Modeling, Parameterization and Damping Optimum

0 downloads 0 Views 3MB Size Report
These have pointed out to precise coordination of the winch electrical servodrive ... module-related rope force control system and sustained power production during ... interacts with the ground station through tether tension force, which also ..... Figure 4. Efficiency map (a) and power losses map (b) of motor/generator and ...
1

Modeling, Parameterization and Damping Optimum-based Control System

2

Design for an Airborne Wind Energy Ground Station Power Plant

3

Danijel Pavković†*, Mihael Cipek†,1 , Mario Hrgetić†, Almir Sedi㇠†

4

Faculty of Mechanical Engineering and Naval Architecture, University of Zagreb, Zagreb, Croatia ‡

5 6

INA – Industrija nafte, d.d., Zagreb, Croatia

emails: [email protected]; [email protected]; [email protected]; [email protected]

7 8

ABSTRACT

9

This paper presents the results of modeling and parameterization of the high-altitude wind energy system

10

ground station power-plant equipped with a generator/motor unit as a primary power source tethered to the

11

airborne module via a winch system, an ultracapacitor energy storage system, and grid inverter connected to the

12

common direct-current link. Consequently, a suitable ground station power plant control strategy is designed,

13

comprising the generator/motor speed control and cable tension control system, direct-current link power flow

14

coordination control, and grid-side inverter control strategy. Control system design is exclusively based on the

15

damping optimum criterion which provides a straightforward way of closed-loop damping tuning. The

16

effectiveness of the proposed ground station control strategy is verified by means of comprehensive computer

17

simulations. These have pointed out to precise coordination of the winch electrical servodrive with the airborne

18

module-related rope force control system and sustained power production during the airborne module ascending

19

phase in the presence of high-altitude wind disturbances, and continuous power delivery to the grid-side inverter,

20

facilitated by the utilization of ultracapacitor energy storage. This indicates rather robust behavior of the overall

21

ground station control system under anticipated external disturbance conditions.

22 23

KEYWORDS

24 25

High-altitude wind energy; ultracapacitor energy storage; grid-tied inverter; power flow control strategy; power source coordination

26 27

1

Corresponding author 1

1

1. INTRODUCTION

2

Even though the possibility of harnessing of the relatively steady high-altitude/high-speed wind power has

3

been continually studied since the early 1980’s (see e.g. [1]), it has become increasingly attractive over the last

4

decade. This is primarily due to inherent limiting factors of ground-based wind-turbine systems related to the

5

size constraints of the turbine blade and the generator, high investment costs, and relatively unpredictable nature

6

of near-surface winds. One of the key advantages of high-altitude wind energy (HAWE) systems over traditional

7

wind turbine-based systems is that the HAWE system power-plant is located at the ground level, so that the

8

winch machine size and power ratings are no longer an issue. Hence, a number of studies have been carried out

9

up to date, concerning many theoretical aspects of high-altitude wind power system modeling and airborne

10

module (ABM) control, and various practical aspects of airborne module vs. ground station interaction and

11

airborne module trajectory optimization.

12

In particular, reference [2] has shown that detailed modeling of airborne module aerodynamic behavior

13

represents the key prerequisite for the development of suitable ABM guidance strategies and flight control

14

trajectory optimization based on nonlinear model predictive control (MPC) approach. The effectiveness of MPC-

15

based trajectory optimization approach has been subsequently verified in [2] by means of detailed computer

16

simulations and experiments based on a scaled-down high-altitude wind energy system prototype. The airborne

17

module, typically being tethered to the ground station via a winch system and a suitable generator/motor unit,

18

interacts with the ground station through tether tension force, which also mandates a detailed analysis of

19

ABM/tether/winch system, as outlined in [3]. To this end, reference [4] has proposed a multi-segment tether

20

model in order to model the spatially-distributed (so called catenary) shape of the rope in a systematic and

21

straightforward manner, which also inherently includes the tether dynamic behavior and compliance effects.

22

Based on such ABM dynamic models, the airborne unit flight-path-related cycle energy production can be

23

analyzed, as shown in [5], and the energy efficiency of the prospective flight trajectories can be calculated, as

24

illustrated in [6]. Naturally, ABM trajectory (flight path) optimization may also be used for the purpose of on-

25

line maximization of net energy gain [7]. In order to gain the theoretical limit of net energy production, off-line

26

optimization of the airborne module trajectory and the energy production can be based on the solving of the non-

27

linear programming (NLP) problem, as shown in [8]. In particular, the dynamic state equations of the airborne

28

module system and its interaction vs. ground station winch system through the compliant tether have been

29

transformed in [8] into a spatially-discretized grid based on polynomial approximations suitable for NLP

30

optimization problem by using the so-called pseudo-spectral collocation method. Finally, ABM trajectory

2

1

profiles may also be used for the assessment and subsequent selection and optimization of ABM monitoring and

2

control hardware based on the overall system performance requirements [9].

3

An interesting avenue of research in this field has been dedicated to investigation of novel airborne unit

4

designs, such as lighter-than-air turbine systems for fixed-altitude power production [10]. These types of fixed-

5

position high-altitude turbine systems may, in turn, feature highly-specialized turbine configurations and turbine

6

blade designs [11]. Different variable-altitude systems that produce upward lift force by using a parasail-based

7

flying wing configuration have also been considered, either in the form of a single unit [12], or a multiple

8

parasail-unit system [13]. In the latter case, multiple units may provide additional control authority, and

9

continuity of resulting tether pulling force. An alternative approach has been considered in [14], based on

10

positive-buoyancy rotating airborne balloons aimed at exploiting the so-called Magnus’ effect between high-

11

altitude winds and airborne unit rotating body. Alternative uses of high-altitude wind energy harvesting systems,

12

such as those for wind-assisted marine propulsion have been investigated in [15], and their power production

13

potential has been investigated with respect to high-altitude parasail vs. ship’s course and speed. Note also that

14

suitable geographical locations need to be identified before high-altitude wind energy systems are fielded, which

15

suggests certain limitations to the breadth of their implementation [16]. Recently, an emphasis has also been

16

given to optimal choice of airborne wind energy system configurations [17], and respective investigations of

17

fluid mechanics principles governing the dynamic behavior of airborne modules, either realized in the form of

18

variable-altitude units (see e.g. [14]), or those implemented at fixed altitudes [11]. In the latter case, the effect of

19

high-altitude winds intensity and direction variability to wind turbine shell orientation have been analyzed [18],

20

with the aim to define guidelines for stable turbine shell design.

21

Continuous power production may be possible in the case of “stationary” airborne wind turbine [10], and,

22

possibly, in the cases of multi-parasail high-altitude wind energy system [13], and specialized carousel-like

23

ground station configurations [2]. However, many other airborne wind energy harvesting concepts are typically

24

characterized by inherently intermittent power production cycle due to the requirement to eventually spool the

25

tether and return the airborne unit to initial altitude (see e.g. [9]). This is especially emphasized when those

26

systems are used as isolated wind energy production units. In order to mitigate the power production

27

intermittence issues, an energy storage system of sufficiently large storage capacity should be included into the

28

ground-station power-plant [19]. The comprehensive power flow analysis presented in [19] has also shown that

29

ultracapacitor (supercapacitor) energy storage systems (ESS), even though characterized by relatively low

30

energy densities, may still be suitable for HAWE systems characterized by relatively narrow ranges of ABM

3

1

altitude changes (i.e. less than 300 m). This is primarily due to ultracapacitors being characterized by relatively

2

long service and cycle life [20], which would be favorable in the case of rather frequent charging/discharging

3

demands (frequent cable winding/unwinding), likely to be encountered over the aforementioned narrow ranges

4

of HAWE system operating altitudes [19]. In particular, if batteries were used under such frequent energy

5

storage charging/discharging operation, it would likely result in frequent battery replacement, which would

6

increase the investment, operation and maintenance costs, as indicated in [19].

7

In the above literature overview, investigation of different ground station topologies and related modeling and

8

control system design have not been considered. Namely, due to the possibly high level of complexity of the

9

ground station topology, comprising several mutually-interacting mechatronic and power electronics subsystems,

10

it would be advantageous to first derive and parameterize suitable power flow / mechatronic system models.

11

These models could then be used as a basis for the selection of suitable control system structures for individual

12

dynamic subsystems, and subsequent design and verification of dedicated control loops. Moreover, the design of

13

superimposed control system for individual control loops coordination also needs to be performed in a precise

14

and straightforward manner, which further emphasizes the requirement for accurate process modeling. Finally, it

15

would be advantageous if the individual control system design would be based on a straightforward controller

16

tuning methodology. In this way, analytical relationships for power flow control strategy parameters can be

17

provided in relation to key parameters of the controlled plant [21], and the design-specific parameters related to

18

the closed-loop damping [22]. A certain level of robustness of closed-loop damping tuning should also be

19

expected in relation to the anticipated range of process model parameter variations (see e.g. discussion in [23]).

