Next Article in Journal
Ethical Financing in Europe—Non-Parametric Assessment of Efficiency
Previous Article in Journal
Regional Heterogeneity of Migrant Rent Affordability Stress in Urban China: A Comparison between Skilled and Unskilled Migrants at Prefecture Level and Above
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Forecasting Japan’s Solar Energy Consumption Using a Novel Incomplete Gamma Grey Model

1
School of Information and Software Engineering, University of Electronic Science and Technology of China, Chengdu 610054, China
2
School of Science, Southwest University of Science and Technology, Mianyang 621010, China
3
State Key Laboratory of Oil and Gas Reservoir Geology and Exploitation, Southwest Petroleum University, Chengdu 610500, China
*
Author to whom correspondence should be addressed.
Sustainability 2019, 11(21), 5921; https://doi.org/10.3390/su11215921
Submission received: 19 September 2019 / Revised: 19 October 2019 / Accepted: 20 October 2019 / Published: 24 October 2019

Abstract

:
Energy consumption is an essential basis for formulating energy policy and programming, especially in the transition of energy consumption structure in a country. Correct prediction of energy consumption can provide effective reference data for decision-makers and planners to achieve sustainable energy development. Grey prediction method is one of the most effective approaches to handle the problem with a small amount of historical data. However, there is still room to improve the prediction performance and enlarge the application fields of the traditional grey model. Nonlinear grey action quantity can effectively improve the performance of the grey prediction model. Therefore, this paper proposes a novel incomplete gamma grey model (IGGM) with a nonlinear grey input over time. The grey input of the IGGM model is a revised incomplete gamma function of time in which the nonlinear coefficient determines the performance of the IGGM model. The WOA algorithm is employed to seek for the optimal incomplete coefficient of the IGGM model. Then, the validations of IGGM are performed on four real-world datasets, and the results exhibit that the IGGM model has more advantages than the other state-of-the-art grey models. Finally, the IGGM model is applied to forecast Japan’s solar energy consumption in the next three years.

1. Introduction

Under the pressure of environmental pollution and the energy crisis, numerous countries have been devoted to the development of renewable energies such as hydro energy, wind energy, solar energy, geothermal energy, and bio-energy. In order to meet the people’s needs for energy in a sustainable way, the sustainable energy policy plays an important role in developing energies and adjusting the energy mix in a country. Accurately forecasting energy consumption contributes to formulating the sustained development strategy of energy. Nowadays, Japan’s energy structure has significantly changed. Before 2011, Japan mainly focused on developing nuclear energy. However, after the Fukushima nuclear accident, the nuclear energy generation declined rapidly because all the nuclear power plants were closed due to safety in 2012 [1]. Until 2015, several plants were gradually restarted for power generation. Nuclear energy consumption was also decreased due to the low production of nuclear energy. In order to alleviate the pressure of energy supply and reduce the dependence on gas and oil abroad at the same time, Japan has vigorously developed renewable energy sources after the nuclear accident. Meanwhile, the consumption of renewable energy is increasing rapidly due to its production. Renewable energy consumption has accounted for a larger proportion of total energy consumption compared with before the nuclear accident. In particular, Japan’s solar energy consumption is the fastest growing one in its renewable energy consumption, according to the BP Statistical Review of World Energy 2019. The consumption of solar energy in 2018 is 23 times that of 10 years ago. It deserves the attention of energy authorities or energy companies for making energy policy. Nevertheless, correctly predicting the trend of solar energy consumption is a challenging task. The current main prediction models include machine learning models and time series models. However, machine learning models usually need a lot of historical data as a training set used to train a model. The time series models with low demand for historical data are widely used to solve the forecasting problems. One of the most effective time-series prediction methods is the grey prediction model, which can be used to handle the uncertain problem with small samples effectively.
The classic grey model (GM(1,1)), initially designed by Professor Deng in 1982, is used to deal with univariate time series prediction problems. Due to the simplicity of modeling process and the lower demand about raw sequence, GM(1,1) and its extended models have been applied in various fields, such as natural gas consumption [2,3,4], electricity consumption [5,6,7], renewable energy consumption [8,9], tertiary industry [10], oil field production [11,12], carbon emissions [13,14,15], electricity and natural gas demand [16,17], and so on. Though grey prediction models have obtained much success in different fields, the prediction accuracy of grey models still requires to be increased by optimization of existing models. There are many optimization approaches about grey model presented as follows. Firstly, the optimization of background value is an effective approach to promote the prediction accuracy of the classical grey models. Li et al. [18] presented an optimized grey prediction model (TBGM(1,1)) based on the data transformation technique and optimization of grey input to predict the short-term energy consumption of Shanghai China. Zhang et al. [19] improved the classical GM(1,1) model by optimizing the coefficients and the time correlation factor of background values, and applied it to forecast the slope deformation field. Zeng et al. [20] optimized the existing multivariate grey model by dynamic background values and applied it to predict the number of China’s wireless communication users. Ye et al. [21] improved the classical Grey–Markov prediction model based on background-value optimization. Li et al. [22] proposed an optimized GM(1,1) model based on hybridization of background-value optimization and weighted weakening buffer operator. Atalay et al. [23] applied multiple linear regression, time series and grey models to estimate the electricity consumption of the heating, cooling, air conditioning system in a commercial building located in France. Rahman et al. [24] employed an improved GM(1,1) model to analyze and forecast the key performance indicators of a medical institution. Ma et al. [25] modified the classical GM(1,1) model based on the Simpson formula and employed it to forecast the gross domestic product (GDP) of Lanzhou, China.
On the other hand, optimization of initial conditions is also a practical approach. Wang et al. [26] optimized the initial condition of the time response sequence with the weighted sum of the first term and the last term of an accumulation generated sequence to improve the prediction precision of the classical GM(1,1) model. Wang et al. [27] proposed the nash nonlinear grey Bernoulli model (Nash NGBM(1,1)) model by optimization of the initial value in its time response sequence. Zeng et al. [28] proposed an unbiased grey model and optimized the initial value by employing the least-square method.
In addition, various accumulation generators are utilized to improve the traditional grey model. Wu et al. [29] initially introduced a fractional accumulation operator into grey prediction theory. Combining the fractional order differential theory and the fractional accumulating operator, Mao et al. [30] put forward a new fractional grey model that could conquer the disadvantages of the classical grey prediction model. Zeng et al. [28] proposed the NHGM(1,1,k) to deal with the non-homogeneous exponential time series based on a wakening buffer operator. With a weighted accumulation operator, Wu et al. [31] proposed the WDGM model that obtained better prediction precision than other models. Ma et al. [32] defined a comfortable fractional accumulation operator and proposed the comfortable fractional grey model (CFGM) that gained more advantages in fluctuating time series.
Moreover, the classical grey model is generalized to deal with multivariate time series. Tien et al. [33] proposed the GMC(1,n) model which corrected the traditional GM(1,n) model designed by Professor Deng et al. Ding et al. [14] designed a new GM(1,n) model based on the trends of system inputs and background value optimization to predict the carbon emissions from fossil fuel of China. Wu et al. [34] proposed a multivariate grey prediction model with a weakening buffer operator. Pei et al. [35] put forward the TNGM(1,n) model based on a nonlinear least-square approach, which had better performance than the traditional GM(1,n) model. Wu et al. [36] designed the GMCN(1,n) model with the new information priority accumulating operator.
Based on the new information priority operator, Ding et al. [12] modified the classical GM(1,1) to forecast the industrial electricity consumption of China. Xia et al. [37] proposed the new information priority accumulated grey model with time power to forecast the wind turbine capacity. The hybrids of the grey model and other classical approaches are also important methods. Al-shanini et al. [38] proposed the hybrid grey model with the Bayesian network to assist the risk management in the chemical process industries. Another optimization measure is the optimization of grey action quantity which is also an important measure to boost the prediction precision of the classical grey models. Zeng et al. [39] proposed an optimized GM(1, n) model by adding a constant term and a linear term over time in the grey input of a traditional GM(1, n) model. The results showed that the proposed grey model obtained the best precision in predicting a material’s tensile strength compared with other classical models. Yin et al. [40] presented a grey model with an exponential grey quantity of time (EOGM(1,1)) model and predicted the GDP of China’s tertiary industry. Wu et al. [41] improved the classical fractional grey model with a linear grey input over time and forecasted the nuclear energy consumption in China. Ma et al. [42] presented an optimized fractional grey prediction method with a nonlinear time-delayed grey input and forecasted the coal and gas consumption. Zeng et al. [43] proposed a new multivariable grey prediction model (NMGM(1,n)) with structure compatibility by optimizing the grey input of the multivariate grey model. Wang et al. [44] proposed an optimized multivariate grey model based on the background values of all variables in the system and employed it to forecast China’s industrial energy consumption. In addition, many scholars optimized classical grey prediction model with kernel technique [45,46,47], rolling mechanism [48,49], intelligent heuristic algorithm [42,50,51,52,53], data preprocess method [10,54,55], regularization method [56], hybridization of the first principles and the data driven models [57], parameter optimization [58], and so on.
In this study, a novel nonlinear grey input is employed to optimize the traditional grey prediction model. Compared with the nonlinear grey action quantities of the existing grey models [40,42], the proposed grey incomplete gamma input has more complicated curves and nonlinear properties. The incomplete gamma grey model (IGGM) can be utilized to manifest the non-exponential properties of time series. Therefore, one of the highlights is that a novel incomplete gamma grey model is proposed based on the WOA algorithm. The validation results prove that IGGM model obtains better performance than the other existing grey model. The primary function of WOA is to seek for the optimal value of the nonlinear coefficient of grey action quantity. Another highlight is that IGGM model is applied to forecast Japan’s solar energy consumption in the next three years.
The remainders of this paper are organized as follows. The classical GM(1,1) model and its improved grey models based on optimization of grey input are presented in Section 2. Then, a novel incomplete gamma grey prediction model named the IGGM model with a nonlinear grey input is minutely illustrated in Section 3. The method of determining the optimal nonlinear parameter of grey input is described in Section 4. Then, the validation experiments performed on the real-life data sequences and analysis are demonstrated in Section 5. In Section 6, the application in forecasting Japan’s solar energy consumption is stated, and the conclusions are drawn in the last section.

