body der a given set of conditions, and so provides a datum against
y-axis. The z-axis is perpendicular to the x and y axes and is which the efficiency of a real rotor can be measured.
directed upward.
There are four primary ways of modelling the inflow of
coaxial rotors:
    
 îNR  cos ψk − sin ψk 0  îROT 
jˆNR =  sin ψk cos ψk 0 jˆROT (11) 1. The two rotors rotating in the same plane (practically,
k̂NR 0 0 1 k̂ROT they would be very near to each other with minimal rotor
   
spacing) and operated at equal thrusts.
  
 îNAC  − cos ψk sin ψk

0  îROT 
 2. The two rotors in the same plane of rotation and operated
jˆNAC =  sin ψk cos ψk 0  jˆROT (12) at equal and opposite torques, i.e., at a torque balance.
k̂NAC 0 0 −1 k̂ROT
   
3. The two rotors rotating at equal thrusts, with the lower
rotor operating in the fully developed slipstream of the
Blade Coordinate System The blade axes system, denoted upper rotor.

by a prime, flaps with the blade. It is centered at the blade 4. Two rotors rotating at equal and opposite torques, with
hinge point and defined with respect to the rotating hub axes the lower rotor in the fully developed slipstream (i.e., the
system by the flap angle of the kth blade (βk ). The x-axis runs vena contracta) of the upper rotor.
12
where
 
− cos βM 0 sin βM
TNR→B =  0 1 0 
− sin βM 0 − cos βM

The velocity vector V~NR can be written as

V~NR = VNRx îNR + VNRy jˆNR + VNRz k̂NR (16)

We now define the non dimensional forward speed or ad-


vance ratio on the rotor disk as
VNRx
µ= (17)
ΩR

Inflow on Upper Rotor

The inflow on the upper rotor consists of two velocities,


i.e., inflow due to vehicle velocity and inflow due to induced
velocity. Following the momentum theory in hover, let us as-
sume that the induced velocity at the upper rotor disk is νu ,
and in the far wake, wu = 2νu . The induced velocities are di-
rected parallel and opposite to the direction of the thrust vec-
tor, Tu . We define the non dimensional induced velocity on
Fig. 30. Flow model showing the coaxial rotor system the upper rotor as
νu
with lower rotor in the vena contracta of the upper ro- λiu = (18)
tor (Ref. 14) ΩR
Thus, the total non-dimensional inflow on the upper rotor
These cases are discussed in detail along with deriva- can be written as
tions of the induced power requirements and thrust sharing VNRz νu
λu = + (19)
in (Refs. 14, 15). Practically, there is always a finite spacing ΩR ΩR
between the two rotors in a coaxial system to avoid blade col-
lisions between the two rotors. Also, the two rotors will gen- Mass flow rate through the upper rotor is given by
erally operate at equal and opposite torques to provide zero q
net torque to the helicopter when it is operating in steady flight ṁu = ρ A (λu ΩR)2 + µ ΩR)2 (20)
conditions. Therefore, for all the cases mentioned above, Case
4 is of primary practical importance, and is employed in the By momentum conservation
present analysis to calculate inflow through coaxial rotors. q
˙ u 2νu = 2ρ A(ΩR)2λi
Tu = (m) λu2 + µ 2 (21)
Figure 30 shows the flow field of a coaxial rotor with the u

lower rotor operating in the vena contracta of the upper rotor.


Thrusts shared by upper and lower rotors are represented by Non-dimensionalzing thrust on the upper rotor, the thrust
Tu and Tℓ , respectively. Because, the upper rotor is not directly coefficient is given by
affected by the wake of the lower rotor, it can be treated as a q
single rotor that is connected to the lower rotor only through CTu = 2λiu λu2 + µ 2 (22)
the need for a torque balance.
For a vehicle operating at forward speed V at a certain Rearranging equation 22, the induced inflow on the upper
flight path angle(θFP = 0 in this case) from the horizontal rotor can be written as,
plane in gravity axes system, the velocity can be written as
CT
λiu = p u (23)
V~G = −V cos θFP îG + V sin θFP k̂G (14) 2 λu2 + µ 2

