Next Article in Journal
An Increase in the Energy Efficiency of Abrasive Jet Equipment Based on the Rational Choice of Nozzle Geometry
Previous Article in Journal
Part II: State-of-the-Art Technologies of Solar-Powered DC Microgrid with Hybrid Energy Storage Systems: Converter Topologies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Multi-Timescale Lookup Table Based Maximum Power Point Tracking of an Inverse-Pendulum Wave Energy Converter: Power Assessments and Sensitivity Study

1
PowerChina HuaDong Engineering Corporation Limited, Hangzhou 311122, China
2
Hangzhou Huachen Electric Power Control Corporation Limited, Hangzhou 311122, China
3
Key Laboratory of Transients in Hydraulic Machinery, Ministry of Education, Wuhan University, Wuhan 430072, China
4
China Renewable Energy Engineering Institute, Beijing 100120, China
*
Author to whom correspondence should be addressed.
Energies 2023, 16(17), 6195; https://doi.org/10.3390/en16176195
Submission received: 13 June 2023 / Revised: 16 August 2023 / Accepted: 24 August 2023 / Published: 25 August 2023
(This article belongs to the Topic Energy from Sea Waves)

Abstract

:
A novel, inverse-pendulum wave energy converter (NIPWEC) is a device that can achieve natural period control via a mass-position-adjusting mechanism and a moveable internal mass. Although the energy capture capacity of a NIPWEC has already been proven, it is still meaningful to research how to effectively control the NIPWEC in real time for maximum wave energy absorption in irregular waves. This paper proposes a multi-timescale lookup table based maximum power point tracking (MLTB MPPT) strategy for the NIPWEC. The MLTB MPPT strategy was implemented to achieve a theoretical “optimal phase” and “optimal amplitude” by adjusting both the position of the internal mass and linear power take-off (PTO) damping. It consists of two core parts, i.e., internal mass position adjustment based on a 1D resonance position table and PTO damping tuning based on a 2D optimal PTO damping table. Furthermore, power assessments and sensitivity study were conducted for eight irregular-wave sea states with diverse wave spectra. The results show that energy period resonance and the lookup table based PTO damping tuning have the highest possibility of obtaining the maximum mean time-averaged absorbed power. Additionally, both of them are robust to parameter variations. In the next step, the tracking performance of the MLTB MPPT strategy in terms of changing sea states will be studied in-depth.

1. Introduction

Wave energy is a carbon-free renewable resource that possesses the advantages of a high energy density and has outstanding availability and better predictability than wind or solar energy [1]. Besides, the global wave power is about 29,500 TWh/year, according to the estimation from Ocean Energy Systems [2]. If wave energy is fully utilized, the demand for global electricity consumption can be sufficiently satisfied. The data given by the International Energy Agency show that the global electricity consumption in 2019 reached 22,848 TWh [3]. Therefore, wave energy has enormous development potential.
However, there still exist several technical and non-technical challenges for wave energy utilization (WEU) [1,4]. The wave-by-wave, hour-by-hour, and site-by-site variations of the level of wave power represent one of the main obstacles restricting technological convergence and maturity. Therefore, many researchers focus on how to improve the performance of wave energy converters (WECs), especially wave conversion efficiency, under complex sea states [5].
WEC operation control has always been deemed a research hot spot since a successful control method can significantly enhance the wave conversion efficiency of an existing WEC with low additional investment [1]. To date, plenty of control methods have been proposed based on different actuators and control algorithms, wherein the phase control methods (PCMs), natural period control methods (NPCMs), and maximum power point tracking (MPPT) are related to the research content of this paper. Hence, the aforementioned three control methods are introduced respectively.
PCMs, e.g., latching control (LC) and declutching control (DeC), were first presented between the end of the 1970s and the beginning of this century. PCMs are usually based on PTO force tuning, which can compulsively synchronize WEC movement with the variations in wave excitation force to achieve an “optimal phase”. The theoretical basis of PCMs was summarized by Falnes [6] as a universal theory on ocean waves and oscillating systems. Although the theory is linear, it qualitatively points out the core idea of most control methods, not only PCMs. Herein, the core idea means that the maximum absorbed power will appear when both “optimal phase” and “optimal amplitude” are achieved. “Optimal phase” denotes a resonant state where the WEC velocity and wave excitation force are in phase. Meanwhile, “optimal amplitude” refers to damping matching; e.g., the linear damping of a power take-off (PTO) should be equal to radiation damping in regular waves. Nowadays, LC or DeC, with or without the knowledge of future excitation force, still attracts many researchers [7,8].
NPCMs are another solution for achieving “optimal phase”. They alter the natural period/frequency of a WEC to be identical to the period/frequency of an incident wave. Three ways can be used to realize natural period control. The first is inertia adjustment [9,10], the second can be called stiffness adjustment (SA) [11,12,13,14,15,16], and the third is known as surface control [17]. SA is the most common NPCM. It refers to adjusting the inherent stiffness term in a WEC. Marei et al. [11] changed compressed gas stiffness to achieve the resonance of an Archimedes wave swing. Additionally, Chen et al. [12] implemented the SA of a built-in horizontal axis pendulum to achieve resonance. Moreover, Refs. [13,14,15,16] pointed out that the hydrostatic restoring stiffness of an oscillating wave surge converter (OWSC), e.g., a solo duck WEC or an inverse-pendulum-type WEC, can be effectively regulated for resonance by using a moveable mass.
When compared to PCMs, NPCMs are mainly based on slow tuning processes. It means that energy capture performance can be effectively improved as long as a WEC adjusts its natural period to the wave characteristic period of an irregular wave within a large duration. In addition, NPCMs do not need to tune PTO force at a wide range, which reduces the control difficulty and improves control reliability. However, NPCMs are only able to realize “optimal phase”. If “optimal amplitude” is required, PTO force tuning should be inevitably adopted. Therefore, how to comprehensively design and utilize two control mechanisms to achieve both “optimal phase” and “optimal amplitude” could become an issue worthy of in-depth research when applying NPCMs.
MPPT is a real-time search technology for the maximum power points of clean energy power, such as wind power, photovoltaics, thermoelectricity [18], etc. In 2009, Amon et al. [19] first introduced the concept of MPPT to WEU, conducting a perturbation and observation (P&O) algorithm on the duty cycle of a Buck circuit as well as the simulated resistance of a three-phase pulse width modulation (PWM) rectifier. Nowadays, the MPPT algorithm family includes a fixed-step P&O [19,20,21,22], a variable-step P&O [23,24], a segmental fixed-step P&O [25], a lookup-table-based (LTB) MPPT [26,27], diverse heuristic algorithms [28,29], etc. The fixed-step P&O is the most common algorithm and often serves as a reference for other algorithms. Besides, LTB MPPT is an ordinary but practical algorithm that has been applied to an oscillating water column-type WEC and a floating-pendulum WEC [26,27]. Yue et al. [27] proved that a two-dimensional (2D) LTB MPPT algorithm, which contains an optimal duty cycle table derived from regular wave simulations, is a simple, reliable, and excellent solution for WEC output power enhancement in irregular waves. Refs. [30,31] constructed an emulator for a point absorber WEC (PAWEC) and reviewed the MPPT techniques or hybrid strategies of PAWECs.
This paper focuses on a novel inverse-pendulum wave energy converter (NIPWEC), which can conduct SA via a mass-position-adjusting mechanism (MPAM) and a moveable internal mass. The structural details of a NIPWEC can be seen in Section 2. Similar structures can be seen in Refs. [14,15,16,32]. Although the energy capture capacity of a NIPWEC has already been proven via simulations in a numerical wave tank, the prototype tests in a physical wave flume and the annual average power assessments of a 1:16 full-scale prototype [14,15] for how to effectively control the NIPWEC in real-time is still a meaningful research topic. The inefficient implementation of a control method can lead to ordinary or even weakened energy capture performance. Hence, this paper researches an implementation strategy for SA and PTO force tuning, aiming to find a suitable control algorithm for NIPWECs for real irregular waves.
Dong et al. [32] proposed a frequency and amplitude control (FAC) strategy to realize the MPPT of a NIPWEC. Herein, frequency control means adjusting the internal mass position according to the peak frequency of a wave spectrum. Meanwhile, amplitude control refers to tuning the PTO force to find the inherent impendence of the NIPWEC, which is approximately calculated by a weighted average method. The premise of FAC is the fast Fourier transformation (FFT) and frequency-domain analysis of the wave excitation moment within the past hundreds of seconds at least. This may lead to the following two issues.
(1) The nonlinear factors, e.g., an endstop moment for the swing amplitude limitation and a sine function hydrostatic restoring moment, are ignored by frequency domain analysis. This may result in the deterioration of MPPT performance when the ideal NIPWEC inherent impendence is applied to PTO force tuning.
(2) The real-time tracking performance may be weakened since both frequency control and amplitude control are implemented every few hundred seconds.
In contrast to Ref. [32], this paper presents a simple and reliable control algorithm, i.e., the muti-timescale lookup table based maximum power point tracking (MLTB MPPT), for NIPWECs. This algorithm is constructed and conducted based on a similar idea that can be found in Ref. [27]. MLTB MPPT consists of LTB internal-mass position adjustment and LTB PTO damping tuning, which are obtained via time domain analysis with consideration given to the above nonlinear factors. Additionally, MLTB MPPT is capable of tuning the PTO force wave by wave, which can reinforce its real-time tracking performance. Detailed descriptions of MLTB MPPT are illustrated in Section 2. Furthermore, the power assessments and sensitivity study of MLTB MPPT are comprehensively investigated.
The rest of this paper Is organized as follows. First, the structure and implementation procedure of MLTB MPPT is elaborated in Section 2. Next, a NIPWEC dynamic model is described in Section 3. Then, the parameter settings for the time domain simulations are displayed in Section 4. Afterwards, the maximum-power-point (MPP) analysis in regular waves, as well as the power assessments and sensitivity study for MLTB MPPT in irregular waves, is discussed in Section 5. Last, the conclusions, along with future work discussion, are revealed in the Section 6.

