Next Article in Journal
An Experimental Analysis of Soybean Straw Combustion on Both CO and NOX Emission Characteristics in a Tubular Furnace
Previous Article in Journal
Research on the Combined Control Strategy of Low Temperature Charging and Heating of Lithium-Ion Power Battery Based on Adaptive Fuzzy Control
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Techno-Economic Optimization of an Off-Grid Solar/Wind/Battery Hybrid System with a Novel Multi-Objective Differential Evolution Algorithm

1
School of Energy and Power Engineering, Nanjing Institute of Technology, Nanjing 211167, China
2
College of Water Conservancy and Hydropower Engineering, Hohai University, Nanjing 210098, China
*
Author to whom correspondence should be addressed.
Energies 2020, 13(7), 1585; https://doi.org/10.3390/en13071585
Submission received: 12 March 2020 / Revised: 26 March 2020 / Accepted: 27 March 2020 / Published: 1 April 2020
(This article belongs to the Section F: Electrical Engineering)

Abstract

:
Techno-economic optimization of a standalone solar/wind/battery hybrid system located in Xining, China, is the focus of this paper, and reliable and economic indicators are simultaneously employed to address the problem. To obtain a more precise Pareto set, a novel multi-objective differential evolution algorithm is proposed, where differential evolution with a parameter-adaptive mechanism is applied in the decomposition framework. The algorithm effectiveness is verified by performance comparisons on the benchmark test problems with two reference algorithms: a non-dominated sorting genetic algorithm and a multi-objective evolution algorithm based on decomposition. The applicability of the proposed algorithm for the capacity-optimization problem is also validated by comparisons with the same reference algorithms above, where the true Pareto set of the problem is approximated by combining of the three algorithms through the non-dominant relationship. The results show the proposed algorithm has the lowest inverted generational distance indicator and provides 85% of the true Pareto set. Analyses of the Pareto frontier show that it can produce significant economic benefits by reducing reliability requirements appropriately when loss of power supply probability is less than 0.5%. Furthermore, sensitivity analyses of the initial capital of wind turbine, photovoltaic panel and battery system are performed, and the results show that photovoltaic panel’s initial capital has the greatest impact on levelized cost of electricity, while the initial capital of wind turbine has the least impact.

Graphical Abstract

1. Introduction

To address the energy crisis and environmental pollution caused by the growth of the economy and society, hybrid renewable energy systems (HRES) have become popular and wise choices, especially for remote or island areas, where renewable energy is abundant, and where power electricity construction cost is high [1,2]. With their complementary characteristics and matured technologies, HRES consisting of photovoltaics (PV), wind turbines (WT) and battery systems (BS) have become one of the most popular power generation modes [3]. The optimal capacity of HRES is a key and complicated issue [4] because it always contains multiple optimization objectives and is influenced by many factors, such as characteristics of the energy source, technical specifications and environmental conditions [5].
Many researchers have been keen to address the capacity optimization problem. Scholars have applied mathematical programming to solve capacity optimization problems. Linear programming was utilized for capacity optimization of a standalone PV/WT HRES in [6], mixed-integer linear programming was applied for the design for an isolated PV/WT/BS/DG (diesel generator) HRES in [7] and dynamic programming was utilized for storage capacity optimization of PV/WT/BS HRES in [8]. There are also scholars solving the problem by iteration. An iterative algorithm [9] and an enumeration-based iterative algorithm [10] were implemented for component capacity optimization of micro-grids. However, the methods above have the drawbacks that they can only handle single objective optimization and are susceptible to falling into local optimum solutions [11].
Many other researchers utilized software tools to solve this problem. HOMER [12,13] is one of the most popular tools. It was used for capacity optimization, techno-economic and sensitivity analyses of a PV/BS HRES in [14] and a PV/DG/BS HRES in [15]. In [16], the techno-economic feasibility of replacing a completely DG power supply by a standalone PV/WT/BS/DG was analyzed by HOMER. In [17], the performance of a WT/BS/DG HRES impacted by different battery technologies was analyzed by HOMER. It was applied for the investigation of possible renewable power generation systems for Gadeokdo Island in South Korea in [18]. In [19], component capacities of PV/BS/DG HRES was optimized under different load profiles. In [20], the optimal sizing of fuel cell/PV HRES was performed by HOMER with the components cost calculated by fuzzy logic program. Homer is good software for scenario investigation, but some scenarios need to be re-calculated individually for the specific situation because this tool can only handle single-object optimization [21], the flexibility is limited, and is easy to fall into local optimum solutions [22].
Intelligent optimization algorithms, having the superiority of global optimization and the capacity to handle multi-objective problems easily, have been intensively studied in the optimal capacity problem [11]. In [23], an artificial bee colony algorithm was used for finding the optimal capacities of a PV-biomass HRES with the minimum levelized cost of electricity (LCOE), and the results proved it outperformed HOMER. The discrete bat search algorithm [24] and grasshopper optimization algorithm [25] were used for PV/WT/DS/BS HRES. Many researchers have considered multiple objective functions in capacity optimization, and have applied multi-objective intelligent optimization algorithms to obtain solutions. In [26], a multi-objective differential evolution algorithm was applied for the capacity optimization of a PV/WT/DG/BS HRES in Yanbu, where the objective functions were minimizing LCOE and loss of power supply probability (LPSP) simultaneously. A non-dominated sorting genetic algorithm (NSGA-II) was applied for optimal sizing of an off-grid PV/WT/BS HRES where economic and reliability indicators were simultaneously considered in [27]. Many scholars considered additional indicators, except for economic and reliability indicators; for example, an environmental indicator was involved in the optimal sizing of a PV/WT/fuel cell HRES in [28] and a PV/WT/BS HRES in [29].
With the rapid development of HRES, intelligent optimization algorithms with better performance are urgently needed [11]. Multi-objective evolution algorithms based on decomposition (MOEA/D) [30] provide a new approach to multi-objective optimization [31] and have received growing attention as they can incorporate the techniques used in single-objective optimization algorithms well. In this paper, a novel multi-objective optimization algorithm, namely MOEA/DADE, is proposed for better optimization performance. In this algorithm, a differential evolution mechanism with parameter self-adaptation is integrated in decomposition framework and its effectiveness is verified by algorithm contrasts with NSGA-II and MOEA/D on benchmark problems. Then, the algorithm is applied for techno-economic optimization of a standalone PV/WT/BS HRES located in Xining, China and its applicability for this problem is also validated by comparisons with NSGA-II and MOEA/D. Lastly, techno-economic and sensitivity analyses of the initial capital of wind turbines, photovoltaics and battery systems are performed.
The paper is organized as follows: component models of the HRES and energy management strategy (EMS) are introduced in Section 2, MOEA/DADE algorithm and algorithm comparisons with MOEA/D and NSGA-II on benchmark problems are presented in Section 3, techno-economic optimization and sensitivity analyses of PV/WT/BS HRES are performed in Section 4; finally, conclusions are drawn in Section 5.

