Anda di halaman 1dari 17

AIAA AVIATION Forum 10.2514/6.

2013-4282
August 12-14, 2013, Los Angeles, CA
2013 Aviation Technology, Integration, and Operations Conference

Development of a Sizing and Analysis Tool for


Electrohydrostatic and Electromechanical Actuators
for the More Electric Aircraft
Imon Chakraborty,∗ David Jackson,∗ David Trawick,∗

Dimitri N. Mavris
Aerospace Systems Design Laboratory, School of Aerospace Engineering
Georgia Institute of Technology, Atlanta, GA, USA
Downloaded by BEIHANG UNIVERSITY on December 18, 2018 | http://arc.aiaa.org | DOI: 10.2514/6.2013-4282

This work documents the development of a MATLAB/Simulink based methodology


for the sizing, simulation, analysis, and optimization of electric actuators for the pri-
mary and secondary control surfaces of a More Electric Aircraft. For a given aircraft
and control surface configuration, the control surface flight loads are first evaluated
飞行载荷评估 taking into account their aerodynamic characteristics and the critical flight conditions
relevant to each. With this information, the performance of a given actuator design
can be analyzed via a simulation of the actuator and thermal dynamics. Conversely,
for a given objective function and constraint set, the actuator design can be optimized
through the solution of a constrained optimization problem. This work focuses on
the development of the flight load estimation capability, the modeling and simulation
environment, and the weight estimation method, while a separate work describes the
actuator optimization problem and a study of actuator-to-surface allocation. While
applicable to a wide variety of aircraft, the current work analyzes electrohydrostatic
and electromechanical actuators using the Boeing 737-800 aircraft as a test case.

I. Introduction
A. Electric Actuation vs. Conventional (Centralized) Hydraulic Actuation
The majority of military and commercial aircraft in service today use a centralized hydraulic system
for control surface actuation, where an engine-driven hydraulic pump (sometimes accompanied by an
electric motor-pump) pressurizes hydraulic fluid that acts on hydraulic cylinders to move the control
surfaces (ailerons, elevators, rudder, etc.). Centralized hydraulics have the advantage of being a time-
tested and proven solution whose operational characteristics are well-understood through decades of
aeronautical experience.
Centralized hydraulic systems have nevertheless reached a so-called “technology saturation”, a point
of diminishing returns whereafter further improvements in operational efficiency are increasingly difficult
to achieve. Further, the flight control redundancy mandated by civil and military regulations essen-
tially requires the incorporation of independent hydraulic lines/systems, which adds both weight and
complexity but can still be at risk from common-cause failures (e.g., United Airlines Flight 2321 ).
The possibility of electric control surface actuation offering some design benefits in terms of reduced
weight and operational vulnerability, improved efficiency, etc. had been considered conceptually as far
back as the World War II era.2 At that time however, electric drives did not have the power density
required to make them a feasible design choice, and designers preferred the hydraulic systems which
did. However, advances in electric drive technology over the past few decades that have resulted in
the development of motors with high power densities have made electric actuation a concept worth re-
visiting. It is therefore one of several avenues of research for the More Electric Initiative (MEI), which
aims to replace conventional aircraft subsystems with electric alternatives in order to produce a More
Electric Aircraft (MEA).
Some exploratory research has already been conducted to assess the feasibility of electric technologies
including control surface actuation. Cloyd2 provides an account of past, recent, and proposed programs
∗ Graduate Researcher, Aerospace Systems Design Laboratory, School of Aerospace Engg., Georgia Tech.
† Boeing Chaired Professor, School of Aerospace Engg., and Director, Aerospace Systems Design Laboratory, Georgia
Tech.

1 of 17

American
Copyright © 2013 by the American Institute of Aeronautics and Astronautics, Inc. Institute of Aeronautics and Astronautics
All rights reserved.
and developmental activities of the United States Air Force (USAF) with regards to More Electric tech-
nology. Thompson3 describes two test programs for electric actuation using the C-141 and C-130 aircraft.
In the late nineties, NASA’s F-18 Systems Research Aircraft (SRA) was used to test the performance of
two types of electric actuation, Electrohydrostatic Actuators (EHA)4 and Electromechanical Actuators
(EMA).5 For a thorough description of the layout of these actuators, the interested reader is referred to
Botten et al.6 Among current commercial aircraft, the Airbus A380 features significant electric actuation
in the form of EHAs and Electrical Backup Hydraulic Actuators (EBHAs).7 Among current military
aircraft, they are used for the F-35 Joint Strike Fighter, following the Joint Strike Fighter Integrated
Subsystem Technology (J/IST)8 demonstration program which used the F-16 aircraft as a test-bed.
With electric actuation, a ‘Power-by-Wire’ (PBW) system replaces the centralized hydraulics, making
possible the concept of ‘Power on Demand’ (sometimes the phrase ‘Pressure on Demand’ is also used9 ),
in which the actuator is powered only when a control surface movement is required or a load must be
overcome, largely eliminating the continuous losses that occur within a hydraulic circuit.
More-electric actuation systems promise a significant weight savings, with Kulshreshtha and Char-
rier10 citing a weight-saving of around 1,000 lb for the A380 aircraft directly attributable to the use of
electric actuators. Considering the fact that the A380 design retains centralized hydraulics with a 2H/2E
configuration, it could be argued that the weight savings for an aircraft of that weight category with
fully electric actuation could be even higher. The reduction in aircraft empty-weight directly translates
Downloaded by BEIHANG UNIVERSITY on December 18, 2018 | http://arc.aiaa.org | DOI: 10.2514/6.2013-4282

into reduced cost of ownership and operation.


In addition to the increase in efficiency and the reduction in weight, electric control surface actuators
open up more possibilities for building in redundancy with a lower associated weight penalty, and for
military aircraft, they may potentially allow a reduction in operational vulnerability.
However, a transition to more-electric or all-electric control surface actuation comes with its own
set of challenges. At the technical level, the thermal dynamics of the actuator units become a crucial
consideration, since they act as localized sources of heat. Also, the increased electrical power requirement
and the nature of the power draw requires a more extensive consideration of power quality management11
and electromagnetic interference (EMI). From a design philosophy standpoint,12, 13 there is a disparity
in the amount of historical experience available for hydraulic actuators versus electrical ones, associated
with which is a natural reluctance on the part of airframers and actuator designers to transition towards
electric actuators.