2. Methodology: MPPT

2.1. Three MPPT Implementation Methods for a NIPWEC

We focus on a NIPWEC with a mechanical power take-off. As shown in Figure 1, the NIPWEC is constrained on a fixed supporting shaft and swings along it driven by ocean waves. Next, a speed-increase mechanism accelerates the inverse-pendulum swing and converts it into the fast bidirectional rotation of a permanent magnet synchronous generator (PMSG). Then, the PMSG generates continuous but unstable three-phase alternating-current (AC) electricity. Afterwards, a three-phase PWM rectifier transforms the AC electricity into direct current (DC) and regulates the PTO force via the vector control of a PMSG. Last, the DC bus smooths DC voltage and balances power generation and consumption via an energy storage unit and diverse electric loads.
The NIPWEC can automatically tune its natural period within a wide range by adjusting the vertical position of an internal mass, according to the changing characteristic period of irregular waves. Herein, the mass position-adjusting mechanism (MPAM) with a screw and a motor is adopted. The motor drives the screw to rotate forward/reverse in order to realize the up/down movement of the internal mass. Therefore, the NIPWEC is able to generate electricity at the resonant state for a long time, which leads to the effective improvement of the energy capture/generating capacity.
Falnes [6] pointed out that the single DoF motion in ocean waves can be assumed to be a forced linear oscillation system. Under this assumption, the maximum absorbed power can be obtained under the “optimal phase” condition, which means that the oscillation velocity and excitation force are in phase. However, Falnes’ theory relies on a frequency domain model and ignores the influence of nonlinear factors, e.g., an endstop moment for the swing amplitude limitation and a sine function hydrostatic restoring moment [33]. The two nonlinear factors are able to significantly affect the magnitude of the absorbed power under the “resonance” condition. This means that due to the existence of nonlinear factors, the optimal PTO damping R P T O max is not exactly equal to the radiation damping R 55 ( ω r ) at resonance wave frequency ω r . In addition, the actual ocean waves are irregular waves, for which the wave heights and wave periods change all the time. Therefore, it is meaningful to search for a simple and reliable MPPT algorithm based on a time domain model.
The basic goal of an MPPT algorithm is to achieve “optimal phase”, i.e., resonance, and “optimal amplitude”, i.e., optimal PTO damping, against the maximum absorbed power. For a NIPWEC, the above goal can be theoretically achieved via the following three implementation methods.
Method 1 (M1): adjust the internal mass position X to achieve “optimal phase” and tune PTO damping R P T O to achieve “optimal amplitude”. Herein, X = l M ( t ) l M 0 , where l M 0 means the lower bound for l M ( t ) , and l M ( t ) signifies the distance between the rotation center R and the adjustable internal mass center M. The PTO moment M P T O for M1 can be given by
M P T O = R P T O θ ˙
Method 2 (M2): adjust the PTO stiffness K P T O to achieve “optimal phase” and tune R P T O to achieve “optimal amplitude”. M P T O for M2 is
M P T O = R P T O θ ˙ K P T O θ
Method 3 (M3): M3 is similar to M2; the only difference from M2 is that the sine function hydrostatic restoring moment is considered when adjusting K P T O . M P T O for M3 is
M P T O = R P T O θ ˙ K P T O sin θ

2.2. MLTB MPPT

2.2.1. Overall Structure

Section 2.1 has already pointed out that a feasible MPPT implementation method is M1, i.e., simultaneously adjusting X and R P T O . The details of the real-time MPPT technology in irregular waves are discussed in this section. Herein, we propose a simple and reliable MLTB MPPT (see Figure 2), which is embedded in the wave power controller of a NIPWEC. MLTB MPPT contains two core parts, i.e., LTB internal-mass position adjustment and LTB PTO damping tuning.
In terms of LTB internal-mass position adjustment, the precise adjustment of X can be realized via the MPAM with a screw and a controllable motor, e.g., a servo or stepper motor. The controllable motor drives the screw to rotate and to move the internal mass to a reference position X * , which can be given by querying the resonance position table according to the estimated characteristic period of a real-time wave spectrum. The real-time wave spectrum can be obtained through the FFT for the in situ wave elevation signal (WES) within the past hundreds of seconds. Herein, WES can be dependably achieved via the in situ wave-measuring buoy. Additionally, WEC motion can be utilized to identify wave excitation force/moment [14,34,35]. Thereby, WES could be theoretically estimated via the identified wave excitation force/moment and the corresponding transfer function. However, few papers focus on the relevant issue.
For LTB PTO damping tuning, the accurate tuning of R P T O can be achieved by applying vector control to the PMSG using a three-phase PWM rectifier. The vector controller adopts a 0-d-axis-current control mode. Structurally, it only contains a current loop with the reference d-axis current i d * = 0 . Meanwhile, the reference q-axis current i q * is given by a function of R P T O max , which can be retrieved from the optimal PTO damping table according to the period and height of the last single wave or the statistical period and height of a previous multiple-wave group.
Herein, “multi-timescale” is reflected in two aspects. First, the required historical WES for the LTB internal-mass position adjustment and the LTB PTO damping tuning is different in terms of the signal length. The former needs a several-hundred-second WES to accurately calculate the characteristic period of a real-time wave spectrum, whereas the latter only needs the WES within the duration of a single wave or multiple waves. Second, there also exists a distinction between the implementation time intervals for the LTB internal-mass position adjustment and the LTB PTO damping tuning. Internal mass position can be adjusted at intervals of the kilo-second level since the obvious change of a real-time wave spectrum only occurs hourly [1,36]. Meanwhile, PTO damping should be tuned wave-by-wave or “waves-by-waves”, which means that the implementation time intervals of PTO damping tuning lie at the second or minute level.
Furthermore, “look-up table” refers to the two key tables, i.e., the resonance position table and the optimal PTO damping table, derived from the MPP analysis for regular waves. Both of them are essential for guiding the MPPT in terms of regular waves.
When compared to other MPPT algorithms, MLTB MPPT has the following two features.
(1) MLTB MPPT has high reliability. It can effectively avoid search blindness, as the two look-up tables offer a meaningful reference for improving search accuracy.
(2) The principle of MLTB MPPT is simple. Complex wave prediction is unnecessary. Moreover, the two look-up tables can be easily obtained through a batch of regular wave simulations without regard given to irregular waves.

2.2.2. LTB Internal-Mass Position Adjustment

The internal mass position should be adjusted adaptively, along with the changeable real-time wave spectrum, to make sure that the NIPWEC can generate electricity at the resonant state most of the time. The core of the internal-mass position adjustment locates and keeps the natural period of the NIPWEC consistent with the characteristic period T c . Therefore, the adjusting process can be conducted as follows:
S1. Preset adjusting cycle, Δ T m .
S2. Set sequence number, n t = 0 .
S3. Obtain the WES and save it.
S4. Determine whether running time ≥ n t Δ T m . If the inequality is satisfied, go to S5. Otherwise, keep running S3.
S5. Implement FFT and filtering on the historical WES within the duration of d T m after the instant of n t Δ T m to obtain the real-time wave spectrum.
S6. Calculate the T c of the real-time wave spectrum.
S7. Search the resonance position table and find out the resonance position X max corresponding to T c . If T c cannot be found in the resonance position table, an interpolation should be employed.
S8. Set X * = X max and input X * into the MPAM.
S9. Move the internal mass to the given position X * via an MPAM and keep the position.
S10. Determine whether the program needs to be stopped. If yes, terminate the program. Otherwise, n t = n t + 1 and return to S3.