20

Taking into account the aforementioned issues, the hypothesis of this work is that by designing a

21

comprehensive multi-level control strategy, precise and coordinated operation of individual ground stations’

22

constituent components can be achieved, thus facilitating uninterrupted and smooth power delivery to the main

23

electrical grid by using a suitably-sized ultracapacitor energy storage system. Moreover, the utilization of the so-

24

called damping optimum criterion [22] in control system design should facilitate a straightforward and analytical

25

way of tuning of individual control loops, thereby also assuring fast and well-damped behavior of individual

26

ground station control systems.

27

The paper is organized as follows. In Section 2, the considered HAWE system is outlined and the constituent

28

components of its ground station are modeled and parameterized. Section 3 presents the ground station control

29

strategy design, wherein each control subsystem has been designed with the aim of fast and well-damped

30

response characteristics based on the damping optimum criterion. Section 4 presents the comprehensive

4

1

simulation results of HAWE ground station system under steady-state conditions and during wind power

2

disturbances. Concluding remarks and possible avenues for future research are given in Section 5.

3 4

2. HIGH-ALTITUDE WIND ENERGY GROUND STATION TOPOLOGY

5

This section presents the results of modeling and parameterization of the individual HAWE ground station

6

subsystems including the motor/generator servo unit, ultracapacitor energy storage system equipped with appropriate

7

power converter and grid-tied inverter connected to the common direct-current link.

8 9

2.1. System overview

10

In a typical high-altitude wind energy system, the airborne module is tethered via a cable (rope) to the ground-

11

level winch system, whose topology is outlined in Fig. 1a. It is then used for power production during the airborne

12

module ascending phase, i.e. until the cable is almost completely unwound from the winch, which mandates

13

periodic cable winding (spooling) with ABM descent. Since ground-level power-plant features a reversible

14

electrical machine (generator/motor) unit, the cable can be easily wound at relatively low power settings (Fig.

15

1b) by reducing the aerodynamic lift vs. drag ratio of the airborne module [14]. An ultracapacitor energy storage

16

system has been included within the ground station. Its primary function is to store electrical energy during the

17

ABM ascending (power production) phase, and to subsequently provide sufficient power for cable winding and

18

electrical grid supply during the airborne module descending phase (Fig. 1b). The ultracapacitor ESS and other

19

ground station subsystems are connected to the common direct-current (DC) link via appropriate power

20

electronics systems (power converters). Their topologies are based on recommendations for wind power systems

21

presented in [25], and those related to energy storage system interfacing circuits given in [26].

5

1 2

Figure 1. Schematic representation of considered HAWE system (a) and illustration of its power flow cycle (b).

3 4

2.2. Modeling of individual subsystems

5

The ground station power-plant mechatronic system is modeled herein, based on the dynamical behavior of the

6

winch servodrive control system, the ultracapacitor energy storage system power flow relationships and the electrical

7

circuit representation of the common DC link and the grid inverter inductive-capacitive-inductive (LCL) filter.

8

2.2.1. Winch electrical drive

9

The structure of the winch motor/generator speed control loop with proportional-integral (PI) speed controller is

10

shown in Fig. 2. PI speed controller is traditionally used in controlled electrical drives [26] and power converter

11

control [27], and it is typically embedded within the servomachine power converter [28]. The electrical

12

servomachine torque limits are dependent on the winch drive operating point (rotational speed g), due to

13

servomachine power converter voltage limitations (see e.g. [26]). In the above model, the load torque at the winch

14

side related to cable pulling force Fr is given by w = FrrW.

6

1

The total winch inertia Jtot referred to the electrical machine shaft in the case of the direct drive (no gearbox) is

2

calculated in the following manner, which is valid under the assumption of negligible rope compliance and rather

3

small rope mass when compared to the mass of the airborne module mABM: J tot  J m  J w  mABM rW2 ,

4

(1)

5

where Jm and Jw in equation (1) are electrical machine and winch moments of inertia, respectively, mABM is the ABM

6

mass and rW is the winch radius.

7 8

Figure 2. Block diagram of motor/generator speed control system.

9 10 11 12 13

2.2.2. Ultracapacitor energy storage system The ultracapacitor-based ESS is typically modeled as a lumped-parameter resistor-capacitor electrical circuit [29], whose current vs. voltage relationship under negligible charge leakage conditions is given by: u uc 

Quc 1  iuc R s  C uc C uc

 iuc dt  iuc R s ,

(2)

14

where iuc in equation (2) is ultracapacitor current (iuc < 0 for discharging), uuc is terminal voltage, Cuc and Rs are

15

ultracapacitor capacitance and series resistance, respectively, and Quc is the accumulated electrical charge.

16 17 18 19 20

The ultracapacitor electrical circuit representation in equation (2) can be translated into an equivalent ultracapacitor power flow model by using the following ultracapacitor accumulated energy relationship [29]: Wuc 

2 2 Quc  2Qmax ,  2Cuc 2Cuc

(3)

where Qmax in equation (3) is the ultracapacitor charge capacity, and  = Quc/Qmax is the ultracapacitor state-of-charge. Since ultracapacitor input power Pc is related to the ultracapacitor current iuc = dQuc/dt as follows:

7

2 Pc  iuc Rs 

1 2

dWuc , dt

(4)

the final nonlinear model relating ultracapacitor input power Pc and state-of-charge  reads as follows: 2 2  2 Q max  4 R s C uc Pc  Q max d  f ( , Pc )  . dt 2 R s C uc

3

(5)

4 5 6 7

2.2.3. Direct-current link power flow model Based on the DC link electrical circuit representation in Fig. 1a, the DC link accumulated energy can be described by the following first-order model: dWdc  Pdc,mg  PESS  PL , dt

8 9 10

(6)

where Pdc,mg, PESS and PL are motor/generator inverter load, energy storage system power flow and grid-tied inverter load, respectively, and Wdc is the accumulated energy of the DC link capacitor related to its voltage udc as follows:

Wdc 

11

2 C dc u dc . 2

(7)

12 13

2.2.4. Grid inverter harmonic filter model

14

Three-phase four-quadrant grid inverter in Fig. 1a comprises an LCL harmonic filter for the purpose of suppressing

15

the higher-order current harmonics [30], due to its output voltage u1 being controlled by means of pulse-width

16

modulation (PWM) switching action. In the case of symmetrical grid-side load, the dynamic behavior of the three-

17

phase LCL harmonic filter can be analyzed based on the equivalent single-phase electrical circuit [31], which is

18

shown in Fig. 3a. The corresponding dynamic model can be conveniently represented by the block diagram shown in

19

Fig. 3b.

20 21 22

In the case when the LCL filter capacitive impedance 1/(2fgridC3f) is rather large (i.e. due to relatively small C3f choice), the LCL filter Laplace-domain transfer function can be simplified to a simple first-order lag model [31]: G f (s) 

i1 ( s ) 1  , u1 ( s )  u 2 ( s ) R f  L f s

(8)

23

where Rf = R1f + R2f and Lf = L1f + L2f in equation (8) represent the overall resistance and inductance of inverter-side

24

and grid-side reactors, which can be conveniently used in the grid inverter control system design.

8

1 2

Figure 3. Single-phase representation of LCL harmonic filter (a) and its block-diagram model (b).

3 4

2.3. Component sizing and parameterization

5

The individual subsystems of the HAWE system ground station power-plant are sized and parameterized based on

6

the HAWE system power production cycle requirements, and the specific considerations related to efficiencies of

7

individual ground station components, and the requirement for continuous grid power delivery.

8 9

2.3.1. Grid inverter power output and energy storage requirement

10

The parameters of HAWE system power cycle specification (Fig. 1b) and average efficiencies of its constituent

11

subsystems used in this work are listed in Table 1. These parameters are used to determine the ESS energy capacity

12

requirements and anticipated grid power delivery under steady-state HAWE power cycle (Table 1). The grid power

13

delivery Pgrid and energy storage requirement Wst are determined based on the procedure presented in [19], and

14

summarized in equations (9) and (10):

Pgrid   grid

15

Wst 

16

2 2 2 W2  MG  FC  ES Pm,asc Tasc  | Pm,des | Tdes , 2 Tdes  Tasc ES

(9)

2  Pgrid Tdes  Tasc ) | Pm,des | (Tdes   ES 1  . 2 2 2 2  grid ES  W  MG FC (W  MG FC ES Pm,ascTasc  | Pm,des | Tdes ) 

(10)

17

Based on these relationships, the energy storage capacity requirement Wst and anticipated (theoretical) grid power

18

delivery Pgrid are calculated and also listed in Table 1.

