wang2020
wang2020
2020-2666
June 15-19, 2020, VIRTUAL EVENT
AIAA AVIATION 2020 FORUM
Abstract
This paper numerically studies the Reynolds number effect for a 2D Co-Flow Jet (CFJ) airfoil and a 3D
wing at freestream Mach number of 0.46. The Reynolds averaged Navier-Stokes equations solver with Spalart-
Allmaras one-equation turbulence model is utilized for the simulation. The Reynolds number is decreased from
4.1 × 106 to 0.18 × 106 to mimic the altitude change from 10000m to 30000m. The results show that, at Reynolds
number of 4.1 × 106 , the best CFJ airfoil corrected aerodynamic efficiency ((CL /CD )c ) of 81.64 occurs at Cµ of
0.03 at AoA of 6◦ . When the Reynolds number is reduced to 0.18 × 106 , the (CL /CD )c has a 39.5% decrease.
It is because of the dramatic increase of the viscous drag by 66.2%. However, the lift coefficient drops only by
5.1%. The second reason for the decrease of the CFJ airfoil corrected aerodynamic efficiency is that the power
coefficient P c is increased significantly by 85% with the decrease of the Reynolds number. It is caused mostly
by the increased total pressure ratio of CFJ pumping due to higher loss. The 3D wing with an aspect ratio of
20 based on the same 2D airfoil is also studied for the same Reynolds numbers and freestream Mach number.
Similar to the 2D cases, the wing at low Reynolds number suffers significantly increased viscous drag and energy
loss of the CFJ.
Nomenclature
CF J Co-flow jet
AoA Angle of attack
LE Leading Edge
TE Trailing Edge
S Planform area
c Airfoil chord
U Flow velocity
q Dynamic pressure 0.5 ρ U 2
p Static pressure
η Pump efficiency
ρ Air density
ṁ Mass flow
M Mach number
ω Pitching Moment
∗
Graduate Student
†
Professor, ASME Fellow, AIAA associate Fellow
Copyright © 2020 by all the authors of this paper . Published by the American Institute of Aeronautics and Astronautics, Inc., with permission.
P Pumping power
∞ Free stream conditions
j Jet conditions
CL Lift coefficient L/(q∞ S)
CD Drag coefficient D/(q∞ S)
CM Moment coefficient
Cµ Jet momentum coef. ṁj Uj /(q∞ S)
(CL /CD )c CFJ airfoil corrected aerodynamic efficiency L/(D + P/V∞ )
Pc Power coefficient L/(q∞ S V∞ )
PR Total pressure ratio between injection and suction
Mis Isentropic Mach Number
M∞ Freestream Mach Number
Downloaded by UNIVERSITY OF GLASGOW on June 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2666
1 Introduction
Low Reynolds number airfoil performance is important for various applications [1], including high altitude long
endurance aircraft, wind turbines, Martian aerial vehicles, etc. A Reynolds number below 106 is considered as
the range of the low Reynolds number. The airfoil performance with the Reynolds number between 100,000 and
500,000 is in particular not well studied [1]. The flow on the suction surface (upper surface) tends to have laminar
flow separation bubbles and turbulent reattachment, which deteriorates the boundary layer, decreases the L/D,
creates the unsteadiness with hysteresis, and substantially reduces the airfoil operation range.
Recently, the Co-Flow Jet (CFJ) active flow control airfoil developed by Zha et al. [2, 3, 4, 5, 6, 7, 8, 9, 10,
11, 12, 13] provides a promising concept to increase maximum lift coefficient, stall AoA, and cruise efficiency at
low AoA. In a CFJ airfoil, an injection slot near the leading edge (LE) and a suction slot near the trailing edge
(TE) on the airfoil suction surface are created. As shown in Fig. 1, a small amount of mass flow is drawn into
the suction duct, pressurized and energized by the micro compressor, and then injected near the LE tangentially
to the main flow via an injection duct. The whole process does not add any mass flow to the system and hence is
a zero-net-mass-flux(ZNMF) flow control. The CFJ airfoil is demonstrated to achieve radical lift augmentation,
stall margin increase, drag reduction and moderate nose-down moment for stationary and pitching airfoils.
The CFJ airfoil has a unique low energy expenditure mechanism, because the jet gets injected at the leading
edge peak suction location, where the main flow pressure is the lowest and makes it easy to inject the flow, and it
gets sucked at the trailing edge, where the main flow pressure is the highest and makes it easy to draw the flow.