2.2.3. LTB PTO Damping Tuning

If we assume that each single wave in an irregular wave sequence can be seen as a regular wave within the duration of a wave period, the corresponding R P T O max given by an optimal PTO damping table can be applied wave-by-wave, i.e., every other wave period, to approach the real-time “optimal amplitude” condition. Although a single wave is not completely sinusoidal, the aforementioned assumption is essential and suitable for practical engineering. Furthermore, a 2D LTB MPPT algorithm with the same assumption has already been proposed and verified by Ref. [27]. The only difference between the 2D LTB MPPT and the LTB PTO damping tuning here is that the control variable of the 2D LTB MPPT refers to the duty cycle of a buck-boost converter.
Similar to the 2D LTB MPPT, LTB PTO damping tuning can be implemented wave-by-wave or waves-by-waves, i.e., at intervals of multiple wave periods. The first implementation strategy is named single-wave-period tuning (SWPT), whereas the second can be called multiple-wave-period tuning (MWPT). If frequent adjustment is not wanted, MWPT could be an alternative with comparable performance. Moreover, since there is not much difference in the periods/heights of the two adjacent single waves as well as the statistical periods/heights of the two adjacent multiple-wave groups, the corresponding R P T O max of the last single wave or multiple-wave group is utilized for the present one here in order to avoid complex wave prediction.
The procedure of SWPT is described below.
S1. Preset the single wave order number to n = 1 .
S2. Obtain the WES and save it.
S3. Determine whether it reaches the n+1-th up-crossing zero point Z u p , n + 1 . If yes, n = n + 1 and go to S4. Otherwise, return to S2.
S4. Determine whether n > 0 . If yes, go to S5. Otherwise, return to S2.
S5. Calculate the height H n and period T n of the n-th single wave.
S6. Search the optimal PTO damping table and find out the R P T O max corresponding to ( H n , T n ) . If ( H n , T n ) cannot be found exactly in the optimal PTO damping table, an interpolation should be employed.
S7. Set i q * = f ( R P T O max ) and input it into the vector controller. Since the 0-d-axis-current control mode is adopted, f ( R P T O max ) can be expressed as
f ( R P T O max ) = R P T O max ω m 1.5 n p ϕ f k g 2
where ω m is the PMSG angular velocity, n p is the number of pole pairs, ϕ f is the rotor flux, and k g is the acceleration ratio of a speed increase mechanism.
S8. Tune R P T O to R P T O max via the PMSG vector control and keep R P T O .
S9. Determine whether the program needs to be stopped. If yes, terminate the program. Otherwise, return to S2.
In addition, the procedure of MWPT is as follows:
S1. Preset the single wave order number to n = 1 .
S2. Define the number of single waves k w ( k w > 1 ).
S3. Obtain the WES and save it.
S4. Determine whether it reaches the n+1-th up-crossing zero point Z u p , n + 1 . If yes, n = n + 1 and go to S5. Otherwise, return to S3.
S5. Determine whether n > 0 . If yes, go to S6. Otherwise, return to S3.
S6. Calculate the height H n and period T n of the n-th single wave.
S7. Determine whether n is the multiple of k w . If yes, go to S8. Otherwise, return to S3.
S8. Calculate the statistical height H k w and statistical period T k w of the multiple-wave group containing the last k w single waves. Herein, the mean wave height and period for the first one-third of the single waves of the multiple-wave group, which is rearranged in descending order of k w wave heights, is chosen as H k w and T k w .
S9. Search the optimal PTO damping table and find the R P T O max corresponding to ( H k w , T k w ) . If ( H k w , T k w ) cannot be exactly found in the optimal PTO damping table, an interpolation should be employed.
S10. Set i q * = f ( R P T O max ) and input it into the vector controller.
S11. Tune R P T O to R P T O max via the PMSG vector control and keep R P T O .
S12. Determine whether the program needs to be stopped. If yes, terminate the program. Otherwise, return to S3.

2.2.4. The Two Other PTO Damping Tuning Algorithms

The two other PTO damping tuning algorithms were also researched for a comparison with the LTB PTO damping tuning. The first one is a fixed-step P&O algorithm, and the second is a variable-step P&O algorithm, i.e., the admittance differentiation method (ADM).
The structure of a fixed-step P&O algorithm is shown in Figure 3. Herein, R P T O is perturbed with a fixed-step Δ R P T O in order to search the maximum absorbed power P P T O in real time. The update duration of R P T O is set as Δ T . This means R P T O is automatically updated at time intervals of Δ T . The perturbation direction of R P T O will be flipped once the P P T O value at present is lower than that at the previous moment. Similar algorithms were adopted in Refs. [19,20,21,22].
ADM is based on the moving-least-square (MLS) power forecast and maximum capture width ratio (CWR) tracking. An analogous algorithm is described in Ref. [24]. The structure of an ADM is displayed in Figure 4. Herein, R P T O is perturbed with a variable step size Δ R P T O _ A D M . Δ R P T O _ A D M is given by
Δ R P T O _ A D M = a C W R R P T O _ A D M , 1 R P T O _ A D M , 2 = a ( P P T O P P T O _ f o r e c a s t ) / P w a v e R P T O _ A D M , 1 R P T O _ A D M , 2
where a means the step size coefficient, CWR the capture width ratio, R P T O _ A D M , 1 or R P T O _ A D M , 2 the PTO damping at the last moment or the moment before last, P P T O _ f o r e c a s t the absorbed power predicted by a MLS power forecast, and P w a v e as the input wave power P w a v e = J i r b . J i r is the wave power per unit width for irregular waves, and b is the NIPWEC dimension perpendicular to the wave propagation direction.

3. Model: NIPWEC Dynamics

Based on Figure 5, the dynamic model of a NIPWEC can be expressed as the following Cummins’ equation [37].
( J P + M ( l M ( t ) ) + J 55 ) θ ¨ ( t ) + 0 t K ( t τ ) θ ˙ ( τ ) d τ + ( F B l B G P l P G M l M ( t ) ) sin θ ( t ) = M E + M P T O + M e n d
where J P + M is the total moment of inertia, J 55 the additional mass at the wave frequency of + , θ is the swing angle, K ( t ) is the impulse response function of K ( ω ) , F B is the buoyancy force, G P is the pendulum-hull gravity force, G M is the internal-mass gravity force, l B or l P is the distance between the rotation center R and the buoyancy center B or the pendulum-hull mass center P, M E is the wave excitation moment, M P T O is the PTO moment, M e n d is the endstop moment to prevent θ from exceeding its limits.
J P + M is the function of l M ( t ) , and can be described by
J P + M = J P + J M 0 + m M l M 2 ( t )
where J P is the pendulum-hull moment of inertia against R, J M 0 is the internal mass moment of inertia against M, m M is the internal-mass mass.
K ( ω ) can be represented as
K ( ω ) = R 55 ( ω ) + j ω ( J 55 ( ω ) J 55 )
where R 55 ( ω ) and J 55 ( ω ) denote the radiation damping and additional mass against R, respectively.
According to Ref. [38], M E can be written as
M E ( t ) = j = 1 m f e 5 ( ω j ) A ( ω j ) cos ( ω j t + θ e 5 ( ω j ) + 2 π rand ( ) )
where m is the number of wave frequency elements, ω j the j -th wave frequency element, f e 5 and θ e 5 are the amplitude and phase angle of the excitation-force coefficient, respectively, A is the wave amplitude, and rand ( ) is a random number in (0,1). The expression of A has already been revealed in Ref. [27]. This expression is based on the random interval division for wave frequency elements in order to avoid reproducing the WES or M E ( t ) .
M P T O and M e n d can be given by
M P T O = k g M g
M e n d = r e n d F e n d
where M g is the PMSG moment, and r e n d is the distance from R to the action point of the endstop force F e n d . Detailed formulas for M g and F e n d can be seen in Ref. [39].

4. Parameter Settings

4.1. NIPWEC

We took the NIPWEC with a width of 5 m and a side area of 20 m2 as a research case. The overall appearance and dimensions are illustrated in Figure 6. The detailed design parameters of this NIPWEC are listed in Table 1. In addition, the frequency-domain curves for the hydrodynamic parameters are shown in Figure 7.

4.2. Irregular Wave Environments

Irregular wave simulations were implemented in MATLAB/Simulink. In order to study the applicability of MLTB MPPT for diverse irregular waves, the following three common wave spectrums are introduced to simulate irregular wave environments.
(1) JONSWAP spectrum [40]
{ S ω ( ω ) = α H s 2 ( 2 π T p ) 4 ω 5 exp ( 5 4 ( 2 π T p ) 4 ω 4 ) γ a a = exp ( 1 2 σ 2 ( ω 2 π T p 1 ) 2 )
where H s and T p are the significant wave height and the peak period, respectively, γ (1 < γ < 10) means the peak elevation factor, σ signifies the peak shape parameter, and α is the energy scale parameter that ensures the validity of the following equation:
H s 2 = 16 0 S ω ( ω ) d ω
wherein γ is positively correlated with the slenderness of a spectrum peak. When γ = 3.3, the wave spectrum can also be called a “standard JONSWAP spectrum”. Moreover, σ is given segmentally in the frequency domain. The formula of σ is
σ = { 0 . 07 , ω 2 π T p 0 . 09 , ω > 2 π T p
(2) Pierson-Moskowitz (P-M) spectrum [40]:
{ S ω ( ω ) = 1 4 π H s 2 ( 2 π T z ) 4 ω 5 exp ( 1 π ( 2 π T z ) 4 ω 4 ) T z = 0 . 7105 T p
where T z is the zero-crossing period.
(3) Ochi-Hubble spectrum [41]:
S ω ( ω ) = 1 4 i = 1 2 ( λ i + 1 4 ) λ i Γ ( λ i ) ( 2 π T p , i ) 4 λ i H s , i 2 ω 4 λ i + 1 exp ( ( λ i + 1 4 ) ω 4 ( 2 π T p , i ) 4 )
where H s , i , T p , i , and λ i ( i = 1 or 2 ) are the significant wave height, peak period, and shape parameter of the low/high frequency section, respectively. Herein, these six parameters can be separately given by a function of the significant wave height H s ( H s = H s , 1 2 + H s , 2 2 ) for the Ochi-Hubble spectrum.
According to the above three wave spectra, eight sea states, i.e., SS1–SS8, can be defined in Table 2. Among them, SS1–SS5 are set as the standard JONSWAP spectra with different significant wave heights H s and peak periods T p in order to study the MPPT performance for the same wave-spectrum type but different H s or T p . Meanwhile, SS6–SS8 are defined as the JONSWAP spectrum, P-M spectrum, and Ochi-Hubble spectrum with the same H s 2 , respectively, for the sake of investigation into the MPPT performance of different wave-spectrum types. Moreover, the other investigated characteristic periods T c of each sea state are also shown in Table 2. The selection criteria of an investigated T c is based on whether it is common and easily available from the wave spectra of irregular wave environments, wherein the energy period T e , the mean period T m , and the zero-crossing period T z can all be conveniently obtained via spectral moments. The definition for a n -th spectral moment m n is
m n = 0 ω n S ( ω ) d ω
Thereby T e , T m , and T z can be given by
{ T e = 2 π m 1 / m 0 T m = 2 π m 0 / m 1 T z = 2 π m 0 / m 2
In addition, the wave spectrum density functions of the eight sea states are all displayed in Figure 8.

