Você está na página 1de 28

Journal of Intelligent & Robotic Systems

Theoretical and experimental investigations on the effect of overlap and offset on the
design of a novel quadrotor configuration, VOOPS
--Manuscript Draft--

Manuscript Number: JINT-D-16-00451

Full Title: Theoretical and experimental investigations on the effect of overlap and offset on the
design of a novel quadrotor configuration, VOOPS

Article Type: Full/Regular paper

Keywords: Quadrotors; Overlapped Propeller; Unmanned Aerial Vehicles; PID control; Vertically
Offset Overlapped Propulsion System; Quadrotor Payload Capacity; Quadrotor
Endurance

Abstract: The theoretical and experimental investigations of a novel configuration for a Vertical
Take Off and Landing (VTOL) quadrotor system with a Vertically Offset Overlapped
Propulsion System (VOOPS) is presented in this paper. The objective of the VOOPS
concept is to improve the load carrying capability of the VTOL system without any
increase in the footprint or reduction in endurance. This has been accomplished by
overlapping the propeller blades with a vertical offset such that they don't intersect.
This way VOOPS can accommodate a larger propeller size without increase in
footprint. The aerodynamics of the overlap and offset on the quadrotor performance
has been analysed using Blade Element and Momentum Theory (BEMT) and a
methodology has been proposed to find their effects on thrust, drag, and power. The
important design constraints for VOOPS are offset and overlap of the propellers,
whose limits are evaluated using the blade-bending profile and the geometry of the
propeller, respectively. The mathematical model and the design of a VOOPS quadrotor
system is also presented. A quadrotor with VOOPS configuration has been designed
and experimental studies have shown that the theoretical results obtained are in good
agreement with the experiments. The practical implementation of VOOPS configuration
and the consequences of the design changes are also presented.

Powered by Editorial Manager and ProduXion Manager from Aries Systems Corporation
covering letter Click here to download attachment to manuscript
CoverLetter.pdf

From
Prof. Asokan Thondiyath
Department of Engineering Design
Indian Institute of Technology Madras

To
The Editor-in Chief
Journal of Intelligent & Robotic Systems

Sub:

Dear Editor,

Please find enclosed our manuscript, Theoretical and experimental investigations on the effect
of overlap and offset on the design of a novel quadrotor configuration, VOOPS which we
would like to submit for publication in the Journal of Intelligent & Robotic Systems.

In this paper we propose a novel design of quadrotor to increase the endurance and payload. The
new design, Vertically Offset Overlapping Propulsion System (VOOPS), enables the use of larger
propellers in a quadrotor, without increasing the form factor. This is achieved by allowing the
propellers to overlap. Theoretical framework for analysis of overlapping propellers is developed
and experimentally verified by conducting bench tests. As a proof of concept, an experimental
quadrotor with VOOPS was designed and experiments were carried out on the same to prove the
effectiveness of the design. We believe that VOOPS design will bring major changes in the way
quadrotors are built, especially in the design of micro aerial robot.

We believe our findings would appeal to the readership of the Journal of Intelligent & Robotic
Systems. We confirm that this manuscript has not been published or being under consideration by
another journal. All authors have approved the manuscript and agree with its submission to the
Journal of Intelligent & Robotic Systems. We will be thankful if you could review the paper and
give your comments. Looking forward to hearing from you at your earliest convenience.

Thank you.
Manuscript Click here to download Manuscript Final Paper IntelRobSys
V4.pdf

Noname manuscript No.


(will be inserted by the editor)

Theoretical and experimental investigations on the


effect of overlap and offset on the design of a novel
quadrotor configuration, VOOPS

Ganeshram Nandakumar Arjun


Srinivasan Asokan Thondiyath

Received: date / Accepted: date

Abstract The theoretical and experimental investigations of a novel config-


uration for a Vertical Take Off and Landing (VTOL) quadrotor system with
a Vertically Offset Overlapped Propulsion System (VOOPS) is presented in
this paper. The objective of the VOOPS concept is to improve the load car-
rying capability of the VTOL system without any increase in the footprint
or reduction in endurance. This has been accomplished by overlapping the
propeller blades with a vertical offset such that they dont intersect. This way
VOOPS can accommodate a larger propeller size without increase in footprint.
The aerodynamics of the overlap and offset on the quadrotor performance has
been analysed using Blade Element and Momentum Theory (BEMT) and
a methodology has been proposed to find their effects on thrust, drag, and
power. The important design constraints for VOOPS are offset and overlap
of the propellers, whose limits are evaluated using the blade-bending pro-
file and the geometry of the propeller, respectively. The mathematical model
and the design of a VOOPS quadrotor system is also presented. A quadrotor
with VOOPS configuration has been designed and experimental studies have
shown that the theoretical results obtained are in good agreement with the
experiments. The practical implementation of VOOPS configuration and the
consequences of the design changes are also presented.

Ganeshram Nandakumar
Department of Engineering Design, Indian Institute of Technology Madras, Chennai 600036,
India
E-mail: ganeshramnandakumar@gmail.com
Arjun Srinivasan
Department of Mechanical Engineering, National Institute of Technology Tiruchirap-
palli,Tiruchirappalli 620015, India
Asokan Thondiyath
Department of Engineering Design, Indian Institute of Technology Madras, Chennai 600036,
India
2 Ganeshram Nandakumar et al.

