You are on page 1of 13

Journal of Wind Engineering & Industrial Aerodynamics 217 (2021) 104726

Contents lists available at ScienceDirect

Journal of Wind Engineering & Industrial Aerodynamics


journal homepage: www.elsevier.com/locate/jweia

Comparative assessment of transitional turbulence models for airfoil


aerodynamics in the low Reynolds number range
Yue Liu a , Peifeng Li b , Kaiyong Jiang a ,∗
a Fujian Key Laboratory of Special Energy Manufacturing, Huaqiao University, Fujian, China
b
James Watt School of Engineering, University of Glasgow, Glasgow, United Kingdom

ARTICLE INFO ABSTRACT

Keywords: A comparative assessment of the prediction capability of transitional turbulence models was carried out in the
Transitional turbulence models chord Reynolds number range 𝑅𝑒𝑐 = 2.3 × 104 − 2 × 105 with the NACA 0012 and NACA 4415 airfoils. Four
Airfoil aerodynamics transition models based on the 𝑘 − 𝜔 framework, i.e., the low-Reynolds number correction, 𝑅𝑒𝜃 − 𝛾, 𝛾 and
Computational fluid dynamics
𝑘 − 𝑘𝑙 − 𝜔 were chosen to simulate the flow and aerodynamic characteristics. It is found that the prediction
Angle of attack
capability of a model is related to the nature of the flow structure and affected by the 𝑅𝑒𝑐 and angle of attack
Small wind turbines
(𝐴𝑂𝐴). Among all the models, the 𝑅𝑒𝜃 − 𝛾 model has the best prediction capability and is recommended for
low 𝑅𝑒𝑐 applications such as small wind turbines. The 𝑘 − 𝑘𝑙 − 𝜔 model has the secondary capability for some
𝑅𝑒𝑐 and low 𝐴𝑂𝐴𝑠. The 𝛾 model has a similar frame to the 𝑅𝑒𝜃 − 𝛾 model but cannot reliably predict the flow
at larger 𝐴𝑂𝐴𝑠 and lower 𝑅𝑒𝑐 . The low-Reynolds number correction model is not recommended for the low
𝑅𝑒𝑐 range.

1. Introduction simulation (DNS) or large eddy simulation (LES). But the relatively
high computational cost restricts their extensive industrial applications.
Pollution caused by traditional energy supplying systems brings A semi-empirical transition modeling approach based on the linear
urgent needs to develop green energy. Among the renewable energy stability theory — the eN method can rapidly predict the airfoil per-
schemes, wind turbine farms have had rapid growth for the past formance in transition (Ingen, 2008). But the semi-empirical model
decades. But the potential impact of wind turbine farms on climate only predicts the onset of transition and is difficult to implement in
conditions leads to the risk of global warming. An alternative scheme
a general-purpose CFD code. The Reynolds-averaged Navier–Stokes
to meet increasing power demand without substantial side effects is
(RANS) method coupled with transition models can be an effective tool
to employ decentralized small wind turbines (Tummala et al., 2016;
to predict large-scale flow structures where averaged turbulent param-
Baidya Roy, 2011).
Compared to large wind turbines, the small-scale characteristic of eters are important. This method has thus been actively developed for
small wind turbines degrades their service performance, as they work engineering aerodynamics. Di Pasquale et al. (2009) reviewed eight
under a low chord Reynolds number condition (1 × 104 < 𝑅𝑒𝑐 < different methods to model transition and pointed out that transition
5×105 ). The 𝑅𝑒𝑐 can be classified into ultra-low, low, moderate and high models can be implemented into a general RANS environment with
regimes in terms of the dependence of maximum lift coefficient, 𝐶𝐿,𝑚𝑎𝑥 , sufficient engineering accuracy.
on 𝑅𝑒𝑐 (Wang et al., 2014). Under low 𝑅𝑒𝑐 conditions, complex scenery A number of previous works focused on transition models based on
including laminar separation, transition, reattachment and turbulent the framework of 𝑘−𝜔 two-equation turbulence model and investigated
separation can occur in the boundary layer (Gad-el Hak, 1990), and their capability to predict low Reynolds number airfoil flows (Langtry
long laminar separation bubbles frequently form on airfoil surfaces. et al., 2006; Wolgemuth and Walters, 2007; Walters and Cokljat, 2008;
These complex phenomena subsequently deteriorate the aerodynamics Langtry and Menter, 2009; Fürst et al., 2013; Khayatzadeh and Nadara-
of small wind turbines. jah, 2014; Menter et al., 2015; Choudhry et al., 2015; Lin and Sarlak,
Computational fluid dynamics (CFD) is a promising and reliable tool
2016; Xia and Chen, 2016; Zhang et al., 2017; Sanei and Razaghi,
to study the aerodynamics of small wind turbines. But the difficulty
2018; Wauters et al., 2019; Lopes et al., 2020; Liu et al., 2020). These
arises when it models the transition in the low Reynolds number range.
airfoils included the classic NACA 0012, the relatively thick NACA
The transition process can generally be solved by direct numerical

∗ Corresponding author.
E-mail address: jiangky@hqu.edu.cn (K. Jiang).

https://doi.org/10.1016/j.jweia.2021.104726
Received 11 January 2021; Received in revised form 15 July 2021; Accepted 15 July 2021
Available online 2 August 2021
0167-6105/© 2021 Elsevier Ltd. All rights reserved.
Y. Liu et al. Journal of Wind Engineering & Industrial Aerodynamics 217 (2021) 104726

Fig. 1. (a) The 2D computational domain with boundary conditions, and details of the grids (b) in the computational domain, (c) around the airfoil surface, (d, e) around the
leading and trailing edges.

0018, NACA 0021, S809, S862, and DU91-W2-250 for medium and the low 𝑅𝑒𝑐 = 104 − 105 . The influence of inflow conditions on the flow
large wind turbines, and the low Reynolds number airfoils E387 and in this Reynolds number range was not well documented. Moreover,
SD7003. The 𝑅𝑒𝑐 of these airfoils ranged from 6 × 104 − 3 × 106 , mainly there is still a lack of studies regarding the prediction capability of these
in the order of 105 –106 . Their results demonstrate that the transition models for different 𝑅𝑒𝑐 and 𝐴𝑂𝐴𝑠.
models are able to predict the transition and the formation of laminar In this paper, the RANS method was coupled with four frequently
separation bubbles on airfoil surfaces at certain 𝑅𝑒𝑐 numbers. For larger used transition models in the 𝑘 − 𝜔 SST (shear stress transport) turbu-
𝑅𝑒𝑐 (106 ), the predicted time-averaged pressure (𝐶𝑝 ) distribution on lence framework. The inflow conditions and the related decay behavior
airfoil surface coincides with the experimental data. However, for small of the models were first discussed. The prediction capability of the
wind turbines, the interested 𝑅𝑒𝑐 range is relatively low in the order of four models was then compared and evaluated based on the results
104 –105 , and the used airfoil such as NACA 4415, NACA 4412, E387, of the time-averaged pressure distribution on the airfoil surfaces, the
SG6040, SH3055 and S822 is relatively thin to gain more lift. These separation, transition and reattachment locations, the instantaneous
conditions have not been broadly investigated in the previous studies vorticity and the time-averaged streamlines.
to the authors’ knowledge.
Transition models commonly implemented into the RANS method 2. Problem definition and transition models
include low-Reynolds number damping models (low-Reynolds number
correction, LowRe), correlation-based models with intermittency (𝑅𝑒𝜃 − 2.1. Flow configuration
𝛾 and 𝛾) and physics-based models (𝑘 − 𝑘𝑙 − 𝜔). It has been reported
that compared to the original full turbulence model, a correlation-based The NACA 4415 and NACA 0012 airfoils were selected for the
transition model is more reliable especially in small angles of attack simulations, as the NACA 4415 airfoil is commonly used in small
(𝐴𝑂𝐴𝑠) due to its capability to predict separation bubbles (Wolgemuth wind turbines and the NACA 0012 airfoil has plenty of experimental
and Walters, 2007; Sanei and Razaghi, 2018). But according to the and simulation data to verify the simulations. The simulations were
results of Wauters et al. (2019), the low-Reynolds number damping carried out in 2D to reduce the computational cost. The RANS method
model fails to predict laminar separation bubbles while the 𝑅𝑒𝜃 − 𝛾, averages out the 3D effect especially in small 𝐴𝑂𝐴𝑠 as reported by
𝛾 and 𝑘 − 𝑘𝑙 − 𝜔 models well simulate the flow in lower 𝐴𝑂𝐴𝑠. In other researchers (Wauters et al., 2019; Lopes et al., 2020). The same
particular, the flow after turbulence separation can be better predicted flow domain, boundary conditions and grid topology were applied
by the 𝑅𝑒𝜃 − 𝛾 and 𝛾 models. It was also reported that 𝑘 − 𝑘𝑙 − 𝜔 to both airfoils as illustrated in Fig. 1. The velocity inlet boundary
can better predict laminar separation bubbles than 𝑅𝑒𝜃 − 𝛾 (Choudhry condition is applied to the front, top and bottom boundaries as it
et al., 2015; Lin and Sarlak, 2016). Lopes et al. (2020) pointed out that can give a reasonable result with less computational efforts, and the
the correlation-based (𝑅𝑒𝜃 − 𝛾 and 𝛾) and physics-based (𝑘 − 𝑘𝑙 − 𝜔) flow is considered to be incompressible at low Reynolds numbers. The
models are sensitive to inlet boundary conditions, but the 𝑘 − 𝑘𝑙 − 𝜔 simulations using different transition models were compared with the
model cannot accurately simulate the flow in 𝑅𝑒𝑐 < 105 . However, these available data in the literature (Table 1) to evaluate their prediction
comparative works except that by Lopes et al. (2020) did not consider capability.

2
Y. Liu et al. Journal of Wind Engineering & Industrial Aerodynamics 217 (2021) 104726

Table 1
Experimental and numerical data to verify the transition models.
Case Airfoil 𝑅𝑒𝑐 (×105 ) 𝐴𝑂𝐴 (◦ ) 𝑇 𝑢∞ (%) Method
(Saliveros, 1988) NACA 4415 0.5, 0.75, 1, 2 0, 4, 8, 12 0.4 Experimental
(Jones et al., 2018) NACA 4415 0.5 0 – DNS
(Kim et al., 2009) NACA 0012 0.23, 0.48 6 (5.5) <0.4 Experimental
(Jones et al., 2008) NACA 0012 0.5 5 – DNS

2.2. Transition models in the 𝑘 − 𝜔 SST turbulence framework Table 2


Predicted drag and lift coefficients with three different grid densities.
Case Number of cells Ns Nr 𝐶𝑑 𝐶𝑙
The transition models extensively used in general-purpose indus-
trial CFD simulations are based on the RANS 𝑘 − 𝜔 SST turbulence Fine 462747 1370 290 0.0305 1.3178
Medium 230907 1010 200 0.0315 1.3177
framework. The RANS equations are derived based on Reynolds de- Coarse 114212 690 145 0.0321 1.3333
composition. An instantaneous turbulent quantity 𝑢𝑖 is decomposed (Saliveros, 1988) – – – 0.03 1.3
into a time-averaged part 𝑢̄𝑖 and a fluctuating part 𝑢′𝑖 . The additional
Reynolds stress terms −𝑢′𝑖 𝑢′𝑗 derived in the RANS equations would
close with the linear eddy viscosity model (Boussinesq hypothesis). The
models described above were implemented in the solver. The discretiza-
𝑘 − 𝜔 SST turbulence model is a modified 𝑘 − 𝜔 based eddy viscos-
tion accuracy was second order for all terms. The non-dimensional time
ity model. The model introduces an improved definition of the eddy
step 𝛥𝑡 = 𝛿𝑡𝑈∞ ∕𝑐 was set as 0.0005 for the 𝑅𝑒𝜃 − 𝛾 and 𝛾 models, and
viscosity (Menter, 1994), considers the transport of turbulence shear
reduced to 0.0001 for the 𝑘 − 𝑘𝑙 − 𝜔 and LowRe models to meet the
stresses for applications with flow separation, and therefore avoids
convergence criteria. The residual for all parameters was 10−5 as it can
overpredicting the eddy viscosity. This model is mainly appropriate for
give sufficient accuracy in reasonable computational time.
fully turbulent flows but not for transitional flows (Rumsey and Spalart,
The grid density dependency was studied for the NACA 4415 airfoil
2009). Therefore, transition models should be coupled into the 𝑘 − 𝜔
using the 𝑘 − 𝑘𝑙 − 𝜔 model with the highest 𝑅𝑒𝑐 number, 𝑅𝑒𝑐 = 2 × 105 ,
SST model to predict the transition effect under low-Reynolds numbers.
and medium 𝐴𝑂𝐴 = 10◦ . The adjacent grid density (Table 2) was
The four transition models used in this paper are briefly described as
doubled by the total number of cells, and controlled by the number of
follows.
nodes along the airfoil surface (Ns) and the number of nodes along the
The low-Reynolds number correction model (LowRe) was first intro-
radial direction away from the airfoil (Nr). The predicted time-averaged
duced by Wilcox (1994). The turbulent viscosity 𝜈𝑡 , typically defined
drag (𝐶𝑑 ) and lift (𝐶𝑙 ) coefficients are compared with the experimental
as 𝜈𝑡 = 𝑘∕𝜔, is artificially damped with a coefficient 𝛼 ∗ to simulate
data by Saliveros (1988). Moreover, Fig. 2 compares the time-averaged
low-Reynolds number phenomena. In this framework, a laminar flow
pressure coefficient (𝐶𝑝 ) and wall shear stress (𝜏𝑤𝑎𝑙𝑙−𝑥 ) predicted using
is described as the eddy viscosity ratio 𝜈𝑡 ∕𝜈 ≪ 1, the onset of transition
the three grids. Both the fine and medium grid densities produce nearly
is approximated as 𝜈𝑡 ∕𝜈 ≈ 1, and the turbulent flow is fully developed
identical simulation results. According to the previous study on the
as 𝜈𝑡 ∕𝜈 ≫ 1.
chordal node number (Wauters et al., 2019), the Ns > 200 could give
The correlation-based transition SST model (𝑅𝑒𝜃 − 𝛾) (Langtry and a relatively similar prediction of 𝐶𝑑 for all the transition models. The
Menter, 2009) has been developed to sufficiently capture the ma- predicted difference in 𝐶𝑑 between the models using fine and coarse
jor effects of transition triggered by different mechanisms (natural, grids is 16 drag counts (1 drag count is equal to a 𝐶𝑑 = 0.0001), which
bypass and separation induced). The concept of a local variable (vor- can be considered acceptable as it is less than 20 counts according
ticity Reynolds number) relates the non-local transition onset Reynolds to Wauters et al. (2019). The medium grid was thus chosen for the
number and the local boundary-layer quantities, both of which deter- simulations.
mine the flow status together. This overcomes the shortcoming of the The first layer of the grid to the wall was set to 𝛿𝑦 = 10−4 𝑐 and
non-local formulation of other correlation-based models. 4×10−5 𝑐 (𝑐 is the airfoil chord) to study the sensitivity of dimensionless
The intermittency model (𝛾) was formulated by Menter et al. (2015). wall distance 𝑦+ (calculated with local wall friction velocity). A wall-
The 𝛾 model is similar to the 𝑅𝑒𝜃 − 𝛾 model. But In the 𝛾 model, normal grid expansion ratio of 1.1 was utilized in both cases. Fig. 3
the transition source term 𝑝𝛾 is slightly simplified. Moreover, instead shows the 𝑦+ and time-averaged 𝐶𝑝 distribution along the airfoil surface
of using a transport equation, a critical Reynolds number 𝑅𝑒𝜃𝑐 is with the two different 𝛿𝑦. As the simulation was done with the highest
calculated algebraically to avoid the use of the streamwise velocity 𝑈 Reynolds number and medium 𝐴𝑂𝐴, the 𝑦+ value is less than 1.0 in
which causes the Galilean invariance problem in the 𝑅𝑒𝜃 − 𝛾 model. most simulation cases. The similar results of time-averaged 𝐶𝑝 suggest
The critical 𝑅𝑒𝜃𝑐 is the momentum-thickness Reynolds number where that 𝑦+ < 2 can give an accurate result, and 𝛿𝑦 = 10−4 𝑐 was finally
the intermittency first starts to increase in the boundary layer. chosen for the simulations.
The 𝑘−𝑘𝑙 −𝜔 model is based on the phenomenon that pre-transition
laminar fluctuations can lead to turbulence breakdown. The concept of 3. Results and discussion
laminar kinetic energy (Mayle and Schulz, 1996) has been implemented
in the 𝑘 − 𝜔 framework to predict the transition region (Walters and 3.1. Inflow conditions and decay behavior
Cokljat, 2008). In the 𝑘−𝑘𝑙 −𝜔 model, Reynolds stresses are determined
by both the turbulent and laminar fluctuations. Although the 𝑘 − 𝑘𝑙 − 𝜔 In the low-Reynolds number range, the free-stream turbulence char-
model is based on the 𝑘−𝜔 model, most terms in the transport equations acteristics can significantly affect the flow field, so the boundary con-
are modified. The lack of consideration of separation-induced transition ditions should be properly set. The turbulence intensity (𝑇 𝑢) and eddy
in the 𝑘 − 𝑘𝑙 − 𝜔 model may cause the incapability of prediction in some viscosity ratio (𝜈𝑡 ∕𝜈) were chosen to represent the free-stream turbu-
situations. lence characteristic in this paper. In the simulation of the external
airfoil flow, the domain around the airfoil should be large enough
2.3. Numerical details to obtain the free-stream condition at the outer boundaries, and thus
eliminate the boundary influence on simulation results. The turbulence
The simulations were performed in the Ansys Fluent solver, a gen- intensity set at the inlet boundary would decay rapidly as the flow
eral CFD code based on the finite volume method. The transition approaches the airfoil leading edge due to the destruction term in

3
Y. Liu et al. Journal of Wind Engineering & Industrial Aerodynamics 217 (2021) 104726

Fig. 2. Predicted time-averaged values of (a) pressure coefficient 𝐶𝑝 and (b) wall shear stress 𝜏𝑤𝑎𝑙𝑙−𝑥 with three different grid densities.

Fig. 3. Sensitivity study of (a) 𝑦+ distribution and (b) time-averaged 𝐶𝑝 to first layer distance 𝛿𝑦 along the airfoil surface.

the transport equations. To avoid this inconsistency, the boundary Table 3


The setting of inflow turbulence intensity.
conditions set at the inlet should be adjusted to meet the experimental
𝑇 𝑢𝑖𝑛 (%) 𝜈𝑡 ∕𝜈 𝑇 𝑢(𝑋) (%)
or atmospheric conditions. The decay of 𝑇 𝑢 level based on the 𝑘 − 𝜔
framework can be calculated (Spalart and Rumsey, 2007): 1 5 0.5890
0.1 5 0.0990
[ ( ) ]−𝛽 ∗ ∕𝛽 ⎞0.5 0.01 5 0.0100
⎛ 1.5𝑋𝛽𝑅𝑒𝑐 𝑇 𝑢2𝑖𝑛
𝑇 𝑢(𝑋) = ⎜𝑇 𝑢2𝑖𝑛 1 + ( ) ⎟ (1)
⎜ 𝜈𝑡 ∕𝜈 𝑖𝑛 ⎟
⎝ ⎠
with 𝛽 ∗ = 0.0828, 𝛽 = 0.09 and 𝑋 = 𝑥∕𝑐 is the non-dimensional distance 3.1.1. Effect of free-stream turbulence intensity
from the inlet to the airfoil leading edge. According to Eq. (1), the decay Breuer (2018) utilized LES to study the effect of inflow turbulence
rate increases with 𝑅𝑒𝑐 . Higher turbulence intensity 𝑇 𝑢𝑖𝑛 and lower intensity (0% ≤ 𝑇 𝑢∞ ≤ 11.2%) on the flow around the SD7003 airfoil
eddy viscosity ratio 𝜈𝑡 ∕𝜈 lead to rapid decay. at 𝑅𝑒𝑐 = 6 × 104 and 𝐴𝑂𝐴 = 4◦ . The results indicate that the transition
region in the flow field could be very different especially under 𝑇 𝑢∞ ≤
Fig. 4 illustrates the influence of other parameters on the decay. The
1%. The 𝑇 𝑢∞ is physically the velocity fluctuations in the free-stream.
comparison is demonstrated by the 𝑘 − 𝑘𝑙 − 𝜔 and 𝑅𝑒𝜃 − 𝛾 models. The
But in the√ RANS, the velocity fluctuation is not directly solved and the
other two models (LowRe and 𝛾) have the same destruction term as the
𝑇 𝑢∞ = 2𝑘∕3 is converted to the kinetic energy 𝑘 to represent the
𝑅𝑒𝜃 − 𝛾 model and they are expected to have the same decay behavior.
turbulence characteristic. The influence of 𝑇 𝑢∞ is distinct among the
Fig. 4(a) shows that the decay behavior is not influenced by angles of
results of LES, DNS and RANS, especially when the transition occurs.
attack. Nearly identical decay rates are observed for both models. As
Different 𝑇 𝑢𝑖𝑛 values were set with a fixed viscosity ratio 𝜈𝑡 ∕𝜈 = 5
the total kinetic energy is determined by the laminar (𝑘𝑙 ) and turbulent (Table 3) to compare the model sensitivity to the free-stream turbulence
(𝑘𝑡 ) kinetic energies in the 𝑘 − 𝑘𝑙 − 𝜔 model, the 𝑘𝑙 value (8 × 10−6 ) intensity and to help specify their values at the inlets for the RANS
set at the inlet causes an increase on 𝑇 𝑢. The 𝑇 𝑢 suddenly increases transition turbulence models. The inlet laminar kinetic energy 𝑘𝑙 was
near the leading edge in the 𝑘 − 𝑘𝑙 − 𝜔 model as this model includes set to be zero in the 𝑘 − 𝑘𝑙 − 𝜔 model to ensure the identical 𝑇 𝑢 at
additional near-wall anisotropic dissipation terms. Fig. 4(b) indicates the boundaries. Table 3 also lists the corresponding 𝑇 𝑢 at the leading
that the flow time does not influence the decay behavior while the edge of the airfoil calculated by Eq. (1). The NACA 0012 airfoil at
decay rate increases with 𝑇 𝑢𝑖𝑛 . The laminar kinetic energy does not 𝑅𝑒𝑐 = 5 × 104 and 𝐴𝑂𝐴 = 5◦ was chosen to study the effect of 𝑇 𝑢∞ and
decay as there is no destruction term in the laminar transport equation 𝜈𝑡 ∕𝜈 as it gives nearly identical results of time-averaged 𝐶𝑝 distribution
(Fig. 4(c)). when 𝑇 𝑢𝑖𝑛 = 0.1% for all models.

4
Y. Liu et al. Journal of Wind Engineering & Industrial Aerodynamics 217 (2021) 104726

Fig. 4. (a) Decay behavior of different 𝐴𝑂𝐴𝑠 predicted by the 𝑘 − 𝑘𝑙 − 𝜔 and 𝑅𝑒𝜃 − 𝛾 models. (b) Decay behavior of different flow time 𝑇 and 𝑇 𝑢𝑖𝑛𝑙𝑒𝑡 predicted by the 𝑘 − 𝑘𝑙 − 𝜔
model. (c) The turbulent kinetic energy and laminar kinetic energy in the 𝑘 − 𝑘𝑙 − 𝜔 model under 𝑅𝑒𝑐 = 2 × 105 .

Figs. 5(a) and 5(b) shows the sensitivity to inflow turbulence inten- Table 4
The setting of eddy viscosity ratio.
sity in the four models. The 𝐶𝑝 distribution is almost identical for all
𝑇 𝑢𝑖𝑛 (%) 𝜈𝑡 ∕𝜈 𝑇 𝑢(𝑋) (%)
models under lower 𝑇 𝑢 (0.01–0.1%). For the high 𝑇 𝑢 level(> 0.5%), the
low Re and 𝑅𝑒𝜃 − 𝛾 models predict more distinct transition in the flow 0.724 10 0.5890
1 5 0.5890
field compared to the low 𝑇 𝑢 level (Fig. 5(b)). This suggests that the
1171.2 1 0.5890
𝑇 𝑢𝑖𝑛 should be carefully set in the LowRe and 𝑅𝑒𝜃 − 𝛾 models. Both
the 𝑘 − 𝑘𝑙 − 𝜔 and 𝛾 models seem insensitive to the inflow turbulence
intensity.
obtain a high 𝑇 𝑢(𝑋), a small eddy viscosity ratio (𝜈𝑡 ∕𝜈 = 1) can lead to
an unrealistic high 𝑇 𝑢𝑖𝑛 due to the rapid decay.
3.1.2. Effect of eddy viscosity ratio
Fig. 5(c) shows the influence of the eddy viscosity ratio on the 𝐶𝑝
The eddy viscosity ratio at boundaries triggers turbulence in RANS
distribution in the transition region. Both the 𝑅𝑒𝜃 − 𝛾 and 𝛾 models are
as 𝜈𝑡 determines the Reynolds shear stress. Generally, a proper eddy
insensitive to the eddy viscosity ratio. However, very different results in
viscosity ratio should be defined to obtain realistic results. Spalart and
the transition were predicted by the LowRe model using different 𝜈𝑡 ∕𝜈.
Rumsey (2007) recommended the 𝜈𝑡 ∕𝜈 ≈ 2 × 10−7 × 𝑅𝑒𝑐 for a full
The results of the 𝑘 − 𝑘𝑙 − 𝜔 model are only slightly different between
turbulence model at a Reynolds number in the order of 106 − 107 . If
𝜈𝑡 ∕𝜈 = 5 and 𝜈𝑡 ∕𝜈 = 10. Although the LowRe and 𝑘 − 𝑘𝑙 − 𝜔 models show
this recommendation is applied to the lower 𝑅𝑒𝑐 range (104 − 105 ), some sensitivity to 𝜈𝑡 ∕𝜈, the influence of 𝜈𝑡 ∕𝜈 on the predicted results
the resultant small 𝜈𝑡 ∕𝜈 (10−3 − 10−2 ) can cause a very rapid decay, is less than the effect of the 𝑇 𝑢𝑖𝑛 level. The value of 5 < 𝜈𝑡 ∕𝜈 < 10 could
especially with a relatively higher 𝑇 𝑢. The 𝜈𝑡 ∕𝜈 recommended by Ansys give reasonable simulation results in the low-Reynolds number range.
Fluent is 1 − 10 and has been used in many simulations of low Reynolds
number airfoil flow (Walters and Cokljat, 2008; Langtry and Menter, 3.2. Prediction capability at different Reynolds numbers and angles of
2009; Fürst et al., 2013; Khayatzadeh and Nadarajah, 2014; Lopes attack
et al., 2020). To evaluate the eddy viscosity ratio effect, different 𝜈𝑡 ∕𝜈
values were set with a fixed 𝑇 𝑢(𝑋) = 0.5890% at the leading edge 3.2.1. Case of 1 × 104 < 𝑅𝑒𝑐 < 5 × 104
of the airfoil (Table 4). The relatively high 𝑇 𝑢 level was set as the At the Reynolds number between 1 × 104 and 5 × 104 , the airfoil
environmental turbulence intensity can be high in the application of aerodynamic performance is poor due to the large separation of the
wind turbines (Devinant et al., 2002). Table 4 also illustrates that to boundary layer (without reattachment or with reattachment to form a

5
Y. Liu et al. Journal of Wind Engineering & Industrial Aerodynamics 217 (2021) 104726

Fig. 5. Predicted time-averaged 𝐶𝑝 distribution on the NACA 0012 airfoil surface at 𝑅𝑒𝑐 = 5 × 104 and 𝐴𝑂𝐴 = 5◦ using different inflow parameters: (a) overview and (b)(c) details
near the transition range. Turbulence intensity 𝑇 𝑢 in (c) is different from those in (a) and (b).

long laminar separation bubble). The flow is sensitive to environmental 3.2.2. Case of 5 × 104 < 𝑅𝑒𝑐 < 5 × 105
disturbances and a variation of 𝐴𝑂𝐴𝑠. Fig. 6 shows the mean pressure The four models were rerun for the NACA 4415 airfoil with 𝑅𝑒𝑐 =
coefficient distribution predicted by the four models and compared to 0.75×, 1.0×, 2.0 × 105 and 𝐴𝑂𝐴 = 0◦ , 4◦ , 8◦ , 12◦ . The predicted
the experimental and/or DNS data in this 𝑅𝑒𝑐 range. time-averaged 𝐶𝑝 distribution on the airfoil surface is compared to
Under the condition shown in Figs. 6(a) and 6(b), the boundary the experimental results of Saliveros (1988) (Fig. 7). Fig. 8 illustrates
layer separates from the airfoil surface without reattachment. Similar the predicted instantaneous vorticity at the non-dimensional time 𝑇 =
results are predicted by the LowRe and 𝑘 − 𝑘𝑙 − 𝜔 models. The 𝑅𝑒𝜃 − 𝛾 40 (𝑇 = 𝑡𝑈∞ ∕𝑐) and time-averaged streamlines (averaged between
and 𝛾 models give similar results but with a difference in the region T=20–40) at 𝐴𝑂𝐴 = 0◦ under different 𝑅𝑒𝑐 .
close to the trailing edge. The 𝛾 model tends to predict higher negative With 𝑅𝑒𝑐 ≤ 1 × 105 (Figs. 7(a) and 7(b)), the results predicted by
time-averaged 𝐶𝑝 close to the trailing edge. It may be hard to tell which four models are similar at lower 𝐴𝑂𝐴𝑠(0 − 4◦ ). Slight difference occurs
model has better predictions under this condition. For the NACA 4415 near the trailing edge on the suction surface. The 𝑘 − 𝑘𝑙 − 𝜔 model
airfoil (Fig. 6(a)), the scatter in experimental data makes it difficult to predicts higher negative time-averaged 𝐶𝑝 near the trailing edge than
compare the four models. The 𝑅𝑒𝜃 − 𝛾 and 𝛾 models agree qualitatively other models. As the angle of attack increases, the predicted results of
better with the experiments. The 𝛾 model gives a better prediction each model deviate more from each other and all models fail to predict
compared to the DNS data (Jones et al., 2018). For the NACA 0012 at the large 𝐴𝑂𝐴 (≥ 12◦ ). The results predicted by the LowRe and 𝛾
airfoil, the LowRe and 𝑘 − 𝑘𝑙 − 𝜔 models seem to perform better models are similar but different from the experimental data in all the
(Fig. 6(b)). cases. The 𝑘 − 𝑘𝑙 − 𝜔 model performs better than the LowRe and 𝛾
Under the condition of 𝑅𝑒𝑐 ≈ 5 × 104 and 𝐴𝑂𝐴 = (5, 5.5)◦ for models. The 𝑅𝑒𝜃 − 𝛾 model has the best prediction results among the
the NACA 0012 airfoil (Figs. 6(c) and 6(d)), the laminar boundary four models.
layer separates near the leading edge and reattaches to form a long At the Reynolds number 𝑅𝑒𝑐 = 2 × 105 (Fig. 7(c)), the prediction
laminar separation bubble. The simulation results are compared to the capability of all four models enhances significantly compared to the
experimental and DNS results. The prediction capability improves com- 𝑅𝑒𝑐 ≤ 1 × 105 conditions. Almost identical results are predicted by the
pared to the condition without reattachment. But all the models tend 𝑅𝑒𝜃 − 𝛾, 𝛾 and 𝑘 − 𝑘𝑙 − 𝜔 model at small 𝐴𝑂𝐴𝑠. But the LowRe model
to underestimate the negative time-averaged 𝐶𝑝 over the long laminar fails to predict the transition at all 𝐴𝑂𝐴𝑠. At 𝐴𝑂𝐴 ≥ 12◦ , the 𝑘 − 𝑘𝑙 − 𝜔
separation bubble. The 𝑘−𝑘𝑙 −𝜔 model predicts a similar behavior to the model deviates from the experimental data as the turbulence separation
2d-DNS (Fig. 6(d)). In the low Reynold number range, some anisotropic appears.
behavior may occur especially in the separation region (over the long The prediction of the four models can be further compared by the
laminar separation bubble). The isotropic assumption in the RANS may boundary layer characteristic shown in Fig. 8. At 𝑅𝑒𝑐 = 0.75 × 105
fail to predict such behavior. and 1 × 105 (Figs. 8(a) and 8(b)), the LowRe and 𝛾 models tend to

6
Y. Liu et al. Journal of Wind Engineering & Industrial Aerodynamics 217 (2021) 104726

Fig. 6. Time-averaged 𝐶𝑝 distribution with four models on (a) the NACA 4415 airfoil surface at 𝑅𝑒𝑐 = 5 × 104 , 𝐴𝑂𝐴 = 0◦ and the NACA 0012 airfoil surface at (b)
𝑅𝑒𝑐 = 2.3 × 104 , 𝐴𝑂𝐴 = 5.5◦ , (c) 𝑅𝑒𝑐 = 4.8 × 104 , 𝐴𝑂𝐴 = 5.5◦ , (d) 𝑅𝑒𝑐 = 5 × 104 , 𝐴𝑂𝐴 = 5◦ .

Table 5
Summary of flow structures and the recommended model under different conditions (𝑅𝑒𝑐 , 𝐴𝑂𝐴 and airfoil shape)..
𝑅𝑒𝑐 (×105 ) 𝐴𝑂𝐴 (◦ ) Airfoil Flow structureb Recommended model
0.23 5.5 NACA 0012 Separation without reattachment 𝑘 − 𝑘𝑙 − 𝜔a , LowRea
0.5 0 NACA 4415 Separation without reattachment 𝛾
0.48/0.5 5.5/5 NACA 0012 LoLSB 𝑅𝑒𝜃 − 𝛾 a
0.75 0 NACA 4415 TE LoLSB 𝑅𝑒𝜃 − 𝛾 a , 𝑘 − 𝑘𝑙 − 𝜔a
4 TE LoLSB 𝑅𝑒𝜃 − 𝛾
8 MC LoLSB 𝑅𝑒𝜃 − 𝛾 a
12 LE LoLSB 𝑅𝑒𝜃 − 𝛾 a
1 0 NACA 4415 TE LoLSB 𝑅𝑒𝜃 − 𝛾 a
4 TE LoLSB 𝑅𝑒𝜃 − 𝛾
8 MC LoLSB 𝑅𝑒𝜃 − 𝛾
12 LE LoLSB 𝑅𝑒𝜃 − 𝛾 a , 𝑘 − 𝑘𝑙 − 𝜔a
2 0 NACA 4415 TE LoLSB 𝑅𝑒𝜃 − 𝛾, 𝛾, 𝑘 − 𝑘𝑙 − 𝜔
4 MC LoLSB 𝑅𝑒𝜃 − 𝛾, 𝛾, 𝑘 − 𝑘𝑙 − 𝜔
8 MC LoLSB 𝑅𝑒𝜃 − 𝛾, 𝛾, 𝑘 − 𝑘𝑙 − 𝜔
12 LE LSB, TS 𝑅𝑒𝜃 − 𝛾 a , 𝛾 a
a
The model predicts the flow quantitatively with some inconsistency about the time-averaged 𝐶𝑝 value or separation/reattachment point.
b TE,MC and LE mean bubbles form near the trailing-edge, middle-chord and leading-edge, respectively. TS means the turbulence separation.
LSB and LoLSB mean laminar separation bubbles and long laminar separation bubbles, respectively.

predict larger separation near the trailing edge. The 𝑘 − 𝑘𝑙 − 𝜔 model 3.3. Prediction capability for different airfoil shapes
is very sensitive to the 𝑅𝑒𝑐 as different boundary layer characteristics
are predicted at the two 𝑅𝑒𝑐 . This may be because these 𝑅𝑒𝑐 values are As discussed in Section 3.2, the prediction capability of the four
transition models varies with the 𝑅𝑒𝑐 number and 𝐴𝑂𝐴. More accurate
around the critical Reynolds number where the flow would reattach to
results are predicted at larger 𝑅𝑒𝑐 and smaller 𝐴𝑂𝐴 due to reduced
the airfoil surface at the trailing edge for the NACA 4415 airfoil. At the
flow separation on the suction surface of an airfoil. Hence, the model
𝑅𝑒𝑐 value 2 × 105 (Fig. 8(c)), the separation bubble forms on the middle prediction capability is inherently related to the boundary layer char-
suction surface, and the bubble length is reduced. The models except acteristic. If a flow tends to separate and become complex, it is difficult
the LowRe model predict similar flow structures. for a model to predict the flow accurately. The geometrical features of

7
Y. Liu et al. Journal of Wind Engineering & Industrial Aerodynamics 217 (2021) 104726

middle suction surface with no further separation. The flow structures


and the 𝐶𝑝 distribution on the airfoil surface (Fig. 6) predicted by the
four transition models for NACA 0012 are more consistent than NACA
4415.
As the airfoil thickness and camber increase (e.g., from NACA 0012
to NACA 4415), the separation tends to occur, and the laminar separa-
tion point tends to remain around the trailing edge. Better prediction
can thus be obtained by the transition models in a thinner and less
cambered airfoil (e.g., NACA 0012) under low Reynolds numbers.

3.4. Prediction capability of separation, transition and reattachment loca-


tions

The laminar separation, transition and reattachment locations were


obtained from the distribution of predicted time-averaged wall shear
stress in the flow direction (𝜏𝑤𝑎𝑙𝑙−𝑥 ) along the airfoil upper surface. The
separation and reattachment points occur where the wall shear stress
changes its direction (i.e., positive or negative). The transition point is
located at the minimum 𝜏𝑤𝑎𝑙𝑙−𝑥 with a sudden rise of the wall shear
stress after laminar separation. The deviation coefficients are defined
to measure the difference between the predicted and experimental
location results as follows:
𝐿𝑠𝑝 − 𝐿𝑠𝑝𝑒𝑥𝑝
𝜖𝑙𝑚 = (2)
𝐿𝑠𝑝𝑒𝑥𝑝
𝑃 𝑡𝑟 − 𝑃 𝑡𝑟𝑒𝑥𝑝
𝜖𝑡𝑟 = (3)
𝑃 𝑡𝑟𝑒𝑥𝑝
𝑃 𝑟𝑎 − 𝑃 𝑟𝑎𝑒𝑥𝑝
𝜖𝑟𝑎 = (4)
𝑃 𝑟𝑎𝑒𝑥𝑝
where 𝐿𝑠𝑝, 𝑃 𝑡𝑟 and 𝑃 𝑟𝑎 are the laminar separation point, transition
point and reattachment point predicted by the simulation, respectively.
𝐿𝑠𝑝𝑒𝑥𝑝 , 𝑃 𝑡𝑟𝑒𝑥𝑝 , 𝑃 𝑟𝑎𝑒𝑥𝑝 are the respective experimental data. The pre-
dicted laminar separation point (Fig. 10), transition point (Fig. 11) and
reattachment point (Fig. 12) are compared to the available experimen-
tal data in a range of 𝐴𝑂𝐴𝑠 and 𝑅𝑒𝑐 . Failure to predict the location is
not shown in these figures.
The predicted laminar separation point (Fig. 10) well agrees with
the experimental data at small 𝐴𝑂𝐴𝑠 especially with higher 𝑅𝑒𝑐 =
2×105 . But at large 𝐴𝑂𝐴𝑠, all models tend to predict a forward laminar
separation location (negative deviation coefficient 𝜖𝐿𝐴𝑀 ). It is hard to
tell which model can better predict the laminar separation point in the
𝐴𝑂𝐴𝑠 and 𝑅𝑒𝑐 range.
There is a lack of experimental data of transition point at 𝑅𝑒𝑐 =
2 × 105 (Fig. 11). The 𝑅𝑒𝜃 − 𝛾 model can better predict transition points
in the 𝐴𝑂𝐴𝑠 and 𝑅𝑒𝑐 . But the LowRe and 𝛾 models fail to simulate the
sudden rise of the wall shear stress and could be incapable of reliably
predicting transition in this Reynolds number range.
As the 𝑅𝑒𝑐 number reaches a threshold (𝑅𝑒𝑐 = 2×105 ), all the models
can predict the reattachment point (Fig. 12). In particular, the 𝑅𝑒𝜃 − 𝛾
model seems to have the best prediction capability under all conditions.
Fig. 7. Time-averaged 𝐶𝑝 distribution on the NACA 4415 airfoil surface with four
models at different 𝐴𝑂𝐴𝑠, (a) 𝑅𝑒𝑐 = 7.5 × 104 , (b) 𝑅𝑒𝑐 = 1 × 105 and (c) 𝑅𝑒𝑐 = 2 × 105 .
3.5. Selection of transitional turbulence models at low Reynolds numbers

The simulation results were analyzed to give some guidance on


an airfoil such as thickness and camber affect the separation behavior selecting transitional turbulence models in the low 𝑅𝑒𝑐 range (1 × 104 −
of the boundary layer and consequently the prediction capability by 5 × 105 ) for the small wind turbine application. Table 5 summarizes
various models. the predicted flow characteristic under different conditions of chord
Fig. 9 shows the instantaneous vorticity and time-averaged stream- Reynolds numbers (𝑅𝑒𝑐 ), angles of attack (𝐴𝑂𝐴𝑠) and airfoil shapes.
lines of the NACA 4415 and NACA 0012 airfoils at 𝑅𝑒𝑐 ≈ 5 × 104 . In The optimum models for each condition are specified based on their
NACA 4415, the large-scale separation occurs even at small 𝐴𝑂𝐴 = 0◦ capability of predicting both the time-averaged 𝐶𝑝 value and separa-
around the trailing edge and significantly influences the wake flow. tion/reattachment locations. The formation of long laminar separation
Moreover, no reattachment occurs at this Reynolds number. But in bubbles (LoLSB, length> 20%𝑐) causes prediction uncertainty in this
NACA 0012, the separation is small. As the angle of attack increases, the Reynolds number range. The prediction capability of various transition
laminar separation occurs near the leading edge and reattaches on the models is enhanced as the 𝑅𝑒𝑐 increases or the 𝐴𝑂𝐴 decreases.

8
Y. Liu et al. Journal of Wind Engineering & Industrial Aerodynamics 217 (2021) 104726

Fig. 8. Distribution of instantaneous vorticity and time-averaged streamlines over the NACA 4415 airfoil predicted by four models at 𝐴𝑂𝐴 = 0◦ , (a) 𝑅𝑒𝑐 = 7.5 × 104 , (b) 𝑅𝑒𝑐 = 1 × 105
and (c) 𝑅𝑒𝑐 = 2 × 105 .

Fig. 13 recommends the transitional turbulence models in the dif- at a low 𝐴𝑂𝐴 < 8◦ , the 𝑅𝑒𝜃 − 𝛾 model is capable to give a reliable
ferent ranges of 𝑅𝑒𝑐 and 𝐴𝑂𝐴𝑠. For 𝑅𝑒𝑐 < 0.5 × 105 or 0.5 × 105 < prediction. At the 𝑅𝑒𝑐 ≥ 2 × 105 , the 𝑅𝑒𝜃 − 𝛾, 𝛾, 𝑘 − 𝑘𝑙 − 𝜔 models are all
𝑅𝑒𝑐 < 2 × 105 at a large 𝐴𝑂𝐴 > 8◦ , the four transitional models cannot
predict the flow accurately, and so the RANS method is not appropriate. reliable at 𝐴𝑂𝐴 < 12◦ without turbulence separation. If the turbulence
Instead, LES or DNS can be used. For 0.5 × 105 < 𝑅𝑒𝑐 < 2 × 105 boundary separates from the trailing edge at higher 𝐴𝑂𝐴, the 𝑘 − 𝑘𝑙 − 𝜔

9
Y. Liu et al. Journal of Wind Engineering & Industrial Aerodynamics 217 (2021) 104726

Fig. 9. Distribution of instantaneous vorticity and time-averaged streamlines predicted by four models over (a) the NACA 4415 airfoil at 𝑅𝑒𝑐 = 5 × 104 and 𝐴𝑂𝐴 = 0◦ , (b) the
NACA 0012 airfoil at 𝑅𝑒𝑐 = 4.8 × 104 , 𝐴𝑂𝐴 = 0◦ , and (c) the NACA 0012 airfoil at 𝑅𝑒𝑐 = 4.8 × 104 , 𝐴𝑂𝐴 = 5.5◦ .

model overestimates the minimum negative time-averaged 𝐶𝑝 value, wind turbines usually work. The NACA 0012 and NACA 4415 airfoils
and therefore the 𝑅𝑒𝜃 −𝛾, 𝛾 models are recommended for this situation. were chosen for the simulations under the 𝑅𝑒𝑐 = (0.5 − 2) × 105 and
𝐴𝑂𝐴 = (0 − 12)◦ conditions. The conclusions are drawn as follows.
4. Conclusions In the low Reynolds number range, the simulation results by the
RANS method are more sensitive to the inlet turbulent intensity than
This paper evaluated the prediction capability of four frequently the viscosity ratio set at the boundaries. The inlet turbulent intensity
used transitional turbulence models (LowRe, 𝑅𝑒𝜃 − 𝛾, 𝛾, 𝑘 − 𝑘𝑙 − 𝜔) in should be carefully set to meet the simulation situation. The eddy
the RANS framework in the low Reynolds number range in which small viscosity ratio between 5–10 is recommended.

10
Y. Liu et al. Journal of Wind Engineering & Industrial Aerodynamics 217 (2021) 104726

Fig. 10. Deviation of predicted laminar separation point from experimental data on the NACA 4415 airfoil surface with four models at different 𝐴𝑂𝐴𝑠 (a) 𝑅𝑒𝑐 = 7.5 × 104 , (b)
𝑅𝑒𝑐 = 1 × 105 and (c) 𝑅𝑒𝑐 = 2 × 105 .

Fig. 11. Deviation of predicted transition point from experimental data on the NACA 4415 airfoil surface with four models at different 𝐴𝑂𝐴𝑠 (a) 𝑅𝑒𝑐 = 7.5×104 and (b) 𝑅𝑒𝑐 = 1×105 .

As the 𝑅𝑒𝑐 number increases, the prediction capability of all models satisfaction in a certain 𝑅𝑒𝑐 range and at low 𝐴𝑂𝐴𝑠 without turbulence
improves due to the reduced size of laminar separation bubbles. At separation. Although the 𝛾 model is derived directly from the 𝑅𝑒𝜃 − 𝛾
lower 𝑅𝑒𝑐 , the large-scale separation makes it difficult for a model model, it cannot reliably predict the flow field at the relatively large
to predict the flow accurately. Better prediction can be obtained in a 𝐴𝑂𝐴 and relatively low 𝑅𝑒𝑐 . The prediction capability of the LowRe
thinner and less cambered airfoil at higher 𝑅𝑒𝑐 and small 𝐴𝑂𝐴. model is questionable among all the models.
The 𝑅𝑒𝜃 − 𝛾 model has the most promising prediction capability The 𝑅𝑒𝜃 − 𝛾 and 𝛾 models may be further modified with empirical
among all the models as it considers different kinds of transition paths correlations to improve prediction capability in different applications.
in the transition model formulation. The 𝑘 − 𝑘𝑙 − 𝜔 model based on The physics-based 𝑘−𝑘𝑙 −𝜔 model can be improved to meet the general
the pre-transition laminar fluctuation phenomenon gives a secondary use of transition predictions.

11
Y. Liu et al. Journal of Wind Engineering & Industrial Aerodynamics 217 (2021) 104726

Fig. 12. Deviation of predicted reattachment point from experimental data on the NACA 4415 airfoil surface with four models at different 𝐴𝑂𝐴𝑠 (a) 𝑅𝑒𝑐 = 7.5×104 , (b) 𝑅𝑒𝑐 = 1×105
and (c) 𝑅𝑒𝑐 = 2 × 105 .

Declaration of competing interest

The authors declare that they have no known competing finan-


cial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Acknowledgment

The authors acknowledge the financial support of the National


Natural Science Foundation of China (51475174).

References

Baidya Roy, S., 2011. Simulating impacts of wind farms on local hydrometeorology.
J. Wind. Eng. Ind. Aerodyn. 99 (4), 491–498. http://dx.doi.org/10.1016/j.jweia.
2010.12.013.
Breuer, M., 2018. Effect of inflow turbulence on LES of an airfoil flow with laminar
separation bubble. Flow Turbul. Combust. 101, 433–456. http://dx.doi.org/10.
1007/978-3-030-04915-7_46.
Fig. 13. Recommendation for the selection of transitional turbulence models in a range Choudhry, A., Arjomandi, M., Kelso, R., 2015. A study of long separation bubble
of low Reynolds number and angle of attack. on thick airfoils and its consequent effects. Int. J. Heat Fluid Flow 52, 84–96.
http://dx.doi.org/10.1016/j.ijheatfluidflow.2014.12.001.
Devinant, P., Laverne, T., Hureau, J., 2002. Experimental study of wind-turbine airfoil
aerodynamics in high turbulence. J. Wind. Eng. Ind. Aerodyn. 90 (6), 689–707.
CRediT authorship contribution statement http://dx.doi.org/10.1016/S0167-6105(02)00162-9.
Di Pasquale, D., Rona, A., Garrett, S.J., 2009. A selective review of CFD transition
models. In: 39th AIAA Fluid Dynamics Conference. San Antonio, Texas.
Fürst, J., Příhoda, J., Straka, P., 2013. Numerical simulation of transitional flows.
Yue Liu: Conceptualization, Methodology, Software, Validation, Computing 95 (suppl. 1), S163–S182. http://dx.doi.org/10.1007/s00607-012-0266-
Formal analysis, Visualization, Investigation, Writing – original draft. 0.
Peifeng Li: Writing – review & editing, Supervision, Resources. Kaiy- Gad-el Hak, M., 1990. Control of low-speed airfoil aerodynamics. AIAA J. 28 (9),
1537–1552. http://dx.doi.org/10.2514/3.25250.
ong Jiang: Conceptualization, Supervision, Resources, Funding acqui- Ingen, J.L.V., 2008. The eN method for transition prediction. Historical review of work
sition. at TU Delft. In: 38th Fluid Dynamics Conference and Exhibit. Seattle, Washington.

12
Y. Liu et al. Journal of Wind Engineering & Industrial Aerodynamics 217 (2021) 104726

Jones, L.E., Sandberg, R.D., Sandham, N.D., 2008. Direct numerical simulations of Rumsey, C.L., Spalart, P.R., 2009. Turbulence model behavior in low Reynolds number
forced and unforced separation bubbles on an airfoil at incidence. J. Fluid Mech. regions of aerodynamic flowfields. AIAA J. 47 (4), 982–993. http://dx.doi.org/10.
602, 175–207. http://dx.doi.org/10.1017/S0022112008000864. 2514/1.39947.
Jones, G., Santer, M., Papadakis, G., 2018. Control of low Reynolds number flow around Saliveros, E., 1988. The Aerodynamic Performance of the NACA-4415 Aerofoil Section
an airfoil using periodic surface morphing: A numerical study. J. Fluids Struct. 76, At Low Reynolds Numbers (Ph.D. thesis). University of Glasgow, Glasgow, UK.
95–115. http://dx.doi.org/10.1016/j.jfluidstructs.2017.09.009. Sanei, M., Razaghi, R., 2018. Numerical investigation of three turbulence simulation
Khayatzadeh, P., Nadarajah, S., 2014. Laminar-turbulent flow simulation for wind models for S809 wind turbine airfoil. Proc. Inst. Mech. Eng. A 232 (8), 1037–1048.
turbine profiles using the 𝛾 − 𝑅𝑒̃ 𝜃𝑡 transition model. Wind Energy 17, 901–918. http://dx.doi.org/10.1177/0957650918767301.
http://dx.doi.org/10.1002/we. Spalart, P.R., Rumsey, C.L., 2007. Effective inflow conditions for turbulence models in
Kim, D.H., Yang, J.H., Chang, J.W., Chung, J., 2009. Boundary layer and near-wake aerodynamic calculations. AIAA J. 45 (10), 2544–2553. http://dx.doi.org/10.2514/
measurements of NACA 0012 airfoil at low Reynolds numbers. In: 47th AIAA 1.29373.
Aerospace Sciences Meeting. Orlando, Floridas, http://dx.doi.org/10.2514/6.2009- Tummala, A., Velamati, R.K., Sinha, D.K., Indraja, V., Krishna, V.H., 2016. A review
1472. on small scale wind turbines. Renew. Sustain. Energy Rev. 56, 1351–1371. http:
Langtry, R., Gola, J., Menter, F., 2006. Predicting 2D airfoil and 3D wind turbine //dx.doi.org/10.1016/j.rser.2015.12.027.
rotor performance using a transition model for general CFD codes. In: 44th AIAA Walters, D.K., Cokljat, D., 2008. A three-equation eddy-viscosity model for Reynolds-
Aerospace Sciences Meeting and Exhibit. Reno, Nevada, http://dx.doi.org/10.2514/ averaged Navier–Stokes simulations of transitional flow. J. Fluids Eng. 130, 121401.
6.2006-395. http://dx.doi.org/10.1115/1.2979230.
Langtry, R.B., Menter, F.R., 2009. Correlation-based transition modeling for un- Wang, S., Zhou, Y., Alam, M.M., Yang, H., 2014. Turbulent intensity and Reynolds
structured parallelized computational fluid dynamics codes. AIAA J. 47 (12), number effects on an airfoil at low Reynolds numbers. Phys. Fluids 26, 115107.
2894–2906. http://dx.doi.org/10.2514/1.42362. http://dx.doi.org/10.1063/1.4901969.
Lin, M., Sarlak, H., 2016. A comparative study on the flow over an airfoil using Wauters, J., Degroote, J., Vierendeels, J., 2019. Comparative study of transition models
transitional turbulence models. AIP Conf. Proc. 1738, http://dx.doi.org/10.1063/ for high-angle-of-attack behavior. AIAA J. 57 (6), 1–16. http://dx.doi.org/10.2514/
1.4951806. 1.J057249.
Liu, Y., Li, P., He, W., Jiang, K., 2020. Numerical study of the effect of surface grooves Wilcox, D.C., 1994. Simulation of transition with a two-equation turbulence model.
on the aerodynamic performance of a NACA 4415 airfoil for small wind turbines. AIAA J. 32 (2), 247–255. http://dx.doi.org/10.2514/3.59994.
J. Wind. Eng. Ind. Aerodyn. 206 (January), 104263. http://dx.doi.org/10.1016/j. Wolgemuth, N.M., Walters, D.K., 2007. CFD Prediction of boundary layer transition and
jweia.2020.104263. separation on an airfoil at varying angle of attack. In: Proceedings of FEDSM2007.
Lopes, R., Eça, L., Vaz, G., 2020. On the numerical behavior of RANS-based transition San Diego, California, http://dx.doi.org/10.1115/fedsm2007-37111.
models. J. Fluids Eng. 142 (5), 1–14. http://dx.doi.org/10.1115/1.4045576. Xia, C., Chen, W., 2016. Boundary-layer transition prediction using a simplified
Mayle, R.E., Schulz, A., 1996. The path to predicting bypass transition. In: International correlation-based model. Chin. J. Aeronaut. 29 (1), 66–75. http://dx.doi.org/10.
Gas Turbine and Aeroengine Congress & Exhibition. Birmingham, UK, http://dx. 1016/j.cja.2015.12.003.
doi.org/10.1115/96-GT-199. Zhang, Y., Sun, Z., van Zuijlen, A., van Bussel, G., 2017. Numerical simulation of
Menter, F.R., 1994. Two-equation eddy-viscosity turbulence models for engineering transitional flow on a wind turbine airfoil with RANS-based transition model. J.
applications. AIAA J. 32 (8), 1598–1605. http://dx.doi.org/10.2514/3.12149. Turbul. 18 (9), 879–898. http://dx.doi.org/10.1080/14685248.2017.1334908.
Menter, F.R., Smirnov, P.E., Liu, T., Avancha, R., 2015. A one-equation local correlation-
based transition model. Flow, Turbul. Combust. 95 (4), 583–619. http://dx.doi.org/
10.1007/s10494-015-9622-4.

13

You might also like