Next Article in Journal
Clustering Analysis for Active and Reactive Energy Consumption Data Based on AMI Measurements
Previous Article in Journal
Research on Opening Reignition Characteristics and Suppression Measures of 750 kV AC Filter Circuit Breakers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fast Simulation of the Flow Field in a VAWT Wind Farm Using the Numerical Data Obtained by CFD Analysis for a Single Rotor

1
Advanced Mechanical and Electronic System Research Center (AMES), Faculty of Engineering, Tottori University, 4-101 Koyama-Minami, Tottori 680-8552, Japan
2
Graduate School of Engineering, Tottori University, 4-101 Koyama-Minami, Tottori 680-8552, Japan
3
Department of Mechanical and Physical Engineering, Tottori University, 4-101 Koyama-Minami, Tottori 680-8552, Japan
4
Department of Mechanical Engineering, National Institute of Technology (KOSEN), Kagawa College, 355 Chokushi, Takamatsu 761-8058, Japan
*
Author to whom correspondence should be addressed.
Energies 2025, 18(1), 220; https://doi.org/10.3390/en18010220
Submission received: 25 November 2024 / Revised: 27 December 2024 / Accepted: 31 December 2024 / Published: 6 January 2025
(This article belongs to the Section A3: Wind, Wave and Tidal Energy)

Abstract

:
The effects of an increase in output power owing to the close arrangement of vertical-axis wind turbines (VAWTs) are well known. With the ultimate goal of determining the optimal layout of a wind farm (WF) for VAWTs, this study proposes a new method for quickly calculating the flow field and power output of a virtual WF consisting of two-dimensional (2-D) miniature VAWT rotors. This new method constructs a flow field in a WF by superposing 2-D velocity numerical data around an isolated single VAWT obtained through a computational fluid dynamics (CFD) analysis. In the calculation process, the VAWTs were gradually increased one by one from the upstream side, and a calculation subroutine, in which the virtual upstream wind speed at each VAWT position was recalculated with the effects of other VAWTs, was repeated three times for each arrangement with a temporal number of VAWTs. This method includes the effects of the velocity gradient, secondary flow, and wake shift as models of turbine-to-turbine interaction. To verify the accuracy of the method, the VAWT rotor power outputs predicted by the proposed method for several types of rotor pairs, four-rotor tandem, and parallel arrangements were compared with the results of previous CFD analyses. This method was applied to four virtual WFs consisting of 16 miniature VAWTs. It was found that a layout consisting of two linear arrays of eight closely spaced VAWTs with wide spacing between the arrays yielded a significantly higher output than the other three layouts. The high-performance layout had fewer rotors in the wakes of the other rotors, and the induced flow speeds generated by the closely spaced VAWTs probably mutually enhanced their output power.

1. Introduction

To realize a decarbonized society, the early introduction of various forms of renewable energy is necessary. Small wind turbines have advantages over large wind turbines, such as ease of installation; however, the high cost of power generation remains an issue. In recent years, closely spaced arrangements of small vertical-axis wind turbines (VAWTs) have attracted attention because of their power output enhancement effect. If VAWTs are effectively installed with close spacing to suit the shape of the planned site and wind conditions, not only will the amount of electricity generated per unit installation area increase, but construction, maintenance, and removal will also be more efficient, which is expected to significantly reduce power generation costs. However, because the flow field around a VAWT is more complex than that around a horizontal-axis wind turbine (HAWT), the computational cost of numerical analysis is high, and the development of methods for determining the optimal VAWT layout and simple wake models of the VAWT has not progressed sufficiently.
In a pioneering study of VAWT-WF, Rajagopalan et al. [1] performed a two-dimensional (2-D) computational fluid dynamics (CFD) analysis of several cases, including up to 21 VAWTs, assuming a 2-D laminar flow. They showed that VAWTs located in a backward oblique direction could generate higher output power than front VAWTs exposed to undisturbed wind. Whittlesey et al. [2] proposed a method that included the velocity deficit effect of the turbine wake in the potential flow, and showed that a wind farm (WF) consisting of VAWTs closely arranged like a school of fish, in which rows of VAWTs rotating in opposite directions are aligned alternately, could potentially produce greater output per unit installation area than conventional large wind turbines consisting of HAWTs. Dabiri [3] conducted field experiments with small VAWTs with a diameter of 1.2 m and demonstrated that a WF with a grid-like unit or cluster arrangement of a pair of counter-rotating VAWTs could have a higher energy density than a conventional WF with large HAWTs. In addition, he presented measurement results showing the wind-direction dependence of the output of a pair of VAWTs. Kinzel et al. [4] constructed a WF by arranging three pairs × three pairs of counter-rotating VAWT pairs, using the same 1.2 m diameter rotor as that used by Dabiri, in a grid pattern, and experimentally showed that the wake recovery in VAWT-WFs could be faster than that of HAWT-WFs. Zanforlin et al. [5] performed a 2-D CFD analysis using a turbulence model for a 2-D rotor equivalent of the actual rotor used by Dabiri (diameter: 1.2 m). Their CFD analysis showed that a pair of closely spaced counter-rotating VAWTs produced more average power than a single turbine. De Tavernier et al. [6] conducted a 2-D numerical analysis based on the panel vortex method to investigate the effects of rotor loading, rotor spacing, and wind direction on the flow field and output of a pair of VAWTs (each rotor diameter was 20 m) with close spacing. Hezaveh et al. [7] performed a three-dimensional (3-D) CFD analysis using large eddy simulation (LES) on a WF consisting of three small VAWTs as a unit. More recent research using numerical calculations for VAWT clusters has also been conducted. Sahebzadeh et al. [8] conducted a high-fidelity CFD analysis of a pair of VAWTs with one blade for each turbine, and Hansen et al. [9] performed a 2-D CFD analysis of a pair of three-blade VAWTs with a diameter of 20 m and a three-unit arrangement. In each study, the layout conditions and power-increase ratios when a WF or cluster produced more average power than a single VAWT were investigated in detail. Silva et al. [10] conducted a 2-D CFD analysis of the arrangements of three VAWTs (diameter: 3 m, three-bladed) and showed that when the three turbines were arranged in a straight line, the power output was improved compared to other arrangements that were not in a straight line. Xia et al. [11] attempted a 3-D CFD analysis of a pair of nautilus-shaped drag-type VAWTs. They also conducted simple experiments with 3-D printing models and proposed several layouts for WFs. Recently, wind tunnel experiments on clusters of VAWTs have also been conducted by several research groups. Ahmadi-Baloutaki et al. [12] performed wind tunnel experiments on configurations of two or three small VAWTs with five blades and a diameter of 0.3 m, and investigated the optimal spacing of turbines in the trio arrangements where the third turbine was located midway downstream of the counter-rotating pair. Vergaerde et al. [13,14] conducted wind tunnel experiments on a pair of small VAWTs (two blades and a diameter of 0.5 m) to confirm that the power increase was consistent with numerical predictions. They also observed the synchronization of a pair of VAWTs. Su et al. [15] used a model of an H-type VAWT with a diameter of 0.24 m to conduct wind tunnel experiments on the output of two- and three-unit clusters. Jodai et al. [16] applied miniature butterfly-type VAWT models (diameter: 50 mm) with three blades to wind tunnel experiments for six-rotor configurations consisting of a VAWT pair or trio as the unit (cluster). The experiments demonstrated the possibility of increasing the output per unit installation area by reducing the cluster spacing.
Regarding the optimal placement of wind turbines in a WF, Mosetti et al. [17] conducted research on the placement of HAWT WFs in 1994 using a genetic algorithm (GA), which is a heuristic optimization method. Numerous studies have been conducted on HAWT WFs (Gonzales et al. [18], Wagner et al. [19], Gao et al. [20,21], Sun et al. [22]). For example, Chen et al. [23] developed a 3-D greedy algorithm for optimizing the WF of HAWTs and claimed that their method has a lower computational cost than the GA. Gualtieri [24] considered the optimal HAWT configuration and optimal HAWT specifications, inputting data from 200 land-based commercial HAWTs, and used a self-organizing map (SOM) to optimize the layout of a 2 × 2 km square WF. In their study, the Jensen model was used as the wake model for HAWT. To reduce the computational cost of CFD analysis of the wake of HAWTs, the actuator disk model (ADM) and actuator line model (ALM) have been studied, which can simulate a flow field that is closer to reality than the simple Jensen model [25]. ADM regards the wind turbine rotor as a disk and represents the disk as a number of discrete points, whereas ALM represents a blade as a linear combination of discrete points and rotates some blades to simulate rotor rotation. Both ADM and ALM typically use the aerodynamic data of the blade to calculate the aerodynamic forces corresponding to the local wind speed at each discrete point, which are then introduced into the CFD analysis as external forces acting on the fluid. Recently, Malecha and Dsouza [26] proposed an ADM that provides power and thrust coefficients as functions of wind speed without using aerodynamic data.
The wake modeling of wind turbines has also progressed for HAWTs. Göçmen et al. [27] described six widely used wake models developed at the Technical University of Denmark, including the simplest and most widely used Jensen model. They described four wake-effect superposition approaches (geometric sum, linear sum, energy balance, and quadratic sum) that considered the effects of other turbines located upstream of a turbine. In a recent study on HAWT wake models, Shao et al. [28] proposed a correction to the energy conservation sum when using the Jensen model to consider the increased turbulence when the distance to the upstream turbine is short. However, there have been few studies on the optimization of a VAWT-WF and wake models. Chen and Agarwal [29] used a GA to compare the optimization of both HAWT-WFs and VAWT-WFs. They used the simple Jensen model as the wake model for VAWTs and stated that their research was a preliminary attempt. Lam and Peng [30] developed a biased wake model for a five-bladed H-shaped VAWT model (diameter: 0.3 m, height: 0.3 m) through wind tunnel experiments, and presented the results of the optimization of the virtual VAWT-WFs (diameter: 3.0 m, rated power: 3.0 kW) using a covariance matrix adaptive evolution strategy (CMA-ES). Talamalek et al. [31] investigated the effects of turbulence on the wake of a small VAWT pair (two-bladed, diameter: 0.5 m) by conducting wind tunnel experiments, and found that turbulence is beneficial for improving the power output and wake recovery of the VAWT pair. Cazzaro et al. [32] performed a heuristic optimization based on a variable neighborhood search (VNS) implemented with a Gaussian wake model with bias for the Troposkien VAWT. The optimal arrangement of the WF, including 111 VAWTs (with different rotational directions), was reported.
The flow field around a VAWT differs from that around an HAWT. Not only is there a bias (or shift) in the VAWT wake owing to rotation, but an increase in the flow velocity is also observed on both sides of the VAWT. The size and location of the velocity-increased regions change because of the induced velocity associated with rotation. Buranarote et al. [33] attempted to develop a method to mimic the flow field obtained from the 2-D CFD analysis of a VAWT. They used the potential flow as a base flow and added an equation representing the velocity-increased regions of a VAWT to a wake model (Shapiro et al. [34]) of an HAWT. The method successfully reproduced the wind-direction dependence of the output of pairs or trios of VAWTs with a certain degree of accuracy; however, the number of parameters was very large, the program became complex, and the calculation costs were somewhat high, although not as high as those of CFD.
In this study, we fundamentally change the method proposed by Buranarote et al., and propose a new method that significantly reduces the number of parameters by directly using the numerical data of the flow field obtained from a CFD analysis of an isolated single 2-D VAWT rotor and represents the WF flow field by superposing the CFD-based flow velocity data. The possibilities of the new method to significantly reduce the calculation time and improve the prediction accuracy compared to the previous method are shown. Furthermore, this method was applied to four layouts of virtual WFs consisting of 16 miniature 2-D VAWT rotors, and the characteristics and possibilities of closely spaced VAWT arrangements are demonstrated.