4.3. MPPT

4.3.1. MPPT Implementation Methods

In order to examine the effectiveness of the three implementation methods, the NIPWEC time-domain response in the regular wave with a wave period of 6 s and a wave height of 0.5 m was simulated. All the parameters are set for the optimal phase and amplitude. For M1, X = 1 m , R P T O =49,000 Nms. With regard to M2, K P T O = −1.055 × 106 Nm, and R P T O = 79,000 Nms. In terms of M3, K P T O = −1.055 × 106 Nm, and R P T O = 43,000 Nms. Moreover, the calculation formulas for the instantaneous input power P I N , absorbed power P P T O , PTO active power P R , and PTO reactive power P K are
{ P I N = M E θ ˙ P P T O = M P T O θ ˙ P R = R P T O θ ˙ 2 P K = K P T O θ θ ˙ or P K = K P T O sin θ θ ˙
According to M1, X max , R P T O max , the maximum time-averaged absorbed power P ¯ P T O max , and the maximum capture widths C W max corresponding to the diverse NIPWEC MPPs were obtained by a series of time-domain simulations in the regular waves of different wave periods and wave heights. Herein, C W max is computed by
C W max = P ¯ P T O max / J r = R P T O max θ ˙ 2 ¯ / J r
where J r signifies the input wave power per unit width for a certain regular wave.

4.3.2. MLTB MPPT

The resonant states where the NIPWEC natural period is consistent with the four T c , i.e., T p , T e , T m , and T z , were comprehensively investigated for power assessments in terms of irregular waves. Herein, three prerequisites are set uniformly. First, the simulation duration of each case is set as 1000 s. Second, the optimal, fixed PTO damping R P T O _ O F D against a resonant state is searched ergodically from 0 to 2 × 105 Nms (with an interval of 1 × 104 Nms) in order to find the maximum time-averaged absorbed power P ¯ P T O _ O F D under the mode of fixed PTO damping. Third, for the sake of the sensitivity study, the resonant states of the NIPWEC natural period are equal to 0.95–1.05 T c (with an interval of 0.025 T c ); these were researched. In addition, the sum δ v a r of the absolute values of the maximum positive/negative relative deviations was chosen as a unified sensitivity indicator for a certain variable v a r , e.g., P ¯ P T O _ O F D or R P T O _ O F D . A larger δ v a r means that v a r is more sensitive to a certain T c .
In order to study the time-averaged absorbed power and parameter sensitivity of different PTO damping tuning algorithms conveniently, the following three prerequisites are set to be consistent. First, the simulation duration of each case is set as 1000 s as well. Second, R P T O is restricted between 20,000 Nms and 200,000 Nms for whichever algorithm is adopted. Thirdly, all studied cases are based on the resonant state, where the natural period of a NIPWEC is identical to the energy period T e .
The selection criteria of an investigated key parameter for PTO damping tuning simulations are represented by whether the parameter is able to influence the MPPT tracking performance.
In terms of LTB PTO damping tuning, the investigated parameter is the number of single waves k w . A k w that is too large could weaken the real-time tracking performance and restrict the ability to obtain globally optimal solutions. However, if k w is too small, the negative power deviation for tuning processes could be accumulated quickly. This may also limit the capability of optimizing globally. Herein, the cases of k w = 1–10, i.e., the cases of the variable update duration mainly distributed within 4–90 s, are investigated for the comprehensive power assessments and sensitivity study.
The investigated parameters for a fixed-step P&O algorithm are the PTO-damping step size Δ R P T O and the update duration Δ T . A too large Δ R P T O and a too small Δ T will make the search process unstable. Meanwhile, a too small Δ R P T O and a too large Δ T will result in a reduction in the search speed. Therefore, the cases of Δ R P T O = 10,000 Nms and Δ T = 5–90 s (with an interval of 5 s), as well as the cases of Δ R P T O = 1000–20,000 Nms (with an interval of 1000 Nms) and Δ T = 10 s, were simulated in order. Moreover, the initial PTO damping was set to 20,000 Nms.
For an ADM, the step size coefficient a is chosen as an investigated parameter. Like the Δ R P T O for a fixed-step P&O algorithm, a too large a can make the search process unstable, whereas a too small a can result in a reduction in search speed. Therefore, the cases of a = 500–10,000 Nms3 (with an interval of 500 Nms3) were thoroughly researched. Moreover, the ADM sampling time was determined as 0.01 s, whereas its initial PTO damping was also set as 20,000 Nms.
In addition, the overall sensitivity value | Λ P | ¯ is proposed as a unified sensitivity indicator of the three PTO damping tuning algorithms. | Λ P | ¯ can be calculated as
| Λ P | ¯ = 1 m Λ k = 1 m Λ | ( P ¯ P T O P ¯ P T O _ O F D ) k ( P ¯ P T O P ¯ P T O _ O F D ) k 1 p k p k 1 |
where m Λ is the total number of first-order difference quotients, P ¯ P T O is the time-averaged absorbed power against different PTO damping tuning algorithms, and p the normalized value of a parameter. The normalization method of p is given by
p = p min { p 0 , , p m Λ } max { p 0 , , p m Λ } min { p 0 , , p m Λ }
A larger | Λ P | ¯ means that the relative fluctuation of the time-averaged absorbed power is more sensitive to a certain algorithm parameter.

5. Results and Discussion

5.1. Comparison of Three MPPT Implementation Methods

Figure 9 shows the NIPWEC time-domain response against three MPPT implementation methods. The findings are as follows.
(1) All three methods can achieve a basic in-phase situation in terms of the angular velocity θ ˙ and the wave excitation moment M E , which results in the positive value of P I N most of the time (see Figure 9a–c).
(2) Both M1 and M3 can make the NIPWEC operate at the larger θ ˙ , leading to the higher P R , when compared to M2.
(3) Both M2 and M3 generate enormous M P T O or P K , which substantially increases the difficulty of the PTO moment control. The M P T O amplitudes for M2 and M3 can reach approximately 10 times that of M E , while the P K amplitudes for them can attain more than five times that of P I N . However, M P T O for M1 has the same magnitude as M E . Moreover, the P K for M1 disappears all the time.
In general, M1 can achieve high active power without reactive power generation and the occurrence of an excessive PTO moment. Therefore, the subsequent MPPT research is based on M1.

5.2. MPP Analysis in Regular Waves

The resonance position X max , the optimal PTO damping R P T O max , the maximum time-averaged absorbed power P ¯ P T O max , and the maximum capture widths C W max against different MPPs are listed in Table 3, Table 4, Table 5 and Table 6. It should be noted that the wave steepness of some regular waves, e.g., the regular wave of the wave period T = 3 s and the wave height H = 2.5 m, is greater than 0.1412. According to Schwartz’s theory [42], the aforementioned regular waves do not exist since wave breaking happens. Therefore, the corresponding R P T O max , P ¯ P T O max , and C W max are all set as zero. Three key findings emerge.
(1) X max is only related to T , but R P T O max depends on both T and H . As T increases, R P T O max shows a trend of first increasing and then decreasing. However, as H increases, R P T O max only shows an increasing trend.
(2) P ¯ P T O max is related to both T and H . Moreover, P ¯ P T O max appears to have the same variation law as R P T O max .
(3) C W max is also related to both T and H . Nevertheless, the variation law for C W max is different from that of R P T O max or P ¯ P T O max . As T increases, C W max increases first and then decreases. When H increases, C W max gradually decreases.
Moreover, both P ¯ P T O max and C W max were compared with the corresponding solutions from the preliminary frequency-domain calculations in order to verify the correctness of the time-domain simulations (see Figure 10). The results indicate that P ¯ P T O max or C W max is always less than or equal to the frequency-domain solution. When both T and H are small, the frequency-domain solution fits P ¯ P T O max or C W max better than other cases. Moreover, the capture widths derived from the time/frequency-domain model are all below the theoretical upper bound. Hence, the time-domain simulation correctness is effectively verified.
In order to further explain the variation law of MPPs, the regular waves of T = 6 s and H = 0.5–3 m, as well as the ones of T = 3–11 s and H = 1.5 m are taken as examples. The time-averaged absorbed power curves against different H or T are displayed in Figure 11. The results show that R P T O max is usually larger than R 55 ( ω r ) . The difference between R P T O max and R 55 ( ω r ) becomes more obvious, along with the increase in H since R 55 ( ω r ) is independent of H , whereas R P T O max increases synchronously. In addition, the relationship between R P T O max and P ¯ P T O max presents an approximately quadratic correlation under the circumstance of the same T combined with a different H . Meanwhile, there exists a hysteresis phenomenon in the R P T O max - P ¯ P T O max relationship, under the circumstance of the same H combined with a different T .
In summary, the MPPs in regular waves can be effectively gained via time-domain simulations. Moreover, the R P T O max and R P T O max - P ¯ P T O max relationship is totally different from the frequency-domain solutions.

