Modeling Controller Design and Simulation Groundwo
Modeling Controller Design and Simulation Groundwo
Article
Modeling, Controller Design and Simulation Groundwork on
Multirotor Unmanned Aerial Vehicle Hybrid Power Unit
Matija Krznar , Danijel Pavković * , Mihael Cipek and Juraj Benić
Faculty of Mechanical Engineering and Naval Architecture, University of Zagreb, Ivana Lučića 5,
10000 Zagreb, Croatia; [email protected] (M.K.); [email protected] (M.C.); [email protected] (J.B.)
* Correspondence: [email protected]; Tel.: +385-(0)1-6168325
Abstract: This paper presents the results of modeling, control system design and simulation verifi-
cation of a hybrid-electric drive topology suitable for power flow control within unmanned aerial
vehicles (UAVs). The hybrid power system is based on the internal combustion engine (ICE) driving
a brushless DC (BLDC) generator supplying the common DC bus used for power distribution within
the aircraft. The overall control system features proportional-integral-derivative (PID) feedback
control of the ICE rotational speed using a Luenberger estimator for engine-generator set rota-
tional speed estimation. The BLDC generator active rectifier voltage and current are controlled by
proportional-integral (PI) feedback controllers, augmented by estimator-based feed-forward load
compensators. The overall control system design has been based on damping optimum criterion,
which yields straightforward analytical expressions for controller and estimator parameters. The
robustness to key process parameters variations is investigated by means of root-locus methodology,
and the effectiveness of the proposed hybrid power unit control system is verified by means of
comprehensive computer simulations.
Citation: Krznar, M.; Pavković, D.; Keywords: unmanned aerial vehicles; engine-based hybrid power unit; speed estimation and control;
Cipek, M.; Benić, J. Modeling, direct-current bus control
Controller Design and Simulation
Groundwork on Multirotor
Unmanned Aerial Vehicle Hybrid
Power Unit. Energies 2021, 14, 7125. 1. Introduction
https://doi.org/10.3390/en14217125
Nowadays multi-rotor unmanned aerial vehicles (UAVs) are used in many specialized
roles, among which search and rescue missions [1], border patrol and surveillance [2],
Academic Editor: Mario Marchesoni
aerial photography [3,4], inspection of critical infrastructure [5], and agriculture [6] are
most prominently featured. However, their more widespread utilization is related to
Received: 17 August 2021
Accepted: 20 October 2021
energy storage capacity limitations of the state-of-the-art lithium batteries for small-scale
Published: 1 November 2021
aircraft propulsion, and consequent fight autonomy and maneuvering capability issues [7].
Alternative aircraft propulsion systems utilizing internal combustion engine (ICE) as prime
Publisher’s Note: MDPI stays neutral
mover and liquid fuel as the energy source, introduce a number of constraints on the
with regard to jurisdictional claims in
UAV performance, as indicated in [8]. In particular, such propulsion systems significantly
published maps and institutional affil- increase the UAV mass due to the mass of the engine and the quite complex mechanical
iations. transmission system between the engine and the propellers, while also requiring quite
complex engine controls, while the flight dynamics are typically degraded compared to
fully electric propulsion due to slow dynamics of the ICE [9], which may even result in
aircraft stabilization issues [9].
Copyright: © 2021 by the authors.
Taking into consideration the aforementioned issues of both purely electric and purely
Licensee MDPI, Basel, Switzerland.
conventional (ICE-based) propulsion systems, a hybrid propulsion approach may be
This article is an open access article
considered instead [10]. Such solutions are recently becoming increasingly attractive
distributed under the terms and as research topics, in particular those using light-weight ICE coupled to an electricity
conditions of the Creative Commons generator [11], which have shown clear benefits in terms of UAV flight range extension
Attribution (CC BY) license (https:// and mission endurance [12]. Similar to the hybrid electric road vehicles, a suitably-sized
creativecommons.org/licenses/by/ battery may also be used to deal with propulsion system load transients, whereas the
4.0/). ICE can be operated in the vicinity of the specific fuel consumption optimal operating
point [13], thus increasing the UAV flight autonomy and mission endurance. Current
state of research in hybrid aircraft propulsion is mainly focused on regular fixed-wing
aircraft and those with VTOL (vertical takeoff and landing) capability [14–16], along with
lighter than air (dirigible) UAVs [17,18]. Due to specific energy and power requirements of
multi-rotor UAVs compared to the aforementioned aircraft types and their quite different
flight dynamics, the aforementioned results cannot be directly translated to multi-rotor
UAVs. Therefore, detailed simulation analyses and bench tests are typically required
in order to find a suitable energy management strategy for the hybrid propulsion UAV
configuration [19]. For that purpose, the hybrid propulsion system simulation model needs
to be coupled with the appropriate aircraft flight dynamics control system simulation model,
which closely matches typical situations encountered in flight [20]. The rigid body system
dynamic model derivation may be based on Newton-Euler approach [20] or Lagrange’s
method [21]. The target control strategy ought to take into account the specific requirements
of underactuated mechatronic systems (with UAVs being their typical representatives),
such as trajectory planning [21], with off-line and on-line optimization-based control
techniques employed to achieve precise trajectory tracking [21–23].
On the energy management control level, the rotational speed and torque of the
engine-generator set need to be coordinated by the higher-level (supervisory) control strat-
egy, commanding suitable target values to the engine speed and generator current/torque
control systems [13]. Consequently, precise control of these ICE-generator set quantities
is crucial for obtaining its fuel-efficient operation. For instance, an off-line optimization
approach based on dynamic programming (DP) can be used to obtain a benchmark against
which to compare a real-time hybrid power-train control strategy, such as that based on
fuzzy logic control [24]. In order to improve the accuracy of the DP offline optimization
result without increasing its computational load, a hybrid optimization algorithm com-
bining DP and gradient-based optimization has been presented in [25]. Nevertheless, the
aforementioned approaches cannot be used for online hybrid power-train control variable
optimization due to the requirement for full preview of the vehicle mission [26]. For online
power flow optimization and energy consumption minimization in hybrid powertrains, the
so-called equivalent consumption minimization strategies (ECMS) are used, either based
on Pontryagin’s principle [27] or a combined approach featuring online optimization as an
extension of a rule-based control strategy [28], which generate the necessary commands for
the hybrid power-train components. A practical implementation of a power flow control
system would also require on-line prediction of the unmanned aerial vehicle flight trajec-
tory, in order to optimize the energy distribution during aircraft maneuvering, for example
using the Kalman filtering methodology [29].