2. Component Models and Energy Management Strategy

The HRES is consisted of five major components: PV systems, WT systems, a converter, BS and grid loads. Its structure diagram is shown in Figure 1.

2.1. Wind turbine

The relationship between a WT’s output power ( P WT ) and wind speed can be described by a piecewise function as Equation (1) [13]:
P WT ( t ) = { 0 v v cut - in   or   v v cut - out   P r ( v 3 v cut - in 3 v r 3 v cut - in 3 ) v cut - in v v r P r v r v v cut - out
where P r (kW) represents the WT’s rated output power, v (m/s) represents wind speed at the turbine hub height, v cut - in (m/s) is the cut-in speed, v r (m/s) is the rated speed, and v cut - out (m/s) is the cut-out speed. v cut - in , v r and v cut - out are assumed to be 2 (m/s), 9 (m/s) and 24 (m/s) respectively.
As the anemometer is not at the same altitude as the turbine hub, v can be obtained by Equation (2):
v = v ref × ( h h ref ) γ
where h (m), h ref (m) are the installation altitude of the turbine hub and the anemometer respectively, v ref (m/s) is the wind speed measured by anemometer, γ is a constant number between 0.1 and 0.25 [13]. In this paper, the anemometer height is 10 m, the hub height is 25 m and γ is assumed 0.25.

2.2. PV Panel

Ignoring the temperature effects, the PV’s output power ( P PV ) can be calculated by Equation (3) [13],
P PV ( t ) = Y PV f PV ( G t G STC )
where Y PV (kW) and G STC (1 kW/m2) represent the power output and solar radiation under standard test conditions, G t (kW/m2) represents the actual solar incident radiation on the PV array in time t, and f PV (%) is the PV derating factor.

2.3. Battery System

The electric power stored into the BS in time t is described as Equation (4):
P B ( t ) = P WT ( t ) + P PV ( t ) P L ( t ) / η inv
where P L (kW) represents the electrical load, η inv represents the converter efficiency.
If P B ( t ) is greater than 0, the BS will be charged to store the surplus power energy, otherwise, it will be discharged to make up for the electricity shortage. The BS’s state of charge (SOC) for t period is shown as Equation (5):
S O C B ( t ) = ( 1 σ ) S O C B ( t 1 ) + α ( t ) P B ( t ) Δ t
where S O C B (kW∙h) is the BS’s SOC, t represents time index, and σ is the dissipation coefficient. Assume that the charge efficiency ( η c ) and discharge efficiency ( η d ) remain unchanged, α ( t ) can be obtained by Equation (6).
α ( t ) = { η c P B ( t ) > 0   1 η d P B ( t ) < 0
Moreover, the BS’s SOC should satisfy the constraint described as Equation (7):
( 1 D O D ) E bmax S O C ( t ) E bmax
where E bmax (kW∙h) means the maximum allowable amounts of energy that can be stored by the BS, DOD (%) is the BS’s allowable depth of discharge.

2.4. Load Profile

To make the inherently statistical power load more realistic, the load in each time step is obtained by multiplying its annual average value with a perturbation factor k cv [32] shown as Equation (8):
k cv = 1 + α d + α t
where α d ~ N ( 0 , δ d 2 ) denotes the daily variation percent, α t ~ N ( 0 , δ t 2 ) denotes the hourly variation percent.

