You are on page 1of 19

AIAA 2001-2974

Effective Flow Control for Rotorcraft


Applications at Flight Mach Numbers
(Invited)

Hassan Nagib
John Kiedaisch
David Greenblatt
Illinois Institute of Technology, Chicago, IL

Israel Wygnanski
Tel Aviv University, Israel
University of Arizona, Tuscon, AZ

Ahmed Hassan
The Boeing Company, Mesa, AZ

31st AIAA Fluid Dynamics


Conference & Exhibit
11-14 June 2001 / Anaheim, CA
AIAA-2001-2974

EFFECTIVE FLOW CONTROL FOR ROTORCRAFT APPLICATIONS AT


FLIGHT MACH NUMBERS (Invited)

Hassan Nagib∗, John Kiedaisch†, David Greenblatt†‡


Illinois Institute of Technology, Chicago, IL

Israel Wygnanski‡§
University of Arizona, Tuscon, AZ

Ahmed Hassan¶
The Boeing Company, Mesa, AZ

Abstract changing flow conditions, and flow control may be


Pulsating zero-mass flux jets introduced from achieved through solid or fluidic actuators. In either
spanwise slots at various locations on the upper surface case, the flow field is changed globally by locally acting
of an oscillating VR-7 airfoil model are shown to be on the flow. This may be achieved by arrays of
effective in controlling lift, moment and drag actuators that energize the boundary layer with steady or
coefficients over the range of Mach numbers from 0.1 pulsed momentum through the wing surface. These
to 0.4. This control is demonstrated over a wide range types of controlled actuation change the velocity,
of mean angles of attack of the oscillating airfoil from pressure, and vorticity field around the wing to achieve
light to deep stall conditions. Maintaining the non- the desired objectives.
dimensional frequency and amplitude of the forcing It has been observed that relatively large quantities
unchanged results in comparable modifications of the of steady blowing (Cµ = 10%) near the point of
aerodynamic coefficients throughout this Mach number separation can reattach the flow and increase lift, but the
range. Therefore, it appears that this active-flow control steady blowing may also cause a thickening of both the
technique is only limited by the ability to generate the boundary layer and the wake behind the airfoil which
adequate forcing conditions at the higher Mach numbers leads to increased drag.
required for applications such as rotorcraft. In contrast to steady blowing, the oscillatory
blowing takes advantage of inherent local instabilities in
1. Introduction & Background the near-wall shear layer that causes the selective
In modern design of military aircraft, separation amplification of the input oscillation frequency. These
control is vital to improving the flight characteristics of amplified disturbances convect downstream along the
airfoils whether the application is highly maneuverable airfoil as coherent large structures that serve to mix the
fighters, stealth bombers, or micro air vehicles. When boundary layer flow and delay separation. The
air separates from a wing in flight, the result is loss of efficiency of the mixing provides substantial increases
lift and increase in drag that threatens the stability of the in lift, while concomitantly reducing drag even when
aircraft and the safety of the pilot. Separation is one assumes that the entire momentum added recovered
typically avoided by geometric changes and by flying as thrust. The unsteady forcing can also be tailored to
the aircraft within the flight envelope. affect some global instabilities of the separated flow and
Alternatives to fixed-geometry aircraft are flexible lead to reduction in the size of the region of separation
structures, where the wings and fuselage are constructed and improved performance of the wing.
from materials that can be deformed to adapt to


Rettaliata Professor, Mechanical, Materials & Aerospace Engineering Department. Associate Fellow AIAA.

Senior Research Associate. Mechanical, Materials & Aerospace Engineering Department. Member AIAA.

Department of Fluid Mechanics & Heat Transfer, Tel Aviv University.
§
Professor, Fellow AIAA.

Associate Technical Fellow, Associate Fellow AIAA.
Copyright  2001 by the American Institute of Aeronautics and Astronautics Inc. All rights reserved.