In order to implement the ICE-based hybrid power unit control strategy, it is necessary
to possess an accurate model of the engine-generator set, wherein ICE is typically modeled
by using the mean value engine modeling (MVEM) approach [30]. In this approach,
key parameters of the model are given as static maps, typically being experimentally
recorded [31] using a relatively large number of engine dynamometer tests. In that respect,
a nonlinear optimization-based calibration process [32] can be used to shorten the ICE
model development phase. Typically, engine rotational speed is controlled by means of
linear feedback controllers, such as the proportional-integral (PI) and proportional-integral-
derivative (PID) feedback controller [33], whose load suppression performance can be
further improved by adding a feed-forward load compensator [34]. Naturally, accurate real-
time information on engine rotational speed is required for accurate engine speed control.
However, typical engine control applications for road vehicles are characterized by rather
low-resolution of engine crankshaft position/speed measurement [30]. Moreover, electrical
machine current (torque) control depends on the type of machine, with alternating current
(AC) machines requiring a relatively complex field-oriented (vector) control [35]. Since
UAV applications typically feature specialized brushless direct current (BLDC) electrical
machines for propulsion due to their light weight [36], such machines could also be used
as electricity generators in hybrid propulsion UAVs, provided that active rectification is
Energies 2021, 14, 7125 3 of 26
facilitated using suitable phase voltage modulation techniques (see, e.g., [37]). This should
offer more flexibility in power flow control when compared to uncontrolled (passive)
diode-based rectification presented in [38]. However, the possible absence or low resolution
of engine-generator set position/speed sensors due to mounting space restrictions and
weight limitations may require that BLDC machine rotor position/speed sensor-less control
methods are used [39]. These are typically based on stator (armature) back electromotive
force (EMF) estimation [40], e.g., by using Kalman filtering approach, as shown in [41,42].
This paper proposes the low-level control system design aimed for power flow control
application within the hybrid propulsion UAV direct current (DC) power distribution
system, which utilizes the ICE plus BLDC machine generator set as the primary power
source. Engine speed control is based on PID speed controller with speed sensor-less
estimation of engine speed utilizing Luenberger estimator [43], with simultaneous closed-
loop control of the common DC bus voltage with additional load suppression based on the
Luenberger estimator. The proposed control systems and estimators are designed based
on suitable averaged linearized control-oriented process models and utilizing damping
optimum criterion [44]. The proposed UAV hybrid propulsion control system concept is
analyzed with respect to sensitivity to modeling errors, and ultimately validated through
simulations within MATLAB/Simulink software environment.
Figure 2. Mean value engine model (MVEM) of the internal combustion engine (a), and correspond-
ing linearized model in the vicinity of ICE operating point (b).
where l = {1, 2, 3} represents the phase number and p is the number of pole pairs, ω g is the
rotor rotational speed, Rϕm is the rotor magnetic flux at stator side, which is dependent on
the rotor position αg = ω g dt (mechanical angle), and Ke is the EMF constant per phase.
Energies 2021, 14, 7125 5 of 26
Figure 3. Electrical schematic of the BLDC machine with active rectifier at common DC bus (a) and current and voltage
waveforms for simultaneous conducting of two BLDC generator phases (b).
On the other hand, the total electromagnetic torque of the brushless DC machine may
be expressed in terms of individual phase currents il as follows:
3
∑ φm
τg = Ke pα g − 2π (l − 1)/3 il . (2)
l =1
switches [46]. This process is then repeated for different pairs of phases with respect to
rotor electrical position, as shown in Table 1 (see, e.g., [41]).
Table 1. DC bus connection sequence of phase windings with respect to rotor electrical angle.
Figure 3b shows the phase voltage waveforms for the case of PWM modulation
of phase voltages and generator operation of the brushless DC machine. According to
Figure 3a,b, i2 = –i1 = ieq is valid for the generator operating regime. Hence, the BLDC
machine may be viewed as a DC machine during this particular time frame. The equivalent
DC machine model armature circuit is characterized by the following values of equivalent
inductances and resistances given [41]:
eeq = Keq ω g = 2Ke φmn ω g , τg = Keq ieq = 2Ke φmn ieq , (4)
where ieq is the instantaneous line (phase-to-phase) current (Figure 3a), and φmn is the
constant-valued magnetic field flux of the rotor permanent magnets.
Note that the torque and EMF expressions given in Equation (4) are valid for the
case of operation below the rated rotational speed, i.e., when rotor field flux weakening
control is not applied, and which is assumed herein. If air gap field flux weakening is
needed, it is typically carried out by means of armature voltage commutation angle (phase
advance) variation with respect to EMF [47], which would result in values of EMF and
torque constants that are lower than the nominal ones given in (4).
The equivalent line voltage ur (i.e., voltage across two stator phases) is obtained by
pulse-width modulation (PWM) of the DC bus voltage with the suitably chosen switching
frequency (fsw = 1/Tsw ). Its average (mean) value can be expressed by using the DC bus
voltage magnitude udc and the PWM duty cycle (0 ≤ d ≤ 1) as follows [46]:
−
u r = (2d − 1)udc . (5)
Based on the above relationships, the BLDC machine stator winding model may be
expressed in the following form in order to obtain the averaged armature model during
simultaneous conduction of two phases:
dieq −
Leq + Req ieq = u r − eeq . (6)
dt
According to Figure 3b and analysis presented in [36], the brushless DC machine
equivalent armature current (line current) ieq is related to the active rectifier current ir at
the DC bus side as follows:
ir = (2d − 1)ieq . (7)
Energies 2021, 14, 7125 7 of 26
Based on the above equivalent DC bus current formulation, the DC bus voltage
dynamic equation is given as (Figure 3a):
1
Z
udc = − (ir + irL )dt, (8)
Cdc
with irL being the DC bus load current (i.e., feeding the propeller drives and other loads).
The above relatively simple analysis yields the equivalent DC model of the brushless
DC machine and active rectifier connected to the common DC bus shown in Figure 4, which
is valid for operation below the rated speed. Apart from the armature (stator winding)
phase-to-phase (line) resistance and inductance Req and Leq (calculated according to (3)), the
model also includes the back electromotive force and developed mechanical torque of the
equivalent DC machine model, which are calculated according to Equation (4). Thus, the
mechanical part of the model (generator mechanical torque and rotational speed) can be
easily integrated with the engine model in Figure 2 through the gearbox ratio ig , as shown
in Figure 4.