2.5. Economic Model

The life cycle cost (LCC) of the k-th component of the HRES is described as Equation (9). It includes initial capital cost ( I C k ), maintenance and operation cost ( O & M k ), replacement cost ( R p k ) and salvage value ( R V k ).
L C C k = I C k + O & M k + R p k R V k
I C k , O & M k , R p k , R V k can be calculated according to Equations (10)–(15), where C k (kW for WT, PV and converter, and kW∙h for BS) means the component capacity, γ k , θ k and π k mean initial capital cost, maintenance and operation cost, replacement cost per unit respectively and their units are $/kW for WT, PV and converter and $/kW∙h for BS, N (year) and Nk (year) mean the system life time of the system and the k-th component respectively, N k rem (year) means the surplus life of the k-th component when the system ends, R k rp (year) means the last replaced time of the k-th component, INT(.) is a function that returns the smallest integer that is greater than or equal to the input number, and the relationship of the real discount rate i r (%), nominal discount rate r (%) and expected inflation rate u (%) is shown as Equation (16):
I C k = γ k C k
O & M k = θ k C k j = 1 N ( 1 1 + i r ) j
R p k = π k C k m = 1 INT [ N N k ] 1 ( 1 1 + i r ) m × N k
R V k = π k C k N k rem N k ( 1 1 + i r ) N
N k rem = N k ( N R k rp )
R k rp = N k INT ( N N k )
i r = r u r + u
Capital recovery factor (CRF) shown as Equation (17) is applied to convert LCC into the annualized cost of the HRES. Assuming that the amount of electricity generated by the HRES per year stays the same over the project’s lifetime, LCOE is shown as Equation (18) [24]:
CRF ( i r , N ) = i r ( 1 + i r ) N ( 1 + i r ) N 1
LCOE = CRF ( i r , N ) k = 1 N C L C C k E load
where NC is the number of components in the HRES and E load (kW∙h) is the annual power output.

2.6. Rule-Based Energy Management Strategy

EMS is one main criterion for HRES [25]. To coordinate various components’ output power, a rule-based EMS is designed and its flow chart is shown as Figure 2.
Firstly, a binary variable (Sp) is defined to represent whether the electric power generated is sufficient, and a binary variable (Sc) is defined to represent whether the converter capacity is sufficient. Their definitions are shown as Equations (19) and (20).
S p = P PV ( t ) + P WT ( t ) P L ( t ) / η inv
S c = C inv P L ( t ) / η inv
Secondly, according to the combination values of Sp and Sc, the rule-based EMS is designed with four case scenarios as follows: Case 1,the electric power generated and the converter capacity are both sufficient ( S p > 0   and   S c > 0 ); Case 2, the electric power generated is sufficient while the converter capacity is insufficient ( S p > 0   and   S c < 0 ); Case 3, the electric power generated is insufficient while the converter capacity is sufficient ( S p < 0   and   S c > 0 ); and, Case 4, the electric power generated and the converter capacity are both insufficient ( S p < 0   and   S c < 0 ).
Finally, update the BS’s SOC and calculate the loss of power supply (LPS) for different cases according to the follow rules:
Case 1: the load will be completely satisfied (LPS = 0), and the extra power used for charging the battery is calculated by Equation (4) and a new temporary SOC of the BS (soc_new) can be obtained by Equation (5), if s o c _ n e w < E bmax , set s o c ( t + 1 ) = s o c _ n e w , otherwise, set s o c ( t + 1 ) = E bmax and the extra energy will be discarded.
Case 2: the power output of the HRES will be η inv C inv , LPS and the power used for charging the BS ( P B ) can be calculated by Equations (21) and (22) respectively.
L P S ( t ) = P L ( t ) C inv η inv
P B ( t ) = P PV ( t ) + P WT ( t ) C inv
then, update the BS’s SOC as Case 1.
Case 3: the battery will be discharged to supply as much electricity as possible and whether the load can be satisfied with the BS’s support is calculated according to Equation (23).
S bp = P PV ( t ) + P WT ( t ) + s o c ( t ) E bmin P L
If S bc > 0 , the load will be completely satisfied (LPS = 0), and the charge power of BS ( P B ) and BS’s SOC are calculated by Equations (4) and (5) respectively, otherwise, set s o c ( t + 1 ) = E bmin and calculate LPS by Equation (24).
L P S ( t ) = P L ( t ) ( P PV ( t ) + P WT ( t ) + s o c ( t ) E bmin ) η inv
Case 4: the load will not be completely satisfied and the battery will be discharged. Firstly, whether the power output can reach to converter capacity with the BS’s support is calculated according to Equation (25).
S bc = P PV ( t ) + P WT ( t ) + s o c ( t ) E bmin C inv
If S bc > 0 , the power output of the HRES will be η inv C inv , LPS is calculated by Equation (21), the charge power of BS ( P B ) and BS’s SOC can be calculated by Equation (22) and Equation (5) respectively, otherwise, set s o c ( t + 1 ) = E bmin and calculate LPS by Equation (24).

2.7. Objective Function and Constraints

The objective functions described as Equation (26) are to maximize system reliability (represented by minimizing LPSP) and economy (represented by minimizing LCOE) simultaneously,
f = min { LCOE ( C pv , C wt , E b , C inv ) ,      LPSP ( C pv , C wt , E b , C inv ) } 0 C pv C pv max 0 C wt C wt max 0 E b E b max 0 C inv C inv max
where C pv , C wt , C inv , and E b are decision variables that mean the capacity of PV, WT, converter and BS respectively, C pv max , C inv max and E b max are their upper bounds, and LPSP is calculated as Equation (27),
LPSP = t = 1 T L P S ( t ) t = 1 T P L ( t )
where t is a time period index, T (8760 h) represents the total hours of a year.

3. Optimization Algorithm

3.1. MOEA/DADE

MOEA/D whose detailed information can be found in [30] provides a new approach for multi-objective optimization. It has an advantage in that it can incorporate the techniques in single-objective optimization algorithms well, while its performance is greatly impacted by the new solution generation mechanism. For better optimization performance, an adaptive differential evolution mechanism is implemented in MOEA/D and a new algorithm, namely MOEA/DADE, is proposed in this paper.