B. Electrohydrostatic vs. Electromechanical actuation

(a) Electrohydrostatic Actuator (EHA) (b) Electromechanical Actuator (EMA)

Figure 1. Typical EHA and EMA layout (courtesy TRW Aeronautical Systems6 )

Designs for electric control surface actuators fall into two main categories, the key point of difference
being the means by which the motion of the electric motor is transferred to the movement of the actuator
ram. Figure 1 shows the typical layouts of these units (courtesy TRW Aeronautical Systems6 ).
• Electrohydrostatic Actuator (EHA) - This is a self-contained unit in which an electric motor (typ-
ically bi-directional) drives a hydraulic pump (typically fixed displacement) that pressurizes a
hydraulic cylinder and moves a piston that is connected to the control surface. An alternative im-
plementation of this concept is the Integrated Actuator Package (IAPa ),6 containing essentially the
IAP电机是单向定速旋转,液压泵是双向变量泵
a IAP is a registered trademark of TRW Aeronautical Systems

2 of 17

American Institute of Aeronautics and Astronautics


same elements (motor, pump, piston), but in which the motor is a fixed-speed unidirectional motor
and the pump is a variable displacement type whose swashplate angle is controlled to determine
flow rate and direction.
• Electromechanical Actuator (EMA) - This is a self-contained unit in which the electric motor is
coupled to a reduction gearbox, and the rotational motion of the gearbox output shaft is converted
to a linear (translational) motion of the actuator ram via a kinematic converter such as a ball-
screw or a roller-screw. For motors with suitable torque-speed characteristics, it may be possible
to eliminate the reduction gearbox, and directly mate the motor to the control surface.14
Relative advantages and disadvantages have been claimed for both actuator types, and are well-
documented in Botten et al.,6 Garcia et al.,15 etc, to which the interested reader is referred. Here, it
will only be mentioned that despite being the heavier and more complicated alternative, the EHA seems
to be favored for the actuation of especially the flight-critical control surfaces. This is due to the fact
that it is easy to ensure that in case of actuator failure, the unit will revert to a damped mode, allowing
the control surface to be blown back to a faired position. Further, the EHA design is inherently more
suited for side-by-side implementation with conventional hydraulics, as demonstrated in the Airbus A380
design by the EBHA units.7 Some historical experience with hydraulic components is also transferable
Downloaded by BEIHANG UNIVERSITY on December 18, 2018 | http://arc.aiaa.org | DOI: 10.2514/6.2013-4282

to the design of EHAs.


The EMA, despite being the simpler and lighter alternative, has thus far been considered mainly for
secondary flight control surfaces (spoilers etc.) due to the risk of mechanical jamming which represents a
means for a single-point failure of the actuator. Associated with this is the risk that an EMA-controlled
primary control surface could jam in a deflected position, potentially resulting in a serious flight control
problem. Techniques have been suggested to mitigate the effect of actuator jamming,16 but these include
the integration of additional components (adding to weight and complexity) that potentially detract
from the EMA’s simplicity. Additionally, when the NASA F-18 SRA aircraft was used to test an EMA
actuator (Jensen et al.5 ), the actuator displayed some overheating problems. Similar problems were not
encountered when the same control surface of the same aircraft was actuated using an EHA (Navarro4 ).

C. The Proposed Approach


The authors previously addressed the actuator analysis and optimization problem by considering the
optimal design of EHAs for the primary flight control surfaces of the USAF F-16 Fighting Falcon and
the A-10 Thunderbolt II aircraft (Chakraborty et al.,17 Hegde et al.18 ). This work will will extend the
analysis and optimization capability to EMAs and also consider secondary flight controls. The general
sequence of the study is presented below:

1. Estimation of control surface flight loads


(a) Determination of hinge moment coefficients / actuation force coefficients
(b) Determination of critical flight conditions yielding maximum hinge moments
2. Actuator governing relationships and weight estimation
(a) Development of EHA and EMA mathematical models and weight estimation
(b) Weight estimate calibration - using published actuator data
3. Framing and solving the optimization problem
(a) Identification of control variables and constraints
(b) Selection of objective function
(c) Solution of constrained optimization problem for each control surface
4. Actuator-to-surface allocation study

This paper will focus on Items 1 and 2, the development of the core simulation and weight estimation
methodology. Item 3, the constrained optimization of the actuator design, will form the focus of the
second paper of this pair.19 Finally, Item 4 is a study of the allocation of optimized actuator types (either
EHA or EMA) to the aircraft control surfaces, which is significant since the primary and secondary flight
control systems of a modern commercial aircraft consist of a large number of control surfaces, and it is
quite possible that a MEA of the future may use both EHA and EMA design types for flight control
actuation. Clearly then, the design of the flight control system of such an aircraft will involve not only
the optimal design of the EHA and EMA units used, but also their optimal allocation to control surfaces.

3 of 17

American Institute of Aeronautics and Astronautics


II. Test Aircraft Selection and Description
Downloaded by BEIHANG UNIVERSITY on December 18, 2018 | http://arc.aiaa.org | DOI: 10.2514/6.2013-4282

Figure 2. Boeing 737 flight control surfaces20

The Boeing 737-800 aircraft was chosen for this analysis, and Fig. 2 shows the layout of its control
surfaces. Pitch control is achieved by two elevators (interconnected by a torque tube) and a trimmable
horizontal stabilizer. Roll control is provided by two ailerons located on the outboard sections of the
wing and augmented by four flight spoilers located on the upper surface of each wing. Yaw control is
provded by a single rudder. In flight, the speedbrake function is served by the symmetric deployment of
the flight spoilers, while on the ground, the ground spoilers deploy in addition to the flight spoilers.
The high-lift system consists of leading edge (LE) and trailing edge (TE) flaps. The LE devices
comprise two krueger flaps (inboard of the engine) and four slats (outboard of the engine) on each wing.
The krueger flaps have two positions: retracted and extended. The slats have three positions: retracted,
extended (where they form a sealed LE flap), and fully extended (where they form a slotted LE flap).
The TE devices consist of double-slotted flaps inboard and outboard of each engine. The deployment
and retraction of the LE devices is sequenced with that of the TE devices.20

III. Determination of Control Surface Flight Loads


A. Trailing Edge Primary Control Surfaces - Ailerons, Elevators, and Rudder

bm /2
yo
yi b2m
AR = Sm

yi
ηi = bm /2
λc/4
yo
ηo =
Sm /2 bm /2

c̄ c̄f =
cf

λhl cf ch ch
Sf c̄h = cf

Figure 3. Control surface geometry for hinge moment coefficient estimation

For control surfaces hinged to the trailing edges of lifting surfaces (i.e, ailerons, elevators, and
rudder), an overall hinge moment coefficient can be determined that comprises a baseline contribution

4 of 17

American Institute of Aeronautics and Astronautics


due to control surface camber (if any), a contribution due to local effective incidence angle (αef f ), and
a contribution due to control surface deflection (δ),
Ma = q̄Sf cf Ch
Ch = Ch,0 + Ch,α (M ) αef f + Ch,δ (M, δ) δ, (1)
where q̄ is the dynamic pressure, and Sf and cf are respectively the planform area and the chord of the
control surface. A slight manipulation of Eq. 1 yields the following:
Ma = q̄Sf cf (Ch,0 + Ch,α (M ) αef f + Ch,δ (M, δ) δ)
= [Ch,0 , Ch,α (M ), Ch,δ (M, δ)] . [1, αef f , δ]T q̄ . Sf c f (2)
The first group of parameters consists of the dimensionless hinge moment coefficients which were
estimated empirically using the methodology of Roskam.21 A MATLAB script utilizes information such
as the planform area (Sf ) and mean aerodynamic chord of the control surface (cf ), flap chord ratio (c̄f ),
airfoil shape, hinge-line sweep (λhl ), main surface sweep (λc/4 ), spanwise coordinates of the inboard
and outboard extremities of the surface (ηi , ηo ), and distance between the hinge-line and the leading-
edge of the control surface (c̄h ) (Fig. 3) to generate a 1-dimensional lookup table for Ch,α (M ) and a
2-dimensional one for Ch,δ (M, δ). In a previous work (Chakraborty et al.17 ), the hinge moment estimates
Downloaded by BEIHANG UNIVERSITY on December 18, 2018 | http://arc.aiaa.org | DOI: 10.2514/6.2013-4282