Figure 4. The averaged equivalent DC model of the brushless DC machine with simplified model of
connection to the DC bus via a PWM-controlled switch-mode power converter (active rectifier).
with Te being the closed-loop system equivalent time constant, and D2 , D3 , . . . , Dn being
the so-called damping optimum characteristic ratios.
In the so-called “optimal” case Di = 0.5 (i = 2 . . . n), the closed-loop system step
response (for any closed-loop system order n) is characterized by an overshoot of approxi-
mately 6% (thus emulating a second-order system behavior with the damping ratio ζ = 0.707)
with the approximate step response rise time tr = 2·Te . In general, larger Te values corre-
Energies 2021, 14, 7125 8 of 26
spond to improved control system robustness and noise suppression ability, albeit with
slower closed-loop response, and vice versa.
From the standpoint of designing of a superimposed controller, the inner closed-loop
system tuned according to the damping optimum criterion may be approximated by the
equivalent first-order lag transfer function characterized by the closed-loop equivalent
time constant Te (under assumption of unit gain of the inner closed loop model):
1
Ge (s) = . (10)
Te s + 1
Figure 5. Block-diagram representation of the Luenberger estimator for rotary speed estimation
based on brushless DC armature equivalent DC model.
Energies 2021, 14, 7125 9 of 26
Based on the known (and readily available) PWM switching command duty cycle
reference dR , the active rectifier output voltage estimate urm can be reconstructed based on
Equation (5) and used as estimator input, thus finally yielding the following Luenberger
estimator state-space formulation (Figure 5):
dı̂eq /dt − Req /Leq −1/Leq ı̂eq 1/Leq Kei
= + urm + ıeqm − ı̂eq , (12)
dêeq /dt 0 0 êeq 0 −Kee
The above state-space estimator representation can be transformed into its Laplacé
s-domain transfer function counterpart as follows:
i g urm (s) − Leq s + Req ieqm (s)
ω̂ (s) = , (13)
Keq Ao (s)
with the estimator transfer function characteristic polynomial Ao (s) given as:
Thus, the above estimator model outputs a scaled and low-pass filtered value of the
“raw” reconstruction of the equivalent electromotive force (EMF) according to Equation (6),
i.e., eeq = ur –Leq dieq /dt–Req ieq , which is additionally filtered by the voltage/current measure-
ment filters with the time constant Tf (Figure 5).
Luenberger estimator is designed by equating the above low-pass filter transfer func-
tion denominator with the second order damping optimum characteristic polynomial
according to Equation (9), which yields the following results for Luenberger estimator
gains Kee and Kie :
Leq 1 Req
Kee = 2
, Kie = − , (15)
D2o Teo D2o Teo Leq
with the following feasibility condition imposed upon the equivalent time constant Teo :
1 Leq
Teo < · . (16)
D2o Req
Figure 6. Block diagram of the linearized engine speed control system with PID feedback controller.
Energies 2021, 14, 7125 10 of 26
Based on the linearized ICE speed control system representation in Figure 6, PID con-
troller design according to the damping optimum criterion yields the following expressions
for the controller parameters (i.e., equivalent time constant Teω , gain KR , and integral and
derivative time constants TI and TD ) [48]:
1 TΣθ + Teo + T f ( Td + Tm ) + Td Tm
Teω ≥ Teωmin = , (17)
D2ω D3ω D4ω TΣθ + Teo + T f + Td + Tm
Jt TΣθ + Teo + T f + Td + Tm
KR = 2 D T2
− Kp, (18)
D2ω 3ω eω Kmt
K p −1
TI = Teω 1 + , (19)
KR
TΣθ + Teo + T f + Td + Tm K
Jt p
TD = − 1 − TΣθ + Teo + T f , (20)
Kmt K R D2ω D3ω Teω KR
Figure 7. Block diagram representation of the brushless DC machine current control system (a) and
DC bus voltage control system (b).
Energies 2021, 14, 7125 11 of 26
The electromotive force (EMF) “disturbance” within the current control loop can be
dealt with by using the readily available EMF estimate from the engine-generator speed
estimator (see Section 3.2), whereas in the case of DC bus load, a Luenberger state estimator
can be used to reconstruct the load current irL from the available measurements of active
rectifier current irm and DC bus voltage udcm :
dûdc /dt 0 −1/Cdc ûdc −1/Cdc Kdce
= + irm + (udcm − ûdc ). (21)
dı̂rL /dt 0 0 ı̂rL 0 −K Le
Again, the above Luenberger estimator can be represented by its transfer function
model, and the resulting DC bus load estimate vs. estimator inputs model has the form of
low-pass filtered “raw” reconstruction of load current irL = –(ir + Cdc dudc /dt) according to
the DC bus model in Equation (8):
[irm (s) + Cdc sudcm (s)] [irm (s) + Cdc sudcm (s)]
ı̂rL (s) = − =− . (22)
A L (s) s2 Cdc + s Cdc Kdce + 1
K Le K Le
Damping optimum criterion yields the final expressions between the DC bus load
estimator gains, DC bus capacitance Cdc and damping optimum parameters (characteristic
ratio D2L and equivalent time constant TeL ):
Cdc 1
K Le = 2
, Kdce = . (23)
D2L TeL D2L TeL
In the BLDC armature current control loop design, the rather small PWM switching
delay and controller sampling delay (if digital controller is considered) are lumped into
the parasitic time constant T Σi (Figure 7a). This lag is augmented by the current sensor
lag Tf in order to obtain the so-called “lumped” first-order lag term with the equivalent
time constant Tpi = T Σi + Tf which approximates the fast closed-loop dynamics. Using
this approximation, the following closed-loop transfer function model (similar to the case
presented in [50]) is obtained, and used in BLDC machine current control system design:
ir ( s ) 1
Gci (s) = = . (24)
irR (s) Tpi Leq Tci s3 ( Req Tpi + Leq )Tci s2 ( Req +Kci )Tci s
Kci + Kci + Kci +1
By applying the damping optimum criterion, the following results for current con-
troller parameters are obtained:
1 Tpi
Tei ≥ Tei,min = , (25)
D2i D3i 1 + Tpi Req /Leq
!