Using equation 23 and equation 19, the inflow equation for


Using equations 7 and 10, we can write the flow velocity the upper rotor can be written as
on the rotor disk as
VNRz CT
−1  λu = + p u (24)
V~NR = TNR→B TG→B V~G
 
(15) ΩR 2 λu2 + µ 2
13
Inflow on Lower Rotor Thus, the inflow equation on the lower rotor then becomes

The lower rotor is identical to the upper rotor rotating in VNRz


the opposite direction. The wake from the upper rotor now λℓ = λiℓ + λiu + (32)
ΩR
strikes the lower rotor parallel and opposite to the thrust vec-
tor Tℓ with velocity wu . At infinity, wu = 2νu and the wake
A Blade Element Theory
area contracts from A to . Since, the lower rotor is not at
2
A A detailed and systematic derivation of blade element theory
infinity, we assume a contraction factor fc = , where Ac
Ac is discussed in (Refs. 14, 16). The rotor blade is idealized as
is the area of the wake from the upper rotor that strikes the a rigid beam element undergoing only flapping motion. The
lower rotor. The inflow field on the lower rotor now consists position vector of any point P on the kth blade, of an Nb bladed
of two regions, i.e., the inner inflow field of area Ac which ex- rotor system, in the deformed state can be written as
periences wake from the upper rotor and the outer inflow field
which is unaffected by the wake of the upper rotor.
r~p = riˆ′ = r cos βk îROT + r sin βk k̂ROT (33)
From continuity, the velocity of the wake of the upper rotor
is fc νu . We assume that the induced inflow on the lower rotor
is νℓ . The angular velocity of the rotor blade is taken as ω ~ =
Ωk̂ROT . It is assumed that the angular velocity of the rotor is a
The non dimensional inflow on the inner region can be d~ω
written as constant, and hence, the angular acceleration = 0.
VNRz νℓ fc νu dt
λℓinner = + + (25) The absolute velocity of point P has two components: one
ΩR ΩR ΩR
due to rotation and another due to the flapping motion in the
and on the outer region can be written as rotation frame. The absolute velocity of point P is given by
VNRz νℓ
λℓouter = + (26)
ΩR ΩR
 
d~
rp
v~p = ~ × ~r p
+ω (34)
dt rel
The non dimensional induced velocity on the lower rotor
can be defined as
ν
λiℓ = ℓ (27)
ΩR d βk
v~p = −r sin βk îROT
Total Mass flow through the lower rotor can then be written dt
as + Ωr cos βk jˆROT (35)
q
2 d βk
ṁℓ = ρ Ac (λℓouter ΩR)2 + VNR + r cos βk k̂ROT
q
x
(28) dt
2
+ ρ (A − Ac) (λℓinner ΩR)2 + VNRx
The absolute acceleration of point P is given by
By conservation of momentum, the thrust of the lower ro-
tor can be expressed as,
d 2 r~p
 
Tℓ = ṁℓ wℓ − ρ Ac( fc νu )( fc νu ) (29) a~p =
dt 2 rel
 
ω
d~ d~
rp (36)
where wℓ is the velocity of the wake of the lower rotor far + ω×
× ~r p + 2~
downstream. The work done by the lower rotor can be written dt dt rel
as ~ × (~
+ω ω × ~r p )
ṁw2ℓ ρ Ac ( fc νu )( fc νu )2
Tℓ (νu + νℓ ) = − (30)
2 2
Eliminating wℓ from equation 29 and 30, and substituting 
d βk
2
ṁℓ from eq. 28, we get the following expression for induced a~p = −r cos βk
dt
inflow on the lower rotor.
d 2 βk

−r sin βk 2 − Ω2r cos βk îROT
  dt
  (37)
d βk ˆ
1  fc (CTℓ + fc λi2u )2  + −2Ωr sin βk jROT
λiℓ = − fc2 λi3u  dt
 
p
2CTℓ  (λℓ + ( fc − 1)λiu )2 + µ 2  (31) !
d βk 2 d 2 βk
p  
+ (λℓ − λiu ) + µ 2 2
+ −r sin βk + r cos βk 2 k̂ROT
dt dt
− λiu
14
Flapping Equation of Motion
~h = µ ΩR cos ψk îROT
V
The flapping equation of motion for a blade is derived by mak-
− µ ΩR sin ψk jˆROT (44)
ing the total moment about the flap hinge at the root equal to
zero. − λ ΩRk̂ROT