obtained by this method had been compared to published flight-test hinge moment measurements from
the NASA F-18 SRA program,4 and the agreement was found to be acceptable.
The parameters in the second group collectively define a flight condition. The maximum hinge
moment corresponds to a flight condition that maximizes the product of the first two groups. Given the
Mach number and deflection angle dependencies of the coefficients of the first group, finding this critical
flight condition requires the exploration of multiple points in the aircraft altitude-speed flight envelope
in search of the peak hinge moment. The guidelines are provided in Part 25 of the Federal Aviation
Regulations22 (FARs), and are further discussed in the works of Scholz.23–25 Only a brief outline is
provided here.
• Ailerons - The maneuvers that generate the maximum aileron hinge moments are described in FAR
25.349, and correspond to an abrupt full aileron deflection at the design maneuver speed (VA ), an
aileron deflection that generates the same roll rate as this case but at the design cruise speed (VC ),
and a deflection that generates a third of this roll rate at the design dive speed (VD ), evaluated at
two load factors.
• Elevators - Maneuvers resulting in peak elevator hinge moments are outlined in FAR 25.255, and
the hinge moments must be evaluated at different altitudes along the maximum operating speed
/ Mach number (VM o /MM o ) boundary of the aircraft flight envelope. Notably, this FAR refers to
a “runaway” of the trim system, in case of the Boeing 737-800 the trimmable horizontal stabilizer
(THS), and elevator authority required to generate a specified recovery load factor in response to
this occurrence.
• Rudder - The maximum rudder hinge moments can be found from a combined consideration of the
flight conditions outlined in FAR 25.149 and FAR 25.351, the latter along the VM o /MM o boundary,
and evaluated with maximum available/permissible rudder deflection. In this search, the reduction
in rudder authority (by design) of the Boeing 737-800 was taken into account.26
Finally, the quantities in the third group pertain to the control surface geometry, and were measured
off a CAD drawing of the Boeing 737-800 available from the Boeing website.27

B. Spoilers
The Boeing 737-800 has two types of spoilers, flight spoilers and ground spoilers, the latter deploying
only when the aircraft is on the ground. As noted previously, flight spoilers can be deployed symmetrically
to serve as speedbrakes, which are rated up to the aircraft’s design dive speed (VD ).
In contrast to hinge moment estimation methods for trailing-edge surfaces (shown above), there is
not an abundance of handbook methods for the calculation of spoiler hinge moments. Scholz23 reviewed
available data pertaining to spoilers from literature and also aircraft manufacturers and suggested the
following two relationships applicable to deployed and retracted spoiler panels.
1 2 1
Mext = ρV ( CD sin2 δsp )Ssp csp
2 l 2 s  2
csp Ssp W nz 2ysp
Mret = Ksp 1− (3)
c(ysp )bπ b

5 of 17

American Institute of Aeronautics and Astronautics


where ρ is the air density, Ssp and csp respectively the spoiler planform area and chord, δsp the spoiler
deflection, b and c(ysp ) the wingspan and the wing chord at the spanwise location of the spoiler (ysp ),
W and nz the aircraft weight and load factor. The author suggests CD = 1.8 as a representative spoiler
drag coefficient, Vl = 1.14 times the freestream velocity, and Ksp = 1.5 which gives a good match with
available spoiler data.
Table 1 shows the maximum deflections of the primary flight control surfaces and the spoilers. How-
ever, for the primary controls, the flight condition corresponding to peak hinge moment does not nec-
essarily correspond to full control surface deflection. Note: Spoilers are numbered starting from the
outboard left wing, and moving right. For the control surface deflection rates, Scholz23 states that for
a preliminary estimate, a maximum angular rate numerically equal to the angle between the surface’s
neutral and hardover positions may be used, but points out that automatic flight control system (AFCS)
requirements may necessitate a significantly higher rate requirement. This is confirmed by Kulshreshtha
and Charrier,10 who give a value of 60◦ /s, which has been used for this work.
Figure 4 shows the occurrence of the peak hinge moments for the control surfaces listed in Table 1.
For the ailerons, the maximum hinge moment occurs at the design cruise speed VC . For the elevators and
the rudder, the critical flight condition corresponds to the simultaneous occurrence of VM o and MM o .
For the flight spoilers, the maximum hinge moment occurs at the design dive speed VD when they are
used for the speedbrake function. Since the ground spoilers are meant to deploy only when the aircraft
Downloaded by BEIHANG UNIVERSITY on December 18, 2018 | http://arc.aiaa.org | DOI: 10.2514/6.2013-4282

is on the runway, the maximum rated tire speed was used to determine their critical hinge moment.
Table 2 contains the maximum hinge moments encountered for each control surface type corresponding
to these flight conditions.

Table 1. Control Surface Deflection Ranges26, 28, 29

Control Max. deflection (deg)


Ailerons ± 20
Elevators ± 20
Rudder ± 18
Flight spoilers (2,3,10,11) / (4,5,8,9) 33 / 38
Ground spoilers (1,12) / (6,7) 60 / 52

Aileron − blue, Rudder − red, Elevator − cyan, Flt. spoiler − black


500 500

450 450

400 400

350 350

300 300
Altitude (FL)

Altitude (FL)

250 250

200 200

150 150

100 100

50 50

0 0
120 160 200 240 280 320 360 400 150 200 250 300 350 400 450 500
Vmax (KEAS) Vmax (KTAS)

Figure 4. Occurrence of maximum hinge moments for ailerons, elevators, rudder, and flight spoiler

C. High Lift Devices


Figure 5 shows the high lift system of the Boeing 737. The trailing edge flaps (TEFs) are driven by
a centrally located Power Drive Unit (PDU) through a system of torque-tubes and gearboxes which
guarantees flap synchronization mechanically. Since this arrangement prevents a potentially hazardous
“flap disagree” condition, it is commonly used for commercial aircraft. This form of mechanical syn-
chronization is also found for the leading edge flaps (LEFs) of most commercial transports, however, for

6 of 17

American Institute of Aeronautics and Astronautics


Table 2. Maximum hinge moments and flight conditions for ailerons, elevators, rudder, and spoilers

Control Max. hinge moment (Nm) Flight condition


Ailerons 4,200 V = VC , sea-level
Elevators 7,600 V = V M o , M = MM o
Rudder 8,200 V = V M o , M = MM o
Flight spoilers 4,200 V = VD , sea-level
Ground spoilers 3,800 V = 196 KEAS, sea-level
Downloaded by BEIHANG UNIVERSITY on December 18, 2018 | http://arc.aiaa.org | DOI: 10.2514/6.2013-4282