D2i Tei
Tci = Tei 1 − , (26)
Tpi + Leq /Req
equivalent time constant Tei and the controller sampling time T (if digital controller is
used). This results in the overall first-order lag term with the time constant Tpu = T Σu + Tf
describing the “fast” voltage control loop dynamics, and the corresponding closed-loop
system transfer function model is given as follows:
udcm (s) 1
Gudc (s) = = . (28)
udcR (s) Cdc ( TΣu + T f ) Tcu 3 Cdc Tcu 2
Kcu s + Kcu s + Tcu s + 1
By applying the damping optimum criterion, the following expressions for DC bus
voltage PI controller parameters are obtained (see, e.g., [49]):
TΣu + T f
Tcu = Teu = , (29)
D2u D3u
Cdc
Kcu = . (30)
D2u Teu
Finally, the feed-forward compensator (lead-lag filter) is designed with the aim of
canceling out the main dynamics of the inner current control loop (i.e., its equivalent
lag Tei ). The zero-pole canceling approach yields the so-called “lead” time constant
TF = Tei with the filtering pole sF = 1/(αTF ) scaling factor α typically chosen in the range
α = 0.1 . . . 0.6.
After the aforementioned modeling error is included within the closed-loop transfer
function model (24), the closed-loop characteristic polynomial reads as follows:
Tpi Leq Tci s3 R∗eq Tpi + Leq Tci s2 R∗eq + Kci Tci s
Aci (s) = + + + 1. (32)
Kci Kci Kci
which can be used to illustrate the armature resistance variation effects to the closed-loop
damping (and stability) by means of root-locus plots.
The resulting root locus plots for the relative armature resistance variations ∆Req /Req
from −25% to +50% are shown in Figure 8, with these boundary cases represents a realistic
scenario in practical applications, as indicated in [48]. As indicated by closed-loop pole
locations in Figure 8, if armature resistance is increased above the nominal value, this
results in increased closed-loop damping, whereas armature resistance decrease may
result in decreased closed-loop damping. The latter scenario is usually related to ambient
temperature decrease and would typically have less effect during brushless DC generator
operation due to unavoidable machine internal heat losses. Thus, it may be surmised that
Energies 2021, 14, 7125 13 of 26
favorable closed-loop damping of BLDC generator armature current control system should
be maintained for the particular realistic range of armature resistance variations.
Figure 8. Root locus plots of the BLDC generator current control system dominant closed-loop poles
subject to armature resistance variations.
As shown in [49], the feed-forward load compensator action subject to current and
voltage sensor gain errors according to Equations (33) and (34) can be divided into the
“nominal” feed-forward action and the “parasitic” sensor error-related dynamics. The latter
may be considered as additional virtual feedback paths of current ir and voltage udc signals
Energies 2021, 14, 7125 14 of 26
(see Figure 9) which may affect the closed-loop robustness (i.e., damping) [49]. Voltage
offset and gain errors may also result in steady-state control error, while the current offset
error (∆irm ) would be suppressed by the integral action of the DC bus voltage PI controller.
Figure 9. Block diagram of the DC bus voltage control system subject to voltage and current
sensor errors.
Root locus plots of the DC bus voltage closed-loop control system subject to volt-
age/current sensor gain errors are shown in Figure 10. The case of ±10% gain errors is
considered herein as a worst-case scenario, as suggested in [49]. As expected, the closed-
loop pole locations are shifted from the nominal positions (corresponding to zero sensor
gain errors), with dominant pole locations being characterized by less well-damped closed-
loop pole locations for the case of negative gain errors (εi = εu = −0.1). The less dominant
closed-loop poles behave in an opposite manner, i.e., they are shifted towards less well-
damped locations (i.e., towards the origin of the s-plane) in the case of positive sensor gain
errors (εi = εu = +0.1), but they remain well-damped for the considered hypothetical range
of sensor gain errors.
Figure 10. Root locus plots of DC bus voltage control system dominant closed-loop poles subject to
voltage/current sensor gain errors.
Energies 2021, 14, 7125 15 of 26
4.3. Engine Speed Control System Robustness to Torque Gain and Manifold Lag Errors
It is assumed that the engine torque gain Kmt and manifold time constant (lag) Tm of
the linearized engine model (Figure 2b) are subject to additive errors ∆Kmt and ∆Tm , thus
resulting in the following torque gain and time constant K* mt and T* m values:
∗
Kmt = Kmt + ∆Kmt , (36)
ω (s) 1 1
Gc (s) = = = , (38)
ω R (s) Ac (s) 1 + aω1 s + aω2 s + aω3 s3 + aω4 s4 + aω5 s5
2
where:
Kp
aω1 = 1+
TI , (39)
KR
h i
∗ K T +K T +T +T
Jt + Kmt R D p Σθ f eo
aω2 = ∗ K TI , (40)
Kmt R
Based on the above closed-loop system model, Figure 11 shows the closed-loop pole
locations for the torque gain Kmt and time constant Tm relative errors of 50% and 100%.
The root locus analysis results are obtained for engine speed PID controller tuned with
Teω = Teωmin (Equation (17)), which relates to a fast and well-damped response of the
engine speed control loop. The results in Figure 11a show that the intake manifold time
constant error causes the dominant conjugate-complex pole pair to be shifted towards
the origin of the s-plane. A similar result is also obtained for linearized engine model
torque gain parameter variation (Figure 11b), with the dominant conjugate-complex poles
being characterized by decreased damping and larger imaginary parts. However, in both
cases the closed-loop pole damping ratio is kept above ζ = 0.5, which points out to rather
favorable robustness of the proposed tuning approach.
Figure 11. Locations of dominant closed-loop poles of the linearized ICE speed control loop subject to manifold time
constant error (a) and torque gain error (b).
If the aforementioned relationships Equations (31), (44) and (45) are taken into account
within the overall Luenberger estimator model (Equation (13)), the following dynamic
model of generator speed estimation is obtained after some manipulation and rearranging
(cf. also Figure 5):
ω (s) + ∆ω (s) + ωo f f
ω̂ (s) = , (46)
Ao (s) T f s + 1
where the constant-valued speed measurement (estimation) offset ω off and speed estimation
dynamic error ∆ω(s) are given as follows:
ig
∆urm − Req + ∆Req ∆ieq ,
ωo f f = (47)
Keq
ig h − i
∆ω (s) = ε u u r (s) − Leq s + Req ε i ieqm (s) − ∆Req (1 + ε i )ieqm (s) .