5.3. Power Assessments and Sensitivity Study for the LTB Internal-Mass Position Adjustment

The resonant states against different characteristic periods T c of the eight sea states were simulated. Moreover, the resonant states against 0.95–1.05 T c were also considered for the sensitivity study. The statistical results of P ¯ P T O _ O F D and R P T O _ O F D are listed in Table 7 and Table 8, respectively. The largest A P as well as the corresponding A R of each sea state is in bold and is underlined. Herein, A P signifies the arithmetic mean of P ¯ P T O _ O F D , whereas A R means the arithmetic mean of R P T O _ O F D . Moreover, Figure 12 visually displays the energy capture performance of different T c resonances in the form of a histogram. The maximum P ¯ P T O _ O F D and the minimum P ¯ P T O _ O F D are represented by error bars. The results demonstrate the following three things.
(1) Although there is a certain possibility for all T c resonances to obtain the largest A P , it is better to keep resonant with T e because T e resonance can lead 50% of sea states to achieve the largest A P , as well as allowing other sea states to obtain a submaximal A P .
(2) When the largest A P is achieved, the unified sensitivity indicator δ P for P ¯ P T O _ O F D is relatively low. This means that P ¯ P T O _ O F D is not sensitive to the fine tuning of T c when P ¯ P T O _ O F D appears around the largest A P . Moreover, T e resonance owns the minimum or subminimum δ P in six out of eight sea states. Hence, T e resonance is not sensitive to parameter changes in most sea states.
(3) Despite being in the same sea state, A R against different T c is quite different, and the change law of A R along with the variations in T c is elusive. Moreover, in contrast to P ¯ P T O _ O F D , R P T O _ O F D is still sensitive to the fine tuning of T c when P ¯ P T O _ O F D appears around the largest A P .
In summary, T e resonance is preferable in terms of power assessments when compared to other T c resonances. Moreover, the performance reliability of the T e resonance is high since its parameter sensitivity is low in most sea states. Hence, the subsequent simulations for three PTO damping tuning algorithms are all based on the T e resonance.

5.4. Power Assessments and Sensitivity Study for the LTB PTO Damping Tuning

Power assessments for the three PTO damping tuning algorithms, i.e., the fixed-step P&O algorithm and the ADM and LTB PTO damping tuning algorithms, are listed in Table 9, Table 10 and Table 11. The upper bound of the maximum or mean P ¯ P T O of three algorithms in each sea state is in bold and underlined. The results indicate the following three facts.
(1) In terms of the mean P ¯ P T O against the different parameters, LTB PTO damping tuning can obtain the maximum mean value in five out of eight sea states. Moreover, it can also achieve submaximum mean values in the other sea states. However, the fixed-step P&O algorithm can only get the maximum mean value in SS2 or SS3. Moreover, ADM has the best tracking performance for the SS8 of an Ochi-Hubble spectrum.
(2) For the maximum value of the time-averaged absorbed power against the different parameters, the obtaining probability of the three algorithms are not significantly different.
(3) With regard to | Λ P | ¯ , both the | Λ P | ¯ of the fixed-step P&O algorithm and the | Λ P | ¯ of the ADM are much larger than that of the LTB PTO damping tuning. Therefore, the former two algorithms are more sensitive to parameter changes than the LTB PTO damping tuning. Moreover, the fixed-step P&O algorithm is more sensitive to Δ T than to Δ R P T O .
Figure 13 visually displays the P ¯ P T O of the three algorithms in the form of a histogram, wherein the maximum P ¯ P T O and the minimum P ¯ P T O are represented by error bars. From Figure 13, it can be seen that the bar length, i.e., the mean P ¯ P T O , of the LTB PTO damping tuning is generally higher, and its error bar is the narrowest. Although the fixed-step P&O algorithm and ADM can obtain the maximum average absorbed power under certain sea states through parameter tuning, such as the fixed-step P&O algorithm in SS2, SS3, or SS7 (with the optimal { Δ R P T O , Δ T } = {5000 Nms, 10 s}, {8000 Nms, 10 s} or {10,000 Nms, 80 s}, respectively), and the ADM in SS1 or SS8 (with the optimal a = 1500 Nms3 or 4000 Nms3, respectively), the mean P ¯ P T O of them is generally low and owns a wider error bar.
Furthermore, we take SS2, SS5, SS6, and SS7 as examples. The variations in P ¯ P T O as a function of the algorithm parameters against the above sea states are displayed in Figure 14 for the sake of the visual indication of the specific impact of parameter changes on algorithm performance. It can be seen that P ¯ P T O fluctuates at a wide range with no obvious change pattern when the Δ T or Δ R P T O alters equidistantly. When a changes in an equal interval, the fluctuation range of the average absorption power of the ADM is smaller than that of the fixed-step P&O algorithm. However, there exist several sudden increases at certain values of a . When k w increases one by one, the variations in P ¯ P T O against the LTB PTO damping tuning is the smoothest without significant fluctuations.
In summary, the LTB PTO damping tuning is the least sensitive to parameter changes and has a high average absorbed power value. The overall reliability of this algorithm is high. Although the fixed-step P&O method and admittance differentiation method can search for larger power points against some specific parameters and sea states, the optimal parameters corresponding to the above power points have significant differences in different sea states, which makes it difficult to determine the universal optimal parameters. Moreover, the high sensitivity to parameter changes is another limitation.

6. Conclusions

This paper attempts to find a suitable MPPT technology for the optimized operation of a NIPWEC. First, a unique MPPT implementation method, i.e., adjusting the internal-mass position to achieve “optimal phase” and tuning the linear PTO damping to achieve “optimal amplitude”, was proposed. Then, MPPs in regular waves were discussed, and two lookup tables, i.e., a 1D resonance position table and a 2D optimal PTO damping table, were determined. Afterwards, an MLTB MPPT algorithm was put forward for the NIPWEC real-time operation in irregular waves, according to the aforementioned two lookup tables. MLTB MPPT consists of two core parts, i.e., the LTB internal-mass position adjustment and the LTB PTO damping tuning. Finally, the power assessments and sensitivity study for the LTB internal-mass position adjustment and the LTB PTO damping tuning were respectively conducted for eight simulated irregular-wave sea states of diverse wave spectra. The findings are as follows.
(1) It is preferable to choose M1, i.e., simultaneously adjusting the internal mass position and PTO damping, to realize MPPT. First, M1 has the same active power generation performance as the PTO impedance adjustment, with consideration given to the nonlinear effect of the hydrostatic restoring moment. Second, M1 does not generate excessive reactive power and PTO moment.
(2) For the LTB internal-mass position adjustment in irregular waves, energy period resonance possesses the highest possibility to obtain the maximum mean time-averaged absorbed power when compared with other characteristic period resonances. Moreover, the energy period resonance has high reliability since its parameter sensitivity is low in most sea states.
(3) When compared to the fixed-step P&O algorithm and a typical variable-step P&O algorithm, i.e., ADM, LTB PTO damping tuning is the most robust to parameter changes and has the highest mean time-averaged absorbed power in the majority of sea states.
In the future, the tracking performance of MLTB MPPT in changing sea states will be studied in-depth in order to further validate its applicability.