Resolving these velocity components in the blade-fixed


(M̄I ) f lap + (M̄ext ) f lap = 0 (38)
system, we have
The inertia moment about the root is given as ~h = (µ ΩR cos ψk cos βk − λ ΩR sin βk )iˆ′
V
Z R − µ ΩR sin ψk ĵ′ (45)
M̄I = r~p × (−ρ dr~
a p) (39)
0 − (µ ΩR cos ψk sin βk + λ ΩR cos βk )kˆ′

Integrating the term over the length of the blade, the inertia Invoking a small angle assumption, equation 45 can be
moment in flap motion can be expressed as written as

d 2 βk ~h = (µ ΩR cos ψk − λ ΩRβk )iˆ′


V
(M̄I ) f lap = Ib 2
 dt  (40) − µ ΩR sin ψk ĵ′ (46)
Sβ kβ
+ 1+e + Ω2 Ib sin βk cos βk
Ib Ib Ω2 − (µ ΩR cos ψk βk + λ ΩR)k̂′

where Ib = 0R ρ r2 dr is the mass moment of inertia of the


R The relative air velocity at the blade cross section at radial
blade about the flap hinge at the center of the hub. station r from the hub center is given as

Invoking small angle assumption for flap angle βk , the in- ~Vrel = V
~h − v~p (47)
ertia moment in the flap can be simplified as
Substituting equations 43 and 46, we get

d 2 βk ~Vrel = (µ ΩR cos ψk − λ ΩRβk )iˆ′


 
Sβ kβ
(M̄I ) f lap = Ib 2 + 1 + e + Ω2 Ib βk (41)
dt Ib Ib Ω2 − (µ ΩR sin ψk + Ωr) ĵ′ (48)

− (µ ΩR cos ψk βk + λ ΩR − Ωrβk)kˆ′
 
Sβ kβ
We define, 1 + e + = νβ2 , which allows the in-
Ib Ib Ω2
ertia moment to be expressed as The velocity components can then be expressed as
Tangential velocity component at any radial location r:
d 2 βk
(M̄I ) f lap = Ib + νβ2 Ω2 Ib βk (42) r
dt 2

UT = Ωr + µ ΩR sin ψk = ΩR + µ sin ψk (49)
R
The reduced expression, assuming βk to be small, for the
velocity of point P, in the primed blade axes system, can be Radial velocity component along the blade:
written as
UR = µ ΩR cos ψk − λ ΩRβk (50)
d βk ˆ′
v~p = Ωr ĵ′ + r k = Ωr ĵ′ + rβ˙k kˆ′ (43)
dt The normal velocity component at r:

For the sake of consistency and convenience, the time UP = λ ΩR + Ωrβk + βk µ ΩR cos ψk (51)
∗ d βk
derivative of flap βk is non-dimensionalized as βk = Ω =
dΩt The resultant velocity of the oncoming flow can be written,
d βk
Ω = Ωβ˙k where ψ is denoted as nondimensional time Ωt. by assuming UP < UT , as

