10 1016@j Renene 2019 12 060

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 31

Journal Pre-proof

The design of a small lab-scale wind turbine model with high performance
similarity to its utility-scale prototype

B. Li, D.L. Zhou, Y. Wang, Y. Shuai, Q.Z. Liu, W.H. Cai

PII: S0960-1481(19)31932-9
DOI: https://doi.org/10.1016/j.renene.2019.12.060
Reference: RENE 12769

To appear in: Renewable Energy

Received Date: 15 September 2019


Accepted Date: 12 December 2019

Please cite this article as: B. Li, D.L. Zhou, Y. Wang, Y. Shuai, Q.Z. Liu, W.H. Cai, The design of a
small lab-scale wind turbine model with high performance similarity to its utility-scale prototype,
Renewable Energy (2019), https://doi.org/10.1016/j.renene.2019.12.060

This is a PDF file of an article that has undergone enhancements after acceptance, such as the
addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive
version of record. This version will undergo additional copyediting, typesetting and review before it
is published in its final form, but we are providing this version to give early visibility of the article.
Please note that, during the production process, errors may be discovered which could affect the
content, and all legal disclaimers that apply to the journal pertain.

© 2019 Published by Elsevier.


Journal Pre-proof

The design of a small lab-scale wind turbine model with high


performance similarity to its utility-scale prototype

B. Li1, D. L. Zhou1, Y. Wang2, Y. Shuai1, Q. Z. Liu1*, W. H. Cai2*

1 Harbin Institute of Technology, School of Energy Science and Engineering, 92 West Dazhi
Rd, Harbin, China
2 Northeast Electric Power Univeristy, School of Energy and Power Engineering, 169

Changchun Rd, Jilin City, China


Corresponding author: W. H. Cai, caiwh@hit.edu.cn Q. Z. Liu, liuquanzhong@hit.edu.cn

Abstract

This paper discussed the procedure of an optimization design method involving a small lab-

scale wind turbine rotor. To simulate the aerodynamic performance of a prototype wind turbine

at utility-scale by scaled model experiments (especially, the visualization measurements of

wake flow in small wind tunnels), an appropriate similar design of the small lab-scale rotor is

needed. However, the similar design of blade performance from a prototype wind turbine for

small scaled model is rare. In this paper, using a 2.5 MW utility-scale wind turbine as the

prototype, we set up a new optimization process of similar design for a model wind turbine

with a 320 mm rotor based on the Lifting-Line Theory (LLT) with wake-induced corrections.

Results show a dramatic deviation in performance of the geometric scaled model compared to

the objective values, which demonstrates the significance of the Reynolds number effects. With

the optimized blade, the distribution of normal thrust force is very similar to the objective

values. Geometric characteristics variations along the blade span are similar to their prototype.

1
Journal Pre-proof

The tip speed ratio runs at a roughly matched rated condition. The optimized rotor designed

model gives a better performance than other models used in previous wind tunnel studies.

Keywords: model design; wind turbine; similarity; utility-scale; lab-scale

1 Introduction

The development of sustainable renewable energy has become the strategy of energy

development in different countries [1,2]. It is important with an in-depth understanding in terms

of wake development mechanism that would help to improve wind turbine or wind farm

optimal operation strategies [3,4]. Model experimental studies in a lab-scale wind tunnel is also

necessary for the validation of numerical simulation studies, and aid the design of wind turbines.

The rotor of a wind turbine, consisting of the blade and hub rotor assembly, operates in an

atmospheric boundary layer (ABL) characterized by highly turbulent airflow and is subjected

to various wind induced interactions. Besides utility-scale experiments, wind tunnel

experiments can be an important method for studying those extremely complex interactions.

Different sizes of model wind turbines have been used, such as large-scale model turbine [5,6,7],

meter-scale model turbine [8,9,10], and small size model turbine [11,12]. Model experimental

studies with precise measurement methods, such as PIV (Particle image velocimetry), in wind

tunnels can also quantify the performance of a turbine and the correlation with its wake

development accurately [13]. However, both the large scale and the meter scale models are

oversized for most of the wind tunnels due to blockage effect, and its large wake area is

unfavorable for PIV measurement. Therefore, there are many studies adopt small sized model

turbines for wake flow study with wind tunnel experiments.

2
Journal Pre-proof

Due to the different blade types used in wind tunnel experiments at small size of lab-

scale, related researches have generally investigated four types of wind turbine models namely;

direct geometric scale models [14,15], commercial propeller models [16], parametric design

models [17,18], and simple rotor models [19-21]. First, using a geometric scale model based

on the 2MW ERS-100, Yuan W et al. [14] study the influence of the opposite rotation direction

of blades on the wake flow of a series wind turbine. Taking 1/350 as the geometric scaled

factor, the rotor radius is 127mm and the hub height is 225 mm. Tian W et al. [15] use the same

model as Yuan et al. to study the influence of wind conditions on the wake characteristics and

dynamic wind load of the wind turbine model. The difference is that the geometric scale factor

is 1/320 with a smaller 140 mm rotor radius. Results show correlation between the tip vortex

shedding and inflow, but has no comparison with the prototype wind turbine. Second, without

a wind turbine prototype, Hu H et al. [16] installed a propeller model rotor MA0530TE while

studying the aerodynamic characteristics near the wake of the horizontal-axis wind turbine

model in the atmospheric boundary layer wind tunnel. The diameter of the impeller was 127mm,

the twisted blades ranged from 20° at the root to 10°at the tip, which was reversely installed to

simulate the wind turbine. Third, Mctavish et al. [17, 18] designed a three-bladed rotor

specifically for operation conditions under Reynolds number of order 104. It is a parametric

design that matches the tip speed ratio and thrust coefficient to common utility-scale wind

turbines. Fourth, studying the vertical transport of momentum and kinetic energy in the wind

turbine boundary layer, Cal R et al. [19] designed a simple wind turbine model for wind tunnel

experiments; the model blade was cut and twisted from uniform metal sheets. The twist angle

varied from 15° at the root to 10° at the top with a blade thickness of 0.48 mm. Lebrón J et al.