The low energy expenditure is a key factor enabling the CFJ airfoil to achieve ultra-high cruise efficiency [14] at
low AoA when the flow is benign.
Even though the CFJ active flow control is not specially aimed at low Reynolds number airfoil, the injection jet
mixing at near leading edge appears to force the boundary layer transition to turbulent via large vortex structures
[15] and avoid the major low Reynolds number airfoil problems. The three wind tunnel tests by Zha and his team
have the Reynolds number of 0.38 × 106 [5], 0.2 × 106 [15], and 0.2 − 0.7 × 106 [16] respectively. They are all in the
range of low Reynolds number flow. No obvious unsteadiness due to low Reynolds airfoil behaviors are observed
2
in the experiments. Since those studies are not specially designed to investigate low Reynolds number airfoil, it
may be inclusive to claim that CFJ airfoil has no problems for low Reynolds number flows. However, in this paper
we do assume the claim is true and simulate the flow with RANS model with full truculent boundary layer with
no laminar to transition modeling.
The purpose of this paper is to investigate the performance of the CFJ airfoil with the Reynolds number reduced
from 4.1 × 106 to 0.18 × 106 to mimic the effect of altitude change from 10000m to 30000m. Both 2D airfoil and
3D wing are simulated at a freestream Mach number of 0.46 with different Reynolds number.
Downloaded by UNIVERSITY OF GLASGOW on June 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2666
2 Methodology
The momentum and pressure at the injection and suction slots produce a reactionary force, which is automati-
cally measured by the force balance in wind tunnel testing. However, for CFD simulation, the full reactionary force
needs to be included. Using control volume analysis as shown in Fig. 2, the reactionary force can be calculated
using the flow parameters at the injection and suction slot opening surfaces. Zha et al. [3] give the following
formulations to calculate the lift and drag due to the jet reactionary force for a CFJ airfoil. By considering the
effects of injection and suction jets on the CFJ airfoil, the expressions for these reactionary forces are given as :
3
Fxcf j = (ṁj Vj1 + pj1 Aj1 ) ∗ cos(θ1 − α) − (ṁj Vj2 + pj2 Aj2 ) ∗ cos(θ2 + α) (1)
Fycf j = (ṁj1 Vj1 + pj1 Aj1 ) ∗ sin(θ1 − α) + (ṁj2 Vj2 + pj2 Aj2 ) ∗ sin(θ2 + α) (2)
where the subscripts 1 and 2 stand for the injection and suction respectively, and θ1 and θ2 are the angles between
the injection and suction slot’s surface and a line normal to the airfoil chord. α is the angle of attack.
The total lift and drag on the airfoil can then be expressed as:
where Rx0 and Ry0 are the surface integral of pressure and shear stress in x (drag) and y (lift) direction excluding
the internal ducts of injection and suction. For CFJ wing simulations, the total lift and drag are calculated by
integrating Eq. (3) and Eq. (4) in the spanwise direction.
The jet momentum coefficient Cµ is a parameter used to quantify the jet intensity. It is defined as:
ṁVj
Cµ = 1 2 (5)
2 ρ ∞ V∞ S
where ṁ is the injection mass flow, Vj is the mass-averaged injection velocity, ρ∞ and V∞ denote the free stream
density and velocity, and S is the planform area.
CFJ is implemented by mounting a pumping system inside the wing that withdraws air from the suction slot
and blows it into the injection slot. The power consumption is determined by the jet mass flow and total enthalpy
change as the following:
where Ht1 and Ht2 are the mass-averaged total enthalpy in the injection cavity and suction cavity respectively,
P is the Power required by the pump and ṁ the jet mass flow rate. Introducing Pt1 and Pt2 the mass-averaged
total pressure in the injection and suction cavity respectively, the pump efficiency η, and the total pressure ratio
of the pump Γ = PPt2t1
, the power consumption is expressed as:
4
where γ is the specific heat ratio equal to 1.4 for air. The power coefficient is expressed as:
P
Pc = 1 3
(8)
2 ρ∞ V∞ S
L
(9)
D
Downloaded by UNIVERSITY OF GLASGOW on June 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2666
For the CFJ wing, the ratio above still represents the pure aerodynamic relationship between lift and drag.