Keywords Quadrotors Overlapped Propeller Unmanned Aerial Vehicles


PID Control Vertically Offset Overlapped Propulsion System Quadrotor
Payload Capacity

1 Introduction

In recent years, there has been a rapid development in autonomous Unmanned


Aerial Vehicle (UAV) due to easy availability of lighter materials, improved
autonomous control devices, and robust electronics [1]. VTOLs have gained im-
portance and are being applied in the fields of aerial photography, 3D-mapping,
precision-farming, surveillance, etc. [2]. Several configurations of VTOL sys-
tems exists and quadrotor system is found to be the most reliable for small
area landing, take off, and hovering capabilities [3]. Quadrotor systems are
multidisciplinary, requiring attention from researchers in aerospace, mechan-
ical, electrical, computer science, etc. Modelling the dynamics of quadrotor
system and design of controllers were the major research focus in the recent
past [411]. The influence of aerodynamics and proximity effects in quadrotor
flight were presented in [12]. Design of appropriate control laws for improved
handling and flight performance were the focus of research in [1315]. VTOL
systems being small in size, compared to the other type aerial robots, handling
of external disturbances such as wind, and additional payloads like a robotic
arm brings in additional complexities in flight control. Some of these issues
were addressed in [16]. Design modifications of a quadrotor to increase the
flight performance is one of the areas of research which focusses on improved
designs to increase payload, stability, endurance etc. [1722]. A variable pitch
quadrotor was proposed by Michini et al. [17]. They showed that the dynamic
performance of the quadrotor can be improved by dynamically changing the
pitch of the propeller. The effects of tilting the rotors inward were analysed
in a recent work by Efraim et al. [18]. It was shown that the stability of the
quadrotor can be improved through this design modification. Quadrotors with
downward mounted propellers claim better performance as shown in [23], and
several commercial designs are available with this configuration [24]. There
were not many reported works that focus on design modifications for improved
payload and endurance for a given foot-print of the quadrotor. Most of the
missions conducted by a quadrotor system demands better payload carrying
capability and it is an important criteria for designers. Conventional quadrotor
systems have all four rotors placed in one single plane. The thrust generated
by these rotors helps in lifting the quadrotor and its payload. One of the meth-
ods suggested for improving the payload capacity was the decoupling of the
stabilization and the lifting tasks, by using separate propellers for each task
[19, 20]. The design of compact VTOL UAV with an improved endurance and
payload capability is challenging as they are conflicting design requirements.
Attempts to improve endurance by discarding consumed batteries were also
reported [25]. In this paper, we propose a deviation from the conventional ways
of improvement in a quadrotor system through innovative design changes, in
Title Suppressed Due to Excessive Length 3

order to improve two or more of the above mentioned factors. To improve the
payload carrying capability of a VTOL system, either larger propellers or more
number of propellers have to be used. Large propellers improves efficiency and
endurance, but increases the footprint of the VTOL system. As reported in
[22], using more propellers enhances payload capability (50% more for a 6-
rotor system) but also increases the footprint of the VTOL system (the size
increases by 24%) and a 50% increase in the power consumption is observed.
Therefore, larger propellers or more number of propellers in a VTOL system
is not a suitable solution.

Conventional design of quadrotor does not allow a lesser footprint while


using
large propellers, as the size of the system cannot be less than (1 + 2) times
the diameter of the propeller used. Hence, a new design approach, VOOPS,
is proposed to overcome this limitation, wherein the propellers are allowed
to overlap without colliding with each other, which is achieved by placing
them in different planes [26]. In this way, propellers with larger diameter can
be accommodated within smaller footprints, enabling better payload carrying
capability and endurance for the same footprint. The following sections will
analyse VOOPS and show that VOOPS configuration has benefits in payload
and endurance.

2 VOOPS concept

Designing a quadrotor with high payload capacity for a given footprint is an


onerous task and demands novel design approach. Hence, VOOPS was pro-
posed [26], inspired by the tandem rotors configuration found in Piasecki HRP
Rescuer [27]. The main objective here is to accommodate larger propeller by
making design changes in the conventional quadrotor based VTOL system.
The propellers are allowed to overlap by placing them at different planes sep-
arated by a distance in such a way that they do not interfere with each other,
as shown in Figure 1.
VOOPS is designed as a bi-layered frame, with the layers offset by a dis-
tance Z vertically. Each layer accommodates two rotors, U 1 and U 2 on top
layer and L1 and L2 in bottom layer. The crucial factors in the VOOPS con-
figuration are the overlap and offset between the blades. The following sections
explain the methods used to analyse the effect of overlap and offset on thrust,
drag torque, and power consumption in the VOOPS design. A brief account
of the aerodynamics of the propellers is provided in section 3.

3 Aerodynamic analysis

Determination of the aerodynamic forces acting on a propeller involves ana-


lyzing individual elemental sections of the propeller along the span. Based on
4 Ganeshram Nandakumar et al.

(a)

(b)
Fig. 1 VOOPS Configuration (a) Top view (b) Side view

the Blade Element Theory (BET), the forces acting on each elemental sections
are integrated along the span to find the net aerodynamic forces. The main
difficulty in BET modelling is with the estimation of the induced flow created
by the rotating propeller. Since the BEMT addresses the problem by incor-
porating geometric parameters of the propeller blade into a momentum flow
model, it is used here to analyze the effects of offset and overlap.
Title Suppressed Due to Excessive Length 5

3.1 Aerodynamics of single propeller

The thrust T produced by a propeller is given by [28],

T = A(R)2 Ct (1)

where, is the density of air, A is the area of the propeller disk, is the
angular velocity of propeller, R is the radius of the propeller, and Ct is the
coefficient of thrust.
A propeller blade is assumed to be made of a series of airfoil sections from hub
to tip. For a particular section at a distance r from the hub, the solidity ratio
can be given as = nc(r)/R, where n is the number of blades in a propeller,
c(r) is the chord length of the propeller at r, (r) is the pre-twist, otherwise
known as pitch of the blade, and (r) is the inflow ratio (the ratio between
inflow velocity (also known as induced velocity) and the blade tip velocity,
vinduced /R).
Using BET, the coefficient of thrust dCt , for an incremental section dr at a
distance r from the hub is given by
1
dCt = a(r2 (r) r(r))dr (2)
2
where a is the lift curve slope for the propeller, and the inflow ratio, (r)
can be obtained by combining the principles of blade element and momentum
theory [28] and is given as:
r
1 32r(r)
(r) = a(1 + 1 + ) (3)
16 a
The overall coefficient of thrust, Ct , can be obtained by integrating equation
(2) over the length of the blade and averaging the result over one complete
revolution (0,2), as given in equation (4).
Z 2 Z R
1 1
Ct = a(r2 (r) r(r))drd (4)
2 0 0 2
Equations (1), (3), and (4) can be used to calculate the thrust produced by the
propeller in the normal configurations. However, the thrust produced changes
in the VOOPS configuration and this need to be analysed by considering the
effect of offset and overlap on the inflow ratio and the thrust coefficient.

