Next Article in Journal
The Impact of Ethical Leadership on Financial Performance: The Mediating Role of Environmentally Proactive Strategy and the Moderating Role of Institutional Pressure
Previous Article in Journal
Intellectual Property Pledge Financing and Enterprise Innovation: Based on the Perspective of Signal Incentive
Previous Article in Special Issue
Effective Mooring Rope Tension in Mechanical and Hydraulic Power Take-Off of Wave Energy Converter
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Study of a New Wave Energy Converter with Perturb and Observe Maximum Power Point Tracking Method

by
José Carlos Ugaz Peña
,
Christian Luis Medina Rodríguez
and
Gustavo O. Guarniz Avalos
*
Faculty of Engineering, Universidad Tecnológica del Perú, Lima 15046, Peru
*
Author to whom correspondence should be addressed.
Sustainability 2023, 15(13), 10447; https://doi.org/10.3390/su151310447
Submission received: 19 May 2023 / Revised: 12 June 2023 / Accepted: 29 June 2023 / Published: 3 July 2023

Abstract

:
Ocean waves contain the highest energy density among renewable energy sources. However, harnessing the energy from ocean waves represents a challenge because wave energy converters (WECs) must be designed to have great survivability and efficiency. The power production challenge of any WEC depends on the power take-off (PTO) system efficiency. Maximum power point tracking (MPPT) algorithms have been widely applied in renewable energy from photovoltaic and wind sources, and have subsequently been adapted to wave energy converters (WECs). Energy extraction is optimized by applying MPPT, resulting in an increase in efficiency. This study aims to address the analysis of the influence of the perturb and observe MPPT in the electrical power performance of a WEC composed of a point absorber, a hinged arm and a direct mechanical drive PTO system. The PTO is characterized by a pulley system, a counterweight, one-way bearings, a gearbox, a flywheel and an electric generator; in the present study it is considered to be a cylindrical point absorber. The linear theory and the viscous damping effect are applied to analyze the hydrodynamic behavior of the point absorber. Regarding the two generators considered in the present study, the contribution of MPPT is greater for the low power generator; the high values of the capture width ratio (CWR) occur at low values of period and wave height, showing the maximum value in the high-power generator.

1. Introduction

The use of fossil fuels is the main cause of environmental damage and renewable energies represent the main and safest alternative to counteract the effects of fossil fuels [1,2]. The most available renewable energy is marine energy; the ocean makes up 71% of the Earth and it could meet the electricity demand of the whole world [3]. The motion of gravity waves is studied through different numerical methods [4,5,6]. Wind and solar energy have undergone great technological development; however, the production of wave energy is more constant and predictable [7]. Waves have the highest energy density of all renewable energy sources [8]. The technological development of wave energy converters (WECs) is increasing worldwide [9]. Due to its small size, low complexity and low cost, the point absorber is the most studied WEC type [10,11,12,13]. The accessibility of these devices is important for maintenance and repair, as environmental conditions can make this task difficult [14]. For this reason, WECs based on a point absorber connected to shore via an articulated arm are generating much interest due to their easy installation and maintenance [15]. A very important component of the WEC is the power take-off (PTO) system, which represents the mechanism that transforms the kinetic energy of the component that interacts with the wave to generate electricity. The efficiency of the WEC depends mainly on this system; it can be considered as the brain [11] because it can influence the dynamics of the component that interacts with the wave to maximize energy capture. The PTO system is based on a hydraulic [16], pneumatic [17,18] or mechanical system [9,19]; hydraulic and pneumatic components decrease the efficiency of WECs [9]. For this reason, in recent years, the application of the direct drive mechanical system and the direct electrical drive system has increased [11]. As reported in [9], PTO systems based on direct mechanical drive are very well-known and used by 31 developers from all over the world. This system can be composed of a rack and pinion, unidirectional bearings, a belt drive system, a pulley system or a screw mechanism [20,21,22,23,24,25]. The maintenance cost of these components is the challenge of this system because they are exposed to high load cycles; the size of the gearbox represents another problem in some cases [9].
The control system in WECs is introduced to maximize energy capture and provide a physical device constraint that must not be violated. This objective is achieved by the force/torque manipulation of the PTO system. However, the goal should not be maximizing conversion efficiency, but minimizing the cost of converted energy [26]. During the last few years the control system has progressed; however, there is still the challenge of reproducing the non-linearity of the WEC system with respect to the hydrodynamics and the PTO system [26,27]. One of the most important characteristics to take into account when extracting wave energy is its high variation over time. For the efficient operation of WEC systems, all the available wave energy should be converted into electricity, even in the face of a variable wave regime [28]. This variability of the available power is common to other renewable energy sources, such as photovoltaics and wind, where researchers have introduced the concept of maximum power point tracking (MPPT) to adapt the converter to the variation of the energy source and achieve higher efficiency [29,30]. Most of these methods, such as perturb and observe, are heuristics algorithms where no model of the plant is required [31,32] and optimization is based on a gradient-ascent approach [28]. MPPT methods have been adapted for WECs in [28,32,33,34]. The implementation of MPPT methods imposes certain behavior on the generator currents that are related to the torque and force of the PTO, thus influencing the converter performance. Furthermore, operational limits, such as speed, torque and power, are critical considerations when designing and operating electrical generators [35,36]. These limits determine the maximum output capacity and performance of a generator, and exceeding them may result in system failures, damage to equipment and even safety hazards. Speed limits, for instance, are essential to prevent mechanical stress, overheating and even rotor or stator damage. Torque limits, on the other hand, determine the maximum force that a generator can produce and transmit without causing damage or malfunction. Therefore, WECs should be designed within this operational limit to ensure reliable and safe energy production and avoid costly and dangerous accidents. Furthermore, the proposed MPPT method also considers these limits by disabling the optimization when any of them is reached.
The aim of this study is to demonstrate the new WEC’s performance and the influence of the MPPT perturb and observe (P&O) method on electric power generation. The WEC is based on a cylindrical point damper, articulated arm and direct mechanical drive PTO system. The power take-off is made up of a pulley system, a counterweight, one-way bearings, a flywheel, a gearbox and an electric generator. The main components of the power take-off that influence the generation of energy, the variation of the transmission ratio and the power of the electric generator are analyzed. In the present study, regular waves of three different periods T p and heights H w are considered; linear wave theory and the effect of viscous damping are applied to calculate the hydrodynamic force at the point of absorption. This force is considered in the wave-to-wire model to describe the operation of the WEC; this explains the coupling and decoupling between the point absorber and the electric generator. The increase in power generation due to MPPT and the WEC capture width ratio are detailed.