3.1.1. Differential Evolution Mechanism

The differential evolution mechanism is applied for new individual’s generation as Equation (28),
y k = { x i ( t ) + F i ( x r 1 ( t ) x r 2 ( t ) ) rand C r i   or   j = k x i , j ( t ) otherwise    
where F i and C r i , in the range of [0,1], represent the scale factor and crossover rate respectively, r 1 and r 2 are randomly selected from set P. The definition of P is shown as Equation (29),
P = { B i if   rand < δ { 1 , , N } otherwise
where B i represents the neighbor set of the individual i and { 1 , , N } represents the collection of the all individuals, and δ ( 0 , 1 ) . Equation (29) means the individuals used for differential information generation can be selected from the whole individuals according to a tiny probability ( 1 δ ) and this is helpful for maintaining individual diversity.

3.1.2. Parameter Adaptive Mechanism

In differential evolution mechanism, the control parameters ( F i and C r i ) have great impact on the quality of new individuals, and their values should be different for different stages of the evolution process. Thus, a parameter adaptive mechanism is applied to the algorithm [33].
Firstly, at each generation, the control parameters ( F i and C r i ) of each individual are independently generated according to Equations (30) and (31), respectively,
C r i = randn i ( μ Cr , 0.1 )
F i = randc i ( μ F , 0.1 )
where randn i ( μ Cr , 0.1 ) means a standard normal distribution with mean and variance of μ Cr and 0.1, randc i ( μ F , 0.1 ) means a Cauchy distribution with mean and variance of μ F and 0.1. The control parameters will be reinitialized until they are in range of [0,1].
Secondly, record the control parameters that generate better individuals by set SCr and SF.
Lastly, update μ Cr and μ F at the end of each generation according to Equations (32) and (33),
μ Cr = ( 1 c ) μ Cr + c mean A ( S Cr )
μ F = ( 1 c ) μ F + c mean L ( S F )
where c is a weighting factor, mean A ( S Cr ) means the average value of the elements in set SCr and mean L ( S F ) is calculated by Equation (34).
mean L ( S F ) = F S F F 2 F S F F

3.2. Algorithm Contrast

To verify the effectiveness of MOEA/DADE, algorithm contrasts with MOEA/D and NSGA-Ⅱ are performed on PlatEMO v2.1 [34]. In the contrast test, the benchmark problems are ZDT1~ZDT4 and ZDT6 as proposed in [35], their dimensions (D) are 30 and objective function numbers are 2. Parameters for each algorithm are represented as Table 1, the maximum evaluation number for each algorithm are 10,000 and all three algorithms run 30 times. A comprehensive indicator IGD [36] defined as Equation (35), which can reflect both the diversity and convergency of the Pareto set (PS), is chosen to evaluate the algorithm performance.
IGD ( P * , P ) = x * P * min x P d ( x * , x ) | P * |
where P* represents sampling points from the true Pareto frontier (PF), P represents the PF obtained by the optimization algorithm, d ( x * , x ) represents the Euler distance between any two elements in P* and P, and | P * | denotes the number of elements of P*.
The test results are presented in Table 2. From the table, we can see that MOEA/DADE outperforms MOEA/DADE and NSGA-II on all the functions except for ZDT4. It indicates MOEA/DADE is a cost-effective multi-objective optimization algorithm for most of the considered test problems, but not for all considered functions. This behavior is not an exception but actually verifies the statement of No Free Lunch theorem. The reasons for superior performance may be concluded as two aspects: (1) the algorithm is implemented in the decomposition framework where evolution is achieved through the cooperation of neighboring individuals, and this mechanism is different from NSGA-II; (2) a differential evolution mechanism with parameter self-adaptation is applied whose effectiveness has been demonstrated in single-objective optimization.

4. Case Study

4.1. Data

The HRES studied in this paper is located in Xining ( 47 29 N ,   104 17 E ), China. Its nameplate life is 25 years. The technical and economical parameters are presented as Table 3 [37]. A year’s data of solar radiation and wind speed were obtained from HOMER Pro [13] and their annual average values are 4.67 (kW∙h/m2/day) and 6.63 (m/s) respectively. The load profile is obtained by Equation (8) where δ d and δ t are assumed 10% and 20% respectively. The data profile of load, solar radiation and wind speed of each hour are shown as Figure 3.

4.2. Techno-Economic Analysis

To validate the applicability of MOEA/DADE for the capacity optimization problem, comparisons with MOEA/D and NSGA-Ⅱ were performed. The processes of evaluating individuals of MOEA/DADE, MOEA/D and NSGA-Ⅱare the same and are shown in Figure 4.
The decision variables’ ranges are presented as C pv [ 0   500 ] , C inv [ 0 , 500 ] , E b [ 0 , 3600 ] . For this problem, its true PS and PF are unknown. To evaluate algorithm performance, a method combining of MOEA/DADE, MOEA/D and NSGA-Ⅱis proposed to approximate the true PS and PF. Firstly, all the three algorithms are run 20 times respectively to obtain their corresponding PSs; secondly, a temporary set is created by merging all the PSs; finally, the true PS and PF are obtained from the temporary set according to the non-dominant relationship.
The PFs obtained by each algorithm and the approximate true PF are represented as Figure 5. From it, we can see the PF obtained by MOEA/DADE almost coincides with the approximate true PF. There are 81 groups of unduplicated solutions in the approximately true PS, among them MOEA/DADE provides 67 groups (82.72%), NSGA-Ⅱ provides 13 groups (16.05%) and MOEA/D provides 1 group (1.23%).
IGD indicators calculated based on the approximate true PS are shown as Table 4. From it, we can see MOEA/DADE has the minimal IGD. From analysis of Figure 5 and Table 4, it can be concluded that MOEA/DADE is more suitable for this capacity optimization problem than NSGA-II and MOEA/D.
To analyze the relationship of the reliability (LPSP) and economy (LCOE), the slope of LCOE is calculated according to Equation (36).
s l o p e = Δ L C O E Δ L P S P
The slope of LCOE varying with LPSP is presented as Figure 6. From it, we can see LCOE and LPSP are negatively related, and the absolute value of the slope decreases gradually with the increase of LPSP when LPSP is less than 2%, and almost stays at a small value when LPSP is in the range of 2% to 5%. The absolute value of the slope decreases rapidly and falls from 0.1231 ($/kW∙h) to 0.0033 ($/kW∙h) when LPSP is less than 0.5%, slows down and stays in the range of [0.0110,0.0335] ($/kW∙h) when LPSP is in the range of [0.5%,2%], and is almost unchanged when LPSP is larger than 2% with a small average value 0.0081 ($/kW∙h). From Figure 6, it can be concluded that the economic benefits are significant by reducing reliability requirements when LPSP is less than 0.5%, and not obvious when LPSP is larger than 0.5%. For example, the system’s LCOE can fall from 0.2348 ($/kW∙h) to 0.2225 ($/kW∙h), falling by 0.0123 ($/kW∙h) as LPSP increases by 0.1% when LPSP is 0; however, it can only fall from 0.2041 ($/kW∙h) to 0.2008 ($/kW∙h), falling by 0.0033 ($/kW∙h) as LPSP increases by 0.1% when LPSP is 0.5%.