3
Journal Pre-proof

[20] used the same blade design method to study the interaction between the wind turbine array

and the turbulent boundary layer. Hamilton et al. [21] also used a similar design as the scaled

model of a 1 MW horizontal axial wind turbine for the study of the boundary layer of wind

turbine arrays. However, due to Reynolds number effect, the performance of a small lab-scale

model of wind turbine can be significantly different from its prototype [22,23]. For the wake

studies on a small lab-scale model of wind turbine, whether the geometric scaled rotor,

commercial impeller, parametric design rotor, or simple rotor represent the real wind turbine

rotor remains uncertain.

There is a lack of practical methods for the design of a model wind turbine at the lab-scale

to match the wake flow of its prototype real turbine. It greatly restricts the development of wind

tunnel experiments for relative studies of wake flow. To solve this problem, a design method

of wind turbine, especially the blade, at small scale with high performance of similarity to its

prototype should be put forward. In this study, an innovative design method is proposed. In

section 2, the conversion relationship between prototype and model parameters is deduced with

similarity conditions; the lifting-line theory based model with wake-induced corrections is set

up for the calculation of blade performance. In section 3, we use a similar hybrid blade as the

prototype with the optimization variable at each section of the model blade reaching the

objective value. Finally, a blade model with high performance simulation of the aerodynamic

characteristics of the prototype is obtained.

2 Method

In this study, an onshore 2.5-MW turbine was used as the prototype. The lab-scale turbine

has a geometric scaling factor of λ = 1: 300. In view of the big difference in unit performance
4
Journal Pre-proof

between the direct geometric scale model and the prototype, an optimized method of blade

design is established in this paper. The general procedure of the design of a lab-scale wind

turbine model mainly consists of three steps (as shown in Fig.1). The first step, is to calculate

the detailed performance of the prototype wind turbine, and derive the objective values of unit

thrust force at each section following the similarity criterion. The Second, is to take the direct

geometric scale value as the initial condition, and set up constraints for optimization. The final

step, is to calculate and validate the results of the optimization. In details, the radial distribution

of chord length and twist angle of the blade were obtained from the optimization calculations

using a lifting-line theory based model with wake-induced corrections. Results of the model

and prototype wind turbines were compared with the fully coupled simulation tool (Qblade)

[24]. Finally, we can get an optimized model with aerodynamic performance similar to the

prototype.

Fig. 1 Flow chart of the blade similar design

2.1 General Characteristics of the prototype 2.5-MW turbine

Table 1 illustrates the primary parameters of the three-bladed horizontal-axis 2.5-MW turbine
5
Journal Pre-proof

used as prototype in this study. The diameter of the rotor is 96 m, and the rated operating speed

is 11 m/s with design tip speed ratio of TSR = 8 ~ 12. The twist angle of the blade ranges from

9.5 ° at root to -2 ° near the tip and 2 ° at the tip. It is a hybrid designed blade with a circle

cross-section at the root and a high aerodynamic performance airfoil near the tip. The chord-

based Reynolds number at the tip region is approximately 6.3 ×106. The turbine essentially

consists of three pitchable blades to control the aerodynamic torque; the collective pitch control

gives the best efficiency for power generation at a fine pitch of 1 °.

Table 1. General characteristics of the reference 2.5-MW wind turbine

Parameters size unit


Rotor diameter 96 m
Hub height 80 m
Rated wind speed 11 m/s
Rotation, view from upwind Clockwise
Position of rotor Upwind, 3 blades
Cut in Wind Speed 4 m/s
Cut out Wind Speed 25 m/s

2.2 Similarity conditions

A similar design of a lab-scale wind turbine model to its referenced utility-scale wind

turbine should employ the following similarity conditions [25, 26]: geometric similarity,

kinematic similarity and dynamic similarity.

Geometric similarity: It is the baseline for a similar design such that the model and its

prototype are geometrically similar. For linear scale parameters, geometric similarity

conditions must be satisfied, such as diameter, hub height, hub diameter, etc. The geometric

scale ratio λL is defined as follows:

𝜆𝐿 = 𝐷m 𝐷P = 𝜆 (1)

where, D is rotor diameter, subscript m denotes model, subscript p denotes prototype.


6
Journal Pre-proof

Kinematic similarity: If the corresponding points of the prototype flow field and the model

flow field have the same velocity, and the velocity vector is geometrically similar, the kinematic

similarity is satisfied. Obviously, it is impossible to match the Reynolds number because of the

huge difference in lengths between the model and the prototype. However, blade tip speed ratio

(TSR = 𝜔𝑅 𝑉𝑖𝑛), which represents the dimensionless number of the ratio of the blade tip linear

velocity to incoming wind speed Vin, can generally characterize the kinematic similarity for the

wind turbine design. In order to meet the similarity condition, the TSR of the model blade and

the prototype should be consistent that is, TSRp = TSRm. The rotating motion of the blade is

periodic. Thus, the model blade and prototype must keep the value of Strouhal number equal:
𝑉𝑝 𝑉𝑚
Sr𝑏𝑙𝑎𝑑𝑒 = = (2)
𝑛𝑝𝐷𝑝 𝑛𝑚𝐷𝑚

where n is the revolutions per minute (RPM). With 𝜔 = 2𝜋𝑛 60 for the blade, it has TSR =

𝜋𝑛𝐷 60𝑉in, then Eq.(2) indicates Sr = 𝜋 60 ∙ TSR. Therefore, the similarity of TSR can match

the similarity of Sr. Thus, the RPM of model turbine can be achieved by the design of reference

velocity scaling ratio:

𝜆𝑉 = 𝑉m 𝑉P (3)

Dynamic similarity: Once we study the wake area of a wind turbine, the thrust coefficient

can be the most important parameter derived by many wake models [27, 28]. According to the

Jensen wake model [29], the velocity distribution in the wake area mainly relates to the axial

thrust coefficient of the wind turbine and assumes that the wake flow increases linearly with

the increasing distance from the rotor. According to Jensen’s wake model, the wake velocity

can be expressed as:

7
Journal Pre-proof

  r  
2

v  v0 1  (1  1  CT )    (4)
  r   x  

Where, v0 is the local wind speed, v and r are respectively the wake velocity and radius at the

distance x downstream of the wind turbine, CT is the thrust coefficient,  is the dimensionless

attenuation constant, which relates to ambient turbulence and turbine-induced turbulence.