2. Wave-to-Wire Model

The proposed wave energy converter is shown in Figure 1, the PTO system is composed of a pulley system, a counterweight, a gearbox, a flywheel and an electric generator, see Figure 2a. The aim of the counterweight is to maintain tension on the pulley system. The pulley system transforms the oscillation movement of the point absorber into a rotation movement; the cable wraps around the main pulleys in order to transmit torque to the primary shaft, see Figure 2a; the main pulleys use one-way bearings to transmit rotation in one direction. The gearbox increases the rotation velocity; the output shaft of the gearbox is connected to the secondary shaft through a coupling based on one-way bearing that defines the coupling and uncoupling of the point absorber and the electric generator. An MPPT control is implemented to compute the torque of the electric generator. The flywheel is mounted on the secondary shaft to store the captured kinetic energy. The stored energy is used when the point absorber and the electric generator are uncoupled. The point absorber considered in the present work has been studied by [20] and is described in Table 1.
The complexity of the WEC operation is simplified in Figure 3. The schematic model considers a counterweight for each mean pulley, unlike Figure 2. The main dimensions of the schematic model are shown in Table 2.

2.1. Dynamic Model of the WEC

As a first approximation to the analysis of a new WEC, it is assumed that the point absorber is always in a vertical position and the weight of the articulated arm is neglected [37,38]. The equation of motion of the point absorber around the hinge O is defined by Equation (1).
m 1 g R s i n θ + F h R s i n θ F c 1 R s i n α 1 + F c 2 R s i n α 2 = J θ ¨
where J is the moment inertia of the center of gravity of the point absorber with respect to the point O; θ ¨ is the angular acceleration; m 1 is the mass of the point absorber; g is the gravity acceleration; F c 1 and F c 2 are the cable tensions of cable c1 and cable c2, respectively; F h is the hydrodynamic force; R is the length of the hinged arm ( O A ¯ distance). The variables α 1 and α 2 are described in Figure 3.
In Figure 3, the cable length L 1 between points A and B is defined in Equation (2). The relation of the time variation of the length L 1 to the angular velocity δ ˙ 1 of the main pulley 1 of radius r is described in Equation (3). The cable length L 2 between points A and C is defined in Equation (4). The relation of the time variation of the length L 2 to the angular velocity δ ˙ 2 of the main pulley 2 of radius r is described in Equation (5). The relation of the vertical projection of the hinged arm to the z displacement of the point absorber is defined in Equation (6).
L 1 = R 2 + h 2 2 2 R h 2 c o s θ 1 / 2
L ˙ 1 = δ ˙ 1 r
L 2 = R 2 + h 3 2 2 R h 3 c o s π θ 1 / 2
L ˙ 2 = δ ˙ 2 r
R c o s θ = h 1 + h 2 z
The linear potential flow theory is applied to calculate the hydrodynamic force. The direction of the wave considered in the present study is shown in Figure 3b; therefore, the oscillation of the point absorber is generated solely by the wave force in heave. The hydrodynamic force is defined in Equation (7).
F h = F b + F r + F e + F v
where F b is the buoyance force,   F r is the radiation force, F e is the excitation force and F v is the viscous damping force. The wave elevation is described by Equation (8).
z w = A w c o s ω t
where A w is the wave amplitude and ω is the wave angular frequency.
The gravity and buoyancy force define the restoring force F r s , see Equations (9) and (10).
F r s = F b m 1 g = c 33 z
c 33 = ρ g A
where ρ is the density of the water, A is the cross section of the point absorber and c 33 is the stiffness coefficient.
F r is defined according to [39], see Equation (11), whereby the radiation convolution integral is solved by a direct method.
F r = a z ¨ 0 t K r t τ z ˙ τ d τ
where K r is the retardation function and a is the added mass. According to [40], K r can be calculated by Equation (12).
K r t = 2 π 0 b 33 ω cos ω t d ω
where b 33 is the damping coefficient. The retardation function is shown in Figure 4, and the damping coefficient and the added mass of the point absorber are shown in Figure 5.
The excitation force F e is given by Equation (13).
F e = H w 2 F 33 cos ω t ϕ 33
where H w is the wave height, F 33   is the wave force and ϕ 33 is the phase shift. The last two parameters are obtained in Figure 6.
The viscous damping force F v is calculated based on the drag force described in [41], but rewritten according to [42], see Equation (14).
F v = 1 2 ρ C d A z ˙ u z ˙ u
where   ρ is the water density, C d is the drag coefficient, z ˙ is the body velocity,   u is the undisturbed velocity of the fluid and A is the cross-sectional area of the floating body. The value of the drag coefficient C d assumed in this study is 1.85; the drag coefficient of the point absorber considered has been calculated by [20].