3.2 Effect of overlap and offset on inflow velocity

The inflow ratio varies in the radial direction as shown in equation (3) and also
in vertical direction which will be discussed below. In an overlapped configu-
ration, the inflow velocity (u (r)u R) from the top propeller influences the
bottom propeller and this region of influence varies with the extent of overlap
d and offset Z. The inflow velocity variations due to the overlap and offset can
6 Ganeshram Nandakumar et al.

be estimated with the help of Figure 2. As shown in the figure, the stream tube
envelope contracts in the vertical direction and the inflow velocity distribution
changes. Using the law of mass conservation, the contracted radius Rz of the
envelope at distance Z below the propeller can be obtained as:
R
Rz = r (5)
1+ p Z 2
R (1+ Z 2 )
R

Using Actuator Disk Vortex Theory [29], an analytic expression for the vari-
ation of inflow velocity vz along the axis of the propeller at a distance Z can
be derived as:
Z
vz = vu (1 + q ) (6)
Z2
R (1 + R 2)

where, vu - Induced Velocity in the upper propeller at Z = 0.

Fig. 2 Inflow at the overlapped region

Using equations (4), (5) and (6), we can derive the variation of inflow velocity
distibution and the region of overlap at any offset distance from the propeller
plane.
The total inflow velocity seen at the overlapped region is the summation of
inflow velocity vl , of the lower propeller, and the inflow velocity vz , of the
upper propeller at an offset distance Z.
Hence, the inflow ratio, 0 , at the overlapped region (inflow velocity v0 divided
by the tip velocity (l R) of the lower propeller) can be calculated as:
vl + vz
0 =
l R
u
= l + z (7)
l
Title Suppressed Due to Excessive Length 7

where, l , u are the inflow ratio of the lower and upper propeller respectively,
u , l are the angular velocities of the top and the bottom rotor respectively.
Due to the changes in the inflow ratio over the overlapped region, the thrust
coefficient also varies and as a result, the aerodynamic thrust produced also
changes. An analytical frame work for studying the effect of overlap and offset
on the thrust is developed using the BEMT in the following section. This has
been verified using experimental analysis in section 4. This analysis enables
us to choose optimal overlap and offset (as required by the mission) using
the geometric parameters of the propeller rather than conducting rigorous
experiments.

3.3 Thrust variation due to overlap and offset

The impact of overlap on the thrust coefficient of the lower propeller can be
analysed by considering the three possible cases of overlap as shown in Fig-
ure 3. These cases are derived under the assumption that the offset distance
Z < 2R. The circle L in the Figure 3 represents the bottom propeller of radius
R, circle U is the contracted envelope of inflow with radius Rz from the upper
propeller. The existence of the above cases are based on the extent of overlap
d and the offset distance Z as shown in Figure 3. In Case 1, as shown
q in Figure
3(a), when the overlap distance d varies from 0 to R + Rz (R2 + Rz 2 ),
the circle L will have two distinct regions A and B1 (shaded regions in Figure
3(a)). Region A is unaffected by overlap and Region
q B1 is affected by overlap.
In Case 2, when overlap varies from R + Rz (R2 + Rz 2 ) to R, it leads to
three separate regions shown as A, B1 and B2 as shown in Figure 3(b). In
this case, it is possible that a propeller blade can have two unaffected regions
and one affected region, depending on the angular position of the blade. Here
in the B2 region, the propeller is having a small portion under the influence
of upper propeller. In Case 3, the overlap is very high and varies from R to
R + Rz and the lower propeller could be either fully or partially under the
influence of the upper propeller as shown in Figure 3(c), shown as regions A,
B1 and B3 . In region B3 , the lower propeller is fully under the influence of the
upper one.
In order to calculate the thrust generated by the bottom propeller, it is nec-
essary to estimate the thrust coefficient Ct of the lower propeller in regions A
and B. To analyze this, consider one blade of the rotor at an angular position,
shown as LP in Figure 3(a). In cases 1 and 2, LM is the region close to hub
not affected by the inflow (in case 3 LM does not exist), M N is the region
affected by the inflow from the upper propeller, and N P is again not affected
by upper propeller. The points M and N will vary depending on the overlap,
and the distance LM is defined as X1 and LN as X2 . Thrust produced in the
overlapped region B can be found by integrating thrust produced by the blade
sections M N in the overlapped regions B1 , B2 and B3 . Using equations (4)
and (7), the coefficient of thrust in the overlapped region Bi can be expressed
8 Ganeshram Nandakumar et al.

(a)

(b)

(c)

Fig. 3 Schematic representation of blade overlap (a) Case 1 (b) Case 2 (c) Case 3

as in equation (8), where i can be 1, 2 or 3 based on the part of the overlapped


region in which Ct is calculated and the boundaries of these regions are defined
by the integration limits discussed below.
Z 2 Z X2
1 1
CtBi = (2 l a(r2 l r0 )drd) (8)
2 1 X1 2

Substituting equation (7) in (8) and rearranging the terms will lead to the form
where the thrust loss due to overlap can be separated as shown in equation(9).
Z 2 Z X2 Z 2 Z X2
1 2 1 u
CtBi = l a(r l rl )drd) ( ) l au rdrd
2 1 X1 2 l 1 X1
(9)
Title Suppressed Due to Excessive Length 9

which is of the form :