q
We follow from the equations derived in earlier, that the U = UT2 + UP2 ≈ UT (52)
flow velocity over the rotor disk consists of a horizontal com-
ponent or advance ratio(µ ), and a normal component or inflow
The inflow angle is given by
ratio(λ ). The relative air velocity at the blade section due to
the motion of the helicopter and the total induced flow can be UP
written as components along the kth blade axes system as tan φ = (53)
UT
15
and for small angles The velocity components can be written in non dimen-
sional form(from equations 49 and 51) as
UP
φ≈ (54)
UT UT r
uT = = + µ sin ψk (68)
ΩR R
The sectional aerodynamic lift and drag acting on the air-
foil are given as UP r ∗
uP = = λ + βk + βk µ cos ψk (69)
1 ΩR R
L = ρ U 2CCl (55)
2 UP
Substituting Cl = aαe and αe = θ − φ = θ − in equa-
UT
1 tions 52 and 55, the aerodynamic lift per unit span of the blade
D = ρ U 2CCd (56)
2 can be written as
Resolving these forces along the normal and in-plane di- The blade forces per unit span of the blade are
rections of the blade fixed frame, 1
FiROT = − ρ Ca(ΩR)2 [u2T θ − uPuT ]βk (70)
2
Fk′ = L cos φ − D sin φ (57) 1 Cd
FjROT = ρ Ca(ΩR)2 [uT uP θ − u2P + u2T ] (71)
2 a
Fj′ = −(Lsinφ + D cos φ ) (58) 1
FkROT = ρ Ca(ΩR)2 [u2T θ − uPuT ] (72)
2
The components of these distributed aerodynamic loads in
the rotating blade frame, after neglecting radial drag effects, The external flap moment due to the distributed aerody-
can be given as namic lift actig on the blade can be written, using equations
57 and 65, as

FiROT = −Fk′ sin βk (59) Z R


(Mext ) f lap = (riˆ′ × Fk′ k̂′ )dr
e
Z R (73)
FjROT = Fj′ (60)
= (−(r − e)Fk′ )dr
e

FkROT = Fk′ cos βk (61) Using equations 62, 38 and 41, we get the flap equation as
 2  ZR
d βk 2 2
Invoking small angle approximation and assuming L>D, Ib + νβ Ω β k = (r − e)FkROT dr (74)
dt 2 e
the aerodynamic force components can be approximated as
where νβ is the rotating natural frequency of flap dynamics
FkROT ≈ Fk′ = L (62) Non-dimensionalising the time derivative term on the LHS
d 2 βk ∗∗
2 β and the integral on the RHS with respect to
as = Ω k
dt 2
r
FjROT ≈ −(Lφ + D) (63) the rotor radius R as x = . The flap equation can be written
R
in symbolic form as
FiROT = −Fi′ βk (64) ∗∗
βk + νβ2 βk = γ M̄ f lap (75)
The aerodynamic root moment can be obtained as
ρ aCR4
Here, γ = , the aerodynamic flap moment is then,
~ A =~r × ~F = (r cos βk îROT + r sin βk k̂ROT ) × ~F
M (65) Ib

1 R Z
In component form, M̄ f lap = (r − e)FkROT
ρ aCΩ2 R4 e
~ A = −FjROT r sin βk îROT Z 1 (76)
M 1 e
= (x − )(u2T θ − uPuT )dx
+ (FiROT r sin βk − r cos βk FkROT ) jˆROT (66) 2 Re R
+ r cos βk FjROT k̂ROT This is a simplified flap equation for a blade with an hinge
offset (e) and a flap spring of stiffness kβ . This equation
The blade pitch angle consists of the pilot input and a linear is solved numerically using Newmark’s Algorithm to obtain
geometric twist of the blade a steady-state solution while assuming that the aerodynamic
loads lag the flap response. The solution obtained is then used
θ = θo + θ1c cos ψk + θ1s sin ψk (67) to calculate rotor hub forces.
16
Drees Inflow
1
Z R dF
jROT
Cs j′ = 2
dr
During the transition from hover into level forward flight, that ρ A(ΩR) e dr
is, within the range 0.0 ≤ µ ≤ 0.1, the induced velocity in (80)
aC 1 Cd
Z
the plane of the rotor is the most uniform, it being strongly = (uP uT θ − u2P + u2T )dx
2π R Re a
affected by the presence of discrete tip vortices that sweep
downstream near the rotor plane. Following Glauert’s result
for longitudinal inflow in high speed forward flight and con- R dF
1
Z
kROT
sidering a longitudinal and a lateral variation in the inflow, the Csk′ = 2
dr
induced inflow ratio can be written as ρ A(ΩR) e dr
(81)
aC 1 2
Z
= (u θ − uP uT )dx
2π R Re T
λi = λo (1 + kx r cos ψ + ky r sin ψ ) (77)