Figure 5. High lift system on Boeing 737 (Rudolph30 )

the Boeing 737-800, the leading edge slats and Kruegers are actuated by individual hydraulic actuators,
with synchronization achieved electronically rather than mechanically.
While reliable, mechanically linked flap systems are limited in terms of operational flexibility, and
more recently distributed flap drive concepts have been proposed.31–35 Potential benefits include a
reduction in weight, simplication and localization of the flap system, reduction in total parts count, and
possibility of Differential Flap Settings (DFS).16
Unfortunately, the estimation of actuation loads for the high lift devices is more complex than that
for the hinged control surfaces treated previously. Both TEFs and LEFs deploy along flap tracks or
guide mechanisms, and the actuation loads are strong functions of the specifics of the mechanism (which
vary from aircraft to aircraft). It is therefore difficult to formulate a method for estimating these loads
that is general enough to be applicable to a wide variety of aircraft without specific consideration of the
flap deployment mechanism. For the purposes of this work, the electric actuator for the TEFs will be
chosen to match the published specifications of the lead-screw used for the Boeing 737-800.36 For the LE
slats, representative force and moment coefficients obtained from wind tunnel experiments by Kelly and
McCullough37 were used after scaling the presented figures for the slat dimensions of the Boeing 737-800
aircraft. For the Krueger flaps, the wind tunnel data presented by Moraris et al.38 were used after
similar scaling and a reasonable correction for the effect of the difference in Reynolds number. These
flight load coefficients were functions of the flap deployment angle and the angle of attack.
For obtaining the constraining actuation loads for the LEFs, the aircraft’s maximum flap extension
airspeed (VF E ) was used. The magnitude of the actuation loads was found to be consierably more for
the Krueger flaps than for the slats. In addition, the data for the Krueger flaps showed load reversal
with both changes in angle of attack and deployment angle, which is not very desirable.30
Table 3 contains the computed actuation load requirements for the high lift devices. Note that since
the Krueger flap is hinged, the optimization of its actuator is done based on the hinge moment, while
optimization of the slat and TEF actuators is done based directly on stall loads.

7 of 17

American Institute of Aeronautics and Astronautics


Table 3. Actuator specifications for high lift system

High-lift device 2-slotted TEF LE slats Krueger flaps


Flaps/Aircraft 4 8 4
Actuators/flap 2 1 1
Sizing load 51 kN 6.3 kN 5.6 kNm
Output speed 102 mm/s 60 mm/s 16 deg/s

IV. Actuator Mathematical Models


A. Component Models
Table 4 and Figs. 6 and 7 show the main components of the EHA and EMA designs, which are
successively described in this section. Additional details regarding the EHA formulation are available in
Pachter and Houpis,39 Rongjie et al.,40 and Crowder and Maxwell.41
Downloaded by BEIHANG UNIVERSITY on December 18, 2018 | http://arc.aiaa.org | DOI: 10.2514/6.2013-4282

Table 4. EHA and EMA Main Components

Component EHA EMA


Kinematic linkage X X
Ball-screw X
Gearbox X
Hydraulic cylinder X
Hydraulic pump X
Electric motor X X

∆δ
Gk = ∆x Q 1 mt = mp + mj + m b Q 2
∆x
Airframe ∆P = P1 − P2
Control surface
x

t
Uni P1
mp
P2
EHA ∆δ mb mj FL

(a) Kinematic Linkage (Gearing) (b) Hydraulic cylinder

R L

di
V V = E + L dt + Ri E
i

dpp α θ̇p = θ̇m


Ipump Imotor
dp τm
τe

(c) Hydraulic pump (d) Electric motor

Figure 6. Schematic diagrams of EHA components

8 of 17

American Institute of Aeronautics and Astronautics


ωin

1 vout

∆δ
Gk = ∆x
∆x
Airframe
Control surface Gbs = vout /ωin

t 2
Uni ωin
E MA ∆δ

vout
(a) Kinematic Linkage (Gearing) (b) Ball-screw

φd2
Downloaded by BEIHANG UNIVERSITY on December 18, 2018 | http://arc.aiaa.org | DOI: 10.2514/6.2013-4282

ωint L
R
φd3 Iint

di
L V V = E + L dt + Ri E
Iout Iin
i
ωgb,out φd1 ωgb,in
φd4
(c) Gearbox (d) Electric motor

Figure 7. Schematic diagrams of EMA components

1. Kinematic Linkage
This accounts for the conversion of the linear motion of the actuator jack to the angular motion of
the control surface. Since a detailed design of this mechanism would require knowledge of the available
installation space, in this work the conversion is represented using a gearing ratio, Gk (Fig. 6a), relating
the control surface deflection and its derivatives to the jack displacement and its derivatives.

∆δ δ̇ δ̈
Gk = = = . (4)
∆x ẋ ẍ
Energy conservation applied to the linkage results in the following relationship between the external
hinge moment (Mh ) and the actuator jack load (FL ), where ηm < 1 represents the mechanical efficiency
of the linkage.

ηm FL ∆x = Mh ∆δ,
Mh ∆δ Gk
=⇒ FL = = Mh . (5)
ηm ∆x ηm

2. Ball-screw
The ballscrew is used in the EMA for the conversion of the rotary input coming from the gearbox
into a linear output. The two main components are a screw and a nut, and for the mechanism to operate
correctly, one component must be allowed freedom in rotation (with translational motion restricted),
while the other must be allowed translational freedom (with rotational motion restricted). For the
actuation of hinged control surfaces, Configuration 1 (Fig. 7b) is used, in which the actuator ram is
connected to the control surface via a revolute joint, and the rotation of the screw is restricted using
an anti-rotation device (for schematics and diagrams, the interested reader is referred to Jensen et al.5 ).
Configuration 2 can be used for the actuation of high lift devices, especially trailing-edge flaps, in which

9 of 17

American Institute of Aeronautics and Astronautics


the flap carriage is attached to the nut, which in this case has a translational motion, while the screw is
the driven (rotating) member. The ballscrew gearing is defined as
vout v̇out
Gbs = = . (6)
ωin ω̇in
The input torque required to drive the mechanism against an external load FL is given by
mrec G2bs
 
Gbs (mrec v̇out + FL ) FL Gbs
τbs,in = + Irot ω̇in = Irot + ω̇in + , (7)
ηbs ηbs ηbs

where (depending on the configuration) mrec is the mass of the reciprocating component, Irot is the
moment of inertia of the rotating component, and ηbs is the ballscrew efficiency which is given by
tan αbs
ηbs = , (8)
tan(αbs + tan−1 (µbs ))
where αbs is the ballscrew helix angle and µbs is the rolling friction coefficient.
Downloaded by BEIHANG UNIVERSITY on December 18, 2018 | http://arc.aiaa.org | DOI: 10.2514/6.2013-4282

3. Gearbox
The EMA gearbox transfers the rotational motion of the motor output shaft to the ballscrew via a
speed reduction. The gearbox ratio is defined as
ω̇gb,in ωgb,in ωm
Ggb = = = , (9)
ω̇gb,out ωgb,out ωbs
where ωm and ωbs are respectively the angular speed of the motor output shaft and the ballscrew rotating
member. The gearbox may have multiple stages, and in this case the overall gearing ratio is the product
of the stage gearing ratios. For example, in (Fig. 7c), a two-stage system is shown. Here, the stage
gearing ratios are given by G1 = ωgb,in /ωint = d2 /d1 , and G2 = ωint /ωgb,out = d4 /d3 , and clearly
Ggb = G1 G2 .
The input torque required to drive the gearbox against an external load (in this case, the ballscrew
input torque, τbs,in ), taking into account the gearbox inertia, is given by
ef f
τbs,in /Ggb + Igb ω̇gb,in
τgb,in = τm = , (10)
ηgb
ef f
where ηgb is the gearbox efficiency, and Igb is the effective gearbox inertia with respect to the input
shaft, which takes into account the speed reduction ratio of the gearbox. For the two-stage system shown
in the figure, the effective inertia is given by

ef f Iint Iout
Igb = Iin + + 2 2, (11)
G21 G1 G2
where Iin , Iout , and Iint are respectively the moments of inertia of the input, output, and intermediate
shafts. For additional stages, the effective moment of inertia with respect to the gearbox input shaft is
obtained through an extended application of (11).