However since CFJ active flow control consumes energy, the ratio above is modified to take into account the
energy consumption of the pump. The formulation of the corrected aerodynamic efficiency for CFJ wings is:
L CL
( )c = (10)
D CD + Pc
where V∞ is the free stream velocity, P is the pumping power, and L and D are the lift and drag generated by
the CFJ wing. The formulation above converts the power consumed by the CFJ into a force VP∞ which is added to
the aerodynamic drag D. If the pumping power is set to 0, this formulation returns to the aerodynamic efficiency
of a conventional wing.
To compare aircraft that have the same ratio of initial weight to final weight with the same engine fuel con-
sumption or battery energy density, the productivity efficiency CL2 /CD is introduced to measure the productivity
of an airplane represented by its range multiplied by its weight [17].
The productivity efficiency CL2 /CD = CL (CL /CD ) is a more comprehensive parameter than the conventional
aerodynamic efficiency CL /CD to measure the merit of an airplane aerodynamic design for cruise performance.
The former includes not only the information of CL /CD , but also the information of the aircraft weight CL . For
example, for two airplane designs having the same CL /CD with one CL twice larger than the other, if the wing
sizes are the same, one airplane will be able to carry twice more weight than the other with productivity and wing
loading increased by 100%. Such a large difference is not reflected by CL /CD , but very well reflected by CL2 /CD .
The definition of CL /CD in general is a suitable measure of merit for conventional aircraft design. This is
because at a certain Mach number regime, the maximum CL /CD is usually achieved at low angle of attack within
the drag bucket and is more or less the same for different airfoil designs. In other words, for the same optimum
CL /CD , the CL is about the same. A typical CL for subsonic airfoil is about 0.4 and for transonic airfoil is about
0.7.
For CFJ airfoil, the minimum CFJ pumping power occurs at a fairly high AoA [8, 15]. With the augmentation
of CFJ, the subsonic cruise lift coefficient of a CFJ airfoil is typically 2 to 3 times higher than the conventional
airfoil with about the same (CL /CD )c [13]. Such a high lift coefficient is unattainable for conventional airfoil since
5
they would be either stalled or near stalled with very high drag. Hence for CFJ aircraft design, the productivity
efficiency CL2 /CD = CL (CL /CD ) is more informative to be used to reflect the aerodynamic performance. The
corrected productivity efficiency for CFJ airfoils is (CL2 /CD )c = CL2 /(CD + Pc ).
The FASIP (Flow-Acoustics-Structure Interaction Package) CFD code is used to conduct the numerical sim-
ulation. The 3D Reynolds Averaged Navier-Stokes (RANS) equations with one-equation Spalart-Allmaras [18]
turbulence model is used. A 3rd order WENO scheme for the inviscid flux [19, 20, 21, 22, 23, 24] and a 2nd order
central differencing for the viscous terms [19, 23] are employed to discretize the Navier-Stokes equations. The low
diffusion E-CUSP scheme used as the approximate Riemann solver suggested by Zha et al [20] is utilized with the
Downloaded by UNIVERSITY OF GLASGOW on June 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2666
WENO scheme to evaluate the inviscid fluxes. Implicit time marching method using Gauss-Seidel line relaxation
is used to achieve a fast convergence rate [25]. Parallel computing is implemented to save wall clock simulation
time [26].
The 3rd order accuracy no slip condition is enforced on the solid surface with the wall treatment suggested
in [27] to achieve the flux conservation on the wall. The computational mesh is shown in Fig. 3. The ducts
geometries are predetermined according to our previous designs. Total pressure, total temperature and flow angles
are specified at the injection duct inlet, as well as the upstream portion of the far field. Constant static pressure is
applied at the suction duct outlet as well as the downstream portion of the far field. The total mesh size is 47,200
points, split into 10 domains for the parallel computation. The first grid point on the wing surface is placed at
y + ≈ 1.
Table. 1 gives the detailed parameters of CFJ6421 airfoils with the injection and suction slot size normalized by
airfoil chord length (C). The CFJ6421-SST150-SUC247-INJ117 airfoil is optimized by Wang and Zha for its high
6
lift and cruise efficiency [28, 29, 30]. The CFJ6421-SST150-SUC247-INJ117 airfoil has a larger injection slot size
of 1.17%C and suction slot size of 2.47%. The suction surface translation (SST) of 1.50%C is the same as that of
CFJ6421-SST150-SUC133-INJ065 airfoil.