2.2. PTO System

The PTO system depends on the cable tensions F c 1 and F c 2 , see Figure 3. The cable wraps around each of the main pulleys and transmits torque to the shaft; it is assumed that the cable does not slip on the main pulleys. The operation of the PTO system is defined by the coupling and uncoupling of the point absorber and the electric generator. Considering the coupling of the PTO in the upward movement corresponding to the main pulley 1 (the main pulley 2 rotates freely), and neglecting the torque generated by the support bearings of the primary shaft (see Figure 7), the equations of motion of the main pulley 1 and the counterweight according to the free body diagram in Figure 8 are defined by Equations (15) and (16), respectively.
J 1 + c 2 J 2 δ ¨ 1 = F c 1 r F c 01 r c T g + 3 T b
m 0 δ ¨ 1 r = F c 01 m 0 g
The expression in brackets on the left side of Equation (15) is the equivalent inertia referred to the primary shaft (low-speed shaft). J 1 corresponds to the sum of the inertia of the main pulley ( J p ), the inertia of the primary shaft ( J s 1 ) and the inertia of the gearbox ( J g b ). J 2 corresponds to the sum of the inertia of the generator ( J g ), the inertia of the secondary shaft ( J s 2 ) and the inertia of the flywheel ( J f ). δ ¨ 1 is the angular acceleration of the main pulley 1, r is the radio of the main pulley 1, c is the gear ratio (high angular velocity/low angular velocity), T b is the torque of each secondary shaft bearing, T g is the torque of the generator and F c 01 is the cable tension between the main pulley 1 and the counterweight. If the point absorber and the electric generator are uncoupled, Equation (15) is rewritten as Equation (17).
J 1 δ ¨ 1 = F c 1 r F c 01 r
Using the same analysis when the WEC captures energy only in the downward movement corresponding to the main pulley 2 (the main pulley 1 rotates freely), the equation of motion of the main pulley 2 for the coupling and uncoupling of the PTO and the counterweight are shown in Equations (18)–(20), respectively.
J 1 + c 2 J 2 δ ¨ 2 = F c 2 r F c 02 r c T g + 3 T b
J 1 δ ¨ 2 = F c 2 r F c 02 r
m 0 δ ¨ 2 r = F c 02 m 0 g
where δ ¨ 2 is the angular acceleration of the main pulley 2 and F c 02 is the cable tension between the main pulley 2 and the counterweight.
Two electric generators of low and high power are considered to analyze the performance of the WEC. The generators are permanent magnets [43], and the characteristics of the electric generators are shown in Table 3 while the power curves are shown in Figure 9. The generator torque without MPPT is calculated by Equation (21).
T g = P g   / ( η g ω g )
where η g ,   P g and ω g are the efficiency, power and speed of the generator.
The torque T b is approached using Petroff’s method of lubrication analysis [44], see Equation (22).
T b = 4 π 2 r s 3 l μ N c s
where r s is the shaft radius, l is the bearing length, μ is the dynamic viscosity of the lubricator, N is the shaft speed and c s is the radial clearance.
The coupling of the PTO system via the main pulley 1 is described by Equation (23). When the output shaft speed c δ ˙ 1 and the generator speed ω g satisfy this equation, the generator and the point absorber are coupled, then the generator shaft speed takes the value of c δ ˙ 1 . Similarly, for the main pulley 2, coupling occurs when the output shaft speed c δ ˙ 2 and ω g satisfy Equation (24).
  ω g < c δ ˙ 1  
ω g < c δ ˙ 2  
If the values of Equation (23) or Equation (24) are not met, the generator and point absorber are uncoupled. Then, the speed of generator is defined by Equation (25), if the generator is within its operating speed range (see Figure 7), otherwise it is given by Equation (26).
ω g t + Δ t = ω g t T g + 3 T b J 2 Δ t
ω g t + Δ t = ω g t 3 T b J 2 Δ t
where t and Δ t are the time and time step variables of the numerical simulation.
The power is obtained from the generator or the flywheel kinetic energy; the flywheel inertia considered in the present study is 40 kgm 2 . The generator runs only within its operating speed range; therefore, in Equations (15) and (18), T g depends on the generator shaft speed. When the generator exceeds its maximum speed, the torque T b of the bearings reduces the generator shaft speed, see Equation (26).
The mean power P m is determined by Equation (27).
P m = 1 T 0 T P g   d t
where T represents the last 10 periods of the numerical simulation and P g is the instantaneous power of the generator. The numerical simulation time considers 10,000 s; the first 1000 s are discarded to avoid transient instabilities.
The capture width ratio of the WEC is calculated by Equation (28).
C W R = P m P
where P is the incident wave power defined by Equation (29).
P = 1 8 ρ g H w 2 c 2 1 + 2 k d sin h 2 k d D
where the first factor is the mean wave-energy density per unit horizontal area, the second factor is the group velocity [45], c is the phase velocity, k is the wave number, d is the water depth and D is the diameter of the floating body.