(48)
Keq
Obviously, the pure offset ω off of engine-generator set speed estimation is caused by
armature current and voltage measurement offset errors and armature resistance variations,
ultimately resulting in engine-generator set closed-loop steady-state error. On the other
hand, the speed estimation error component ∆ω(s) due to voltage/current measurement
gain errors may affect the engine speed control system closed-loop performance during
transients and in the engine generator set steady-state. These speed estimation errors are
examined in more detail by means of simulations in the next section.
5. Simulation results
5.1. Simulation Model Parameterization
The proposed control systems for the presented hybrid power supply are verified by
means of simulations carried out in MATLAB/Simulink. Simulation models are parameter-
Energies 2021, 14, 7125 17 of 26
ized based on data from [31,34,53–56], with final parameters of linearized process models
listed in Table 2. MVEM model maps used for ICE simulation are obtained by re-scaling
the engine maps from [31,34] using the methodology proposed in [57], and these maps
are shown in Figure 12. The values of parameters of individual controllers and estimators
presented in this work and used throughout the simulation study are also listed in Table 2.
Table 2. Parameters of linearized process models, estimators and controllers used in simulations.
Simulation results of the engine-generator set speed control system for the case of
brushless DC generator armature resistance mismatch and armature current and voltage
measurement (sensor) errors are shown in Figure 14. These results cover the worst-case
scenarios in terms of closed-loop system robustness analyzed in the previous section
Energies 2021, 14, 7125 19 of 26
(i.e., ±10% voltage/current sensor gain errors εu and εi and -armature resistance varia-
tion from the nominal value ∆Req /Req = 25% are considered), along with relatively small
current/voltage measurement offset errors (∆urm = 0.2 V, and ∆irm = 0.2 A) The results
in Figure 14 point out that the anticipated range of process model parameter variations
introduces notable steady-state engine control error, whereas negative speed offset is ob-
tained for positive sensor gain errors (Figure 14a), while positive speed offset is associated
with negative gain error values (Figure 14b). Since only positive voltage/current offset
errors are introduced in the simulation, their effect is visible in the magnitude of the actual
engine-generator set speed offset from the target value, wherein larger speed control errors
are obtained for the same sign of sensor gain and offset error, and vice versa (cf. top
plots in Figure 14a,b). The effect of voltage/current sensor gain error is also manifested
in the closed-loop system speed transient damping after a sudden load torque change,
wherein positive gain errors (εu = εi = 0.1) decrease the level of closed-loop damping, which
is characterized by somewhat larger load response overshoot compared to the nominal
case (cf. middle and bottom plots of developed engine torque and throttle command in
Figures 13a and 14a), whereas negative gain errors (εu = εi = −0.1) tend to increase the level
of closed-loop damping (cf. Figures 13a and 14b).
Figure 13. Closed-loop responses of the proposed hybrid propulsion control system for abrupt DC
bus load change (nominal case–no modeling and sensor errors): engine quantities (a) and DC bus
quantities (b).
the robustness analysis in the previous section, voltage offset and gain errors result in
steady state DC bus control error (cf. top plots in Figure 15a,b). Again, larger control error
absolute values are obtained for the same sign of sensor gain and offset error, and vice versa.
Moreover, the current and voltage reconstruction errors may also affect the steady-state
accuracy of the load current estimation based on Luenberger estimator, as shown in bottom
plots in Figure 15a,b. However, due to integral action of the DC bus PI controller, the
active rectifier current is correctly commanded in order to maintain the steady-state DC
bus voltage (nonetheless being affected by the accuracy of DC bus voltage measurement).
The effects of sensor gain errors to closed-loop damping match the findings of the closed-
loop system robustness analysis, wherein closed-loop damping level is decreased for
negative sensor gain errors (εi = εu = −0.1) compared to nominal case (cf. top plots in
Figures 13b and 15a), as opposed to the case of positive sensor gain errors (εi = εu = 0.1),
see top plot in Figure 15b.
Figure 14. Closed-loop responses of engine-generator set speed control system based on EMF
estimation subject to armature resistance error ∆Req /Req = −0.25, sensor offset errors ∆urm = 0.2 V
and ∆ieqm = 0.2 A, and sensor gain errors: εi = εu = 0.1 (a), and εi = εu = −0.1 (b).
Energies 2021, 14, 7125 21 of 26
Figure 15. Closed-loop responses of the DC bus voltage control system subject to sensor offset errors
∆udcm = 0.2 V and ∆irm = 0.2 A and sensor gain errors: εi = εu = 0.1 (a), and εi = εu = −0.1 (b).
6. Discussion of Results
Robustness analysis of the overall control system to parameters variations within
process models (i.e., parameters of the controlled hybrid power-plant) has yielded the
following findings:
(i) Brushless DC machine armature current control system should be fairly robust to
armature resistance variations over the expected range of its variations;
(ii) DC bus control system sensitivity to voltage and current sensor gain and offset error
may manifest in closed-loop voltage control error, with possibly decreased level of
damping of the dominant closed-loop poles in the case of negative sensor gain errors;
(iii) The engine-generator set speed control system should be robust to a relatively large
change of manifold lag and equivalent engine torque gain parameter.
(iv) The generator current and voltage sensor errors may affect both the steady-state
and transient accuracy of the engine-generator speed estimation. The steady-state
estimation error is solely affected by the current/voltage sensor offset errors and
generator armature resistance mismatch with respect to its nominal value.
The results of comprehensive simulation analysis have shown good load disturbance
ability of the overall control system in terms of fast and well damped control system
responses to sudden DC bus load changes, along with favorable robustness to hybrid
power-plant process model parameter variations, in particular:
(i) The engine speed control system with PID controller is capable of suppressing the
load disturbance within 1 s, with only a moderate engine speed drop of 15.6% from
the target value of 4500 rpm;
(ii) The DC bus voltage/current control system has been characterized by 80 ms recovery
time and 200 ms settling time after the sudden DC bus load change, and is also
characterized by a non-emphasized drop in the DC bus voltage (10.4% of the target
value of 48 V);
(iii) The engine-generator set speed control system based on Luenberger estimator of
brushless DC machine electromotive force estimation may be affected by the armature
Energies 2021, 14, 7125 22 of 26
7. Conclusions
The paper has presented the detailed control system design and robustness analysis
for the hybrid propulsion system suitable for unmanned aerial vehicles, which is based
on internal combustion engine plus brushless DC generator set power supply of the
common DC bus used for power distribution within the aircraft. The overall control
system has featured (i) the internal combustion engine speed control system based on a
PID feedback controller and (ii) the brushless DC generator active rectifier voltage/current
control based on PI feedback controllers, with both feedback loops also featuring load
disturbance estimators based on the Luenberger estimator methodology. The design of
feedback control systems and estimators has been based on damping optimum criterion
which yields straightforward analytical expressions for controller and estimator parameters.