1
American Institute of Aeronautics and Astronautics
Pulsed blowing has proven to be a reliable technique Recently, periodic excitation (or forcing) has been
for separation control. Recent work by Seifert et al. demonstrated as an effective, efficient and practical
(1993, 1996, 1998, 1999), Wygnanski (1997), and Hites method for controlling incompressible dynamic stall
et al. (1997, 2001) has shown repeatedly that low (e.g. Greenblatt & Wygnanski, 1999). Based on these
amplitude, oscillatory blowing can delay separation and results it is obvious to expect that the technique may
enhance lift over a wide range of Reynolds numbers also be effective in improving the performance of a
including those corresponding to aircraft takeoff and wide range of airfoils used in the rotorcraft industry
landing. These experiments have demonstrated several such as those carefully documented by McAlister et al.
consistent results, including: 1) the most effective (1982). Carr (1988) shows, however, that
location for unsteady forcing is near the point of compressibility can have a profound effect on dynamic
separation, 2) the optimum reduced frequency for the stall, even at relatively moderate Mach numbers, i.e.
oscillations is about F+ = fc/U ≈1, and 3) the amplitude M = 0.3, when the flow can be supersonic in the
of the oscillations required for effective separation leading-edge region. Although in Carr’s case the airfoil
control is about two orders of magnitude lower than that stall was dominated by leading edge separation, where
for steady blowing. When oscillatory separation control compressibility effects may be exaggerated by a shock
is applied to an airfoil that is not separated, the effects boundary layer interaction, Carr and Chandrasekhara
are not detrimental to lift. (1996) projected that such effects would lead to the
McManus and Magill (1997) performed experiments failure of flow control methodologies in all unsteady
on a 3-D, finite lambda wing model and demonstrated airfoil applications for Mach numbers equal to or larger
the effectiveness of pulsed vortex jets up to M = 0.2. than about 0.3. When attempting dynamic stall control,
The improvement in the lift coefficient was less then compressibility must be seriously considered because
that of Seifert et al.; however, the method and the typical full-scale Mach numbers on a rotorcraft
amplitude of the blowing were different. In the retreating blade in the vicinity of dynamic stall are in
McManus and Magill experiments, four round jets were the range from 0.3 to 0.5. (Apparently, the effect of
used to pulsate the flow locally just downstream of the Reynolds number is less understood due to the difficulty
leading edge, whereas the work by Seifert used a linear of varying Reynolds number significantly without
slot over the entire span of the airfoil at essentially x/c = introducing compressibility effects.)
0. McManus introduced streamwise vorticity into the The review by Carr and Chandrasekhara
boundary layer while Seifert introduced an unsteady (1996) shows that no definitive conclusions can be
vorticity with a predominantly spanwise component. drawn concerning the impact of the supersonic flow in
In a numerical investigation by Towne and the leading-edge region. One of the problems is that
Buter (1994), subharmonic pulsed blowing applied from various investigators report different findings for
the leading edge in an M = 0.2 simulation. Their results experiments at nominally the same conditions on
showed that the oscillations were ineffective at identical airfoils. Nevertheless, two main conclusions
controlling dynamic stall, whereas steady blowing was are apparent: Firstly, there is a limit to the magnitude of
effective. This contradicts recent experimental findings the suction peak that can be reached at high incidence in
with pulsed blowing that demonstrate more efficient compressible flow and this limit is due to the local
prevention of separation than steady blowing in airfoils. supersonic velocity on the surface of the airfoil.
Although the Mach number was similar to the Secondly, compressibility effects can completely
experimental studies, the chord Reynolds number was change the flow behavior compared to that observed at
about three orders of magnitude lower than the low Mach numbers (M < 0.3).
experimental work of Seifert. Also the dimensionless The first attempt made to carefully and
blowing frequency, F+, was about 10-3 in Towne and independently examine the conclusion by Carr and
Buter’s study, several orders of magnitude below the Chandrasekhara (1996), regarding the detrimental
optimal pulsing frequency of other researchers. A effects of compressibility on control of separation, and
comprehensive computational model has yet to be to examine the effectiveness of pulsed jets in the
developed to describe the effects of oscillatory blowing enhancement of airfoil performance at Mach numbers
on stall, although Wu et al. (1997) have begun to with compressibility effects, is contained in the thesis of
propose schemes which address non-linear mode Hites (1997); see also Hites et al. (2001). These
interaction, resonance, and vortex dynamics. Very experiments demonstrated effective control up to Mach
recently, extensive numerical simulations by various numbers above 0.4. More recently, Seifert & Pack
groups have achieved more success in representing the (1999) investigated the effect of periodic forcing on a
various flow modules of such complex flow, e.g., see static NACA 0015 airfoil under the conditions
Hassan and Munts (2000). 0.28 ≤ Ma ≤ 0.55 . For all cases, forcing had a
beneficial effect on lift and drag, with the most