Therefore, the thrust coefficient is an important coefficient for the wake flow. The Frandsen’s

wake model [27] also demonstrates a similar conclusion. Therefore, in this study, the thrust

coefficient CT is defined as the major criteria of dynamic similarity, and it is a good

combination of both the lift coefficient CL and the drag coefficient CD because of the triangle

of velocities. In all, aerodynamic parameters of the blade can be scaled down according to the

above similarity conditions.

2.3 LLT based model

In wind turbine studies, the Blade Element Momentum (BEM) theory is widely used for

blade design of wind turbine, because of its high computational efficiency. However, it also

has limitations and is only applicable to the situation that the flow field around the airfoil is

always in equilibrium or a large number of empirical correction models have to be added to

achieve a tolerable accuracy. In fact, the wake of a rotor running naturally presents relatively

lagged unsteady aerodynamic characteristics [30]. Therefore, the calculation error of BEM in

rotor coupling analysis is large [31]. Moreover, the lifting-line theory [32], also known as the

Lanchester-Prandtl wing theory, is an important mathematical theory to solve the problems

related to the lift of blade. According to the geometric shape of the wing (the distribution of

chord, airfoil and twist angle along the aspect of the chord) and the flow condition, the lift force

8
Journal Pre-proof

distribution along the aspect of the chord is obtained. Afjeh [33] corroborated the high accuracy

of the lifting line model in his study of the influence of wakes on the aerodynamic performance

of rotors. In addition, to consider the effect of wake-induced corrections, the resultant flow

condition is updated to generate a more accurate blade performance. Therefore, in this study,

the LLT is improved to calculate the aerodynamic performance of the rotor. The new version

of QBlade (v0.963) also includes a new aerodynamic module with a new advanced Lifting-line

theory module.

Based on the generalized Lifting-line theory, with incompressible and inviscid flow, the

high-aspect-ratio wing is modelled as a single bound vortex line located at the 1/4 chord

position and an associated shed vortex sheet. The lifting line combines the attached vortices on

the blade into a vortex line with a variable strength. The lift of each section acts on the line. For

the first part, the wake model assumes that the wake is negligible and remains in a plane parallel

to the free stream. The model wake using single vortex sheet starting at the quarter chord, is

shown in Fig. 2.

Fig. 2 The coordinate system of blade wake model

Assuming that the Kutta-Joukowsky condition is satisfied at local sections, then the lift L

is expressed as:

9
Journal Pre-proof

𝑠
𝐿 ≈ 𝜌𝑉∞ ∫𝛤𝑑𝑦 (5)
―𝑠

where ρ is the air density (kg/𝑚3), V∞ is the freestream velocity (m/s), 𝛤 is

the circulation (m²/s). The downwash at y due to entire wake is given by:
―𝑑Γ
𝑠 𝑑𝑦| 𝑑𝑦1
𝑦1 (6)
𝑤(y) = ∫ ―𝑠 4𝜋(𝑦 ― 𝑦1)

where w is the downwash velocity, y1 is the location of the section dy1, s is the half span.

In addition, the wake-induced correction to w can be improved by considering the free

vortex filaments in the wake area. Under induced flow, both the aerodynamic performance and

vortex filament nodes changes. Assuming it moves as Lagrangian lattices, the corrected

equation of flow velocity is given by Eq. (7),

∂𝒙 ∂𝒙
𝑉' = 𝑉∞ + 𝑉𝑒 + 𝑉𝑇 = n( + ) (5)
∂𝜓 ∂𝜁

where, ψ is the azimuthal rotor position, ζ is the wake age, Ve is the induced effective velocity,

VT is the turbine rotor speed contribution.

For the second part, the section model assumes that flow over each section are determined

by downwash at ¼ chord, according to the thin airfoil theory. The sectional lift coefficient (2D

lift coefficient) is given by:

𝑙 𝜌𝑉'Γ
C𝑙 = =
1 '2 1 '2 (6)
𝜌𝑉 𝑐 𝜌𝑉 𝑐
2 2

where c is the chord length of the local section,l is the lift force per unit span.

Γ = π𝑉'(𝛼 ― 𝛼0)𝑐 + 𝜋𝑤𝑐 (7)

where α0 is zero-lift angle of attack of that section (depends on the airfoil geometry),α is the

geometric angle of attack.

In this study, in order to make the aerodynamic performance of the blade more accurate,
10
Journal Pre-proof

the relative calculation of performance is based on LLT instead of BEM.

3 Modeling procedures and solutions

3.1 Objective value of parameters

According to the geometric, motion and dynamic similarity conditions given above, the

conversion relationship that each physical quantity should satisfy for the model turbine can be

deduced as below in Table 2. Thus, the model turbine has a rotor diameter of 320 mm, hub

height of 270 mm, and TSR of 7. With an 11 m/s incoming wind speed in the lab experiment,

the rated RPM can go up to 4650. The lab-scale experiment can implement a low incoming

wind speed to achieve a low RPM that is suitable with the use of small drive-motor, but it may

generate a low power coefficient. Moreover, the objective values of unit thrust force of all sections

on the blade can be derived based on the similarity scale factor as shown in Fig. 6. The chord based

Reynolds number at the tip region of the model turbine is scaled to approximately 2 ×104 which

demonstrates a low-Re condition.