2. Methods

First, we provide an overview of the calculation method for the WF flow field using the method proposed in Section 2.1. In Section 2.2, we explain the model representing the interference effects between multiple VAWTs and the method for determining the characteristics of each wind turbine. In Section 2.3, the processes used to build a flow field are described.

2.1. Calculation of WF Velocity Field Using Single-VAWT CFD Data

This study aims to predict the flow field and output power of a WF consisting of multiple VAWTs installed at a site in a certain area. However, for simplicity, a 2-D flow field is assumed. In the previous method developed by Buranarote et al. [33], to reproduce the flow field around a 2-D VAWT, which is equivalent to the results obtained by CFD analysis, a unique function representing the velocity-increased regions existing at both sides of the VAWT was added to the wake model (super-Gaussian function) of Shapiro et al. [34]. This model simulates the transformation in the distribution of the velocity deficit from a top-hat type to a Gaussian type in the near-wake region. Our new method does not use functions to represent the velocity field (velocity deficit) around an isolated single VAWT but directly uses the numerical flow velocity data obtained from a CFD analysis. However, the CFD analysis is performed as an unsteady calculation (Hara et al. [35]) in advance under the condition of a specific constant upstream wind speed U∞0, and the results are averaged to obtain the 2-D velocity field (uCFD, vCFD) as a steady field. Figure 1 illustrates the proposed method. In the example in Figure 1, the white circles represent the 2-D VAWT rotors, and all the rotors rotate in the counterclockwise direction. The coordinate system has an x-axis in the mainstream direction and a y-axis perpendicular to the mainstream direction. Two color maps in Figure 1 show the distributions of the u and v components of flow velocity near a single miniature VAWT rotor (diameter: 50 mm) obtained by CFD analysis under the condition U∞0 = 10 m/s. The upstream wind speed U∞0 when performing CFD analysis does not have to be the same as the upstream wind speed U of any WF being calculated. If there are CFD analysis data for a specific wind speed, the data can be used to calculate the WF under any upstream wind speed condition. In addition, the effects of the other rotors in the WF on the surrounding flow field of the rotor are calculated based on the prepared CFD data of a single isolated VAWT.
The rotor diameter of the VAWT to be calculated is assumed to be D. The velocity data of the prepared CFD results around a single isolated VAWT are defined as ( u C F D ,   v C F D ). When the coordinate origin in the CFD analysis is set at the center of the VAWT rotor and the normalized coordinate system divided by the rotor diameter is expressed as ( x n 0 ,   y n 0 ), x n 0 and y n 0 are defined by Equations (1) and (2), respectively:
x n 0 = x D
y n 0 = y D
The normalized velocity deficit field ( d u d i f ,   d v d i f ) generated around a single isolated VAWT is defined by Equations (3) and (4) when the upstream wind speed is U∞0.
d u d i f x n 0 ,   y n 0 = 1 U 0 u C F D x n 0 ,   y n 0 U 0
d v d i f x n 0 ,   y n 0 = v C F D x n 0 ,   y n 0 U 0
In this study, the entire computational domain, including the WF area, was assumed to be a square with side L. The origin of the coordinate system for the WF calculation was placed at the center of the entire computational domain, and the size of the computational domain was given by the parameter ROC = L/(2D), which is the length from the center to the boundary divided by the turbine diameter. The computational grid is divided into m parts, vertically and horizontally. The grid cell size Celln_D, which is nondimensionalized by the wind turbine diameter, is defined as follows:
C e l l n _ D = L m D = 2 R O C m
The CFD data (as input data) used in the calculation were limited to those in the ranges shown in Equations (6) and (7). Figure 2 shows the area of the CFD data to be input using a color map (only the u component is shown as an example).
2 x n 0   10
8 y n 0   8
Figure 2. Extrapolation regions [A] to [H].
Figure 2. Extrapolation regions [A] to [H].
Energies 18 00220 g002
In the proposed method, when the center of a single rotor or any rotor in the WF is set as the origin of the local coordinate system, the normalized velocity deficit field at any point P(xnp, ynp) existing in the area from [A] to [H] outside the input range of the CFD data in Figure 2 is calculated by extrapolating from the value at point B(xnb, ynb) on the boundary of the input range of the CFD data using Equations (8) and (9): in Equations (8) and (9), disPB is the distance between a point P and a point B, and τout is a parameter that represents the rate of decay of the normalized velocity deficit.
d u d i f x n p ,   y n p = d u d i f x n b ,   y n b × e d i s P B / τ o u t
d v d i f x n p ,   y n p = d v d i f x n b ,   y n b × e d i s P B / τ o u t
Figure 2 shows three examples (points P, P′, and P″) where point P exists in different regions. When a point P is in the regions [A] to [D] shown in Equations (10)–(13) (e.g., point P or point P′ in Figure 2), point B is a point on the boundary of the input range of the CFD data (e.g., point B or point B′ in Figure 2).
region   A :   x n 0 < 2   and 8 y n 0   8
region   B :   x n 0 > 10   and 8 y n 0   8
region   C : 2 x n 0   10   and   y n 0 < 8
region   D : 2 x n 0   10   and   y n 0 > 8
When a point P is in the regions [E] to [H] shown in Equations (14)–(17) (e.g., point P′′ in Figure 2), point B is a point on the boundary of region [C] or [D] (e.g., point B″ in Figure 2).
region   E :   x n 0 < 2   and   y n 0 < 8
region   F :   x n 0 > 10   and   y n 0 < 8
region   G :   x n 0 < 2   and   y n 0 > 8
region   H :   x n 0 > 10   and   y n 0 > 8
Assuming that N VAWTs of the same size exist in a WF, the proposed method assumes that the flow velocity components u and v at any position (x, y) in the WF can be calculated using Equations (18) and (19), respectively:
u x ,   y = U + j = 1 N U F ( j ) d u d i f x n ( j ) ,   y n ( j )
v x ,   y = j = 1 N U F ( j ) d v d i f x n ( j ) ,   y n ( j )
where U is the upstream wind speed and UF(j) is the virtual upstream wind speed of the j-th VAWT located at position (xj, yj). The decision on the value of UF(j) is explained in the following section. The arguments for the normalized velocity deficits xn(j) and yn(j) are given by Equations (20) and (21), respectively.
x n ( j ) = x x j D
y n ( j ) = y y j D

