You are on page 1of 15

The Twelfth Scandinavian International Conference on Fluid Power, May 18-20, 2011, Tampere, Finland

MODEL BASED DESIGN OF EFFICIENT POWER TAKE-OFF


SYSTEMS FOR WAVE ENERGY CONVERTERS
Rico H. Hansen, Torben O. Andersen, Henrik C. Pedersen
Aalborg University
Department of Energy Technology
Pontoppidanstrde 101, 9220 Aalborg, Denmark
Phone: + 45 9940 9240, Fax: + 45 9815 1411
E-mail: rhh@et.aau.dk, toa@et.aau.dk, hcp@et.aau.dk

ABSTRACT
The Power Take-Off (PTO) is the core of a Wave Energy Converter (WECs), being the
technology converting wave induced oscillations from mechanical energy to electricity.
The induced oscillations are characterized by being slow with varying frequency and
amplitude. Resultantly, fluid power is often an essential part of the PTO, being the
only technology having the required force densities. The focus of this paper is to
show the achievable efficiency of a PTO system based on a conventional hydro-static
transmission topology. The design is performed using a model based approach. Generic
component models are developed and combined into a PTO system, describing the
dynamics and power losses from wave to grid. Using the model, components sizes and
control are optimised and the achievable performance of the PTO is identified.
KEYWORDS: PTO, Hydraulics, Fluid Power, Point absorber, WEC, Wave Energy

1. INTRODUCTION
Numerous types of Wave Energy Converters (WECs) are under development for harvesting the energy of the ocean waves, where several have reached the proof-of-concept
stage, showing it is possible to produce power, see e.g. [1] and [2] for a survey. A large
group of WECs bases on directly converting the waves into an oscillating mechanical
motion, e.g. point absorbers and multiple point absorber systems, see Fig. 1. Converting
the mechanical motion into electricity is performed by the Power Take-Off (PTO).
Today, PTO solutions for such systems are characterized by poor efficiencies and
reliabilities. The reason is that the waves induce slow irregular oscillations, which
requires processing of large alternating forces in order to extract power [3]. Resultantly,
fluid power is often a crucial part in the PTO system, being the only technology having
the required force densities. However, fluid power systems are often characterized by
poor efficiencies, especially at part load which is an inherit property of wave power.
A ratio of ten between peak and mean absorbed wave power is common, [4].

PTO
Pin=tPTO warm

Reciprocating
motion

Pout

warm

text

tPTO

Point absorber

Multiple point absorber

WaveVtar WEC

(a)

(b)

(c)

Figure 1: Point absorber type WECs and the Wavestar 600 kW WEC [10].
The PTO extracts energy from the waves by applying a damping torque PTO to the
float, such that the float performs work on the PTO system. As a result, the power
Phav = PTO arm is extracted from the waves. In order to maximise the amount of
harvested power Phav , PTO has to be controllable. How to generate the optimal PTO
force trajectory is referred to as Wave Power Extraction Algorithms (WPEA). It can
be shown, see e.g. [4] and [5], that the WPEA optimizing the amount of harvested
power requires the PTO to periodically transfer power to the float, i.e. four-quadrant
behaviour. This type of WPEA is termed reactive control. The main reason for using
reactive control is that the floats natural frequency should match the wave frequency.
As the wave frequency varies, the PTO is used to change the natural frequency, which
requires reactive power. This will described more thoroughly in section 3.2.
The focus of this paper is to investigate a PTO system for a multiple point absorber
WEC, based on using conventional hydraulic and electrical components. The PTO
system should be able to perform reactive control. The evaluation is performed for
the Wavestar 600 kW WEC [9] seen in Fig. 1c, consisting of 20 hemisphere shaped
floats, each 5 m in diameter. The advantage of a multiple point absorber system is the
increased power smoothing, [5].
The investigated PTO system is seen in Fig. 2. The PTO torque is produced by a
symmetrical cylinder, which is operated in closed circuit with a variable displacement
swash-plate pump/motor. The motor can stroke both positive and negative. Thus, the
motor converts the bi-directional cylinder flow into a uni-directional high speed rotation
for powering a generator. Pressure control is performed by the swash-plate motor in
order to control the torque PTO applied to the float. Accordingly, a bent-axis motor is
not utilized as its bandwidth is too low to perform pressure control. To avoid oversizing
motor and generator, an energy overflow system is added such that cylinder flows
exceeding the motor capacity are combined in a common line, powering an extra
generator. Similar PTO concepts to Fig. 2 are suggested in e.g. [6]. More simplified
hydraulic system have been proposed in [7] and [8], however, these can only provide
a constant PTO torque.

ext

Flushing
valve

Symm
cylind etric
er

pEO

Float
Float
module module
AC

DC bus

DC

Inverter
Energy overflow

Float
module

Energy overflow
Float
Float
module module

Float module

Figure 2: Investigated PTO system.