CtBi = CtBi Actual CtBi loss (10)

where (CtBi Actual ) represents the coefficient of thrust if there was no top rotor
and (CtBi loss ) represents the loss in the coefficient of thrust due to the interac-
tion of the inflow from the top rotor. Once we identify the integration limits in
equation (9), it is possible to calculate the above coefficients. The integration
limits for r and in the overlapped regions can be estimated using standard
trigonometric relationships and are given in Table 1. Here, is the angle at
which the X1 and X2 are equal and the blade becomes a tangent to the inflow
envelope of top rotor. Value of can be estimated as
R
= [sin1 ( )] (11)
d
and the angle (6 QLU , Q is the point of intersection of the two circles L and
U ), can be obtained as:

= cos1 (d2 + R2 Rz2 )/(2Rd) (12)

The distances X1 and X2 can be obtained as a function of the blades angular


position and the extent of overlap d as shown below.
1
X1 = (d + d cos(2)+
2
p
2 d2 cos2 () + 2R2 cos2 () + d2 cos2 () cos(2) sec())
1
X2 = (d + d cos(2)
2
p
2 d2 cos2 () + 2R2 cos2 () + d2 cos2 () cos(2) sec()) (13)
Using equations (11) to (13) and Table 1, it is possible to estimate the Ctloss

Table 1 Overlapped region boundaries

Region Boundary
B1 (0, r), r (X1 , X2 )
B2 (, ), r (X1 , X2 )
B3 (, ), r (X1 , X2 )

at various regions of overlap. Thus the loss in thrust coefficient in the three
cases of overlap can be calculated as:

Case 1: CtB loss = CtB1 loss


Case 2: CtB loss = CtB1 loss + CtB2 loss (14)
Case 3: CtB loss = CtB1 loss + CtB3 loss
10 Ganeshram Nandakumar et al.

3.4 Effect of overlap and offset on rotor power and drag torque

When the propeller is spinning, the drag experienced by every airfoil section
collectively contributes to a torque Q at the center of the propeller, as given
by equation (15). This torque is responsible for the degree of freedom in the
yaw axis. Power P can be directly calculated using equation (16) as:

Q = A 2 R3 Cq (15)

P = Q (16)
where, Cq is the coefficient of torque. Using BET, the incremental coefficient
of torque dCq can be obtained by equating it to the incremental coefficient of
power, dCp , for an incremental section dr at a distance r from the hub. It can
be calculated using the equivalence of dCq and dCp as given in [9] as:

dC q = dC p = dC p0 + dCpi (17)

Here, dC p is a combination of profile power dC p0 given by dC p0 = 12 Cd r3 dr


and the induced power dC pi given by dC pi = (r)dC t , where, Cd is coefficient
of drag (assumed to be a constant). Therefore, the torque coefficient for the
rotor can be obtained by substituting equation (2) in (17) and integrating over
the rotor length and angle of rotation as:
Z Z
1 1 1
Cq = ((r)( a(r2 (r) r(r))) + Cd r3 )drd (18)
2 2 2
The loss in torque and power due to overlap and offset can be obtained by
estimating the loss of Cq in the regions of overlap as discussed in the previous
sections. The integration limits of equation (18) can be appropriately chosen
to arrive at the loss in Cq in overlapped regions in a similar way as discussed
in the previous section and hence the torque and power loss due to overlap
and offset can be estimated using equations (15) and (16).

3.5 Simulation results

Numerical simulations were carried out using the analytical relations derived
in sections 3.3 and 3.4. Propellers of diameter 15 inch were considered for
simulations and the overlap was varied from 0 to 100% (when the overlap is
equal to the propeller radius, it is considered as 100% overlap) with vertical
offset varying from 2 to 5 cm. Figure 4 shows the thrust lost due to overlap as
a percentage of the total thrust. It can be seen that as the overlap increases,
there is a loss of thrust and it increases with the overlap. The losses due to
overlap are less than 5% up to an overlap of 30%, which is in tune with the
rigorous theoretical analysis reported [30]. As seen in the result, the effect of
vertical offset is minimal on the thrust loss. As the offset increases, the thrust
loss increases.
Title Suppressed Due to Excessive Length 11

Fig. 4 Thrust loss at various overlap and offset

Fig. 5 Percentage loss in drag torque

The simulation results for drag loss as a percentage of total drag is given
in Figure 5. It is observed that the drag losses are not very significant and
12 Ganeshram Nandakumar et al.

does not vary much with overlap and offset, except when the overlap is close
to 100% (due to large area exposed to the flow from the propeller on the top).
As the practical design of VOOPS will not be possible for such high over-
laps, we can confidently assume that the drag loss is negligible for overlapped
configuration of propellers. The above analyses and results provide the theoret-
ical framework for the design of VOOPS quadrotor configuration. Experiments
were conducted to verify and compare with the simulation outcome before the
design and testing of VOOPS quadrotor.

4 Experimental verification

A test rig was built to measure the thrust, drag and power of the bottom
propeller at various offsets and overlaps. The bottom propeller is fixed to a
slider which can move on a vertical and horizontal guide for varying the offset
and overlap, and the top propeller was fixed above, as shown in Figure 6. A

Fig. 6 Experimental setup

two-axis load cell (MBA500), which can measure thrust and the drag torque
produced by the rotating blades, is attached to the bottom propeller. The
block digram in Figure 7 shows the instrumentation used for the experiments.
A microcontroller is used to communicate motor speeds from the computer.
An amplifier, Hx711, is used for the load cells, and the data is then transferred
to the computer which then stores the data. Experiments were conducted with
a BLDC motor (Tiger Navigator) with a speed constant of 400 rotations/Volt
as the prime mover, spinning 15 inch propellers. The power supply used was
a 22volt DC at 30amps. The effect of offset and overlap was studied by con-
ducting the experiments by varying offset from 2cm to 5cm with a step size of
Title Suppressed Due to Excessive Length 13

1cm and overlap from 0% to 100% at intervals of 2 cm.