3. Maximum Power Point Tracking

Most WECs based on rotating generators are considered to be permanent magnet synchronous generators (PMSGs); thus, in this work, a generic model of a PMSG using the P&O MPPT method is considered in order to compute the torque according to Equations (30) and (31), where P g is the output electrical power; K is the back electromotive force constant of the generator in vs/rad; N p is the pole pairs number; L s and R s are the stator inductance and resistance, respectively; η g is the efficiency of the generator and R g is the generator resistance, which is the parameter controlled by the MPPT algorithm.
P g = K 2 ω g 2 R g 2 N p ω g L s 2 + R s + R g 2
T g = P g η g ω g = K 2 ω g R g 2 N p ω g L s 2 + R s + R g 2
The MPPT P&O algorithm is performed to extract the maximum available power by periodically adjusting the R g value, thus modifying the generator torque under a gradient-ascent approach as detailed in [33]. The control of R g is realized by means of a suitable electronic converter, in this way the electric side of the WEC acts like a variable resistor. A flowchart of the P&O MPPT method considered in this work is presented in Figure 10. Variations in R g are discrete with the R s t e p value, after each variation in the mean power P g is computed and compared to the previous value to set the next value of R g . Since the tracking of the maximum power is limited by the generator-rated value, the maximum speed and the extreme allowable value for RG (Rmin and Rmax), the algorithm considers these constraints.

4. Simulation Results

The simulation results are presented in two sections. First, an analysis of the influence of different parameters, such as the gear ratio, wave period and height are presented; this is conducted based on two generators (G1 and G2), for which the technical details are listed in Table 3. Second, the simulation results of the time series of the main variables are presented for certain operational parameters with generator G1.

4.1. WEC Performance: Influence of Key Parameters

Under defined H w and T p conditions, operating with a higher gear ratio (GR) implies a higher generator speed. Since the rotor speed is variable, there are time intervals of maximum speed. As the GR is increased, these intervals of maximum speed can mean operating above the generator speed limit. This operation is undesirable; thus, in these conditions, the system disconnects the generator output, resulting in zero generation. There is a critical value of GR above which intervals of zero generation begin to occur. As the GR increases over this critical value, the intervals of zero generation become longer, thus reducing the average generated power until it reaches zero. This is noticeable in Figure 11, Figure 12, Figure 13 and Figure 14, which present the mean power vs. the GR for generators G1 and G2 under different conditions of wave period and height. It is also noticeable that a higher incident power (i.e., higher H w and lower T p ) leads to a lower critical value of GR. This is predictable, since higher power is related to higher generator speed; thus, a lower GR is expected.
When operating with MPPT, the value of Rg is adjusted to increase both the speed and the torque, aiming to extract the maximum available power. Thus, the critical value of GR is expected to be less than or equal to operation without MPPT. Moreover, for a GR lower than the critical value, operation with MPPT should produce a higher mean power than without MPPT. For GR values above the critical value, MPPT operation is expected to produce the same power (or even a bit lower) as without MPPT. Since for both operation conditions (with or without MPPT) the incident power is the same, the CWR for MPPT operation is also expected to be higher. On the other hand, for a GR above the critical value, the MPPT method is disabled, and the converter operates the same as without MPPT.
For the G2 generator without MPPT, the power takes the value of zero in some cases. This behavior occurs because the operating speed of the generator in a steady state exceeds the nominal speed of the generator; therefore, there is no power generation. This behavior is removed when the WEC operates with MPPT.
From Figure 11 and Figure 13 it is noticeable that, for GR values lower than the critical value, the proposed WECs produce more power when operating under the MPPT algorithm. Above the critical GR, the power for operation with MPPT is almost equal to that without MPPT. These are the expected results. For the best operating conditions of the two generators, the maximum achieved value in terms of mean power is about 50% of the rated power according to the datasheet: 44 kW for G2 and 12 kW for G1, both for Hw = 1.5 and Tp = 10 s. Moreover, the CWR is also higher for operation with MPPT, as can be seen in Figure 12 and Figure 14, achieving values of up to 14% for G1 and 53% for G2 when operating with Tp = 12 s and Hw = 1 m. These results may lead one to think that G2 is the most suitable generator for the WEC. However, the results obtained with generator G1 are quite promising, particularly considering that operating with smaller generators (up to 25 kW) represents a significant opportunity as this implies a substantial reduction in costs. Furthermore, the results confirm that the inclusion of the MPPT method allows the WECs to achieve the same mean power with a lower GR than without MPPT. This means lower losses and higher efficiency.