Float
module
PTO System

2. METHODS
The evaluation of the PTO system in Fig. 2 is based on performing time simulation
of the WEC and PTO for different wave inputs. Hence, to evaluate and optimize the
PTO system, a time-efficient system model, which reflects the WEC and described
PTO concept is developed. The components are modelled generically to accommodate
free and rapid change of component sizes and properties, and only the dominating
dynamics and power losses are included. The component models are combined into a
complete model from wave to grid. A suitable WPEA is identified for the system, and
three different overall PTO control strategies is developed and evaluated. To evaluate
the performance, the system is simulated for three representative sea states.
The system is divided into the sub-models seen in Fig. 3. The sub-models are arranged
in four main blocks, i.e. Wave Input, Wave and Float Interaction, Main PTO System
and Energy Overflow. The Wave Input and Float Arm blocks are fixed, whereas the
remaining blocks consist of components, where component sizes and control is to be
optimized to minimize losses and maximize power output of the WEC.
PTO,ref

arm WPEA
arm
arm
arm

PTO
ext

PTO
Float/arm

ext

Control

Gn,ref
ref

ref
Main PTO system
pB
pB
pB M
I,V,v I
tot
pA QB
darm
xc
QBc
V
G
arm
M
QAc QB
v
G
G
Fc
p
A
Fc
QA QA
Kinematics
Cylinder
Hyd. motor
Generator
pB pA
QA QB

Gn,ref

xc
vc p
A

QA
Hs, TP

v
P1

QB
PEO

Wave and float


interaction

Pout

Inverter

Inverter

PEO

pEO

Wave

QEO
Manifold
Energy overflow conversion
Energy overflow system

Figure 3: System model for evaluating and optimising PTO performance.


Multiple levels of optimization is performed, see Fig. 4. The WPEA algorithm calculates the time-varying torque ref to maximize the expected energy output of the WEC,
taking into account the PTO efficiency. Hence, an initial efficiency guess is required.
Next, the PTO control is tuned and the system is evaluated, and with a new efficiency,
the WPEA algorithm is updated. With a converged solution of WPEA and control, the
losses of the components are identified. With this knowledge the component sizes and
properties are re-adjusted and the loop is performed again.
Initial
system

Manual
update
system
System

Update
WPEA
WPEA

Update control

N
PTO

PTO
control

Evaluate

control
OK?

N
Eff.
convergY
ed?

N
PTO
Sys.
perform. evaluated
conv.? Y

Figure 4: Strategy for optimising and evaluating the PTO system.


3. MODELLING
The system is modelled in the following according to the sub-systems in Fig. 3.

3.1.

Wave Model

0.2
Wave Frequency [Hz]

0.1

0.3

Wave height h
[m]

Spectral density for representive sea states


4
State 3:
Hm0=2.50m Tp=6.44s
(Large)
Hm0=1.75m Tp=5.57s

Spectral density example SA(f)


Tp
Hm0= 2 m
Df TP = 6 s

4
2
0

1
0
-1
0

20

40
60
Time [s]

80

100

SA(f) [m2s]

[m s]

Ocean waves are irregular waves, i.e. waves with varying frequency and amplitude.
As a result, irregular waves are described by a wave amplitude spectrum, see Fig. 5.
A sea state or spectrum is usually represented by two quantities, the significant wave
height Hs and the peak wave period Tp . The significant wave height is the average of
the wave heights of the one-third highest waves, and Tp is the period where the waves
at average are highest. The Pierson-Moskowitz spectrum is utilized [11].

State 2:
2 (Medium)
State 1:
(Small)
0

Hm0=1.00m Tp=4.62s

0.4
0.1
0.3
0.2
Wave Frequency [Hz]

0.5

Figure 5: Wave spectra for sea states and an example of a corresponding wave.
From a spectrum, the individual wave components can be extracted as,

w,i (t) = 2SA (fi )f sin(2fi t)

[ m ] (1)

and a time series of an irregular wave, can be generated as a sum of wave components,
n

w (t) =
2SA (fi )f sin(2fi t + rand,i )

[ m ] (2)

i=1

where rand,i is a random phase for each component. This is known as the randomphase method. However, a more random wave which represents sea waves better is
obtained by filtering white noise using proper digital filters designed according to the
spectra, see [11] for reference. This method is utilized instead to generate time-series
of waves.
To evaluate the performance of the PTO, the three sea states shown in Fig. 5 are used,
which represents the range of waves in which the WEC should be able to produce
power. A wave time series has been generated for each sea state for evaluation of PTO
performance.
The following section describes how the wave interacts with the float.
3.2.

Wave and Float Interaction

The equation of motion for a float is given as,


Jarm+float arm (t) = wave (t) G (t) PTO (t)

[ Nm ] (3)

where Jarm+float is the inertia of float and arm, wave is the torque due to wave-float
interaction, G is the torque due to gravity and PTO is the torque applied by the PTO
system to the float arm.
To describe the interaction between wave and float, wave (t), linear wave theory is
often applied, as it gives an adequate description in the conditions in which a WEC
is producing energy, [14]. In linear wave theory simplified fluid dynamics is assumed

in order to apply linear potential theory. Resultantly, the wave-float interaction can be
described by superimposing three torques,
wave (t) = rad (t) + Arch (t) + ext (t)