Fig. 7 Experimental setup block diagram

Measurements were taken at a fixed RPM corresponding to standard payload


(Rotor RPM 50% of max RPM) and at an offset of 3cm. Experiments were
also conducted at the decided overlap and offset discussed in section 5.1 for
all possible combinations of RPM of both the rotors. (This data was stored as
look up tables Ti and Hi for thrust and drag respectively, which is used in the
simulations discussed in section 6.2 to compute the external inputs, the RHS
of the dynamic equations).

Figure 8(a) shows the experimental thrust loss for an offset of 3cm. It is found
that there is a good match between the simulation results in Figure 4 and the
experimental results, barring few outliers which may be due to experimental
errors. This confirms that the mathematical frame work developed in the pre-
vious section can effectively model the thrust loss. Figure 8(b) shows that the
drag power remains almost constant from no overlap to full overlap, validating
the mathematical model. The above results confirm that the overlap and off-
set of propellers do not affect the aerodynamics of VTOL systems significantly
and the VOOPS configuration can be used for design of efficient quadrotors.
The design and experimental verification of VOOPS quadrotor is presented in
the following section.

4.1 Design constraints on overlap and offset

The VOOPS configuration discussed here is a quadrotor system with two lay-
ers, where the alternate rotors are kept at different planes which are offset
vertically, enabling propellers to overlap as shown in Figure 9. Rotors U 1 and
U 2 are kept on the top layer and L1 and L2 are kept in the bottom layer
and each rotor is of a diameter 2R. The offset between the propellers can be
14 Ganeshram Nandakumar et al.

(a)

(b)

(c)

Fig. 8 (a) Variation in thrust at an offset of 3 cm, (b) Variation in drag at an offset 3 cm
and (c) Variation in power at an offset 3 cm
Title Suppressed Due to Excessive Length 15

varied from 0 to 2R and the overlap can be varied from 0 to 100%. The entire
range of overlap and offset cannot be realized due to structural and geometri-
cal reasons. The adjacent propellers can be overlapped only till the diagonally
opposite propellers
interfere. It can be geometrically shown that a maximum
overlap of R 2( 2 1) is possible, which is 29.29%.

Fig. 9 Physical constraints in propeller overlap

A loss in thrust (3-4%) is observed as the offset distance between the propeller
planes become larger as discussed in section 3.5. Though minimal offset be-
ing better in terms of efficiency, the blades can bend due to the lift produced
during rotation and might collide with each other. To analyze the minimum
offset to be provided to prevent the blade tip collision due to bending, a flexure
modelling of the blade along with aerodynamics is carried out to model the
tip difflection profile. The distribution of lift on the propeller blade is derived
using BEMT and then the tip difflection profile is determined using Euler-
Bernoulli beam theory as explained below.

Referring to Figure 10, the inflow velocity Vi at each elemental section of the
propeller blade can be evaluated using BEMT using equation (3), by multi-
plying it with l R (tip speed). Bending moment due to this loading is given
as [31]:
1 R
Z
Mb = CL c(r)(Vi2 + Vt2 )cos((r) )dr (19)
2 0
where, Vt is the tangential velocity seen by the blade due to rotation, is the
angle of attack, CL is the coefficient of lift, and r is the radius of the blade. The
propeller blade is considered as an Euler-Bernoulli beam and the relationship
between bending moment and deflection zd , can be represented as:
d2 zd Mb
2
= (20)
dr EI
16 Ganeshram Nandakumar et al.

Fig. 10 Blade cross section

Table 2 Overlapped region boundaries

Parameters Value
Diameter 0.381 m
Maximum angular velocity 680 rad/s
Air density 1.225 kg/m3
Lift coefficient 1.21
Elastic modulus of blade material 1.9 GPa
Moment of inertia about the bending axis 1.442x1012 m4

where, I is the moment of inertia of the propeller blade and E is the elas-
tic modulus of the blade material. Solving the differential equation (20) will
provide the propeller blade bending profile as shown in the Figure 11 and
the maximum deflection of blade along Z can be determined for a particu-
lar propeller. During the flight, the propellers can spin at different Rotations
Per Minute(RPM) and there is a possibility of having considerable difference
between them. Therefore an added factor of safety was implemented and the
offset between the two layers of the VOOPS design was decided to be 3 cm.
The parameters given in Table 2 were used to derive the blade bending pro-
file from which maximum blade deflection was found to be 2.25 cm at the
maximum RPM.

4.2 Design of VOOPS quadrotor system

To evaluate the performance of VOOPS quadrotor system over the commer-


cially available vehicles of the same class, an experimental quadrotor with
VOOPS configuration was designed. This vehicle was planned for executing
missions within 4 km radius with a minimum payload of 500 grams and an
endurance of 40 minutes. To meet these requirements, propeller sizing and
actuator sizing were performed. Propeller size of 15 inches was chosen and
BLDC motors (MN4014 400kv) were selected as actuators. Based on the ex-
Title Suppressed Due to Excessive Length 17

Fig. 11 Propeller bending along the its length

perimental analysis presented in section 4.1, an offset of 3 cm and an overlap


of 7.62 cm (20%) was decided. The tip to tip size of the VOOPS quadrotor
was found to be 81.5 cm with 20% overlap. A conventional quadrotor with
a 14 inch propeller will have a tip to tip size of 85.84 cm, which is 4.34 cm
larger than the VOOPS quadrotor. A conventional quadrotor with 15 inch
propeller will be 10.7 cm larger than VOOPS configuration. The chosen off-
set and overlap in the VOOPS configuration gives the advantage of reduced
size, and can accommodate a larger propeller and hence better performance. A
smaller footprint is always advantageous in situations where the quadrotor has
to navigate through tighter spaces. The side and top view of the experimental
VOOPS quadrotor is shown in Figure 12.