2.2. Interference Model and Wind Turbine Characteristics

In the proposed method, the virtual upstream wind speed UF(j) of each VAWT is predicted using Equation (22).
U F j = u a v e j + u 1 j u 2 ( j )  
where uave(j) is the u component of the average wind speed at the position where the j-th VAWT is installed before installation. As for the VAWTs other than the subject VAWT, some were installed and others were not, and their situations differed in the calculation process. In this study, uave(j) is defined as the average of the u component at 11 points on a line segment of the same length as the diameter in the y direction, which includes the center of the j-th VAWT. ∆u1(j) and ∆u2(j) in Equation (22) represent the influence of the surrounding VAWTs on the j-th VAWT and are defined in Equations (23) and (24), respectively.
u 1 j = k c 1 E s s S r d ( j ) k u d u d y 1 S r d ( j ) k v v a v e ( j )
u 2 j = u a v e j × e v a v e ( j ) k 1 + d m i n k 2 D
vave(j) in the above equations is the v component of the average wind speed at the position where the j-th VAWT is to be installed before installation. In this study, vave(j) is defined as the average of the v component at 11 points on a line segment of the same length as the diameter in the y direction, which includes the center of the j-th VAWT.
u1(j) in Equation (23) is introduced as the increment in the virtual upstream wind speed UF(j) to represent the resulting increase in the power output of the j-th VAWT due to the effects of other VAWTs in the vicinity, which cause a velocity gradient dudy or a secondary flow component vave(j) at the position where the j-th VAWT is installed. In this study, wind turbines with a large solidity σ are considered to be of the drag type, and those with a small solidity σ are considered to be of the lift type. Figure 3 illustrates the predicted effects of the velocity gradient du/dy and the secondary flow component v on the output power of each type of wind turbine.
As shown in Figure 3, the drag- and lift-type outputs are expected to respond differently to the velocity gradient, whereas it is assumed that both types of wind turbines provide the same response to secondary flow. In Equation (23), the effects of the velocity gradient and secondary flow are not treated independently, but are given as the product of both, because the results of a preliminary comparison with the CFD analysis of a pair of VAWTs, described below, showed that Equation (23) was better than the case in which the two effects were treated independently. Although Figure 3 shows only the cases of counterclockwise (CCW) rotation, the parameter Srd(j), which represents the rotation direction of the j-th VAWT and is defined by Equations (25) and (26), is introduced to properly handle clockwise (CW) rotation.
S r d j = 1   for   CCW   rotation
S r d j = 1   for   CCW   rotation
The results of the CFD analyses previously conducted in our laboratory on four differently sized rotor pairs of VAWTs with different solidities are shown in Figure 4. Table 1 summarizes the specifications and calculation conditions of the four cases (L, ML, M, and S).
Although the details are not shown, Figure 4b shows the solidity dependence of the three types of paired VAWT arrangements (counter-down (CD), counter-up (CU), and co-rotation (CO)) shown in Figure 4a, with the vertical axis representing the output power averaged between the two rotors (Pn_ave), which is normalized by dividing by an isolated single-rotor power. The horizontal axis represents the solidity (σ = Bc/(πD)) of each VAWT rotor with a logarithmic coordinate axis. CFD analyses were performed only under the condition of gap/D = 0.2. As shown in Figure 4b, when the solidity is large (S-size), the output of the CD arrangement is larger than that of the CU arrangement, and the CO arrangement is biased but lies between the CD and CU arrangements. However, when the solidity was small (L-size), the output of the CU arrangement was larger than that of the CD arrangement, indicating the opposite tendency. This tendency has also been shown in the results of numerical analyses conducted by other researchers. For example, in the study by Zanforlin et al. [5], CD > CU results were obtained for a VAWT pair with high solidity (B = 3, c = 0.128 m, D = 1.2 m, σ = 0.102), whereas, in the research of De Tavernier et al. [6], CU > CD results were obtained for a VAWT pair with low solidity (B = 2, c = 1 m, D = 20 m, σ = 0.0318) under the condition of gap/D = 0.2.
Referring to Figure 4b, the reversal of the CD and CU outputs occurs around the solidity (σ = 0.057) of the ML-size VAWT pair. Therefore, in the present model, we define a parameter Ess in Equation (27) that expresses the effects of solidity, including the reversal of the CD-CU output power. This parameter is multiplied by a factor indicating the effects of the velocity gradient in ∆u1(j) of Equation (23), which is expected to produce different responses between the drag type and lift type.
E s s = σ 0.057 0.325
kc, ku, and kv in Equation (23) are the fitting parameters that were determined in this study by comparison with a previous CFD analysis of a pair of VAWTs (Hara et al. [35]).
Next, ∆u2(j) in Equation (24) is explained. ∆u2(j) is a correction introduced from a comparison of a CFD analysis of paired VAWTs (Hara et al. [35]) and the predictions for 16 wind directions by the present method. In the preliminary comparison, it was found that when a VAWT was located diagonally behind another VAWT located upstream, the method tended to predict a higher output power than the CFD analysis. Because the state of overprediction corresponded to a region where the absolute value of the secondary velocity was small to some extent, the virtual upstream wind speed was reduced by the amount calculated using Equation (24). Note that dmin in Equation (24) is the distance to the nearest VAWT upstream of the j-th VAWT (center-to-center), and if this value is large, the amount of correction by ∆u2(j) will be small. k1 and k2 in Equation (24) are the fitting parameters determined by comparison with the CFD analysis for a pair of VAWTs.
Induced velocities occur around the VAWTs owing to the rotation of the rotors. Consequently, the wake of the VAWT shifts in a direction perpendicular to the main flow. The CFD analysis results (input data) for a single isolated VAWT include this shift; however, the amount of wake shift changes owing to the influence of the other VAWTs. To consider this effect, in the proposed method, argument yn(j) of the normalized velocity deficit for the j-th VAWT introduced in Equation (21) is replaced by Equation (28).
y n ( j ) = y y j k s v a v e ( j ) 1 e x n ( j ) D       ( x n j > 0 )
The above equation is applied only to the wake region where xn(j) is positive, and Equation (21) is used upstream. Equation (28) assumes that a shift proportional to the secondary flow component vave(j) at the center of the j-th VAWT occurs in the wake region; however, an exponential function is introduced to prevent a discontinuous shift near the rotor. ks is a fitting parameter that adjusts the amount of the shift.
In this method, once the virtual upstream wind speed UF(j) is obtained using Equation (22), it is converted into the characteristics of the j-th VAWT in the WF based on the pre-entered power curve and the relationship between the operating rotor speed and wind speed to obtain the prediction results. At this time, the input characteristics, such as the power curve, may be experimental data instead of theoretically predicted values. In this study, we focused on the 2-D rotor of a miniature VAWT with a diameter of 50 mm, for which we previously conducted a CFD analysis. Therefore, the input power curve and the relationship between the rotation speed and wind speed are given as a curve and a straight line passing through the power (Ps_CFD = 177 mW) and rotation speed (Ns_CFD = 3483 rpm) obtained in a previous CFD analysis of an isolated single 2-D rotor under the condition of a wind speed U= 10 m/s. Figure 5 shows the power curve and the relationship between the rotation speed and wind speed, which were used as input data in this study.
The target wind turbine in this study is the same as that assumed in reference [35], and the torque characteristics have been obtained as the results of CFD analysis for four upstream wind speeds U = 6, 8, 10, and 12 m/s, as shown in Figure 2 of reference [35] (see Appendix A). In this study, the load torque characteristics are defined as a curve that is proportional to the square of the angular velocity ω passing through the state of 95% of the maximum output at U = 10 m/s, and are given by Equation (1) in reference [35] or Equation (A1) in Appendix A. The rotation speed Ns [rpm] and power Ps [mW] of the isolated single wind turbine shown in Figure 5 are approximated by Equations (29) and (30), respectively, as functions of the upstream wind speed U.
N s = 376.34 U 280.58   [ rpm ]
P s = 0.2459 U 3 0.8344 U 2 + 1.4757 U 0.2466   [ mW ]
From Equation (29), when the wind speed is below 0.7455 m/s, the rotational torque of the single turbine obtained by the CFD analysis is smaller than the load torque. For simplicity, in this study, it is assumed that the target wind turbine cannot rotate at wind speeds below 2 m/s, and the input power curve data are limited to U = 2 m/s or more.
The proposed method was developed as an engineering model with low computational costs that can be applied to explore the optimal layout of VAWTs. Therefore, ideally, it is desirable to fix the upstream wind speed for the prepared CFD analysis to a single value of the rated wind speed of the target wind turbine or the average wind speed of the site (U = 10 m/s in this study). The extent to which this method remains reliable when the upstream wind speed changes must be confirmed by conducting additional CFD analyses for various configurations; however, this is a topic for future work.

2.3. Flow Field Construction Process

In the previous section, we explained the outline of the analytical formula used in the calculations of the proposed method; however, the calculation procedure affected the results. In this method, the flow field of a WF with multiple VAWTs was constructed by placing the turbines individually in the flow field, starting with the furthest upstream and gradually constructing the entire flow field. Figure 6 illustrates the case in which the total number of VAWTs in the WF is N = 4. As shown in Figure 6a, initially, only one VAWT is placed at the most upstream position (temporary number of VAWTs: Ntemp = 1), and a flow field proportional to the CFD analysis obtained by substituting U into UF(1) in Equations (18) and (19) is assumed. Using the assumed flow field, the average wind speed components (uave, vave) at the location where the second VAWT from the upstream was to be installed were calculated, and the virtual upstream wind speed UF(2) was obtained from Equation (22). Once UF(2) is determined, temporarily assuming N = Ntemp = 2 in Equations (18) and (19), the flow field, including the two VAWTs in the WF, is calculated by superimposing the effects of each VAWT. Using the resultant flow field, the average wind speed (uave, vave) at the third VAWT position upstream was calculated, and the virtual upstream wind speed UF(3) was obtained using Equation (22). Thus, the number of VAWTs was gradually increased to determine the virtual upstream wind speed, UF(j), for all the wind turbines. Finally, using the power curve data shown in Figure 5, all UF(j) were converted into the output power and rotation speed of each VAWT to obtain a prediction result.
Although the process of building the flow field is described in Figure 6, in the case where some VAWTs have the same x coordinates, there is a degree of freedom in the ordering from the upstream side. In addition, the effects of the downstream turbines on the upstream turbines must be considered. Therefore, in the proposed method, although the calculation cost increases, once all virtual upstream wind velocities UF(j) for the temporary wind turbine number (Ntemp) have been obtained for the first time, several iterations of the calculation routines, as shown in the example in Figure 7, are inserted, and the calculations are repeated until the flow field for the temporary number state of the VAWTs converges. Figure 7 shows the case in which Ntemp = 3. After the virtual upstream wind speed UF(3) is calculated for the first three VAWTs from upstream, each VAWT turbine is removed individually from upstream, and the virtual upstream wind speed UF(j) (j = 1, 2, 3) is recalculated. This recalculation process was performed for all intermediate states, and the flow field was gradually constructed. Although this iterative calculation is a drawback of the proposed method, it is important for constructing a reliable flow field.

3. Results and Discussion

In this section, we apply the proposed method (the new model) to specific clusters of VAWTs and demonstrate their accuracy and features. Moreover, we apply the method to a WF consisting of many miniature VAWT rotors in specific arrangements to illustrate the expected output power of the WF. In Section 3.1, we compare the results of a previous CFD analysis with the predictions calculated using the new model regarding the 16-wind-direction dependence of the output of paired VAWTs. In Section 3.2, the CFD analysis results of a four-turbine cluster are compared with those of the proposed method. In Section 3.3, the method is applied to four layouts of a virtual WF consisting of 16 VAWTs, and the annual power generation for each layout is forecasted. Section 3.4 describes the computational cost of the proposed method.