4. Hydraulic Cylinder
A balanced, double-acting cylinder is considered for EHAs, with total translating mass mt (Fig. 6b).
The inverse simulation of its dynamics (Eq. (12) and Fig. 6b) gives the instantaneous pressure difference
(∆p(t)) required to produce the desired control surface motion (δ(t)).
1
∆p = (mt ẍ + cẋ + FL ). (12)
Ap

5. Hydraulic Pump
A fixed-displacement variable-speed piston-pump (Fig. 6c) directly mated to the motor is considered
for the EHA. The total required flow rate is given by the sum of the theoretical pump flow rate and the
leakage flow rate, the latter modeled using a constant leakage coefficient cleak . The total flow is given by

Qtotal = Qideal + Qleak = Ap ẋ + cleak ∆p. (13)

10 of 17

American Institute of Aeronautics and Astronautics


As a result of the pump leakage, the pump continues to rotate at a small RPM when the control surface
is stationary under loaded condition. The leakage coefficient is a function of pump design, and in the
analysis it is automatically set to give a leakage-induced pump-speed of around 100 RPM under stall
load.41 Alternatively, the maximum permissible leakage flow rate can be specified as a percentage of the
maximum theoretical flow rate. The instantaneous pump speed (ωp ) required to produce the required
instantaneous flow rate Qtotal is given by
Qtotal
ωp = θ̇p = , (14)
Dpump

where Dpump is the pump displacement, defined as the fluid volume displaced by the pump per unit
revolution. The pump shaft torque (τm ) is obtained by using Eq. (14) in an energy balance for the
pump, that takes into account its hydraulic efficiency (ηh ).

∆pDpump
ηh τm θ̇p = ∆pQtotal = ∆pDpump θ̇p , =⇒ τm = (15)
ηh

6. Electric Motor
Downloaded by BEIHANG UNIVERSITY on December 18, 2018 | http://arc.aiaa.org | DOI: 10.2514/6.2013-4282

A brushless DC (BLDC) motor has been considered for this work,40 although other motor types are
also applicable.42 Frischemeier43 provides a typical torque-speed characteristic for this type of motor,
which is reproduced in Fig. 8. For the EHA, the motor is directly mated to the pump, and so in that
case the pump and motor shaft speeds are the same, i.e. ωm = ωp (Fig. 6d). For the EMA, the motor
speed is the same as the gearbox input shaft speed, i.e. ωm = ωgb,in . The instantaneous electrical torque
of the motor, τe , is given by
τe = I θ̈m + Kvm θ̇m + τm , (16)
where I includes the mass moment of inertia of the rotating motor components (and includes the mass
moment of inertia of the rotating pump components as well for the EHA), and Kvm represents a viscous
resistance. τm is the required pump shaft torque in case of the EHA, and the required gearbox input
shaft torque in case of the EMA. The required current (i), the back-emf (E), and the required voltage
(V ) are then given by
τe di
i= , E = Ke θ̇m , V = E + L + Ri. (17)
Kt dt
where L and R are the effective inductance and resistance of the motor, and Ke and Kt are the motor
speed and torque constants. The mechanical power output (Pm ), electrical power requirement (Pe ), and
the rate of heat dissipation (Pheat ) are given by

Pm = τm θ̇m , Pe = V i, Pheat = i2 R. (18)

Figure 8. Typical torque-speed characteristics for BLDC motor43

11 of 17

American Institute of Aeronautics and Astronautics


T0 , ambient
R0

T1 , case
R1
R4 T2 , stator iron
R3 R2
T3 , stator teeth
T4 , stator copper R5
T5 , rotor windings

R6
T6 , rotor iron
R8 R7
T7 , shaft
Downloaded by BEIHANG UNIVERSITY on December 18, 2018 | http://arc.aiaa.org | DOI: 10.2514/6.2013-4282

Figure 9. Motor Thermal Circuit

B. Motor Thermal Model


Since EHA and EMA units act as localized sources of heat, their thermal behavior under different
operational scenarios is an extremely important consideration. Excessive temperature buildup can affect
the health of the motor and associated power electronics, as evidenced by the F-18 SRA EMA test
program.5 Here, certain flight tests had to be terminated prematurely because of motor overheating
problems that occurred when the control surface was locked in deflected position for extended periods
of time (i.e., when the flaperon was used as a flap).
The thermal model of the motor that has been used in this work is based on the “lumped parameter”
or “lumped capacitance” method.44, 45 In this method, key parts of the motor (outer casing, rotor, stator,
etc.) are represented by nodes with thermal capacities (Fig. 9), which are thermally connected to other
nodes as per the motor layout. Heat transfer rates among nodes are governed by node temperatures and
the thermal resistances existing between two nodes (corresponding to the various heat transfer processes).
These thermal resistances are computed based on empirical relationships45, 46 that show good agreement
with experimental results. With this setup, the temperature of the ith node, Ti , is governed by an ODE
of the form
X Tj − Ti
mi ci Ṫi = + ui , . . . i = 1, 2, ..., 7, (19)
Rij
j, j 6= i

where mi and ci represent respectively the mass and specific thermal capacity of the node, Rij the
equivalent thermal resistance of all heat transfer processes between nodes ‘i’ and ‘j’, and ui is the
external heat input to the node ‘i’.

C. Simulink Model
In addition to being used for sizing the components of the EHA and EMA designs, the mathematical
relationships presented in the previous sections are incorporated into Simulink models for the EHA and
the EMA. The general layout of the model is shown in Fig. 10. The top layer contains a block that
reads in the relevant flight condition, a second block that calculates the applicable flight load based on
this flight condition and the control surface characteristics, and a third block that contains the actuator
dynamics. The contents of the actuator block are shown as well, and contain modules for the relevant
actuator components (EHA: linkage, cylinder, pump, motor. EMA: linkage, ballscrew, gearbox, motor).
The final subsystem simulates the thermal dynamics of the motor based on the instantaneous heat loss.