19 20

Table 1. Basic data of hypothetical HAWE system.

1 2

Maximum anticipated mechanical power1 at winch during ascending phase Pm,asc

170 kW

Maximum anticipated mechanical power at winch during descending2 phase |Pm,des|

58 kW

Based on maximum mechanical power data for a suitably chosen PMSM servomachine from [28]. Based on analysis presented in [19] 9

Altitude range h

300 m

Ascending vs. descending phase duration Tasc / Tdes 3

60 s / 40 s

Electrical machine + power converter average efficiency g = MGFC 4

0.84

Winch system average efficiency W 5

0.95

Grid inverter average efficiency grid 6

0.96

Energy storage system average efficiency ES 7

0.87

Energy storage minimum capacity requirement Wst

1.32 kWh

Theoretical grid power delivery Pgrid

32 kW

Winch radius rW 8

0.325 m

Total inertia referred to the motor/generator shaft Jtot 9

47 kgm2

1 2

However, the ultracapacitor energy storage capacity needs to be adjusted based on the requirement on the

3

minimum allowable state-of-charge min in order to avoid ultracapacitor system operating in the highly-inefficient

4

low state-of-charge region. Moreover, the storage system capacity should also be oversized with respect to

5

anticipated aging-related 20% capacity loss during the ultracapacitor system lifetime [36]. This is especially likely to

6

occur under highly-variable charging/discharging regimes, such as in electric vehicle applications [37]. Based on

7

equation (3), the following expression is derived for the adjusted ultracapacitor storage capacity [19]: Wst , adj 

8

 osWst  osWst  , 2 1   min 1  (1   max ) 2

(11)

9

where max = 1 – min is the maximum depth-of-discharge of the ultracapacitror energy storage system, and os = 1/0.8

10

= 1.25 is the energy storage oversizing factor related to the aforementioned aging-related 20% capacity loss. Based

11

on the above relationship, the adjusted energy storage capacity in the case when operation below 60% of state-of-

12

charge is not desired (min = 0.6) amounts to 2.6 kWh.

13 14

3

Based on analysis presented in [9] Estimated based on efficiency maps presented in [32], and anticipated machine operating points from [28]. 5 Based on winch transmission efficiency from [33] and winch PMSM electrical drive efficiency from [34] 6 Anticipated average efficiency of grid-connected inverters according to analysis presented in [35]. 7 Estimated efficiency of ultracapacitors according to [36] and DC/DC power converter according to [29]. 8 Based on winch speed and rope spooling data from [9]. 9 Based on motor inertia Jm = 4.5 kgm2 [28] and estimated winch inertia Jw = 26.7 kgm2 and ABM mass of 150 kg, as indicated in [9]. 4

10

1

2.3.2. Winch electrical drive rating and power losses assessment

2

A direct-drive permanent-magnet synchronous machine (PMSM) motor/generator with matching inverter is chosen

3

in this study, having in mind the above specifications in Table 1. The rated and maximum torque curves of the

4

motor/generator + inverter system are shown in Fig. 4a (cf. [28]). These curves have been derived by scaling up the

5

available efficiency map of a similar lower-power electrical drive, which has been obtained from the available

6

literature [38]. The results in Fig. 4a indicate that inverter-driven PMSM can be characterized by relatively high

7

efficiencies in high-power operating regimes.

8 9 10

Based on the overall efficiency map g(g, g) the power losses map Ploss(g, g) in Fig. 4b is calculated in the following form, which is more convenient for simulation purposes: Ploss ( g ,  g )  Pg

1   g ( g ,  g )

 g ( g ,  g )

  gg

1   g ( g ,  g )

 g ( g ,  g )

,

(12)

11 12

Figure 4. Efficiency map (a) and power losses map (b) of motor/generator and inverter unit according to (12).

13 14

2.3.3. Ultracapacitor energy storage system

15

The ultracapacitor ESS parameterization and sizing is based on the data for commercial modules from [39]. These

16

modules are characterized by 125 V voltage rating, rated capacitance Cmod = 62 F, series resistance Rmod = 15 m,

17

and maximum short-term discharge rate of 1850 A. In order to obtain the adjusted energy storage capacity Wst,adj of

18

2.6 kWh according to equation (11), a total of 20 ultracapacitor modules are required. However, HAWE system DC-

19

link voltage also needs to be matched fairly closely to the operating range of the motor/generator inverter unit DC

20

bus (which typically operates between 460 V and 690 V [28]). For the choice of ultracapacitor idling voltage Uc0 =

21

500 V, and energy capacity of 2.6 kWh, four groups of five parallel-connected ultracapacitor modules need to be

22

stacked in a series connection. In that case, the overall ultracapacitor system capacitance Cuc and series resistance Rs

23

can be calculated as follows: 11

Cuc 

1

5 4 C mod  77.5 F , R s  Rmod  12 m . 4 5

(13)

2

According to equation (13), the maximum ultracapacitor charge Qmax = CucUc0 = 10.8 Ah is obtained based on

3

ultracapacitor stack idle voltage Uc0 = 500 V and equivalent capacitance Cuc = 77.5 F.

4

Note that fixed efficiency DC of the DC/DC power converter has been included in the above overall energy storage

5

system efficiency ES. However, in real life applications, the DC/DC converter efficiency is modeled by a static

6

curve, which is typically above 0.92 (see [40] and references therein) during high-power charging and discharging.

7

This would result in overall energy storage system efficiency ES  0.87 for the typical case of ultracapacitor system

8

efficiency exceeding 0.95 [41].

9 10

2.3.4. Direct-current link capacitance and grid-side harmonic filter

11

The DC link capacitance Cdc is chosen based on the analysis of requirement of power transmission between the DC

12

link and the grid [40], which is expressed in the following inequality condition related to the magnitude of allowable

13

DC link voltage variations [42]: C dc 

14

S grid

 grid U dc U dc

,

(14)

15

where Sgrid is the anticipated grid apparent power delivery, grid = 2fgrid is the grid voltage frequency, Udc is DC link

16

average voltage, and Udc is the acceptable magnitude of DC link voltage ripple. By inserting Sgrid = 32 kVA, Udc =

17

690 V, fgrid = 50 Hz, and voltage variation Udc/Udc = 2%, the DC link capacitance needs to satisfy Cdc  11 mF.

18

The LCL filter design can be based on the recommendations in [43], wherein the filter bandwidth and higher-order

19

harmonic suppression ability are taken into account. Based on the procedure presented therein, the final LCL filter

20

parameters are listed in Table 2, along with anticipated resistances R1f and R2f of individual LCL filter reactors.

21 22

Table 2. LCL filter parameters for HAWE system grid inverter.

Parameter

C3f

R3f

L1f

L2f

R1f

R2f

Value

40 F

0.63 

3.3 mH

0.33 mH

100 m

10 m

23 24

12

1

3. CONTROL SYSTEM DESIGN

2

This section presents the design of individual control systems aimed at coordinated winch servodrive and tether

3

pulling force control, and DC link power flow control for the purpose of facilitating steady grid inverter power

4

delivery.

5 6

3.1. Damping optimum criterion

7

In order to obtain a well-damped speed control system response, the controller tuning may be carried out according

8

to the damping optimum criterion [22]. This is a practical pole-placement-like analytical method of design of linear

9

continuous-time closed-loop systems, which results in straightforward analytical relationships between the controller

10

parameters and the parameters of the process model. The tuning procedure is based on the following closed-loop

11

characteristic polynomial:

12

Ac ( s )  D2n 1 D3n2  DnTen s n    D2Te2 s 2  Te s  1 ,

(15)

13

where Te is the closed-loop equivalent time constant, and D2, D3,..., Dn are the damping optimum characteristic ratios.

14

The damping optimum criterion has been proposed in [22] with the aim to find analytical relations between the

15

coefficients of characteristic polynomial of an arbitrary order, so that the system has an optimal damping which

16

corresponds to the so-called “optimal” damping ratio  = 0.71 of the second-order dynamic system. In the damping

17

optimum based-design, the aforementioned requirement can be generalized for the linear system of any order. In the

18

so-called optimal case Di = 0.5 (i = 2 … n), resembling the behavior of a second-order system with damping ratio 

19

= 0.71, the closed-loop system has a quasi-aperiodic step response characterized by an overshoot of approximately

20

6% and the approximate rise time (1.8 – 2.1)Te. This choice of characteristic ratios Di may be considered optimal in

21

those applications where small overshoot and related well-damped behavior are critical.