5 Mathematical modelling of VOOPS quadrotor

5.1 Kinematics

To analyse and study the geometry of motion of the VOOPS Quadrotor with
respect to time, a body-fixed coordinate frame {X0 Y0 Z0 } and an earth-fixed
frame {Xe Ye Ze } is assigned.
The pose vector of the quadrotor is represented as =[ T1 , T2 ]T , where 1
is the position vector given by 1 =[x,y,z]T , and 2 is the orientation vector
given by 2 =[, , ]T , where , , are roll, pitch and yaw respectively. V is
the velocity vector represented by v=[vT1 ,vT2 ]T where v1 is the linear velocity
vector given by v1 =[u,v,w]T and v2 is the angular velocity vector given by
v2 =[p,q,r]T .
Velocities in the earth frame are converted to body frame using the following
relationship [15]:

= J( 2 )v (21)
18 Ganeshram Nandakumar et al.

(a)

(b)

Fig. 12 VOOPS quadrotor (a) top view (b) side view

where, J( 2 ) is the Jacobian matrix given by


 
J1 ( 2 ) O33
J( 2 ) = (22)
O33 J2 ( 2 ) 66

Matrices J1 ( 2 ) and J2 ( 2 ) are given by



c c s c + c s s s s + c c s
J1 ( 2 ) = s c c c + s s s c s + s s s (23)
s c s c c
Title Suppressed Due to Excessive Length 19


1 s t c t
J2 ( 2 ) = 0 c s (24)
s c
0 c c

5.2 Dynamics

The behavior of the system under the influence of external forces can be derived
using Newton Euler approach [33], which helps in analysing the dynamics of
VOOPS quadrotor. The equations of motion are given as
= M + C + G (25)
where, M is the inertia matrix, C is a vector of centripetal and coriolis terms
and G is gravitational vector. The external forces and moments acting on the
quadrotor is given by the vector =[f 0 ,m0 ]T , where f 0 is the external force
vector in x, y, and z directions given by [fx , fy , fz ]T and m0 is external mo-
ment vector about x, y and z axes given by [mx , my , mz ]T . Since the thrust
produced in the VOOPS configuration for the bottom propellers are different
from the propellers on the top, their forces and moments are estimated using
the theoretical framework discussed in section 3.3.

To find the solution for equation (25), initial conditions of the system i.e.
the initial position and velocities of the system are required.

The thrust generated by each propeller sums up to the total lifting force of
the quadrotor in its body frame.

f 0 = [0, 0, fz ]T (26)
X
Z = AC t (w32 + w42 ) + Ti (1 , 3 , 4 ) (27)
i=1,2
The first expression in equation (27), derived from (1) is the thrust produced
by the propellers U 1 and U 2 in the top layer which depends only on their
rotational speeds. The second expression is the thrust (Ti ) produced by the
propellers L1 and L2 in the bottom layer influenced by propellers U 1 and U 2
on top. As Ti is a function of propeller speeds, and depends on the overlap
and offset, it can be estimated using the theoretical frame work presented in
section 3.3. Simulations were performed based on the model given in equation
(25) for various propeller speeds and proposed overlap and offset and a Look
Up Table (LUT) was computed for the current VOOPS configuration. The
thrust values (Ti ) were used from this LUT for simulation.

The moments m0 was also calculated in a similar way. The yaw moment
acting about Z0 of the body frame is a consequence of the difference in drag
on each propeller at different angular velocities [34] and is given as:
X
mz = AR3 Cq (w32 + w42 ) Hi (1 , 3 , 4 ) (28)
i=1,2
20 Ganeshram Nandakumar et al.

The first expression in equation (28) was derived from equation (15) and it
represents the net yaw moment generated by the propellers U 1 and U 2 in the
top layer, whereas, the second expression represents the yaw moment generated
by the propellers L1 and L2 in bottom layer which are under the influence of
the propellers U 1 and U 2. The moment due to overlapped/offset configuration
was estimated using the model presented in section 3.3, through experimental
studies. The moments about Y0 and X0 of the body frame are given by my
and mx respectively as in equations (29) and (30).

my = b 2[ACt (w32 + w42 ) + T1 (1 , 3 , 4 ) T2 (2 , 3 , 4 )] (29)

mx = b 2[ACt (w32 w42 ) T1 (1 , 3 , 4 ) + T2 (2 , 3 , 4 )] (30)

6 Controller design

The VOOPS quadrotor requires a controller for stabilization for which a cas-
caded PI-PID controller [35] was implemented. The inner loop is a velocity
loop which resists the change in angular velocities of the quadrotor in roll,
pitch and yaw using the gyroscope data. This is achieved by comparing the
desired rate of angular rotation 2 and the gyroscope output. The rate dif-
ference is fed back into the inner controller and sent to the motors to correct
the attitude by minimizing the difference. This makes the quadrotor stiff and
reluctant to change in attitude. The outer loop controls the angular position
of the quadrotor which helps it stabilize and compensates the drift in the in-
ner loop due to the sensor noise and other disturbances. The outer loop also
receives the set points from the joystick and is used to move the quadrotor
around.
The outer PI loop generates the desired rate of angular rotation by taking the
difference between the desired angle and the actual angle form the accelerom-
eter data, which is fed to the inner loop as a set point after transforming the
angular velocities from earth to body frame. Setting the right amount of gains
for these loops are critical to achieve stability. The gains were tuned for the
inner loop first and followed by the outer loops. Ziegler-Nichols Method [36]
was used to tune the gains. The same gains were used in the experimental
vehicle.

7 Simulation and experimental studies

Numerical simulations were performed using the mathematical model to study


the behavior of the VOOPS quadrotor. Simulations were carried out using
M AT LAB Simulink T M . A series of inputs on roll axis as explained below,
were given to change the roll of quadrotor.
1. An increasing ramp signal for first 5 seconds and a decreasing ramp for
next 5 seconds
Title Suppressed Due to Excessive Length 21

Fig. 13 PI-PID control loop

2. An impulse at time t=15 seconds


3. A step up at t=20 seconds and step down at t=25 seconds
The simulation results are shown in Figure 14. It was found that the vehicle
follows the commanded input and performs well with minimal errors. Also, it
was found that very jerky inputs are rejected by the vehicle due to the inertia
of the system.