V. Component Weight Estimation


Since actuator weight is an important consideration for both civil and military applications of elec-
tric actuation, the EHA and EMA sizing codes also include weight estimation subroutines. Once key
design parameters for the actuators (gearbox ratio, pump displacement, etc.) are selected (either by

12 of 17

American Institute of Aeronautics and Astronautics


Downloaded by BEIHANG UNIVERSITY on December 18, 2018 | http://arc.aiaa.org | DOI: 10.2514/6.2013-4282

Figure 10. Simulink model layout - a common layout is shown here for brevity. The components in
parantheses are present in the EMA-Simulink model.

the user/analyst or by an optimizer during optimization), the weight estimation subroutines essentially
‘design’ the layout of the actuator components either from mechanical engineering principles, or us-
ing parametric models of the components in question. For continuity of presentation, specific details
regarding the weight estimation methods are deferred to the Appendix.
The weight estimation subroutines are able to account for the weights of key components, but they
are incapable of predicting the weights of fittings, seals, attachment points to the airframe, etc. In order
to account for these parts, “mass calibration factors” (MCFs) were assigned to each key component. The
values of these MCFs were determined by comparing predicted actuator weights with published figures
from several EHA and EMA research programs (with the code sizing for the same actuator specifications
as these programs).
The actuator performance figures from these programs (stall load, maximum ram velocity, etc.) were
fed to the EMA and EHA sizing codes, along with information regarding the actuator layout (dual-
tandem cylinder, number of electromechanical channels, etc.). The MCF values were selected such that
the predicted actuator weights showed reasonable agreement with the published weights from these
programs with the same values of the MCFs applied in each case.
The actuator specifications and actual and predicted weights that were used for the calibration are
shown in Tables 5 and 6 for EHAs and EMAs respectively. The expansions of the acronyms used are
presented in the footnote b .

VI. Conclusion and Future Work


This paper describes the development of a MATLAB/Simulink based tool for the analysis of electric
control surface actuators. The tool, which is based off a previous tool created by the authors for the
analysis of electrohydrostatic actuators (EHA), now has the capability of analyzing electromechanical
actuators (EMA) as well.
Using the control surface layout of a given aircraft or aircraft concept, it is capable of estimating the
aerodynamic loads (hinge moments) for ailerons, elevators, rudder, and spoilers in accordance with the
relevant flight conditions for each type of surface specified in FAR 25, and using an empirical methodology
b M - motor, BLDCM - brushless DC motor, PMM - permanent magnet motor, SRM - switched reluctance machine, BS

- ballscrew, GB - gearbox, FDP - fixed displacement pump, TC - tandem cylinder, φ - motor phase

13 of 17

American Institute of Aeronautics and Astronautics


Table 5. EHA Weight Estimate - Calibration

Program F-18 SRA flaperon Typical tandem EHA


Source Navarro4 Sadeghi & Lyons47
Stall load (lbf) 13,300 32,000
No load rate (in/s) 7.7 8.0
Stroke (in) 4.5 -
Architecture 1 3φ PMM → 1 FDP Dual EH channel, 2 TC/surface
Published weight (lb) 41.5 159.5
Predicted weight (lb) 42 164

Table 6. EMA Weight Estimate - Calibration

Program F-18 SRA flaperon Transport spoiler C-141 aileron


Source Jensen et al.5 Fronista & Bradbury42 Thompson3
Downloaded by BEIHANG UNIVERSITY on December 18, 2018 | http://arc.aiaa.org | DOI: 10.2514/6.2013-4282

Stall load (lbf) 13,200 50,000 19,050


No load rate (in/s) 6.7 7.0 4.65
Stroke (in) 4.125 6.0 5.43
Architecture 2 3φ BLDCM → 1 BS 1 5φ SRM → 1 BS 2 M/GB → 1 BS
Published weight (lb) 26 39 35
Predicted weight (lb) 26 40 37

for the computation of the control surface hinge moment coefficients. While actuator specifications for
the high lift devices are drawn up in this paper, the tool does not yet have a generalized (i.e. non aircraft-
specific) methodology for the estimation of the actuation loads for these surfaces. The enhancement of
the tool’s capability in this regard may form an avenue for future work.
With a certain limited number of actuator design variables specified by the user (or by an optimizer
during optimization), the tool is capable of sizing the major components of the EHA and the EMA in
accordance with relevant engineering principles in order to internally satisfy several design constraints,
and provide the user (or optimizer) with information pertaining to others.
Further, once the major components of the actuator are sized, the tool is capable of evaluating the
weight of each component, and then estimating the weight of the actuator as a whole taking into account
the user-specified design architecture (component redundancy, etc.). Since it is not possible to predict
the weight of each and every actuator sub-component using a conceptual design-stage analysis tool, the
weight predictions of the tool were calibrated against published actuator weight data after instructing
the code to match the performance specifications of each published design. This was done for both EHA
and EMA actuators, and the calibration factors were found to be acceptable in magnitude and consistent
across the calibration test set. This then sets the stage for the second paper of this set, that deals with
the constrained optimization of the control surface actuators and analyzes the allocation of actuator
types to control surfaces.
Further extensions or avenues for future work include further enhancements to the aerodynamic loads
prediction capability. The current methodology uses hinge moment estimation based on an empirical
methodology. However, a future version of the environment may contain a surrogate model of flight loads
computed using TORNADO,48 a software package utilizing the vortex-lattice method for aerodynamic
predictions. In addition, another possible line of future work may be the incorporation of a method
to estimate the flap system actuation loads using only limited information regarding the specific flap
mechanism.

Acknowledgments
The authors would like to thank Dr. Elena Garcia of the Aerospace Systems Design Laboratory
(ASDL) for her technical advice during this investigation.

14 of 17

American Institute of Aeronautics and Astronautics


Appendix - Component Weight Estimation
A. Hydraulic cylinder
The hydraulic cylinder is first sized using the actuator design stall load F0 , stroke ∆xmax , and
maximum operating pressure ∆pmax . The stall load and the stroke are used to compute the piston-rod
(ram) diameter (d) based on considerations of buckling, using the more conservative of the minimum
diameters predicted by Euler and Rankine formulae.49 Then the effective piston area (Ap ) is computed
as that required to generate the design stall load when acted upon by the maximum operating pressure
difference: Ap = κF0 /∆pmax , where κ > 1 represents a reasonable excess margin. With the piston
rod diameter andp the required effective piston area known, the piston outer diameter (D) is readily
computed as D = d2 + 4 Ap /π. With the cylinder bore now known, the required cylinder wall thickness
is computed based on circumferential (Hoop) stress. The end-cap thickness is computed using the
more conservative of the thicknesses required to withstand tensile and shear stresses. The current
methodology sizes “balanced” cylinders, i.e., a counterbalance rod is included by default. Two variations
of cylinder design are currently available: a single (simple) cylinder arrangement, and a tandem cylinder
arrangement, and the user can select the desired layout during the problem setup phase. Once the
dimensions of the piston and cylinder are known, representative material densities are used in order to
Downloaded by BEIHANG UNIVERSITY on December 18, 2018 | http://arc.aiaa.org | DOI: 10.2514/6.2013-4282

compute the mass of cylinder, piston, and hydraulic fluid. Finally, the cylinder is assigned a dedicated
mass calibration factor (MCF) which was set to obtain reasonable agreement with published actuator
weight data (as described in the text).