22

In general, the response damping of any closed-loop mode may be adjusted through varying the characteristic ratios

23

D2, D3, … Dn. The damping of dominant closed-loop dynamics (i.e. the dominant closed-loop poles damping ratio) is

24

primarily influenced by the most dominant characteristic ratio D2, whereas the damping of higher-frequency modes

25

is adjusted by means of higher (so-called less dominant) characteristic ratios D3, … Dn. For example, by reducing the

26

characteristic ratio D2 to approximately 0.35 the fastest (boundary) aperiodic step response without overshoot is

27

obtained. On the other hand, if D2 is increased above 0.5 the closed-loop system response damping decreases, and

28

larger step response overshoots are obtained [23].

13

1

Regarding the choice of the closed-loop equivalent time constant Te, a larger Te value generally results in improved

2

control system robustness and reduced noise sensitivity. However, this also results in response slowdown and less

3

efficient disturbance rejection, because Te increase results in narrower closed-loop system bandwidth.

4 5

3.2. Winch electrical drive speed control system

6

Assuming small values of power converter and speed measurement delays in the block diagram representation of

7

the winch servomachine speed control loop in Fig. 2, these delays can be lumped into a single lag term characterized

8

by time constant T = T + Tm. This approximation results in the following form of closed-loop transfer function

9

characteristic polynomial:

10

Ac ( s ) 

J tot T Tc 3 J tot Tc 2 s  s  Tc s  1 . K c K c

(16)

11

By equating the coefficients of the closed-loop transfer function denominator (16) with the coefficients of the third-

12

order damping optimum polynomial (n = 3 in equation (15)), and rearranging, the following expressions for the PI

13

speed controller parameters are obtained:

14

Tc  Te 

T J tot , K c  , D2 D3 D2 Te

(17)

15

wherein by setting D2 = D3 = 0.5 (“optimal” tuning case) a fast and well-damped speed control loop response is

16

obtained with respect to external disturbance (cable force Fr = wrW) variations.

17 18

3.3. Winch drive superimposed control and rope tightening control

19

The superimposed motor/generator control strategy, shown in Fig. 5, coordinates the winch electrical servosystem

20

and ABM-based rope force control subsystems by commanding appropriate references to the winch speed control

21

loop and the ABM lift/drag control system (see e.g. [4]). The coordination strategy needs to determine the ABM

22

mode of operation based on the open-loop reconstruction of the unwound rope length. The unwinding (un-spooling)

23

is initiated when the estimated unwound rope length lˆr is below the lower threshold lmin. In that case, the winch

24

control system is commanded by the positive speed reference (wR = vasc/rW; vasc = 5 m/s), thus initiating the ABM

25

ascending phase. On the other hand, when the estimated unwound rope length exceeds the upper threshold lmax, a

26

negative (downward motion) speed reference is commanded (wR = vdes/rW < 0; vdes = –7.5 m/s). The generated winch

27

drive speed reference R0 is fed through a first-order low-pass filter, in order to facilitate relatively smooth winch

28

drive reversing, thus preventing likely excitation of rope vibration modes and undesired winch stresses. The 14

1

coordination strategy may also forward its commands to the ABM flight control strategy, alongside the pulling force

2

target FR, which may be used to enhance the power production when additional energy is required for energy storage

3

system recharging (e.g. due to DC link power requirement PdcR + PL in Fig. 5).

4

5 6

Figure 5. Block diagram of superimposed winch coordination strategy.

7 8

In order to properly facilitate the winch winding/unwinding, it is necessary to determine if rope tension is sufficient

9

to avoid the possible entanglement associated with rope slackening. Simple threshold logic is used within the

10

coordination strategy in Fig. 5, where the non-zero winch speed reference is commanded only if the estimated rope

11

force Fˆr is larger than a predefined positive threshold Fthr. The rope force estimator is based on the winch inertia

12

dynamic model (Fig. 2), implemented within the so-called Luenberger observer (see e.g. [44]). This is done by

13

modeling the rope force as a-priori unknown and slowly-varying disturbance (dFr/dt  0), and by extending the

14

overall multi-variable model with correction terms based on the winch speed tracking error [45]:

15

16

ˆ g 

1 J tot

(rW Fr   g )  K e ( g  ˆ g ) ,

(18)

ˆ Fr  K eF ( g  ˆ g ) ,

(19)

17

where ˆ g and Fˆr are estimated winch speed and rope force, respectively, and parameters Ke and KeF in equations

18

(18) and (19) are the so-called estimation error correction gains.

19 20

The Luenberger estimator design based on the damping optimum criterion, considers the following estimator characteristic polynomial, obtained by combining equations (16) and (17):

15

1 2 3 4

Ac ( s ) 

J tot J K s 2  tot e s  1 , rW K eF rW K eF

(20)

which is, in turn, equated to the damping optimum characteristic polynomial (n = 2) in equation (15). After some manipulation and rearranging, the final expressions for estimator correction gains read as follows: K e 

J tot 1 , K eF  . D2 o Teo D2oTeo2 rW

(21)

5 6

3.4. Direct-current link and ultracapacitor energy storage system power flow control

7

Figure 6 shows the block diagram of combined control of DC link and ultracapacitor ESS, which is aimed at

8

maintaining the DC link accumulated energy Wdc (i.e. DC link voltage udc) according to equations (6) and (7) within

9

narrow limits in order to minimize the voltage stresses to the common DC link capacitor bank [46]. The DC link

10

control system is based on the DC link PI feedback controller commanding appropriate charging/discharging

11

command to the energy storage system. The ultracapacitor ESS control system includes a proportional (P) controller

12

of ultracapacitor state-of-charge and which is combined with the superimposed charging/discharging demand from

13

the DC link control loop which command appropriate charging/discharging power commands to the ultracapacitor

14

DC/DC power converter (see e.g. [40]).

15

Since the DC link process model resembles in the structural sense the process model found in speed-controlled

16

electrical servo-drives and is characterized by second-order integral-lag process dynamics, PI controller can be

17

used instead of the more sophisticated controllers, such as the PID controller (see e.g discussion in [23]). In

18

particular, utilization of derivative action within PID controller might increase the closed-loop noise sensitivity,

19

which may play an emphasized role when power converter control applications are considered.

20

In the DC link PI controller design, the parasitic lag Tu represents the ultracapacitor ESS response lag. Based on

21

the block diagram representation in Fig. 6 (and assuming no interaction from the ultracapacitor state-of-charge

22

controller, P = 0), application of the damping optimum tuning procedure (see Subsection 3.1) results in the

23

following expressions for the DC link PI controller parameters:

24

TIdc  Tedc 

Tu 1 , K Rdc  . D2 dc D3dc D2 dc Tedc

(22)

25

Indeed, these expressions resemble the result of winch servomachine speed controller design (cf. equations (16) and

26

(17)). Again, by setting the characteristic ratios to D2dc = D3dc = 0.5 a fast and well-damped DC link control loop

27

behavior is obtained with respect to external disturbance (Pdc,mg and PL).

16

1 2

Figure 6. Block diagram of DC link and ultracapacitor ESS control system.

3 4

The ultracapacitor ESS control system commands appropriate charging/discharging power commands to the

5

ultracapacitor DC/DC power converter (see e.g. [40]) by means of a proportional (P) controller of ultracapacitor

6

state-of-charge and a superimposed charging/discharging demand from the DC link control system. The state-of-

7

charge feedback controller is enabled (ENchg flag is ON) when generator/motor unit produces power (ABM

8

ascending phase, Pdc,mg > 0). During ABM descending phase (Pdc,mg < 0), the feedback controller is disabled, and only

9

the superimposed command (demanding discharging) is active. However, if the ultracapacitor becomes fully-charged

10

( = R), further charging is disabled, and the excess power (discharging command P < 0) is re-routed towards the

11

grid inverter (by means of simple switching logic also shown in Fig. 6).

12 13 14

For the purpose of ultracapacitor state-of-charge controller design, the nonlinear ultracapacitor model in equation (5) is linearized in the vicinity of the anticipated operating point (0, Pc0) as follows:  

f ( 0 , Pc 0 ) f ( 0 , Pc 0 )   Pc  b  K p Pc ,  Pc

(23)

15

where  =  –0 and Pb = Pc – Pc0 correspond to perturbation variables in the linearized model (Fig. 6), with model

16

parameters (damping parameter b and gain Kp) given as follows:

17

b

18

1 1   2 Rs Cuc 2 Rs Cuc

Kp 

1 4R C 2 P 1  s2 uc2 c 0  0 Qmax

Cuc 2 Qmax  02 Qmax  4 Rs Cuc2 Pc 0

17

.