Table 2. The conversion relationship in the model turbine design


Parameters Field scale Scale factor Lab-scale
Rated power 2.5 MW 𝜆2𝐿𝜆3𝑉 27.8 w
Blades number 3 1 3
Diameter 96 m λL 320 mm
Hub height 80 m λL 270 mm
Rated wind speed 11 m/s 1 11 m/s
Rated rpm 15.5 𝜆𝐿―1𝜆𝑉 4650
TSR 7 1 7
Rated thrust force 388 KN 𝜆2𝐿𝜆2𝑉 4.31 N

The comparison of aerodynamic parameters among different types of blade design models

is illustrated in Fig. 3. Here, the theoretical scaled value at lab-scale is converted directly from

11
Journal Pre-proof

the prototype based on the scale factor; the geometric scaled value is the aerodynamic

performance of the lab-scale model blade which is geometrically scaled from the prototype;

DU96W180 and NACA4412 are two blades with corresponding uniform airfoil along the entire

span. It should be mentioned that the comparison in Fig. 3 is based on the same model size.

Moreover, we used Qblade to calculate the aerodynamic performance of the turbine at different

scales. As it can be seen from Fig. 3a, the Ct of the geometric scaled model is dramatically

smaller than the theoretically scaled one, which means that experimental results of the

geometric scaled model should have considerable deviations. This demonstrates the significant

Reynolds number effects. Therefore, the necessity of optimization of the direct geometric scale

model is obvious, especially for a wake area study. Fig. 3a also illustrates that Ct of blades with

uniform airfoil of DU96W180 and NACA4412 at their entire span are smaller than the

theoretical scaled one, but NACA4412 obviously has a better performance under this low-Re

condition. A similar conclusion can be drawn from Fig. 3b, where lift coefficients in other blade

models are much smaller than the theoretical scaled value. In addition, the NACA4412 model

gives a high Cl for entire blade with fine pitch at lab-scale, and the geometric scaled model

coming last shows significant deviation. From Fig. 3c, Cd of the geometric scaled model has

similar deviations as the DU96W180 and NACA4412, but illustrates a large gap compare to

the theoretical scaled model. In view of the huge difference in aerodynamic performance of the

geometric and the theoretical scaled model, undertaking an optimization design of the model

turbine is a desiderate task for wake related studies.

In the optimization procedure, we would like to embed the improved LLT-based model in

MATLAB as the main process tool of the performance calculation. However, a validation is

12
Journal Pre-proof

required and consists of two main steps in this test. The first step is to run a set up with the

prototype turbine under its rated condition. This gives a good agreement in terms of the rated

power output of around 2.5 MW. The second step compares the results of the MATLAB code

to Qblade. As shown in Fig. 4, the figure demonstrates that the results of the LLT based

MATLAB code and those of Qblade at utility-scale (e.g. theoretical scaled data) are quite

consistent. Moreover, there is a significant deviation of Ct near root area of the span at the lab-

scale (e.g. geometric scaled data). As the LLT based MATLAB code simulates a reasonably

similar performance between the geometric scaled blade and the blade with uniform airfoil of

DU96W180 in Fig. 3, and together with the deviation of lab-scale model in Fig.4, we would

not like to trust the Qblade under these low-Re conditions. Therefore, our optimization code

should embed the calculation of the LLT based model performance. At the same time, both

calculation methods show that there is a considerable variation in the thrust coefficient of the

geometric scaled turbine when compared to the theoretical scaled value.

a) Theoretical scaled
Geometric scaled
0.25 DU96W180
NACA4412
0.20

0.15
Ct

0.10

0.05

0.00

0.0 0.2 0.4 0.6 0.8 1.0


r/R

13
Journal Pre-proof

b) Theoretical scaled
c) 2.0
2.0 Geometric scaled Theoretical scaled
1.8
DU96W180 Geometric scaled
NACA4412 DU96W180
1.6 1.5
NACA4412
1.4

1.2
1.0
1.0

Cd
Cl

0.8

0.6 0.5
0.4

0.2
0.0
0.0

-0.2
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0

r/R r/R

Fig. 3 Aerodynamic performance of different turbine models. Values of the theoretical scaled model
are directly converted from the values of the prototype turbine by scale factor; Values of the geometric
scaled model are given by the aerodynamic performance of the lab-scale model blade which is
geometrically scaled from the prototype; Values of the other two models are calculated from blades
consist of uniform airfoils of DU960W180 and NACA4412 respectively in the entire span.

Theoretical scaled-Qblade-BEM
0.25 Geometric scaled-Qblade-BEM
Theoretical scaled-Qblade-LLT
Geometric scaled-Qblade-LLT
Theoretical scaled-WInDS-LLT
0.20 Geometric scaled-WInDS-LLT

0.15
Ct

0.10

0.05

0.00

0.0 0.2 0.4 0.6 0.8 1.0


r/R

Fig. 4 Comparison of thrust coefficient among different blade models.

3.2 Optimization setup

In order to achieve a more comparable aerodynamic performance of the lab-scale wind

turbine to the prototype, a new optimized design method is developed in the current paper. In

addition, it should be noted that the utility-scale wind turbine design generally uses hybrid blade.

A hybrid blade can combine advantages of both strong structure at root and high performance

at near tip sections. In fact, it is also important for lab-scale model design, because the use of

thin airfoil with high performance at root area would result in a very weak structure of blade at

14
Journal Pre-proof

small size. It may not run at high revolving speed and is difficult to manufacture. For this reason,

the optimized design of a hybrid blade at lab-scale is conducted in this study.

The aerodynamic design of blades is directly based on the geometric scaled model, that

is, the distribution of chord length and the twist angle along the span are geometrically similar

to its prototype. In addition, in the optimized design process, there are mainly three problems

worthy of explanation. Firstly, the design of blades should follow the above-mentioned

similarity conditions. Secondly, the blade optimal design should be carried out based on the

geometric similarity, e.g. blade shape and the airfoil transformation. Thirdly, due to the fact

that the first several sections of the 2.5-MW turbine blade do not contribute significantly to the

performance, and also because the chord length changes dramatically, we clipped them off in

order to simplify the optimization process. Therefore, the redesigned wind turbine maintains

those near root sections by geometrically scaling profiles. In addition, considering the general

method of blade design of a traditional wind turbine [34], we select chord length and twist angle

as variables of the optimization process.