[ Nm ] (4)

where ext (t) is the excitation torque an incoming regular wave applies to a float held
fixed, rad (t) is the radiation toque experienced from oscillating the float in otherwise
calm water, and Arch is the torque due to the Archimedes force, i.e. buoyancy.
The sum of the three torques gives,
wave (t) = J arm (t) hrad (t) arm (t) +Arch (t) + ext (t)
|
{z
}

[ Nm ] (5)

rad

where hrad is the impulse response function from float velocity to torque, describing
the hydrodynamic damping. The impulse response hrad can be viewed as a high order
damping term. The inertia term J represents the added mass, which contains the
effect, that when oscillating a float, it will appear to have a greater mass due to the
water being displaced along with the float.
The coefficients of Eq.5 are identified by applying the numerical tool WAMIT to
the float. WAMIT is a computer program for computing wave loads and motions of
structures in waves [12]. WAMIT also outputs a force filter, which can be applied to
w (t) to find ext (t).
Inserting Eq.5 into Eq.3 gives the equation of motion for the float,
(
)
1
arm =
kres arm (t) hrad (t) arm (t)+ ext (t) PTO (t) [ sm2 ] (6)
Jarm+float + J
where the sum of gravity and Archimedes term has been linearised around the draft of
the float, res (t) = Arch (t) G (t) arm (t)kres . Thus the input to float-arm subsystem
are the torques ext and PTO , and the output is the angular position and velocity of the
arm. To avoid the convolution term hrad (t) arm (t), the impulse response has been
fitted with a fifth-order system using Pronys methods [15].
The power extracted or harvested from a wave Phav is the product of the PTO torque
PTO and arm velocity arm . Hence, PTO should be controlled such that harvested energy
Ehav is optimised:

Ehav =
PTO (t)arm (t)dt
[ J ] (7)
0

In general, to maximise Eq.7 the float should be in phase with the exiting wave torque
ext , i.e. the natural frequency of the float and arm should match the incoming wave.
However, as the dominating wave frequency varies, the natural frequency will not
match. Consequently, the PTO system is used to move the natural frequency to increase
power capture. This is depicted in Fig. 6a, where the reference of the PTO torque is
generated by a damping term bPTO and a spring term kPTO :
PTO,ref = kPTO arm + bPTO arm

[ Nm ] (8)

The effect of including the stiffness term is illustrated in Fig. 6c, where a regular
wave is applied to the system. In the first case, linear damping is utilized, i.e. kPTO
is zero, and in the second case kPTO is non-zero (Reactive control). When using the
linear damping, the power Phav is always positive. In the second case, the power is

periodically negative, but the average harvested power is twice compared to linear
damping. Hence, more power is extracted, but it requires a PTO with four-quadrant
operation as reactive power is involved.
The optimal control in regard to power output depends on the PTO efficiency PTO ,
as the loss associated with the power required to move the natural frequency begins
to outweigh the extra harvested power. To find the optimal parameters bPTO and kPTO
as a function of PTO efficiency PTO , time-series simulation has been executed for
different efficiencies and sea states, and the values of bPTO and kPTO maximising the
average power Pout,avg have been found. These were found by using a simplex-based
optimisation algorithm. The simulation model is seen in Fig. 6a. Note that also a
saturation limit has been added to the PTO torque of 1 MNm, as it has been found to
be a reasonable limit in regard to harvested power versus requirements of the PTO.
The limit have been found by multiple simulations,where the limit have gradually
decreased. The results for the optimal values of kPTO and bPTO for sea state 1 and sea
state 2 is seen in Fig. 6b as a function of efficiency. Also the expected power outputs
are shown.
WPEA b
PTO
+ +
S

tPTO,max

Linear Damping (kPTO = 0)

warm
qarm

kPTO

Phav

tPTO
text

tPTO
text + S

qarm

Reactive Control

hPTO Pout
1
hPTO
1
1
s Eout tend Pout,avg

(a)

Pout,avg 5

bPTO
0
70
80 90 100
PTO efficiency hPTO [%]

bPTO , -kPTO

bPTO , -kPTO

-kPTO

1e6 Sea state 2


30
10
P out,avg
-kPTO
bPTO 15
5
0
70
80 90 100
PTO efficiency hPTO [%]

Pout,avg [kW]

10

Pout,avg [kW]

1e6 Sea state 1


10

(b)

Time [s]

(c)

Time [s]

Figure 6: Power extraction from waves, optimising the WPEA as a function of PTO .
3.3.

Hydraulic System - Main PTO System

The hydraulic system consist of a closed circuit pump/motor and a cylinder. The output
of the hydraulic system is the torque M for driving the generator, and the cylinder
force Fc acting on the float and arm.
To simplify dynamics, hose losses are neglected, such that the pressure in cylinder and
at motor are equal. Power loss associated due to hose loses will be discussed afterwards
when evaluating efficiency.
Using the flow continuity equation the following expression is obtained for the symmetric cylinder,
e,1
pB =
(QB v c Ac )
[ bar
] (9)
s
Ac xc + V0,1
e,2
pA =
(QA + v c Ac )
[ bar
] (10)
s
Ac (xc,max xc ) + V0,2
where Ac is the cylinder area, xc,max is the maximum stroke of cylinder and the volumes
V0,1 and V0,2 are volumes of hoses.