The robustness of the proposed control systems to process model parameter variations
has been analyzed by closed-loop root locus plots, which have indicated that favorable
closed-loop damping obtained through controller/estimator tuning according to damping
optimum criterion ought to be preserved for the anticipated range of modeling errors.
The effectiveness of the proposed hybrid propulsion control system suitable for UAV
applications has been verified by means of comprehensive simulations. Results have
pointed out that the overall control system is characterized by rather fast and effective
recovery with respect to load disturbance from the common DC bus, with the “slower”
engine speed control system being characterized by approximately 1 s long engine speed
recovery transient, whereas the “faster” DC bus voltage control system is capable of
recovering the DC bus voltage in approximately 80 ms, with both control loops suffering
only moderate excursions from their respective set-point (reference) values. These control
system characteristics have been achieved due to accurate and fast estimation of key hybrid
propulsion system variables, i.e., engine-generator set speed and electromotive force, and
DC bus load, thus enabling effective suppression of control system external disturbances.
Moreover, the simulation analysis has also largely confirmed the results of robustness
analysis in terms of closed-loop systems maintaining their favorable closed-loop damping
properties. Finally, the simulation study has also pointed out to the existence of closed-loop
steady-state control errors in both engine speed and DC bus voltage when voltage and
current sensor gain and offset errors are present. These errors still have limited magnitudes
which are primarily related to magnitudes of sensor offset errors, which directly affect the
engine-generator set speed estimation accuracy within the Luenberger estimator and the
steady-state accuracy of the DC bus voltage feedback.
Future work may involve designing the upper-level supervisory control strategy
aimed at hybrid power-plant energy management and power flow control and building a
down-scaled laboratory setup of the hybrid propulsion system for the purpose of exper-
Energies 2021, 14, 7125 23 of 26
imental verification of individual control system components and overall hybrid power
system control strategy.
Author Contributions: Conceptualization, M.K. and D.P.; methodology, D.P., M.K., M.C. and J.B.;
software, D.P. and M.C.; validation, D.P. and M.K.; formal analysis, D.P.; investigation, D.P. and M.K.;
resources, M.C.; writing—original draft preparation, D.P.; writing—review and editing, M.K., D.P.,
J.B. and M.C.; visualization, D.P. and M.K.; supervision, M.C.; project administration, D.P.; funding
acquisition, D.P. All authors have read and agreed to the published version of the manuscript.
Funding: It is gratefully acknowledged that this research has been supported by the EU European
Regional Development Fund under the Grant KK.01.1.1.01.0009 (DATACROSS).
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Research data can be made available for non-profit use in research and
education upon a formal request made to the corresponding author.
Acknowledgments: We would like to pay our gratitude and our respects to our colleague, Davor
Zorc, who has actively supported this research until his untimely passing away in June of 2020. He
was a tenured professor at the Department of Robotics and Production System Automation at the
Faculty of Mechanical Engineering and Naval Architecture, University of Zagreb, Croatia.
Conflicts of Interest: The authors declare no conflict of interest. The funding body had no role in
the design of the study; in the collection, analyses, or interpretation of data; in the writing of the
manuscript, or in the decision to publish the results.
Nomenclature
Abbreviations:
AC Alternating current
BLDC Brushless direct current (machine/generator)
BPTT Back propagation through time (gradient optimization)
DC Direct current
DP Dynamic programming
ECMS Equivalent consumption minimization strategy
EMF Electromotive force
ICE Internal combustion engine
PI Proportional-integral (controller)
PID Proportional-integral-derivative (controller)
PWM Pulse-width modulation
rpm Revolutions per minute
UAV Unmanned aerial vehicle
D1 . . . D6 Flywheeling diodes
Q1 . . . Q6 MOSFET switches
Variables:
eeq BLDC machine DC model equivalent electromotive force
el BLDC machine electromotive force per phase
dR , d Active rectifier duty cycle reference and actual duty cycle value
i1 , i2 , i3 Instantaneous values of BLDC machine phase currents
ieq , ieqm BLDC machine DC model equivalent current and its measurement
ir , irL Active rectifier output current and DC bus load current
Iph BLDC machine phase current magnitude
pm ICE manifold pressure
s Laplace operator
udc , udcm DC bus voltage and its measurement value
ur Active rectifier output voltage (line voltage of two BLDC phases)
Wi , Wo ICE manifold intake and output air mass flow
Energies 2021, 14, 7125 24 of 26
References
1. Maza, I.; Caballero, F.; Capitán, J.; Martínez-De-Dios, J.R.; Ollero, A. Experimental results in multi-UAV coordination for disaster
management and civil security applications. J. Intell. Robot. Syst. 2011, 61, 563–585. [CrossRef]
2. Kingston, D.; Beard, R.W.; Holt, R.S. Decentralized perimeter surveillance using a team of UAVs. IEEE Trans. Robot. 2008, 24,
1394–1404. [CrossRef]
3. Adams, S.M.; Levitan, M.L.; Friedland, C.J. High resolution imagery collection for post-disaster studies utilizing unmanned
aircraft systems (UAS). Photogramm. Eng. Remote Sens. 2014, 80, 1161–1168. [CrossRef]
4. Ayele, Y.Z.; Aliyari, M.; Griths, D.; Droguett, E.L. Automatic crack segmentation for UAV-assisted bridge inspection. Energies
2020, 13, 6250. [CrossRef]
5. Deng, C.; Wang, S.; Huang, Z.; Tan, Z.; Liu, J. Unmanned aerial vehicles for power line inspection: A cooperative way in platforms
and communications. J. Commun. 2014, 9, 687–692. [CrossRef]
6. Huang, Y.; Hoffmann, W.C.; Lan, Y.; Wu, W.; Fritz, B.K. Development of a spray system for an unmanned aerial vehicle platform.