4.3. Sensitivity Analysis

Sensitivity analysis was performed to investigate the impact of initial capital of WT, PV and BS on the LCOE. The sensitivity analysis for any component is carried out by the following steps: (1) let its initial capital be 80%, 90%, 110% and 120% of its initial value respectively while keeping the other parameters unchanged; (2) solve the capacity optimization problem of the HRES by MOEA/DADE; (3) choose the solution with LPSP equal to 0.5% from Pareto set. The effects of components’ initial capital on LCOE when LPSP is 0.5% are presented as Figure 7. In Figure 7, the influence on LCOE can be reflected by the slope of the fold line, and the higher slope means a greater impact. From Figure 7, we can see that the PV’s line slope is higher than the others, while WT’s is lower than the others—that is to say, PV’s initial capital has the greatest effect while WT’s initial capital has the least effect on LCOE. When the initial capital of PV, BS and WT falls to 80% of its initial value, the LCOE with LPSP equals to 0.5% falls from 0.2348 ($/kW∙h) to 0.2255 ($/kW∙h) for the PV case (falling by 3.97%), to 0.2297 ($/kW∙h) for the BS case (falling by 2.19%) and to 0.2299 ($/kW∙h) for the WT case (falling by 2.07%). When components’ initial capital increases to 120% of its initial value, the LCOE can reach to 0.2490 ($/kW∙h) for the PV case (increasing by 6.05%), to 0.2460 ($/kW∙h) for the BS case (increasing by 4.77%) and to 0.2407 ($/kW∙h) for the WT case (increasing by 2.50%).

5. Conclusions

In this paper, we focused on the techno-economic optimization of a standalone PV/WT/BS HRES in Xining, China. To find out the optimal LCOE under different LPSP, a novel multi-objective optimization algorithm, namely MOEA/DADE, is proposed. In this algorithm, a differential evolution mechanism with parameter self-adaptation is implemented in the decomposition framework. Algorithm comparisons with NSGA-II and MOEA/D on benchmark problems verify that MOEA/DADE is superior to NSGA-II and MOEA/D. The applicability of MOEA/DADE on the capacity optimization problem was also validated by comparisons. Then, MOEA/DADE was applied for techno-economic and sensitivity analyses of the HRES. Techno-economic analyses from the PF shows the economic benefits are significant by reducing reliability requirements when LPSP is less than 0.5%, and are not obvious when LPSP is larger than 0.5%. The system’s LCOE can fall from 0.2348 ($/kW∙h) to 0.2225 ($/kW∙h), falling by 0.0123 ($/kW∙h) as LPSP increases by 0.1% when LPSP is 0, however, it can only fall from 0.2041 ($/kW∙h) to 0.2008 ($/kW∙h), falling by 0.0033 ($/kW∙h) as LPSP increases by 0.1% when LPSP is 0.5%. Sensitivity analyses for the components’ initial capital show PV’s initial capital has the greatest impact while WT’s initial capital has the least impact on LCOE. When the components’ initial capital falls to 80% of its initial value, the LCOE can fall by 3.97% for the PV case, 2.19% for the BS case and 2.07% for the WT case, and when components’ initial capital increases to 120% of its initial value, the LCOE can increase by 6.05% for the PV case, 4.77% for the BS case and 2.50% for the WT case. The results indicate that reducing the PV’s initial capital produces more obvious economic benefits.

Author Contributions

Conceptualization, Y.Y. and R.L.; data curation, R.L.; formal analysis, Y.Y. and R.L.; funding acquisition, Y.Y.; investigation, Y.Y.; methodology, Y.Y.; project administration, R.L.; software, Y.Y. and R.L.; validation, R.L.; visualization, Y.Y.; writing—original draft, R.L.; writing—review & editing, R.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Natural Science foundation of Jiangsu Province (Grant No. BK20181308), and the Fundamental Research Funds for the Central Universities (Grant No. 2018B45414).