2. Basic GM(1,1) and Its Extended Models with Optimization of Grey Action Quantity

Grey prediction theory is one of the most effective approaches to deal with prediction problems with small samples. For a given data sequence χ = x ( 1 ) , x ( 2 ) , , x ( ) , the classical grey prediction model with one variable and first order accumulated generating operator (GM(1,1)) can be represented as
x ( 1 ) ( k ) x ( 1 ) ( k 1 ) + a z ( 1 ) ( k ) = b ,
where x ( 1 ) ( k ) = i = 1 k x ( i ) denotes the first-order accumulation generating value at kth point, z ( 1 ) ( k ) = x ( 1 ) ( k ) + x ( 1 ) ( k 1 ) × 0 . 5 denotes the background value.
The corresponding whitening equation of the classical grey prediction model is defined as
d x ( 1 ) ( t ) d t + a x ( 1 ) ( t ) = b ,
where a and b are development coefficient and grey action quantity, respectively. The core issue is to estimate the parameters a and b of the grey model. By employing the least square (LS) method, the parameters a and b can be calculated as follows:
[ a , b ] T = A A T 1 A T Y ,
where
A = 0 . 5 x ( 1 ) ( 2 ) + x ( 1 ) ( 1 ) 0 . 5 x ( 1 ) ( 3 ) + x ( 1 ) ( 2 ) 0 . 5 x ( 1 ) ( ) + x ( 1 ) ( 1 ) 1 1 1 T
and
Y = x ( 1 ) ( 2 ) x ( 1 ) ( 1 ) x ( 1 ) ( 3 ) x ( 1 ) ( 2 ) x ( 1 ) ( ) x ( 1 ) ( 1 ) T .
After the two parameters are determined, by solving the whitening Equation (2) and substituting the initial condition into the solution of Equation (2), the time response function of GM(1,1) can be represented as follows:
x ( 1 ) ( m + 1 ) = x ( 0 ) ( 1 ) b a e a m + b a
By using inverse accumulation operator, the stored value of GM(1,1) can be calculated as:
x ^ ( m + 1 ) = x ( 0 ) ( 1 ) b a ( 1 e a ) e a m .
The stored value x ^ ( m ) denotes the predicted value.
Though the classical GM(1,1) model has been applied in many fields and obtained numerous achievements in solving prediction problems with poor information, GM(1,1) still requires improvement to promote the prediction accuracy and enlarge the application range. Optimization of grey input is one of the most important measures to increase the accuracy of classical grey model. Many studies have been performed by improving the grey input of grey model. When the term b k takes the place of the grey input of GM(1,1), NGM(1,1,k) model [59] can be obtained. When the term b k + c takes the place of the grey input of GM(1,1), the SAIGM model [60] can be obtained. When the term b t α + c takes the place of the grey input of GM(1,1), the whitening GM(1,1, t α ) model [61] can be obtained. When the term b e a t replaces the grey input of GM(1,1), the whitening equation of EOGM model [40] can be obtained. When the term i = 0 h b i k i , ( h 1 ) replaces the grey input of GM(1,1), the FOTP-GM model [62] can be obtained. When the term b x ( 1 ) ( t ) n replaces the grey input of GM(1,1), the NGBM model [63] can be obtained. When the term b x ( 1 ) ( t ) 2 replaces the grey input of GM(1,1), the GVM model [64] can be obtained.

3. Incomplete Gamma Grey Model

In this section, a novel incomplete gamma grey model called an IGGM model is proposed, in which the grey action quantity of traditional GM(1,1) is taken to replace a new grey input driven by incomplete gamma function.

3.1. IGGM Model

In classical grey prediction theory, the grey action quantity of GM(1,1) model is a constant value. This denotes that the grey input of grey system is stable as time changes. However, the input of the grey system usually changes over time in the real world. Some scholars believe that grey action quantity of the grey system varies linearly with time. The typical grey models such as NGM and SAIGM are proposed to promote the prediction accuracy of classical GM(1,1) model. The grey input of NGM is the term b k . The grey input of SAIGM is the term b k + c . Moreover, some other scholars deem that the grey input of grey model varies nonlinearly over time. The nonlinear term t α is considered the grey action quantity of GM(1,1, t α ) model. The nonlinear term e α t is considered as the grey input of EOGM(1,1) model. Figure 1 shows the properties of these two grey action quantities. These two existing grey models with nonlinear grey action quantity exhibit better performance and applicability compared with the other models. In this section, the incomplete gamma function as the grey input is brought into a grey prediction model. The incomplete gamma function has superior nonlinear characteristics which can enrich the state of grey input at different times and cater more data sequences with different features.
From Figure 2, it can be noticed that the properties of incomplete gamma function are significantly different from those of the above-mentioned linear and nonlinear grey input. The classic grey model can be improved with grey input based on an incomplete gamma function. The definition of the improved grey model is described as follows:
The differential equation
d x ( 1 ) ( t ) d t + a x ( 1 ) ( t ) = b γ ( μ , t ) + c
is defined as a whitening equation of the incomplete gamma grey prediction model called IGGM. The parameter a is the development coefficient of the IGGM model. The term b γ ( μ , t ) + c is a grey input of the IGGM model. The function γ ( μ , t ) is called an incomplete gamma function and is defined as:
γ ( μ , t ) = 0 t t μ 1 e τ d τ ,
where μ is a tunable parameter called an incomplete coefficient, and μ > 0 . The parameter μ determines the property of the incomplete gamma function and profoundly affects the performance of the proposed IGGM model. Through integrating both sides of Equation (9) within interval [ k 1 , k ] , the discrete differential form of the IGGM model can be represented as:
x ( 0 ) ( k ) + a z ( 1 ) ( k ) = 0.5 b γ ( μ , k ) + γ ( μ , k 1 ) + c ,
where γ ( μ , k ) = k μ e k m = 0 Γ ( μ ) k m Γ ( μ + m + 1 ) = k μ e k 1 μ + k μ ( μ + 1 ) + k 2 μ ( μ + 1 ) ( μ + 2 ) + .
When the parameter b is set to 0, the IGGM model can be degenerated into the classical GM(1,1) model. When the incomplete coefficient μ is set to 1, the incomplete gamma function can be represented as γ ( 1 , k ) = 1 e k . In addition, then the IGGM model can be reduced into a kind of special EOGM(1,1) model with the form d x ( 1 ) ( t ) d t + a x ( 1 ) ( t ) = b e a t in which the development coefficient a is equal to 1.

3.2. Estimating Linear Parameter of IGGM Model

In the IGGM model, there are four parameters a , b , c , and μ which should be estimated. The parameters a , b , and c are linear parameters which can be estimated as similar to the traditional GM(1,1) model by using the LS method. However, the coefficient μ is a nonlinear parameter which can be optimized by an intelligent algorithm introduced in Section 4. Assuming that the parameter μ is given, the parameters [ a , b , c ] T can be estimated by the LS method and computed as:
[ a , b , c ] T = ( A T A ) 1 A T Y ,
where
A = z ( 1 ) ( 2 ) 0 . 5 γ ( μ , 2 ) + γ ( μ , 1 ) 1 z ( 1 ) ( 3 ) 0 . 5 γ ( μ , 3 ) + γ ( μ , 2 ) 1 z ( 1 ) ( ) 0 . 5 γ ( μ , ) + γ ( μ , 1 ) 1 , Y = x ( 0 ) ( 2 ) x ( 0 ) ( 3 ) x ( 0 ) ( ) ,
in which is the amount of modeling samples.