,

(24)

(25)

1

Taking into account that the ultracapacitor capacitance Cuc parameter in equations (24) and (25) is typically rather

2

large, the damping parameter b should take on fairly small values in comparison (b  0). Based on the block diagram

3

in Fig. 6 with b  0, the state-of-charge closed-loop transfer function is given by:

4 5 6 7 8

Gc ( s) 

 ( s) 1 .   R ( s) 1  ( K K p ) 1 s  ( K K p ) 1Tu s 2

(26)

By applying the damping optimum methodology to closed-loop equation (26), the following expressions for the closed-loop equivalent time constant Teu and controller gain K are obtained: Teu 

Tu 1 , K  , Teu K p D2,uc

(27)

with D2,uc = 0.35 chosen to obtain closed-loop response without overshoot, as elaborated in [40].

9 10

3.5. Grid-tied inverter control system

11

Grid-tied inverter power delivery can be controlled within the grid-synchronized control framework in the rotating

12

direct-quadrature (d-q) coordinate frame which facilitates decoupled active and reactive power control [47], as shown

13

in Fig. 7a. In the aforementioned cascade control scheme, the superimposed PI controllers of grid active and reactive

14

power delivery command the current references idR and iqR to inner current control loops. These inner control loops

15

correspond to the direct (d) and quadrature/orthogonal (q) current components id and iq, calculated from harmonic

16

phase currents ia, ib and ic by using the direct Park transform [27]:

17

18

i  id  2 cos( ) cos(  2 / 3) cos(  4 / 3)  a  i     i b  ,  q  3  sin( ) sin(  2 / 3) sin(  4 / 3)  i   c

(28)

where  = grid dt is the phase angle of the grid voltage harmonic waveform at grid frequency grid = 2fgrid.

19

The currents id and iq are controlled by respective fast PI controllers, as shown in Fig. 7b, which are typically

20

embedded within the inverter control unit. The current control system also needs to include compensation of d-q axis

21

cross-coupling terms ud and uq based on the available grid frequency grid and current measurements (feed-

22

forward d-q axis decoupling). The current controller outputs, i.e. the voltage references udR and uqR, are then used to

23

calculate the inverter phase voltage references uaR, ubR, and ucR by means of inverse Park transform:

24

cos( ) sin( ) u aR    u   cos(  2 / 3) sin(  2 / 3) u dR  .  bR    u  u cR  cos(  4 / 3) sin(  4 / 3)  qR 

(29)

25

In order to facilitate the calculation of direct and quadrature quantities in the d-q rotating reference frame, and to

26

establish the direct and inverse Park transforms in (28) and (29), it is necessary to synchronize the inverter with

18

1

respect to the grid voltage. For that purpose, a synchronous reference frame phase-locked-loop (SRF-PLL) algorithm

2

[48] is considered in Fig. 7a. The inverter is synchronized herein with respect to the direct component ud through

3

estimation of grid voltage phase angle  and frequency grid. A dedicated PI-type PLL algorithm is used for that

4

purpose, which controls the quadrature component of the rotating-frame grid voltage (uq) to zero [49].

5

Grid-tied inverter power flow control system design is also based on the damping optimum tuning procedure,

6

which starts with the definition of the closed-loop characteristic polynomials for the active and reactive power

7

control loops:

8

Ac , P ( s )  1 

9

Ac ,Q ( s )  1 

2  3U fm K RP 3K RPU fm 3U fm K RQ  2 3K RQU fm

TIP s  TIQ s 

2Tei TIP 2 s , 3K RPU fm 2Tei TIQ 3K RQU fm

s2 ,

(30)

(31)

10

where Ufm is the phase voltage magnitude in d-q frame under synchronized conditions (i.e. ud = Ufm and uq = 0), and

11

Tei is the equivalent lag of the “fast” inner current control loop (Fig. 7a).

12 13 14

15 16

By applying the damping optimum approach to closed-loop characteristic polynomials (30) and (31), the following analytical expressions for PI controller parameters are obtained:  D T  TIP  TIQ  Tec 1  2c ec  , Tei   K RP   K RQ 

2 3U fm

(32)

 Tei   1 ,   D2 c Tec 

(33)

with Tec and D2c in equations (32) and (33) subject to feasibility condition Tei > D2cTec.

17

As indicated by block diagram in Fig. 7b, the current controller design should be the same for the direct and

18

quadrature current control loops due to their symmetrical structures. Moreover, the largest open-loop time constant

19

(i.e. the LCL filter time constant Tf = Lf/Rf herein) is typically cancelled out by the PI controller zero in the current

20

control loop design (see [26]), thus resulting in a straightforward choice for the PI controller integral time constant:

21 22 23

Tci 

Lf Rf



Lf1  Lf 2 Rf1  Rf 2

.

(34)

Due to aforementioned zero-pole canceling, the current control system closed-loop transfer function now reads: 1

Gc ,iq ( s )  Gc,id ( s )  1

R f Tci K ci

T T s  ci i s 2 K ci

.

(35)

24

By equating the denominator of the transfer function model (35) with the second-order damping-optimum

25

characteristic polynomial (n = 2 in equation (15)), the following expressions are obtained for current controller

26

proportional gain Kci and closed-loop equivalent time constant (closed-loop equivalent lag) Tei: 19

T Ti , K ci  ci R f , D 2i Tei

1

Tei 

2

where Rf is the equivalent series resistance of the LCL filter (see Subsection 3.1).

(36)

3

4 5

Figure 7. Block diagram of superimposed grid-tied inverter power flow control system and grid synchronization

6

system (a) and inner current control system (b).

7 8 9

Finally, the damping optimum methodology is also applied in the design of PLL-based grid synchronization loop, wherein the uq “feedback” signal is obtained by using Park transform [27]: u q  U fm cos(ˆ   ) ,

10

11

which is linearized in the vicinity of the steady-state solution ˆ  ˆ   as follows: u q  U fm sin(ˆ)  U fm ˆ .

12 13 14

(37)

(38)

Application of the damping optimum criterion to the PLL system, linearized according to equations (37) and (38), and characterized by the following closed-loop transfer function characteristic polynomial: Apll ( s )  

15

T pll T p U fm K pll

s3 

T pll U fm K pll

s 2  T pll s  1 .

(39)

16

The characteristic polynomial in equation (39) takes on a similar form as in the case of winch servomachine speed

17

control loop, which results in the following expressions for PLL PI control algorithm integral time constant Tpll and

18

proportional gain Kpll:

19

T pll  Te, pll 

Tp D2 pll D3 pll

, K pll  

20 21 20

1 . D2 pll Te, pllU fm

(40)

1

4. SIMULATION RESULTS

2

The ground station and grid inverter control system dynamic models, incorporating the above presented control

3

system sub-models, has been implemented within the Matlab/SimulinkTM simulation software environment. The

4

respective simulation models implemented in SimulinkTM graphical user interface are given in the Apendix in order

5

to illustrate the complexity and mutual interconnections of the individual sub-models of the power flow strategy

6

and grid inverter control system (Fig. A1). The dynamic equations within the simulation model are solved by using

7

the variable-step Dormand-Prince integration method, with relative and absolute tolerances of 10-3 and 10-6,

8

respectively, which should result in favorable model accuracy. The parameters of individual controllers and rope

9

force observer are listed in Table A1 in Appendix.

10

The proposed high-level coordination strategy including the winch servodrive and ABM/rope tension control

11

system and power flow control system has been first tested for the cases of steady power production characterized by

12

constant wind strength and sudden wind disturbance (wind power drop), as illustrated in Fig. 8. In both of these

13

scenarios, the ground station operation starts with rope unwinding until maximum unwound length is achieved,

14

which is followed by periodic winding and unwinding between the minimum and maximum unwound rope lengths

15

(Fig. 8d). Assuming effective ABM vs. rope force control, as outlined in [4], the rope tension force Fr profile (Fig.

16

8a) needs to be countered by the winch servodrive torque (g) under the closed-loop winch drive speed control (see

17

Figs. 8b and 8c). In the case of steady-state power production (blue traces in Fig. 8), winch drive alternates between

18

rather high and relatively low torque values. On the other hand, when wind disturbance (wind strength drop) is

19

simulated by means of rope pulling force reduction, servodrive adjusts the related torque values under the closed-

20

loop speed control, which maintains the predefined altitude rate of change given in Table 1, and discussed in

21

Subsection 3.3.