3.1. A Comparison of Both the Results of CFD and the Proposed Method on Paired VAWTs

Figure 8a,b show the flow field calculated by applying the proposed method to a pair of miniature 2-D VAWTs with each diameter of D = 50 mm rotating in the counterclockwise direction in a CO configuration. The miniature VAWT considered in this study had three blades with an NACA 0018 airfoil cross section and the S-size specifications listed in Table 1. The performance of an isolated single turbine was assumed to be a power curve, as shown in Figure 5. In Figure 8, the upstream wind speed is U = 10 m/s, and the gap length between two rotors is gap = 25 mm (gap/D = 0.5). The angle θ, which represents the wind direction, is defined in Figure 8a. In the calculation program, the mainstream always flows from left to right, and a condition with different wind direction is created by rotating the arrangement of VAWTs counterclockwise by an angle θ. Figure 8a shows the state where θ = 22.5°, and Figure 8b shows the state where θ = 337.5° (−22.5°). In Figure 8a,b, the distribution of the u component of the flow velocity is shown only in the vicinity of the paired VAWTs. However, the entire calculation domain is a square region of L × L = 1 × 1 m. The number of grids in the vertical and horizontal directions was m = 400 and the normalized grid size was Celln_D = 0.05. Figure 8c compares the variation in the output power of Rotor 1 (R1) when the wind direction was changed to 16 directions with the results of a previous CFD analysis (Hara et al. [35]). The radar graph shows the normalized output PR1 of R1 divided by the output of a single wind turbine PSI = 177 mW (at U = 10 m/s). Similarly, Figure 8d shows the 16-direction dependence of the normalized power (Pave/PSI) of the average power Pave for R1 and R2 and is compared with previous CFD analysis results. As shown in Figure 8c, the increase in the output power of Rotor 1 at θ = 180° by the induced velocity caused by Rotor 2 and the decrease at θ = 270° resulting from the wake effects of Rotor 2 are also predicted by the new method.
Figure 9 shows the results of performing the same calculations as in Figure 8 for an inverse-rotation (IR) arrangement in which the rotation directions are opposite to each other. However, Figure 9a is for the CD arrangement with θ = 0°, and Figure 9b corresponds to the CU arrangement with θ = 180°.
In this study, the fitting parameters in the proposed method were determined mainly by comparison with the CFD analysis for the 16-wind-direction output distribution shown in Figure 8 and Figure 9. In all calculations presented in this paper, we set kc = 0.000103, ku = 5.4, kv = 10.4, ks = 0.012, k1 = 1.2, and k2 = 0.64. The τout introduced in Equations (8) and (9) does not have any effect when the VAWTs are close to each other in a small area, as in Figure 8 and Figure 9. In this study, the value of τout was determined to be 10 from the results of the tandem arrangement of four VAWTs described later. The iterative calculation routine performed in each state, consisting of a temporary number of VAWTs described in Figure 7, was fixed at three times, at which convergence was confirmed.
As shown in Figure 8c,d and Figure 9c,d, the degree of agreement (or, in other words, the error) between the CFD analysis results and predictions using the new model varies depending on the wind direction. Even if the values of the aforementioned fitting parameters are fixed, the degree of match varies depending on the arrangement of the target VAWT pairs, gap length, the size of the entire calculation domain, and normalized grid size (resolution). Therefore, in this study, the difference between results by CFD analysis and the new model for the output of R1 in the wind direction θi is defined as Equation (31), and the average value of the difference ∆Pn_R1(θi) over 16 wind directions is defined as Equation (32), and the degree of agreement (error) is quantified and compared using those values. Similarly, the difference between the CFD and our method of the average power Pave of R1 and R2 for each wind direction θi and the average value over the 16 wind directions is defined by Equations (33) and (34), respectively.
P n _ R 1 θ i = P R 1 θ i P R 1 _ C F D θ i P S I  
E r r n _ R 1 = 1 16 i = 1 16 P n _ R 1 θ i × 100   %
P n _ a v e θ i = P a v e θ i P a v e _ C F D θ i P S I  
E r r n _ a v e = 1 16 i = 1 16 P n _ a v e θ i × 100   %
The differences Errn_R1 and Errn_ave between the CFD and model predictions defined in Equations (32) and (34) when the grid number is m = 400 are shown in Figure 10a,b. Six cases are shown with gap lengths of 25, 50, and 100 mm in the CO and IR configurations. For each case, the normalized cell size, Celln_D, was set to 0.05, 0.1, 0.2, 0.4, and 0.75 (corresponding to an ROC of 10, 20, 40, 80, and 150), as shown in the color-coded bar graphs. Similarly to Figure 10, the values of Errn_R1 and Errn_ave when the number of grids is m = 600 are shown in Figure 11. The normalized cell sizes, Celln_D, were set to 0.0333, 0.0667, 0.1333, 0.2667, and 0.5 (corresponding to ROC = 10, 20, 40, 80, and 150).
Figure 10 and Figure 11 show that the normalized errors, Errn_R1 and Errn_ave, decrease as the gap increases. Errn_ave is smaller than Errn_R1 because the normalized error of the output of R1 or R2 alone is partially canceled out owing to the symmetry in the average error of the two rotors. Although this was not observed in all cases, there was a tendency for errors to increase as the normalized cell size increased. In addition, there was little difference in the trends owing to the differences in the number of grids m.

3.2. Comparison of the Predictions by CFD and the Proposed Model on a Four-VAWT Cluster

Previous predictions by CFD analysis (Buranarote et al. [33]) of a four-VAWT cluster, in which all the rotors rotated in the same direction and were arranged in line with a gap of 150 mm (gap/D = 3), were compared with predictions using the proposed method. In subsequent model calculations, we set m = 400 and Celln_D = 0.1 (ROC = 20). Figure 12a is the same as Figure 20a in the paper by Buranarote et al. and shows the unsteady flow field (u-component distribution) obtained by CFD analysis. The upstream wind speed is U = 10 m/s. The values shown in the figure for PR1, PR2, PR3, and PR4 are the predicted output power values for the VAWT rotors (mW). The flow field of the four VAWTs arranged in parallel, as shown in Figure 12b, was predicted using the proposed model. The VAWT output predicted by the model was in good agreement with the CFD analysis, and the errors, which were calculated as the difference between the predicted outputs of the CFD and the model divided by the output of an isolated single VAWT (PSI), were 0.73%, 2.1%, 1.1%, and 8.6% for R1 to R4, respectively.
Figure 13 shows a similar comparison for a four-VAWT tandem arrangement, where Figure 13a is the same as Figure 20b in Buranarote et al. [33]. The difference between the CFD and model predictions for the tandem arrangement was somewhat larger; however, the differences based on the output PSI of a single VAWT were 0.11%, 15.5%, 3.8%, and 5.1% for R1 to R4, respectively. The error for R1, the most upstream VAWT, is small because, as mentioned above, the value of τout was determined so that the output of the lead turbine in this tandem arrangement of four VAWTs matches between the CFD and the model predictions.

3.3. Application to Virtual 16-VAWT Wind Farms

3.3.1. Flow Field Predictions and WF Output Dependence on 16 Wind Directions

In this section, we apply the method to a virtual WF consisting of 16 miniature 2-D VAWTs. The area where the VAWTs were placed was assumed to be a square with dimensions of 500 × 500 mm. The following four layouts of VAWTs were assumed:
  • CO–4 × 4 layout
  • CO–8–pairs layout
  • IR–8–pairs layout
  • CO–8 × 2 layout
The flow field and output of each VAWT obtained when the wind direction was θi = 0° are shown in Figure 14a, Figure 15a, Figure 16a and Figure 17a, in which the details of each layout are also illustrated. The upstream wind speed was assumed to be U = 10 m/s. Figure 14b, Figure 15b, Figure 16b and Figure 17b show the distribution of the average power of the WF, which was normalized by dividing the WF average power P16ave over 16 wind directions by an isolated single VAWT power PSI. For all layouts, the entire computational domain was a square of L × L = 2 × 2 m, and the values of the other parameters were the same as those in the four-VAWT case described in Section 3.2.
Comparing the results of four layouts, only in the case of the CO–8 × 2 layout of Figure 17 did the WF average power P16ave exceed the single VAWT power PSI in the wind-direction range of –22.5° ≤ θi ≤ 22.5° or 157.5° ≤ θi ≤ 202.5°. For the other three layouts, P16ave < PSI for all wind directions. In the case of Figure 17a, the eight VAWTs are arranged in a line perpendicular to the mainstream (parallel arrangement) with very narrow spacing (gap1/D = 0.286), so many of the turbines (actually, half the number) in the WF are not in the wake, and the effects of increasing the output by the induced velocity are also significant. In particular, R1 turbine, which is located at the bottom end of the upwind parallel arrangement, had an exceptionally high output. A similar power output improvement was reported by Mereu et al. [36] in a CFD analysis of a parallel arrangement of eight Savonius turbines; although the turbine spacing dependence differs owing to different solidities, their study and the results of our study may be due to the same effects.

3.3.2. Annual Energy Production Prediction

The proposed method assumes the existence of 2-D flow field data for an isolated single VAWT analyzed by CFD at a specific wind speed U∞0. Assuming the availability of the measurement data of wind-direction occurrence probability (WDP), the probability of obtaining wind direction in a certain wind direction θi among the 16 direction segments is defined as WDP(θi) in this study. For simplicity, we assume that the wind speed occurrence rate is independent of wind direction and is given by a Rayleigh distribution; the probability density function of any wind speed V according to the Rayleigh distribution is expressed as fR(V). Once the average output Pave(U∞0, θi) [W] of a WF consisting of N VAWTs at a specific wind speed U∞0 and a wind direction θi is obtained by the proposed method, the annual energy production (AEP), AEP(U∞0) [kWh], in the bin at the specific wind speed U∞0 when the wind speed bin width is ∆V = 1 m/s, is calculated using Equation (35). The constant 8.76 is 60 × 60 × 24 × 365/(3.6 × 106).
A E P U 0 = i = 1 i = 16 P a v e U 0 ,   θ i × N × W D P θ i × f R U 0 × 8.76
If the cut-in wind speed is V2 = 2 m/s and the cut-out wind speed is V20 = 20 m/s, the AEP of the WF can be easily predicted using the previously calculated average power Pave (U∞0, θi) or AEP(U∞0) using Equation (36) or Equation (37) without any additional flow field calculations.
A E P = j = 2 j = 20 i = 1 i = 16 P a v e U 0 ,   θ i × P S I V j P S I U 0 × N × W D P θ i × f R V j × 8.76  
A E P = j = 2 j = 20   P S I V j P S I U 0 × f R V j f R U 0 × A E P U 0  
Table 2 shows the AEPs calculated for the four layouts of 16 VAWTs shown in Section 3.3.1, assuming a hypothetical uniform wind-direction distribution (WDP(θi) = 0.0625) where the wind-direction occurrence probability is the same in all 16 directions. The wind speed distribution was assumed to be Rayleigh with an average wind speed of 10 m/s. Table 2 lists the AEP per unit area and AEP based on the total AEP of the 16 isolated VAWTs. Even in the CO–8 × 2 layout, which had the highest AEP among the four layouts, the AEP was only 80.9% of the total AEP of the 16 isolated VAWTs.
Figure 18 shows an example of the nonuniform wind-direction occurrence probability, where the prevailing wind direction appears in the direction of 157.5°. The probability of the wind-direction occurrence was measured at the Arid Land Research Center (ALRC) of Tottori University. For the four 16-VAWT layouts shown in Section 3.3.1, in a uniform wind-direction distribution, the CO–4 × 4 layout showed the maximum average WF power in four wind directions, θi = 67.5°, 157.5°, 247.5°, 337.5° (−22.5°), the CO–8–pairs layout showed the maximum average power in two wind directions, θi = 157.5° and 337.5° (−22.5°), the IR–8–pairs layout showed the maximum average power in one wind direction, θi = 337.5° (−22.5°), and the CO–8 × 2 layout showed the maximum average power in two wind directions, θi = 0° and 180°. If the condition that maximizes the average WF output in a uniform wind-direction distribution is aligned with the prevailing wind direction in a nonuniform wind-direction distribution, the maximum AEP of the WF must be obtained. Assuming an extremely nonuniform wind-direction distribution, as shown in Figure 18, and a Rayleigh distribution with an average wind speed of 10 m/s and matching the maximum output direction of each layout of the 16 VAWTs to the prevailing wind direction, the AEPs were calculated and are tabulated in Table 3.
In all layouts, when the maximum average output direction was aligned with the prevailing wind direction, the output increased compared to the case of a uniform wind-direction distribution. By calculating the percentage increase in AEP for each layout based on the total AEP of the 16 isolated VAWTs, the results were 6.7% for CO–4 × 4, 19.1% for CO–8–pairs, 18.2% for IR–8–pairs, and 49.6% for CO–8 × 2. The CO–8 × 2 layout shows an outstanding increase rate, and as a result, it is interesting that we can expect an increase in the WF output of approximately 20% compared with 16 isolated VAWTs.