4 Simulated Cases
Downloaded by UNIVERSITY OF GLASGOW on June 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2666
Table. 2 lists all the freestream conditions and the CFJ momentum coefficients that are studied. Since the focus
is on the cruise performance, the AoA is limited to low value of 0◦ - 14◦ .
Table. 3, Table. 4 and Table. 5 list the 2D CFJ airfoil optimum aerodynamic efficiency at different Cµ . For all
different Reynolds numbers, the optimum corrected aerodynamic efficiency ((CL /CD )c ) is at Cµ of 0.03 and AoA
around 6◦ .
Table 3: Airfoil performance at different Cµ for CFJ airfoil with Reynolds number of 4.1 × 106 .
Cµ CL CD CL /CD (CL /CD )c (CL 2 /CD )c AoA Pc
0.01 1.0271 0.0107 95.762 74.689 76.713 2◦ 0.00303
0.02 1.1604 0.0078 149.708 73.991 85.859 5◦ 0.00969
0.03 1.7041 0.0122 139.689 81.644 139.129 6◦ 0.00867
0.04 1.7963 0.0101 178.583 75.788 136.136 6◦ 0.01377
0.05 1.8443 0.0092 201.004 64.933 119.756 6◦ 0.01923
7
Table 4: Airfoil performance at different Cµ for CFJ airfoil with Reynolds number of 2.6 × 106 .
Cµ CL CD CL /CD (CL /CD )c (CL 2 /CD )c AoA Pc
0.01 1.0211 0.0113 90.668 70.538 72.028 2◦ 0.00321
0.02 1.1553 0.0082 141.301 70.111 80.999 6◦ 0.00830
0.03 1.5922 0.0105 132.489 77.186 122.894 5◦ 0.01015
0.04 1.7894 0.0104 171.440 72.044 128.913 6◦ 0.01440
0.05 1.8419 0.0094 196.940 62.269 114.692 6◦ 0.02023
Table 5: Airfoil performance at different Cµ for CFJ airfoil with Reynolds number of 0.18×6 .
Cµ CL CD CL /CD (CL /CD )c (CL 2 /CD )c AoA Pc
Downloaded by UNIVERSITY OF GLASGOW on June 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2666
Since the the optimum (CL /CD )c is at Cµ of 0.03 for all Reynolds number. The interest and discussion will
be focused on the Cµ of 0.03 and the AoA from 0◦ to 14◦ . Fig. 4 shows the that the lift reduction for Reynolds
number from 4.1 × 106 to 0.18 × 106 is about 3.86%, the lift reduction for Reynolds number from 4.1 × 106 to
2.6 × 106 is less than 1%. However, the drag coefficient is increased by 39.0% for Reynolds number from 4.1 × 106
to 0.18 × 106 , and is increased only by 4% for Reynolds number of 2.6 × 106 . The substantial drag increase at low
Reynolds number is due to the increase of viscous drag.
Table. 6 is the lift and drag force breakdowns with the contributions of pressure, surface viscous friction, and
jet reactionary force. CLP and CLV are the pressure and viscous contribution to lift, CDP and CDV are pressure
and viscous contribution to drag, Fx and Fy are the jet reactionary force contribution to the drag and lift. Table.
6 indicates that when the Reynolds number is decreased from 4.1 × 106 to 0.18 × 106 , the most significant change
is the viscous friction drag contribution CDV , which is increased by 66.2%.
Table 6: Detailed force breakdown for the 2D CFJ airfoil at different Reynolds number.
8
(a) (b) (c)
Downloaded by UNIVERSITY OF GLASGOW on June 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2666
Figure 4: Lift, drag and moment coefficients for the CFJ airfoil at different Reynolds numbers.
Fig. 5 shows that the aerodynamic efficiency, corrected CFJ aerodynamic efficiency, and CFJ productivity
efficiency for Reynolds number of 0.18 × 106 has a 31%, 40%, and 42% reduction respectively compared with the
case of Reynolds number of 4.1 × 106 .
Figure 5: CL /CD , (CL /CD )c , (CL 2 /CD )c VS AoA for the CFJ airfoil at different Reynolds numbers.
Fig. 6 shows that for the CFJ airfoil at low Reynolds number, the injection total pressure is increased by 6.6%
in order to reach the same Cµ . However, the suction static pressure is not decreased as much, only about 2.3% at
AoA of 6◦ . The CFJ total pressure ratio is increased by 9.1%.