21

1 2

Figure 8. Results of winch servodrive and AMB coordination control strategy verification for the case of ABM

3

ascending from ground level and subject to high-altitude wind disturbance effect: rope force (a),

4

generator/motor unit torque (b), generator/motor unit speed (c), and rope length (d).

5 6

Figure 9 shows an initial detail of the rope unwound length and tension force responses for the simulation scenario

7

in Fig. 8, which illustrates the performance of the rope force estimator and related winch unwinding enabling logic.

8

These results show that the proposed rope force estimator (tuned with Teo = 50 ms herein) is characterized by a

9

relatively fast and well-damped response, and is able to track the rope force profile (Fig. 9b) resulting from the winch

10

drive vs. ABM coordination. Based on the estimated force, winch unwinding is enabled only after the rope is

11

sufficiently taut (Fig. 9a); otherwise, the winch is kept at standstill to allow the rope tension to build up until the

12

predefined threshold (Fthr = 3 kN) is exceeded (Fig. 9b).

13

14 15

Figure 9. Details of winch system responses at beginning of rope unwinding: rope unwound length (a), and actual

16

and estimated rope force traces (b).

17 18

Figure 10 shows the power flow control system responses related to the winch servodrive vs. rope tension control

19

system responses in Fig. 8, for the case of constant grid active power target PR,grid = 15 kW. In particular, during the

20

power production phases, the winch motor/generator unit DC-link output power Pdc,mg effectively covers for the power 22

1

requirement PESS of the ultracapacitor ESS and grid inverter DC link load PL. On the other hand, during ABM descent

2

phase, the energy storage system discharge operation covers for the power consumption of the motor/generator unit

3

required for rope winding under low tension conditions, while also supplying the steady grid-side inverter active

4

power delivery (Figs. 10a – 10c). These DC link power flow profiles are reflected in the ESS state-of-charge, which is

5

initially charged from partially discharged state ( = 30%) to near 100% state-of-charge during the initial ABM ascent

6

to maximum altitude (ESS pre-charging under closed-loop control). Since state-of-charge target R is reached during

7

this phase of operation, the excess power needs to be re-routed towards the grid inverter, thus briefly increasing the

8

overall grid power delivery. Since periodic ESS charging and discharging is mandated by the HAWE system operating

9

cycle, the ESS state-of-charge varies in this scenario between 80% and 100% during HAWE system cyclic operation

10

(Fig. 10d). Moreover, during the wind disturbance event, the motor/generator unit power production is notably

11

decreased, which is reflected in reduced ESS charging (and discharging) power (Figs. 10b and 10e), thus ultimately

12

leading to state-of-charge drift towards lower values (Fig. 10d). After the power production increase, the ultracapacitor

13

ESS state-of-charge is steadily increased (i.e. being recharged by additional power from motor/generator) so that it

14

eventually establishes the steady-state charging/discharging cycle between 80% and 100% state-of-charge. DC link

15

accumulated energy under considered highly-varying power profiles is shown in Fig. 10f. The results indicate that the

16

proposed DC link controller is able to maintain the DC link energy level (i.e. voltage) within narrow bounds (-2% and

17

+4% of the target value WdcR = 6.25 kJ herein).

18

The power flow control strategy has also been validated for the case of quasi stochastic wind disturbance and

19

varying grid inverter power delivery, with the main simulation results presented in Fig. 11. In this particular

20

scenario, the speed-controlled winch electrical servodrive maintains the required winch speed profile (Fig. 11a),

21

but the effect of highly-varying wind disturbance manifests itself in large-magnitude quasi-stochastic

22

perturbations in rope force-related winch torque (Fig. 11b). The resulting variability of the winch

23

generator/motor unit power (Fig. 11c) and the varying power demand from the grid inverter (Fig. 11e) needs to

24

be covered by the ultracapacitor ESS (Fig. 11d). In this scenario, characterized by lower average power

25

production, the ultracapacitor ESS state-of-charge also takes on lower values compared to the case of no wind

26

disturbance (cf. Figs. 10d and 11d).

27

Figure 12 shows the responses of grid inverter control strategy shown in Fig. 7 for the case of constant active

28

power command PR,grid = 15 kW with reactive power command kept at zero QR,grid = 0. The active and reactive

29

power responses in Fig. 12a and corresponding direct and quadrature current and voltage responses in Figs. 12b and

30

12c are characterized by fast and well-damped dominant transient behavior, which is characteristic for damping

23

1

optimum tuning method, albeit with some high-frequency transients due to finite time required for the convergence

2

of grid frequency grid estimate based on PLL algorithm (Fig. 12d), which is also characterized by well-damped

3

response shape. Ultimately, the above results facilitate high control performance of individual phase harmonic

4

currents and corresponding phase voltage references (Figs. 12e and 12f), wherein steady-state phase currents are

5

characterized by zero DC bias (offset) and negligible distortion from the target sinusoidal (harmonic) waveforms.

6

7 8

Figure 10. DC link power flow responses without and with stepwise wind disturbance (drop in wind strength): DC

9

link power from generator/motor unit (a), ultracapacitor ESS power profile (b), grid inverter load (c), ESS state-of-

10

charge (d), ESS current (e) and DC link energy (f).

24

1 2

Figure 11. Winch drive and power flow control strategy with quasi-stochastic wind disturbance: generator/motor unit

3

speed (a), generator/motor unit torque (b), DC link power flow from generator/motor and ESS (c), ESS state-of-

4

charge (b), grid inverter load (c), ESS state-of-charge (d), grid inverter load (e) and DC link energy (f).

5 6

Figure 12. Grid inverter responses under power/current cascade control and grid synchronization: active and reactive

7

power (a), d-q frame currents (b), d-q frame voltages (c), grid frequency and its estimate (d), phase currents (e) and

8

phase voltage references (f). 25

1

5. CONCLUSION

2

The paper has presented the modeling, parameterization and control system design for the high-altitude wind

3

energy system ground station power-plant equipped with an ultracapacitor energy storage system. The control strategy

4

has been aimed at coordinating the individual power sources connected to the common direct-current link, in particular

5

the winch motor/generator servodrive, an ultracapacitor energy storage system, and grid-side inverter.

6

The control strategy has included the low-level winch motor/generator unit speed control system equipped with a

7

PI speed controller, and the superimposed control level intended for coordination between the winch servodrive and

8

the cable tension control via airborne module lift force. The upper-level winch control system has been based on a

9

Luenberger observer of rope tension force and relatively simple switching logic. The DC power flow control has

10

been based on a DC link accumulated energy PI controller commanding appropriate upper-level

11

charging/discharging command to the ultracapacitor ESS equipped with a simple proportional-type state-of-charge

12

controller. Finally, the grid-tied inverter has been equipped with active and reactive power flow control system

13

established in the grid-synchronous d-q reference frame, and synchronized to the grid by means of a dedicated phase-

14

locked-loop algorithm. The tuning of aforementioned feedback controllers, Luenberger estimator, and the phase-

15

locked-loop synchronization algorithm has been based on the damping optimum criterion, yielding straightforward

16

analytical relationships for controller and observer parameters, and assuring their well-damped closed-loop behavior.

17

The effectiveness of the proposed ground station control system has been verified by means of comprehensive

18

computer simulations for two characteristic HAWE system operating scenarios, corresponding to steady power

19

production and the case of sudden high-altitude wind disturbance emulated through reduced cable pulling force

20

magnitude. The results have shown that: (i) the winch servodrive can be effectively coordinated with cable tension

21

system so that cable unwinding has been enabled only when the cable has been sufficiently taut; (ii) the DC link control

22

system effectively distributes the power flow between the motor/generator unit, ultracapacitor ESS and grid inverter,

23

while keeping the DC link energy within narrow bounds; and (iii) the grid inverter control system can provide smooth

24

delivery of active and reactive power to the grid. Finally, the control system performance has also shown robustness

25

with respect to quasi-stochastic wind disturbance, wherein energy for maintained grid power delivery has been supplied

26

by the ultracapacitor energy storage system.

27

Future work may be directed towards further refinement of the proposed control strategy, such as coordination

28

between the motor/generator unit, ultracapacitor ESS and grid inverter. These measures may include the variable rope

29

spooling and un-spooling speeds in order to accommodate the complex trajectories made by the airborne module, and

26

1

varying active and reactive power demands from the grid inverter and their coordination with the ABM power

2

production cycle and energy storage system.

3 4

Acknowledgement

5

It is gratefully acknowledged that this work has been supported by the European Commission through the “High

6

Altitude Wind Energy” FP7 project, grant No. 256714. It is also acknowledged that the presented research has been

7

conducted in part within the activities of the Center of Research Excellence for Data Science and Cooperative Systems

8

supported by the Ministry of Science and Education of the Republic of Croatia. Authors would also like to express

9

their appreciation of the efforts of the journal Editor and anonymous reviewers whose comments and suggestions

10

have helped to improve the quality of the presented subject matter.