3.4. Calculation Cost of the Developed Simulation Program

Table 4 lists the computational costs of the simulation program for calculating the WF flow field and the output powers of the VAWTs using the proposed method. Four layouts were considered: one, two, four, and sixteen VAWTs. In this computational cost analysis, calculations were performed under the same conditions: ROC = 20, m = 400, and Celln_D = 0.1. The program was written in C#, and the PC used for the calculations had a 10-core Intel Core (TM) i9-10900 2.81 GHz processor. Table 4 lists the time required for the calculation of one wind direction and the time required for the continuous calculation of sixteen wind directions. However, to calculate the 16 wind directions, parallel calculations are performed using the “Parallel.For” method in the C# language. As shown in Section 2.3, in this calculation method, the flow field is constructed by increasing the number of VAWT rotors one by one, and the sub-calculation routine is repeated three times for each temporary number state of the rotors. Therefore, as shown in the calculation results for one wind direction in Table 4, the calculation time increased at a rate of approximately N(N + 1) with increasing N, the number of VAWTs. However, owing to the effect of parallel computing, the calculation time for 16 wind directions was approximately twice that for one wind direction, even in the case of 16 VAWTs.

4. Conclusions

The ultimate goal of this study was to determine the optimal layout of VAWTs for a WF. As preliminary research, we proposed a method to predict the flow field and output of the VAWT-WF in a short time. The proposed method assumes the existence of numerical data obtained by CFD analysis of the velocity field around an isolated single two-dimensional VAWT rotor. By superimposing the data of all VAWT rotors in a WF, a significant reduction in calculation time and an improvement in accuracy were achieved compared with previous algebraic wake modeling methods. The turbine-to-turbine interaction model in this method includes the effects of velocity gradient, secondary flows, and wake shift. We also proposed a parameter (Ess) to represent the effect of solidity, but further research is needed to extend this to solidities other than the miniature wind turbine (σ = 0.382) targeted in this study.
This method is also unique in the WF flow field building process. When constructing the flow field of a WF with a total VAWT number of N, the temporary number of VAWTs is gradually increased from one to N, while at each temporary rotor number state, one turbine is removed to re-calculate the virtual upstream wind speed UF of the removed turbine repeatedly. Therefore, the computational cost for a particular wind direction is approximately proportional to N (N + 1). However, the calculation time for 16 wind directions was reduced by parallelizing the calculations.
Comparing the previous CFD analysis of a pair of VAWTs with the results predicted by the proposed method, the normalized error tended to decrease as the gap length between the two turbines increased. In addition, the error tended to increase as the normalized cell size increased; however, there was little difference in the trends owing to the differences in the number of grids.
A four-VAWT linear arrangement (parallel and tandem) was also compared between the CFD and model analyses, and showed relatively good agreement.
Current model parameter values vary depending on the size, structure, and solidity of the wind turbine. If an operation curve is determined for a single wind turbine, the model parameters can be obtained by conducting a CFD analysis for a minimum of 17 cases, including eight wind directions for each (from which the power dependence of each rotor pair on the 16 wind directions is derived from symmetry) of representative closely spaced rotor pairs of CO and IR arrangements (e.g., gap = 0.5D) and one wind direction for a representative tandem layout (e.g., four rotors with gap = 3D). In the future, if a CFD analysis of various wind turbines is performed to clarify the dependence of each parameter on the turbine size and solidity, it may be possible to express the parameter in a universal mathematical formula.
When this method was applied to four virtual WFs with 16 2-D miniature VAWTs arranged, it was found that a layout (CO–8 × 2) in which the spacing between neighboring turbines (gap1/D = 0.286) was shortened and eight turbines were arranged in a line in two rows that were sufficiently separated (gap2/D = 8) showed significantly better performance than the other layouts. This could be attributed to the fact that there were many wind turbines that were not in the wake, and the close spacing resulted in a significant increase in the output owing to the induced velocity on each turbine.
In the case of the CO–8 × 2 layout of the 2-D VAWTs with large solidities discussed in this paper, an approximately 20% increase in AEP could be expected compared to the total AEP of isolated turbines by matching the condition that maximizes the average WF output to the prevailing wind direction in a nonuniform wind-direction distribution.
It may be difficult to obtain results similar to those described above for practical-sized lift-type VAWTs because of their low solidity. The optimal layout for lift-type VAWTs should be investigated in the future by combining the proposed method with a GA and other methods.

Author Contributions

Conceptualization, Y.H.; methodology, Y.H.; software, Y.H.; validation, M.S.M. and A.I.; formal analysis, Y.H.; investigation, Y.H., M.S.M. and A.I.; resources, Y.H.; data curation, Y.H.; writing—original draft preparation, Y.H.; writing—review and editing, M.S.M., A.I. and Y.J.; visualization, Y.H.; supervision, Y.H.; project administration, Y.H.; funding acquisition, Y.H. and Y.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research was conducted jointly at Tottori University and Nikkeikin Aluminum Core Technology Co., Ltd. and was supported by JSPS KAKENHI Grant Number JP22K12456.

Data Availability Statement

The data presented in this study are available upon request from the corresponding author.

Acknowledgments

The CFD analysis investigating the solidity dependence of the power output of the pairs of 2-D VAWTs shown in Table 1 and Figure 4 in this paper was primarily carried out by Tomoyuki OKINAGA and Taichi MATSUDA, who are acknowledged for their work.

Conflicts of Interest

The authors declare no conflicts of interest.

Appendix A

In this study, the CFD analysis results for a vertical-axis miniature rotor with diameter D = 50 mm studied in references [33,35] were used. Appendix A presents an overview of the CFD analysis of the miniature rotor. Figure A1a shows a 3-D CAD image of a miniature butterfly wind turbine model for wind tunnel testing, as described in [16]. The present computational study targeted a two-dimensional rotor equivalent to the equatorial cross section of the experimental turbine model (Figure A1b). However, an 18 mm diameter hub was not considered in the 2-D rotor. Three blades were used, and the cross section was a symmetrical airfoil NACA 0018. The chord length was set as c = 20 mm.
Figure A1. (a) A 3-D model of a miniature butterfly wind turbine. (b) The 2-D rotor used for 2-D CFD analysis.
Figure A1. (a) A 3-D model of a miniature butterfly wind turbine. (b) The 2-D rotor used for 2-D CFD analysis.
Energies 18 00220 g0a1
STAR-CCM+ was used for the CFD analysis. Figure A2 shows the computational domain and mesh state generated in the CFD analysis used to determine the wind turbine characteristics. The 2-D rotor was installed at 20D from the inlet boundary, and the distance from the wind turbine rotor to the outlet boundary was 30D. The distance between the rotor and the side boundary was 20D. A rotational region was created using the overset mesh method to realize the rotation of the rotor in the computational domain. As boundary conditions, a constant velocity U was assumed at the inlet boundary, and the pressure at the outlet boundary was held to be 0 [Pa] as the gauge pressure. A slip condition was applied to the side boundaries. The fluid was air, and was assumed to be incompressible. For the CFD analysis, the continuity equation and unsteady Reynolds-averaged Navier–Stokes equations (RANS) were solved. The SST kω turbulence model was used. Figure A3 shows an example of the analysis results for U = 10 m/s, showing the distribution of the x-direction component u of the flow velocity. Figure A4 is the same as Figure 2 in reference [35], which shows the torque characteristics of the miniature 2-D rotor for several angular velocities ω at the upstream flow speed U = 6, 8, 10, and 12 m/s. In the two-dimensional analysis using STAR-CCM+, the rotor thickness (i.e., height) was set to 1 m. Therefore, to make the torque equivalent to that of an experimental 3-D model with a rotor height of 43.4 mm, the torque value obtained from the CFD analysis was multiplied by 0.0434, as shown in Figure A4. The mass of the 3-D model is M = 0.014 kg, and the moment of inertia is I = 5.574 × 10−6 kgm2. In the CFD analysis described in Appendix B, these were set as listed in Table A2 (see the S-size column).
Figure A2. The mesh conditions for CFD analysis to investigate the torque performance of a single rotor.
Figure A2. The mesh conditions for CFD analysis to investigate the torque performance of a single rotor.
Energies 18 00220 g0a2
Figure A3. An example of the unsteady flow field around a single rotor obtained by CFD analysis.
Figure A3. An example of the unsteady flow field around a single rotor obtained by CFD analysis.
Energies 18 00220 g0a3
Figure A4. The torque performance of a miniature rotor (S-size) obtained by CFD analysis and an ideal load torque curve defined by Equation (A1).
Figure A4. The torque performance of a miniature rotor (S-size) obtained by CFD analysis and an ideal load torque curve defined by Equation (A1).
Energies 18 00220 g0a4
The curve in red shown in Figure A4 shows the load torque proportional to the square of ω, which passes through 95% of the torque at maximum output when the wind speed U = 10 m/s, and is expressed by Equation (A1).
Q L = C l o a d   ω 2 [N·m]
Here, Cload [kgm2/rad2] is a coefficient that represents the load torque, and in Figure A4, the value of Cload = 3.71 × 10−9 kgm2/rad2 (when the rotor height is 43.4 mm).
The flow field data of the CFD analysis for the single turbine used in this study were obtained through CFD computation using the Dynamic Fluid/Body Interaction (DFBI) model built in STAR-CCM+, which considers the variation in the rotation speed of the turbine according to the flow velocity. In this case, the computational domain was set to twice the size (80D × 100D) of the domain shown in Figure A2 to compute under the same conditions as in the case of multiple wind turbines (see Figure 5a in reference [35]). The boundary conditions are identical. The upstream wind speed was U = 10 m/s, and the coefficient of the load torque curve defined by Equation (A1) was set to Cload = 8.57 × 10−8 kgm2/rad2, which corresponds to that of a rotor of unit height (1 m). The rotor mass and moment of inertia were also converted to unit height, and were set to M = 0.3226 kg and I = 1.284 × 10−4 kgm2, respectively. In the CFD analysis using the DFBI model, calculations are performed while solving the equation of motion (A2) for the wind turbine rotor, and the change in angular velocity ω over time is obtained.
I d ω d t = Q Q L
The time step was set to Δt = 2.5 × 10−5 s. The calculation was performed for up to 9 s, when the angular velocity was close to convergence. During the last second (8–9 s), time averaging was performed for each component of flow velocity (u, v) to create the input data for the proposed method. The characteristics of an isolated single (SI) miniature rotor (S-size) at the wind speed of U = 10 m/s were obtained from the convergence values of the CFD analysis using the above DFBI model as follows: angular velocity: ω = 363.7 rad/s, rotor torque: Q = 4.884 × 10−4 N · m, and rotor power: P = 0.1776 W (see Table A3).