Practical experiments were carried out using the VOOPS quadrotor to ver-
ify the simulated performance. Inputs similar to the simulation studies were
commanded and the response was recorded through onboard sensors. The roll
performance of the quadrotor is shown in Figure 14 along with the simula-
tion results. Except for the time delay (due to communication delays) and the
transients, the performance of the VOOPS quadrotor matches well with the
simulation studies. The quadrotor was able to follow the ramp input with a
slight lag and it also filtered out aggressive inputs like impulse. It responded
well for the step inputs with 20% overshoot and a settling time of less than 5
seconds. The results confirmed the utility of VOOPS configuration for practi-
cal design and development of quadrotors.

8 Benefits of VOOPS over conventional design

A conventional quadrotor with 14 inch propeller is considered for comparison.


This quadrotor is designed such that its tip-to-tip size is (1 + 2) times the
diameter of the propeller, making it the smallest possible design with a 14 inch
propeller. VOOPS experimental vehicle discussed in section 4.2 (20% overlap)
has an advantage of accommodating a larger propeller (15 inch) and still be
5.4% smaller than the conventional quadrotor. VOOPS can even accommodate
16 inch and 17 inch propellers (propellers are available in increments of an inch)
and still be smaller than the 14 inch conventional quadrotor discussed above,
but the overlap, and the losses due to overlap, will be significantly high.
A comparison of the power consumed by the quadrotor for producing thrust
for both the designs is shown in figure 15(a). It was found that the conven-
22 Ganeshram Nandakumar et al.

Fig. 14 Simulated and experimental responses for the given inputs

(a) (b)

Fig. 15 (a)Variation of power consumption with thrust (b) Variation of flight time with
payload

tional quadrotor always consumes more power than the VOOPS quadrotor
to produce a particular thrust. This is mainly due to the smaller propeller
used in the conventional quadrotors. As a result, VOOPS configuration will
always give a better flight time for a given battery capacity. Payload capac-
ity is another important consideration in the quadrotor performance. Figure
15(b) shows the variation of flight time with payload for both the designs.
It can be seen from the plot that VOOPS has better flight time for a given
payload. An analysis of data in Figure 15(a) and (b) showed that a 1kg of ad-
ditional payload will result in the consumption of 70.91 watts of more power
in a conventional quadrotor than the VOOPS quadrotor. This means VOOPS
Title Suppressed Due to Excessive Length 23

will have 18.3% more endurance than a conventional quadrotor. A comparison

Fig. 16 Thrust Comparison Conventional vs VOOPS quadrotor

of the thrust produced with respect to the rotor speed variations is shown
in Figure 16. It was found that the VOOPS quadrotor was able to produce
25.5% more thrust than the conventional design at 75% of the maximum rotor
speed(maximum take-off weight rotor speed percentage, which indirectly de-
cides the maximum payload capacity of the quadrotor). This clearly explains
why the VOOPS configuration enables higher payload and endurance. Since

Fig. 17 Size of VOOPS vs Conventional Quadrotor

one of the advantages of VOOPS is the size reduction, a comparison of the


size variation of the quadrotors with respect to the propeller diameter was
carried out. Figure 17 shows the variation of the minimum footprint for both
24 Ganeshram Nandakumar et al.

the designs with the propeller


diameter.
The footprint of smallest VOOPS can
be derived to be (1+ 2)/(1+ 2) times (17.2%) smaller than the smallest
conventional designs. This confirms that the size reduction is really feasible
and at higher propeller sizes, considerable reduction in the overall size of the
quadrotor is possible in VOOPS configuration.

9 Conclusion

This paper has presented a detailed theoretical and experimental analysis of a


new configuration of quadrotors with propeller overlap and offset. A theoret-
ical framework was proposed for the analysis of thrust losses due to overlap
and offset. It was found that the losses due overlap were minimal up to 25%
overlap and the losses due to offset were almost negligible at rotor operating
velocities. The overlap and offset has motivated VOOPS, a novel quadrotor
VTOL design capable of accommodating larger propeller which can enhance
the payload capability and endurance without increasing the footprint. This
paper also showed the satisfactory performance of a PI-PID controller (as
available on commercial autopilots) for the control of the VOOPS VTOL in
simulation studies and for the experimental vehicle. The experimental vehicle
has shown superior performance in terms of payload capability and endurance
when compared to the existing quadrotor UAV systems of the same class. The
experimental vehicle weighing 3000 grams was able to fly for 60 minutes and
was also capable of carrying additional payload of 2295 grams which is 40%
better than the conventional designs. Future work will study the performance
of VOOPS with advanced controllers in a trajectory tracking mission.

References

1. Gavrilets, V., Martinos, I., Mettler, B., and Feron, E., Control Logic for Automated
Aerobatic Flight of a Miniature Helicopter. AIAA Guidance, Navigation, and Control
Conference and Exhibit (2002).
2. Bouabdallah, S., Murrieri, P., and Siegwart, R., Design and Control of an Indoor Micro
Quadrotor, In proceedings of International Conference on robotics and Automation (2004).
3. Pines, D.J., and Bohorquez, F., Challenges Facing Future Micro-Air-Vehicle Develop-
ment, Journal of Aircraft, 43, 290305 (2006).
4. Bibik, P., Narkiewicz, Zasuwa, M., Zugaj, M., Quadrotor Dynamics and control of precise
handling, Studies in Systems, Decision and Control Innovative Simulation Systems, 33,
335-351 (2015).
5. Goslinski, J., Gardecki, S., Giernacki, W., An Efficient PSO-Based Method for an Iden-
tification of a Quadrotor Model Parameters, Progress in Automation, Robotics and Mea-
suring Techniques, Advances in Intelligent Systems and Computing, 351, 95-104 (2015).
6. Gabrlik, P., Kriz, V., Vomocil, J., Zalud, L., The Design and Implementation of Quadro-
tor UAV, Emergent Trends in Robotics and Intelligent Systems, Advances in Intelligent
Systems and Computing, 316, 47-56 (2015).
7. Kim, T.S., Karl, S., Kecman, V., Control of 3 DOF Quadrotor Model, Lecture Notes in
Control and Information Sciences Robot Motion and Control, 360, 29-38 (2007).
8. Szlchetko, B., Lower, M., On Quadrotor Navigation Using Fuzzy Logic Regulators, Com-
putational Collective Intelligence Technologies and Applications, 210-219 (2012).
Title Suppressed Due to Excessive Length 25