B. Hydraulic pump
The current methodology uses a parametric model of a fixed-displacement, bi-directional piston pump.
The pump displacement is the input to this sizing and weight estimation subroutine, whose output
includes the mass and the moment of inertia of the pump. For the EHA design, the moment of inertia
of the pump is added to that of the motor (since these are mated together coaxially) and is used for
the actuator simulation in Simulink. Future upgrades to the work will feature a pump sizing code that
responds to the pump rated speed (RPM) and input shaft torque in addition to the pump displacement.

C. Ballscrew
The length of the screw is set based on the actuator stroke ∆xmax , and its diameter is determined based
on the fact that it may be simultaneously subjected to a tensile/compressive force F0 and an associated
torque τ0 = F0 Gbs /ηbs (Khurmi49 ). Since the efficiency of the ballscrew is a function of the helix angle
which is also affected by the ballscrew gearing ratio and diameter, these quantities are related by implicit
relationships that are iterated to convergence within the ballscrew sizing routine. Once the dimensions
of the screw have been computed, representative material densities and the ballscrew mass calibration
factor (MCF) are used to determine the mass of the ballscrew.

D. Gearbox
The gearbox sizing and weight estimation code takes as input the required gearbox reduction ratio
(Ggb ). The distance between gearbox shafts is set as a fixed user-specified parameter (L). Based on this
distance, the maximum reduction Gmax available per gearbox stage is computed based on an allowable
minimum diameter of the pinion cog (dmin pin ). The minimum number of stages required in the gearbox
to achieve the required reduction ratio is then computed as n = ceil(log(Ggb )/log(Gmax )), where “ceil”
represents the “ceiling” function that increases a fraction to the next higher integer. Once the number
of stages is known, the stage reduction ratio is related to the final reduction ratio by Ggb = Gnstage .

E. Electric motor
For the purpose of weight estimate calibration, representative values of electric drive power-to-weight
ratio (kW/kg) were used. However, for the optimization of the actuator designs, the motor and power
electronics weight estimation relationships provided by Torabzadeh-Tari50 were used. These empirical
relationships allow motor, inverter, and converter weights to respond to motor torque, speed, and output
power for several motor types (out of which BLDC was used for this work).

15 of 17

American Institute of Aeronautics and Astronautics


References
1 National Transportation Safety Board (NTSB), “Aircraft Accident Report - United Airlines Flight 232, McDonnell

Douglas DC-10-10, Sioux Gateway Airport, Sioux City, Iowa, July 19, 1989,” Tech. Rep. NTSB / AAR-90 / 06, Nov 1
1990.
2 Cloyd, J., “Status of the United States Air Force’s More Electric Aircraft initiative,” Aerospace and Electronic

Systems Magazine, IEEE , Vol. 13, No. 4, April 1998, pp. 17–22.
3 Thompson, K., “Notes on ’The electric control of large aeroplanes’,” Aerospace and Electronic Systems Magazine,

IEEE , Vol. 3, No. 12, dec. 1988, pp. 19 –24.


4 Navarro, R., “Performance of an Electro-Hydrostatic Actuator on the F-18 Systems Research Aircraft,” 16th Digital

Avionics Systems Conference, Irvine, California, October 1997.


5 Jensen, S., Jenney, G., and Dawson, D., “Flight test experience with an electromechanical actuator on the F-18

Systems Research Aircraft,” 19th Digital Avionics Systems Conference, 2000 , Vol. 1, 2000, pp. 2E3/1 – 2E310.
6 Botten, S., Whitley, C., and King, A., “Flight Control Actuation Technology for Next-Generation All-Electric

Aircraft,” Technology Review Journal - Millenium Issue, 2000, pp. 55–68.


7 Van der Bossche, D., “The A380 Flight Control Electrohydrostatic Actuators, Achievements and Lessons Learnt,”

25th International Congress of the Aeronautical Sciences, ICAS 2006 , 2006.


8 Burkhard, A. and Dietrich, R., “Joint Strike Fighter Integrated Subsystems Technology, A Demonstration for In-

dustry, by Industry,” Journal of Aircraft, Vol. 40, 2003, pp. 906–913.


9 Cronin, M., “The all-electric aircraft,” IEE Review , Vol. 36, No. 8, sep 1990, pp. 309 –311.
10 Kulshreshtha, A. and Charrier, J., “Electric Actuation for Flight and Engine Control: Evolution and Challenges,”
Downloaded by BEIHANG UNIVERSITY on December 18, 2018 | http://arc.aiaa.org | DOI: 10.2514/6.2013-4282

SAE-ACGSC Meeting 99 , SAE, Feb 28 - Mar 2 2007.


11 Trainer, D. and Whitley, C., “Electric actuation-power quality management of aerospace flight control systems,”

Power Electronics, Machines and Drives, 2002. International Conference on (Conf. Publ. No. 487), june 2002, pp. 229 –
234.
12 Blanding, D., “Subsystem Design and Integration for the More Electric Aircraft,” 25th International Congress of

the Aeronautical Sciences (ICAS 2006), 2006.


13 Leonard, J., “The All-Electric Fighter Airplane Flight Control Issues, Capabilities, and Projections,” IEEE Trans-

actions on Aerospace and Electronic Systems, Vol. AES-20, 3, May 1984, pp. 234–242.
14 Lyshevski, S., “Electromechanical flight actuators for advanced flight vehicles,” Aerospace and Electronic Systems,

IEEE Transactions on, Vol. 35, No. 2, apr 1999, pp. 511 –518.
15 Garcia, A., Cusido, J., Rosero, J., Ortega, J., and Romeral, L., “Reliable electro-mechanical actuators in aircraft,”

Aerospace and Electronic Systems Magazine, IEEE , Vol. 23, No. 8, aug. 2008, pp. 19 –25.
16 Recksiek, M., “Advanced High Lift System Architecture with Distributed Electrical Flap Actuation,” AST 2009,

Workshop on Aviation System Technology, 2009.


17 Chakraborty, I., Trawick, D., Hegde, C., Choi, H., Mendez-Ramos, E., and Mavris, D. N., “Development of a

Modeling and Simulation Environment for Real-time Performance Analysis of Electric Actuators in Maneuvering Flight,”
51st AIAA Aerospace Sciences Meeting Including The New Horizons Forum and Aerospace Exposition, No. AIAA-2013-
0471, Grapevine, TX, January 2013.
18 Hegde, C., Chakraborty, I., Trawick, D., Choi, H., Mendez-Ramos, E., and Mavris, D., “A Surrogate Model Based

Constrained Optimization Architecture for the Optimal Design of Electrohydrostatic Actuators for Aircraft Flight Control
Surfaces,” 51st AIAA Aerospace Sciences Meeting Including The New Horizons Forum and Aerospace Exposition, No.
AIAA-2013-0470, Grapevine, TX, January 2013.
19 Chakraborty, I., Trawick, D., Jackson, D., and Mavris, D., “Electric Control surface Actuator Design Optimization

and Allocation for the More Electric Aircraft,” AIAA Aviation 2013 Conference, AIAA, Los Angeles, California, August
12-14, 2013.
20 “Boeing 737 NG Systems Summary - Flight Controls,” available online at: www.smartcockpit.com.
21 Roskam, J., Airplane Design Part VI - Preliminary Calculation of Aerodynamic, Thrust and Power Characteristics,

Design Analysis & Research, 1999.