9
(a) (b) (c)
Downloaded by UNIVERSITY OF GLASGOW on June 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2666
Figure 6: Injection total pressure, Suction static pressure and the total pressure ratio between injection and
suction plots for CFJ airfoil at different Reynolds numbers.
Fig. 7 shows that the power coefficient is increased by 75% for Reynolds number of 0.18 × 106 . This is mostly
because the pressure ratio is substantially increased as shown in Fig. 6. The injection total pressure is increased
which caused the injection velocity increases as well under the same static pressure.
Figure 7: CFJ Power coefficient, Normalized mass flow rate and injection velocity VS AoA for the CFJ airfoil at
different Reynolds numbers.
Fig. 8 shows the Mach number contours of the flow field for different Reynolds numbers at AoA of 6◦ . Qualita-
tively, they do not look very different. But quantitatively, Fig. 5 - 7 show that the CFJ airfoil at the low Reynolds
number suffers higher viscous loss and higher CFJ pressure ratio and power.
10
(a) Re = 4.1 × 106 (b) Re = 2.6 × 106 (c)Re = 0.18 × 106
Downloaded by UNIVERSITY OF GLASGOW on June 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2666
Figure 8: Mach contour plots for CFJ airfoil at different Reynolds numbers.
Fig. 9 is the airfoil surface isentropic Mach number distributions at AoA=6◦ for the different Reynolds number.
The difference is not very large as indicated by the lift coefficients in Fig. 4.
To study the tip vortex effect of 3D wing at different Reynolds numbers, a 3D wing is stacked using the same 2D
CFJ airfoil in the last section. The aspect ratio is 20. The Mach number is still 0.46 and the Reynolds numbers
are the same as the 2D cases.
When the AoA varies at cruise, it is advantageous to hold the injection total pressure constant as the control law
for the micro-compressor [30]. This section hence simulates the CFJ wings performance at cruise condition with
the injection total pressure held as constant. The total pressure value fixed is the one at the optimum aerodynamic
efficiency point at AoA of 5◦ for all different Reynolds numbers with Cµ of 0.03. Since the injection total pressure
is constant, the Cµ will vary at different AoA.
11
The Pc and Cµ plots for the CFJ wings at the fixed injection total pressure are shown in Fig. 10. Different
from the constant Cµ cases shown in Fig. 7, both the Pc and Cµ have very low value at low AoA and are linearly
increased with the AoA even when the airfoil is stalled.
There are two reasons that the power coefficient Pc is increased with the AoA when the injection total pressure
is held as constant. First, the mass flow rate is increased due to the leading edge suction pressure decrease. The
increased mass flow increases the power coefficient as shown in Eq. (7) and Eq. (8). Second, the total pressure
ratio loss is increased when the AoA is increased since the injection velocity is higher and the boundary layer goes
through more diffusion before going into the suction duct. With a lower suction total pressure, the pressure ratio
is higher to reach the same injection total pressure, the power coefficient is thus increased. Fig. 10 also indicates
that the CFJ power coefficient is higher at the low Reynolds number for the same reason explained for the 2D
case, which is that the higher viscous loss due to low Reynolds number requires more power to keep the same Cµ .
Downloaded by UNIVERSITY OF GLASGOW on June 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2666
(a) (b)
Figure 10: Pc and Cµ plots for CFJ Wings at different Reynolds number and at M∞ = 0.46.
As shown in Fig. 11 (a), away from AoA of 4◦ , the injection mass flow rate is slightly affected by the Reynolds
number, so is the injection jet velocity as shown in Fig. 11 (b).
(a) (b)
Figure 11: Normalized injection mass flow rate ṁinj and V inj plots for CFJ Wing at different Reynolds number.
12
The lift, drag and moment coefficient vs. AoA for the CFJ wing at a fixed injection total pressure at different
Reynolds number are shown in Fig. 12. At the low Reynolds number, the maximum CL drops significantly at
AoA of 10◦ . The drag coefficient CD at the low Reynolds number is increased across the whole profile as in the
2D case due to the increased viscous drag (Fig. 12 (b)). Since the total pressure is fixed, which allows Cµ varies
by its iterations. The stall AoA is about 10◦ , higher than the 2D cases.
Downloaded by UNIVERSITY OF GLASGOW on June 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2666
Figure 12: Lift, drag and moment coefficients for CFJ airfoil at different Reynolds number.