3.3. Time Response Function and Restored Values

After estimating the optimal parameters of the IGGM model, we can solve Equation (8) and obtain its solution as the time response function of the grey model. Multiplying both sides of Equation (8) by e a t , we have
e a t d x ( 1 ) ( t ) d t + a x ( 1 ) ( t ) = e a t b γ ( μ , t ) + c .
Rearrange Equation (13) and obtain
e a t d x ( 1 ) ( t ) d t + x ( 1 ) ( t ) d e a t d t = e a t ( b γ ( μ , t ) + c ) .
Then, we have
d e a t x ( 1 ) ( t ) d t = e a t b γ ( μ , t ) + c .
Integrating both sides of Equation (15) within interval [ 1 , t ] , the equation can be obtained as
e a t x ( 1 ) ( t ) t = 1 t = t = 1 t e a x b γ ( μ , x ) + c d x .
Substituting the initial condition x ^ ( 1 ) ( 1 ) = x ( 0 ) ( 1 ) , we have
e a t x ( 1 ) ( t ) e a x ( 0 ) ( 1 ) = 1 t e a x b γ ( μ , x ) + c d x .
Rearrange Equation (17) and obtain
x ( 1 ) ( t ) = e a t e a x ( 0 ) ( 1 ) + 1 t e a x b γ ( μ , x ) + c d x .
Then, we have
x ( 1 ) ( t ) = x ( 1 ) ( 1 ) e a ( t 1 ) + 1 t e a ( t x ) b γ ( μ , x ) + c d x .
Using the Gaussian formula, the convolution integral term of Equation (19) can be discrete as:
1 t e a ( t x ) b γ ( μ , x ) + c d x = m = 2 k 1 2 e a ( k m + 0 . 5 ) f ( m ) + f ( m 1 ) ,
where f ( t ) = b γ ( μ , t ) + c . Substituting Equation (20) into Equation (19), the time response function of the IGGM model can be represented as:
x ^ C ( 1 ) ( k ) = x ( 0 ) ( 1 ) e a ( k 1 ) + m = 2 k 1 2 e a ( k m + 0 . 5 ) f ( m ) + f ( m 1 ) .
By using inverse first order accumulated operation, the restored values of the IGGM model can be calculated as follows:
x ^ C ( 0 ) ( k ) = x ^ C ( 1 ) ( k ) x ^ C ( 1 ) ( k 1 ) ,
where k = 2 , 3 , , , r + p , in which is the total amount of fitted values while p is the number of predicted values.

4. Determining the Nonlinear Parameter of an Incomplete Gamma Grey Model

4.1. Constructing Optimization Problem for Searching the Optimal Nonlinear Parameter

In Section 3, the methodology of linear parameter estimation of the IGGM model has been introduced based on the hypothesis of given nonlinear parameter μ . In fact, the parameter μ profoundly determines the accuracy of the two proposed grey models. In order to determine the optimal value of μ , a constrained optimization problem based on the above-mentioned process of building the IGGM model is constructed. The objective function of this optimization problem is to minimize the fit error during the simulation stage. Usually, the fit error can be calculated with mean absolute percentage error. The constraint of the optimization problem is a serial of equality constraint derived from the process of modeling in Section 3. Mathematically, the optimization problem is represented as follows:
min E r r o r ( μ ) = 1 1 m = 2 x ^ C ( 0 ) ( m ) x ( 0 ) ( m ) x ( 0 ) ( m ) s . t . [ a , b , c ] T = ( A T A ) 1 A T Y A u = z ( 1 ) ( 2 ) 0 . 5 γ ( μ , 2 ) + γ ( μ , 1 ) 1 z ( 1 ) ( 3 ) 0 . 5 γ ( μ , 3 ) + γ ( μ , 2 ) 1 z ( 1 ) ( ) 0 . 5 γ ( μ , ) + γ ( μ , 1 ) 1 , Y = x ( 0 ) ( 2 ) x ( 0 ) ( 3 ) x ( 0 ) ( ) x ^ C ( 1 ) ( m ) = x ( 0 ) ( 1 ) e a ( m 1 ) + i = 2 m 1 2 e a ( m i + 0 . 5 ) f ( i ) + f ( i 1 ) x ^ C ( 0 ) ( m ) = x ^ C ( 1 ) ( m ) x ^ C ( 1 ) ( m 1 ) , m = 2 , 3 , , γ ( μ , m ) = m μ e k i = 0 Γ ( μ ) m i Γ ( μ + i + 1 ) f ( m ) = γ ( μ , m ) + c ,
where denotes the number of modeling samples. Obviously, it can be found that problem (23) is a nonlinear programming problem with an equality constraint. It is difficult to solve such problem (23) exactly because of the complexity of its objective function and constraint. Some scholars have constructed similar optimization problems to search for the optimal parameters. For example, Wang et al. [44] established an optimization problem with constraint to search for the generation coefficient λ i of background values. Mao et al. [30] employed the similar measure to choose the optimal order r of the accumulated operator and the order q of differential equations. These studies prove that it is an effective measure to search for optimal parameters to enhance the accuracy of their proposed model.

4.2. Searching Optimal Nonlinear Parameter by the Whale Optimization Algorithm

From the expression of problem (23), it is a nonlinear programming problem with constraints. Obviously, it can be solved by employing a heuristic search algorithm. The Whale Optimization Algorithm (WOA) firstly proposed by Mirjalili et al. in 2016 [65] is employed to search for the optimal value of the incomplete coefficient. The WOA algorithm has been widely employed to settle the optimization problems such as classification, feature extraction, image segment, and so on [66,67]. The inspiration of WOA is to mimic the predation behavior of humpback group. When humpback whales prey on the fish, they encircle the prey and spirally move toward the prey synchronously. Mathematically, the predation behavior of humpback group can be formulated as:
L ( κ + 1 ) (24a) L * ( κ ) ( 2 r f ( κ ) f ( κ ) ) 2 r L * ( κ ) L ( κ ) , p r o b < 0 . 5 , (24b) cos ( 2 π τ ) e β τ L * ( κ ) L ( κ ) + L * ( κ ) , p r o b 0 . 5 ,
where L ( κ ) denotes the location of the humpback at time κ , L * ( κ ) denotes the optimal location of humpback at time κ , the coefficient r is randomly generated in interval [ 0 , 1 ] , the coefficient τ is a random value in [ 1 , 1 ] , the coefficient β is a constant, p r o b is a probability of choosing behavior, and f ( κ ) = 2 2 κ / T at time κ . When p r o b < 0.5 , WOA mimics the encircling behavior formulated by Equation (24a). When p r o b 0.5 , WOA mimics the spiral movement behavior represented by Equation (24b). In order to enhance the global search capability of WOA, the location of a randomly chosen humpback replaces the best location of whale when the norm of ( 2 r f ( κ ) f ( κ ) ) is greater than 1. Mathematically, it can be represented as:
L ( κ + 1 ) = L r a n d ( κ ) ( 2 r f ( κ ) f ( κ ) ) 2 r L r a n d ( κ ) L ( κ ) ,
where L r a n d ( κ ) denotes the location of a randomly chosen humpback at time κ . In order to employ WOA to seek for the optimal incomplete coefficient, the necessary parameters are set as follows: the maximum number of iteration T is 300, the size of humpback group is set to 30, the lower bound of the designed parameter is 0, and the upper bound of the designed parameter is set to 10. WOA was initially used to solve the unconstraint optimization problem. Thus, WOA should be revised to solve problem (23) with complex equality constraints. The main computational steps of revised WOA algorithm are summarized as follows:
Step 1: Initialize each humpback’s location L ( κ ) randomly at κ = 1 in search space [ 0 , 10 ] .
Step 2: Calculate the fitness values of humpback group. According to the objective function of Equation (23), the fitness function can be defined as:
f i t n e s s κ = 1 1 m = 2 x ^ C ( 0 ) ( m ) x ( 0 ) ( m ) x ( 0 ) ( m ) .
Then, the location of a humpback that has the lowest fitness value is considered as the optimal location L * ( κ ) at κ st iteration.
Step 3: Update the locations of all humpbacks. When p r o b 0.5 , each humpback’s location is updated by using Equation (24b). When p r o b < 0.5 and 2 r f ( κ ) f ( κ ) < 1 , the location of each humpback is updated by using Equation (25).The location of each humpback is updated by using Equation (24a) in the other condition. Then, each humpback’s fitness is recalculated according to the new locations of humpbacks.
Step 4: If there is a better location that generated a lower fitness than the fitness of last iteration, the best location L * ( κ + 1 ) is updated.
Step 5: The procedure iterates and continues from Step 3 until the stop criterion is reached. The stop criterion is that the maximum number of iterations has been reached.

4.3. Computational Procedure of the Grey Model

In order to utilize the IGGM model to forecast the future, the key issue is to build an IGGM model with optimal parameters. Firstly, the nonlinear parameter μ can be sought out by solving the optimization problem (23). Then, the linear parameters a , b , and c can be estimated by the LS method after determining the coefficient μ . Finally, the future values can be predicted by using the time response function of IGGM with optimal parameters. The overall procedure of the IGGM model can be summarized in Figure 3.

5. Validation of an Incomplete Gamma Grey Model

5.1. Evaluation Metrics and Benchmark Grey Models

In order to evaluate the performance of the proposed grey models, two common performance metrics Absolute Percentage Error (APE) [68] and Mean Absolute Percentage Error (MAPE) [69] are adopted. The formula of APE is represented as:
A P E ( m ) = x ( 0 ) ( m ) x C ( 0 ) ( m ) x ( 0 ) ( m ) × 100 % , m = 1 , 2 , , , , + p ,
where refers to the number of samples for building the model while the other p samples are used to test the model; x ( 0 ) ( m ) and x p ( 0 ) ( m ) respectively denote the actual value and the produced value by grey models at the period of m. Lower APE indicates that the predicted value by the grey model is more approximate to the actual value. The evaluation criterion MAPE is usually used to evaluate the performance of prediction model. Mathematically, the MAPE of fitting is defined as:
M A P E f i t t i n g = 1 i = 1 A P E ( i ) .
Meanwhile, MAPE can also be used to evaluate prediction performance of the prediction method. MAPE for prediction is defined as:
M A P E p r e d i c t i o n = 1 p i = + 1 + p A P E ( i ) .
In addition, overall MAPE of fitting and prediction is defined as:
M A P E o v e r a l l = 1 + p i = 1 + p A P E ( i ) .
Lower MAPE indicates that the prediction model has better performance. To reveal the prediction advantages of the IGGM model, six state-of-the-art grey models including GM(1,1) [70], ARGM [2], NGM [59], DGM [71], SAIGM [60], and the GVM [64] model are used as baseline prediction models that are applied in validation cases and application compared with the two proposed grey models.