Appendix B

To investigate the solidity dependence of the power output of paired VAWT rotors with the three configurations (CD, CU, and CO) shown in Figure 4a, the Renewable Energy Engineering Laboratory of Tottori University conducted a series of CFD analyses. In the analysis, the VAWT-paired arrangements were assumed to consist of rotors of three sizes (M, ML, and L), as shown in Figure A5, in addition to a miniature rotor with D = 50 mm. In all cases, the distance between the two rotors was set to gap = 0.2D. The specifications of the four rotors are listed in Table 1 in the main text.
Figure A5. Three 2-D rotors with different solidities. (a) M size; (b) ML size; and (c) L size.
Figure A5. Three 2-D rotors with different solidities. (a) M size; (b) ML size; and (c) L size.
Energies 18 00220 g0a5
For example, Figure A6 shows the state of the computational grid when an M-sized rotor is arranged in the CD layout. With respect to the main flow from left to right, the upper wind turbine was R1 and the lower wind turbine was R2. All four turbines had an NACA 0018 airfoil profile as the cross-section of each blade. The computational domains were 80D × 100D for the S, M, and ML turbine pairs, and 60D × 80D (the distance between the inlet boundary and the rotor was 32D) was used only for the L size. A trim mesh was used as the computational grid and a prism layer mesh was used only near the blade surface. Table A1 summarizes the number of cells in the computational grid used to calculate each paired wind turbine size. Table A2 shows the rotor mass: M, moment of inertia: I, load torque coefficient: Cload, Reynolds number: Re, and time step: Δt at the unit height (1 m) set for each case.
Figure A6. The mesh near a rotor pair arranged as the CD layout (M-size, gap = 0.2D).
Figure A6. The mesh near a rotor pair arranged as the CD layout (M-size, gap = 0.2D).
Energies 18 00220 g0a6
Table A1. Number of cells for calculation grid for CFD analysis of each rotor pair.
Table A1. Number of cells for calculation grid for CFD analysis of each rotor pair.
RegionRotor: SRotor: MRotor: MLRotor: L
Static179,000179,000179,000176,000
Rotational98,000334,000650,000278,000
Whole278,000515,000829,000454,000
Table A2. Setting of each rotor (unit height) of four rotor pairs. Reynolds numbers are based on each rotor diameter.
Table A2. Setting of each rotor (unit height) of four rotor pairs. Reynolds numbers are based on each rotor diameter.
RotorSMMLL
Mass: M [kg]0.3226 7.858 13.04 12.57
Moment of inertia: I [kgm2]1.284 × 10−47.885 81.58 314.5
Coefficient of load torque: Cload [kgm2/rad2]8.57 × 10−80.0291 0.758 3.36
Reynolds number: Re [-]3.3 × 1041.3 × 1063.3 × 1064.0 × 106
Time step: Δt [s]2.5 × 10−58.0 × 10−51.0 × 10−42.0 × 10−4
Figure A7 shows the distribution of the x-component (u) of the flow velocity for each wind turbine pair size in the CO arrangement. An example of the time variation in the angular velocity and torque of the wind turbine rotors (R1, R2) obtained by CFD analysis using the DFBI model is shown in Figure A8, in which the data for the ML-size (D = 5 m) rotors in the CO arrangement are illustrated. Figure A9 shows the time variation in the output power obtained from the product of the angular velocity and torque. The M, ML, and L wind turbines had large moments of inertia because their rotor masses were determined by assuming the presence of aluminum blades. Consequently, although the torque converged relatively quickly, the angular velocity required a longer time for convergence. Therefore, the angular velocities of the M-, ML-, and L size wind turbines for which the convergence was insufficient were approximated using an exponential function to estimate the convergence value. Table A3 summarizes the characteristics of each wind turbine rotor in the three configurations (CD, CU, and CO) of wind turbine pairs (gap = 0.2D), consisting of four rotor sizes (S, M, ML, and L). The data in red indicate the normalized average power (Pn_ave) of the two rotors (R1 and R2) for each configuration shown in Figure 4b in the main text.
Figure A7. Distributions of x-direction component (u) of flow velocity of CO-arrangement in each rotor pair case. (a) S size; (b) M size; (c) ML size; and (d) L size.
Figure A7. Distributions of x-direction component (u) of flow velocity of CO-arrangement in each rotor pair case. (a) S size; (b) M size; (c) ML size; and (d) L size.
Energies 18 00220 g0a7
Figure A8. Variation in angular velocity and torque of each rotor (R1, R2) of ML size in CO arrangement. (a) t = 0 to 25 s, (b) t = 23 to 25 s (partial expansion of (a)).
Figure A8. Variation in angular velocity and torque of each rotor (R1, R2) of ML size in CO arrangement. (a) t = 0 to 25 s, (b) t = 23 to 25 s (partial expansion of (a)).
Energies 18 00220 g0a8
Figure A9. Variation in mechanical power (product of ω and Q) of each rotor (R1 and R2) of ML size in CO arrangement. (a) t = 0 to 25 s, (b) t = 23 to 25 s (partial expansion of (a)).
Figure A9. Variation in mechanical power (product of ω and Q) of each rotor (R1 and R2) of ML size in CO arrangement. (a) t = 0 to 25 s, (b) t = 23 to 25 s (partial expansion of (a)).
Energies 18 00220 g0a9
Table A3. The obtained performance of each rotor in three layouts (CD, CU, and CO) of four-sized rotor pairs (S, M, ML, and L). The values in red show the normalized averaged power of two rotors (R1 and R2) of each pair (Pn_ave), shown in Figure 4b.
Table A3. The obtained performance of each rotor in three layouts (CD, CU, and CO) of four-sized rotor pairs (S, M, ML, and L). The values in red show the normalized averaged power of two rotors (R1 and R2) of each pair (Pn_ave), shown in Figure 4b.
Rotor: SSICD-R1CD-R2CD-Ave.CU-R1CU-R2CU-Ave.CO-R1CO-R2CO-Ave.
ω [rad/s]363.7 380.5 380.4 380.4 355.1 355.8 355.4 355.8 366.9 361.4
Q [N·m]0.0004884 0.0005781 0.0005250 0.0005516 0.0004723 0.0004575 0.0004649 0.0005027 0.0004575 0.0004801
P [W]0.1776 0.2200 0.1997 0.2098 0.1677 0.1628 0.1652 0.1789 0.1679 0.1734
Pnorm [-]1.000 1.239 1.124 1.1810.944 0.916 0.9301.007 0.945 0.976
Cp [-]0.00549 0.00680 0.00617 0.00648 0.00518 0.00503 0.00510 0.00552 0.00519 0.00535
λ [-]0.909 0.951 0.951 0.951 0.888 0.889 0.889 0.889 0.917 0.903
Rotor: MSICD-R1CD-R2CD-Ave.CU-R1CU-R2CU-Ave.CO-R1CO-R2CO-Ave.
ω [rad/s]24.85 25.76 25.72 25.74 25.45 25.47 25.46 25.59 25.53 25.56
Q [N·m]17.90 19.21 19.18 19.20 19.05 18.72 18.89 18.81 19.14 18.97
P [W]444.6 495.0 493.3 494.2 485.0 476.9 480.9 481.3 488.5 484.9
Pnorm [-]1.000 1.113 1.110 1.1111.091 1.072 1.0821.083 1.099 1.091
Cp [-]0.343 0.382 0.381 0.382 0.375 0.368 0.371 0.372 0.377 0.374
λ [-]2.485 2.576 2.572 2.574 2.545 2.547 2.546 2.559 2.553 2.556
Rotor: MLSICD-R1CD-R2CD-Ave.CU-R1CU-R2CU-Ave.CO-R1CO-R2CO-Ave.
ω [rad/s]11.96 12.25 12.21 12.23 12.28 12.28 12.28 12.28 12.18 12.23
Q [N·m]108.5 114.8 113.1 114.0 113.9 114.0 114.0 114.1 112.5 113.3
P [W]1298 1406 1382 1394 1398 1400 1399 1402 1371 1386
Pnorm [-]1.000 1.083 1.065 1.0741.078 1.079 1.0781.080 1.056 1.068
Cp [-]0.401 0.434 0.427 0.431 0.432 0.433 0.432 0.433 0.423 0.428
λ [-]2.991 3.061 3.054 3.058 3.071 3.070 3.070 3.071 3.045 3.058
Rotor: LSICD-R1CD-R2CD-Ave.CU-R1CU-R2CU-Ave.CO-R1CO-R2CO-Ave.
ω [rad/s]5.433 5.469 5.455 5.462 5.520 5.519 5.520 5.526 5.478 5.502
Q [N·m]99.48 100.53 101.51 101.02 102.21 102.20 102.20 102.98 100.92 101.95
P [W]540.5 549.7 553.7 551.7 564.2 564.1 564.1 569.1 552.8 561.0
Pnorm [-]1.000 1.017 1.025 1.0211.044 1.044 1.0441.053 1.023 1.038
Cp [-]0.386 0.393 0.396 0.394 0.403 0.403 0.403 0.407 0.395 0.401
λ [-]4.528 4.557 4.546 4.551 4.600 4.599 4.600 4.605 4.565 4.585