The CL /CD , (CL /CD )c , (CL 2 /CD )c at different Reynolds number for the CFJ wing are shown in Fig. 13. The
low Reynolds number case has the substantially reduced efficiency for all the AoA as in the 2D cases. As shown in
Fig. 13 (a), all the peak CL /CD occurs at AoA of 2◦ . However, the peak (CL /CD )c occurs has a plateau between
AoA of 2◦ and 6◦ . That is the advantage of holding the injection total pressure constant as indicated by Wang
and Zha [30]. The productivity efficiency has a plateau between AoA of 4◦ and 10◦ for the high Reynolds number
cases. For the low Reynolds number case, the productivity efficiency drops sharply at AoA greater than 6◦ .
Figure 13: CL /CD , (CL /CD )c , (CL 2 /CD )c plots for CFJ airfoil at different Reynolds number.
Fig. 14 (a) shows the suction static pressure versus AoA at different Reynolds numbers. The injection total
pressure are 4.07, 3.87, and 3.84 for Reynolds number of 0.18 × 106 , 2.6 × 106 , and 4.1 × 106 respectively, which
decreases when the Reynolds number increases. Fig. 14 (b) shows the pressure ratio, with the decrease of the
13
Reynolds number, the P R increases due to more viscous loss.
Downloaded by UNIVERSITY OF GLASGOW on June 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2666
(a) (b)
Figure 14: P t and P R plots for CFJ airfoil at different Reynolds number.
Fig. 15 shows the Isentropic Mach number distribution of the wing section from the wing root to tip at different
span locations. At the outer span close to the wing tip, the loading is decreased due to wing tip vortices. The
Isentropic Mach number distribution are similar for different Reynolds numbers.
Figure 15: Isentropic Mach Number distributions for the 3D CFJ wings at different Reynolds number at M∞ of
0.46 and AoA of 5◦ .
Fig. 16 shows the Mach contours at 10%, 50%, 75% and 95% spanwise location for the CFJ wing at different
Reynolds numbers, AoA = 5◦ , M∞ = 0.46. The loading decreases toward the tip. Similar to the 2D case, the
flow fields look qualitatively very similar for the different Reynolds numbers, but the drag and energy loss is
significantly more for the low Reynolds number case.
14
(a) Re = 0.18e6 (b) Re = 2.6e6 (c) Re = 4.1e6
Downloaded by UNIVERSITY OF GLASGOW on June 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2666
Figure 16: Mach contours at 10%, 50%, 75% and 95% spanwise location for the CFJ wing at different Reynolds
numbers, and at AoA = 5◦ , M∞ = 0.46.
7 Conclusion
This paper numerically studies the Reynolds number effect for a 2D Co-Flow Jet (CFJ) airfoil and a 3D wing
at freestream Mach number of 0.46. The Reynolds averaged Navier-Stokes equations solver with Spalart-Allmaras
turbulence model is utilized for the simulation. The Reynolds number is decreased from 4.1 × 106 to 0.18 × 106 to
mimic the altitude change from 10000m to 30000m. The results show that, at Reynolds number of 4.1 × 106 , the
15
best CFJ airfoil corrected aerodynamic efficiency ((CL /CD )c ) of 81.64 occurs at Cµ of 0.03 at AoA of 6◦ . When
the Reynolds number is reduced to 0.18 × 106 , the (CL /CD )c has a 39.5% decrease. It is observed that the reason
is due to the dramatic increase of the viscous drag by 66.2% when the Reynolds number is decreased. However,
the lift coefficient drops only by 5.1%. The second reason for the decrease of the CFJ airfoil corrected aerodynamic
efficiency is that the power coefficient P c is increased significantly with the decrease of the Reynolds number. It
is caused mostly by the increased total pressure ratio of CFJ pumping due to higher loss. The power coefficient
is increased by 85% when the Reynolds number is reduced from 4.1 × 106 to 0.18 × 106 . The 3D wing with an
aspect ratio of 20 based on the same 2D airfoil is also studied for the same Reynolds numbers and freestream
Mach number. Similar to the 2D cases, the wing at low Reynolds number suffers significantly increased viscous
drag and energy loss of the CFJ.
Downloaded by UNIVERSITY OF GLASGOW on June 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2666
8 Acknowledgment
The simulations are conducted on Pegasus super computing system at the Center for Computational Sciences
at the University of Miami.