5.2. Example A: Forecasting the Average Daily Electricity Consumption of China

In this validation study, the raw sequence is China’s average daily electricity consumption (ADEC) from 2006 to 2016 tabulated in Table 1. The raw consumptions are collected from the National Bureau of Statistics of China. The ADEC from 2006 to 2013 are used to build the model while the rest of the raw data are used to examine the prediction performance of the grey model. The proposed IGGM model and the other classic grey models are applied to predict ADEC of China. The linear parameters of grey models are directly estimated by using the LS method. For the IGGM model, the optimal value of incomplete coefficient μ is firstly sought out by the WOA algorithm. The optimal incomplete coefficient of the IGGM model is 5.7157 . Then, the linear parameters are calculated based on the optimal coefficient μ . The results of all grey models are listed in Table 2. The APE and MAPE of all grey models are listed in Table 3. Fitting MAPE of IGGM is 0.7079 . Prediction MAPE of the IGGM model is 0.9664 . From Figure 4, it can be obviously noticed that the IGGM model achieves the best prediction performance and obtains the best fit compared with the other six classic grey models. Above all, our proposed IGGM model is superior to the other six popular grey models in Example A.

5.3. Example B: Forecasting the Nuclear Energy Consumption of China

In this subsection, the nuclear energy consumption (NEC) of China collected from the BP Statistical Review of World Energy (BPSRWE) 2019 are considered as a raw sequence to validate the performance of our proposed grey models. The raw data listed in Table 4 are divided into two groups. The first one, including the first nine digits, is regarded as a training set to build the model for the proposed grey models and the other six state-of-the-art existing grey models. The second group, including the last three digits, is used to examine the prediction performance of all grey models. By using an intelligent algorithm WOA, the optimal incomplete coefficient of IGGM is sought out. Its optimal value is 3.0101 . Then, all linear parameters of grey models are estimated by the LS method. The results of all grey models are placed in Table 5 To evaluate the fit and prediction performance, the APE of each period and MAPEs of fitting and predicting for all models are computed, which are tabulated in Table 6. The fit MAPE of SAIGM model is the lowest among the seven grey models while its MAPE of prediction reaches 18.2042%. It indicates that the overfitting phenomenon appears in the modeling process of SIAGM. Though IGGM model’s MAPE of fitting are not best, its overall MAPE and MAPE of prediction are smaller than the six other contrastive grey models. From Figure 5 and Table 6, it can be apparently found that the IGGM model obtains the lowest MAPE of prediction. In conclusion, our proposed IGGM prediction model significantly outperforms the other six grey models in this validation.

5.4. Example C: Forecasting the Hydro Electricity Consumption of China

The raw data are collected form BPSRWE 2019, which is tabulated in Table 7. In order to evaluate the grey models, the raw data are partitioned into two sets as training set and test set. The training set, including the first ten samples, is used to build a model for different grey models, respectively. The test set, including the other two samples, is used to validate the prediction performance of these grey models. The nonlinear parameter μ of IGGM is searched by WOA. Its optimum value is 7.6901 . In addition, the linear parameters of all models are easily determined by the LS method. The fitted and predicted values of different grey models are tabulated in Table 8. The errors of fitting and forecasting during the modeling stage and forecasting stage are placed in Table 9. The IGGM model obtains the best fitting performance and the best prediction performance which are 2.5863 and 1.1741 , respectively. From Figure 6 and Table 9, it can be apparently found that the proposed model outperforms the other six grey models.

5.5. Example D: Forecasting the Oil Production of Oil Fields

In this section, the raw data are obtained from the paper [11] and listed in Table 10. The first eight samples are used to build model while the other four samples are used to validate the prediction performance of the proposed model. Similar to the previous three validation experiments, the incomplete coefficient of IGGM is sought out by the WOA algorithm. In addition, then the linear parameters of all grey models are estimated by the LS method. The optimal value of the incomplete coefficient is 2.8614 . The fitted and predicted results of different grey models are shown in Table 11. The APE and MAPE of grey models are tabulated in Table 12. Though the fitting MAPE of IGGM is larger than that of the ARGM and SAIGM model, the IGGM model’s MAPE of prediction is lower than that of the two models. From Table 12 and Figure 7, it can be noticed that the prediction performance of the IGGM model is superior to the other six grey models in this validation experiment.

5.6. Analysis and Discussion

From the results of four validation experiments, it can be noticed that the proposed IGGM model has advantages over the other six classical grey models. In Example A, the MAPEs of GM, ARGM, DGM, SAIGM, and IGGM are less than 2% in terms of data fit while the MAPEs of GM, DGM, and SAIGM are larger than 10% in terms of prediction. It indicates that the over-fitting problem exists in modeling processes of GM, DGM, and SAIGM. The MAPE of fitting and prediction of GVM model are much larger than 10%. Though the prediction abilities ARGM and IGGM are excellent according to Lewis’ criterion, the MAPE of IGGM is significantly lower than that of the ARGM model. In Example B, only fit errors of SAIGM and IGGM model are apparently lower than those of other models and close to each other. However, the prediction’s MAPE of SAIGM is 18.2042%. It denotes that the over-fitting phenomenon occurs in the SAIGM model. In Example C, the MAPEs of fitting in GM, ARGM, DGM, SAIGM, and IGGM are relatively low and very close to each other. However, the prediction’s MAPE of GM, DGM, and SAIGM is markedly higher than 10%. Similarly, there are over-fitting problems in such three models. Only the IGGM model achieves excellent fitting and prediction results. In Example D, ARGM, SAIGM, and IGGM models obtain good fitting and prediction performance simultaneously. Though the MAPEs of ARGM and SAIGM for fitting are slightly lower than the MAPE of the IGGM model, the MAPE of the IGGM model for prediction is distinctly lower than those of ARGM and SAIGM. Above all, it can be drawn that IGGM has more advantages in prediction than the other six grey models. Notably, the various kinds of energy consumption prediction are involved in Example A, B, and C. To some extent, it indicates that the IGGM model has the potential of energy prediction.

6. Applications

In the past decade, Japan’s energy consumption mix has undergone significant changes. In addition, the total primary energy consumption of Japan decreased from 504.7 TKW in 2010 to 454.1 TKW in 2018, according to BP Statistical Review of World Energy (BPSRWE) 2019. Especially after the Fukushima Daiichi nuclear crisis, nuclear energy consumption declined rapidly. Japan’s nuclear power consumption was previously the third-highest level in the world but is now the 14th highest of the 30 nuclear energy-producing countries [72]. The energy consumption transition toward renewable energy is another of the most important phenomena in Japan. From BPSRWE 2019, it can be noticed that the solar consumption of Japan has increased rapidly from 0.7 million tonnes oil equivalent in 2009 to 16.2 million tonnes oil equivalent in 2018, while other renewable energy consumption of Japan changed only slightly in the past few decades. Thus, special attention should be paid to solar energy consumption because it is currently the only rapidly growing one of the renewable energies in Japan. Meanwhile, accurate prediction of energy consumption plays an essential role in making energy policy and supply plans by energy authorities and energy companies. Therefore, the IGGM model is employed to predict Japan’s solar energy consumption based on historical consumption data in the last ten years.

6.1. Raw Data of Japan’s Solar Energy Consumption

The raw data used to build Japan’s solar energy consumption prediction model are collected from the BP Statistical Review of World Energy 2019. The raw data are partitioned into two subsets, including modeling dataset and test dataset. The modeling dataset used to build model includes the first eight years’ solar energy consumption from 2009 to 2016. The test dataset used to validate the prediction performance of the built model includes the two years’ solar energy consumption from 2017 to 2018. The raw data are listed in Table 13.

6.2. Construct and Verify Grey Models

In the grey prediction model, the key issue is to determine the optimal parameters of the model. The incomplete coefficient of the IGGM model can be sought out by using WOA. Then, the other linear parameters of IGGM can be estimated by the LS method. The formulation of the IGGM model used to predict Japan’s solar energy consumption can be represented as d x ( 1 ) ( t ) d t 0.0905 x ( 1 ) ( t ) = 0.0056 γ ( 7.6901 , t ) + 0.7836 . The parameters of the other classical grey models can be directly estimated by the LS method. The whitening equation of the GM model can be formulated as d x ( 1 ) ( t ) d t 0.4206 x ( 1 ) ( t ) = 0.4581 . The ARGM model can be represented as x ( 0 ) ( k ) = 1.4205 x ( 0 ) ( k 1 ) + 0.2398 . The DGM model can be formulated as x ( 1 ) ( k ) = 1.5310 x ( 1 ) ( k 1 ) + 0.5917 . The NGM model can be formulated as x ( 0 ) ( k ) 0.3806 z ( 1 ) ( k ) = 0.1698 k . The SAIGM model can be formulated as x ( 0 ) ( k ) 0.3620 z ( 1 ) ( k ) = 0.2633 k 0.3089 . The whitening equation of the GVM model can be formulated as d x ( 1 ) ( t ) d t 0.5529 x ( 1 ) ( t ) = 0.0050 x ( 1 ) ( t ) 2 . The produced results of these grey models are tabulated in Table 14. The valuation results of all grey models are placed in Table 15. The MAPEs of different grey models for fitting and prediction are plotted in Figure 8. From Table 15 and Figure 8, it can be noticed that the MAPEs of the IGGM model for fitting and prediction are the lowest while the fitting and prediction performance of the other six grey models are significantly worse than those of the IGGM model. According to Lewis’ criterion of prediction valuation [73], the IGGM model has excellent prediction ability because its MAPE of prediction is far less than 10%. Therefore, the IGGM model can be used to forecast the short-term solar energy consumption of Japan.