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

DODBS’s allowable depth of discharge
f PV PV derating factor (%)
G t solar incident radiation on the PV (kW/m2)
G STC solar radiation under standard test conditions ( G STC = 1 kW/m2)
h turbine hub altitude (m)
h ref anemometer altitude (m)
I C initial capital cost ($/kW)
i r real discount rate (%)
k cv perturbation factor of load
O & M maintenance and operation cost ($/kW)
P L electrical load (kW)
P PV output power of PV (kW)
P r rated output power of WT (kW)
P WT output power of WT (kW)
r nominal discount rate (%)
R p replacement cost ($/kW)
R V salvage value($/kW)
S c a binary variable denoting whether the converter capacity is sufficient
S p a binary variable denoting whether the electric power generated is sufficient
S O C B SOC BS’s state of charge
uexpected inflation rate (%)
v wind speed at the turbine hub altitude (m/s)
v cut - in cut-in speed (m/s)
v cut - out cut-out speed (m/s)
v r nominal speed (m/s)
v ref wind speed measured by anemometer (m/s)
α d daily variation percent of load
α t hourly variation percent of load
η c charge efficiency of BS
η d discharge efficiency of BS
η inv converter efficiency

Abbreviation

BSbattery system
CRFcapital recovery factor
DGdiesel system
EMSenergy management strategy
HREShybrid renewable energy systems
LCClife cycle cost
LCOElevelized cost of electricity
LPSPloss of power supply probability
WTwind turbine
PFPareto frontier
PSPareto set