For the optimal calculation of the hybrid blade design, the characteristics of different

airfoils used in the hybrid blade are the most remarkable input parameters, that is, the design

of transition airfoils, and aerodynamic parameters of all airfoils. In addition, the prototype

turbine consists of 27 sections with 16 different airfoils. It changes from a circle at root to the

DU96W180 at tip gradually. The circle section improves the strength at root area with large

thickness airfoils; the DU96W180 airfoil improves the performance at tip area with its specific

aerodynamic characteristics. Accordingly, the model blade design considers a similar design

of airfoils as its prototype, for which geometry of model blade generally changes from a circle

15
Journal Pre-proof

to the NACA4412 airfoil. In particular, it should be noted that the NACA4412 is much thinner

than those transition DUINT-series airfoils that are used in the prototype blade. Therefore, we

design a new transition airfoil of TEST3415 (Camber 3.13%, at 39.2%, thickness 15.46%) to

transfer smoothly from DUINT190 to NACA4412. The comparison of airfoils in prototype and

model blade are shown in Fig. 5. The model turbine has a much simple structure with only 8

kinds of airfoils for its small size. To improve the performance of sections at middle area of the

blade (≤50% span), more NACA4412 sections are used. Then, aerodynamic characteristics of

all airfoils are evaluated by Qblade respectively as parameters of the optimization model.

Fig. 5 Comparison of airfoils in prototype and model blade; left is the prototype and right is the model.

The initial guess of the two variables for the optimization process is to make variables the

same as in the geometric scaled blade. However, in order to ensure the continuity of the chord

length and the twist angle of two adjacent sections of different airfoils, the least square method

is used to fit the data along the span wise distribution. Prior to this, Kim B [35] also performed

similar blade parameter fitting work. Thus, for those two geometric parameters of the geometric

scaled model blade in the current study, the span wise distribution of the chord length and the

twist angle can be roughly fitted into two third-order polynomials as shown in Fig. 6. In addition,

the last point of the chord length and the last two points of the twist angle are ignored in the
16
Journal Pre-proof

optimization process because of its small size and sharp change at the tip sections. Those

sections are directly copied to the model design.

a)

b)

Fig. 6 Fitting of distributions for a) chord length and b) twist angle along the span wise direction

The optimization method is set up to search for an optimized result of the blade shape. On

the basis of the specific airfoil at each section, non-unique results of the combinations of chord

lengths and twist angles can generate a similar thrust force. However, ranges of optimized chord

lengths and twist angles are subjected to additional restrictions of similarity of the blade shape

to the prototype. Therefore, the goal of the optimization method is to both satisfy the similarity

of shape to the geometric scaled model and match the objective values. Here, the objective

value is the unit thrust force (t0) of the theoretical scaled values. As a basic rule of constrained

optimization problems, the mathematical optimization with m constrain equations uses the

following general form:

min f ( X );

 s.t. (10)
 g ( x)  0, i  1, 2,..., m
 i  

Here we generally consider that the optimized blade should follow geometrically within a

constraint region similar to its prototype (for example positive chord length and twist angle).
17
Journal Pre-proof

Therefore, the specified optimization model for the blade thrust can be expressed as follows:

 Nsec

 T  X   t   i  1, 2, , N 
2
min f ( X )  i 0i sec
 i 1

 s.t. (11)

 LB  X  UB
 AX  b

where Nsec is the number of sections of the blade in the span wise direction; t0i is the objective

value of unit thrust force at each section from the similarity law; LB and UB are the lower and

upper limits for X respectively. In addition, after several pretests to ensure similarities in

geometry and having a better optimal calculation speed, the lower and upper bounds of the

variable were set to the range of plus or minus 1.9 times of X0. Moreover, AX ≤ b are the

constraint equations that will be expressed in details later. Ti is the unit thrust force at each

section, depending on the chord length (ci) and twist angle (θi) as shown below:

Ti ( X )  Ti (ci , i )

ci  x4  x3 ( r )  x2 ( r ) 2  x1 ( r )3
 R R R (12)

i  x8  x7 ( r )  x6 ( r ) 2  x5 ( r )3
 R R R
 X  R8

In all, we established a constrained optimization problem, and the pattern search method

was good for the calculations. The setting of constraints for optimization equation comprises

of two main guidelines. First, it is essential to have similar trends between distributions of

geometric parameters of the optimized blade and the geometric scaled blade. Second, to be

reasonable, the chord length should be small at the tip, and the twist angle cannot be far negative;

the twist angle in the first section and the chord length at the tip should not be too large.

Accordingly, linearly unequal constraints AX ≤ b can be characterized by the following details:

the twist angles should be within the range of (-9, 12) for the reliable Cl and Cd data, and the
18
Journal Pre-proof

chord length should be small but positive. Therefore, these constraint conditions can be

compiled as by Eq. (13)

b  9 12 0
T

 0.953 0.952 0.95 1 0 0 0 0  (13)


 
A   0.33 0.32 0.3 1 0 0 0 0
 0 0 0 0 1 1 1 1

3.3 Results of the optimization

Solving the optimization problem, we deduced the optimized distribution of the chord

length and twist angle, as shown in Fig. 7. Comparatively, we also plotted the geometric scaled

values with the corresponding fitting lines. The values for the optimized blade parameters are

also listed in Table 3. Moreover, Fig. 8 illustrates simulation results to compare the

performance difference between the optimized model turbine and the objective values

(theoretical scaled values). The thrust coefficient of the optimized blade is approximately the

same as that of the objective value. The power coefficient also is a relatively good match to the

prototype. It is important to note that the power coefficient for the lab-scale model is far less

than the theoretical scaled value for the Reynolds effects [22]. Therefore, the optimization

method of this study gave a good comparative results for the model turbine to the objective

values. In addition, as shown in Fig. 9, the optimized model turbine in this study has a higher

power coefficient than previous studies [14][36]. The observed inconsistency in Cp of the

optimized model and the prototype is as a result of the Reynolds number effect, and it requires

more implementations of high performance low-Re airfoils in the future. Moreover, it gives a

similar range of high performance of Cp as the prototype 2.5-MW turbine at utility-scale. The

high-performance range of TSR is approximately 5 to 7, which is the designed rated condition


19
Journal Pre-proof

of the model turbine as we mentioned before. It ensures that the optimized turbine operates at

the optimum TSR with the maximum power output under designed incoming flow condition.