6.3. Forecasting Japan’s Solar Energy Consumption with Grey Models

This study forecasts Japan’s short-term solar energy consumption during the period 2019–2021. The predicted results of different grey models are filled in Table 16. The predicted consumptions of GM, ARGM, DGM, and the NGM model in 2021 rapidly grow more than 70 million tonnes oil equivalent, which is more than four times the consumption in 2018. From the empirical point of view, these results are not reasonable in the real world. The predicted consumptions of GVM are declined from 15.15592 to 10.38506. These results of GVM are also not reasonable based on existing photovoltaic capacity and future ongoing investment. The predicted consumptions from 2019 to 2021 by the IGGM model are 20.12687, 22.94959, and 25.73219, respectively. Based on the trend of solar energy consumption and the renewable energy mix of Japan, the results of the IGGM model are much more reasonable than those of the other six classical grey models.

7. Conclusions

Forecasting energy consumption is one of the critical issues to formulate sustainable energy policy. The grey model is an effective method to solve the prediction problem with small samples. However, there is still room to promote its prediction performance. Optimization of grey action quantity is one of the most effective approaches to enhance the prediction performance of the grey model. In this study, a novel incomplete gamma grey prediction model named the IGGM model is put forward by improving the grey input of the classical grey prediction model. The incomplete gamma grey input of the IGGM model is an incomplete gamma function of time, which is formulated as b γ ( μ , t ) + c . It is apparently a nonlinear grey input of which the nonlinear coefficient plays a significant role in boosting the prediction performance of the IGGM model. In addition, the optimal value of the incomplete coefficient is searched by employing WOA. On the other hand, the IGGM model can be degenerated into some classical grey models, such as GM(1,1) and the FOTP-GM model, when the coefficient μ is set to different values. In other words, the IGGM model has the characteristics of the above-mentioned grey models. Finally, the IGGM model is applied to forecast the short-term solar energy consumption of Japan. The MAPE of IGGM for building the solar energy consumption prediction model is 2.206%, which is far superior to those of the other six competitive grey models. The MAPE of IGGM for verifying the prediction performance is 3.5056%, which is also far better than the other six competitive grey models. Japan’s solar energy consumptions forecasted by the IGGM model from 2019 to 2021 are respectively 20.12687, 22.94959, and 25.73219 million tonnes oil equivalent. They can serve as reference information to make energy policy and energy plan. As future work, the IGGM model can be applied in more application fields such as gas and oil prediction [74,75]. Meanwhile, the nonlinear grey action quantity can be adopted to optimize the fractional grey prediction model in the future.

Author Contributions

Conceptualization, P.Z. and X.M.; methodology, P.Z; software, P.Z.; validation, P.Z.; formal analysis, X.M.; investigation, P.Z.; data curation, X.M.; writing—original draft preparation, P.Z; writing—review and editing, P.Z.; visualization, P.Z.; supervision, X.M.; project administration, K.S.; all authors read and approved the final manuscript.

Funding

This work is funded by the National Natural Science Foundation of China (No. 71901184, No. 61672136 and No. 1872323), the Humanities and Social Science Project of the Ministry of Education of China (No. 19YJCZH119), the Open Fund (PLN 201710) of the State Key Laboratory of Oil and Gas Reservoir Geology and Exploitation (Southwest Petroleum University), and the Doctoral Research Foundation of Southwest University of Science and Technology (No. 16zx7140).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
GM(1,1)Grey model
APEAbsolute percentage error
MAPEMean absolute percentage error
SECSolar energy consumption
COPCumulative oil production
HECHydro electricity consumption
NECNuclear energy consumption
ADECAverage daily electricity consumption
LSLeast square
WOAWhale optimization algorithm
BPSRWEBP Statistical Review of World Energy
TBGMGrey model based on both data transformation and background-value optimization
GDPGross domestic product
Nash NGBM(1,1)Nash nonlinear grey Bernoulli model
NHGM(1,1,k)Non-homogeneous exponential grey model
GMC(1,n)Grey model with convolution integral
TNGM(1,n)Transformed nonlinear grey multivariable
GMCN(1,n)New information priority accumulated grey multivariable convolution model
EOGM(1,1)Exponential optimization grey model
NMGM(1,n)New multivariable grey prediction model with structure compatibility
GM(1,n)The traditional multivariate grey model
IGGMIncomplete gamma grey model
NGM(1,1,k)Non-homogeneous grey model
SAIGMSelf-adapting intelligent grey model
GM(1,1, t α )Grey model with time power
FOTP-GMGrey model with full-order time power terms
DGMDiscrete grey model
ARGMAutoregressive grey model
GVMGrey Verhulst model
AGOAccumulated generation operator
WDGMDiscrete grey model with the weighted accumulation
CFGMComfortable fractional grey model
NGBM(1,1)Nonlinear grey Bernoulli model