Author Contributions

Conceptualization, X.Y. and Q.C.; methodology, X.Y.; software, X.Y. and D.G.; validation, J.Z., F.M. and D.G.; formal analysis, X.Y.; investigation, D.G.; data curation, J.L.; writing—original draft preparation, X.Y.; writing—review and editing, D.G.; visualization, X.Y.; supervision, Q.C. and J.L.; funding acquisition, J.Z. and F.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Grant No. 52179088) and the Science & Technology Project of Enterprise (Grant No. ZKY2023-HCDK-02-01).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Guo, B.; Ringwood, J.V. A Review of Wave Energy Technology from a Research and Commercial Perspective. IET Renew. Power Gener. 2021, 15, 3065–3090. [Google Scholar] [CrossRef]
  2. Huckerby, J.; Jeffrey, H.; de Andres, A.; Finlay, L. An International Vision for Ocean Energy; OES: Lisbon, Portugal, 2017. [Google Scholar]
  3. IEA Electricity Consumption—Electricity Information: Overview. Available online: https://www.iea.org/reports/electricity-information-overview/electricity-consumption#abstract (accessed on 31 May 2023).
  4. López, I.; Andreu, J.; Ceballos, S.; Martínez De Alegría, I.; Kortabarria, I. Review of Wave Energy Technologies and the Necessary Power-Equipment. Renew. Sustain. Energy Rev. 2013, 27, 413–434. [Google Scholar] [CrossRef]
  5. Sheng, W. Wave Energy Conversion and Hydrodynamics Modelling Technologies: A Review. Renew. Sustain. Energy Rev. 2019, 109, 482–498. [Google Scholar] [CrossRef]
  6. Falnes, J. Ocean Waves and Oscillating Systems: Linear Interaction Including Wave Energy Extraction; Cambridge University Press: Cambridge, MA, USA, 2002; ISBN 0511030932. [Google Scholar]
  7. Wu, J.; Yao, Y.; Zhou, L.; Göteman, M. Real-Time Latching Control Strategies for the Solo Duck Wave Energy Converter in Irregular Waves. Appl. Energy 2018, 222, 717–728. [Google Scholar] [CrossRef]
  8. Garcia-Rosa, P.B.; Fosso, O.B.; Molinas, M. Switching Sequences for Non-Predictive Declutching Control of Wave Energy Converters. IFAC-PapersOnLine 2020, 53, 12295–12300. [Google Scholar] [CrossRef]
  9. Costa, P.R.; Garcia-Rosa, P.B.; Estefen, S.F. Phase Control Strategy for a Wave Energy Hyperbaric Converter. Ocean Eng. 2010, 37, 1483–1490. [Google Scholar] [CrossRef]
  10. Qiu, S.; Ye, J.; Wang, D.; Liang, F. Experimental Study on a Pendulum Wave Energy Converter. China Ocean Eng. 2013, 27, 359–368. [Google Scholar] [CrossRef]
  11. Marei, M.I.; Mokhtar, M.; El-Sattar, A.A. MPPT Strategy Based on Speed Control for AWS-Based Wave Energy Conversion System. Renew. Energy 2015, 83, 305–317. [Google Scholar] [CrossRef]
  12. Chen, Q.; Shang, X.; Huang, X.; Tan, T.; Chen, X.; Xu, Z. Resonant Wave Energy Power Take-off System Based on Hydraulic Transmission. Autom. Electr. Power Syst. 2016, 40, 115–120. [Google Scholar]
  13. Wu, J.; Yao, Y.; Li, W.; Zhou, L.; Göteman, M. Optimizing the Performance of Solo Duck Wave Energy Converter in Tide. Energies 2017, 10, 289. [Google Scholar] [CrossRef]
  14. Cai, Y.; Huo, Y.; Shi, X.; Liu, Y. Numerical and Experimental Research on a Resonance-Based Wave Energy Converter. Energy Convers. Manag. 2022, 269, 116152. [Google Scholar] [CrossRef]
  15. Cai, Y.; Shi, X.; Huo, Y.; Liu, Y. Experimental Study on a Pitching Wave Energy Converter with Adjustable Natural Period. Ocean Eng. 2022, 261, 112128. [Google Scholar] [CrossRef]
  16. Zheng, Y.; Chen, Q.; Yue, X.; Geng, D.; Niu, Y.; Gu, X. Active-Resonance Technology for Wave Energy Utilization: An Efficient Offshore Distributed Renewable Power Generation Alternative. Energy Convers. Manag. 2023, 292, 117359. [Google Scholar] [CrossRef]
  17. Tom, N.; Lawson, M.; Yu, Y.-H.; Wright, A. Spectral Modeling of an Oscillating Surge Wave Energy Converter with Control Surfaces. Appl. Ocean Res. 2016, 56, 143–156. [Google Scholar] [CrossRef]
  18. Qasim, M.A.; Alwan, N.T.; PraveenKumar, S.; Velkin, V.I.; Agyekum, E.B. A New Maximum Power Point Tracking Technique for Thermoelectric Generator Modules. Inventions 2021, 6, 88. [Google Scholar] [CrossRef]
  19. Amon, E.A.; Schacher, A.A.; Brekken, T.K.A. A Novel Maximum Power Point Tracking Algorithm for Ocean Wave Energy Devices. In Proceedings of the 2009 IEEE Energy Conversion Congress and Exposition, ECCE 2009, San Jose, CA, USA, 20–24 September 2009; pp. 2635–2641. [Google Scholar]
  20. Mendes, R.; Calado, M.D.R.; Mariano, S. Maximum Power Point Tracking for a Point Absorber Device with a Tubular Linear Switched Reluctance Generator. Energies 2018, 11, 2192. [Google Scholar] [CrossRef]
  21. Artal-Sevil, J.S.; Martínez-López, D.; Guillén-Asensio, A.; Domínguez-Navarro, J.A. Wave Energy Converter Model Based on a Decentralized Hybrid Energy Storage System with MPPT Control Algorithm. In Proceedings of the 2020 Fifteenth International Conference on Ecological Vehicles and Renewable Energies (EVER), Monte-Carlo, Monaco, 10–12 September 2020; pp. 1–11. [Google Scholar]
  22. 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, 447. [Google Scholar] [CrossRef]
  23. Yang, J. Research on Key Technologies of Horizontal Pendulum Wave Energy Converter; Zhejiang University: Hangzhou, China, 2018. [Google Scholar]
  24. Zhou, Y. Research on the Power Control Technology of Floating Pendulum Wave Energy Converter; Zhejiang University: Hangzhou, China, 2019. [Google Scholar]
  25. Xu, J.; Yang, Y.; Hu, Y.; Xu, T.; Zhan, Y. MPPT Control of Hydraulic Power Take-off for Wave Energy Converter on Artificial Breakwater. J. Mar. Sci. Eng. 2020, 8, 304. [Google Scholar] [CrossRef]
  26. Mishra, S.; Purwar, S.; Kishor, N. Maximizing Output Power in Oscillating Water Column Wave Power Plants: An Optimization Based MPPT Algorithm. Technologies 2018, 6, 15. [Google Scholar] [CrossRef]
  27. Yue, X.; Geng, D.; Chen, Q.; Zheng, Y.; Gao, G.; Xu, L. 2-D Lookup Table Based MPPT: Another Choice of Improving the Generating Capacity of a Wave Power System. Renew. Energy 2021, 179, 625–640. [Google Scholar] [CrossRef]
  28. Zhao, A.; Wu, W.; Sun, Z.; Zhu, L.; Lu, K.; Chung, H.; Blaabjerg, F. A Flower Pollination Method Based Global Maximum Power Point Tracking Strategy for Point-Absorbing Type Wave Energy Converters. Energies 2019, 12, 1343. [Google Scholar] [CrossRef]
  29. M’zoughi, F.; Garrido, I.; Garrido, A.J.; De La Sen, M. Rotational Speed Control Using ANN-Based MPPT for OWC Based on Surface Elevation Measurements. Appl. Sci. 2020, 10, 8975. [Google Scholar] [CrossRef]
  30. Agyekum, E.B.; PraveenKumar, S.; Eliseev, A.; Velkin, V.I. Design and Construction of a Novel Simple and Low-Cost Test Bench Point-Absorber Wave Energy Converter Emulator System. Inventions 2021, 6, 20. [Google Scholar] [CrossRef]
  31. Yao, G.; Luo, Z.; Lu, Z.; Wang, M.; Shang, J.; Guerrerob, J.M. Unlocking the Potential of Wave Energy Conversion: A Comprehensive Evaluation of Advanced Maximum Power Point Tracking Techniques and Hybrid Strategies for Sustainable Energy Harvesting. Renew. Sustain. Energy Rev. 2023, 185, 113599. [Google Scholar] [CrossRef]
  32. Dong, F.; Pan, S.; Gong, J.; Cai, Y. Maximum Power Point Tracking Control Strategy Based on Frequency and Amplitude Control for the Wave Energy Conversion System. Renew. Energy 2023, 215, 118973. [Google Scholar] [CrossRef]
  33. Ouyang, Z.; Zheng, Y.; Yue, X.; Chen, Q. Wave-to-Wire Dynamic Modeling and Time-Domain Characteristic Analysis for the Novel Inverse-Pendulum Wave Energy Converter. J. Ocean Technol. 2023, 42, 94–104. [Google Scholar] [CrossRef]
  34. Peña-Sanchez, Y.; Windt, C.; Davidson, J.; Ringwood, J.V. A Critical Comparison of Excitation Force Estimators for Wave-Energy Devices. IEEE Trans. Control Syst. Technol. 2020, 28, 2263–2275. [Google Scholar] [CrossRef]
  35. Hillis, A.J.; Brask, A.; Whitlam, C. Real-Time Wave Excitation Force Estimation for an Experimental Multi-DOF WEC. Ocean Eng. 2020, 213, 107788. [Google Scholar] [CrossRef]
  36. Lobeto, H.; Menendez, M.; Losada, I.J. Future Behavior of Wind Wave Extremes Due to Climate Change. Sci. Rep. 2021, 11, 7869. [Google Scholar] [CrossRef]
  37. Cummins, W.E. The Impulse Response Function and Ship Motion. Schiffstechnik 1962, 9, 101–109. [Google Scholar]
  38. Sheng, W.; Alcorn, R.; Lewis, A. A New Method for Radiation Forces for Floating Platforms in Waves. Ocean Eng. 2015, 105, 43–53. [Google Scholar] [CrossRef]
  39. Xuhui, Y.; Qijuan, C.; Zenghui, W.; Dazhou, G.; Donglin, Y.; Wen, J.; Weiyu, W. A Novel Nonlinear State Space Model for the Hydraulic Power Take-off of a Wave Energy Converter. Energy 2019, 180, 465–479. [Google Scholar] [CrossRef]
  40. Molin, B. Hydrodynamique des Structures Offshore; Editions Technip: Paris, France, 2002; ISBN 2710808153. [Google Scholar]