References

  1. Huang, Y.; Li, S.; Ding, P.; Zhang, Y.; Yang, K.; Zhang, W. Optimal Operation for Economic and Exergetic Objectives of a Multiple Energy Carrier System Considering Demand Response Program. Energies 2019, 12, 3995. [Google Scholar] [CrossRef] [Green Version]
  2. Yang, Y.; Guo, S.; Liu, D.; Li, R.; Chu, Y. Operation optimization strategy for wind-concentrated solar power hybrid power generation system. Energy Convers. Manag. 2018, 160, 243–250. [Google Scholar] [CrossRef]
  3. Kyritsis, A.; Voglitsis, D.; Papanikolaou, N.; Tselepis, S.; Christodoulou, C.; Gonos, I.; Kalogirou, S.A. Evolution of PV systems in Greece and review of applicable solutions for higher penetration levels. Renew. Energy 2017, 109, 487–499. [Google Scholar] [CrossRef]
  4. Anoune, K.; Bouya, M.; Astito, A.; Abdellah, A.B. Sizing methods and optimization techniques for PV-wind based hybrid renewable energy system: A review. Renew. Sustain. Energy Rev. 2018, 93, 652–673. [Google Scholar] [CrossRef]
  5. Zahraee, S.M.; Khalaji Assadi, M.; Saidur, R. Application of Artificial Intelligence Methods for Hybrid Energy System Optimization. Renew. Sustain. Energy Rev. 2016, 66, 617–630. [Google Scholar] [CrossRef]
  6. Nogueira, C.E.C.; Vidotto, M.L.; Niedzialkoski, R.K.; de Souza, S.N.M.; Chaves, L.I.; Edwiges, T.; dos Santos, D.B.; Werncke, I. Sizing and simulation of a photovoltaic-wind energy system using batteries, applied for a small rural property located in the south of Brazil. Renew. Sustain. Energy Rev. 2014, 29, 151–157. [Google Scholar] [CrossRef]
  7. Malheiro, A.; Castro, P.M.; Lima, R.M.; Estanqueiro, A. Integrated sizing and scheduling of wind/PV/diesel/battery isolated systems. Renew. Energy 2015, 83, 646–657. [Google Scholar] [CrossRef] [Green Version]
  8. Li, J.L.; Guo, B.Q.; Niu, M.; Xiu, X.Q.; Tian, L.T. Optimal Configuration Strategy of Energy Storage Capacity in Wind/PV/Storage Hybrid System. Trans. China Electrotech. Soc. 2018, 33, 1189–1196. [Google Scholar]
  9. Smaoui, M.; Abdelkafi, A.; Krichen, L. Optimal sizing of stand-alone photovoltaic/wind/hydrogen hybrid system supplying a desalination unit. Sol. Energy 2015, 120, 263–276. [Google Scholar] [CrossRef]
  10. Bhuiyan, F.A.; Yazdani, A.; Primak, S.L. Optimal sizing approach for islanded microgrids. IET Renew. Power Gener. 2015, 9, 166–175. [Google Scholar] [CrossRef]
  11. Al-falahi, M.D.A.; Jayasinghe, S.D.G.; Enshaei, H. A review on recent size optimization methodologies for standalone solar and wind hybrid renewable energy system. Energy Convers. Manag. 2017, 143, 252–274. [Google Scholar] [CrossRef]
  12. HOMER Pro Version 3.7 User Manual. Available online: https://www.homerenergy.com (accessed on 15 December 2019).
  13. Lambert, T.; Gilman, P.; Lilienthal, P. Micropower System Modeling with HOMER. 2006. Available online: http://homerenergy.com/documents/MicropowerSystemModelingWithHOMER.pdf (accessed on 15 December 2019).
  14. Shahzad, M.K.; Zahid, A.; ur Rashid, T.; Rehan, M.A.; Ali, M.; Ahmad, M. Techno-economic feasibility analysis of a solar-biomass off grid system for the electrification of remote rural areas in Pakistan using HOMER software. Renew. Energy 2017, 106, 264–273. [Google Scholar] [CrossRef]
  15. Halabi, L.M.; Mekhilef, S.; Olatomiwa, L.; Hazelton, J. Performance analysis of hybrid PV/diesel/battery system using HOMER: A case study Sabah, Malaysia. Energy Convers. Manag. 2017, 144, 322–339. [Google Scholar] [CrossRef]
  16. Hossain, M.; Mekhilef, S.; Olatomiwa, L. Performance evaluation of a stand-alone PV-wind-diesel-battery hybrid system feasible for a large resort center in South China Sea, Malaysia. Sustain. Cities Soc. 2017, 28, 358–366. [Google Scholar] [CrossRef]
  17. Li, C.; Zhou, D.; Wang, H.; Lu, Y.; Li, D. Techno-economic performance study of stand-alone wind/diesel/battery hybrid system with different battery technologies in the cold region of China. Energy 2020, 192, 116702. [Google Scholar] [CrossRef]
  18. Park, E.; Kwon, S.J. Towards a Sustainable Island: Independent optimal renewable power generation systems at Gadeokdo Island in South Korea. Sustain. Cities Soc. 2016, 23, 114–118. [Google Scholar] [CrossRef]
  19. Rajbongshi, R.; Borgohain, D.; Mahapatra, S. Optimization of PV-biomass-diesel and grid base hybrid energy systems for rural electrification by using HOMER. Energy 2017, 126, 461–474. [Google Scholar] [CrossRef]
  20. Singh, A.; Baredar, P.; Gupta, B. Techno-economic feasibility analysis of hydrogen fuel cell and solar photovoltaic hybrid renewable energy system for academic research building. Energy Convers. Manag. 2017, 145, 398–414. [Google Scholar] [CrossRef]
  21. Zahboune, H.; Zouggar, S.; Krajacic, G.; Varbanov, P.S.; Elhafyani, M.; Ziani, E. Optimal hybrid renewable energy design in autonomous system using Modified Electric System Cascade Analysis and Homer software. Energy Convers. Manag. 2016, 126, 909–922. [Google Scholar] [CrossRef]
  22. Sinha, S.; Chandel, S.S. Review of recent trends in optimization techniques for solar photovoltaic–wind based hybrid energy systems. Renew. Sustain. Energy Rev. 2015, 50, 755–769. [Google Scholar] [CrossRef]
  23. Singh, S.; Kaushik, S.C. Optimal sizing of grid integrated hybrid PV-biomass energy system using artificial bee colony algorithm. IET Renew. Power Gener. 2016, 10, 642–650. [Google Scholar] [CrossRef]
  24. Shivaie, M.; Mokhayeri, M.; Kiani-Moghaddam, M.; Ashouri-Zadeh, A. A reliability-constrained cost-effective model for optimal sizing of an autonomous hybrid solar/wind/diesel/battery energy system by a modified discrete bat search algorithm. Sol. Energy 2019, 189, 344–356. [Google Scholar] [CrossRef]
  25. Bukar, A.L.; Tan, C.W.; Lau, K.Y. Optimal sizing of an autonomous photovoltaic/wind/battery/diesel generator microgrid using grasshopper optimization algorithm. Sol. Energy 2019, 188, 685–696. [Google Scholar] [CrossRef]
  26. Ramli, M.A.M.; Bouchekara, H.R.E.H.; Alghamdi, A.S. Optimal sizing of PV/wind/diesel hybrid microgrid system using multi-objective self-adaptive differential evolution algorithm. Renew. Energy 2018, 121, 400–411. [Google Scholar] [CrossRef]