References

  1. Rajagopalan, R.G.; Rickerl, T.L.; Klimas, P.C. Aerodynamic interference of vertical axis wind turbines. J. Propul. Power 1990, 6, 645–653. [Google Scholar] [CrossRef]
  2. Whittlesey, R.W.; Liska, S.; Dabiri, J.O. Fish schooling as a basis for vertical axis wind turbine farm design. Bioinspir. Biomim. 2010, 5, 035005. [Google Scholar] [CrossRef]
  3. Dabiri, J.O. Potential order-of-magnitude enhancement of wind farm power density via counter-rotating vertical-axis wind turbine arrays. J. Renew. Sustain. Energy 2011, 3, 043104. [Google Scholar] [CrossRef]
  4. Kinzel, M.; Mulligan, Q.; Dabiri, J.O. Energy exchange in an array of vertical-axis wind turbines. J. Turbul. 2012, 13, N38. [Google Scholar] [CrossRef]
  5. Zanforlin, S.; Nishino, T. Fluid dynamic mechanisms of enhanced power generation by closely spaced vertical axis wind turbines. Renew. Energy 2016, 99, 1213–1226. [Google Scholar] [CrossRef]
  6. De Tavernier, D.; Ferreira, C.; Li, A.; Paulsen, U.S.; Madsen, H.A. Towards the understanding of vertical-axis wind turbines in double-rotor configuration. J. Phys. Conf. Ser. 2018, 1037, 022015. [Google Scholar] [CrossRef]
  7. Hezaveh, S.H.; Bou-Zeid, E.; Dabiri, J.; Kinzel, M.; Cortina, G.; Martinelli, L. Increasing the Power Production of Vertical-Axis Wind-Turbine Farms Using Synergistic Clustering. Bound.-Layer Meteorol. 2018, 169, 275–296. [Google Scholar] [CrossRef]
  8. Sahebzadeh, S.; Rezaeiha, A.; Montazeri, H. Towards optimal layout design of vertical-axis wind-turbine farms: Double rotor arrangements. Energy Convers. Manag. 2020, 226, 113527. [Google Scholar] [CrossRef]
  9. Hansen, J.T.; Mahak, M.; Tzanakis, I. Numerical modelling and optimization of vertical axis wind turbine pairs: A scale up approach. Renew. Energy 2021, 171, 1371–1381. [Google Scholar] [CrossRef]
  10. Silva, J.E.; Danao, L.A.M. Varying VAWT Cluster Configuration and the Effect on Individual Rotor and Overall Cluster Performance. Energies 2021, 14, 1567. [Google Scholar] [CrossRef]
  11. Xia, G.; Cao, Y.; Qian, Z.; Zhu, Y.; Wang, J.; Guo, T.; Yang, Y.; Zhang, W.; Wang, Y.; Wu, G. Optimization Layout and Aerodynamic Performance Research on Double Nautilus Vertical-Axis Wind Turbine. Appl. Sci. 2023, 13, 10959. [Google Scholar] [CrossRef]
  12. Ahmadi-Baloutaki, M.; Carriveau, R.; Ting, D.S.K. A wind tunnel study on the aerodynamic interaction of vertical axis wind turbines in array configurations. Renew. Energy 2016, 96, 904–913. [Google Scholar] [CrossRef]
  13. Vergaerde, A.; De Troyer, T.; Kluczewska-Bordier, J.; Parneix, N.; Silvert, F.; Runacres, M.C. Wind tunnel experiments of a pair of interacting vertical-axis wind turbines. J. Phys. Conf. Ser. 2018, 1037, 072049. [Google Scholar] [CrossRef]
  14. Vergaerde, A.; De Troyer, T.; Standaert, L.; Kluczewska-Bordier, J.; Pitance, D.; Immas, A.; Silvert, F.; Runacres, M.C. Experimental validation of the power enhancement of a pair of vertical-axis wind turbines. Renew. Energy 2020, 146, 181–187. [Google Scholar] [CrossRef]
  15. Su, H.; Meng, H.; Qu, T.; Lei, L. Wind tunnel experiment on the influence of array configuration on the power performance of vertical axis wind turbines. Energy Convers. Manag. 2021, 241, 114299. [Google Scholar] [CrossRef]
  16. Jodai, Y.; Tokuda, H.; Hara, Y. Experiments on Interaction between Six Vertical-Axis Wind Turbines in Pairs or Trios. J. Phys. Conf. Ser. 2024, 2767, 072003. [Google Scholar] [CrossRef]
  17. Mosetti, G.; Poloni, C.; Diviacco, B. Optimization of wind turbine positioning in large windfarms by means of a genetic algorithm. J. Wind Eng. Ind. Aerodyn. 1994, 51, 105–116. [Google Scholar] [CrossRef]
  18. González, J.S.; Gonzalez Rodriguez, A.G.; Mora, J.C.; Santos, J.R.; Payan, M.B. Optimization of wind farm turbines layout using an evolutive algorithm. Renew. Energy 2010, 35, 1671–1681. [Google Scholar] [CrossRef]
  19. Wagner, M.; Day, J.; Neumann, F. A fast and effective local search algorithm for optimizing the placement of wind turbines. Renew. Energy 2013, 51, 64–70. [Google Scholar] [CrossRef]
  20. Gao, X.; Yang, H.; Lin, L.; Koo, P. Wind turbine layout optimization using multi-population genetic algorithm and a case study in Hong Kong offshore. J. Wind Eng. Ind. Aerodyn. 2015, 139, 89–99. [Google Scholar] [CrossRef]
  21. Gao, X.; Yang, H.; Lu, L. Optimization of wind turbine layout position in a wind farm using a newly-developed two-dimensional wake model. Appl. Energy 2016, 174, 192–200. [Google Scholar] [CrossRef]
  22. Sun, H.; Yang, H.; Gao, X. Investigation into spacing restriction and layout optimization of wind farm with multiple types of wind turbines. Energy 2019, 168, 637–650. [Google Scholar] [CrossRef]
  23. Chen, K.; Song, M.X.; Zhang, X.; Wang, S.F. Wind turbine layout optimization with multiple hub height wind turbines using greedy algorithm. Renew. Energy 2016, 96, 676–686. [Google Scholar] [CrossRef]
  24. Gualtieri, G. A novel method for wind farm layout optimization based on wind turbine selection. Energy Convers. Manag. 2019, 193, 106–123. [Google Scholar] [CrossRef]
  25. Martinez, L.; Leonardi, S.; Churchfield, M.; Moriarty, P. A Comparison of Actuator Disk and Actuator Line Wind Turbine Models and Best Practices for Their Use. In Proceedings of the 50th AIAA Aerospace Sciences Meeting Including the New Horizons Forum and Aerospace Exposition, Nashville, TN, USA, 9–12 January 2012. [Google Scholar] [CrossRef]
  26. Malecha, Z.; Dsouza, G. Modeling of Wind Turbine Interactions and Wind Farm Losses Using the Velocity-Dependent Actuator Disc Model. Computation 2023, 11, 213. [Google Scholar] [CrossRef]
  27. Göçmen, T.; Laan, P.v.d.; Réthoré, P.-E.; Diaz, A.P.; Larsen, G.C.; Ott, S. Wind turbine wake models developed at the technical university of Denmark: A review. Renew. Sustain. Energy Rev. 2016, 60, 752–769. [Google Scholar] [CrossRef]
  28. Shao, Z.; Wu, Y.; Li, L.; Han, S.; Liu, Y. Multiple Wind Turbine Wakes Modeling Considering the Faster Wake Recovery in Overlapped Wakes. Energies 2019, 12, 680. [Google Scholar] [CrossRef]
  29. Chen, X.; Agarwal, R. Optimal Placement of Horizontal- and Vertical-Axis Wind Turbines in a Wind Farm for Maximum Power Generation Using a Genetic Algorithm, In Proceedings of the ASME 2011 5th International Conference on Energy Sustainability. Washington, DC, USA, 7–10 August 2011; pp. 2033–2039. [Google Scholar] [CrossRef]
  30. Lam, H.F.; Peng, H.Y. Development of a wake model for Darrieus-type straight-bladed vertical axis wind turbines and its application to micro-siting problems. Renew. Energy 2017, 114, 830–842. [Google Scholar] [CrossRef]
  31. Talamalek, A.; Runacres, M.C.; De Troyer, T. Effect of turbulence on the performance of a pair of vertical-axis wind turbines. J. Phys. Conf. Ser. 2022, 2265, 022098. [Google Scholar] [CrossRef]
  32. Cazzaro, D.; Bedon, G.; Pisinger, D. Vertical Axis Wind Turbine Layout Optimization. Energies 2023, 16, 2697. [Google Scholar] [CrossRef]
  33. Buranarote, J.; Hara, Y.; Furukawa, M.; Jodai, Y. Method to Predict Outputs of Two-Dimensional VAWT Rotors by Using Wake Model Mimicking the CFD-Created Flow Field. Energies 2022, 15, 5200. [Google Scholar] [CrossRef]
  34. Shapiro, C.R.; Starke, G.M.; Meneveau, C.; Gayme, D.F. A Wake Modeling Paradigm for Wind Farm Design and Control. Energies 2019, 12, 2956. [Google Scholar] [CrossRef]
  35. Hara, Y.; Jodai, Y.; Okinaga, T.; Furukawa, M. Numerical Analysis of the Dynamic Interaction between Two Closely Spaced Vertical-Axis Wind Turbines. Energies 2021, 14, 2286. [Google Scholar] [CrossRef]
  36. Mereu, R.; Federici, D.; Ferrari, G.; Schito, P.; Inzoli, F. Parametric numerical study of Savonius wind turbine interaction in a linear array. Renew. Energy 2017, 113, 1320–1332. [Google Scholar] [CrossRef]
Figure 1. The proposed method for simulating a VAWT wind farm using the numerical velocity data obtained by CFD analysis for a single isolated rotor.
Figure 1. The proposed method for simulating a VAWT wind farm using the numerical velocity data obtained by CFD analysis for a single isolated rotor.
Energies 18 00220 g001
Figure 3. The effects of velocity gradient du/dy and secondary component v on wind turbine rotor power. In the figure, the red arrows indicate the direction of rotation of the wind turbine rotor, and the green arrow indicates the velocity vector of the inflow wind (the combination of u and v components).
Figure 3. The effects of velocity gradient du/dy and secondary component v on wind turbine rotor power. In the figure, the red arrows indicate the direction of rotation of the wind turbine rotor, and the green arrow indicates the velocity vector of the inflow wind (the combination of u and v components).
Energies 18 00220 g003
Figure 4. (a) Three typical arrangements of VAWT–rotor pairs. (b) The effects of rotor solidity on the normalized average power of a rotor pair.