11 12

NOMENCLATURE Abbreviations:

Parameters:

ABM

airborne module

C3f

LCL filter capacitance [F]

DC

direct current

Cdc

DC link capacitance [F]

ESS

energy storage system

Cuc

ultracapacitor ESS capacitance [F]

HAWE high-altitude wind energy

Di

damping optimum characteristic ratios (i = 2…n)

LCL

inductive-capacitive-inductive

fgrid

grid voltage frequency [Hz]

PI

proportional-integral

Jtot

total inertia at winch side [kgm2]

PLL

phase-locked loop

Jw, Jm

winch and generator inertia [kgm2]

PMSM permanent-magnet synchronous machine

Kc

PI speed controller proportional gain

PWM

pulse-width modulation

Kci

current controller proportional gain

SRF

synchronous reference frame

Ke, KeF Luenberger observer correction gains

d-q

direct-quadrature

Kpll

PLL proportional gain

Dynamic variables:

KRdc

DC link PI controller proportional gain

^

estimated value

KRP,KRQ grid inverter power controller proportional gains

Ac(s)

closed-loop characteristic polynomial

K

state-of-charge controller proportional gain

Fr, FR

rope pulling force and rope force reference [N]

L1f, L2f

LCL filter reactor inductances [H]

ia, ib, ic

instantaneous phase currents [A]

Ldc

storage system inductances [H]

iuc

ultracapacitor current [A]

mABM

airborne module mass [kg]

iESS

storage system current [A]

Qmax

ultracapacitor charge capacity [As]

i1, i2, if

LCL filter current components [A]

rW

winch radius [m]

id, iq,

direct and quadrature current components

R3f

LCL filter capacitive branch resistance []

lr

unwound rope length [m]

R1f, R2f

LCL filter reactor Ohmic resistances []

Pc

ultracapacitor power [W]

Rs

ultracapacitor ESS series resistance []

27

Pdc,mg

motor/generator power [W]

Tasc, Tdes duration of ABM ascending and descending [s]

PdcR

DC link power reference [W]

Tc

PI speed controller integral time constant [s]

PESS

energy storage system power [W]

Tci

current controller integral time constant [s]

Pgrid

grid inverter active power [W]

Te

damping optimum equivalent time constant [s]

Pg

generator/motor mechanical power [W]

TIdc

DC link PI controller integral time constant [s]

Ploss

power losses map [W]

TIP,TIQ

grid inverter power controller time constants [s]

PL

grid inverter load [W]

Wst,

ESS energy storage requirement [J], [kWh]

Pm,asc

power at winch during ABM ascending [W]

Wst,adj

ultracapacitor energy storage capacity [J], [kWh]

|Pm,des|

power at winch during ABM descending [W]

h

ABM altitude range [m]

Qgrid

grid inverter active power [VAr]

W

winch system mechanical efficiency

Sgrid

grid inverter apparent power [VA]

MG

generator/motor efficiency

s

-1

Laplace operator [s ]

FC

generator/motor power converter efficiency

u1, u2

LCL filter input and output voltage [V]

ESS

energy storage system efficiency

ua,ub,uc instantaneous phase voltages [V]

grid

grid inverter efficiency

udc, Udc instantaneous and average DC link voltage [V]

g

generator/motor and power converter efficiency

ud, uq

os

energy storage over-sizing factor



closed-loop damping ratio

direct and quadrature voltage components [V]

ud,uq d-q axis cross-coupling terms [V] uuc

ultracapacitor terminal voltage [V]

g, w

generator and winch angular velocity [rad/s]

Uc0

ultracapacitor stack idle voltage [V]

R,

generator angular velocity reference [rad/s]

grid,

grid voltage frequency [rad/s]

vasc, vdes ABM ascending and descending speed [m/s] Wdc, Wuc DC link and ultracapacitor energy [J], [kWh]

g, w

generator and winch torque [Nm]



ultracapacitor state-of-charge



grid voltage phase angle [rad]

1 2

REFERENCES

3

1.

Loyd, M. L., Crosswind Kite Power, Journal of Energy, Vol. 4, No. 3, pp. 106-111, 1980.

4

2.

Canale, M., Fagiano, L., Milanese, M., High Altitude Wind Energy Generation Using Controlled Power

5 6

Kites, IEEE Transactions on Control Systems Technology, Vol. 18, No. 2, pp. 279 – 293, 2010. 3.

7 8

and fast motion transfer, Energy, Vol. 55, pp. 987-995, 2013. 4.

9 10 11

Goldstein, L., Theoretical analysis of an airborne wind energy conversion system with a ground generator

Milutinović, M., Kranjčević, N., and Deur, J., Multi-mass dynamic model of a variable-length tether used in a high altitude wind energy system, Energy Conversion and Management, Vol. 87, pp. 1141-1150, 2014.

5.

Roberts, B.W., Shepard, D.H., Caldeira, K., Cannon, M.E., Eccles, D.G., Grenier, A.J., Harnessing HighAltitude Wind Power, IEEE Transactions on Energy Conversion, Vol. 22, No. 1, pp. 136 – 144, 2007.

28

1

6.

2 3

generator, Renewable Energy, Vol. 34, No. 6, pp. 1525 – 1532, 2009. 7.

4 5

8 9 10

Fechner, U., Schmehl, R., Design of a Distributed Kite Power Control System, Proceedings of 2012 IEEE International Conference on Control Applications, pp. 800-805, Dubrovnik, Croatia, October 3-5, 2012.

8.

6 7

Argatov, I., Rautkorpi, P., Silvennoinen, R., Estimation of the mechanical energy output of the kite wind

Williams, P., Lansdorp, B., Ockels, W., Optimal Crosswind Towing and Power Generation with Tethered Kites, Journal of Guidance, Control, and Dynamics, Vol. 31, No. 1, pp. 81 – 93, 2008.

9.

Milutinović, M., Čorić, M., and Deur, J., Operating cycle optimization for a Magnus effect-based airborne wind energy system, Energy Conversion and Management, Vol. 90, pp. 154-165, 2015.

10. Saeed, M., and Kim, M.-H., Aerodynamic performance analysis of an airborne wind turbine system with NREL Phase IV rotor, Energy Conversion and Management, Vol. 134, pp. 278-289, 2017.

11

11. Saleem, A., and Kim, M.-H., Effect of rotor axial position on the aerodynamic performance of an airborne

12

wind turbine system in shell configuration, Energy Conversion and Management, Vol. 151, pp. 587-600,

13

2017.

14 15 16 17

12. Canale, M., Fagiano, L., and Milanese, M., KiteGen: A Revolution in Wind Energy Generation, Energy, Vol. 34, No. 3, pp. 355-361, 2009. 13. Lansdorp, B., Ockels, W. J., The Laddermill - Innovative Wind Energy from High Altitudes in Holland and Australia, Global Windpower ‘06 Conference & Exhibition, Adelaide, Australia, September, 2006.

18

14. Perković, L., Silva, P., Ban, M., Kranjčević, N., and Duić, N., Harvesting high altitude wind energy for

19

power production: The concept based on Magnus’ effect, Applied Energy, Vol. 101, pp. 151-160, 2013.

20

15. Kim, J., and Park, C., Wind Power Generation with a Parawing on Ships, a Proposal, Energy, Vol. 35, No.

21

3, pp. 1425 – 1432, 2010.

22

16. Lunney, E., Ban, M., Duić, N., and Foley, A., A state-of-the-art review and feasibility analysis of high

23

altitude wind power in Northern Ireland, Renewable and Sustainable Energy Reviews, Vol. 68, pp. 899 –

24

911, 2017.

25 26

17. Ban, M., Perković, L., Duić, N., and Penedo, R., Estimating the spatial distribution of high altitude wind energy potential in Southeast Europe, Energy, Vol. 57, pp. 24-29, 2013.

27

18. Saeed, M., and Kim, M.-H., Airborne wind turbine shell behavior prediction under various wind conditions

28

using strongly coupled fluid structure interaction formulation, Energy Conversion and Management, Vol.

29

120, pp. 217-228, 2016.

29

1 2

19. Pavković, D., Hoić, M., Deur, J., and Petrić, J., Energy Storage Systems Sizing Study for a High-Altitude Wind Energy Application, Energy, Vol. 76, pp. 91–103, 2014.

3

20. Krajačić, G., Duić, N., Vujanović, M., Kiklis, S, Rosen, M. A, and Al-Nimr, M. A., Sustainable

4

development of energy, water and environment systems for future energy technologies and concepts,

5

Energy Conversion and Management, Vol. 125, pp. 1-14, 2016.