  41. Golpira, A.; Panahi, R.; Shafieefar, M. Developing Families of Ochi-Hubble Spectra for the Northern Parts of the Gulf of Oman. Ocean Eng. 2019, 178, 345–356. [Google Scholar] [CrossRef]
  42. Schwartz, L.W. Computer Extension and Analytic Continuation of Stokes’ Expansion for Gravity Waves. J. Fluid Mech. 1974, 62, 553–578. [Google Scholar] [CrossRef]
  43. Babarit, A. A Database of Capture Width Ratio of Wave Energy Converters. Renew. Energy 2015, 80, 610–628. [Google Scholar] [CrossRef]
Figure 1. Mechanical structure and electrical topology of the NIPWEC with a DC bus.
Figure 1. Mechanical structure and electrical topology of the NIPWEC with a DC bus.
Energies 16 06195 g001
Figure 2. Schematic diagram of the control structure for MLTB MPPT.
Figure 2. Schematic diagram of the control structure for MLTB MPPT.
Energies 16 06195 g002
Figure 3. Block diagram of the fixed step P&O algorithm.
Figure 3. Block diagram of the fixed step P&O algorithm.
Energies 16 06195 g003
Figure 4. Block diagram of the ADM. Note: Δ R P T O 0 = 50 Nms; Δ T 0 = 0.01 s.
Figure 4. Block diagram of the ADM. Note: Δ R P T O 0 = 50 Nms; Δ T 0 = 0.01 s.
Energies 16 06195 g004
Figure 5. Force analysis of the NIPWEC. Herein, F E , F R , G M , G P , and F B mean the wave excitation force, radiation force, internal-mass gravity force, pendulum-hull gravity force, and buoyancy force, respectively, M E , M R , M M , M P , and M B are the corresponding moments against R.
Figure 5. Force analysis of the NIPWEC. Herein, F E , F R , G M , G P , and F B mean the wave excitation force, radiation force, internal-mass gravity force, pendulum-hull gravity force, and buoyancy force, respectively, M E , M R , M M , M P , and M B are the corresponding moments against R.
Energies 16 06195 g005
Figure 6. The appearance and dimensions of the researched NIPWEC. (a) A three-dimensional view. (b) The basic dimensions of the side profile.
Figure 6. The appearance and dimensions of the researched NIPWEC. (a) A three-dimensional view. (b) The basic dimensions of the side profile.
Energies 16 06195 g006
Figure 7. Frequency-domain curves for the hydrodynamic parameters of the researched NIPWEC. (a) Excitation force coefficient amplitude: f e 5 . (b) Excitation-force-coefficient phase angle: θ e 5 . (c) Radiation damping: R 55 . (d) Additional mass: J 55 .
Figure 7. Frequency-domain curves for the hydrodynamic parameters of the researched NIPWEC. (a) Excitation force coefficient amplitude: f e 5 . (b) Excitation-force-coefficient phase angle: θ e 5 . (c) Radiation damping: R 55 . (d) Additional mass: J 55 .
Energies 16 06195 g007
Figure 8. Wave spectrum density functions of eight sea states.
Figure 8. Wave spectrum density functions of eight sea states.
Energies 16 06195 g008
Figure 9. NIPWEC time-domain response against three MPPT implementation methods. (a) Angular velocity: θ ˙ . (b) PTO moment: M P T O and wave excitation moment: M E . (c) Input power: P I N . (d) Absorbed power: P P T O . (e) PTO active power: P R . (f) PTO reactive power: P K .
Figure 9. NIPWEC time-domain response against three MPPT implementation methods. (a) Angular velocity: θ ˙ . (b) PTO moment: M P T O and wave excitation moment: M E . (c) Input power: P I N . (d) Absorbed power: P P T O . (e) PTO active power: P R . (f) PTO reactive power: P K .
Energies 16 06195 g009
Figure 10. Variations in the time-averaged absorbed power (a) and the capture width C W (b) as the function of T against different H . Note: solid lines: resonance and R P T O = R P T O max ; Dashed lines: resonance and R P T O = max { R 55 ( ω r ) , R P T O _ min ( ω r ) } ; Red dotted line: the theoretical upper bound of C W , i.e., λ / π [43]. Herein, R P T O _ min ( ω r ) means the minimum PTO damping against ω = ω r , which is derived from the upper limitation for the swing angle amplitude. λ is the wavelength.
Figure 10. Variations in the time-averaged absorbed power (a) and the capture width C W (b) as the function of T against different H . Note: solid lines: resonance and R P T O = R P T O max ; Dashed lines: resonance and R P T O = max { R 55 ( ω r ) , R P T O _ min ( ω r ) } ; Red dotted line: the theoretical upper bound of C W , i.e., λ / π [43]. Herein, R P T O _ min ( ω r ) means the minimum PTO damping against ω = ω r , which is derived from the upper limitation for the swing angle amplitude. λ is the wavelength.
Energies 16 06195 g010
Figure 11. Variations in the time-averaged absorbed power as the function of the linear PTO damping R P T O . (a) T = 6 s and H = 0.5–3 m. (b) T = 3–11 s and H = 1.5 m. Note: “*”: R P T O = R 55 ( ω r ) ; solid triangle points: R P T O = R P T O max .
Figure 11. Variations in the time-averaged absorbed power as the function of the linear PTO damping R P T O . (a) T = 6 s and H = 0.5–3 m. (b) T = 3–11 s and H = 1.5 m. Note: “*”: R P T O = R 55 ( ω r ) ; solid triangle points: R P T O = R P T O max .
Energies 16 06195 g011
Figure 12. Histogram about the maximum time-averaged absorbed power P ¯ P T O _ O F D at different resonant states against different characteristic periods, T c . Herein, the upper and lower bounds are displayed in the form of error bars.
Figure 12. Histogram about the maximum time-averaged absorbed power P ¯ P T O _ O F D at different resonant states against different characteristic periods, T c . Herein, the upper and lower bounds are displayed in the form of error bars.
Energies 16 06195 g012
Figure 13. Time-averaged absorbed power histogram against different PTO damping tuning algorithms. Herein, P ¯ P T O _ O F D against optimal, fixed PTO damping R P T O _ O F D is also illustrated as a reference. Moreover, the upper and lower bounds are displayed in the form of error bars.
Figure 13. Time-averaged absorbed power histogram against different PTO damping tuning algorithms. Herein, P ¯ P T O _ O F D against optimal, fixed PTO damping R P T O _ O F D is also illustrated as a reference. Moreover, the upper and lower bounds are displayed in the form of error bars.
Energies 16 06195 g013
Figure 14. Variations in P ¯ P T O as a function of the parameters of different PTO damping tuning algorithms in four sea states. (a) Δ T of the fixed-step P&O algorithm. (b) Δ R P T O of the fixed-step P&O algorithm. (c) a of the ADM. (d) k w of the LTB PTO damping tuning.
Figure 14. Variations in P ¯ P T O as a function of the parameters of different PTO damping tuning algorithms in four sea states. (a) Δ T of the fixed-step P&O algorithm. (b) Δ R P T O of the fixed-step P&O algorithm. (c) a of the ADM. (d) k w of the LTB PTO damping tuning.
Energies 16 06195 g014
Table 1. The design parameters of the researched NIPWEC.
Table 1. The design parameters of the researched NIPWEC.
ParameterValueUnit
J P 5.138 × 104[kg·m2]
J M 0 1.025 × 105[kg·m2]
m M 9.306 × 104[kg]
l M 0 0.2[m]
J 55 1.844 × 105[kg·m2]
F B 1.014 × 106[N]
l B 1.726[m]
G P 1.015 × 105[N]
l P 1.12[m]
G M 9.129 × 105[N]
m 300[-]
Distance between R and the still water level6[m]
k g 120[-]
n p 3[-]
ϕ f 0.6[Wb]
Table 2. Parameter settings for the simulated sea states.
Table 2. Parameter settings for the simulated sea states.
Sea StateWave-Spectrum Type H s   or   H s , 1 / H s , 2 (m) T p   or   T p , 1 / T p , 2 (s) λ 1 / λ 2 T e (s) T m (s) T z (s)
SS1Standard JONSWAP spectrum1.54/3.613.353.15
SS2Standard JONSWAP spectrum1.56/5.425.014.69
SS3Standard JONSWAP spectrum1.59/8.137.517.01
SS4Standard JONSWAP spectrum0.56/5.425.014.69
SS5Standard JONSWAP spectrum2.56/5.425.014.69
SS6JONSWAP spectrum ( γ = 10 )1.56/5.665.385.14
SS7JONSWAP spectrum ( γ = 1 ), i.e., P-M spectrum1.56/5.144.644.29
SS8Ochi-Hubble spectrum1.12/1.038.36/4.763.43/2.046.355.645.29
Table 3. Resonance position table.
Table 3. Resonance position table.
T (s)33.544.555.566.57
X max (m)0.0090.3180.5240.6820.8090.91311.0711.131
T (s)7.588.599.51010.511
X max (m)1.1811.2251.2611.2931.321.3441.3651.383
Table 4. Optimal PTO damping table. Herein background color is utilized to illustrate the value difference. The deeper red means the larger value, while the deeper blue refers to the smaller one.
Table 4. Optimal PTO damping table. Herein background color is utilized to illustrate the value difference. The deeper red means the larger value, while the deeper blue refers to the smaller one.
R P T O max (Nms) H (m)
0.511.522.53
T (s)338,00041,00049,00059,00000
3.551,00056,00069,00083,00097,0000
449,00064,00083,000101,000119,000135,000
4.546,00071,00094,000115,000135,000154,000
546,00076,000101,000124,000145,000164,000
5.548,00079,000105,000128,000149,000168,000
649,00080,000106,000129,000150,000169,000
6.549,00080,000105,000128,000148,000167,000
749,00079,000104,000126,000146,000164,000
7.548,00077,000101,000122,000141,000159,000