References

  1. Hayashi, M.; Hughes, L. The policy responses to the Fukushima nuclear accident and their effect on Japanese energy security. Energy Policy 2013, 59, 86–101. [Google Scholar] [CrossRef]
  2. Wu, L.; Liu, S.; Chen, H.; Zhang, N. Using a Novel Grey System Model to Forecast Natural Gas Consumption in China. Math. Probl. Eng. 2015, 2015, 1–7. [Google Scholar] [CrossRef]
  3. Ding, S. A novel self-adapting intelligent grey model for forecasting China’s natural-gas demand. Energy 2018, 162, 393–407. [Google Scholar] [CrossRef]
  4. Zeng, B.; Li, C. Forecasting the natural gas demand in China using a self-adapting intelligent grey model. Energy 2016, 112, 810–825. [Google Scholar] [CrossRef]
  5. Xu, N.; Dang, Y.; Gong, Y. Novel grey prediction model with nonlinear optimized time response method for forecasting of electricity consumption in China. Energy 2017, 118, 473–480. [Google Scholar] [CrossRef]
  6. Kumar, U.; Jain, V.K. Time series models (Grey-Markov, Grey Model with rolling mechanism and singular spectrum analysis) to forecast energy consumption in India. Energy 2010, 35, 1709–1716. [Google Scholar] [CrossRef]
  7. Hu, Y. Electricity consumption prediction using a neural-network-based grey forecasting approach. J. Oper. Res. Soc. 2017, 68, 1259–1264. [Google Scholar] [CrossRef]
  8. Wu, W.; Ma, X.; Zeng, B.; Wang, Y.; Cai, W. Forecasting short-term renewable energy consumption of China using a novel fractional nonlinear grey Bernoulli model. Renew. Energy 2019, 140, 70–87. [Google Scholar] [CrossRef]
  9. Tsai, S.B. Using grey models for forecasting China’s growth trends in renewable energy consumption. Clean Technol. Environ. Policy 2016, 18, 563–571. [Google Scholar] [CrossRef]
  10. Wang, Q.; Liu, L.; Wang, S.; Wang, J.Z.; Liu, M. Predicting Beijing’s tertiary industry with an improved grey model. Appl. Soft Comput. J. 2017, 57, 482–494. [Google Scholar] [CrossRef]
  11. Ma, X.; Liu, Z. Predicting the Cumulative Oil Field Production Using the Novel Grey ENGM Model. J. Comput. Theor. Nanosci. 2016, 13, 89–95. [Google Scholar] [CrossRef]
  12. Ding, S.; Hipel, K.W.; Dang, Y.G. Forecasting China’s electricity consumption using a new grey prediction model. Energy 2018, 149, 314–328. [Google Scholar] [CrossRef]
  13. Hamzacebi, C.; Karakurt, I. Forecasting the energy-related CO2 emissions of Turkey using a grey prediction model. Energy Sources Part A Recover. Util. Environ. Eff. 2015, 37, 1023–1031. [Google Scholar] [CrossRef]
  14. Ding, S.; Dang, Y.G.; Li, X.M.; Wang, J.J.; Zhao, K. Forecasting Chinese CO2 emissions from fuel combustion using a novel grey multivariable model. J. Clean. Prod. 2017, 162, 1527–1538. [Google Scholar] [CrossRef]
  15. Lin, C.; He, R.; Liu, W. Considering multiple factors to forecast CO2 emissions: A hybrid multivariable grey forecasting and genetic programming approach. Energies 2018, 11, 3432. [Google Scholar] [CrossRef]
  16. Shaikh, F.; Ji, Q.; Shaikh, P.H.; Mirjat, N.H.; Uqaili, M.A. Forecasting China’s natural gas demand based on optimised nonlinear grey models. Energy 2017, 140, 941–951. [Google Scholar] [CrossRef]
  17. Lin, J.; Zhu, K.; Liu, Z.; Lieu, J.; Tan, X. Study on a simple model to forecast the electricity demand under China’s new normal situation. Energies 2019, 12, 2220. [Google Scholar] [CrossRef]
  18. Li, K.; Zhang, T. A novel grey forecasting model and its application in forecasting the energy consumption in Shanghai. Energy Syst. 2019. [Google Scholar] [CrossRef]
  19. Zhang, W.; Xiao, R.; Shi, B.; Zhu, H.H.; Sun, Y.J. Forecasting slope deformation field using correlated grey model updated with time correction factor and background value optimization. Eng. Geol. 2019, 260, 105215. [Google Scholar] [CrossRef]
  20. Zeng, B.; Li, C. Improved multi-variable grey forecasting model with a dynamic background-value coefficient and its application. Comput. Ind. Eng. 2018, 118, 278–290. [Google Scholar] [CrossRef]
  21. Ye, J.; Dang, Y.; Li, B. Grey-Markov prediction model based on background value optimization and central-point triangular whitenization weight function. Commun. Nonlinear Sci. Numer. Simul. 2018, 54, 320–330. [Google Scholar] [CrossRef]
  22. Li, L.; Wang, H. A VVWBO-BVO-based GM (1,1) and its parameter optimization by GRA-IGSA integration algorithm for annual power load forecasting. PLoS ONE 2018, 13, e0196816. [Google Scholar] [CrossRef] [PubMed]
  23. Atalay, S.D.; Calis, G.; Kus, G.; Kuru, M. Performance analyses of statistical approaches for modeling electricity consumption of a commercial building in France. Energy Build. 2019, 195, 82–92. [Google Scholar] [CrossRef]
  24. Rahman, M.H.; Tumpa, T.J.; Ali, S.M.; Paul, S.K. A grey approach to predicting healthcare performance. Meas. J. Int. Meas. Confed. 2019, 134, 307–325. [Google Scholar] [CrossRef]
  25. Ma, X.; Wu, W.; Zhang, Y. Improved GM(1,1) model based on Simpson formula and its applications. arXiv 2019, arXiv:1908.03493. [Google Scholar]
  26. Wang, Y.; Dang, Y.; Li, Y.; Liu, S. An approach to increase prediction precision of GM(1,1) model based on optimization of the initial condition. Expert Syst. Appl. 2010, 37, 5640–5644. [Google Scholar] [CrossRef]
  27. Wang, Z.X. An optimized Nash nonlinear grey Bernoulli model for forecasting the main economic indices of high technology enterprises in China. Comput. Ind. Eng. 2013, 64, 780–787. [Google Scholar] [CrossRef]
  28. Zeng, B.; Duan, H.; Bai, Y.; Meng, W. Forecasting the output of shale gas in China using an unbiased grey model and weakening buffer operator. Energy 2018, 151, 238–249. [Google Scholar] [CrossRef]
  29. Wu, L.; Liu, S.; Yao, L.; Yan, S.; Liu, D. Grey system model with the fractional order accumulation. Commun. Nonlinear Sci. Numer. Simul. 2013, 18, 1775–1785. [Google Scholar] [CrossRef]
  30. Mao, S.; Gao, M.; Xiao, X.; Zhu, M. A novel fractional grey system model and its application. Appl. Math. Model. 2016, 40, 5063–5076. [Google Scholar] [CrossRef]
  31. Wu, L.; Zhao, H. Discrete grey model with the weighted accumulation. Soft Comput. 2019, 3. [Google Scholar] [CrossRef]
  32. Ma, X.; Wu, W.; Zeng, B.; Wang, Y.; Wu, X. The conformable fractional grey system model. ISA Trans. 2019. [Google Scholar] [CrossRef] [PubMed]
  33. Tien, T.L. A research on the grey prediction model GM(1,n). Appl. Math. Comput. 2012, 218, 4903–4916. [Google Scholar] [CrossRef]
  34. Wu, L.; Liu, S.; Yang, Y.; Ma, L.; Liu, H. Multi-variable weakening buffer operator and its application. Inf. Sci. 2016, 339, 98–107. [Google Scholar] [CrossRef] [Green Version]
  35. Pei, L.; Li, Q.; Wang, Z. The NLS-based nonlinear grey multivariate model for forecasting pollutant emissions in China. Int. J. Environ. Res. Public Health 2018, 15, 471. [Google Scholar] [CrossRef]
  36. Wu, L.; Zhang, Z. Grey multivariable convolution model with new information priority accumulation. Appl. Math. Model. 2018, 62, 595–604. [Google Scholar] [CrossRef]
  37. Xia, J.; Ma, X.; Wu, W.; Huang, B.; Li, W. Application of a new information priority accumulated grey model with time power to predict short-term wind turbine capacity. J. Clean. Prod. 2020, 244, 118573. [Google Scholar] [CrossRef]
  38. Al-shanini, A.; Ahmad, A.; Khan, F.; Oladokun, O.; Mohd Nor, S.H. Alternative Prediction Models for Data Scarce Environment. In Computer Aided Chemical Engineering; Elsevier: Amsterdam, The Netherlands, 2015; Volume 37, pp. 665–670. [Google Scholar] [CrossRef]
  39. Zeng, B.; Luo, C.; Liu, S.; Bai, Y.; Li, C. Development of an optimization method for the GM(1,N) model. Eng. Appl. Artif. Intell. 2016, 55, 353–362. [Google Scholar] [CrossRef]
  40. Yin, K.; Geng, Y.; Li, X. Improved grey prediction model based on exponential grey action quantity. J. Syst. Eng. Electron. 2018, 29, 560–570. [Google Scholar] [CrossRef]
  41. Wu, W.; Ma, X.; Zeng, B.; Wang, Y.; Cai, W. Application of the novel fractional grey model FAGMO(1,1,k) to predict China’s nuclear energy consumption. Energy 2018, 165, 223–234. [Google Scholar] [CrossRef]
  42. Ma, X.; Mei, X.; Wu, W.; Wu, X.; Zeng, B. A novel fractional time delayed grey model with Grey Wolf Optimizer and its applications in forecasting the natural gas and coal consumption in Chongqing China. Energy 2019, 178, 487–507. [Google Scholar] [CrossRef]
  43. Zeng, B.; Duan, H.; Zhou, Y. A new multivariable grey prediction model with structure compatibility. Appl. Math. Model. 2019, 75, 385–397. [Google Scholar] [CrossRef]
  44. Wang, Z.; Hao, P. An improved grey multivariable model for predicting industrial energy consumption in China. Appl. Math. Model. 2016, 40, 5745–5758. [Google Scholar] [CrossRef]
  45. Ma, X. A Brief Introduction to the Grey Machine Learning. J. Grey Syst. 2019, 31, 1–12. [Google Scholar]
  46. Ma, X.; Liu, Z.B. The kernel-based nonlinear multivariate grey model. Appl. Math. Model. 2018, 56, 217–238. [Google Scholar] [CrossRef]
  47. Wu, L.; Huang, G.; Fan, J.; Zhang, F.; Wang, X.; Zeng, W. Potential of kernel-based nonlinear extension of Arps decline model and gradient boosting with categorical features support for predicting daily global solar radiation in humid regions. Energy Convers. Manag. 2019, 183, 280–295. [Google Scholar] [CrossRef]