The cylinder force is calculated as,

c ) ; pAc vc > 0
tanh(avc )|pAc |(1
(
)
Fc = pAc Ffric ; Ffric =
1
tanh(avc )|pAc |
1 ; pAc vc 0
c

[N]

(11)

where p = pB pA . The cylinder friction Ffric is defined such that the cylinder has a
constant efficiency of c . The function tanh(avc ) is used instead of sign(vc ) to avoid
discontinuity, where a adjust the steepness around zero velocity.
The hydraulic motor is a closed-circuit swash-plate pump. The model is based on
measured efficiency plots for different pump sizes. A typical efficiency plot is seen in
Fig. 7a for 100% and 50% displacement respectively. Using Schlsser formula [13] for
friction and flow loss, the following expression are used for motor torque and flow:
QM,nom (M , p) = D M pCQ1
)
(
2
M,nom (M , p) = D p C 1 + C 2 p + C 3 M + C 4 M

[ ms ] (12)
[ Nm ] (13)

The efficiency of the fitted model is seen in Fig. 7b. If the fitted pump is referred to
as the nominal size, other motor size is obtained by scaling this model,
)
(
D,new rated,new
rated,nom
3
QM,new (M , p) =
QM,nom
M , p
[ ms ] (14)
D,nom rated,nom
rated,new
)
(
D,new
rated,nom
M,new (M , p) =
M,nom
M , p
[ Nm ] (15)
D,nom
rated,new
where D,new and rated,new is the displacement and rated speed of the new motor
respectively.

20

0.89

30

=0.5
0.

89

05
0.9
0.91

80
0. .815
0
0.82

20

0.9

10

10
0
0

40

0.6
8
0.
75

=1
80
0. .825
0 4
0.8

0.77
0.8
4
0.8
8

0.7
0.
75

=0.5
5
.87

p [bar]

p [bar]

30

0.7
5
0.8
25
0.
85

=1
40

1000
2000 3000 0
n [RPM]
(a)

1000
2000 3000
n [RPM]

0
0

1000

2000 3000 0
n [RPM]
(b)

1000

2000 3000
n [RPM]

Figure 7: Efficiency plots for a typical swash-plate pump, measurements and model.
To replenish the fluid leaked by the motor and to give the necessary flushing for cooling
and filtering, a charge/booster pump is installed. The charge pump is set to maintain
a pressure of 17 bar, which is a low but sufficient charge pressure. According to data
sheets a rule-of-thumb is to size the charge pump to be 10% of the installed displacement. Consequently, to model the required power for flushing, a fixed displacement
pump running together with the motor is used:
Pflush = D,charge pcharge M = 0.1 D,M pcharge M

[W]

(16)

Regarding swash-plate dynamics, it is assumed that the swash-plate control is fast


enough for controlling the pressure as the pressure reference is dictated by the wave
frequency, which is low. Hence swash-plate dynamics is omitted.

3.4.

Generator and Inverter - Main PTO System

The generator setup consists of an asynchronous generator and an inverter for grid
connection and to enable variable speed control. The input to the generator is the
hydraulic motor torque and the output of the power system is the angular velocity of
the generator M and output power.
The electrical properties of the generator is modelled in steady state. This assumed to
be adequate, as the inverter handles the dynamics. An equivalent circuit for a phase of
a three-phase Delta-connected induction motor is seen in Fig. 8b, where denotes the
motor slip. The slip is defined as,
npp Gn
[ ] (17)
=1
V
where npp is the number of pole pairs and V is the frequency of the supply voltage. The
resistor R2 1
represents the mechanical input to the generator, i.e. Gn Gn = I2 R2 1
,

for reference see e.g. [16].


As the current I2 IM , the generator torque is given as,
2
3npp R2
VRMS
Gn =
(
)2
V
R2 + (V (L1 + L2 ))2
R1 + R2 + 1

[ Nm ]

(18)

where VRMS is the RMS-value of the line-to-line voltage.


The steady-state phase current IP of the generator is given as:


Vp
VP
Z2 (s)ZM (s)
IP =
, HGn (jV ) =
=
+ Z1 (s)
|HGn (jV )|
IP
Z2 (s) + ZM (s)
s=jV
where Z1 = R1 + L1 s, Z2 =

R2

+ L2 s and ZM =

(19)

RFe sLM
.
RFe +sLM

The electrical output power of the generator is three times the power per phase:

[ W ] (20)
PGn,out = 3VP IP cos( HGn (jV )) = 3VL IL cos( HGn (jV ))
As hydraulics motors are typically operated in the range of 500-2500 RPM, a 4-poled
generator is used, as it operates at 1500 RPM at a voltage frequency of 50 Hz. A
nominal model is based on the measurement of an asynchronous high-efficiency 4-pole
55-kW generator, where the parameters of the equivalent circuit has been identified.
The model result is seen in Fig. 8c, where the efficiency is plotted as a function of load
at 1500 RPM. The torque characteristic as a function of slip produced by the model
seen in Fig. 9. Other generator sizes are obtained by scaling the nominal model similar
to the methods applied to the hydraulic motor.
The angular velocity of the generator is given by,
1
Gn =
(M Gn )
[ s12 ] (21)
JGn + JM
where JGn and JM is the inertia of the rotor of the generator and hydraulic motor
respectively.
The torque of the generator is controlled by an inverter. The inverter is modelled
with a constant efficiency inv of 95%. To control the torque of the generator, the
expressions for finding the appropriate voltage and voltage frequency for the generator
is implemented in the inverter, see Fig. 9. The modelled 55 kW generator may be
operated at 150% load for two minutes, and may in shorter periods also be operated
at 200% load. Consequently, the inverter is limited to run the generator at 200% load.