Here kx and ky can be viewed as weighting factors and rep- 1


Z R
dMiROT
resent the deviation of the inflow from the uniform value pre- Cni′ = dr (82)
ρ A(ΩR)2 e dr
dicted by the simple momentum theory. A parsimonious lin-
ear inflow model frequently employed in basic rotor analysis
is attributed to Drees(1949). In this model, the coefficients of R
1 dFkROT
Z
the linear part of the inflow are given by : Cn j′ = (r − e) dr
ρ A(ΩR) e2 dr
(83)
aC 1 e
Z
1 − cos χ − 1.8µ 2 (1 − )(u2T θ − uPuT )dx
 
4 =
kx = and ky = −2 µ (78) 2π R R e R
3 sin χ
 
µx R dFjROT
tan−1 1
Z
where the wake skew angle or χ = . Cnk′ = (r − e) dr
µz + λ i ρ A(ΩR)2 e dr
(84)
aC 1 e Cd
Z
Rotor Load Calculations = (1 − )(uP uT θ − u2P + u2T )dx
2π R Re R a

The distribution of aerodynamic and centrifugal forces along Root Loads in Rotating Hub Axes System The root loads in
the span, and the structural dynamics of the blade in response rotating hub axes system are obtained by simply transfering
to these forces create shear loads and bending loads at the the root loads to the hub. For a non-zero offset, the rotating
blade root. hub loads are:
For a zero hinge offset, the blade root is at the center of
rotation. For a non-zero hinge offset, it is at a distance e out-
board from the center of rotation. By ’loads’ we mean ’reac- CFiROT = Csi′ CMiROT = Cni′ (85)
tion’ forces generated by the net balance of all forces acting CFjROT = Cs j′ CM jROT = Cs j′ + eCsk′ (86)
over the blade span. Let si′ , s j′ , and sk′ be the three shear
loads, in-plane, radial, and vertical. Let ni′ , n j′ , and nk′ be the CFk = Csk′ CMk = −Csk′ − eCs j′ (87)
ROT ROT
bending loads, flap bending moment, torsion moment (posi-
tive for leading edge up), and chord bending moment (positive The rotating frame hub loads for a counter rotating rotor
in lag direction). They occur at the blade root, rotate with the can be written as
blade, and vary with the azimuth angle. Thus they are called
the rotating root loads or root reactions. Since the forces and
moments are calculated by integrating over a complete rota- CFiROT = Csi′ CMiROT = Cni′ (88)
tion, the centrifugal and trigonometric terms that integrate to
zero over the limits 0 to 2π are ignored in the formulation. CFjROT = −Cs j′ CM jROT = Cs j′ + eCsk′ (89)
The root forces and moments are calculated in the non dimen- CFk = Csk′ CMk = Csk′ + eCs j′ (90)
ROT ROT
sional form in the blade axes system are given as follows:
Loads in Non-rotating Hub Axes System Root loads in the
non-rotating hub axes system can be obtained by simply re-
1 R dF
Z
i solving them in the NR frame for each blade and adding them
Csi′ = − ROT dr
ρ A(ΩR)2 e dr together. Let m = 1, 2, ...Nb be the blade number. ψm be the
(79)
aC 1 azimuthal location of each blade m. The non-dimensional
Z
= −βk (u2T θ − uPuT )dx
2π R Re loads can be expressed as,
17
where ~ru denote the position vector of the hub centre of
Nb rotor 1 or upper rotor. Similiarly, for lower rotor, the moments
CFiNR = ∑ (CFiROT cos ψm + CFjROT sin ψm ) (91) can be written as
m=1
Nb
~ B = TNR→B M~ NR +~rℓ × ~FB
 