  48. Duan, H.; Xiao, X. A multimode dynamic short-term traffic flow grey prediction model of high-dimension tensors. Complexity 2019, 2019. [Google Scholar] [CrossRef]
  49. Xu, N.; Ding, S.; Gong, Y.; Bai, J. Forecasting Chinese greenhouse gas emissions from energy consumption using a novel grey rolling model. Energy 2019, 175, 218–227. [Google Scholar] [CrossRef]
  50. Liu, H.; Guo, W.; Zhang, C.; Yang, H. Research on the Grey Verhulst Model Based on Particle Swarm Optimization and Markov Chain to Predict the Settlement of High Fill Subgrade in Xiangli Expressway. Math. Probl. Eng. 2019, 2019. [Google Scholar] [CrossRef]
  51. Hu, Y.C.C. A genetic-algorithm-based remnant grey prediction model for energy demand forecasting. PLoS ONE 2017, 12, e0185478. [Google Scholar] [CrossRef]
  52. Ma, X.; Xie, M.; Wu, W.; Zeng, B.; Wang, Y.; Wu, X. The novel fractional discrete multivariate grey system model and its applications. Appl. Math. Model. 2019, 70, 402–424. [Google Scholar] [CrossRef]
  53. Ding, S. A novel discrete grey multivariable model and its application in forecasting the output value of China’s high-tech industries. Comput. Ind. Eng. 2019, 127, 749–760. [Google Scholar] [CrossRef]
  54. Bahrami, S.; Hooshmand, R.A.; Parastegari, M. Short term electric load forecasting by wavelet transform and grey model improved by PSO (particle swarm optimization) algorithm. Energy 2014, 72, 434–442. [Google Scholar] [CrossRef]
  55. Niu, X.; Wang, J. A combined model based on data preprocessing strategy and multi-objective optimization algorithm for short-term wind speed forecasting. Appl. Energy 2019, 241, 519–539. [Google Scholar] [CrossRef]
  56. He, Z.; Shen, Y.; Li, J.; Wang, Y. Regularized multivariable grey model for stable grey coefficients estimation. Expert Syst. Appl. 2015, 42, 1806–1815. [Google Scholar] [CrossRef]
  57. De Prada, C.; Hose, D.; Gutierrez, G.; Pitarch, J.L. Developing Grey-Box Dynamic Process Models. IFAC-PapersOnLine 2018, 51, 523–528. [Google Scholar] [CrossRef]
  58. Ma, X.; Liu, Z. The GMC (1, n) Model with Optimized Parameters and Its Application. J. Grey Syst. 2017, 29, 122–138. [Google Scholar]
  59. Chen, P.Y.; Yu, H.M. Foundation Settlement Prediction Based on a Novel NGM Model. Math. Probl. Eng. 2014, 2014, 1–8. [Google Scholar] [CrossRef]
  60. Zeng, B.; Liu, S. A self-adaptive intelligence gray prediction model with the optimal fractional order accumulating operator and its application. Math. Methods Appl. Sci. 2017, 40, 7843–7857. [Google Scholar] [CrossRef]
  61. Qian, W.; Dang, Y.; Liu, S. Grey GM(1,1,t^∖alpha) model with time power and its application. Syst. Eng. Theory Pract. 2012, 32, 2247–2252. [Google Scholar] [CrossRef]
  62. Li, S.; Ma, X.; Yang, C. A novel structure-adaptive intelligent grey forecasting model with full-order time power terms and its application. Comput. Ind. Eng. 2018, 120, 53–67. [Google Scholar] [CrossRef]
  63. Lu, J.; Xie, W.; Zhou, H.; Zhang, A. An optimized nonlinear grey Bernoulli model and its applications. Neurocomputing 2016, 177, 206–214. [Google Scholar] [CrossRef]
  64. Kayacan, E.; Ulutas, B.; Kaynak, O. Grey system theory-based models in time series prediction. Expert Syst. Appl. 2010, 37, 1784–1789. [Google Scholar] [CrossRef]
  65. Mirjalili, S.; Lewis, A. The Whale Optimization Algorithm. Adv. Eng. Softw. 2016, 95, 51–67. [Google Scholar] [CrossRef]
  66. Gharehchopogh, F.S.; Gholizadeh, H. A comprehensive survey: Whale Optimization Algorithm and its applications. Swarm Evol. Comput. 2019, 48, 1–24. [Google Scholar] [CrossRef]
  67. Fan, J.; Wu, L.; Ma, X.; Zhou, H.; Zhang, F. Hybrid support vector machines with heuristic algorithms for prediction of daily diffuse solar radiation in air-polluted regions. Renew. Energy 2020, 145, 2034–2045. [Google Scholar] [CrossRef]
  68. Zhang, P.; Ma, X.; She, K. A novel power-driven grey model with Whale Optimization Algorithm and its application of forecasting the residential energy consumption in China. Complexity 2019. [Google Scholar] [CrossRef]
  69. Li, J.; Wang, R.; Wang, J.; Li, Y. Analysis and forecasting of the oil consumption in China based on combination models optimized by artificial intelligence algorithms. Energy 2018, 144, 243–264. [Google Scholar] [CrossRef]
  70. Long, D.J. Control problems of grey systems. Syst. Control Lett. 1982, 1, 288–294. [Google Scholar] [CrossRef]
  71. Cui, J.; Liu, S.F.; Zeng, B.; Xie, N.M. A novel grey forecasting model and its optimization. Appl. Math. Model. 2013, 37, 4399–4406. [Google Scholar] [CrossRef]
  72. Kharecha, P.A.; Sato, M. Implications of energy and CO2 emission changes in Japan and Germany after the Fukushima accident. Energy Policy 2019, 132, 647–653. [Google Scholar] [CrossRef]
  73. Xiong, P.; Yan, W.; Wang, G.; Pei, L. Grey extended prediction model based on IRLS and its application on smog pollution. Appl. Soft Comput. J. 2019, 80, 797–809. [Google Scholar] [CrossRef]
  74. Lu, H.; Guo, L.; Azimi, M.; Huang, K. Oil and Gas 4.0 era: A systematic review and outlook. Comput. Ind. 2019, 111, 68–90. [Google Scholar] [CrossRef]
  75. Ding, X.; Qu, D.; Qiu, H. A New Production Prediction Model Based on Taylor Expansion Formula. Math. Probl. Eng. 2018, 2018, 1–12. [Google Scholar] [CrossRef]
Figure 1. The curves of the grey input functions k α and e α t .
Figure 1. The curves of the grey input functions k α and e α t .
Sustainability 11 05921 g001
Figure 2. The curves of the grey input function γ ( μ , k ) .
Figure 2. The curves of the grey input function γ ( μ , k ) .
Sustainability 11 05921 g002
Figure 3. Flowchart of the incomplete gamma grey model.
Figure 3. Flowchart of the incomplete gamma grey model.
Sustainability 11 05921 g003
Figure 4. MAPEs of different grey models in Example A.
Figure 4. MAPEs of different grey models in Example A.
Sustainability 11 05921 g004
Figure 5. MAPEs of different grey models in Example B.
Figure 5. MAPEs of different grey models in Example B.
Sustainability 11 05921 g005
Figure 6. MAPEs of different grey models in Example C.
Figure 6. MAPEs of different grey models in Example C.
Sustainability 11 05921 g006
Figure 7. MAPEs of different grey models in Example D.
Figure 7. MAPEs of different grey models in Example D.
Sustainability 11 05921 g007
Figure 8. MAPEs of different grey models in the case study.
Figure 8. MAPEs of different grey models in the case study.
Sustainability 11 05921 g008
Table 1. China’s average daily electricity consumption (ADEC) from 2006 to 2016 (108 kilowatts hours).
Table 1. China’s average daily electricity consumption (ADEC) from 2006 to 2016 (108 kilowatts hours).
YearADECYearADECYearADECYearADECYearADECYearADEC
200678.3200894.42010114.920121362014154.52016167.5
200789.62009101.52011128.82013148.52015159
Table 2. Fitted and predicted values of different grey models in Example A.
Table 2. Fitted and predicted values of different grey models in Example A.
YearRaw DataGMARGMDGMNGMSAIGMGVMIGGM
200678.378.378.378.378.378.378.378.3
200789.687.603987.151387.678948.598287.544839.895189.6
200894.495.734996.37695.818281.092895.785557.307994.0626
2009101.5104.6205105.9896104.713102.5308104.753279.1194102.6355
2010114.9114.3308116.0087114.4336116.6741114.5121103.4536114.3785
2011128.8124.9424126.4503125.0565126.005125.1321126.0643126.9591
2012136136.539137.3323136.6656132.1609136.689141.0617138.3768
2013148.5149.2118148.6733149.3524136.2222149.2656143.547147.6738
2014154.5163.0609160.4925163.2168138.9016162.9518132.634154.777
2015159178.1954172.8101178.3684140.6693177.8456112.0574160.0748
2016167.5194.7345185.6473194.9264141.8355194.053487.732164.0766
Table 3. APEs and MAPEs of different grey models in Example A.
Table 3. APEs and MAPEs of different grey models in Example A.
YearRaw DataGMARGMDGMNGMSAIGMGVMIGGM
200678.30000000
200789.62.22782.73292.144145.7612.293755.47420
200894.41.41412.09321.502314.09661.467739.29250.3575
2009101.53.07444.42333.16551.01553.205122.04981.1188
2010114.90.49540.96490.40591.54410.33769.9620.4539
2011128.82.9951.82432.90642.172.84782.1241.4293
20121360.39630.97970.48942.82290.50663.72181.7476
2013148.50.47930.11670.5748.26790.51563.33540.5564
2014154.55.5413.87865.64210.09615.470414.15270.1793
201515912.07268.685612.181411.528811.852629.52370.676
2016167.516.259410.834216.37415.322115.852847.62272.0438
M A P E f i t t i n g 1.38531.64191.39859.45971.396816.9950.7079
M A P E p r e d i c t i o n 11.2917.799511.399112.315711.058630.4330.9664
M A P E o v e r a l l 4.08683.32124.125910.23864.031820.65990.7784
Table 4. China’s nuclear energy consumption (NEC) from 2008 to 2018 (million tonnes of oil equivalent).
Table 4. China’s nuclear energy consumption (NEC) from 2008 to 2018 (million tonnes of oil equivalent).
YearNECYearNECYearNECYearNECYearNECYearNEC
200714.1200915.9201119.5201325.3201538.6201756.1
200815.5201016.7201222201430201648.3201866.6
Table 5. Fitted and predicted values of different grey models in Example B.
Table 5. Fitted and predicted values of different grey models in Example B.
YearRaw DataGMARGMDGMNGMSAIGMGVMIGGM
200714.114.114.114.114.114.114.114.1
200815.512.821314.563612.88646.437815.26655.346315.5266