Appl. Eng. Agric. 2014, 25, 803–809. [CrossRef]
7. Boukoberine, M.N.; Zhou, Z.; Benbouzid, M. A critical review on unmanned aerial vehicles power supply and energy manage-
ment: Solutions, strategies, and prospects. Appl. Energy 2019, 255, 113823. [CrossRef]
8. Pang, T.; Peng, K.; Lin, F.; Chen, B.M. Towards long-endurance flight: Design and implementation of a variable-pitch gasoline-
engine quadrotor. In Proceedings of the IEEE International Conference on Control and Automation (ICCA 2016), Kuching,
Malaysia, 23–27 May 2016; pp. 767–772.
9. Sheng, S.; Sun, D. Control and optimization of a variable-pitch quadrotor with minimum power consumption. Energies 2016, 9,
232. [CrossRef]
10. Hung, J.Y.; Gonzalez, L.F. On parallel hybrid-electric propulsion system for unmanned aerial vehicles. Prog. Aerosp. Sci. 2012, 51,
1–17. [CrossRef]
11. Lu, W.; Zhang, D.; Zhang, J.; Li, T.; Hu, T. Design and implementation of a gasoline-electric hybrid propulsion system for a micro
triple tilt-rotor VTOL UAV. In Proceedings of the 6th Data Driven Control and Learning Systems Conference, Chongqing, China,
16 October 2017; pp. 433–438.
12. Fredericks, W.J.; Moore, M.D.; Busan, R.C. Benefits of hybrid-electric propulsion to achieve 4x cruise efficiency for a VTOL UAV.
In Proceedings of the 2013 International Powered Lift Conference, Los Angeles, CA, USA, 12–14 August 2013; p. 21.
13. Cipek, M.; Pavković, D.; Petrić, J. A control-oriented simulation model of a power-split hybrid electric vehicle. Appl. Energy 2013,
101, 121–133.
14. Abdul Sathar Eqbal, M.; Fernando, N.; Marino, M.; Wild, G. Hybrid propulsion systems for remotely piloted aircraft systems.
Aerospace 2018, 5, 34. [CrossRef]
15. Friedrich, C.; Robertson, P.A. Hybrid-electric propulsion for aircraft. J. Aircr. 2015, 52, 176–189. [CrossRef]
16. Finger, D.F.; Braun, C.; Bil, C. A Review of Configuration Design for Distributed Propulsion Transitioning VTOL Aircraft.
In Proceedings of the 2017 Asia-Pacific International Symposium on Aerospace Technology (APISAT 2017), Seoul, Korea,
16–18 October 2017; p. 15.
17. Recoskie, S.; Fahim, A.; Gueaieb, W.; Lanteigne, E. Hybrid power plant design for a long-range dirigible UAV. IEEE Trans.
Mechatron. 2013, 19, 606–614. [CrossRef]
18. Recoskie, S.; Fahim, A.; Gueaieb, W.; Lanteigne, E. Experimental testing of a hybrid power plant for a dirigible UAV. J. Intell.
Robot. Syst. 2013, 69, 69–81. [CrossRef]
19. Depcik, C.; Cassady, T.; Collicott, B.; Burugupally, S.P.; Li, X.; Alam, S.S.; Hobeck, J. Comparison of lithium ion batteries, hydrogen
fueled combustion engines, and a hydrogen fuel cell in powering a small Unmanned Aerial Vehicle. Energy Convers. Manag. 2020,
207, 112514. [CrossRef]
20. Benić, Z.; Krznar, M.; Kotarski, D. Mathematical modelling of unmanned aerial vehicles with four rotors. Interdiscip. Descr.
Complex Syst. 2016, 14, 88–100. [CrossRef]
21. Wang, X.; Liu, J.; Zhang, Y.; Shi, B.; Jiang, D.; Peng, H. A unified symplectic pseudospectral method for motion planning and
tracking control of 3D underactuated overhead cranes. Int. J. Robust Nonlinear Control 2019, 29, 1–18. [CrossRef]
22. Peng, H.; Li, F.; Liu, J.; Ju, Y. A symplectic instantaneous optimal control for robot trajectory tracking with differential-algebraic
equation models. IEEE Trans. Ind. Electron. 2020, 67, 3819–3829. [CrossRef]
23. Li, F.; Peng, H.; Song, X.; Liu, J.; Tan, S.; Ju, Z. A physics-guided coordinated distributed MPC method for shape control of an
antenna reflector. IEEE Trans. Cybern. 2021, 1–13. [CrossRef]
24. Guemri, M.; Neffati, A.; Caux, S.; Ngueveu, S.U. Management of distributed power in hybrid vehicles based on D.P. or Fuzzy
Logic. Optim. Eng. 2014, 15, 993–1012. [CrossRef]
25. Cipek, M.; Kasać, J.; Pavković, D.; Zorc, D. A novel cascade approach to control variables optimisation for advanced series-parallel
hybrid electric vehicle power-train. Appl. Energy 2020, 276, 115488. [CrossRef]
26. Esser, A.; Eichenlaub, T.; Schleiffer, J.E.; Jardin, P.; Rinderknecht, S. Comparative evaluation of powertrain concepts through an
eco-impact optimization framework with real driving data. Optim. Eng. 2021, 22, 1001–1029. [CrossRef]
27. Sinoquet, D.; Rousseau, G.; Milhau, Y. Design optimization and optimal control for hybrid vehicles. Optim. Eng. 2011, 12, 199–213.
[CrossRef]
Energies 2021, 14, 7125 26 of 26
28. Škugor, B.; Deur, J.; Cipek, M.; Pavković, D. Design of a power-split hybrid electric vehicle control system utilizing a rule-based
controller and an equivalent consumption minimization strategy. Proc. Inst. Mech. Eng. Part. D J. Automob. Eng. 2014, 228,
631–648. [CrossRef]
29. Jimenez, P.; Lichota, P.; Agudelo, D.; Rogowski, K. Experimental validation of total energy control system for UAVs. Energies
2020, 13, 14. [CrossRef]
30. Ulsoy, A.G.; Peng, H.; Çakmakci, M. Automotive Control Systems; Cambridge University Press: New York, NY, USA, 2012;
pp. 126–130.