2
American Institute of Aeronautics and Astronautics
significant effects in the post-stall regime. In fact, transducer with a ±12.5 psi range. The airfoil was
Seifert & Pack showed that upper surface suction that is equipped with two oscillatory blowing (forcing) slots, at
attenuated as a result of an increase in Mach number x/c = 0.3 and 0.5, that were used independently. Forcing
from 0.28 to 0.4 could be restored by forcing at low was achieved by means of four large acoustic-speaker-
perturbation amplitudes (Cµ < 0.1%). On the other type devices. These actuators are based on a Boeing
hand, Greenblatt et al. (1999) have shown that the net design and were fabricated by Domzalski Machine of
result of forcing in incompressible flows is essentially Mesa, AZ. The oscillatory flow emanating form the
insensitive to whether an airfoil is stationary or slots were calibrated by a hot-wire at the slot exit. To
oscillating in pitch. This is primarily due to the large obtain the total drag, the wake profile was measured by
disparity between the time-scales characterizing the a fixed Pitot rake with one hundred 0.040” diameter
airfoil pitch oscillations and those characterizing the (OD) tubes located at approximately x/c = 2 behind the
forcing-generated large coherent structures (LCSs). It airfoil trailing edge. An Endevco pressure transducer
can therefore be speculated that periodic forcing is also (model 8514-10, ±10 psi range) was surface-mounted at
effective in controlling dynamic stall under x / c = 0.4 , at the same chordwise location as a pressure
compressible conditions, even at low perturbation port, but at a spanwise distance of 12.7mm from the
amplitudes. pressure port. The dynamic response of the pressure
ports, vinyl tubing and scanivalve, were calibrated and
2. Objectives & Scope modeled (see Greenblatt et al, 2001).
The global objective of this work was to study the
effectiveness of periodic forcing on the control of 4. Computational Grid
rotorcraft dynamic stall at typical flight Mach numbers. The oscillatory jet flow (periodic forcing) was
To this end, both an experimental and a numerical simulated using a time-dependent boundary condition at
investigation were conducted on a VR-7 airfoil. x/c = 0.5 at an angle of 25 degrees relative to the local
It is well known that significant changes in the tangent to the airfoil surface. All computations were
characteristics of stall occur due to compressibility for performed on a C-type mesh having a resolution of
M > 0.3. Hence, the approach adopted in this 297x81, using the NASA Ames hyperbolic grid
investigation is an incremental one: firstly, the effect of generator, HYGRID (Cordova & Barth, 1988). Fig. 1
forcing on dynamic stall under incompressible depicts a close-up view of a typical computational grid
conditions (M = 0.1 and 0.2) was evaluated; and for the VR-7 airfoil with clustering near the x/c = 0.5
secondly, the effect of forcing while gradually position to accurately resolve the synthetic jet flow.
increasing M into the compressible regime is assessed
by comparison with the incompressible cases. Both the 5. Discussion of Experimental Results
experimental data and numerical predictions presented 5.1 Baseline Results
here are of a preliminary nature. Experimental data is a Fig. 2 shows a comparison of the Endevco signal on
summary of a detailed investigation carried out in the the airfoil surface, compared with the measured and
NDF at IIT and considers data for M ≤ 0.4. Numerical corrected pressure-port signals respectively, under the
predictions are of a more preliminary nature and are conditions M = 0.3, k = 0.05 and α = 17°+5°sin(ω t).
presented for M ≤ 0.3. The airfoil spends the majority of its time in the post-
stall regime, some 10° beyond the static stall angle. As
3. Experimental Apparatus such, this deep-stall case may be considered particularly
A VR-7 airfoil with a 22 in. span was mounted challenging for the correction procedure.
vertically in the NDF test section (c.f. Hites et al, 2001). Notwithstanding, the method performs well. High
The airfoil could be dynamically pitched about its ¼ frequency details shown on the Endevco signal result
chord location through ±15° with mean angles ranging from high levels of turbulent flow about the airfoil
from ±0° to 15°. The airfoil was tested in the National surface. These small scales are impossible to detect due
Diagnostic Facility at IIT, and more details can be to the significant attenuation of the signal in the
found in Nagib et al. (1994) and Nagib and Hites (1994) pressure ports and associated tubing. [see Greenblatt et
The airfoil was equipped with 44 static al (2001) for more details.]
pressure ports around its perimeter. Each of the static Figs. 3a and 3b, respectively, show baseline quasi-
pressure port was connected with 0.063" vinyl tubing to static lift and moment coefficient data for M = 0.1 to
a 48 port J9 Scanivalve. A Pitot-static probe located in 0.4. The gradually increasing CL,max from M = 0.1 to 0.3
the free stream provided the reference pressure for each is primarily a low Reynolds number effect (Re =
pressure port, and each pressure port was sampled 700,000 to 2,100,000) due to the increasing free-stream
sequentially using a Validyne DP-103 pressure velocity. The humps in the CM data at approximately