4.2. Time Series for Generator 1

The conditions selected for the time series analysis are: Hw = 1 m, Tp = 12 s and GR = 20. The main waveforms are presented in Figure 15 for steady-state operation (last 60 s of simulation). Results are presented for both operation without MPPT (left) and with MPPT (right). The selected variables are: displacement and velocity of the point absorber, wave excitation force (upper), instantaneous torque and power (middle) and generator speed (lower).
For operation without MPPT, the generator is loaded with a constant impedance, which produces a torque proportional to the speed; in time, the power is proportional to the square of the speed. On the other hand, when operating under the MPPT method, the torque is always higher than without MPPT; the algorithm increases the torque to extract more power. The limit is the rated torque (611 Nm). For both conditions it is noticeable that a minimum generator speed is achieved for maximum displacement, which also corresponds to the minimum torque. However, with MPPT, the torque around the minimum displacement is at its maximum; in this way, operating with MPPT produces more power.

MPPT Optimization of Rg

As detailed, the MPPT algorithm aims to maximize the power generation by means of adjusting the load resistance seen by the generator. Figure 16 shows the time response of the MPPT algorithm. For the first 300 periods an increase in Rg leads to an increase in the mean power. When Rg reaches the value of 0.32 Ω, an increase in Rg results in a decrease rather than an increase in power, thus the MPPT has been achieved. The MPPT algorithm keeps the Rg value oscillating around 0.32 Ω. Therefore, the mean power also oscillates around its maximum value.

5. Conclusions

The influence of the perturb and observe MPPT method applied to a new wave energy converter based on a point absorber, a hinged arm and a direct mechanical drive PTO system is studied herein. The PTO system is composed of a pulley system, a counterweight, a gearbox, a flywheel and an electric generator. The analysis of the WEC performance in regular waves considers two electric generators with a rated power of 22.7 kW and 95.5 kW, respectively. In general, the electrical power generation of the WEC increases with MPPT for lower GR values until a critical GR value is reached, after which the power decreases. The contribution of MPPT is higher for the low-power generator. The maximum mean power achieved is about 50% of the rated power for both generators. The high CWR values of the WEC occur at low values of period and wave height; the maximum value obtained is 53%, corresponding to the high-power generator G2. However, the low-power generator G1 shows a better response with the inclusion of MPPT; that is, having the same mean power with less GR than without MPPT, and smaller generators mean a substantial reduction in costs. In future work, the power performance of the WEC and the contribution of MPPT in irregular waves will be analyzed.

Author Contributions

Methodology, J.C.U.P. and G.O.G.A.; Software, C.L.M.R. and G.O.G.A.; Formal analysis, J.C.U.P.; Writing – original draft, G.O.G.A.; Funding acquisition, G.O.G.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Universidad Tecnológica del Perú (resolución rectoral no. 0047-2022-R-UTP).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article.

Conflicts of Interest

All authors declare that they have no conflict of interest.