Disclosure: The University of Miami and Dr. Gecheng Zha may receive royalties for future commercialization
of the intellectual property used in this study.
References
[1] T. Mueller, “Low Reynolds Number Vehicles.” AGARDograph No. 288, ISBN 92-835-1486-6, Feb. 1985.
[2] G.-C. Zha and D. C. Paxton, “A Novel Flow Control Method for Airfoil Performance Enhancement Using
Co-Flow Jet.” Applications of Circulation Control Technologies, Chapter 10, p. 293-314, Vol. 214, Progress in
Astronautics and Aeronautics, AIAA Book Series, Editors: Joslin, R. D. and Jones, G.S., 2006.
[3] G.-C. Zha, W. Gao, and C. Paxton, “Jet Effects on Co-Flow Jet Airfoil Performance,” AIAA Journal, No.
6,, vol. 45, pp. 1222–1231, 2007.
[4] G.-C. Zha, C. Paxton, A. Conley, A. Wells, and B. Carroll, “Effect of Injection Slot Size on High Performance
Co-Flow Jet Airfoil,” AIAA Journal of Aircraft, vol. 43, 2006.
[5] G.-C. Zha, B. Carroll, C. Paxton, A. Conley, and A. Wells, “High Performance Airfoil with Co-Flow Jet Flow
Control,” AIAA Journal, vol. 45, 2007.
[6] Wang, B.-Y. and Haddoukessouni, B. and Levy, J. and Zha, G.-C., “Numerical Investigations of Injection Slot
Size Effect on the Performance of Co-Flow Jet Airfoil,” Journal of Aircraft, vol. Vol. 45, No. 6,, pp. pp.2084–
2091, 2008.
[7] B. P. E. Dano, D. Kirk, and G.-C. Zha, “Experimental Investigation of Jet Mixing Mechanism of Co- Flow
Jet Airfoil.” AIAA-2010-4421, 5th AIAA Flow Control Conference, Chicago, IL, 28 Jun - 1 Jul 2010.
[8] B. P. E. Dano, G.-C. Zha, and M. Castillo, “Experimental Study of Co-Flow Jet Airfoil Performance Enhance-
ment Using Micro Discreet Jets.” AIAA Paper 2011-0941, 49th AIAA Aerospace Sciences Meeting, Orlando,
FL, 4-7 January 2011.
[9] A. Lefebvre, B. Dano, W. Bartow, M. Fronzo, and G. Zha, “Performance and energy expenditure of coflow
jet airfoil with variation of mach number,” Journal of Aircraft, vol. 53, no. 6, pp. 1757–1767, 2016.
16
[10] A. Lefebvre, G-C. Zha, “Numerical Simulation of Pitching Airfoil Performance Enhancement Using Co-Flow
Jet Flow Control,” AIAA paper 2013-2517, June 2013.
[11] A. Lefebvre, G-C. Zha, “Cow-Flow Jet Airfoil Trade Study Part I : Energy Consumption and Aerodynamic
Performance,” 32nd AIAA Applied Aerodynamics Conference, AIAA AVIATION Forum, AIAA 2014-2682,
June 2014.
[12] A. Lefebvre, G-C. Zha, “Cow-Flow Jet Airfoil Trade Study Part II : Moment and Drag,” 32nd AIAA Applied
Aerodynamics Conference, AIAA AVIATION Forum, AIAA 2014-2683, June 2014.
[13] Lefebvre, A. and Zha, G.-C., “Trade Study of 3D Co-Flow Jet Wing for Cruise Performance.” AIAA Paper
2016-0570, AIAA SCITECH2016, AIAA Aerospace Science Meeting, San Diego, CA, 4-8 January 2016.
[14] Lefebvre, A. and Zha, G.-C. , “Design of High Wing Loading Compact Electric Airplane Utilizing Co-Flow Jet
Downloaded by UNIVERSITY OF GLASGOW on June 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2666
Flow Control.” AIAA Paper 2015-0772, AIAA SciTech2015: 53nd Aerospace Sciences Meeting, Kissimmee,
FL, 5-9 Jan 2015.
[15] Lefebvre, A. and Dano, B. and Bartow, W. and Di Franzo, M. and Zha, G.-C., “Performance Enhancement
and Energy Expenditure of Co-Flow Jet Airfoil with Variation of Mach Number.” AIAA Paper 2013-0490,
AIAA Journal of Aircraft, DOI: 10.2514/1.C033113, 2016.
[16] G.-C. Zha, Y.-C. Yang, Y. Ren, and B. McBreen, “Super-lift and thrusting airfoil of coflow jet-actuated by
micro-compressors.” AIAA Paper 2017-3061, AIAA AVIATION 2018, Atlanta, GA, Submitted for publication
in AIAA Journal , 25 - 29 June 2018.
[17] Yunchao Yang and Gecheng Zha, “Super-Lift Coefficient of Active Flow Control Airfoil: What is the Limit?.”
AIAA Paper 2017-1693, AIAA SCITECH2017, 55th AIAA Aerospace Science Meeting, Grapevine, January
9-13 2017.
[18] P. R. Spalart and S. R. Allmaras, “A one-equation turbulence model for aerodynamic flows,” in 30th Aerospace
Sciences Meeting and Exhibit, Aerospace Sciences Meetings, Reno, NV, USA, AIAA Paper 92-0439, 1992.
[19] Y.-Q. Shen and G.-C. Zha, “Large Eddy Simulation Using a New Set of Sixth Order Schemes for Compressible
Viscous Terms ,” Journal of Computational Physics, vol. 229, pp. 8296–8312, 2010.
[20] Zha, G.C., Shen, Y.Q. and Wang, B.Y., “An improved low diffusion E-CUSP upwind scheme ,” Journal of
Computer and Fluids, vol. 48, pp. 214–220, Sep. 2011.
[21] Y.-Q. Shen and G.-Z. Zha , “Generalized finite compact difference scheme for shock/complex flowfield inter-
action,” Journal of Computational Physics, vol. doi:10.1016/j.jcp.2011.01.039, 2011.
[22] Shen, Y.-Q. and Zha, G.-C. and Wang, B.-Y., “ Improvement of Stability and Accuracy of Implicit WENO
Scheme,” AIAA Journal, vol. 47, No. 2, pp. 331–344, 2009.
[23] Shen, Y.-Q. and Zha, G.-C. and Chen, X.-Y., “ High Order Conservative Differencing for Viscous Terms
and the Application to Vortex-Induced Vibration Flows,” Journal of Computational Physics, vol. 228(2),
pp. 8283–8300, 2009.
[24] Shen, Y.-Q. and Zha, G.-C. , “ Improvement of the WENO Scheme Smoothness Estimator,” International
Journal for Numerical Methods in Fluids, vol. DOI:10.1002/fld.2186, 2009.
[25] G.-C. Zha and E. Bilgen, “Numerical Study of Three-Dimensional Transonic Flows Using Unfactored Upwind-
Relaxation Sweeping Algorithm,” Journal of Computational Physics, vol. 125, pp. 425–433, 1996.
17
[26] B.-Y. Wang and G.-C. Zha, “A General Sub-Domain Boundary Mapping Procedure For Structured Grid
CFD Parallel Computation,” AIAA Journal of Aerospace Computing, Information, and Communication,
vol. 5, No.11, pp. 2084–2091, 2008.
[27] Y.-Q. Shen, G.-C. Zha, and B.-Y. Wang, “Improvement of Stability and Accuracy of Implicit WENO Scheme
,” AIAA Journal, vol. 47, pp. 331–344, 2009.
[28] Y. Wang and G.-C. Zha, “Study of 3D Co-flow Jet Wing Induced Drag and Power Consumption at Cruise
Conditions.” AIAA Paper 2019-0034, AIAA SciTech 2019, San Diego, CA, January 7-11, 2019.
[29] Y. Wang, Y.-C. Yang, and G.-C. Zha, “Study of Super-Lift Coefficient of Co-Flow Jet Airfoil and Its Power
Consumption.” AIAA Paper 2019-3652, AIAA Aviation 2019, AIAA Applied Aerodynamics Conference,
Dallas, Texas, 17-21 June 2019.
Downloaded by UNIVERSITY OF GLASGOW on June 27, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2666
[30] Y. Wang and G.-C. Zha, “Study of Mach Number Effect for 2D Co-Flow Jet Airfoil at Cruise Conditions.”
AIAA Paper 2019-3169, AIAA Aviation 2019, AIAA Applied Aerodynamics Conference, Dallas, Texas, 17-21
June 2019.
18