200915.914.839115.235914.910210.856216.09667.245915.9023
201016.717.174516.210817.251815.067117.30119.711216.7
201119.519.877417.624319.961219.080319.048712.822118.5163
20122223.005719.67423.09622.905221.584616.59921.4709
201325.326.626322.64626.723226.550425.26420.948525.5627
20143030.816626.955630.920130.024530.602825.60530.8306
201538.635.666533.204435.776133.335638.349330.137.3969
201648.341.279642.265441.394736.491149.589233.804345.4683
201756.147.776255.40447.895739.498665.898236.071255.3329
201866.655.295174.455255.417642.364889.562236.448867.3614
Table 6. APEs and MAPEs of different grey models in Example B.
Table 6. APEs and MAPEs of different grey models in Example B.
YearRaw DataGMARGMDGMNGMSAIGMGVMIGGM
200714.10000000
200815.517.28166.04116.862258.46551.506465.5080.1716
200915.96.6724.17646.225531.72211.236554.42840.0147
201016.72.84132.92943.30419.77783.599341.84890
201119.51.93549.61882.36492.15212.314134.24585.0445
2012224.571210.57274.9824.11431.888324.55012.4048
201325.35.242110.48995.62554.94230.142317.19941.0382
2014302.722110.14813.0670.08182.009314.64982.7688
201538.67.599713.97817.315913.63850.649622.02073.1169
201648.314.534912.493914.296724.4492.669230.01195.8628
201756.114.83751.240614.624529.592617.465635.7021.3674
201866.616.974411.794616.790336.389234.477845.2721.1432
M A P E f i t t i n g 5.42957.55055.527413.87721.482930.49461.6177
M A P E p r e d i c t i o n 15.44898.509715.237230.143618.204236.99532.7911
M A P E o v e r a l l 7.93447.79037.954917.94385.663232.11981.9111
Table 7. China’s hydro electricity consumption (HEC) from 2008 to 2018 (million tonnes of oil equivalent).
Table 7. China’s hydro electricity consumption (HEC) from 2008 to 2018 (million tonnes of oil equivalent).
YearHECYearHECYearHECYearHECYearHECYearHEC
2007109.82009139.32011155.72013205.82015252.22017263.6
2008144.120101612012195.22014237.820162612018272.1
Table 8. Fitted and predicted values of different grey models in Example C.
Table 8. Fitted and predicted values of different grey models in Example C.
YearRaw DataGMARGMDGMNGMSAIGMGVMIGGM
2007109.8109.8109.8109.8109.8109.8109.8109.8
2008144.1133.5679131.0898133.785267.7323133.849844.7947144.1
2009139.3145.7886150.9755146.0083113.2576145.975261.6218144.1285
2010161159.1274169.5498159.3481148.6234159.240883.4428150.585
2011155.7173.6867186.899173.9066176.097173.7537110.6087165.6826
2012195.2189.578203.1041189.7953197.4396189.6313142.549187.6051
2013205.8206.9233218.2404207.1356214.0194207.0018177.193211.8547
2014237.8225.8557232.3784226.0601226.8993226.0057210.6237233.98
2015252.2246.5202245.584246.7137236.9049246.7964237.4526251.2433
2016261269.0754257.9187269.2543244.6776269.5422252.2393262.8209
2017263.6293.6943269.4399293.8542250.7158294.4267251.5681269.2077
2018272.1320.5656280.2012320.7016255.4065321.651235.6004271.4988
Table 9. APEs and MAPEs of different grey models in Example C.
Table 9. APEs and MAPEs of different grey models in Example C.
YearRaw DataGMARGMDGMNGMSAIGMGVMIGGM
2007109.80000000
2008144.17.30899.02867.158152.99637.113268.91420
2009139.34.6588.38164.815718.69524.79255.76323.4662
20101611.16315.31041.02617.68731.092748.17226.4689
2011155.711.552120.037911.693413.100211.595228.96046.4114
2012195.22.88014.04922.76881.14742.852826.97293.8908
2013205.80.54586.04490.6493.99390.58413.90042.942
2014237.85.02292.27994.93694.5844.959811.42821.6064
2015252.22.25212.62332.17546.06472.14265.84750.3794
20162613.0941.18063.16266.25383.27293.35660.6977
2017263.611.41662.215411.47734.887811.69454.56452.1274
2018272.117.81172.977317.86176.135118.210613.4140.2209
M A P E f i t t i n g 3.84775.89363.838611.45233.840526.33162.5863
M A P E p r e d i c t i o n 14.61422.596414.66955.511414.95258.98921.1741
M A P E o v e r a l l 5.64215.34415.643710.46215.692523.44122.3509
Table 10. The raw data of the cumulative oil production (COP) in Example D.
Table 10. The raw data of the cumulative oil production (COP) in Example D.
YearCOPYearCOPYearCOPYearCOP
2001195.0592004342.63942007454.0432010552.6569
2002247.85472005382.43122008485.11712011581.6092
2003297.09022006420.03992009519.85082012608.1863
Table 11. Fitted and predicted values of different grey models in Example D.
Table 11. Fitted and predicted values of different grey models in Example D.
YearRaw DataGMARGMDGMNGMSAIGMGVMIGGM
2001195.059195.059195.059195.059195.059195.059195.059195.059
2002247.8547270.5874248.4202271.0127155.6094247.9491110.0805247.8547
2003297.0902299.8087297.2072300.2579250.2939297.165164.6198296.3056
2004342.6394332.1856341.8121332.659319.9912342.0147236.6466342.8489
2005382.4312368.059382.5935368.5565371.2953382.8853321.5642383.328
2006420.0399407.8065419.8791408.3277409.0602420.1301405.075419.0918
2007454.043451.8463453.9685452.3907436.8589454.0705464.2502452.5551
2008485.1171500.6422485.1358501.2085457.3216484.9998477.7854485.6453
2009519.8508554.7075513.6314555.2943472.3841513.1851440.2383519.6086
2010552.6569614.6116539.6843615.2165483.4716538.8699366.1885555.1964
2011581.6092680.9847563.504681.605491.6331562.276279.542592.8695
2012608.1863754.5257585.2818755.1576497.6408583.6055199.7332632.9383
Table 12. APEs and MAPEs of different grey models in Example D.
Table 12. APEs and MAPEs of different grey models in Example D.
YearRaw DataGMARGMDGMNGMSAIGMGVMIGGM
2001195.0590000000
2002247.85479.17180.22819.343437.21750.038155.58670
2003297.09020.9150.03941.066215.75150.025244.58930.2641
2004342.63943.0510.24142.91286.60990.182330.93420.0611
2005382.43123.75810.04243.6282.91190.118815.91580.2345
2006420.03992.91240.03832.78842.6140.02153.56270.2257
2007454.0430.48380.01640.36393.78470.00612.24810.3277
2008485.11713.20030.00393.3175.72970.02421.51130.1089
2009519.85086.70511.19646.8189.13081.282215.31450.0466
2010552.656911.21032.347311.319812.51872.494733.74040.4595
2011581.609217.08633.11317.19315.47023.324151.93651.9361
2012608.186324.06163.76624.165518.17634.041667.15924.0698
M A P E f i t t i n g 2.93650.07622.92759.32740.05219.29350.1528
M A P E p r e d i c t i o n 14.76582.605714.874113.8242.785742.03761.628
M A P E o v e r a l l 6.87960.91946.909710.82630.963226.87490.6445
Table 13. Solar energy consumption (SEC) of Japan from 2009 to 2018 (million tonnes oil equivalent).
Table 13. Solar energy consumption (SEC) of Japan from 2009 to 2018 (million tonnes oil equivalent).
YearSECYearSECYearSECYearSECYearSEC
20090.720111.220132.920157.8201714
20100.920121.720145.3201611201816.2
Table 14. Fitted and predicted values of different grey models for forecasting solar energy consumption of Japan.
Table 14. Fitted and predicted values of different grey models for forecasting solar energy consumption of Japan.
YearRaw DataGMARGMDGMNGMSAIGMGVMIGGM
20090.70.70.70.70.70.70.70.7
20100.90.93551.23420.96340.62770.57530.51110.9
20111.21.42471.99311.4751.12521.15110.87731.1164
20121.72.16973.07112.25831.85311.98131.49221.7278
20132.93.30434.60253.45762.91823.17852.4993.0346
20145.35.03226.77795.29374.47664.90484.07865.1371
20157.87.66379.86818.10486.75697.39416.38487.8769
20161111.671314.257912.408810.093410.98369.369910.9619
20171417.774520.493818.998414.975516.159512.524714.1224
201816.227.069229.352229.087422.11923.62314.804717.1942
Table 15. APEs and MAPEs of different grey models for forecasting solar energy consumption of Japan.
Table 15. APEs and MAPEs of different grey models for forecasting solar energy consumption of Japan.
YearRaw DataGMARGMDGMNGMSAIGMGVMIGGM
20090.70000000
20100.93.94437.13547.045630.250536.07843.20670
20111.218.724466.091622.91846.23214.077226.88946.9703
20121.727.629780.654232.84219.006916.548412.22471.6324
20132.913.941558.706819.22660.62729.604113.82814.6403
20145.35.052627.88460.119215.53577.45623.04473.0727
20157.81.747526.51433.908213.3735.203518.14410.9859
2016116.102329.617412.80768.24140.14914.81940.346
20171426.960546.384535.70316.967915.425210.53790.8739
201816.267.09481.186479.551636.536845.82128.61276.1373
M A P E f i t t i n g 9.642740.825512.358510.40839.889519.01962.206
M A P E p r e d i c t i o n 47.027363.785457.627421.752430.62329.57533.5056
M A P E o v e r a l l 17.119645.417521.412212.677114.036317.13082.4659
Table 16. The predicted solar energy consumption of Japan.
Table 16. The predicted solar energy consumption of Japan.
YearGMARGMDGMNGMSAIGMGVMIGGM
201941.2244841.935944.5339332.5713434.3851715.1559220.12687
202062.7818959.8115968.1832447.8653349.9038213.3947322.94959
202195.6122785.20477104.391370.2436172.2812410.3850625.73219

Share and Cite

MDPI and ACS Style

Zhang, P.; Ma, X.; She, K. Forecasting Japan’s Solar Energy Consumption Using a Novel Incomplete Gamma Grey Model. Sustainability 2019, 11, 5921. https://doi.org/10.3390/su11215921

AMA Style

Zhang P, Ma X, She K. Forecasting Japan’s Solar Energy Consumption Using a Novel Incomplete Gamma Grey Model. Sustainability. 2019; 11(21):5921. https://doi.org/10.3390/su11215921

Chicago/Turabian Style

Zhang, Peng, Xin Ma, and Kun She. 2019. "Forecasting Japan’s Solar Energy Consumption Using a Novel Incomplete Gamma Grey Model" Sustainability 11, no. 21: 5921. https://doi.org/10.3390/su11215921

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