V13

V12

IL1
3

VP=VL
-

R1

V23

L1

L2

I2

R2

IM Z2

Z1
VP

IL3
IL2

IP
+

R2 1-

ZM
LM

RFe

Efficeincy [-]

IP1 =

Back emf

IL1

0.8
0.6
0.4
0.20

(b)

40

(b)

80 120 160 200


Load [%]
(c)

Figure 8: Per phase equivalent circuit for an induction motor.

[Nm]

800 Generator torque at


1500RPM
400

Gn,ref

Inverter

Gn,ref

V= nppGn
1-

-400
Pinv,out

-800
-0.04 -0.02 0 0.02 0.04
Slip [-]

inv
1
inv

Gn 400 V
VL 2
50 Hz
VL
V
V

Generator
VL
V
Gn
PGn,out
M

M
Gn

IL,VL,v
G

Figure 9: Torque characteristic of generator and torque control of generator.


3.5.

Energy Overflow System

If a float cylinder produces more flow than the hydraulic motor can consume, the
pressure builds up in the cylinder until opening the check-valve to the Energy Overflow
(EO) system. Due to accumulators the EO system is assumed to be operating at a
steady high pressure pEO . As simulating the behaviour of the EO system would require
to simulate all 20 floats, a fixed efficiency is assumed instead for the EO system.
A model of a check-valve to connect the cylinder to the EO system is included, which
determines the flow QEO entering the EO system. The remaining EO system consist
of a long pipeline to connect overflow from all floats, a fixed displacement motor, a
generator and an inverter. The following efficiency EO is used for the EO system:
EO = pipe,avg M,avg Gn,avg inv,avg = 0.95 0.90 0.90 0.95 = 0.73

(22)

Hence, if the power delivered to the EO system is PEO,in = pEO QEO , the power delivered
to the grid from the EO system is PEO,out = PEO,in EO .
3.6.

Calculating Efficiencies and Power Losses

To optimize and evaluate the PTO system, power losses and efficiencies of the individual components are calculated. If the instantaneous power in and out of a component
are Pin and Pout respectively, the efficiency and losses are calculated as:
tend
tend
1
1
Pin (t)dt ,
Pout,avg =
Pout (t)dt
(23)
Pin,avg =
tend 0
tend 0
Pin,avg
=
,
Ploss,avg = Pin,avg Pout,avg
(24)
Pout,avg
As power transfer in both direction occur, the efficiency does not specify the component
efficiency but a resulting efficiency, e.g. a component with a fixed efficiency will show
a lower efficiency when reactive power is involved.

4. DESIGN OF PTO
The design of PTO is organized by first taking a brief view on the requirements,
combined with an initial sizing of the PTO components. For evaluating the PTO
performance, three different control strategies for control of motor and generator are
developed and applied. Based on the control strategies and initial sizing, the PTO
system is optimised.
4.1.

Requirements and Initial Sizing of PTO

The requirements of the PTO system is to be able to produce power at sea states
ranging from a significant wave height of about 1 m to 3 m. To design and evaluate the
PTO system, the three representative sea states defined in Fig. 5 are utilized.
In order to quantify the requirements, simulation of cylinder forces Fc and power
input have been made by applying the three different sea states to the float model, see
Fig. 10, where a power extraction algorithm tuned for a PTO with a efficiency of 70%
is applied.
WPEA optmized for 70% efficient PTO
p

1
Ac

kPTO
Pharvested

Float Dynamic Model


Fc

Wave

bPTO

+
+

ext

Fc

ext

xc
vc

Ac

Roshage float

Figure 10: Simulation for estimating requirements of PTO.


The flow and pressure requirements depends on the cylinder size. To minimize flow
losses, the pressure should be as high as possible, however, according to the typical
pump efficiency data displayed in Fig. 7, the efficiency drops above 300 bar. As a result,
the cylinder is designed such that the maximum required force is obtained at a delta
pressure of 300 bar. To give a torque of 1 MNm, a cylinder force of 420 kN is required,
yielding Ac = 140 cm2 . The pressure of the EO system is set to 325 bar. The result
of applying the cylinder area to the simulation is seen in Fig. 11, where the required
pressures and flows are seen.
According to Fig. 11 the peak power in sea state 3 is about 250 kW, however having
e.g. a 250 kW hydraulic motor and generator producing an average power of 29 kW
will lead to very poor efficiencies. As sea state 2 is more frequently occurring, the
first iteration of a PTO is based on this sea state. The peak flows are approximately
250 L/min. If the generator is set to run at a fixed speed of 1500 RPM, a 160 cc motor
is required.
If the generator is assumed to be able run at 100% overload in shorter periods, the
generator matching a hydraulic motor is found as,

where Gn,max

M,max = D pmax
[ Nm ] (25)
1
PGn,norm = M,max Gn,max
[ W ] (26)
2
is the chosen maximum motor/generator speed and D is the stroke