28 35
Chord length (Optimized)
Chord length (Geometric scaled)
24 Chord length (fitting) 30
20
25
16
Chord length (mm)

Twist angle (deg)


12 20
8
15
4
0 10

-4
5
-8
0
-12 Twist angle (Optimized)
Twist angle (Geometric scaled)
-16 Twist angle (fitting) -5
-20
0.0 0.2 0.4 0.6 0.8 1.0
r/R

Fig. 7 Distribution of chord length and twist angle in the span wise direction
Table 3 Optimized model turbine blade parameters
Chord Chord
Twist angle Twist angle
Nodes r/R length Airfoils Nodes r/R length Airfoils
(deg) (deg)
(mm) (mm)
1 0.000 8.00 9.50 circle 15 0.627 11.53 5.44 NACA4412
2 0.036 8.00 9.50 circle 16 0.678 10.77 4.97 NACA4412
3 0.058 8.00 9.50 circle 17 0.714 10.40 4.73 NACA4412
4 0.093 8.00 9.50 circle 18 0.760 9.38 3.98 NACA4412
5 0.130 8.00 9.50 circle 19 0.808 8.84 3.59 NACA4412
6 0.171 9.63 9.50 DUINT760 20 0.833 8.46 3.32 NACA4412
7 0.219 23.14 11.12 DUINT280 21 0.857 7.80 2.88 NACA4412
8 0.271 21.26 9.84 DUINT230 22 0.884 7.15 2.48 NACA4412
9 0.329 19.27 8.79 DUINT190 23 0.910 6.34 2.01 NACA4412
10 0.384 17.53 8.00 TEST3415 24 0.935 5.41 1.52 NACA4412
11 0.447 15.28 7.09 NACA4412 25 0.959 4.44 1.04 NACA4412
12 0.497 14.62 6.82 NACA4412 26 0.980 3.45 0.00 NACA4412
13 0.533 13.43 6.34 NACA4412 27 0.995 0.29 1.82 NACA4412
14 0.579 12.50 5.93 NACA4412          

20
Journal Pre-proof

16 140

14 120

12 100

(N/m)
(N/m)

(W/m)
(W/m)
10 80

force
force

power
Unit power
8 60

thrust
Unit thrust

Normal
6 40
Normal
4 20

2 0
Objective values Objective values
Model values Model values
0 -20
0 0.5 1 0 0.5 1
r/R r/R

Fig. 8 Performance comparison between optimized model blades and prototype (target value)
0.6
Offset case-Hamilton2015
Base case-Hamilton2015
0.5 Tian2014
Present model
Prototype 2.5 MW turbine
0.4

0.3
Cp

0.2

0.1

0.0

0 2 4 6 8 10 12 14 16 18
TSR
Fig. 9 Comparison of the optimized model turbine with the CP of existing model turbines

As the development of wake can give a good demonstration of the model performance,

wake development simulations of the model and prototype turbines was deployed. It gives a

further verification of the similar design which can be seen intuitively from Fig. 10. The

development of the two wakes is roughly similar at near wake area, which is important for near

wake studies in wind tunnel. In addition, it should be mentioned that inflow velocity of both

model and prototype simulations are the same (11 m/s). Assuming that the wake development

length is 2.5 times the rotor diameter, the developing time of the model turbine wake is around

0.073 s, and that of the prototype wake is around 21.82 s. This gives a significant difference in

terms of the downstream wake structure at the far wake area as shown in Fig. 10, because of
21
Journal Pre-proof

the complex interactions in the wake flow.

Overall, from the above analysis, it can be concluded that the optimization model

established in this paper improves the similarity between the model turbine at lab-scale and the

prototype turbine at utility-scale in terms of aerodynamic performance and wake development.

Fig. 10 Comparison of the wake development of the model turbine and prototype turbine

4 Conclusions

It is very important to have a similar design of the model with wind tunnel experiments

for wind turbine studies. However, a similar design of blade performance for a small scale wind

turbine is rare, especially for scaled factor less than 1: 200. This paper describes the similar

design of the blade to enhance the similarity in aerodynamic performance and wake

characteristics of lab-scale wind turbine to its prototype. It improves the comparability between

the wind tunnel experiment of wind turbines at small lab-scale and the field measurements of

its prototype wind turbine at utility-scale, especially, the PIV measurement of entire wake flow

by wind tunnel experiment.

Due to the Reynolds number effect in the wind tunnel experiments for the studies of wake

area of wind turbines, a big performance gap exists between small lab-scale models and utility-

scale prototype. There is also lack of a practical method for the similar design of a small lab-

22
Journal Pre-proof

scale wind turbine model to match the wake flow of its prototype turbine. In this study, a new

method of blade design is proposed, using a real 2.5-MW turbine as the prototype. The model

rotor is 320 mm under a geometric scale ratio of 1:300. The comparison of aerodynamic

parameters of different types of design with objective values illustrates a dramatic deviation of

the geometric scaled model and a preferable performance of NACA4412 at low-Re condition.

From the validation, the application of the lifting-line theory with the improvement of wake-

induced corrections demonstrates an accurate evaluation of the aerodynamic performance of

the rotor. For the blade optimization design, a similar hybrid blade just like the prototype was

implemented, together with transition airfoil of TEST3415 design. The variation in geometric

characteristics along the span of the blade maintains similarities to its prototype. The tip speed

ratio runs at a roughly matched rated condition. In all, the model wind turbine with optimized

hybrid blade depicts a high performance in similarity to the prototype turbine, and gives a better

performance than other models used in previous wind tunnel studies.