3
American Institute of Aeronautics and Astronautics
13° are also low Reynolds number artifacts and Throughout this paper, Cµ and F+ are indicated
diminish with increasing Re. For M > 0.3 the low Re symbolically, not by their exact numeric values. This
effects are effectively eliminated as can be seen by the allows us to present the most recent data obtained from
higher lift curves and elimination of the CM hump. experiments and computations conducted specifically
Increasing Re and M effects appear to offset one for The Boeing Company. The values of the momentum
another as CL,max stays approximately constant between coefficient are indicated by Cµ(1)…Cµ(8), representing
M = 0.3 and 0.4. At M = 0.4, moment stall is much more increasing magnitude. Similarly, increasing values of
severe than that at lower M. the non-dimensional frequency are represented by
Dynamic lift and moment coefficient baseline data F+(1)…F+(5). This allows us to demonstrate the effects
are presented in Figs. 4a and 4b. Qualitatively, the data of each parameter individually, keeping other
are similar to the quasi-static data at low M. However, parameters unchanged.
at M > 0.3, a reduction in CL,max is observed from M = Figs. 6a and 6b show the effect of increasing M
0.3 to 0.4. This was originally presumed to be due to from 0.1 to 0.3 on lift and moment coefficients,
local supersonic flow in the leading-edge region of the respectively, while maintaining both forcing frequency
airfoil. It was discovered, however, that attenuation due and amplitude fixed in relative terms (i.e. constant F+
to wave propagation in the vinyl tubing was the main and Cµ). Note that forcing data is for the slot at x/c =
factor responsible for the lower values (see Greenblatt 0.5. For this light-stall scenario, it can clearly be seen
et al, 2001). Although this introduced an inherent that the increase in M has no deleterious effect
inaccuracy, it was not regarded as a problem, because whatsoever on the effectiveness of forcing. In fact, the
the main objective of this study was to assess the negative moment, relative to the baseline value,
influence of forcing over the baseline case. Thus the introduced at M = 0.1 under these forcing conditions is
factor considered important here was the difference somewhat alleviated. Therefore, while the effectiveness
between baseline and controlled scenarios, and not the of forcing certainly does not diminish, there is evidence
absolute values. to suggest that it may be superior at the higher Mach
number.
5.2 Control of Light Stall Data at a different F+ for increasing forcing
The effect of control on lift and moments at M = 0.1 amplitudes is presented in Figs. 7a and 7b. Once again,
for a single F+ and various increasing (with numerical this is a light stall scenario, but here all data is for M =
assignment) values of Cµ are presented in Figs. 5a and 0.3 and forcing is from the x/c = 0.3 slot. Note that
5b. The maximum incidence angle is 4° above the static CL,max increases continuously with increasing Cµ and
stall angle and thus this can be considered to a case of minimum (negative) CM is only marginally affected. In
so-called light stall. Note that the amplitude of fact, hysteresis associated with CM is significantly
oscillatory forcing is characterized by the momentum attenuated as forcing amplitudes increase. Furthermore,
coefficient: CM excursions (i.e. CM,max–CM,min), which are often used
as an indication of dynamic stall severity are slightly
2
h u  reduced. Thus, it can be concluded from the previous
Cµ = 2  
c  U ∞  two figures, that an increase in M from 0.1 to 0.3 has no
observable deleterious effect on the efficacy of dynamic
stall control in the light stall regime.
where h is the height of the blowing slot, c is the chord
of the control surface, and u is the rms velocity, and 5.3 Control of Deep Stall
F+ ≡ fXTE/U∞, where f is the excitation frequency and Figs. 8a and 8b show a similar comparison to that
XTE is the distance from the slot to the trailing-edge. For presented in Figs 7a and 7b, with the exception that the
the present case, the effect of control becomes more airfoil is moved far into the post-stall regime (with
pronounced with increasing forcing amplitude, maximum incidence angle 10° beyond the static stall
particularly increasing the maximum lift and minimum angle) and at a different forcing frequency. Here, both
moment, until a particular threshold is exceeded. Once lift and moment stall are severe. Introducing and
this threshold is exceeded, lift hysteresis virtually increasing the forcing amplitude has the effect of
disappears and the negative pitching moment is fully increasing CL,max as was shown previously. Here, CL,max
controlled. An examination of pressure distributions is increased in the vicinity of the static stall angle and at
shows that separation is partially controlled at lower the largest Cµ the improvement in CL,max over the
forcing amplitudes, increasing lift as well as the severity baseline dynamic case is approximately 35%. Note,
of stall. Beyond the above-mentioned threshold, however that the minimum moment associated with the
however, the flow remains fully attached throughout the baseline value is not materially affected by the forcing.
cycle, totally eliminating dynamic stall. In fact, it either remains unchanged, or is slightly

4
American Institute of Aeronautics and Astronautics
increased. Moreover, the CM excursions are consistently reduction effect is not large. This is mainly due to the
reduced with increasing Cµ . relatively low excitation amplitudes (Cµ) available as
the free-stream velocity increases.
5.4 Control at M ≥ 0.3 Additional details are revealed through examination
A significant challenge when applying control at of the chordwise pressure distributions. The pressure
high M, is delivering enough amplitude from the slot in distributions in Fig. 17a, show how forcing from the
order to supply sufficient control authority. In order to x/c = 0.3 slot renders the flow locally supersonic close
maintain constant Cµ the slot rms velocity must scale to the leading edge. The effects of flow control are most
linearly with the free stream velocity. The Cµ produced pronounced on the upper surface of the airfoil in the
in this investigation, at M > 0.3 was therefore very low. region between 0.1 < x/c < 0.4. Pressure gradients at
The effect of increasing M from 0.3 to 0.35 for both M = 0.4, shown in Fig. 17b, indicate that the flow is
baseline and identical excitation conditions is shown in supercritical in the leading-edge region, irrespective of
Figs. 9a and 9b. The overall effect on lift in both cases whether control is applied or not. While these figures
is to increase the stall angle by approximately 2°. show no distinct signature of a shock, this may be due to
Minimum moments by-and-large are not significantly the lack of adequate spacial resolution in the pressure
affected. Although control at M = 0.35 produced measurements in the leading-edge region. Also, in
slightly more negative moments, the overall excursions contrast to the NACA 0015 airfoil used by Carr and
are approximately the same. Increasing M from 0.35 to Chandrasekhara (1995 and 1996), the VR-7 airfoil is
0.4 (see Figs 10a and 10b), under the same forcing known to not exhibit a strong leading-edge shock at
conditions showed very little difference to CL,max or these Mach numbers; e.g. see McAlister et al. (1982).
moment excursions. Consideration of Figs. 9a,b and
10a,b illustrate a point that is important to the overall 6. Discussion of Numerical Results
objective of this paper, namely that the time histories of 6.1 Static Airfoil Results
both lift and moment coefficients are affected in the Fig. 18 illustrates comparisons between the
same way regardless of M in the range 0.3 to 0.4. Thus, measured and the predicted sectional lift characteristics
even though the small Cµ is incapable of exerting for the VR-7 airfoil as a function of post-stall angle of
significant control authority, the time histories indicate attack for M = 0.20 and Re = 1,430,000. Note that
that the effect of forcing is not diminished with baseline lift is over-predicted when assuming fully
increasing Mach number. turbulent flow. In this case, the forcing utilized Cµ(8)
and F+(1). Furthermore, user-specified “fixed” laminar-
5.5 Other Experimental Results turbulent transition points were enforced in the CFL3D
Figs. 11a,b and 12a,b show the effect of varying the simulations. For example, “tr:U05_L05” implies that
mean incidence angle at M = 0.3, while maintaining transition was enforced at 5% chord on the upper
incidence angle excursions of 5°, for baseline and surface of the airfoil and at 5% chord on the lower
controlled scenarios. These data, in essence, summarize surface (Fig. 18). Upper and lower surface transition
and confirm the effectiveness of this control technique points at the 5% and 15% chord, respectively, yield the
over a wide range of angles, as observed in the previous best correlation with the experimental data.
three sections, from light stall conditions well into deep
stall. 6.2 Dynamic Stall Results
The deep-stall data for M = 0.1, shown in Figs. 13a Fig. 19 compares data of McAlister et al (1982,
and 13b, demonstrate the effects of control at low Mach frame 47305) with the CFL3D-predicted lift, pressure
number where high values of Cµ can be achieved. drag and moment coefficients, for the baseline VR-7
These figures also show that although CL,max is airfoil under the conditions α = 14.8 ±49.88 degrees,
increased, this is accompanied by a moment penalty. k = 0.10, M = 0.281 and Re = 3,780,000. Initial
Thus, comparing these figures to Fig. 8, described transients in the computations were eliminated by
earlier, we see a definite improvement in the efficacy of performing at least four complete airfoil “cycles”. Fig.
control on CM at M = 0.3 in the deep-stall regime. 19 shows reasonable agreement of the computations
In the discussion to this point, only lift and moment with the data. Specifically, the lift curve slope is
coefficient results have been presented. It is also of adequately represented during the pitch-up motion, The
value to examine the effects of this control technique on growth and shedding of the dynamic stall vortex is
the airfoil drag. Examples of instantaneous wake qualitatively predicted at 23.5 degrees as compared to
profiles (Fig. 14), time-mean momentum deficit in the the experimental angle of 21 degrees. During the pitch-
wake (Fig 15), and computed drag coefficient (Fig 16) down motion, reasonable correlation with the measured
are shown for the light-stall case. In general, the drag- lift values is obtained. Similarly, the sectional pressure

5
American Institute of Aeronautics and Astronautics
drag values and pitching moments are well predicted 8. Conclusions
with the exception of the values near the maximum Pulsating zero-mass flux jets introduced from
angle of attack of 25 degrees. spanwise slots at various locations on the upper surface
Having established the reasonable accuracy of the of an oscillating VR-7 airfoil model are shown to be
solver for dynamic stall computations, the effect of effective in controlling lift, moment and drag
excitation was considered. Fig. 20 illustrates coefficients over the range of Mach numbers from 0.1
comparisons between the measured and the predicted to 0.4. This control is demonstrated over a wide range
lift coefficients for baseline controlled scenarios at M = of mean angles of attack of the oscillating airfoil from
0.10, Re = 714,000, k = 0.05 and α = 11 ±5degrees. It is light to deep stall conditions. Maintaining the non-
seen that invoking forcing with Cµ(7) and F+(2) results dimensional frequency and amplitude of the forcing
in a reduction in lift hysteresis, as a result of the lift unchanged results in comparable modifications of the
increase during the pitch-down part of the cycle. This aerodynamic coefficients throughout this Mach number
behavior is well captured by the modified CFL3D range.
solver. For the baseline airfoil, note that the predicted Therefore, contrary to the earlier speculations of
lift values are higher than those measured especially Carr and Chandrasekhara (1996), it is clear based on the
near the maximum angle of attack of 16 degrees. No present wind-tunnel tests that the pulsed-blowing active
effort was made to investigate the effects of laminar- flow technique can be effective for at least some
turbulent transition on the predicted lift values. – mainly oscillating airfoils operating within the compressible
due to the complex nature of emulating this phenomena regime. Their documented conclusions may only be
on an oscillating airfoil. valid for the particular airfoil used and the forcing
Fig. 21 illustrates the same comparison, but under conditions applied.
the conditions M = 0.3, Re = 2,140,000 million and k = Based on our results, it appears that this active-
0.05. As above, with the exception of the higher levels flow control technique is only limited by the ability to
of the sectional lift values, it is seen that the effect of generate the adequate forcing conditions at the higher
control, with Cµ(5) and F+(1), is to reduce the area of Mach numbers required for applications such as
the loop. Simultaneously, an increase in the lift values rotorcraft. Future research, design and development
during the pitch-up and pitch-down parts of the cycle is should focus on providing actuators that can deliver the
seen for all angles of attack larger than the mean angle pulsating momentum required at the higher Mach
of 11 degrees. Similar to the CFL3D results shown in numbers. Technologies leading to actuators that can be
Fig. 20 for the baseline airfoil, the predicted lift values integrated into the rotorcraft blade, currently under
for angles of attack larger than 13 degrees are also development, should be a high priority to the active-
higher than the measured ones. flow control filed.

7. Summary Acknowledgments
Fig. 22 presents a summary of the experimental The authors greatly appreciate the support of The
results presented in Section 5 by showing CL,max, Boeing Company and the Army VGART Program,
referenced to baseline values, for a wide range of monitored by Dr. Chee Tung. Also, many thanks are
forcing cases. The figure suggests two possible due to Mr. Craig Johnson and Mr. Chris Chistophorou
conclusions that can be drawn from the present work. of the MMAE Department at IIT for their valuable
Firstly, the effectiveness of control improves as the assistance.
airfoil enters further into deep stall. Secondly, it appears
that the efficacy of forcing diminished with increasing References
Mach number, as has been suggested by Carr and Carr, L. W., “Progress in the analysis and
Chandrasekhara (1995 and 1996). However, Fig. 23 prediction of dynamic stall” AIAA Journal of Aircraft,
clearly demonstrates that the main reason for the second Vol. 25, No. 1, 1988, pp. 6-17.
(erroneous) observation is the reduction in the relative Carr, L.W. and Chandrasekhara, M.S., "An
forcing amplitude, Cµ. Assessment of the Impact of Compressibility on
It is clear from Fig. 23 that for comparable values of Dynamic Stall," AIAA-95-0779, 33rd Aerospace
the forcing amplitude, the lift enhancement is Sciences Meeting, Reno, NV, 1995.
independent of Mach number over the range from 0.1 to Carr, L. W. and Chandrasekhara, M. S.,
0.4. This is true for both the light-to-moderate stall “Compressibility effects on dynamic stall”, Progress in
conditions, where the lift enhancement monotonically Aerospace Sciences, Vol. 32, pp. 523-573, 1996.
increases up to approximately 10% improvement, and Cordova, J. Q. and Barth, T. J., “Grid Generation
for the deep stall cases (19 deg. mean angle), where the for General 2-D Regions Using Hyperbolic Equations,”
lift enhancement can be by as much as 30%. AIAA paper 88-0520, Jan. 1988.

6
American Institute of Aeronautics and Astronautics
Gault, D.E., "A Correlation of Low-Speed, Airfoil- Accuracy of Time-Accurate Turbulent Navier-Stokes
Section Stalling Characteristics With Reynolds Number Computations,” AIAA paper 95-1835, June 1995.
and Airfoil Geometry," NACA TN 3963, March 1957. Seifert, A., Bachar, T., Koss, D., Shepshelovich,
Greenblatt, D. and Wygnanski, I. “Parameters M. and Wygnanski, I., "Oscillatory Blowing: A Tool to
affecting dynamic stall control by oscillatory Delay Boundary-Layer Separation," AIAA Journal, Vol.
excitation”, AIAA Paper 99-3121, 17th AIAA 31, No. 11, p. 2052, 1993.
Applied Aerodynamics Conference, Norfolk, VA, Seifert, A., Darabi, B. Nishri, B, and Wygnanski, I,
28 June – 1 July 1999. “The effects of forced oscillations on the performance
Greenblatt, D., Darabi, A., Nishri, B. and of airfoils”, AIAA 93-3264, 1993.
Wygnanski, I. “Some factors affecting stall control with Seifert, A. and Pack, L.G., “Oscillatory control of
particular emphasis on dynamic stall”, AIAA Paper 99- separations at high Reynolds numbers”, AIAA 98-0214,
3504, 30th AIAA Fluid Dynamics Conference, Norfolk, 1998.
VA, 28 June – 1 July, 1999. Seifert, A, Darabi, A., and Wygnanski, “Delay of
Greenblatt, D., Kiedaisch, J. and Nagib, H., airfoil stall by periodic excitation”, Journal of Aircraft,
“Unsteady-Pressure Corrections In Highly Attenuated Vol. 33, No. 4, 1996.
Measurements At High Mach Numbers”, AIAA-2001- Seifert, A. and Pack, L. G., “Oscillatory excitation
2983. of unsteady compressible flows over airfoils at flight
Hassan, A. and Munts E. “Transverse and Near Reynolds numbers” AIAA Paper 99-0925, 37th
Tangent Synthetic Jets for Aerodynamic Flow Control,” Aerospace Sciences Meeting and Exhibit, Reno, NV,
AIAA Paper 2000-4334, Applied Aerodynamic Jan. 11-14, 1999.
Conference, Denver, CO, 14 – 17 August 2000. Spalart, P. And Allmaras, S., “A One-Equation
Hites, M.H., Scaling of High-Reynolds-Number Turbulence Model for Aerodynamic Flows,” AIAA 92-
Turbulent Boundary Layers in the National Diagnostic 0439, 1992.
Facility, Ph.D. Thesis, Illinois Institute of Technology, Wygnanski, I. and Seifert, A., "The Control of
1997. Separation by Periodic Oscillations," AIAA-94-2608,
Hites, M., Nagib, H., Bacher, T., and Wygnanski, 18th AIAA Aerospace Ground Testing Conference,
I., “Enhanced Performance of Airfoils at Moderate Colorado Springs, CO, 1994.
Mach Numbers Using Zero-Mass Flux Pulsed Wygnanski, I, Boundary Layer and Flow Control
Blowing,” AIAA Paper 2001-0734, 39th Aerospace by Period Addition of Momentum (Invited)”, AIAA 97-
Sciences Meeting and Exhibit, Reno, NV, Jan. 12-15, 2117, 1997.
2001 Wu, J-Z., Lu, X-Y., and Wu, J-M., “Post-stall lift
Mazanec, M.J., Design of Research Equipment for enhancement on an airfoil by local unsteady control”,
Use in the National Diagnostic Facility, M.S. Thesis, AIAA 97-2064.
Illinois Institute of Technology, May 1996.
McAlister, K. W., Pucci, S. L., McCroskey, W. J.
and Carr, L. W. “An experimental study of dynamic
stall on advanced airfoil sections. Volume 2. Pressure
and force data” NASA TM 84245, 1982.
McManus, K. and Magill, J., “Airfoil performance
enhancement using pulsed jet separation control”,
AIAA-1971, 1997.
Nagib, H.M., and Hites, M.H., "Measurement of
Disturbance Levels in the National Diagnostic Facility",
AIAA 94-0770, 32nd Aerospace Sci. Mtg., Reno, NV,
1994.
Nagib, H., Hites, M., Gravante, S., and Won, J.,
"Flow Quality Documentation of the National
Diagnostic Facility," AIAA 94-2499, 18th AIAA
Aerospace Ground Testing Conference, 1994.
Peterson, B.A., Design of Experimental Equipment
for Use in the National Diagnostic Facility, M.S. Thesis,
Illinois Institute of Technology, December 1995.
Rumsey, C. L., Sanetrik, M. D., Biedron, R. T.,
Melson, N. D., and Parlette, E. B., “Efficiency and

7
American Institute of Aeronautics and Astronautics
2
VR7: Baseline, Quasi-Static,
1.8 α = 10o+10osin(ωt). (a)

1.6

1.4

1.2
Cl
1
Figure 1. Typical computational grid depicting the
required clustering near the 50% chord to capture the 0.8
details of the synthetic jet flow.
0.6 M = 0.1
M = 0.2
0.4 M = 0.3
-4000
Measured Pressure Signal M = 0.35
0.2
Surface-Mounted Endevco M = 0.4
-3000 20th harmonic correction & model
0
0 5 10 15 20 25
α (deg)
Oscillatory Pressure (Pa)

-2000
0
-1000 (b)
-0.02

0 -0.04

-0.06
1000
-0.08
Cm
2000 -0.1
VR7: Baseline, Quasi-Static,
-0.12 α = 10o+10osin(ωt).
3000
0 120 240 360 -0.14 M = 0.1
Cycle Angle (degrees) M = 0.2
Figure 2. Comparison of measured pressure signal, -0.16 M = 0.3
corrected signal, and the signal from the surface- M = 0.35
mounted transducer for one airfoil pitch-oscillation -0.18
M = 0.4
cycle for Ma = 0.3, k = 0.05, α=17°+5°sin( ω t).
-0.2
0 5 10 15 20 25
α (deg)
Figure 3. Quasi-static lift coefficient (a) and moment
coefficient (b) vs. angle of attack for five different
Mach numbers.

8
American Institute of Aeronautics and Astronautics
2 2
VR7: M=0.1, k=0.05
1.8
(a)
(a)
1.6 1.6

1.4
CL
1.2 1.2
Cl
1

0.8 0.8
VR7: k = 0.05, α=11o+5osin(ωt) Baseline: Quasi-Static
0.6 Baseline
F+(2), Cµ(1)
M = 0.1, Baseline
0.4 0.4 F+(2), Cµ(2)
M = 0.3, Baseline
F+(2), Cµ(4)
M = 0.35, Baseline F+(2), Cµ(6)
0.2
M = 0.4, Baseline F+(2), Cµ(7)
0 0
0 5 10 15 20 25 0 5 10 15 20 25
α (deg) α (degrees)
0 0

-0.02 (b) (b)

-0.04 -0.04

-0.06 CM

-0.08 -0.08
Cm
-0.1 VR7: M=0.1, k=0.05

-0.12 -0.12
Baseline: Quasi-Static
o o
VR7: k = 0.05, α=11 +5 sin(ωt) Baseline
-0.14 F+(2), Cµ(1)
M = 0.1, Baseline
F+(2), Cµ(2)
-0.16 M = 0.3, Baseline -0.16
F+(2), Cµ(4)
M = 0.35, Baseline F+(2), Cµ(6)
-0.18 F+(2), Cµ(7)
M = 0.4, Baseline

-0.2 -0.2
0 5 10 15 20 25 0 5 10 15 20 25
α (degrees)
α (deg)
Figure 5. Unsteady lift (a) and moment (b) coefficients
Figure 4. Unsteady lift (a) and moment (b) coefficients
vs. angle of attack for five excitation amplitudes at low
vs. angle of attack for four Mach numbers; baseline
Mach number compared to the baseline unsteady and
cases (no excitation). (Note: solid lines are upstroke
and dashed lines are downstroke.) quasi-static cases, α=11°+5°sin( ω t). (Forcing slot at
x/c = 0.5.)