[kN]

500

[bar]
[L/min]

Sea State 2

Sea State 3

Fc

Fc

QM

QM

QM

0
-500
300

[kW]

Sea State 1
Fc

0
-300
500
0
-500
250
125
0
0

Phav

Pavg = 4.1kW

10
15
Time [s]

20

Pavg = 15.6kW

Phav

10

20
Time [s]

30

Phav 28.93

10

Pavg = 28.9kW

20
Time [s]

30

Figure 11: Power, flow and pressure at different sea states.


displacement of the hydraulics motor in m3 /s. The pressure pmax is maximum allowed
pressure, in this case 300 bar.
The generator size matching a 160 cc motor is then a 60 kW unit. If the cylinder
produces more flow than the motor can consume, the pressure builds up until opening
the check-valve to the Energy Overflow (EO) system.
4.2.

PTO Control Strategies

The objectives of the PTO control is to track the cylinder force reference produced
by the WPEA algorithm while minimizing power losses. The control inputs are the
displacement control of the hydraulic motor and the generator torque Gn .
Three control strategies are tested:
1) Fixed speed of generator according to sea state and force control using
2) Slowly varying generator speed according to average peak flow requirement.
3) Generator speed is controlled to keep motor displacement at maximum.
A comparison of the three control strategies is shown in Fig. 12. In the first strategy
the generator is running at 1500 RPM, as a result, the motor is at part stroke most of
the time. In strategy 2 the engine speed is varied according to the average required
flow, which leads to the motor being closer to full stroke. In strategy 3 the generator
speed is continues controlled to get the motor to 95% stroke, however this requires the
generator speed to oscillate together with the wave. To avoid using to much electrical
power to accelerate the generator inertia, the generator is not allowed to operate in
motor mode above e.g. 1000 RPM.
4.3.

Evaluation and Optimization of PTO

The initial PTO design is evaluated for each sea state using the three described control
strategies. The results are seen in Tab. 1, Tab. 2 and Tab. 3, where PL and denotes
average power loss [kW] and efficiency [%] respectively, and the columns Pin and Pout
are the average harvested power and power output of the system in [kW]. The overall
efficiency and power output is best for control method three. It overall gives a 2 kW
higher output. Control strategy 1 and 2 give approximately the same power output,

Control Strategy 2

Control Strategy 3

Time [s]

Time [s]

Time [s]

Generator
Displacement
speed Gn [RPM] control [-]

Control Strategy 1

Figure 12: Comparison of the three control strategies.


however, strategy 2 has a better efficiency, so the required cooling would be lower.
From the tables it seen that generally, the hydraulic motor is dropping below 80%
efficiency in sea state 1 and 2. Also, the generator efficiency is dropping low in sea
state 1. Resultantly, both the hydraulic motor and the generator are downsized.

SS
1
2
3

SS
1
2
3

SS
1
2
3

Table 1: Initial system with control strategy 1


Cylinder
Motor
Flush Generator Inverter EO
PL

PL

Pflush PL

PL

PL
0.31 94.4 1.52 69.8 0.34
0.80 74.8 0.27 88.7 0.00
1.32 94.3 5.14 74.3 0.68
1.74 87.7 1.10 91.2 0.29
2.39 94.6 7.26 78.5 0.68
2.21 91.5 1.73 92.7 1.58

Pin
5.48
23.2
44.0

Total
Pout
2.09
12.3
27.1

38.2
52.9
61.6

Table 2:
Cylinder
Motor
PL

PL

0.34 94.3 1.53 72.1


1.20 94.4 3.84 78.8
2.25 94.6 6.12 80.3

Initial system with control strategy 2


Flush Generator Inverter EO
Pflush PL

PL

PL
0.34
0.80 77.9 0.29 89.6 0.00
0.43
1.27 90.8 0.95 92.5 0.34
0.46
1.78 92.7 1.54 93.2 1.74

Total
Pin Pout
5.97 2.52
21.4 12.8
41.9 27.0

42.2
59.7
64.4

Table 3:
Cylinder
Motor
PL

PL

0.37 94.2 1.42 75.8


1.31 94.3 3.75 81.3
2.35 94.6 5.89 82.8

Initial system with control strategy 3


Flush Generator Inverter EO
Pflush PL

PL

PL
0.30
0.75 82.0 0.27 92.1 0.00
0.39
1.24 92.2 0.96 93.4 0.27
0.42
1.79 93.6 1.60 93.9 1.44

Pin
6.39
23.2
43.8

Total
Pout
3.13
14.6
29.3

49.0
63.1
66.9

By optimising on the simulation model, it has been found that to have good efficiencies
at sea state 1 and 2, it is best to utilize a smaller hydraulic motor and instead increase
the speed in high power periods, e.g. up to 2500 RPM. After optimising control and
components, the values in Tab. 4 have been identified.
Table 4: Optimized system parameters.
Cylinder Area
Motor size
Generator size
Maximum speed
2
3
140 cm
80 cm
45 kW
2500 RPM
4.4.