6

21. Pavković, D., Lobrović, M., Hrgetić, M., and Komljenović, A., A Design of Cascade Control System and

7

Adaptive Load Compensator for Battery/Ultracapacitor Hybrid Energy Storage-based Direct Current

8

Microgrid, Energy Conversion and Management, Vol. 114, pp. 154 – 167, 2016.

9

22. Naslin, P., Essentials of Optimal Control, Ch. 2, Iliffe Books Ltd, London, UK, 1968.

10 11 12 13

23. Pavković, D., Polak, S., Zorc. D., PID Controller Auto-Tuning Based on Process Step Response and Damping Optimum Criterion, ISA Transactions, Vol. 53, No. 1, pp. 85–96, 2014. 24. Baroudi, J.A., Dinavahi, V., and Knight, A. M., A Review of Power Converter Topologies for Wind Generators, Renewable Energy, Vol. 32, No. 14, pp. 2369 – 2385, 2007.

14

25. Pernão Pires, V., Romero-Cardaval, E., Vinnikov, D., Roasto, I., and Martins, J.F., Power converter

15

interfaces for electrochemical energy storage systems – A review, Energy Conversion and Management,

16

Vol. 86, pp. 453 – 475, 2014.

17 18

26. Schröder, D., Elektrische Antriebe – Regelung von Antriebssystemen, 3rd ed., Springer-Verlag, Berlin, Germany, 2007.

19

27. Timbus, A., Liserre, M., Teodorescu, R., Rodriguez, P., and Blaabjerg, F., Evaluation of Current

20

Controllers for Distributed Power Generation Systems, IEEE Transactions on Power Electronics, Vol. 24,

21

No. 3, pp. 654 – 664, 2009.

22

28. … SIMOVERT MASTERDRIVES - Compendium Vector Control, Siemens AG, 2004.

23

29. Miller, J. M., Ultracapacitors, Chap. 7 of Handbook of Automotive Power Electronics and Motor Drives,

24 25 26 27 28 29 30

A. Emadi (ed.), Taylor & Francis Group, 2005. 30. Zhong, Q.-C., and Hornik, T., Control of Power Inverters in Renewable Energy and Smart Grid Integration, John Wiley & Sons Ltd / IEEE Press, NY, USA, 2013. 31. Liserre, M., Blaabjerg, F., and Hansen, S., Design and Control of an LCL-Filter-Based Three-Phase Active Rectifier, IEEE Transactions on Industry Applications, Vol. 41, No. 5, pp. 1281 – 1291, 2005. 32. Odvářka, E., Mebarki, A., Gerada, D., Brown, N., and Ondrůšek, Č., Electric Motor-generator for a Hybrid Electric Vehicle, Engineering Mechanics, Vol. 16, No. 2, pp. 131–139, 2009.

30

1

33. Matejić, M., Blagojević, M., Đorđević, Z., Marjanović, N., and Petrović, N., Comparative Analysis of

2

Different Type Reducers for Winch Drum Driving Unit, Proceedings of 7th International Quality

3

Conference, pp. 255-262, University of Kragujevac, Serbia, 2013.

4 5

34. Lamas Pardo, M., Carral, L, Castro-Santos, L., and Carral-Couce, J. C., A review of the drive options of offshore anchor handling winches, Brodogradnja/Shipbuilding, Vol. 68, No 3, pp. 119-134, 2017.

6

35. Rampinelli, G.A., Krenzinger, A. and Chenlo Romero, F., Mathematical models for efficiency of inverters

7

used in grid connected photovoltaic systems, Renewable and Sustainable Energy Reviews, Vol. 34, pp.

8

578–587, 2014.

9 10

36. Miller, J.R., and Burke, A.F., Electrochemical Capacitors: Challenges and Opportunities for Real-World Applications, The Electrochemical Society's Interface, Vol. 17, No. 1, 2008.

11

37. Burke, A.F., Ultracapacitor technologies and application in hybrid and electric vehicles, International

12

Journal of Energy Research, Special Issue: Electrical Energy Storage for Future Transportation and

13

Renewable Energy, Vol. 34, No. 2, pp. 133–151, 2010.

14 15

38. Olszewski, M., Evaluation of 2004 Toyota Prius Hybrid Electric Drive System, Technical report, ORNL/TM-2006/423, Oak Ridge National Laboratory, TN, USA, 2006.

16

39. … NESSCAP Ultracapacitor Products-125V Module Data Sheet (http://www.nesscap.com), 2010.

17

40. Pavković, D., Sedić, A., and Guzović, Z., Oil Drilling Rig Diesel Power-plant Fuel Efficiency

18

Improvement Potentials through Rule-Based Generator Scheduling and Utilization of Battery Energy

19

Storage System, Energy Conversion and Management, Vol. 121, pp. 194 – 211, 2016.

20 21

41. Emadi, A., Ehsani, M., and Miller, J.M., Vehicular Electric Power Systems - Land, Sea, Air, and Space Vehicles, Marcel Dekker Inc., New York, USA, 2004.

22

42. Gohil, G.V., Wang, H., Liserre, M., Kerekes, T., Teodorescu, R., and Blaabjerg, F., Reduction of DC-link

23

Capacitor in Case of Cascade Multilevel Converters by means of Reactive Power Control, Proceedings of

24

the 29th Annual IEEE Applied Power Electronics Conference and Exposition (APEC 2014), pp. 231 – 238,

25

2014.

26

43. Pavković, D., Užarević, V., Kristović, P., Hrgetić, M., and Komljenović, A., Single Phase AC Inverter

27

Current PR Control with Auxiliary PI Controller for DC Current Suppression, In Proc. of IEEE EUROCON

28

2017, Ohrid, Republic of Macedonia, 2017.

29 30

44. Weihrich, G., Drehzahlregelung von Gleichstromantrieben unter Verwendung eines Zustands- und Störgrößen-Beobachters, Regelungstechnik, Vol. 26, pp. 349-354 and pp. 392-397, 1978.

31

1

45. Pavković, D., Deur, J., and Lisac, A., Torque Estimator-based Control Strategy for Oil-Well Drill-string

2

Torsional Vibrations Active Damping Including an Auto-tuning Algorithm, Control Engineering Practice,

3

Vol. 19, No. 8, pp. 836–850, 2011.

4

46. Kolar, J.W., and Round, S.D., Analytical Calculation of the RMS Current Stress on the DC Link Capacitor

5

of Voltage-PWM Converter Systems, IET Electrical Power Applications, Vol. 153, No. 4, pp. 535-543,

6

2006.

7

47. Arboleya, P., Diaz, D., Guerrero, J.M., Garcia, P., Briz, F., Gonzalez-Moran, C., and Gomez Aleixandre, J.,

8

An improved control scheme based in droop characteristic for microgrid converters, Electric Power

9

Systems Research, Vol. 80, pp. 1215 – 1221, 2010.

10

48. Rodríguez, P., Pou, J., Bergas, J., Candela, J.I., Burgos, P., and Boroyevich, D., Decoupled Double

11

Synchronous Reference Frame PLL for Power Converters Control, IEEE Transactions on Power

12

Electronics, Vol. 22, No. 2, pp. 584 – 592, 2007.

13

49. Rodríguez, P., Luna, A., Candela, I., Mujal, R., Teodorescu, R., and Blaabjerg, F., Multiresonant

14

Frequency-Locked Loop for Grid Synchronization of Power Converters Under Distorted Grid Conditions,

15

IEEE Transactions on Industrial Electronics, Vol. 58, No. 1, pp. 127 – 138, 2011.

32

1

Appendix: Simulation model and key parameters

2 3

Figure A1. Ground station power flow control system model (a), and grid inverter control system (b)

4

implemented in Matlab/SimulinkTM software environment. 33

1

Table A1. Parameters of individual controllers and rope force observer.

Speed controller proportional gain Kc = 2332.2

Power controller proportional gain KRP = –KRQ = 5.36·10-4

Speed controller integral time constant Tc = 40 ms

Power controller integral time constant TIR = TIQ = 0.64 ms

Luenberger observer correction gain Ke = 40

Current controller proportional gain Kci = 1.65

Luenberger observer correction gain KeF = 1.15·105

Current controller integral time constant Tci = 33 ms

DC link controller proportional gain KRdc = 66.67

SRF PLL controller proportional gain Kpll = -1.607

DC link controller integral time constant TIdc = 30 ms

SRF PLL controller integral time constant Tpll = 4 ms

State-of-charge controller gain K = 4.5·105 2

34

Suggest Documents