ACKNOWLEDGEMENTS

This work has been supported financially by the National Natural Science Foundation of

China (Grant No.51708151) and the China Postdoctoral Science Foundation (Grant

No.2019T120274), which is gratefully acknowledged by the authors. We are also grateful for

the academic help provided by Dr. Jiarong Hong, Mr. Bingzheng Dou, Dr. Yun Liu, and Dr.

Teja Dasari, who made a lot of improvements in this work.

References
[1] Saidur R, Islam M, Rahim N, Solangi K. A review on global wind energy policy, Renewable and

Sustainable Energy Reviews, 14 (2010) 1744-1762. https://doi.org/10.1016/j.rser.2010.03.007.


23
Journal Pre-proof

[2] Hua Y, Oliphant M, Hu EJ. Development of Renewable Energy in Australia and China: a

Comparison of Policies and Status. Renewable Energy, 85(2016) 1044-1051.

https://doi.org/10.1016/j.renene.2015.07.060

[3] Rockel S, Peinke J, Hölling M, et al. Wake to wake interaction of floating wind turbine models

in free pitch motion: An eddy viscosity and mixing length approach, Renewable Energy, 85 (2016)

666-676. https://doi.org/10.1016/j.renene.2015.07.012.

[4] González-longatt Fall P, Terzija V. Wake Effect in Wind Farm Performance: Steady-state and

Dynamic Behavior. Renewable Energy, 39(2014) 329-338.

https://doi.org/10.1016/j.renene.2011.08.053

[5] Hand M, Simms D, Fingersh L, et al. Unsteady aerodynamics experiment phase Vi: wind tunnel

test configurations and available data campaigns. National Renewable Energy Laboratory, Golden,

Colorado, 2001. https://digital.library.unt.edu/ark:/67531/metadc625067

[6] Snel H, Schepers J, Montgomerie B. The Mexico Project (model experiments in controlled

conditions): the database and first results of data processing and interpretation. Journal of Physics:

Conference Series, IOP Publishing, 75(1) (2007) 012014.

https://iopscience.iop.org/article/10.1088/1742-6596/75/1/012014

[7] Cho T, Kim C. Wind Tunnel Test Results for a 2/4.5 Scale Mexico Rotor. Renewable Energy,

42(2012) 152-156. https://doi.org/10.1016/j.renene.2011.08.031

[8] Bottasso C, Campagnolo F, Petrović V. Wind tunnel testing of scaled wind turbine models:

beyond aerodynamics, Journal of Wind Engineering and Industrial Aerodynamics, 127 (2014) 11-

28. https://doi.org/10.1016/j.jweia.2014.01.009.

[9] Kimball R, Goupee A, Fowler M, et al. Wind/wave basin verification of a performance-matched

scale-model wind turbine on a floating offshore wind turbine platform//ASME 2014 33rd

International Conference on Ocean, Offshore and Arctic Engineering. American Society of

Mechanical Engineers, 2014.

[10] Ryi J, Rhee W, Hwang UC, et al. Blockage effect correction for a scaled wind turbine rotor by

using wind tunnel test data. Renewable Energy, 79(2015) 227-235.

https://doi.org/10.1016/j.renene.2014.11.057

[11] Lignarolo L, Ragni D, Krishnaswami C, et al. Experimental analysis of the wake of a horizontal-

axis wind-turbine model. Renewable Energy, 70 (Supplement C) (2014) 31-46.


24
Journal Pre-proof

https://doi.org/10.1016/j.renene.2014.01.020.

[12] Villegas A, Diez F. On the quasi-instantaneous aerodynamic load and pressure field

measurements on turbines by non-intrusive PIV. Renewable Energy, 2014, 63(1) 181-193.

https://doi.org/10.1016/j.renene.2013.09.015Get rights and content.

[13] Chamorro L, Porté-Agel F. Effects of thermal stability and incoming boundary-layer flow

characteristics on wind-turbine wakes: a wind-tunnel study. Boundary-Layer Meteorology, 136(3)

(2010) 515-533. https://doi.org/10.1007/s10546-010-9512-1

[14] Yuan W, Tian W, Ozbay A, et al. An experimental study on the effects of relative rotation

direction on the wake interferences among tandem wind turbines. Sci. China-physics. Mechanic &

Astronomy, 57(5) (2014) 935-949.

[15] Tian W, Ozbay A, Hu H. Effects of incoming surface wind conditions on the wake

characteristics and dynamic wind loads acting on a wind turbine model, Physics of Fluids 26 (2014)

125-108. https://doi.org/10.1063/1.4904375

[16] Hu H, Yang Z, Sarkar P. Dynamic wind loads and wake characteristics of a wind turbine model

in an atmospheric boundary layer wind. Experiments in Fluids, 52 (2012) 1277-1294.

https://doi.org/10.1007/s00348-011-1253-5.

[17] Mctavish S, Feszty D, Nitzsche F. A Study of the Performance Benefits of Closely-spaced Lateral

Wind Farm Configurations. Renewable Energy, 59(2013) 128-135.

https://doi.org/10.1016/j.renene.2013.03.032

[18] Mctavish S, Feszty D, Nitzsche F. An Experimental and Computational Assessment of Blockage

Effects on Wind Turbine Wake Development. Wind Energy, 17(2014) 1515-1529.

https://onlinelibrary.wiley.com/doi/10.1002/we.1648

[19] Cal RB, Lebrón J, Castillo L, et al. Experimental Study of the Horizontally Averaged Flow

Structure in a model wind-turbine array boundary layer. Journal of Renewable and Sustainable

Energy 2, (2010) 013106. https://doi.org/10.1063/1.3289735.

[20] Lebron J, Castillo L, Cal RB, et al. Interaction between a wind turbine array and a turbulent

boundary layer. 48th AIAA Aerospace Sciences Meeting Including the New Horizons Forum and

Aerospace Exposition, 4-7 (2010). https://doi.org/10.2514/6.2010-824

[21] Hamilton N, Tutkun M, Cal R. Low-order Representations of the canonical wind turbine array

boundary layer via double proper orthogonal decomposition, Physics of Fluids, 28 (2) (2016)
25
Journal Pre-proof

025103. https://doi.org/10.1063/1.4940659.