Optimized Solution

The results for the three control strategies applied to the optimised solution are seen
in Tab. 5, Tab. 6 and Tab. 7. Compared to the previous solution, the harvested power

have been reduced with 2-5 kW, however the output power remains unchanged, which
gives a rise in efficiency. The average efficiency improvement is 5%. The reduction in
harvested power is due to the main PTO system more often becoming saturated, e.g.
the flow from the cylinder cannot be consumed by the motor, leading to the pressure
rising to pEO . As a results, the system cannot track the optimal cylinder force trajectory.
Comparing the results of the three control strategies, strategy 1 and 2 outputs the
same amount of power, but less is harvested in strategy 2, giving it a better efficiency
score. Strategy 2 harvest less power, as the generator speed is too low in periods. To
improve the control, the controller must be improved in predicting when to increase
the generator speed. Control strategy 3 shows the best results, harvesting power as
control strategy 1 while reducing losses. Also this strategy is able to maintain a motor
efficiency above 80% in all sea states.

SS

One of the reasons for the relative good efficiency of strategy 3 is, that the generator
power is mostly positive. This is seen in Fig. 13, where the generator power is compared
for the three strategies. When the system requires reactive power this is actually taken
from the kinetic energy saved in the inertia of motor and generator when the generator
speed is reduced. In the two other strategies, the reactive power is drawn from the
inverter, i.e. the power travels through both the generator and inverter.
Table 5: Optimised system with control strategy 1
Cylinder
Motor
Flush Generator Inverter EO
Total
PL

PL

Pflush PL

PL

PL
Pin Pout
1 0.20 94.7 0.79 77.3 0.17
0.58 76.8 0.17 91.3 0.00 3.83 1.77 46.2
2 1.12 94.5 3.18 80.9 0.45
1.58 87.8 0.88 92.3 0.47 20.2 12.1 60.0
3 2.19 94.7 4.78 83.5 0.50
1.99 91.6 1.47 93.2 2.21 41.2 27.4 66.6

SS
1
2
3

SS
Power output
of generator [kW]

1
2
3

Table 6: Optimised
Cylinder
Motor
Flush
PL

PL

Pflush
0.30 94.4 1.18 76.5 0.28
1.04 94.5 2.57 83.4 0.35
2.05 94.7 3.92 85.0 0.37

system with
Generator
PL

0.86 75.9
1.22 90.3
1.52 93.0

control strategy 2
Inverter EO
PL

PL
0.25 90.6 0.00
0.77 93.2 0.47
1.27 93.8 2.32

Pin
5.45
18.9
38.8

Total
Pout
2.45
12.1
26.7

44.9
64.0
68.9

Table 7: Optimised
Cylinder
Motor
Flush
PL

PL

Pflush
0.34 94.3 1.00 81.8 0.22
1.13 94.5 2.38 86.4 0.30
2.15 94.7 3.77 87.4 0.33

system with
Generator
PL

0.73 82.8
1.10 92.6
1.46 94.3

control strategy 3
Inverter EO
PL

PL
0.25 92.9 0.00
0.74 94.6 0.35
1.29 94.7 1.91

Total
Pin Pout
5.93 3.28
20.5 14.2
40.9 29.3

55.3
68.9
71.8

80

Control strategy 1
Control strategy 2
Control strategy 3

60
40
20
0
-20
5

10

15

20 Time [s]

25

30

35

Figure 13: Power output of the generator for the three control strategies.

40

5. DISCUSSION
Based on the modelled system, the best efficiency achievable of the investigated PTO
system is between 55% and 72% in the production range of the wave energy converter.
These results are obtained using control strategy 3, see the summarized results in Tab.8.
However, control strategy 3 requires the motor and generator to speed up and down
according to the float velocity, e.g. from low to high speed two times per wave period.
Hence, lifetime of the components may be reduced by this scheme. Consequently, an
evaluation of components lifetime should be evaluated before using this control scheme.
Nevertheless, control strategy 3 is good for showing the best achievable efficiency with
off-the-shelf components.

SS
1
2
3

Pin
3.83
20.2
41.2

Table 8: Optimised system performance summary.


Control 1
Control 2
Control 3
Pout

Pin
Pout

Pin
Pout

1.77
46.2
5.45
2.45
44.9
5.93
3.28
55.3
12.1
60.0
18.9
12.1
64.0
20.5
14.2
68.9
27.4
66.6
38.8
26.7
68.9
40.9
29.3
71.8

The output power of control strategy 1 and 2 is in average 2 kW lower, however these
solutions would be easy implementable. As such, control strategy 2 is preferred, as
the speed of the motor and generator are reduced during low power periods, e.g. wear
and tear is reduced. Regarding generator efficiency, when operating at low speeds as
in sea state 1, the generator shows a poor efficiency. This might be raised to about
90% by reducing the magnetizing current, which gives the losses. For same reason, a
permanent magnet generator will show better result at sea state 1, but it will not give
a significant improvement at sea state 2 and 3.
To assess the accuracy of the results, an overview of the model properties is shown
in Tab. 9, along with future improvements. If the transient behaviour was included
for the generator, features as reducing the magnetizing current could be evaluated.
The remaining features to be added represents additional losses. Hence it would be
reasonable to subtract 3-5% from the results.
Table 9: Optimised system with control strategy 1
Included
Add

Cylinder
Pressure dynamics, hose volumes, const. eff.