  27. Kamjoo, A.; Maheri, A.; Dizqah, A.M.; Putrus, G.A. Multi-objective design under uncertainties of hybrid renewable energy system using NSGA-II and chance constrained programming. Int. J. Electr. Power Energy Syst. 2016, 74, 187–194. [Google Scholar] [CrossRef]
  28. Nasiraghdam, H.; Jadid, S. Optimal hybrid PV/WT/FC sizing and distribution system reconfiguration using multi-objective artificial bee colony (MOABC) algorithm. Sol. Energy 2012, 86, 3057–3071. [Google Scholar] [CrossRef]
  29. Mazzeo, D.; Baglivo, C.; Matera, N.; Congedo, P.M.; Oliveti, G. A novel energy-economic-environmental multi-criteria decision-making in the optimization of a hybrid renewable system. Sustain. Cities Soc. 2020, 52, 101780. [Google Scholar] [CrossRef]
  30. Qingfu, Z.; Hui, L. MOEA/D: A Multiobjective Evolutionary Algorithm Based on Decomposition. IEEE Trans. Evol. Comput. 2007, 11, 712–731. [Google Scholar] [CrossRef]
  31. Xinye, C.; Yexing, L.; Zhun, F.; Qingfu, Z. An External Archive Guided Multiobjective Evolutionary Algorithm Based on Decomposition for Combinatorial Optimization. IEEE Trans. Evol. Comput. 2015, 19, 508–523. [Google Scholar] [CrossRef]
  32. Amrollahi, M.H.; Bathaee, S.M.T. Techno-economic optimization of hybrid photovoltaic/wind generation together with energy storage system in a stand-alone micro-grid subjected to demand response. Appl. Energy 2017, 202, 66–77. [Google Scholar] [CrossRef]
  33. Zhang, J.; Sanderson, A.C. JADE: Adaptive differential evolution with optional external archive. IEEE Trans. Evol. Comput. 2009, 13, 945–958. [Google Scholar] [CrossRef]
  34. Tian, Y.; Cheng, R.; Zhang, X.; Jin, Y. PlatEMO: A MATLAB Platform for Evolutionary Multi-Objective Optimization [Educational Forum]. IEEE Comput. Intell. Mag. 2017, 12, 73–87. [Google Scholar] [CrossRef] [Green Version]
  35. Zitzler, E.; Deb, K.; Thiele, L. Comparison of multiobjective evolutionary algorithms: Empirical results. Evol. Comput. 2000, 8, 173–195. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Coello, C.A.; Cortés, N.C. Solving multiobjective optimization problems using an artificial immune system. Genet. Program. Evolvable Mach. 2005, 6, 163–190. [Google Scholar] [CrossRef]
  37. 2019 ATB. Available online: https://atb.nrel.gov/electricity/2019/ (accessed on 12 March 2020).
Figure 1. Structure diagram of photovoltaic/wind turbine/battery system hybrid renewable energy system (PV/WT/BS HRES).
Figure 1. Structure diagram of photovoltaic/wind turbine/battery system hybrid renewable energy system (PV/WT/BS HRES).
Energies 13 01585 g001
Figure 2. The flow chart of the energy management strategy (EMS).
Figure 2. The flow chart of the energy management strategy (EMS).
Energies 13 01585 g002
Figure 3. The data profile of load, solar radiation and wind speed for each hour.
Figure 3. The data profile of load, solar radiation and wind speed for each hour.
Energies 13 01585 g003
Figure 4. The process of evaluating individuals of MOEA/DADE, MOEA/D and NSGA-Ⅱ.
Figure 4. The process of evaluating individuals of MOEA/DADE, MOEA/D and NSGA-Ⅱ.
Energies 13 01585 g004
Figure 5. The Pareto frontier (PF) of NSGA-II, MOEA/D, MOEA/DADE and the approximation.
Figure 5. The Pareto frontier (PF) of NSGA-II, MOEA/D, MOEA/DADE and the approximation.
Energies 13 01585 g005
Figure 6. The slope of minimum levelized cost of electricity (LCOE) varying with loss of power supply probability (LPSP).
Figure 6. The slope of minimum levelized cost of electricity (LCOE) varying with loss of power supply probability (LPSP).
Energies 13 01585 g006
Figure 7. The effects of components’ initial capital on LCOE.
Figure 7. The effects of components’ initial capital on LCOE.
Energies 13 01585 g007
Table 1. Parameters of NSGA-II, MOEA/D and MOEA/DADE.
Table 1. Parameters of NSGA-II, MOEA/D and MOEA/DADE.
AlgorithmParameters
NSGA-IIPc = 0.9, Pm = 1/D, ηc = 20, ηm = 20
MOEA/DT = 10, Pc = 0.9, Pm = 1/D, ηc = 20, ηm = 20
MOEA/DADET = 10, μ Cr = 0.8 , μ F = 0.5 , δ = 0.9
Table 2. Test results of NSGA-II, MOEA/D and MOEA/DADE with D = 30 for each function.
Table 2. Test results of NSGA-II, MOEA/D and MOEA/DADE with D = 30 for each function.
FunctionNSGA-II Mean (std)MOEA/D Mean (std)MOEA/DADE Mean (std)
ZDT11.7218 × 10−1 (1.10 × 10−1) −1.6737 × 10−1 (5.91 × 10−2) −4.7928 × 10−2 (1.62 × 10−2)
ZDT 25.5298 × 10−1 (1.01 × 10−1) −3.2507 × 10−1 (1.88 × 10−1) −1.2261 × 10−1 (1.63 × 10−1)
ZDT 31.3700 × 10−1 (8.06 × 10−2) −2.1748 × 10−1 (1.12 × 10−1) −6.6407 × 10−2 (3.45 × 10−2)
ZDT 41.4230 × 101 (4.08 × 100) +1.4189 × 101 (4.79 × 100) +3.8489 × 101 (9.88 × 100)
ZDT 63.8877 × 100 (3.78 × 10−1) −2.3380 × 100 (4.81 × 10−1) −1.8724 × 100 (3.83 × 10−1)
+/-/=1/4/01/4/0
Table 3. Technical and economical parameters of the HRES.
Table 3. Technical and economical parameters of the HRES.
FactorValue FactorValue
ProjectLifetime (year)25BatteryLifetime (year)10
Discount rate (%)6 Initial capital ($/kW∙h)160
Inflation rate (%)2 Replacement ($/kW∙h)128
PVLifetime (year)25 O&M ($/year/kW∙h)1
Initial capital ($/kW)1857 Round trip efficiency (%)80
Replacement ($/kW)1486ConverterLifetime (year)15
O&M ($/year/kW)18 Initial capital ($/kW)890
Wind TurbineLifetime (year)20 Replacement ($/kW)800
Initial capital ($/kW)1610 O&M ($/year/kW)10
Replacement ($/kW)1288 Efficiency (%)95
O&M ($/year/kW)32
Table 4. IGD indicators for NSGA-II, MOEA/D and MOEA/DADE.
Table 4. IGD indicators for NSGA-II, MOEA/D and MOEA/DADE.
AlgorithmNSGA-IIMOEA/DMOEA/DADE
IGD4.7809 × 10−41.3926 × 10−37.5655 × 10−5

Share and Cite

MDPI and ACS Style

Yang, Y.; Li, R. Techno-Economic Optimization of an Off-Grid Solar/Wind/Battery Hybrid System with a Novel Multi-Objective Differential Evolution Algorithm. Energies 2020, 13, 1585. https://doi.org/10.3390/en13071585

AMA Style

Yang Y, Li R. Techno-Economic Optimization of an Off-Grid Solar/Wind/Battery Hybrid System with a Novel Multi-Objective Differential Evolution Algorithm. Energies. 2020; 13(7):1585. https://doi.org/10.3390/en13071585

Chicago/Turabian Style

Yang, Yong, and Rong Li. 2020. "Techno-Economic Optimization of an Off-Grid Solar/Wind/Battery Hybrid System with a Novel Multi-Objective Differential Evolution Algorithm" Energies 13, no. 7: 1585. https://doi.org/10.3390/en13071585

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