Figure 4. (a) Three typical arrangements of VAWT–rotor pairs. (b) The effects of rotor solidity on the normalized average power of a rotor pair.
Energies 18 00220 g004
Figure 5. The assumed rotor performance (power and rotational speed) of a miniature rotor, which is based on CFD results obtained at wind speed of 10 m/s.
Figure 5. The assumed rotor performance (power and rotational speed) of a miniature rotor, which is based on CFD results obtained at wind speed of 10 m/s.
Energies 18 00220 g005
Figure 6. Main calculation process: (a) Ntemp = 1; (b) Ntemp = 2; (c) Ntemp = 3; (d) N = 4.
Figure 6. Main calculation process: (a) Ntemp = 1; (b) Ntemp = 2; (c) Ntemp = 3; (d) N = 4.
Energies 18 00220 g006
Figure 7. The sub-calculation process (repetition) in the case of temporary rotor number Ntemp = 3: (a) Situation 1 (R1 is removed to calculate the revised UF(1)). (b) Situation 2 (R2 is removed to calculate the revised UF(2)). (c) Situation 3 (R3 is removed to calculate the revised UF(3)).
Figure 7. The sub-calculation process (repetition) in the case of temporary rotor number Ntemp = 3: (a) Situation 1 (R1 is removed to calculate the revised UF(1)). (b) Situation 2 (R2 is removed to calculate the revised UF(2)). (c) Situation 3 (R3 is removed to calculate the revised UF(3)).
Energies 18 00220 g007
Figure 8. A comparison between the new model and CFD in the case of CO–25 mm. (a) A flow field for θ = 22.5°; (b) a flow field for θ = 337.5° (−22.5°); (c) the 16-wind-direction distribution of power of R1; (d) the 16-wind-direction distribution of averaged power of R1 and R2.
Figure 8. A comparison between the new model and CFD in the case of CO–25 mm. (a) A flow field for θ = 22.5°; (b) a flow field for θ = 337.5° (−22.5°); (c) the 16-wind-direction distribution of power of R1; (d) the 16-wind-direction distribution of averaged power of R1 and R2.
Energies 18 00220 g008
Figure 9. A comparison between the new model and CFD in the case of IR–25 mm. (a) A flow field for θ = 0° (CD condition); (b) a flow field for θ = 180° (CU condition); (c) the 16-wind-direction distribution of power of R1; (d) the 16-wind-direction distribution of averaged power of R1 and R2.
Figure 9. A comparison between the new model and CFD in the case of IR–25 mm. (a) A flow field for θ = 0° (CD condition); (b) a flow field for θ = 180° (CU condition); (c) the 16-wind-direction distribution of power of R1; (d) the 16-wind-direction distribution of averaged power of R1 and R2.
Energies 18 00220 g009
Figure 10. (a) Normalized power difference in R1 (Errn_R1) between CFD and present model predictions. (b) Normalized average power difference in R1 and R2 (Errn_ave) between CFD and present model predictions (grid number: m = 400. Colors show the difference in normalized cell size: Celln_D = 0.05, 0.1, 0.2, 0.4, 0.75).
Figure 10. (a) Normalized power difference in R1 (Errn_R1) between CFD and present model predictions. (b) Normalized average power difference in R1 and R2 (Errn_ave) between CFD and present model predictions (grid number: m = 400. Colors show the difference in normalized cell size: Celln_D = 0.05, 0.1, 0.2, 0.4, 0.75).
Energies 18 00220 g010
Figure 11. (a) Normalized power difference in R1 (Errn_R1) between CFD and present model predictions. (b) Normalized average power difference in R1 and R2 (Errn_ave) between CFD and present model predictions (grid number: m = 600. Colors show the difference in normalized cell size: Celln_D = 0.0333, 0.0667, 0.1333, 0.2667, 0.5).
Figure 11. (a) Normalized power difference in R1 (Errn_R1) between CFD and present model predictions. (b) Normalized average power difference in R1 and R2 (Errn_ave) between CFD and present model predictions (grid number: m = 600. Colors show the difference in normalized cell size: Celln_D = 0.0333, 0.0667, 0.1333, 0.2667, 0.5).
Energies 18 00220 g011
Figure 12. A comparison of the power prediction between (a) CFD (unsteady) and the (b) present model (steady) in the case of a four-VAWT parallel arrangement (unit of power: mW).
Figure 12. A comparison of the power prediction between (a) CFD (unsteady) and the (b) present model (steady) in the case of a four-VAWT parallel arrangement (unit of power: mW).
Energies 18 00220 g012
Figure 13. A comparison of the power prediction between (a) CFD (unsteady) and the (b) present model (steady) in the case of a four-VAWT tandem arrangement (unit of power: mW).
Figure 13. A comparison of the power prediction between (a) CFD (unsteady) and the (b) present model (steady) in the case of a four-VAWT tandem arrangement (unit of power: mW).
Energies 18 00220 g013
Figure 14. (a) The flow field and output power of each turbine at θ = 0° in the CO–4 × 4 layout. (b) The wind-direction dependence of the averaged power of 16 VAWTs.
Figure 14. (a) The flow field and output power of each turbine at θ = 0° in the CO–4 × 4 layout. (b) The wind-direction dependence of the averaged power of 16 VAWTs.
Energies 18 00220 g014
Figure 15. (a) The flow field and output power of each turbine at θ = 0° in the CO–8–pairs layout. (b) The wind-direction dependence of the averaged power of 16 VAWTs.
Figure 15. (a) The flow field and output power of each turbine at θ = 0° in the CO–8–pairs layout. (b) The wind-direction dependence of the averaged power of 16 VAWTs.
Energies 18 00220 g015
Figure 16. (a) The flow field and output power of each turbine at θ = 0° in the IR–8–pairs layout. (b) The wind-direction dependence of the averaged power of 16 VAWTs.
Figure 16. (a) The flow field and output power of each turbine at θ = 0° in the IR–8–pairs layout. (b) The wind-direction dependence of the averaged power of 16 VAWTs.
Energies 18 00220 g016
Figure 17. (a) The flow field and output power of each turbine at θ = 0° in the CO–8 × 2 layout. (b) The wind-direction dependence of the averaged power of 16 VAWTs.
Figure 17. (a) The flow field and output power of each turbine at θ = 0° in the CO–8 × 2 layout. (b) The wind-direction dependence of the averaged power of 16 VAWTs.
Energies 18 00220 g017
Figure 18. The nonuniform wind direction probability (WDP) distribution observed in the Arid Land Research Center of Tottori University. Wind-direction angle θi = 0 corresponds to the north in this figure.
Figure 18. The nonuniform wind direction probability (WDP) distribution observed in the Arid Land Research Center of Tottori University. Wind-direction angle θi = 0 corresponds to the north in this figure.
Energies 18 00220 g018
Table 1. Specification and calculation conditions of four rotor pairs.
Table 1. Specification and calculation conditions of four rotor pairs.
SymbolLMLMS
Solidity: σ[–]0.0250.0570.0950.382
Diameter: D [m]10520.05
Number of blade: B [–]2333
Chord length: c [m]0.40.30.20.02
gap/D [–]0.20.20.20.2
Wind velocity: U [m/s]6101010
Table 2. AEPs for four layouts consisting of 16 VAWTs for uniform wind-direction distribution.
Table 2. AEPs for four layouts consisting of 16 VAWTs for uniform wind-direction distribution.
LayoutAEP [kWh]AEP/Area [kWh/m2]AEP/AEP16SI
CO–4 × 425.857103.4280.673
CO–8–pairs21.72586.9000.565
IR–8–pairs22.12188.4840.576
CO–8 × 231.083124.3320.809
Table 3. AEPs for four layouts consisting of 16 VAWTs for nonuniform wind-direction distribution shown in Figure 18.
Table 3. AEPs for four layouts consisting of 16 VAWTs for nonuniform wind-direction distribution shown in Figure 18.
LayoutAEP [kWh]AEP/Area [kWh/m2]AEP/AEP16SI
CO–4 × 427.603 110.4120.718
CO–8–pairs25.868 103.4720.673
IR–8–pairs26.170 104.680.681
CO–8 × 246.508 186.0321.210
Table 4. The calculation cost of a program implemented using the proposed method.
Table 4. The calculation cost of a program implemented using the proposed method.
LayoutNTime (1-dir.) [s]Time (16-dir.) [s]
SI1 0.085 0.385
CO–50 mm2 0.220 1.107
4–rotor parallel4 1.079 2.93
CO–4 × 416 54.425 114.739
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Hara, Y.; Moral, M.S.; Ide, A.; Jodai, Y. Fast Simulation of the Flow Field in a VAWT Wind Farm Using the Numerical Data Obtained by CFD Analysis for a Single Rotor. Energies 2025, 18, 220. https://doi.org/10.3390/en18010220

AMA Style

Hara Y, Moral MS, Ide A, Jodai Y. Fast Simulation of the Flow Field in a VAWT Wind Farm Using the Numerical Data Obtained by CFD Analysis for a Single Rotor. Energies. 2025; 18(1):220. https://doi.org/10.3390/en18010220

Chicago/Turabian Style

Hara, Yutaka, Md. Shameem Moral, Aoi Ide, and Yoshifumi Jodai. 2025. "Fast Simulation of the Flow Field in a VAWT Wind Farm Using the Numerical Data Obtained by CFD Analysis for a Single Rotor" Energies 18, no. 1: 220. https://doi.org/10.3390/en18010220

APA Style

Hara, Y., Moral, M. S., Ide, A., & Jodai, Y. (2025). Fast Simulation of the Flow Field in a VAWT Wind Farm Using the Numerical Data Obtained by CFD Analysis for a Single Rotor. Energies, 18(1), 220. https://doi.org/10.3390/en18010220

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