31. Deur, J.; Ivanović, V.; Pavković, D.; Jansz, M. Identification and speed control of si engine for idle operating mode. In Proceedings
of the SAE 2004 World Congress, Detroit, MI, USA, 8–11 March 2004; SAE Technical Paper No. 2004-01-0898. p. 12.
32. Langouët, H.; Métivier, L.; Sinoquet, D.; Tran, Q.H. Engine calibration: Multi-objective constrained optimization of engine maps.
Optim. Eng. 2011, 12, 407–424. [CrossRef]
33. Lin, C.E.; Supsukbaworn, T.; Huang, Y.C.; Jhuang, K.T.; Li, C.H.; Lin, C.Y.; Lo, S.Y. Engine controller for hybrid powered dual
quad-rotor system. In Proceedings of the 41st Annual Conference of the IEEE Industrial Electronics Society, Yokohama, Japan,
9–12 November 2015; pp. 1513–1517.
34. Pavković, D.; Deur, J.; Kolmanovsky, I. Adaptive Kalman filter-based load torque compensator for improved SI engine idle speed
control. IEEE Trans. Control Syst. Technol. 2009, 17, 98–110. [CrossRef]
35. Schröder, D. Elektrische Antriebe-Regelung von Antriebssystemen; Springer: Berlin/Heidelberg, Germany, 2007; pp. 814–865.
36. Carev, V.; Roháč, J.; Šipoš, M.; Schmirler, M. A multilayer brushless DC motor for heavy lift drones. Energies 2021, 14, 2504.
[CrossRef]
37. Pajchrowski, T.; Krystkowiak, M.; Matecki, D. Modulation variants in DC circuits of power rectifier systems with improved
quality of energy conversion—Part I. Energies 2021, 14, 1876. [CrossRef]
38. Krznar, M.; Piljek, P.; Kotarski, D.; Pavković, D. Modeling, control system design and preliminary experimental verification of a
hybrid power unit suitable for multirotor UAVs. Energies 2021, 14, 2669. [CrossRef]
39. Zhang, X.Z.; Wang, Y.N. A novel position-sensorless control method for brushless DC motors. Energy Convers. Manag. 2011, 52,
1669–1676. [CrossRef]
40. Shao, J.; Nolan, D.; Hopkins, T. A novel direct back EMF detection for sensorless brushless DC (BLDC) motor drives. In
Proceedings of the 17th Annual IEEE Applied Power Electronics Conference and Exposition (APEC 2002), Dallas, TX, USA,
10–14 March 2002; pp. 33–37.
41. Krznar, M.; Kotarski, D.; Pavković, D.; Piljek, P. Propeller speed estimation for unmanned aerial vehicles using Kalman filtering.
Int. J. Autom. Control. 2020, 14, 284–303. [CrossRef]
42. Pavković, D.; Krznar, M.; Cipek, M.; Zorc, D.; Trstenjak, M. Internal combustion engine control system design suitable for hybrid
propulsion applications. In Proceedings of the ICUAS 2020 Conference, Athens, Greece, 1–4 September 2020; pp. 1614–1619.
43. Luenberger, D.G. Observing the state of a linear system. IEEE Trans. Mil. Electron. 1964, 8, 74–80. [CrossRef]
44. Naslin, P. Essentials of Optimal Control; Illife Books: London, UK, 1968; Chapter 2.
45. Mathew, T.; Sam, C.A. Modeling and closed-loop control of BLDC motor using an single current sensor. Int. J. Adv. Res. Electr.
Eng. Instrum. Eng. 2013, 2, 2525–2531.
46. Williams, B.W. Power Electronics: Devices, Drivers, Applications and Passive Components; McGraw-Hill: New York, NY, USA, 1992;
pp. 772–789.
47. Kong, H.; Liu, J.; Cui, G. Study on Field-Weakening Theory of Brushless DC Motor Based on Phase Advance Method. In
Proceedings of the 2010 International Conference on Measuring Technology and Mechatronics Automation, Changsha, China,
3–14 March 2010; pp. 583–586.
48. Pavković, D.; Deur, J. Modeling and Control of Electronic Throttle Drive: A practical Approach-from Experimental Characterization to
Adaptive Control and Application; Lambert Academic Publishing: Saarbrücken, Germany, 2011; pp. 123–131.
49. Pavković, D.; Lobrović, M.; Hrgetić, M.; Komljenović, A. A Design of Cascade Control System and Adaptive Load Compensator
for Battery/Ultracapacitor Hybrid Energy Storage-based Direct Current Microgrid. Energy Convers. Manag. 2016, 114, 154–167.
[CrossRef]
50. Pavković, D.; Cipek, M.; Kljaić, Z.; Mlinarić, T.J.; Hrgetić, M.; Zorc, D. Damping Optimum-Based Design of Control Strategy
Suitable for Battery/Ultracapacitor Electric Vehicles. Energies 2018, 11, 2854. [CrossRef]
51. Pavković, D.; Šprljan, P.; Cipek, M.; Krznar, M. Cross-axis control system design for borehole drilling based on damping optimum
criterion and utilization of proportional-integral controllers. Optim. Eng. 2021, 22, 51–81. [CrossRef]
52. Gu, J.; Ouyang, M.; Li, J.; Lu, D.; Fang, C.; Ma, Y. Driving and braking control of PM synchronous motor based on low-resolution
Hall sensor for battery electric vehicle. Chin. J. Mech. Eng. 2013, 26, 1–10. [CrossRef]
53. Krznar, M. Modelling and Control of Hybrid Propulsion Systems for Multirotor Unmanned Aerial Vehicles. Ph.D. Dissertation,
Faculty of Mechanical Engineering and Naval Architecture, University of Zagreb, Zagreb, Croatia, 2020.
54. FoxTech Co. Available online: https://www.foxtechfpv.com/t-motor-u12ii-kv120.html (accessed on 5 July 2021).
55. DSS2x101-02A High Performance Schottky Diode; Data Sheet No. 20110603A; IXYS Corp.: Leiden, The Netherlands, 2011.
56. Titan ZG 45PCI-HV. Technical Manual; Tony Clark GmbH: Luebbecke, Germany, 2012.
57. Cipek, M.; Petrić, J.; Pavković, D. A Novel Approach to Hydraulic Drive Sizing Methodology and Efficiency Estimation based on
Willans Line. J. Sustain. Dev. Energy Water Environ. Syst. 2019, 7, 155–167. [CrossRef]