References

  1. Sasmaz, M.U.; Sakar, E.; Yayla, Y.E.; Akkucuk, U. The Relationship between Renewable Energy and Human Development in OECD Countries: A Panel Data Analysis. Sustainability 2020, 12, 7450. [Google Scholar] [CrossRef]
  2. Adekoya, O.B.; Olabode, J.K.; Rafi, S.K. Renewable energy consumption, carbon emissions and human development: Empirical comparison of the trajectories of world regions. Renew. Energy 2021, 179, 1836–1848. [Google Scholar] [CrossRef]
  3. Sang, Y.; Karayaka, H.B.; Yan, Y.; Yilmaz, N.; Souders, D. 1.18 Ocean (Marine) Energy. In Comprehensive Energy Systems; Elsevier: Amsterdam, The Netherlands, 2018; Volume 1, pp. 733–769. [Google Scholar] [CrossRef]
  4. Mei, C.C.; Stiassnie, M.A.; Yue, D.K.-P. Theory and Applications of Ocean Surface Waves (In 2 Volumes), 3rd ed.; World Scientific: Singapore, 2018; pp. 1–1171. [Google Scholar] [CrossRef]
  5. Rasool, S.; Muttaqi, K.M.; Sutanto, D.; Iqbal, S. Modeling ocean waves and investigation of oceanic wave spectra for wave-to-wire system. J. Eng. Res. 2022, 10, 1–17. [Google Scholar] [CrossRef]
  6. Abdelwahed, H.G.; Abdelrahman, M.A.E.; Alsarhana, A.F.; Mohamed, K. Investigation of the Ripa Model via NHRS Scheme with Its Wide-Ranging Applications. Fractal Fract. 2022, 6, 745. [Google Scholar] [CrossRef]
  7. González, A.T.; Dunning, P.; Howard, I.; McKee, K.; Wiercigroch, M. Is wave energy untapped potential? Int. J. Mech. Sci. 2021, 205, 106544. [Google Scholar] [CrossRef]
  8. Veerabhadrappa, K.; Suhas, B.; Mangrulkar, C.K.; Kumar, R.S.; Mudakappanavar, V.; Narahari;Seetharamu, K. Power Generation Using Ocean Waves: A Review. Glob. Transit. Proc. 2022, 3, 359–370. [Google Scholar] [CrossRef]
  9. Ahamed, R.; McKee, K.; Howard, I. Advancements of wave energy converters based on power take off (PTO) systems: A review. Ocean Eng. 2020, 204, 107248. [Google Scholar] [CrossRef]
  10. Beatty, S.J.; Bocking, B.; Bubbar, K.; Buckham, B.J.; Wild, P. Experimental and numerical comparisons of self-reacting point absorber wave energy converters in irregular waves. Ocean Eng. 2019, 173, 716–731. [Google Scholar] [CrossRef]
  11. Blanco, M.; Torres, J.; Santos-Herrán, M.; García-Tabarés, L.; Navarro, G.; Nájera, J.; Ramírez, D.; Lafoz, M. Recent Advances in Direct-Drive Power Take-Off (DDPTO) Systems for Wave Energy Converters Based on Switched Reluctance Machines (SRM); Springer International Publishing: Berlin/Heidelberg, Germany, 2022; pp. 487–532. [Google Scholar] [CrossRef]
  12. Guo, B.; Wang, T.; Jin, S.; Duan, S.; Yang, K.; Zhao, Y. A Review of Point Absorber Wave Energy Converters. J. Mar. Sci. Eng. 2022, 10, 1534. [Google Scholar] [CrossRef]
  13. López, I.; Andreu, J.; Ceballos, S.; de Alegría, I.M.; Kortabarria, I. Review of wave energy technologies and the necessary power-equipment. Renew. Sustain. Energy Rev. 2013, 27, 413–434. [Google Scholar] [CrossRef]
  14. Val, D.V. Reliability of Marine Energy Converters. Energies 2023, 16, 3387. [Google Scholar] [CrossRef]
  15. Albert, A.; Berselli, G.; Bruzzone, L.; Fanghella, P. Mechanical design and simulation of an onshore four-bar wave energy converter. Renew. Energy 2017, 114, 766–774. [Google Scholar] [CrossRef]
  16. Jusoh, M.A.; Ibrahim, M.Z.; Daud, M.Z.; Albani, A.; Yusop, Z.M. Hydraulic Power Take-Off Concepts for Wave Energy Conversion System: A Review. Energies 2019, 12, 4510. [Google Scholar] [CrossRef] [Green Version]
  17. Juan, N.P.; Valdecantos, V.N.; Esteban, M.D.; Gutiérrez, J.S.L. Review of the Influence of Oceanographic and Geometric Parameters on Oscillating Water Columns. J. Mar. Sci. Eng. 2022, 10, 226. [Google Scholar] [CrossRef]
  18. Rosati, M.; Henriques, J.; Ringwood, J. Oscillating-water-column wave energy converters: A critical review of numerical modelling and control. Energy Convers. Manag. X 2022, 16, 100322. [Google Scholar] [CrossRef]
  19. Ahamed, R.; McKee, K.; Howard, I. A Review of the Linear Generator Type of Wave Energy Converters’ Power Take-Off Systems. Sustainability 2022, 14, 9936. [Google Scholar] [CrossRef]
  20. Avalos, G.O.G.; Shadman, M.; Estefen, S.F. Application of the Latching Control System on the Power Performance of a Wave Energy Converter Characterized by Gearbox, Flywheel, and Electrical Generator. J. Mar. Sci. Appl. 2021, 20, 767–786. [Google Scholar] [CrossRef]
  21. Binh, P.C.; Tri, N.M.; Dung, D.T.; Ahn, K.K.; Kim, S.; Koo, W. Analysis, design and experiment investigation of a novel wave energy converter. IET Gener. Transm. Distrib. 2016, 10, 460–469. [Google Scholar] [CrossRef]
  22. Kofoed, J.P.; Tetu, A.; Ferri, F.; Margheritini, L.; Sonalier, N.; Larsen, T. Real Sea Testing of a Small Scale Weptos WEC Prototype. In Proceedings of the International Conference on Offshore Mechanics and Arctic Engineering, Madrid, Spain, 17–22 June 2018; Volume 10, pp. 1–9. [Google Scholar] [CrossRef]