[22] McTavish S, Feszty D, Nitzsche F. Evaluating Reynolds number effects in small-scale wind

turbine experiments, Journal of Wind Engineering and Industrial Aerodynamics, 120 (2013) 81-90.

https://doi.org/10.1016/j.jweia.2013.07.006.

[23] Xiao J, Chen L, Xu B, Wu J. Investigation on aerodynamic performance of a 15 MW wind

turbine, Acta Aerodynamic Sinica, 29(4) (2011) 131-136 (in Chinese).

http://en.cnki.com.cn/Article_en/CJFDTOTAL-KQDX201104024.htm

[24] Marten D, Lennie M, Pechlivanoglou G, et al. Implementation, Optimization, and Validation

of a Nonlinear lifting line-free vortex wake module within the wind turbine simulation code Qblade.

Journal of Engineering for Gas Turbines and Power, 138(7) (2016) 72601.

https://doi.org/10.1115/1.4031872

[25] Martin H, Development of a scale model wind turbine for testing of offshore floating wind

turbine systems. Electronic Theses and Dissertations. (2011)

https://digitalcommons.library.umaine.edu/etd/1578

[26] Du W, Design and Analysis of a Model Wind Turbine Blade for Wave Basin Test of Floating

Wind Turbines. Shanghai Jiao Tong University, 2013 (in Chinese).

http://www.wanfangdata.com.cn/details/detail.do?_type=degree&id=D575663

[27] Frandsen S, Barthelmie R, Pryor S, et al. Analytical modelling of wind speed deficit in large

offshore wind farms. Wind Energy, 9 (1-2) (2006) 39-53, https://doi.org/10.1002/we.189

[28] Göçmen T, Van der Laan P, Réthoré P E, et al. Wind turbine wake models developed at the

technical university of Denmark: A review. Renewable and Sustainable Energy Reviews, 60 (2016)

752-769. https://doi.org/10.1016/j.rser.2016.01.113.

[29] Shakoor R, Hassan M Y, Raheem A, et al. Wake effect modeling: A review of wind farm layout

optimization using Jensen’s model, Renewable and Sustainable Energy Reviews, 58 (2016) 1048-

1059. https://doi.org/10.1016/j.rser.2015.12.229.

[30] Leishman J G. Challenges in modeling the unsteady aerodynamics of wind turbines, Wind

Energy, 5 (2-3) (2002) 85-132. https://doi.org/10.1002/we.62

[31] Chen Q. Three-dimensional PIV wind tunnel experimental study on wake characteristics of

horizontal-axis fans: [doctoral dissertation]. Wuhan university, 2014 (in Chinese).

[32] Anderson J, Fundamental of aerodynamics, McGraw-Hill, Boston. ISBN 0-07-237335-0, 360


26
Journal Pre-proof

(2001).

[33] Afjeh A. Wake effects on the aerodynamic performance of horizontal-axis wind turbines:

Toledo Univ., OH (USA), 1984.

[34] Lanzafame R, Messina M. Fluid dynamics wind turbine design: critical analysis, optimization

and application of BEM theory, Renewable Energy, 32 (14) (2007) 2291-2305.

https://doi.org/10.1016/j.renene.2006.12.010.

[35] Kim B, Kim W, Lee S, et al. Development and verification of a performance based optimal

design software for wind turbine blades, Renewable Energy, 54 (2013) 166-172.

https://doi.org/10.1016/j.renene.2012.08.029.

[36] Hamilton N, Melius M, Cal R. Wind Turbine Boundary Layer Arrays for Cartesian and Staggered

Configurations‐part I, Flow Field and Power Measurements, Wind Energy, 18 (2) (2015) 277-295.

http://dx.doi.org/10.1002/we.1697

27
Journal Pre-proof

Author Contribution List

B. Li: substantial contributions to conception, design, analysis, and drafting the

article.

D. L. Zhou: substantial contributions to acquisition of data, interpretation of data

and involving in the drafting.

Y. Wang: substantial contributions to involving in acquisition of data, drafting and

revising the article.

Y. Shuai: substantial contributions to conception, revising it critically for important

intellectual content.

Q. Z. Liu: substantial contributions to analysis and interpretation of data, co-

approval of the version to be published.

W. H. Cai: substantial contributions to design, final approval of the version to be

published
Journal Pre-proof

The design of a small lab-scale wind turbine model with high


performance similarity to its utility-scale prototype

B. Li1, D. L. Zhou1, Y. Wang2, Y. Shuai1, Q. Z. Liu1*, W. H. Cai2*

1 Harbin Institute of Technology, School of Energy Science and Engineering, 92 West Dazhi

Rd, Harbin, China

2 Northeast Electric Power Univeristy, School of Energy and Power Engineering, 169

Changchun Rd, Jilin City, China

Corresponding author:

W. H. Cai, E-mail: caiwh@neepu.edu.cn Tel.: +86-18603617391

Address: School of Energy and Power Engineering, 169 Changchun Rd, Jilin City, China.

Q. Z. Liu, E-mail: liuquanzhong@hit.edu.cn Tel.: +86-18686860762

Address: School of Energy Science and Engineering, 92 West Dazhi Rd, Harbin, China

Acknowledgments

This work has been supported financially by the National Natural Science Foundation of

China (Grant No.51708151) and the China Postdoctoral Science Foundation (Grant

No.2019T120274), which is gratefully acknowledged by the authors. We are also grateful for

the academic help provided by Dr. Jiarong Hong, Mr. Bingzheng Dou, Dr. Yun Liu, and Dr. Teja

Dasari, who made a lot of improvements in this work.


Journal Pre-proof

A new small lab-scale wind turbine similar to the prototype has been designed.

The design process provides a hybrid model blade similar to the prototype.

Performance of geometric scaled model is significantly less than the objective value.

A better performance of the optimized rotor has been comparably reached.

You might also like