Motor
Friction,
flow
losses,
power
for flushing.

Hose loses, cyl.


friction model.

Power for stroke


control.

Generator
Mechanical
dynamics,
steady
state
electrical model.
Electrical
dynamics.

Inverter
Const.
efficiency,
electrical model
of generator ctrl.
Efficiency as a
function of load.

EO
Constant
efficiency.

Model EO with
20 floats

To improve efficiency in the future, the swash-plate motor could be replaced with
upcoming components like digital displacement motors [17], [18], which are characterized by having excellent part load efficiency. Also, the current solution relies on
being a multiple absorber system to provide power smoothing. To change the concept
to include more smoothing would be advantageous, also to reduce component sizes.
Finally, PTO systems characterized by having hydraulic motors for e.g. 2 or 4 floats
on a common shaft to power a larger generator have also been investigated during this
research. This will increase the generator efficiency, but only control strategy 1 would
be applicable. Hence, the motor efficiency drops below 80%, ie. this system setup will
not improve the efficiency.

6. CONCLUSION
By optimizing the proposed PTO system for a multiple point absorber system, the best
achievable efficiency with conventional components is in the range from 52% to 68%
under the different wave conditions at which a wave energy converter is producing
power. This emphasizes the need for new component types like digital-displacement
motors, or entirely new PTO concepts in order to utilize wave energy from point
absorbers. New PTO concepts are therefor currently under investigation.
REFERENCES
[1] A. Muetze and J.G. Vining. Ocean wave energy conversion - a survey. In Industry
Applications Conference, 2006.
[2] B. Drew, A. Plummer and M.N. Sahinkaya. A review of wave energy converter
technology. Proceedings of the Institution of Mechanical Engineers, Part A: Journal
of Power and Energy, 223 (8), 2009.
[3] J Cruz. Ocean Wave Energy: Current Status and Future Perspectives, 2008, Green
Energy and Technology Series, ISBN:978-3-540-74894-6.
[4] S.H. Salter, J.R.M. Taylor and N.J. Caldwell. Power Conversion Mechanisms for
Wave Energy. Proceedings of the Institution of Mechanical Engineers, Part M:
Journal of Engineering for the Maritime Environment, 2002.
[5] J. Falnes. Optimum Control of Oscillation of Wave-Energy Converters. International Journal of Offshore and Polar Engineering, vol. 12 2002.
[6] E. Wood. Power generation systems in buoyant structures. US-patent: US4158780,
1979.
[7] A.F. de O. Falcao. Modelling and control of oscillating-body wave energy converters with hydraulic power take-off. Ocean Engineering, vol. 35, 2008.
[8] A. Babarit,M. Guglielmi and A.H. Clement. Declutching control of a wave energy
converter. Journal of Ocean Engineering, vol. 36, 2009.
[9] L. Marquis, M. Kramer and P. Frigaard. First Power Production figures from the
Wave Star Roshage Wave Energy Converter. 3rd International Conference and
Exhibition on Ocean Energy, 2010.
[10] Wave Star A/S. http://www.wavestarenergy.com/
[11] M.J. Ketabdari and A. Ranginkaman. Simulation of Random Irregular Sea
Waves for Numerical and Physical Models Using Digital Filters. Transaction B:
Mechanical Engineering Vol. 16, No. 3, 2009.
[12] http://www.wamit.com/
[13] K. Huhtala, J. Vilenius, A. Raneda and T. Virvalo. Energy losses of a tele-operated
skid steering mobile machine. Power transmission and motion control, 2002.
[14] J.Falness. Ocean Waves and Oscillating Systems. ISBN: 0-521-78211-2
[15] G. De Backer. Hydrodynamic Design Optimization of Wave Energy Converters
Consisting of Heaving Point Absorbers. Ph.D. thesis, 2009.
[16] A. Hughes. Electrical motor and drives. Third Edition, 2006, ISBN:0-7506-4718-3
[17] G.S. Payne,U.B.P. Stein, M. Ehsan, N.J. Caldwell and W H.S. Rampen. Potential
of Digital Displacement Hydraulics for Wave Energy Conversion. In proc. of the
6th European Wave and Tidal Energy Conference, Glasgow, UK, 2005.
[18] M. Ehsan, W.H.S. Rampen and J.R.M Taylor. Simulation and Dynamic Response
of Computer Controlled Digital Hydraulic Pump/Motor System Used in Wave
Energy Power Conversion. In proc. 2nd European Wave Power Conference, Lisbon,
1995.

You might also like