  23. Sanada, M.; Inoue, Y.; Morimoto, S. Generator design and characteristics in direct-link wave power generating system considering appearance probability of waves. In Proceedings of the 2012 International Conference on Renewable Energy Research and Applications (ICRERA), Nagasaki, Japan, 11–14 November 2012; pp. 1–6. [Google Scholar] [CrossRef]
  24. Shadman, M.; Guarniz Avalos, G.O.; Estefen, S.F. On the power performance of a wave energy converter with a direct mechanical drive power take-off system controlled by latching. Renew. Energy 2021, 169, 157–177. [Google Scholar] [CrossRef]
  25. Sjolte, J.; Sandvik, C.M.; Tedeschi, E.; Molinas, M. Exploring the Potential for Increased Production from the Wave Energy Converter Lifesaver by Reactive Control. Energies 2013, 6, 3706–3733. [Google Scholar] [CrossRef] [Green Version]
  26. Ringwood, J.V. Wave energy control: Status and perspectives 2020. IFAC-PapersOnLine 2020, 53, 12271–12282. [Google Scholar] [CrossRef]
  27. Ringwood, J.V.; Merigaud, A.; Faedo, N.; Fusco, F. An Analytical and Numerical Sensitivity and Robustness Analysis of Wave Energy Control Systems. IEEE Trans. Control Syst. Technol. 2020, 28, 1337–1348. [Google Scholar] [CrossRef] [Green Version]
  28. Ding, B.; Cazzolato, B.S.; Arjomandi, M.; Hardy, P.; Mills, B. Sea-state based maximum power point tracking damping control of a fully submerged oscillating buoy. Ocean Eng. 2016, 126, 299–312. [Google Scholar] [CrossRef]
  29. Pena, J.C.U.; de Brito, M.A.G.; Melo, G.D.A.E.; Canesin, C.A. A comparative study of MPPT strategies and a novel singlephase integrated buck-boost inverter for small wind energy conversion systems. In Proceedings of the COBEP 2011—11th Brazilian Power Electronics Conference, Natal, Brazil, 11–15 September 2011; pp. 458–465. [Google Scholar] [CrossRef]
  30. De Brito, M.A.G.; Galotto, L.; Sampaio, L.P.; Melo, G.D.A.E.; Canesin, C.A. Evaluation of the Main MPPT Techniques for Photovoltaic Applications. IEEE Trans. Ind. Electron. 2013, 60, 1156–1167. [Google Scholar] [CrossRef]
  31. Xiao, X.; Huang, X.; Kang, Q. A Hill-Climbing-Method-Based Maximum-Power-Point-Tracking Strategy for Direct-Drive Wave Energy Converters. IEEE Trans. Ind. Electron. 2016, 63, 257–267. [Google Scholar] [CrossRef]
  32. Amon, E.A.; Brekken, T.K.A.; Schacher, A.A. Maximum Power Point Tracking for Ocean Wave Energy Conversion. IEEE Trans. Ind. Appl. 2012, 48, 1079–1086. [Google Scholar] [CrossRef]
  33. Lettenmaier, T.; Von Jouanne, A.; Brekken, T. A New Maximum Power Point Tracking Algorithm for Ocean Wave Energy Converters. Int. J. Mar. Energy 2017, 17, 40–55. [Google Scholar] [CrossRef] [Green Version]
  34. Hardy, P.; Cazzolato, B.; Ding, B.; Prime, Z. A maximum capture width tracking controller for ocean wave energy converters in irregular waves. Ocean Eng. 2016, 121, 516–529. [Google Scholar] [CrossRef]
  35. Villalba, I.; Blanco, M.; Pérez-Díaz, J.I.; Fernández, D.; Díaz, F.; Lafoz, M. Wave farms grid code compliance in isolated small power systems. IET Renew. Power Gener. 2019, 13, 171–179. [Google Scholar] [CrossRef]
  36. Aderinto, T.; Li, H. Ocean Wave Energy Converters: Status and Challenges. Energies 2018, 11, 1250. [Google Scholar] [CrossRef] [Green Version]
  37. Kim, S.-J.; Koo, W.; Shin, M.-J. Numerical and experimental study on a hemispheric point-absorber-type wave energy converter with a hydraulic power take-off system. Renew. Energy 2019, 135, 1260–1269. [Google Scholar] [CrossRef]
  38. Yang, S.; Chen, H.; Ji, Z.; Li, H.; Xiang, X. Modelling and analysis of inertia self-tuning phase control strategy for a floating multi-body wave energy converter. IET Renew. Power Gener. 2021, 15, 3126–3137. [Google Scholar] [CrossRef]
  39. Cummins, W.E. The Impulse Response Function and Ship Motions. Open J. Fluid Dyn. 1962, 57, 101–109. [Google Scholar]
  40. Ogilvie, T.F.; Model, D.T.; Washington, B.; Scheepsbouwkwide, L.Y. Recent Progress toward the Understanding and Prediction of Ship Motions. In Proceedings of the 5th Symposium on Naval Hydrodynamics, Bergen, Norway, 10–12 September 1964; David, W., Ed.; Taylor Model Basin: Washington, DC, USA, 1964; pp. 3–80. [Google Scholar]
  41. Morison, J.R.; O’Brien, M.P.; Johnson, J.W.; Schaaf, S.A. The forces exerted by surface waves on piles. J. Pet. Trans. 1950, 2, 149–154. [Google Scholar] [CrossRef]
  42. Sumer, B.M.; Fredsøe, J. Hydrodynamics Around Cylindrical Structures; World Scientific: Singapore, 2006. [Google Scholar] [CrossRef]
  43. Frameless Motor for Direct Drive | Alxion | Torque Motors, Permanent Magnet Generators, Resolvers, Servo Drives. Available online: http://www.alxion.com/products/stk-alternators/ (accessed on 14 April 2023).