847,00075,00098,000119,000138,000155,000
8.546,00073,00096,000115,000134,000150,000
944,00071,00093,000112,000129,000146,000
9.543,00068,00089,000108,000124,000140,000
1041,00066,00086,000104,000120,000135,000
10.540,00064,00083,000101,000116,000131,000
1139,00062,00081,00098,000113,000127,000
Table 5. Maximum time-averaged absorbed power table. Herein background color is utilized to illustrate the value difference. The deeper red means the larger value, while the deeper blue refers to the smaller one.
Table 5. Maximum time-averaged absorbed power table. Herein background color is utilized to illustrate the value difference. The deeper red means the larger value, while the deeper blue refers to the smaller one.
P ¯ P T O max (W) H (m)
0.511.522.53
T (s)3287811,36024,50641,14400
3.5485119,06540,76967,95299,4270
4712726,70654,60788,417126,927169,406
4.5971332,97963,907100,355141,269185,993
511,64435,90466,814102,619142,422185,667
5.512,19935,26063,98896,941133,371172,809
611,62432,42058,05087,310119,572154,435
6.510,64628,98551,41676,938105,021135,329
7948925,49845,00367,15891,512117,780
7.5840522,34839,28158,48479,574102,310
8724819,26533,86650,43168,62788,245
8.5637316,82229,48743,83859,59476,572
9554514,63425,65338,14151,84966,624
9.5489412,86322,51133,43745,42858,348
10430511,31019,78829,39039,92851,281
10.53787994417,39725,83635,09845,077
113375882915,42122,88031,06239,875
Table 6. Maximum capture width table. Herein background color is utilized to illustrate the value difference. The deeper red means the larger value, while the deeper blue refers to the smaller one.
Table 6. Maximum capture width table. Herein background color is utilized to illustrate the value difference. The deeper red means the larger value, while the deeper blue refers to the smaller one.
C W max (m) H (m)
0.511.522.53
T (s)33.9113.8593.7003.4940.0000.000
3.55.6505.5515.2764.9474.6320.000
47.2646.8046.1845.6325.1744.796
4.58.7997.4696.4335.6825.1194.680
59.4947.3186.0535.2294.6454.205
5.59.0426.5345.2704.4913.9543.558
67.8985.5074.3823.7083.2502.915
6.56.6774.5453.5833.0162.6352.358
75.5263.7122.9122.4442.1321.905
7.54.5683.0372.3721.9871.7301.545
83.6932.4541.9171.6061.3991.249
8.53.0562.0171.5711.3141.1431.020
92.5121.6571.2911.0800.9390.838
9.52.1001.3801.0730.8970.7800.696
101.7551.1530.8960.7490.6510.581
10.51.4700.9650.7500.6270.5450.486
111.2510.8180.6350.5300.4600.410
Table 7. Power assessments for resonant states against different characteristic periods T c . Herein, a sensitivity study is also conducted for 0.95–1.05 T c . A P means the arithmetic mean of P ¯ P T O _ O F D , and δ P is a unified sensitivity indicator for P ¯ P T O _ O F D . The largest A P of each sea state is in bold and underlined.
Table 7. Power assessments for resonant states against different characteristic periods T c . Herein, a sensitivity study is also conducted for 0.95–1.05 T c . A P means the arithmetic mean of P ¯ P T O _ O F D , and δ P is a unified sensitivity indicator for P ¯ P T O _ O F D . The largest A P of each sea state is in bold and underlined.
Sea State T p T e T m T z
A P (w) δ P A P (w) δ P A P (w) δ P A P (w) δ P
SS115,1300.07211,8730.39783140.41672960.428
SS219,3660.35123,3320.16218,3360.58812,7780.560
SS310,3200.39313,3390.09612,4470.15211,2740.161
SS425320.23820780.43415270.35612340.327
SS544,2260.27954,0060.08850,5350.24741,8040.376
SS623,6790.39427,9020.25329,4880.11825,0300.585
SS714,0740.25715,4190.15812,6090.22510,6180.298
SS882480.28589060.39386420.10992140.125
Table 8. Analysis of the optimal, fixed PTO damping R P T O _ O F D for the resonant states against different characteristic periods T c . Herein, a sensitivity study was also conducted for 0.95–1.05 T c . A R means the arithmetic mean of R P T O _ O F D , and δ R is a unified sensitivity indicator for R P T O _ O F D . The A R against the largest A P of each sea state is in bold and underlined.
Table 8. Analysis of the optimal, fixed PTO damping R P T O _ O F D for the resonant states against different characteristic periods T c . Herein, a sensitivity study was also conducted for 0.95–1.05 T c . A R means the arithmetic mean of R P T O _ O F D , and δ R is a unified sensitivity indicator for R P T O _ O F D . The A R against the largest A P of each sea state is in bold and underlined.
Sea State T p T e T m T z
A R (Nms) δ R A R (Nms) δ R A R (Nms) δ R A R (Nms) δ R
SS176,0000.263104,0000.673170,0000.412188,0000.213
SS282,0000.36652,0000.57730,0000.66768,0001.765
SS364,0000.46942,0000.47636,0000.27832,0000.625
SS434,0000.58862,0000.806102,0000.490132,0000.379
SS5122,0000.24694,0000.31974,0000.27066,0000.455
SS686,0000.34970,0000.28652,0000.76932,0000.938
SS772,0000.41748,0000.62574,0000.811100,0000.500
SS870,0000.28696,0001.354142,0000.634110,0000.364
Table 9. Power assessments of the fixed-step P&O algorithm. Herein, a sensitivity study was also conducted for Δ R P T O = 10,000 Nms with Δ T = 5–90 s, as well as Δ T = 10 s, with Δ R P T O = 1000–20,000 Nms. The upper bound of the maximum or mean P ¯ P T O of three algorithms in each sea state is in bold and underlined.
Table 9. Power assessments of the fixed-step P&O algorithm. Herein, a sensitivity study was also conducted for Δ R P T O = 10,000 Nms with Δ T = 5–90 s, as well as Δ T = 10 s, with Δ R P T O = 1000–20,000 Nms. The upper bound of the maximum or mean P ¯ P T O of three algorithms in each sea state is in bold and underlined.
Sea State Δ R P T O = 10,000 Nms Δ T = 10 sOverall Assessment
Max.
P ¯ P T O (W)
Min.
P ¯ P T O (W)
Mean
P ¯ P T O (W)
| Λ P | ¯ Max.
P ¯ P T O (W)
Min.
P ¯ P T O (W)
Mean
P ¯ P T O (W)
| Λ P | ¯ Max.
P ¯ P T O (W)
Min.
P ¯ P T O (W)
Mean
P ¯ P T O (W)
SS111,391793810,0931.39411,567770610,5650.39111,567770610,341
SS223,68518,33120,8670.87723,68016,36720,3631.06823,68516,36720,602
SS312,47610,26411,7330.78712,759976111,3220.51812,759976111,517
SS42050180619410.6532051183419760.425205118061959
SS550,98629,54240,6151.01652,21123,87344,0611.10552,21123,87342,429
SS626,52518,14621,9731.44625,68816,83123,1590.93526,52516,83122,598
SS716,49413,97615,2390.77115,75613,60914,7190.516,49413,60914,965
SS87449596468330.8567500624571980.282750059647025
Table 10. Power assessments of the ADM. Herein, a sensitivity study was also conducted for a = 500–10,000 Nms3. The upper bound of the maximum or mean P ¯ P T O of three algorithms in each sea state is in bold and underlined.
Table 10. Power assessments of the ADM. Herein, a sensitivity study was also conducted for a = 500–10,000 Nms3. The upper bound of the maximum or mean P ¯ P T O of three algorithms in each sea state is in bold and underlined.
Sea StateOverall Assessment
Maximum   P ¯ P T O (W) Minimum   P ¯ P T O (W) Mean   P ¯ P T O (W) | Λ P | ¯
SS111,923903511,0750.824
SS222,72614,66216,2701.672
SS39923751980020.948
SS41971159316931.159
SS552,90841,72344,6491.321
SS625,14117,42618,9531.358
SS714,88412,58113,1380.686
SS88144782379570.196
Table 11. Power assessments of the LTB PTO damping tuning. Herein, a sensitivity study was also conducted for k w = 1–10. The upper bound of the maximum or mean P ¯ P T O of three algorithms in each sea state is in bold and underlined.
Table 11. Power assessments of the LTB PTO damping tuning. Herein, a sensitivity study was also conducted for k w = 1–10. The upper bound of the maximum or mean P ¯ P T O of three algorithms in each sea state is in bold and underlined.
Sea StateOverall Assessment
Maximum   P ¯ P T O (W) Minimum   P ¯ P T O (W) Mean   P ¯ P T O (W) | Λ P | ¯
SS111,79011,44611,6310.052
SS221,04819,77620,5130.107
SS311,17610,44510,7700.063
SS42091202320550.042
SS553,32950,27251,5880.130
SS627,87126,95227,3870.098
SS715,61114,64215,0170.149
SS88020770579020.102
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

Yue, X.; Zhang, J.; Meng, F.; Liu, J.; Chen, Q.; Geng, D. Multi-Timescale Lookup Table Based Maximum Power Point Tracking of an Inverse-Pendulum Wave Energy Converter: Power Assessments and Sensitivity Study. Energies 2023, 16, 6195. https://doi.org/10.3390/en16176195

AMA Style

Yue X, Zhang J, Meng F, Liu J, Chen Q, Geng D. Multi-Timescale Lookup Table Based Maximum Power Point Tracking of an Inverse-Pendulum Wave Energy Converter: Power Assessments and Sensitivity Study. Energies. 2023; 16(17):6195. https://doi.org/10.3390/en16176195

Chicago/Turabian Style

Yue, Xuhui, Jintao Zhang, Feifeng Meng, Jiaying Liu, Qijuan Chen, and Dazhou Geng. 2023. "Multi-Timescale Lookup Table Based Maximum Power Point Tracking of an Inverse-Pendulum Wave Energy Converter: Power Assessments and Sensitivity Study" Energies 16, no. 17: 6195. https://doi.org/10.3390/en16176195

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