9. Nonami, K., Kendoul, F., Suzuki, S., Wang, W., Nakazawa, D., Autonomous Control
of a Mini Quadrotor Vehicle using LQG Controllers, Autonomous flying robots, 1, 61-76
(2010).
10. Yakovlev, K., Khithov, V., Loginov, M., Petrov, A., Distributed Control and Naviga-
tion System for Quadrotor UAVs in GPS-Denied Environments, Advances in Intelligent
Systems and Computing, 323, 49-56 (2015).
11. Sadeghzadeh, I., Abdolhosseini, Zhang, Y.M., Payload drop applications of unmanned
quadrotor helicopter using gain-scheduled PID and model predictive control techniques,
Intelligent Robots and Applications, 7506, 386-395 (2012).
12. Powers, C., Mellinger, D., Kushleyev, A., Kothmann, B., Kumar, V., Influence of Aero-
dynamics and Proximity Effects in Quadrotor Flight, Experimental Robotics, 88, 289-302
(2013).
13. Yan, H., Li, H., Song, S., Study on Dynamic Balance PID Control Algorithm and its
Application on the Quadrotor, International conference on Electrical Engineering and
Automatic Control, 367, 1071-1078 (2016).
14. Bibik, P., Narkiewicz, J., Zasuwa, M., Zugaj, M., Quadrotor Dynamics and Control
for Precise Handling, Studies in Systems, Decision and Control, Innovative Simulation
Systems, 33, 335-351 (2015).
15. Powers, C., Mellinger, D., Kumar, V., Quadrotor Kinematics and Dynamics, Handbook
of Unmanned Aerial Vehicles, 307-328 (2014).
16. Aydemir, M., Arikan, K.B., Irfanoglu, B., Disturbance rejection control of a Quadrotor
equipped with a 2 DOF manipulator, Machine Vision and Mechatronics in Practice, 91-103
(2015).
17. Michini, B., Redding, J., Ure, N.K., Cutler, M., How, J.P., Design and flight testing of
an autonomous variable-pitch quadrotor, IEEE International Conference on Robotics and
Automation (2011).
18. Efraim, H., Shapiro, A., Weiss, G., Quadrotor with a Dihedral Angle: on the Effects of
Tilting the Rotors Inwards, Journal of Intelligent & Robotic Systems, 80, 313-324 (2015)
19. Verbeke, J., Hulens, D., Ramon, H., Goedem, T., and Schutter, J.D., The Design and
Construction of a High Endurance Hexacopter Suited for Narrow Corridors*, International
Conference on Unmanned Aircraft Systems (ICUAS), 54351 (2014).
20. Driessens, S., Pounds, P., The Triangular Quadrotor: A More Efficient Quadrotor Con-
figuration, IEEE Transactions on Robotics, 31, 1517-1525 (2015).
21. Bouabdallah, S., Bermes, C., Grzonka, S., Gimkiewicz, C., Brenzikofer, A., Hahn,
R., Schafroth, D., Towards Palm-Size Autonomous Helicopters, Journal of Intelligent &
Robotic Systems, 61, 44571 (2010).
22. Driessens, S., and Pounds, P. E. I., Towards a More Efficient Quadrotor Configuration,
IEEE/RSJ International Conference on Intelligent Robots and Systems, 1386-92 (2013).
doi:10.1109/IROS.2013.6696530.
23. Pounds, P., Mahony, R., Gresham, J., Towards Dynamically-Favourable Quad-Rotor
Aerial Robots, Australian Conference on Robotics and Automation (2004)
24. Idea Forge, http://www.ideaforge.co.in/ (2016).
25. Abdilla, Analiza, Richards, A., and Burrow, S., Endurance Optimisation of Battery-
Powered Rotorcraft, Towards Autonomous Robotic Systems, 112 (2015).
26. Nandakumar, G., Ranganathan, T., Arjun, B.J., and Thondiyath, A., Design and Anal-
ysis of a Novel Quadrotor System VOOPS,. IEEE International Conference on Robotics
and Automation (ICRA) (2015).
27. Apostolo, G., The Illustrated Encyclopedia of Helicopters. Bonanza Books, New York
(1984).
28. Leishman, G. J., Principles of Helicopter Aerodynamics. Cambridge University Press,
Cambridge (2002).
29. Johnson, W., Helicopter Theory. New York: Dover Publications, New York (1995).
30. Stepniewski, W.Z., and Keys, C.N., Rotary-Wing. Dover Publications, New York,
(1984).
31. Megson, T., Introduction to Aircraft Structural Analysis. Butterworth-
Heinemann/Elsevier, Amsterdam (2010).
32. Fossen, Thor I., Guidance and Control of Ocean Vehicles. Wiley, Chichester (1994).
33. Murray, R. M. M., Shastry, S., Li, Z., and Sastry, S.S., A Mathematical Introduction to
Robotic Manipulation. CRC Press, Boca Raton (1994).
26 Ganeshram Nandakumar et al.

34. Hoffmann, G., Huang, H., Waslander, S., and Tomlin, C., Quadrotor Helicopter Flight
Dynamics and Control: Theory and Experiment, AIAA Guidance, Navigation and Control
Conference and Exhibit (2007).
35. Craig, J. J., Introduction to Robotics: Mechanics and Control. 3rd ed. Prentice Hall,
United States (2005).
36. Ziegler, J.G., and Nichols, N.B., Optimum Settings for Automatic Controllers, Journal
of Dynamic Systems, Measurement, and Control, 115, 220-22 (1993).

Você também pode gostar