CFjNR = ∑ (CFiROT sin ψm − CFjROT cos ψm ) (92) M ℓ ℓ ℓ
(108)
m=1
Nb Airframe Loads and Moments
CFkNR = ∑ CFkROT (93)
m=1
The trim equations for the aircraft depend on the forces and
Nb
moments of each component. The proprotor contributions are
CMiNR = ∑ (CMiROT cos ψm + CM jROT sin ψm ) (94)
discussed earlier. The methodology used to calculate the air-
m=1
Nb frame forces and moments is to determine the lift and drag of
CM jNR = ∑ (CMiROT sin ψm − CM jROT cos ψm ) (95) each component, and then, using the relative position on the
m=1 aircraft, determine the associated body axis forces and mo-
Nb ments. By definition :
CMkNR = ∑ CMkROT (96)
m=1 1
• Dynamic Pressure: q = ρ V 2
(97) 2
1
The assumption here is that all blades have identical root • Lift: L = ρ V 2 ACL
2
loads, only shifted in phase. In case the blades are similar,
1
the hub loads transmit all the harmonics. Such is the case of • Drag: D = ρ V 2 ACD
damaged or dissimilar rotors. The periodically varying loads 2
over one complete rotation of the rotor blade are The airframe components used for this analysis were the
wing, fuselage, horizontal tail, and vertical tail. For the pur-
pose of this analysis, the lift and drag from the nacelles were
Rotor Drag or H = ρ A(ΩR)2 × CFiNR (98) ignored. Assumptions made to develop these equations are
2 listed below:
Side Force or Y = ρ A(ΩR) × CFjNR (99)
Thrust or T = ρ A(ΩR)2 × CFkNR (100) • CL =
L
and CD =
D
.
qA qA
Roll Moment or Mx = ρ A(ΩR)3 × CMiNR (101)
• CL is linear, therefore, Cℓ = CLα α .
Pitch Moment or My = ρ A(ΩR)3 × CM jNR (102)
Torque or Q = ρ A(ΩR)3 × CMkNR (103) • CM is linear, therefore, CM = CMα α + CM0 .

(104) • Drag Coefficient CD = CD0 + kCℓ2 .


• Wing has a constant airfoil section.
To obtain the steady state components, the loads are aver-
aged out over one complete rotation. This provides us the hub • Proprotor effects on the airflow over the wing and other
loads used to trim the aircraft. aircraft components are negligible.
• The small angle approximation was made for the angle
Loads in Body Axes System In order to get the contribution of of attack and sideslip angles.
these loads in maintaining equilibrium, they need to be trans-
ferred to the body axes system. The forces can be easily trans-
ferred using the rotation matrices in the following way. Wing

FBu = TNR→B ~
~
 
FNRu (105) The equations for lift and drag developed for the wing are
as follows:
Similarly, 1
FBℓ = TNR→B ~
~
 
FNRℓ (106) Lw = ρ V 2 Aw [CLα w (αw − α0L )] (109)
2
The moments in body frame are a sum of the hub moments
1
and moments produced by the hub forces at the CG of the Dw = ρ V 2 AwCDw (110)
aircraft. We can write the moments due to the upper rotor as 2
where
1
~ Bu = TNR→B M
M
 
~ NRu +~ru × ~FBu (107) kw =
π ew ARw
18
The equations for lift and drag developed for the horizontal
CDw = CD0 + kwCL2w
tail are as follows:
1
The wing forces in gravity frame for an airplane at a given LHT = ρ V 2 AHT [CLα HT (αHT − α0L )] (119)
flight path angle of can be written as: 2


−Dw cos θFP − Lw sin θFP
 1
DHT = ρ V 2 A f CDHT (120)
~FGw = 0  (111) 2
−Lw cos θFP + Dw sin θFP where
1
kHT =
And in the body frame as π eHT ARHT
CDHT = CD0 + k f CL2HT
~
FBw = TG→B ~
 
FGw (112)
The horizontal tail forces in gravity frame for an airplane
The moment due to wing forces can be evaluated as at a given flight path angle of can be written as:

~ Bw =~rw × ~FBw
M (113)  
−DHT cos θFP − LHT sin θFP
where~rw is the position vector of the line of aerodynamic ~FGHT = 0  (121)
center of the wing from the aircraft CG. −LHT cos θFP + DHT sin θFP