  44. Budynas, R.; Nisbett, J. Shigley’s Mechanical Engineering Design; McGraw Hill: New York, NY, USA, 2011. [Google Scholar]
  45. Journée, J.; Massie, W. Offshore Hydromechanics; CITG Section Hydraulic Engineering: Delft, The Netherlands, 2000. [Google Scholar]
Figure 1. Components of the wave energy converter.
Figure 1. Components of the wave energy converter.
Sustainability 15 10447 g001
Figure 2. Direct mechanical drive PTO. (a) Back view. (b) Front view.
Figure 2. Direct mechanical drive PTO. (a) Back view. (b) Front view.
Sustainability 15 10447 g002
Figure 3. Schematic model of the wave energy converter. (a) Main dimensions of the WEC. (b) Wave propagation direction near the WEC.
Figure 3. Schematic model of the wave energy converter. (a) Main dimensions of the WEC. (b) Wave propagation direction near the WEC.
Sustainability 15 10447 g003
Figure 4. Retardation function K r .
Figure 4. Retardation function K r .
Sustainability 15 10447 g004
Figure 5. Added mass and radiation damping coefficients obtained via ANSYS/AQWA.
Figure 5. Added mass and radiation damping coefficients obtained via ANSYS/AQWA.
Sustainability 15 10447 g005
Figure 6. Wave force amplitude and wave force phase shift obtained via ANSYS/AQWA.
Figure 6. Wave force amplitude and wave force phase shift obtained via ANSYS/AQWA.
Sustainability 15 10447 g006
Figure 7. PTO system.
Figure 7. PTO system.
Sustainability 15 10447 g007
Figure 8. Diagram of the free body of the main pulley 1 and the counterweight.
Figure 8. Diagram of the free body of the main pulley 1 and the counterweight.
Sustainability 15 10447 g008
Figure 9. Power curves of the generators.
Figure 9. Power curves of the generators.
Sustainability 15 10447 g009
Figure 10. Flowchart of the considered perturb and observe MPPT method.
Figure 10. Flowchart of the considered perturb and observe MPPT method.
Sustainability 15 10447 g010
Figure 11. Mean power vs. GR at different wave heights and at wave period Tp =12 s: (a) generator G1, (b) generator G2.
Figure 11. Mean power vs. GR at different wave heights and at wave period Tp =12 s: (a) generator G1, (b) generator G2.
Sustainability 15 10447 g011
Figure 12. CWR vs. GR at different wave heights and at wave period Tp = 12 s: (a) generator G1, (b) generator G2.
Figure 12. CWR vs. GR at different wave heights and at wave period Tp = 12 s: (a) generator G1, (b) generator G2.
Sustainability 15 10447 g012
Figure 13. Mean power vs. GR at different wave periods and at wave height Hw = 1.5 m: (a) generator G1, (b) generator G2.
Figure 13. Mean power vs. GR at different wave periods and at wave height Hw = 1.5 m: (a) generator G1, (b) generator G2.
Sustainability 15 10447 g013
Figure 14. CWR vs. GR at different wave periods and at wave height Hw = 1.5 m: (a) generator G1, (b) generator G2.
Figure 14. CWR vs. GR at different wave periods and at wave height Hw = 1.5 m: (a) generator G1, (b) generator G2.
Sustainability 15 10447 g014
Figure 15. Main waveforms for G1 operation without MPPT.
Figure 15. Main waveforms for G1 operation without MPPT.
Sustainability 15 10447 g015
Figure 16. MPPT optimization of the generator load.
Figure 16. MPPT optimization of the generator load.
Sustainability 15 10447 g016
Table 1. Characteristics of the buoy.
Table 1. Characteristics of the buoy.
ParameterValue
Mass (m)40.25 t
Diameter (D)5 m
Draft (T)2 m
Natural Period (Tn)3.64 s
Table 2. Main parameters of the WEC.
Table 2. Main parameters of the WEC.
ParameterValue
h 1 3.83 m
h 2 3.83 m
h 3 3.83 m
R 10 m
m 0 0.25 t
Table 3. Parameters of the electric generators.
Table 3. Parameters of the electric generators.
ParametersG1G2
Rated power (W)22,67895,484
Rated speed (rpm)400350
Rotor inertia (kgm2)1.2707.620
Rated torque (Nm)6112771
Efficiency ,   η g (%)8994
Phase resistance (Ohm) at 20 °C0.20.02
Phase inductance (mH)0.940.27
Voltage at no load (back emf) at 20 °C299336
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ugaz Peña, J.C.; Medina Rodríguez, C.L.; Guarniz Avalos, G.O. Study of a New Wave Energy Converter with Perturb and Observe Maximum Power Point Tracking Method. Sustainability 2023, 15, 10447. https://doi.org/10.3390/su151310447

AMA Style

Ugaz Peña JC, Medina Rodríguez CL, Guarniz Avalos GO. Study of a New Wave Energy Converter with Perturb and Observe Maximum Power Point Tracking Method. Sustainability. 2023; 15(13):10447. https://doi.org/10.3390/su151310447

Chicago/Turabian Style

Ugaz Peña, José Carlos, Christian Luis Medina Rodríguez, and Gustavo O. Guarniz Avalos. 2023. "Study of a New Wave Energy Converter with Perturb and Observe Maximum Power Point Tracking Method" Sustainability 15, no. 13: 10447. https://doi.org/10.3390/su151310447

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop