Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

JID:AESCTE AID:4926 /FLA [m5G; v1.250; Prn:9/01/2019; 16:43] P.

1 (1-10)
Aerospace Science and Technology ••• (••••) •••–•••

1 67
Contents lists available at ScienceDirect
2 68
3 69
4 Aerospace Science and Technology 70
5 71
6 72
7
www.elsevier.com/locate/aescte 73
8 74
9 75
10 76
11
12
Design methodology using characteristic parameters control for low 77
78
13 Reynolds number airfoils 79
14 80
15
Sen Zhang ∗ , Huaxing Li, Afaq Ahmed Abbasi 81
16 82
17 School of Aeronautics, Northwestern Polytechnical University, Xi’an Shaanxi 710072, China 83
18 84
19 85
20 a r t i c l e i n f o a b s t r a c t 86
21 87
22 Article history: In the past century, low Reynolds number airfoil design for both military and civil applications has 88
Received 25 April 2018 continued to capture the interest of practitioners of applied aerodynamics. A new design methodology
23 89
Received in revised form 31 December 2018 for airfoils has been developed, which controls the airfoil profile through 8 parameters. The airfoil
24 Accepted 2 January 2019 90
aerodynamic characteristics were predicted through the XFOIL code, to optimize airfoil under mode-
25 Available online xxxx 91
FRONTIER environment. The reliability verification result predicted that the XFOIL code can be used
26 92
Keywords: to determine the airfoil performance in the linear segment of the lift coefficient curve. Application of
27 93
Low Reynolds number this methodology was demonstrated through two cases. For Case-I, airfoils design were based on E387
28
Airfoil design airfoil profile and synthetic aerodynamic performance at different angles of attack for Re = 2.0 × 105 94
29 XFOIL were considered. While for Case-II, designed airfoils were based on PSU 94-097 airfoil with synthetic 95
30 Optimization design aerodynamic performance at different angles of attack for Re = 4.0 × 105 . Results of both cases showed 96
31 enhanced aerodynamic properties for the designed airfoils. 97
32 © 2019 Elsevier Masson SAS. All rights reserved. 98
33 99
34 100
35 101
36 102
37 1. Introduction airfoils. For this reason, the improvement and use of the airfoil de- 103
38 sign method is still an effective solution [22,23]. 104
39 There are many kinds of design methods for airfoils, mainly 105
In the past century, the design of low Reynolds number air-
40 including geometric means guided by experimental research and 106
foils for both military and civil applications has continued to cap-
41 inverse methods. Because of the ability to make the control of 107
ture the interest of practitioners of applied aerodynamics [1–9].
42 aerodynamic performance reach a high level, the inverse method 108
For example, airfoils for U.S. Navy electronic warfare UAVs (e.g.,
43 has certain advantages, but still needs to meet many geometric 109
LAURA [10] and FLYRT [11,12] aircraft) operate in the Reynolds
44 constraints. Inverse design in the classical sense is specifying a de- 110
number range of 2.0 × 105 to 5.0 × 105 . With the development
45 sired velocity distribution, and then solving the development of the 111
of technology in the aeronautics, astronautics and material indus-
46 boundary layer to get the required aerodynamic performance. An 112
tries, together with the rise in oil prices triggered by the global improvement of the inverse design is to set the required bound-
47 113
oil crisis, the new technology of the airship has attracted wide ary layer characteristics directly, which makes it easier to control
48 114
attention [13–16]. The Reynolds number range of the propeller air- the desired aerodynamic performance. Thus, the designer can ef-
49 115
foils for the multi-body advanced airship falls in the 105 order of fectively balance the influence of various factors during the de-
50 116
magnitude [15]. In addition, the application of low Reynolds num- sign process to achieve the performance goals. The hybrid design
51 117
ber airfoils is varied, including remotely piloted vehicles, ultralight method, which combines the geometric mean and the inverse de-
52 118
glider, wind turbine, propeller and so on [17–21]. For specific ap- sign method, has also been widely used, and the beneficial results
53 119
plications, the design requirements of airfoils are different. If airfoil have been obtained. Nevertheless, these four design methods have 120
54
55
design requirements for unique applications are regarded as a set, their own advantages, which can control some characteristics of 121
56
the number of elements will continue to grow [22]. When one airfoils more or less, such as airfoil profile distribution, airfoil sur- 122
57
considers the arrangement of the numerous possible airfoil design face velocity distribution, airfoil boundary layer characteristics and 123
58 requirements, it will soon be found that the required size of air- airfoil aerodynamic performance. Each of the above features is re- 124
59 foil design requirements is much larger than that of the existing garded as a design variable and is ideally incorporated into a single 125
60 design method. 126
61 Generally, the airfoil design requirement is achieving low drag 127
62 * Corresponding author. in the operating range of lift coefficient. Therefore, high lift is 128
63 E-mail address: sen96@mail.nwpu.edu.cn (S. Zhang). rarely the only desirable feature of an airfoil. The lift-to-drag ra- 129
64 130
https://doi.org/10.1016/j.ast.2019.01.003
65 131
1270-9638/© 2019 Elsevier Masson SAS. All rights reserved.
66 132
JID:AESCTE AID:4926 /FLA [m5G; v1.250; Prn:9/01/2019; 16:43] P.2 (1-10)
2 S. Zhang et al. / Aerospace Science and Technology ••• (••••) •••–•••

1 67
2 Nomenclature 68
3 69
4 Cp Pressure coefficient kf2 Trailing shape factor of mean camber line, k f 2 ∈ 70
5 x Airfoil abscissa (−1, 0] 71
6 xt Relative position of airfoil thickness y f |(x f , f ) First-order derivative of y f at the vertex 72
7 xf Relative position of airfoil camber y f Second-order derivative of y f 73
8
y Airfoil ordinate x0 Abscissa 74
9 kt1 Leading shape factor of thickness distribution, kt1 ∈ 75
yf Function of mean camber line
10 (0, ∞) 76
11 yt Function of a symmetrical thickness distribution kt2 Trailing shape factor of thickness distribution, kt2 ∈ 77
12 kf1 Leading shape factor of mean camber line, k f 1 ∈ (0, ∞) 78
13 (−1, 0] Tu Absolute turbulence intensity 79
14 80
15 81
16 82
17 83
18 84
19 85
20 86
21 Fig. 1. Airfoil nomenclature. 87
22 88
23 tio, thickness, pitching moment, stall performance and roughness 89
24 sensitivity are all important factors, which must each be weighed 90
25 separately when one considers selecting or designing an airfoil. 91
26 The objective of this work focuses on a new design philosophy Fig. 2. The mean camber line consisting of two elliptic arcs. 92
27 for low Reynolds number airfoils, which consists of the establish- 93
28 ment of the airfoil profile equation controlled by characteristic 94
tex of the two ellipses. Then the function of the mean camber line
29 parameters and the combination with optimization method. The 95
is given by Eq. (2):
30 airfoil aerodynamic characteristics prediction program used for this 96
31
 (x − x f )2 (1 − k2f 1 )2  2 0.5  97
study is the XFOIL code, which makes use of a panel-method-based f kf1 f
32 potential-flow solution that is coupled with an integral boundary yf = 1− + 98
x2f kf1 +1 kf1 +1
33 layer and is widely used in academia and industry. The following 99


34 sections will provide an overview of the design methodology, op- × 1 − trunc x + (1 − x f ) 100
35 timization strategy and two design cases.  (x − x f )2 (1 − k2f 2 )2  2 0.5 101
36 f 102
+ 1−
37 2. Design methodology (1 − x f )2 kf2 + 1 103
38  104
kf2 f

39
Consider the airfoil sketched in Fig. 1. The mean camber line + trunc x + (1 − x f ) (2) 105
40 kf2 +1 106
is the locus of points halfway between the upper and lower sur-
41 107
faces as measured perpendicular to the mean camber line itself. where x f is the relative position of airfoil camber; k f 1 donates the
42 108
Just like all standard NACA airfoils, the shapes of airfoils can be leading shape factor of mean camber line, k f 1 ∈ (−1, 0]; k f 2 repre-
43 109
generated by specifying the shape of mean camber line and then sents the trailing shape factor of mean camber line, k f 2 ∈ (−1, 0].
44 110
wrapping a specified symmetrical thickness distribution around The trunc( ) function that returns a number after truncated to
45 111
the mean camber line. Therefore, the airfoil profile equation can certain decimal places are defined as presented in Eq. (3):
46 112
be presented as:


⎨ 0 0 ≤ x < xf
47 113

48
1 trunc x + (1 − x f ) = 1 x = x f (3) 114
y = yf ± yt (1) ⎩
49
2 1 xf < x ≤ 1 115
50 116
51 where y f represents the function of mean camber line; yt denotes Because the first-order derivative of Eq. (2) is zero at the vertex, 117
52 the function of a symmetrical thickness distribution. that is y f |(x f , f ) = 0, the mean camber line in Fig. 2 is smooth and 118
53 Eq. (1) shows that the airfoil profile equation can be expressed continuous. And, on the interval [0, 1], the second-order derivative 119
54 as the superposition of y f and yt . Therefore, if the functions of y f of Eq. (2) is constant less than zero, that is y f < 0, which means 120
55 and yt are established, using Eq. (1) to design airfoil profile will be that the mean camber line expressed by Eq. (2) has no inflection 121
56 possible. point. 122
57 123
58 2.1. Function of mean camber line 2.1.2. The influence of k f 1 and k f 2 on the shape of mean camber line 124
59 When f and x f are established, the shape of mean camber line 125
60 2.1.1. Function establishment can be adjusted locally through k f 1 and k f 2 . 126
61 Fig. 2 shows the construction method. The mean camber line Figs. 3(a) and (b) show the mean camber lines for f = 3.43% 127
62 is constructed by the combination of two elliptic arcs, which is and x f = 40%. All the mean camber lines are divided into front 128
63 cut out from two ellipses with the same vertex, respectively. To part and rear part by coordinate (x f , f ). k f 1 only has an impact 129
64 facilitate the description, the airfoil chord length is set to 1. on the front part and has nothing to do with the rear part. On 130
65 The coordinates (0, 0) and (1, 0) represent the leading edge and the other hand, k f 2 only has an influence on the rear part. They 131
66 trailing edge points, respectively. And coordinate (x f , f ) is the ver- have similar effects on the shape of mean camber line. Except for 132
JID:AESCTE AID:4926 /FLA [m5G; v1.250; Prn:9/01/2019; 16:43] P.3 (1-10)
S. Zhang et al. / Aerospace Science and Technology ••• (••••) •••–••• 3


8 3
1 Solving Eq. (8) for m, and substituting the result into n = 9
m, 67
2 gives 68
3 69
4 (kt + d)kt +d 70
5 n=t (9) 71
k
kt t dd 72
6
7 yt = nxkt (1 − x)d (10) 73
8 74
9 In order to control the front and rear parts of the airfoil thickness 75
10 distribution alone, parameters are divided into two cases. Conse- 76
11 quently, Eq. (10) is expressed as: 77
12 78


13 yt = n1 xkt1 (1 − x)d1 1 − trunc x + (1 − xt ) 79
14
80
15 + n2 xkt2 (1 − x)d2 trunc x + (1 − xt ) (11) 81
16 82
17
where 83
18 kt1 +d1 84
(kt1 + d1 )
19 n1 = t 85
k d
20 kt1t1 + d11 86
21 kt2 +d2 87
(kt2 + d2 )
22 Fig. 3. Comparison of mean camber lines for f = 3.43%, x f = 40%. (a) k f 2 = −0.99. n2 = t 88
k d
23 (b) k f 1 = −0.99. kt2t2 + d22 89
24 90
kt1 (1 − xt )
25
coordinate (x f , f ), the ordinates of the front part and the rear part d1 = 91
26 xt 92
get bigger with the increase of k f 1 and k f 2 , respectively.
27 kt2 (1 − xt ) 93
28 d2 = 94
2.2. Function of a symmetrical thickness distribution xt
29 ⎧ 95
30
2.2.1. Function establishment
⎨ 0 0 ≤ x < xt 96
31 trunc x + (1 − xt ) = 1 x = xt 97
The expression for the Joukowsky airfoil profile, as suggested in ⎩
32 1 xt < x ≤ 1 98
the literature [24], is given by Eq. (4):
33 99
 √ and xt is the relative position of airfoil thickness; kt1 donates
34  100
1 1 1 2 3 the leading shape factor of thickness distribution, kt1 ∈ (0, ∞);
35 y= + 2
− x20 − ± t (1 − 2x0 ) 1 − 4x20 (4) 101
4 64 f 8f 9 kt2 represents the trailing shape factor of thickness distribution,
   
36 102
37
  kt2 ∈ (0, ∞). 103
2
1
38 The characteristics of the thickness distribution expressed by 104
39 where x0 denotes the abscissa; t is the airfoil thickness. In Eq. (4), Eq. (11) are similar to that of the mean camber line expressed 105
40 part-1 and part-2 represent the expressions of mean camber line by Eq. (2). The curve of the thickness distribution is smooth and 106
41 and a specified symmetrical thickness distribution, respectively. In continuous, and has no inflection point. 107
42 this study, only the thickness distribution is extracted as the re- 108
43 search object. 2.2.2. The influence of kt1 and kt2 on the thickness distribution 109
44 In order to move the leading edge to coordinate origin, the fol- When t and xt are established, the shape of thickness distribu- 110
45 lowing transformation is made: tion can be adjusted locally through kt1 and kt2 . 111
46 Figs. 4(a) and (b) show thickness distributions for t = 10.0% and 112
47 1 + 2x0 = 2x xt = 40%. The same as the characteristics of mean camber line, the 113
48 1 whole thickness distributions are also divided into front part and 114
49 x0 = x − (5) rear part by coordinate (xt , t). kt1 only has an impact on the front 115
2
50 part and kt2 only has an influence on the rear part. They have a 116
51 Substituting Eq. (5) into part-2 of Eq. (4), gives similar effect on the thickness distribution. Except for coordinate 117
52 √ (x f , f ), the ordinates of the front part and the rear part get bigger 118
8 3 0 .5 1 .5
53 yt = tx (1 − x) (6) with the increase of kt1 and kt2 , respectively. 119
54 9 Eventually, the design methodology for low Reynolds number 120
55 The form of Eq. (6) is changed to: airfoils proposed in this paper is established. In this method, the 121
56
√ airfoil profile equation composed of the function of mean cam- 122
57 8 3 kt d ber line and the function of a symmetrical thickness distribution 123
58 yt = mx (1 − x) (7) is presented in Eq. (1), which is controlled by 8 parameters. Those 124
9
59 parameters have clear physical meanings, four of which are air- 125
By solving yt = 0, the expressions of the airfoil thickness and the
60 foil thickness, airfoil camber, relative position of airfoil thickness 126
61 relative position of airfoil thickness are obtained as follows: and relative position of airfoil camber. Using the design method- 127
62 √ k ology, a designer can obtain a desired airfoil by precise control 128
8 3 kt t dd
63
t = y t | y  =0 = m (8) of characteristic parameters. However, it is obvious that it is dif- 129
64 t 9 (kt + d)kt +d ficult to complete the design by manually selecting the proper 130
65 kt characteristic parameters, so the optimization algorithm needs to 131
66 xt = be introduced. 132
kt + d
JID:AESCTE AID:4926 /FLA [m5G; v1.250; Prn:9/01/2019; 16:43] P.4 (1-10)
4 S. Zhang et al. / Aerospace Science and Technology ••• (••••) •••–•••

1 displacement thickness is guaranteed by the iteration between the 67


2 inner and outer flow solutions, the XFOIL code can provide an 68
3 accurate prediction for viscous pressure distributions, which cap- 69
4 70
ture the influence of laminar separation bubbles and limited trail-
5 71
ing edge separation [26,27]. The XFOIL code uses the e N based
6 72
approximate envelope method to predict laminar-turbulent transi-
7 73
tion, which is described by Drela and Giles [28]. Instead of tracking
8 74
the amplification rates of all frequencies (or a range of frequen-
9 75
10
cies), this method tracks only the most amplified frequency at a 76
11
given point on the airfoil downstream from the point of instabil- 77
12 ity to obtain the amplitude of the entire frequency envelope. The 78
13 magnification rates used in the XFOIL code are the linear approx- 79
14 imation of the actual magnification envelopes as predicted from 80
15 Falkner–Skan boundary layer profiles. As these simplifications are 81
16 based on the constant-shape-factor boundary layer development, 82
17 rapid changes in the shape factor in any part of the transition anal- 83
18 ysis may introduce errors relative to a full e N approach. The proper 84
19 N for XFOIL calculations can be calculated by Eq. (12) as described 85
20 by Van Ingen [29]. 86
21 87
22 N = −8.43 − 2.4 ln( T u ) (12) 88
23 89
24 where T u is the absolute turbulence intensity. In present research 90
25 Fig. 4. Comparison of thickness distributions for t = 10.0%, xt = 40%. (a) kt2 = 0.5. work, N is set to a default value of 9, which is appropriate for a 91
26 (b) kt1 = 0.5. smooth wing surface in a low turbulence intensity free stream. 92
27 93
28 3. Design optimization 3.1.2. Reliability verification for XFOIL code 94
29 3.1.2.1. E387 airfoil In the mid-1960s, Richard Eppler designed the 95
30 In order to achieve the optimum design, a method that can E387 airfoil specifically for use on model sailplanes [30]. The E387 96
31 quickly obtain the airfoil performance data must be introduced. airfoil was designed for proper aerodynamic performance and op- 97
32 There are many theoretical methods for predicting airfoil aerody- erating conditions by inverse conformal transformation. Much re- 98
33 namic characteristics, such as Drela and Youngren’s XFOIL code, search aimed at ameliorating aerodynamics of low Reynolds num- 99
34 Eppler’s PROFIL code, and some computational fluids dynamics ber airfoils has taken the E387 airfoil as a benchmark. The exper- 100
35 (CFD) solvers. In this research, Drela and Youngren’s XFOIL code iment data of the airfoil obtained by the NASA Langley Research 101
36 was used to predict airfoil aerodynamic characteristics. Center low turbulence wind tunnel (LTPT) in 1980s is still con- 102
37 sidered to be the standard for comparison of other low Reynolds 103
38 3.1. XFOIL code number airfoil experiments. 104
39 In Figs. 5(a) and (b), the aerodynamic characteristics predicted 105
40 3.1.1. Brief introduction for Re = 2.0 × 105 by the XFOIL code are compared to the NASA 106
41 The XFOIL code [25] is presented for viscous/inviscid analy- Langley Research Center LTPT measurements [31]. 107
42 sis and mixed-inverse design of subcritical airfoils. The code was The results presented in Fig. 5(a) show that the XFOIL code 108
43 widely used for fast aerodynamic performance prediction of low provides a reasonable prediction for the linear segment of lift coef- 109
44 Reynolds number airfoils. As its convergence on the boundary layer ficient curve, although with a little overprediction of the maximum 110
45 111
46 112
47 113
48 114
49 115
50 116
51 117
52 118
53 119
54 120
55 121
56 122
57 123
58 124
59 125
60 126
61 127
62 128
63 129
64 130
65 Fig. 5. Comparison of aerodynamic characteristics predicted by the XFOIL code and measured at the NASA Langley Research Center LTPT [31] for the E387 airfoil, Re = 131
66 2.0 × 105 . (a) C l vs α . (b) C l vs C d . 132
JID:AESCTE AID:4926 /FLA [m5G; v1.250; Prn:9/01/2019; 16:43] P.5 (1-10)
S. Zhang et al. / Aerospace Science and Technology ••• (••••) •••–••• 5

1 67
2 68
3 69
4 70
5 71
6 72
7 73
8 74
9 75
10 76
11 77
12 78
13 79
14 80
15 Fig. 6. Pressure coefficient distributions of the E387 airfoil measured at the NASA Langley Research Center LTPT [31] compared with the XFOIL prediction result, Re = 2.0 × 105 . 81
16 (a) α = 0◦ . (b) α = 3◦ . 82
17 83
18 84
19 85
20 86
21 87
22 88
23 89
24 90
25 91
26 92
27 93
28 94
29 95
30 96
31 97
32 98
33 99
34 100
35 101
36 102
37 103
Fig. 7. Comparison of aerodynamic characteristics predicted by the XFOIL code and measured at the Pennsylvania State University Low-Speed, Low-Turbulence Wind Tunnel
38 for the PSU 94-097 airfoil [32], Re = 4.0 × 105 . (a) C l vs α . (b) C l vs C d . 104
39 105
40 106
lift coefficient. In Fig. 5(b), the prediction also showed good agree- 3.1.2.2. PSU 94-097 airfoil The PSU 94-097 airfoil was designed to
41 107
ment with the experiment through the entire low drag region and be used on winglets of high-performance sailplanes at the Penn-
42 108
captured the sharpness both of the upper and lower corners. Fur- sylvania State University (PSU) employing the Eppler Airfoil Design
43 109
thermore, the predicted angles of attack for the linear segment of and Analysis Code (PROFIL98) [32]. The airfoil was designed specif-
44 110
lift coefficient curve range from −4 to 9 degrees. The maximum ically for the purpose of making a winglet operate as a wing does.
111
45
difference in the linear region (−4 to 9 degrees) of the lift coef- The PSU 94-097 airfoil has a high maximum lift coefficient and
46 112
ficient curve is not more than C l = 0.05. After 9 degrees, as the low drag performance within the range of operation. The operating
47 113
angle is further increased, the maximum difference is more than Reynolds number range of the PSU 94-097 airfoil is from 7.2 × 104
48 114
C l = 0.15. Therefore, if the angle of attack exceeds 9 degrees, (wing tip at low speed) to 1.0 × 106 (wing root at high speed).
49 115
In Figs. 7(a) and (b), the aerodynamic characteristics predicted
50 the XFOIL code will not accurately predict the aerodynamic perfor- 116
for Re = 4.0 × 105 by the XFOIL code are compared to the measure-
51 mance of the E387 airfoil. 117
ments, which were obtained in the Pennsylvania State University
52 The pressure coefficient distributions for angles of attack of 0 118
53 Low-Speed, Low-Turbulence Wind Tunnel. 119
and 3 degrees are shown in Figs. 6(a) and (b), respectively. The re-
54 The results presented in Fig. 7(a) show that the predicted re- 120
sults obtained from the above numerical method were compared
55 sults agree well with the experimental data in the linear region of 121
with the measurements provided by NASA Langley Research Center
56 lift coefficient curve, although the maximum lift coefficient is over 122
LTPT. At angles of attack of 0 and 3 degrees, the pressure coeffi-
57 predicted. In Fig. 7(b), as with the E387 airfoil, the XFOIL code also 123
cient predicted by the XFOIL code is in good agreement with the shows very good agreement with experiment through the entire
58 124
experimental results. The consistency of these pressure coefficient low drag region and well captures the sharpness both of the up-
59 125
60
distributions is precisely the reason why the predicted aerody- per and lower corners, although the drag at higher lift coefficients 126
61 namic performance of the linear section of lift coefficient curve is over predicted. Moreover, the predicted angles of attack for the 127
62 agrees well with those of the experiment. linear segment of lift coefficient curve range from −3 to 8 degrees. 128
63 These comparisons presented in Fig. 5 and Fig. 6 indicate that The maximum difference in the linear section (−3 to 8 degrees) of 129
64 the XFOIL code can well predict the aerodynamic performance of lift coefficient curve is not more than C l = 0.07. However, as the 130
65 the E387 airfoil in the linear section of lift coefficient curve for angle is further increased, the maximum difference increases up to 131
66 Re = 2.0 × 105 . C l = 0.16. Therefore, if the angle of attack is over 8 degrees, the 132
JID:AESCTE AID:4926 /FLA [m5G; v1.250; Prn:9/01/2019; 16:43] P.6 (1-10)
6 S. Zhang et al. / Aerospace Science and Technology ••• (••••) •••–•••

1 67
2 68
3 69
4 70
5 71
6 72
7 73
8 74
9 75
10 76
11 77
12 78
13 79
14 80
15 Fig. 8. Pressure coefficient distributions of the PSU 94-097 airfoil measured at the Pennsylvania State University Low-Speed, Low-Turbulence Wind Tunnel [32] compared with 81
16 the XFOIL prediction result, Re = 4.0 × 105 . (a) α = 2◦ . (b) α = 5◦ . 82
17 83
18 XFOIL code will not well predict the aerodynamic performance of 84
19 the PSU 94-097 airfoil. 85
20 The pressure coefficient distributions for angles of attack of 2 86
21 and 5 degrees are shown in Figs. 8(a) and (b), respectively. The 87
22 results obtained from the above numerical method were compared 88
23 with the measurements taken in the Pennsylvania State University 89
24 Low-Speed, Low-Turbulence Wind Tunnel. At angles of attack of 2 90
25 and 5 degrees, the pressure coefficient distributions predicted by 91
26 the XFOIL code are in excellent agreement with the experimental 92
27 results. The consistency of these pressure coefficient distributions 93
28 ensures the reliability of the prediction results of the aerodynamic 94
29 performance for the PSU 94-097 airfoil. 95
30 The comparisons of airfoil aerodynamic characteristics obtained 96
31 by experiment and prediction show that the XFOIL code can well 97
32 predict the aerodynamic performance of the PSU 94-097 airfoil in 98
33 the linear segment of lift coefficient curve for Re = 4.0 × 105 , al- 99
34 though the maximum lift coefficient tends to be over predicted. 100
35 101
36 3.2. Optimization procedure 102
37 103
38 The airfoil profile equation presented in Eq. (1) has eight un- 104
39 known variables. Therefore, an optimization procedure is needed 105
40 to ensure the best airfoil designed by Eq. (1) is possible. The 106
41 proposed methodology of optimization contains three main as- 107
42 pects [33–36]. First, the aerodynamic characteristic is calculated by 108
Fig. 9. Flow-diagram for optimization design.
43 XFOIL code. Second, an optimization algorithm is necessary to op- 109
44 timize the objective function. Last, a fully automatic optimization 110
45 platform that combines the optimization algorithm and flow field In present research work, an objective equation is proposed by 111
46 calculation is required. considering the aerodynamic performance at different angles of 112
47 ModeFRONTIER offers various optimization algorithms to se- attack. The aerodynamic characteristics at angles of attack of αi 113
48 lect from and has powerful post-processing capability [37]. In this (i = 1, 2, . . . , n) are selected as the design objective. n is the num- 114
49 paper, the optimization design was carried out in ModeFRON- ber of concerned angles of attack. The designer can choose the 115
50 TIER2016 environment with new airfoil design procedure devel- number of n according to the specific design requirements. The 116
51 oped in fortran2013 environment. The work flow for optimization objective function is defined as presented in Eq. (13): 117
52 design is shown in Fig. 9. 118
53 The flow diagram presented in Fig. 9 can be described in detail 
n
C l |α i 
n
C d |α i 119
max : f = wi − w n +i
54 in the following section. Firstly, an input file that contains the eight C l |Exp,αi C d |Exp,αi 120
55 parameters is created. Secondly, the airfoil data file that contains i =1 i =1 121
56 the coordinates of the designed airfoil is obtained by the exe- C l |αi > C l |Exp,αi i = 1, 2, . . . , n 122
57 cutable program. Thirdly, the aerodynamic performance at design 123
C d |αi > C d |Exp,αi (13)
58 angles of attack is predicted by the XFOIL code using the airfoil 124
data file. And lastly, designed algorithm determines the optimized n Exp (|Exp ) denotes expected; and w i represents
where subscript
59 125
results based on optimization algorithm. the weight, i =1 w i = 1 (i = 1, 2, . . . , 2n). Designers can take the
60 126
61 For different applications, the working environment of airfoils weight w i as the constant value. However, weight w i can also be 127
62 is also different. Some airfoils always work at a specific angle of determined according to the importance of the design objective. 128
63 attack, while others need to maintain good performance within The design problem is set as a multi-objective optimization 129
64 a range of angles of attack. Therefore, objective equations were problem, which helps to obtain better airfoils. In addition, based 130
65 proposed to fulfill the optimum design for airfoils with specific ap- to special applications, practitioners can add more airfoil features 131
66 plication. into Eq. (13), such as lift-to-drag ratio, thickness, pitching moment, 132
JID:AESCTE AID:4926 /FLA [m5G; v1.250; Prn:9/01/2019; 16:43] P.7 (1-10)
S. Zhang et al. / Aerospace Science and Technology ••• (••••) •••–••• 7

1 Table 1 Table 2 67
2 Constraints of design variables for Design-I and Design-II. Optimization results for Design-I and Design-II. 68
3 E387 Design-I Design-II E387 Design-I Design-II 69
4 70
t (%) 8.88 8.88 8.88 t (%) 8.88 8.88 8.88
5 xt (%) 28.80 [20, 50] [20, 50] xt (%) 28.80 23.26 25.80 71
6 kt1 [0.15, 0.6] [0.15, 0.6] kt1 / 0504 0.566 72
7 kt2 [0.15, 0.4] [0.15, 0.4] kt2 / 0.400 0.400 73
f (%) 3.83 3.83 3.83 f (%) 3.83 3.83 3.83
8 74
x f (%) 39.20 [15, 60] [15, 60] x f (%) 39.20 45.28 40.89
9 kf1 [−0.99, −0.2] [−0.99, −0.2] kf1 / −0.792 −0.377 75
10 kf2 [−0.99, −0.2] [−0.99, −0.2] kf2 / −0.471 −0.785 76
11 C l |α =−1◦ 0.2909 0.3977 0.3579 77
12 C l |α =0◦ 0.401 0.5045 0.4705 78
C l |α =5◦ 0.9501 1.0477 0.9872
13 etc., this will help them to obtain the airfoil with better perfor- 79
C d |α =−1◦ 0.0093 0.0107 0.0092
14 mance for their desired application. In present research work, the C d |α =0◦ 0.0098 0.0087 0.0089 80
15 Multi-Objective Genetic Algorithm II (MOGA-II) optimization algo- C d |α =5◦ 0.0126 0.0125 0.0125 81
16 rithm is employed for the optimization problem. 82
17 The MOGA-II is the proprietary version of the multi-objective 83
18 genetic algorithm. It uses an intelligent and efficient multi-search 84
19 elitism to maintain excellent (Pareto or non-dominated) solutions 85
20 without premature convergence to a local optimum. Elitism im- 86
21 proves the convergence of the algorithm and ensures that the 87
22 fitness of each new generation is greater than that of the par- 88
23 ent generation. MOGA-II handles constraints by applying penalty 89
24 strategies. Error designs are always dominated by valid designs, for 90
25 either feasible or unfeasible design. Unfeasible designs can be lo- 91
26 cated at the Pareto front only if there is no feasible design. When 92
27 a generation is complete, all of its designs, even error designs, will Fig. 10. Airfoil shapes used for Case-I and Case-II.
93
28 be considered for the computation of the new generation to ensure 94
29 the continuation of optimization [37]. 95
with E387 airfoil. For Design-I, lift coefficient is increased for an-
30 96
gles of attack of 0 and 5 degrees by 25.81% and 10.27%, respec-
31 4. Results and discussion 97
tively. And, the reduction of drag coefficient is 11.22% and 0.79%
32 98
for 0 and 5 degrees, respectively. Whereas for Design-II, the in-
33 To validate the design procedure proposed in section 3.2, two 99
crease of lift coefficient at angles of attack of −1, 0 and 5 degrees
34 100
cases were carried out by taking the E387 airfoil and the PSU is 23.03%, 17.33% and 3.9%, respectively. And, the reduction of drag
35 101
94-097 airfoil as references, respectively. In order to increase the coefficient is 1.08%, 9.18% and 0.79%, respectively. In addition, the
36 102
reliability of calculation by the XFOIL code, the design angles of designed airfoils also present better lift coefficient characteristics
37 103
attack were selected among the linear segments of lift coefficient than the E387 airfoil at all other predicted angles of attack.
38 104
curve. The designed airfoil is not a special application airfoil. The High lift is rarely the only desirable feature of an airfoil; how-
39 105
same Reynolds numbers were used for both cases i.e. Case-I and ever, the drag coefficient is an important factor. In Fig. 11(b), the
40 106
Case-II, as used previously for E387 and PSU 94-097 in the Sec- reduction of drag coefficient can be observed at angles of attack
41 107
tion 3.1.2 respectively. greater than 0 degrees. It can be observed that, for angles of at-
42 108
tack less than 0 degrees, Design-I shows a worse drag performance
43 Case-I: 109
than the E387 airfoil. This poor drag performance can be justi-
44 In this section, a new airfoil, designated as Design-I, was de- 110
fied; the reason for this reduction is that angles of attack less
45 signed considering the synthetically aerodynamic performance at 111
than 0 degrees are not taken into account for the optimum de-
46 angles of attack of 0 and 5 degrees for Re = 2.0 × 105 . The con- 112
sign. Furthermore, the drag coefficient behaves differently with the
47 straints of design variables are shown in Table 1 and all weights 113
lift coefficient in the linear segment of lift curve. Therefore, the
48 take the constant value, namely w i = 0.25 (i = 1, 2, 3, 4). For fur- 114
aerodynamic performance at angle of attack of 0 degrees was con-
49 ther consideration of the aerodynamic performance at negative 115
sidered in the optimum design of Design-II. Design-II shows better
50 angle of attack, another new airfoil, designated as Design-II, was 116
drag performance at the lower corner than Design-I, although its
51 designed considering the synthetically aerodynamic performance 117
drag performance is still slightly worse than the E387 airfoil. The
52 at angles of attack of −1, 0 and 5 degrees for Re = 2.0 × 105 . The 118
result shows that the special angles of attack should be taken into
53 constraints of design variables for Design-II are the same as those 119
account to the optimum design in practical applications.
54 of Design-I and the weights are w i = 1/6 (i = 1, 2, . . . , 6). The vari- 120
55 ables t and f of the two new airfoils are the same as those of Case-II: 121
56 the E387 airfoil. Furthermore, the aerodynamic performance of the In Case-II, the other two airfoils, Design-III and Design-IV, were 122
57 E387 airfoil that predicted by the XFOIL code at the selected angles designed considering the synthetically aerodynamic performance 123
58 of attack is taken as the expected parameters. Table 2 shows the at different angles of attack Re = 4.0 × 105 . For Design-III, the de- 124
59 optimization results. In addition, the shapes of the Design-I airfoil sign angles of attack are 0 and 8 degrees. The constraints of design 125
60 and the Design-II airfoil are presented in Fig. 10. variables are shown in Table 3 and all weights take the constant 126
61 In Figs. 11(a) and (b), the aerodynamic characteristics of value, namely w i = 0.25 (i = 1, 2, 3, 4). While for Design-IV, the 127
62 Design-I and Design-II for Re = 2.0 × 105 predicted by XFOIL code design angles of attack are −2, 0 and 8 degrees, which is for 128
63 were compared with those of the E387 airfoil. further consideration of the aerodynamic performance at negative 129
64 The results presented in Figs. 11(a) and (b) show that, the lift angle of attack. The constraints of design variables for Design-IV 130
65 coefficient characteristics of Design-I and Design-II have been im- are the same as those of Design-III and the weights are w i = 1/6 131
66 proved significantly at the design angles of attack, in comparison (i = 1, 2, . . . , 6). The variables t and f of Design-III and Design-IV 132
JID:AESCTE AID:4926 /FLA [m5G; v1.250; Prn:9/01/2019; 16:43] P.8 (1-10)
8 S. Zhang et al. / Aerospace Science and Technology ••• (••••) •••–•••

1 67
2 68
3 69
4 70
5 71
6 72
7 73
8 74
9 75
10 76
11 77
12 78
13 79
14 80
15 81
16 82
17 83
18 84
19 85
20 86
Fig. 11. Aerodynamic characteristics of Design-I and Design-II compared with those of the E387 airfoil for Re = 2.0 × 10 . (a) C l vs
5
α . (b) C l vs C d .
21 87
22 88
Table 3 signed airfoils also show an improvement of lift coefficient at all
23 89
Constraints of design variables for Design-III and Design-IV. other predicted angles of attack.
24 90
PSU 94-097 Design-III Design-IV Performance improvement can also be observed in Fig. 12(b).
25 91
t (%) 9.71 9.71 9.71 For Design-III, the drag coefficients at angles of attack of 0 and 8
26 92
xt (%) 32.70 [20, 50] [20, 50] degrees decrease by 4.17% and 4.02%, respectively. In addition, the
27 93
kt1 / [0.15, 0.6] [0.15, 0.6] reduction of drag coefficient can also be observed at other angles
28 kt2 / [0.15, 0.4] [0.15, 0.4] 94
of attack, especially at the upper corner. However, the Design-III 95
29 f (%) 4.02 4.02 4.02
30 x f (%) 43.30 [15, 60] [15, 60] airfoil shows a little worse drag performance than the PSU 94-097 96
31 kf1 / [−0.99, −0.2] [−0.99, −0.2] airfoil at the lower corner. While for Design-IV, the drag coeffi- 97
32
kf2 / [−0.99, −0.2] [−0.99, −0.2] cients at angles of attack of −2, 0 and 8 degrees decrease by 16%, 98
33
4.17% and 6.9%, respectively. Moreover, the results at other angles 99
34
of attack also present a drag coefficient reduction, especially at 100
35 Table 4 the upper corner. Because of the consideration of the aerodynamic 101
36
Optimization results for Design-III and Design-IV. performance at angle of attack of −2 degrees in the process of 102
37 PSU 94-097 Design-III Design-IV optimum design, Design-IV shows better drag performance at the 103
38 t (%) 9.71 9.71 9.71
lower corner than the PSU 94-097 airfoil. The above result also il- 104
39 xt (%) 32.70 24.13 24.52 lustrates the importance of the reasonable selection of the design 105
40
kt1 / 0.569 0.540 angle of attack. 106
kt2 / 0.400 0.387
41 107
f (%) 4.02 4.02 4.02
42 x f (%) 43.30 51.81 50.49
5. Conclusion 108
43 kf1 / −0.201 −0.311 109
44 kf2 / −0.268 −0.201 A new methodology has been developed for airfoil design. This 110
C l |α =−2◦ 0.2811 0.3987 0.3563
45 design methodology controls the airfoil profile through eight pa- 111
C l |α =0◦ 0.4942 0.6975 0.6529
46 C l |α =8◦ 1.3117 1.4750 1.4256 rameters. Four of the eight parameters, airfoil thickness, airfoil 112
47 C d |α =−2◦ 0.0100 0.0098 0.0084 camber, relative position of airfoil thickness and relative position 113
48 C d |α =0◦ 0.0072 0.0069 0.0069 of airfoil camber, have clear physical meaning and the others can 114
C d |α =8◦ 0.0174 0.0167 0.0162
49 adjust the airfoil profile locally. 115
50 To make the airfoil optimization design under modeFRONTIER 116
51 environment possible, the airfoil aerodynamic characteristics were 117
52 are the same as those of the PSU 94-097 airfoil. In addition, the predicted through the XFOIL code. The reliability verification was 118
53 aerodynamic performance of the PSU 94-097 airfoil that predicted conducted for the XFOIL code. The result shows that the XFOIL 119
54 by the XFOIL code at the selected angles of attack is taken as the code can well predict the aerodynamic performance of airfoils in 120
55 expected parameters. Table 4 shows the optimization results. Fur- the linear segment of the lift coefficient curve for low Reynolds 121
56 thermore, the shapes of the Design-III airfoil and the Design-IV number, although the maximum lift coefficient tends to be over 122
57 airfoil are presented in Fig. 10. predicted. 123
58 In Figs. 12(a) and (b), the aerodynamic characteristics of Application of this work was demonstrated through two cases. 124
59 Design-III and Design-IV for Re = 4.0 × 105 predicted by XFOIL For Case-I, the E387 airfoil was taken as the reference airfoil and 125
60 code were compared to those of the PSU 94-097 airfoil. The de- two airfoils were designed considering the synthetic aerodynamic 126
61 signed airfoils show a better performance at the design angles of performance at different angles of attack for Re = 2.0 × 105 . The 127
62 attack in comparison with PSU 94-097 airfoil. For Design-III, the design airfoils show better aerodynamic characteristics than those 128
63 lift coefficients increase at angles of attack of 0 and 8 degrees by of the E387 airfoil at the design angles of attack. Similarly, con- 129
64 41.14% and 12.45%, respectively. While for Design-IV, the lift co- sidering the synthetically aerodynamic performance at different 130
65 efficients at angles of attack of −2, 0 and 8 degrees increase by angles of attack for Re = 4.0 × 105 , Case-II shows two airfoils that 131
66 26.75%, 32.11% and 8.68%, respectively. Furthermore, the two de- achieved better aerodynamic characteristics than those of the PSU 132
JID:AESCTE AID:4926 /FLA [m5G; v1.250; Prn:9/01/2019; 16:43] P.9 (1-10)
S. Zhang et al. / Aerospace Science and Technology ••• (••••) •••–••• 9

1 67
2 68
3 69
4 70
5 71
6 72
7 73
8 74
9 75
10 76
11 77
12 78
13 79
14 80
15 81
16 82
17 83
18 84
19 85
20 86
Fig. 12. Aerodynamic characteristics of Design-III and Design-IV compared with that of the PSU 94-097 airfoil for Re = 4.0 × 10 . (a) C l vs
5
α . (b) C l vs C d .
21 87
22 88
23
94-097 airfoil at the design angles of attack. Furthermore, the re- [11] C. Bovais, P. Davidson, Flight testing the flying radar target (FLYRT), in: Bien-
89
sults also show that the selection of the design angle of attack has nial Flight Test Conference, American Institute of Aeronautics and Astronautics,
24 90
1994.
25
influence on the performance of the designed airfoil. In practical 91
[12] R. Foch, K. Ailinger, Low Reynolds number long endurance aircraft design, in:
26
applications, the special angles of attack should be taken into ac- Aerospace Design Conference, American Institute of Aeronautics and Astronau- 92
27
count to the optimum design. As a result of this work, it is clear tics, 1992. 93
28
that the design methodology of this work is an excellent airfoil de- [13] J. Morgado, M. Abdollahzadeh, M.A.R. Silvestre, J.C. Páscoa, High altitude pro- 94
sign tool. peller design and analysis, Aerosp. Sci. Technol. 45 (2015) 398–407, https://
29 95
doi.org/10.1016/j.ast.2015.06.011.
30 [14] R. Ma, B. Zhong, P. Liu, Optimization design study of low-Reynolds-number 96
31 Conflict of interest statement high-lift airfoils for the high-efficiency propeller of low-dynamic vehicles in 97
32 stratosphere, Sci. China, Technol. Sci. 53 (2010) 2792–2807, https://doi.org/10. 98
The authors declare that there is no conflict of interests regard- 1007/s11431-010-4087-0.
33 99
[15] J. Morgado, M. Silvestre, J. Páscoa, Parametric study of a high altitude airship
34 ing the publication of this article. 100
according to the multi-body concept for advanced airship transport, in: MAAT,
35 MEFTE 2012 – IV Conferência Nacional em Mecânica dos Fluidos, Termod- 101
36 Acknowledgements inâmica e Energia, Laboratório Nacional de Engenharia Civil, Lisboa, Portugal, 102
37 2012. 103
[16] S.G. Kontogiannis, D.E. Mazarakos, V. Kostopoulos, ATLAS IV wing aerodynamic
38 The present research work was supported by National Natural 104
design: from conceptual approach to detailed optimization, Aerosp. Sci. Tech-
39 Science Foundation of China [Grant No. 11172243]. nol. 56 (2016) 135–147, https://doi.org/10.1016/j.ast.2016.07.002.
105
40 [17] M.D. Maughmer, Design of winglets for high-performance sailplanes, J. Aircr. 106
41 References 40 (2003) 1099–1106, https://doi.org/10.2514/1.10817. 107
42 [18] H. Youngren, Multi-point design and optimization of an natural laminar flow 108
airfoil for a mission adaptive compliant wing, in: 46th AIAA Aerospace Sciences
43 [1] A. Gopalarathnam, B. Broughton, B. Mcgranahan, M. Selig, Design of low 109
Meeting and Exhibit, American Institute of Aeronautics and Astronautics, 2008.
44 Reynolds number airfoils with trips, J. Aircr. 40 (2003) 768–775, https:// 110
doi.org/10.2514/2.3157. [19] M. Selig, B. McGranahan, Wind tunnel aerodynamic tests of six airfoils for use
45 on small wind turbines, in: 42nd AIAA Aerospace Sciences Meeting and Exhibit, 111
[2] D. Arivoli, I. Singh, Self-adaptive flaps on low aspect ratio wings at low
46
Reynolds numbers, Aerosp. Sci. Technol. 59 (2016) 78–93, https://doi.org/10. American Institute of Aeronautics and Astronautics, 2004. 112
47 1016/j.ast.2016.10.006. [20] H. Li, S. Guo, Y.L. Zhang, C. Zhou, J.H. Wu, Unsteady aerodynamic and opti- 113
48 [3] M.S. Selig, J.J. Guglielmo, High-lift low Reynolds number airfoil design, J. Aircr. mal kinematic analysis of a micro flapping wing rotor, Aerosp. Sci. Technol. 63 114
34 (2012) 72–79, https://doi.org/10.2514/2.2137. (2017) 167–178, https://doi.org/10.1016/j.ast.2016.12.025.
49 115
[4] A. Gopalarathnam, M. Selig, Low speed NLF airfoils – a case study in inverse [21] B. Becker, M. Reyer, M. Swoboda, Steady and unsteady numerical investigation
50 of transitional shock-boundary–layer-interactions on a fan blade, Aerosp. Sci. 116
airfoil design, in: 37th Aerospace Sciences Meeting and Exhibit, American In-
51 stitute of Aeronautics and Astronautics, 1999. Technol. 11 (2007) 507–517, https://doi.org/10.1016/j.ast.2007.05.002. 117
52 [5] X.S. Meng, H.Y. Hu, X. Yan, F. Liu, S.J. Luo, Lift improvements using duty-cycled [22] M.S. Selig, Low Reynolds Number Airfoil Design Lecture Notes, Vki Lecture Se- 118
53 plasma actuation at low Reynolds numbers, Aerosp. Sci. Technol. 72 (2018) ries, 2003. 119
123–133, https://doi.org/10.1016/j.ast.2017.10.038. [23] J.W. van der Burg, J.E.J. Maseland, F.J. Brandsma, Low speed maximum lift and
54 120
[6] H. Ishikawa, Y. Ueda, N. Tokugawa, Natural laminar flow wing design for a low- flow control, Aerosp. Sci. Technol. 8 (2004) 389–400, https://doi.org/10.1016/j.
55 ast.2004.01.004. 121
boom supersonic aircraft, in: 55th AIAA Aerospace Sciences Meeting, American
56 Institute of Aeronautics and Astronautics, 2017. [24] C.N. Dong, Z.X. Zhang, Non Viscous Fluid Mechanics, Tsinghua University Press, 122
57 [7] M.S. Selig, J.J. Guglielmo, A.P. Broeren, P. Giguere, Summary of Low Speed Air- Beijing, 2003. 123
58 foil Data, SoarTech Publications, Virginia, USA, 1995. [25] M. Drela, XFOIL: An Analysis and Design System for Low Reynolds Number 124
[8] M.S. Selig, A. Gopalarathnam, P. Giguere, C.A. Lyon, Systematic airfoil design Airfoils, Springer, Berlin, Heidelberg, 1989, pp. 1–12.
59 125
studies at low Reynolds numbers, in: Fixed and Flapping Wing Aerodynamics [26] J.G. Coder, M.D. Maughmer, Comparisons of theoretical methods for predicting
60 126
for Micro Air Vehicle Applications, American Institute of Aeronautics and As- airfoil aerodynamic characteristics, J. Aircr. 51 (2014) 183–191, https://doi.org/
61 tronautics, 2001, pp. 143–167. 10.2514/1.c032232. 127
62 [9] G.K. Ananda, P.P. Sukumar, M.S. Selig, Measured aerodynamic characteristics [27] J. Morgado, R. Vizinho, M.A.R. Silvestre, J.C. Páscoa, XFOIL vs CFD performance 128
63 of wings at low Reynolds numbers, Aerosp. Sci. Technol. 42 (2015) 392–406, predictions for high lift low Reynolds number airfoils, Aerosp. Sci. Technol. 52 129
https://doi.org/10.1016/j.ast.2014.11.016. (2016) 207–214, https://doi.org/10.1016/j.ast.2016.02.031.
64 130
[10] R.J. Foch, P.L. Toot, Flight testing navy low Reynolds number (LRN) unmanned [28] M. Drela, M. Giles, Viscous-inviscid analysis of transonic and low Reynolds
65 131
aircraft, in: T.J. Mueller (Ed.), Low Reynolds Number Aerodynamics, Springer, number airfoils, in: 4th Applied Aerodynamics Conference, American Institute
66 Berlin, Heidelberg, 1989, pp. 407–417. of Aeronautics and Astronautics, 1986. 132
JID:AESCTE AID:4926 /FLA [m5G; v1.250; Prn:9/01/2019; 16:43] P.10 (1-10)
10 S. Zhang et al. / Aerospace Science and Technology ••• (••••) •••–•••

1 [29] J. van Ingen, The eN method for transition prediction: historical review of work [34] A. Abdoli, G.S. Dulikravich, Multi-objective design optimization of branch- 67
2 at TU Delft, in: 38th Fluid Dynamics Conference and Exhibit, American Institute ing, multifloor, counterflow microheat exchangers, J. Heat Transf. 136 (2014) 68
3 of Aeronautics and Astronautics, 2008. 101801–101810, https://doi.org/10.1115/1.4027911. 69
[30] R. Eppler, Airfoil Design and Data, Springer, Berlin Heidelberg, 1990. [35] J. Liu, M. Song, K. Chen, B. Wu, X. Zhang, An optimization methodology for
4 70
[31] R.J. Mcghee, B.S. Walker, B.F. Millard, Experimental results for the Eppler 387 wind lens profile using computational fluid dynamics simulation, Energy 109
5 airfoil at low Reynolds numbers in the Langley low-turbulence pressure tunnel, (2016) 602–611, https://doi.org/10.1016/j.energy.2016.04.131. 71
6 NASA Tech. Memo. 4062 (1988). [36] R. Duvigneau, M. Visonneau, Simulation and optimization of stall control for 72
7 [32] M.D. Maughmer, T.S. Swan, S.M. Willits, Design and testing of a winglet airfoil an airfoil with a synthetic jet, Aerosp. Sci. Technol. 10 (2006) 279–287, https:// 73
for low-speed aircraft, J. Aircr. 39 (2002) 654–661, https://doi.org/10.2514/2. doi.org/10.1016/j.ast.2006.01.002.
8 74
2978. [37] Esteco, ModeFRONTIER 2016 User’s Manual, 2016.
9 [33] S. Gaggero, D. Villa, G. Tani, M. Viviani, D. Bertetta, Design of ducted propeller 75
10 nozzles through a RANSE-based optimization approach, Ocean Eng. 145 (2017) 76
11 444–463, https://doi.org/10.1016/j.oceaneng.2017.09.037. 77
12 78
13 79
14 80
15 81
16 82
17 83
18 84
19 85
20 86
21 87
22 88
23 89
24 90
25 91
26 92
27 93
28 94
29 95
30 96
31 97
32 98
33 99
34 100
35 101
36 102
37 103
38 104
39 105
40 106
41 107
42 108
43 109
44 110
45 111
46 112
47 113
48 114
49 115
50 116
51 117
52 118
53 119
54 120
55 121
56 122
57 123
58 124
59 125
60 126
61 127
62 128
63 129
64 130
65 131
66 132

You might also like