Fuselage And in the body frame as


~FBHT = TG→B ~FGHT
 
The equations for lift and drag developed for the fuselage (122)
are as follows:
The moment due to horizontal tail forces can be evaluated
1 as
L f = ρ V 2 A f [CLα f (α f − α0L )] (114)
2
~ BHT =~rHT × ~FBHT
M (123)
1
D f = ρ V 2 A f CD f (115) where~rHT is the position vector of the line of aerodynamic
2
center of the horizontal tail from the aircraft CG.
where
1
kf = Vertical Tail
π e f AR f
CD f = CD0 + k f CL2 f The equations for lift and drag developed for the vertical
tail are as follows:
The fuselage forces in gravity frame for an airplane at a 1
given flight path angle of can be written as: LV T = ρ V 2 AV T [CLαV T (αV T − α0L )] (124)
2
 
−D f cos θFP − L f sin θFP 1
~FG = 0 (116) DV T = ρ V 2 A f CDV T (125)
2

f
−L f cos θFP + D f sin θFP where
1
And in the body frame as kV T =
π eV T ARV T
FB f = TG→B ~
~ CDV T = CD0 + k f CL2V T
 
FG f (117)
The vertical tail forces in gravity frame for an airplane at a
The moment due to fuselage forces can be evaluated as given flight path angle of can be written as:
 
~ B f =~r f × ~FB f −DV T cos θFP − LV T sin θFP
M (118) ~FGV T = 0  (126)
−LV T cos θFP + DV T sin θFP
where ~r f is the position vector of the line of aerodynamic
center of the fuselage from the aircraft CG. And in the body frame as
~FBV T = TG→B ~FGV T
 
Horizontal Tail (127)
19
The moment due to vertical tail forces can be evaluated as

Table 4. Fuselage Parameters


~ BV T =~rV T × ~FBVT
M (128)
Parameter Description Value Units
where~rV T is the position vector of the line of aerodynamic Flat Plate Drag, f 1.0 m2
center of the vertical tail from the aircraft CG. Lift Curve Slope, a f 0.286 /rad
Zero lift angle of attack, α0L f -0.14 rad
Equilibrium Equations Zero AOA Moment Coeffficient, CM0 f -0.0096 kg-m
Moment Coeffience vs. AOA, CMα f 1.145 /rad
The equations of equilibrium can be derived by simply adding
all the forces and moments in the body frame. The resultant
forces and moments in the body frame can be written as a sum
of all the components of the aircraft.
Table 5. Wing Parameters
Parameter Description Value Units
~ ~ = 0 (129)
FB = ~FBu + ~FBℓ + ~FBw + ~FB f + ~FBHT + ~FBVT + W Wing Area, Aw 16.0 m2
Span, bw 10 m
Aspect Ratio, AR 6.25 –
Incidence angle, iw 0 rad
~B =M
M ~ Bu + M
~ B +M
~ Bw + M
~ Bf + M ~ BVT = 0 (130)
~ BHT + M Lift Curve Slope, aw 5.31 /rad

Zero-Lift Angle of attack, α0L -0.07 rad
The components of forces and moments in body frame to- Profile Drag Coefficient, Cd0w 0.017 –
gether are the six equilibrium equations which are used for Wing chord, cw 1.6 m
trim analysis using standard Newton-Raphson technique. Wing efficiency factor, ew 0.9 0
Taper Ratio, λ 1 –
APPENDIX B Moment coefficient vs. AOA, CMαw 0 rad