22 “Federal Aviation Regulations (FAR) Part 25 - Airworthiness Standards: Transport Category Airplanes, Federal

Aviation Administration (FAA),” available online at http://www.ecfr.gov/.


23 Scholz, D., “Development of a CAE-Tool for the Design of Flight Control and Hydraulic Systems,” AeroTech ’95,

Birmingham,, 1995.
24 Scholz, D., Equations for a Preliminary Actuator Design, Technical Note, Deutsche Airbus.
25 Scholz, D., “Computer Aided Engineering for the Design of Flight Control and Hydraulic Systems,” ICAS Proccedings

1996 (20th Congress of the International Council of the Aeronautical Sciences), 1996.
26 “Use of Rudder on Transport Category Airplanes (For Boeing Aircraft Only),” Tech. rep., International Federation

of Air Line Pilots’ Associations (IFALPA), 2002.


27 “CAD 3-View Drawings for Airport Planning Purposes, Boeing Commercial Airplanes,” available online at

http://www.boeing.com/boeing/commercial/airports/.
28 “The Boeing 737 Technical Site (Aircraft Systems / Flight Controls),” available online at: http://www.b737.org.uk/.
29 “United Kingdom Air Accidents Investigation Branch (AAIB) Bulletin No: 5/97 Ref: EW/C96/9/7Category: 1.1,”

available online at: http://www.aaib.gov.uk/cms resources.cfm?file=/dft avsafety pdf 501724.pdf.


30 Rudolph, P., “High Lift Systems on Subsonic Commercial Airliners,” NASA Contractor Report 4746 CONTRACT

A46374D(LAS), National Aeronautics and Space Administration, September 1996 1996.


31 Bennett, J., Mecrow, B., Jack, A., Atkinson, D., Sewell, C., Mason, G., Sheldon, S., and Cooper, B., “Choice of

drive topologies for electrical actuation of aircraft flaps and slats,” Power Electronics, Machines and Drives, 2004. (PEMD
2004). Second International Conference on (Conf. Publ. No. 498), Vol. 1, 2004, pp. 332–337 Vol.1.
32 Bennett, J., Mecrow, B., Jack, A., Atkinson, D., Sheldon, S., Cooper, B., Mason, G., Sewell, C., and Cudley, D.,

“A prototype electrical actuator for aircraft flaps and slats,” Electric Machines and Drives, 2005 IEEE International
Conference on, 2005, pp. 41–47.

16 of 17

American Institute of Aeronautics and Astronautics


33 Bennett, J., Mecrow, B., Jack, A., and Atkinson, D., “A Prototype Electrical Actuator for Aircraft Flaps,” Industry

Applications, IEEE Transactions on, Vol. 46, No. 3, 2010, pp. 915–921.
34 Bennett, J., Mecrow, B., Atkinson, D., and Atkinson, G., “Safety-critical design of electromechanical actuation

systems in commercial aircraft,” Electric Power Applications, IET , Vol. 5, No. 1, 2011, pp. 37–47.
35 Bennett, J., Mecrow, B., Atkinson, D., and Atkinson, G., “Failure mechanisms and design considerations for fault

tolerant aerospace drives,” Electrical Machines (ICEM), 2010 XIX International Conference on, 2010, pp. 1–6.
36 “High Lift Actuation Systems, Triumph Gear Systems Inc., A Triumph Group Company,” available online at:

http://www.triumphgroup.com/Libraries/Triumph Gear Systems - Park City/Hi Lift Actuation.sflb.ashx.


37 Kelly, J. and McCullough, G., “Aerodynamic Loads on a Leading-edge Flap and a Leading-edge Slat on the NACA

64A010 Airfoil Section (NACA TN 3220),” Tech. rep., National Advisory Committee for Aeronautics (NACA), 1954.
38 Moraris, V., Lawson, N., and Garry, K., “Aerodynamic and Performance Characteristics of a Passive Leading Edge

Krueger Flap at Low Reynolds Numbers,” The Aeronautical Journal, Vol. 116 (1181), 2012, pp. 759–769.
39 Pachter, M. and Houpis, C., “Modeling of Electrohydraulic and Electrohydrostatic Actuators,” CD-ROM, The

Measurement, Instrumentation and Sensors Handbook, CRC Press1999.


40 Rongjie, K., Zongxia, J., Shaoping, W., and Lisha, C., “Design and Simulation of Electro-hydrostatic Actuator with

a Built-in Power Regulator,” Chinese Journal of Aeronautics, Vol. 22, 2009, pp. 700 – 706.
41 Crowder, R. and Maxwell, C., “Simulation of a prototype electrically powered integrated actuator for civil aircraft,”

Proceedings of the Institution of Mechanical Engineers, Part G, Journal of Aerospace Engineering, Vol. 211, June 1997,
pp. 381–394.
42 Fronista, G. and Bradbury, G., “An electromechanical actuator for a transport aircraft spoiler surface,” Energy

Conversion Engineering Conference, 1997. IECEC-97., Proceedings of the 32nd Intersociety, Vol. 1, 1997, pp. 694–698
Downloaded by BEIHANG UNIVERSITY on December 18, 2018 | http://arc.aiaa.org | DOI: 10.2514/6.2013-4282

vol.1.
43 Frischemeier, S., “Electrohydrostatic actuators for aircraft primary flight control - types, modelling and evaluation,”

5th Scandinavian International Conference on Fluid Power, SICFP 1997, Linkoping, Sweden, May 1997.
44 Mellor, P., Roberts, D., and Turner, D., “Lumped parameter thermal model for electrical machines of TEFC design,”

IEE Proceedings-B , Vol. 138, September 1991, pp. 205 – 218.


45 Boglietti, A., Cavagnino, A., Lazzari, M., and Pastorelli, M., “A Simplified Thermal Model for Variable-Speed Self-

Cooled Industrial Induction Motor,” IEEE Transactions on Industry Applications, Vol. 39, No. 4, July / August 2003,
pp. 945 – 952.
46 Staton, D., Boglietti, A., and Cavagnino, A., “Solving the More Difficult Aspects of Electric Motor Thermal Analy-

sis,” IEEE Transactions on Electronic Computers, Vol. 20, No. 3, 2005, pp. 620–628.
47 Sadeghi, T. and Lyons, A., “Fault Tolerant EHA Architectures,” IEEE AES Systems Magazine, March 1992, pp. 32–

42.
48 Online: http://www.redhammer.se/tornado/DL.html, User’s Guide - Reference Manual, TORNADO 1.0, Release

2.3 .
49 Khurmi, R. and J.K., G., A Textbook of Machine Design, Eurasia Publishing House (Pvt.) Ltd., 2005.
50 Torabzadeh-Tari, M., Dimensioning Tools of MEA Actuator Systems, Including Modeling, Analysis and Technology

Comparison, Ph.D. thesis, KTH Electrical Engineering, 2008.

17 of 17

American Institute of Aeronautics and Astronautics

Anda mungkin juga menyukai