Table 3. Rotor Parameters


Parameter Description Value Units
Table 6. Vertical Tail Parameters
Radius, R 4.0 m Parameter Description Symbol Value Units
Rotor Speed, Ω 589 RPM
Area AV T 2.4 m2
No. of blades, Nb 3 –
Span bV T 2.4 m
Chord, C 0.4 m
Aspect Ratio ARV T 2.4 –
Taper ratio, TR 1 –
Lift Curve Slope aV T 3.06 /rad
Lift Curve Slope, a 5.73 /rad
Zero-Lift Angle of attack α0LV T 0 rad
Profile Drag Coefficient, Cd 0.01 –
Profile Drag Coefficient Cd0V T 0.0017 –
Blade Flapping Inertia, Ib 140 kg − m2
Oswald’s efficiency factor eV T 1 –
Flapping Spring Constant, Kβ 1800 kg-m/rad
Distance from aircraft CG XV T 14.5 m
Hinge Offset, e/R 0.15 –
Twist, θtw -41 deg
Twist at hub, θtw0 40 deg
Rotor(upper) height from CG, h1 2 m
Rotor(lower) height from CG, h2 1.5 m Table 7. Horizontal Tail Parameters
X distance of mast from CG, Xcg 0 m Parameter Description Symbol Value Units
Y distance of mast from CG, Ycg 0 m Area AHT 4.5 m2
Rate of Nacelle Movement, β̇M 0 rad/s Span bHT 4 m
Aspect Ratio AR 3.55 –
Incidence angle iHT 0 rad
Lift Curve Slope aHT 4.03 /rad
Zero-Lift Angle of attack α0L 0 rad
Profile Drag Coefficient Cd0HT 0.0088 –
Efficiency factor eHT 0.8 0
Distance from aircraft CG XHT 14.2 m

20
REFERENCES 15 Syal, M., Contributions to the aerodynamic optimisation of

a coaxial rotor system, Ph.D. thesis, University of Maryland,


1 Maisel, M. D., Giulianetti, D. J., and Dugan, D. C., “The College Park, 2008.
History of the XV-15 Tilt Rotor Research Aircraft: From Con-
16 Venkatesan,C., Fundamentals of Helicopter Dynamics,
cept to Flight,” , 2000.
CRC Press, 2014.
2 Maisel, M., “NASA/ARMY XV-15 Tiltrotor Research Air-

craft Familiarization Document,” Technical report, Ames Re-


search Center, Moffett Field, California, January 1975.
3 Kleinhesselink, K. M., Stability and control modelling of
tiltrotor aircraft, Ph.D. thesis, University of Maryland, Col-
lege Park, 2007.
4 Churchill, G.
B. and Dugan, D. C., “Simulation of the XV-
15 Tilt Rotor Research Aircraft,” Technical report, Ames Re-
search Center, Moffett Field, California, March 1982.
5 Juhasz,O., Flight Dynamic Simulation Modelling And
Control of a Large Flexible Tiltrotor Aircraft, Ph.D. thesis,
University of Maryland, College Park, 2014.
6 Deckert,
W. H. and Ferry, R. G., “Flight Evaluation of the
XV-3,” Technical report, Air Force Flight Test Center, Ed-
wards Air Force Base, California, July 1959.
7
Quigley, H. C. and Koenig, D. G., “A Flight study of
the dynamic stability of a tilting-rotor convertiplane,” Techni-
cal report, Ames Research Center, Moffett Field, California,
April 1961.
8 Johnson, W., Yamauchi, G. K., and Watts, M. E., “NASA

Heavy Lift Rotorcraft Systems Investigation,” Technical re-


port, Ames Research Center, Moffett Field, California, De-
cember 2005.
9
Ferguson, S. W., “Development and Validation of a Sim-
ulation for a generic tiltrotor aircraft,” Technical report, Sys-
tems Technology Inc, Mountain View, California, April 1989.
10
Ferguson, S. W. and Hanson, G., “Generic Tilt-Rotor Sim-
ulation (GTRSIM) User’s and Programmer’s Guide,” Techni-
cal report, Systems Technology Inc, Mountain View, Califor-
nia, October 1983.
11 Praetor,R., Conceptual Design Studies of a Coaxial Mono
Tiltrotor, Ph.D. thesis, University of Maryland, College Park,
2005.
12 Melanson, T., “Advantages of a Future Amphibious
Tiltrotor Craft: An Angel in the Sky,” Technical report,
http://www.aeronautics.nasa.gov/, June 2010.
13 Harrington, R. D., “Full Scale Tunnel investigation of the
static thrust performance of a coaxial helicopter rotor,” Tech-
nical report, Langley Aeronautical Laboratory, Langley Field,
Vancouver, March 1951.
14 Leishman,
J. G., Principles of Helicopter Aerodynamics,
Cambridge University Press, 2006.
21

You might also like