9
American Institute of Aeronautics and Astronautics
2 2
VR7: Effect of "M", k=0.05
Excitation from 50% slot (a) 1.8 (a)
1.6 1.6

1.4
CL
1.2 1.2
Cl
1 VR7: M=0.3, k = 0.05,
α=11o+5osin(ωt)
0.8 0.8 Forcing slot at 30% chord.

Quasi-Static: M=0.1 Baseline


0.6
Quasi-Static: M=0.3 F+(3), Cµ(1)
Baseline: M=0.1 0.4 F+(3), Cµ(2)
0.4
Baseline: M=0.3 F+(3), Cµ(3)
F+(2), Cµ(4): M=0.1 0.2 F+(3), Cµ(5)
F+(2), Cµ(4): M=0.3 Baseline, Quasi-Static
0 0
0 5 10 15 20 25 0 5 10 15 20 25
α (degrees) α (deg)
0.04 0
VR7: Effect of "M", k=0.05
Excitation from 50% slot (b) -0.02 (b)
0
-0.04
CM
-0.06
-0.04
-0.08
Cm
-0.08 -0.1
VR7: M=0.3, k = 0.05,
α=11o+5osin(ωt)
-0.12
Forcing slot at 30% chord.
-0.12 Quasi-Static: M=0.1
-0.14 Baseline
Quasi-Static: M=0.3
F+(3), Cµ(1)
Baseline: M=0.1
-0.16 F+(3), Cµ(2)
-0.16 Baseline: M=0.3
F+(3), Cµ(3)
F+(2), Cµ(4): M=0.1
-0.18 F+(3), Cµ(5)
F+(2), Cµ(4): M=0.3 Baseline, Quasi-Static
-0.2 -0.2
0 5 10 15 20 25 0 5 10 15 20 25
α (degrees) α (deg)
Figure 6. Unsteady lift (a) and moment (b) coefficients Figure 7. Unsteady lift (a) and moment (b) coefficients
vs. angle of attack, baseline and with excitation, for vs. angle of attack for different excitation amplitudes
two Mach numbers, α=11°+5°sin( ω t). Quasi-static compared to baseline unsteady and quasi-static cases.
cases also shown.

10
American Institute of Aeronautics and Astronautics
2 2
VR7: M=0.3, k = 0.05,
1.8 α=19o+5osin(ωt) (a) 1.8
(a)
Forcing slot at 30%
1.6 chord. 1.6

1.4 1.4

1.2 1.2
Cl Cl
1 1

0.8 0.8 VR7: k = 0.05, α=11o+5osin(ωt)


Forcing slot at 30% chord.
0.6 Baseline 0.6
F+(1), Cµ(1) M = 0.3 Baseline
0.4 0.4
F+(1), Cµ(3) M = 0.3, F+(1), Cµ(1)

0.2 F+(1), Cµ(5) M = 0.35 Baseline


0.2
Baseline, Quasi-Static M = 0.35, F+(1), Cµ(1)
0 0
0 5 10 15 20 25 0 5 10 15 20 25
α (deg) α (deg)
0 0

-0.02 (b) -0.02 (b)

-0.04 -0.04

-0.06 -0.06

-0.08 -0.08
Cm Cm
-0.1 -0.1
VR7: M=0.3, k = 0.05,
-0.12 -0.12 VR7: k = 0.05, α=11o+5osin(ωt)
α=19o+5osin(ωt)
Forcing slot at 30% chord. Forcing slot at 30% chord.
-0.14 -0.14
Baseline M = 0.3 Baseline
-0.16 F+(1), Cµ(1) -0.16 M = 0.3, F+(1), Cµ(1)
F+(1), Cµ(3)
M = 0.35 Baseline
-0.18 F+(1), Cµ(5) -0.18
Baseline, Quasi-Static M = 0.35, F+(1), Cµ(1)
-0.2 -0.2
0 5 10 15 20 25 0 5 10 15 20 25
α (deg) α (deg)
Figure 8. Unsteady lift (a) and moment (b) coefficients Figure 9. Unsteady lift (a) and moment (b) coefficients
vs. angle of attack for different excitation amplitudes vs. angle of attack at Ma = 0.3 and Ma = 0.35; cases
compared to baseline unsteady and quasi-static case; with equivalent excitation momentum coefficients
deep stall cases. compared to baseline cases.

11
American Institute of Aeronautics and Astronautics
2 2
VR7: M=0.3, k = 0.05, α=αm+5osin(ωt)
1.8
(a) 1.8 (a)
1.6 1.6

1.4 1.4

1.2 1.2
Cl Cl
1 1

0.8 VR7: k = 0.05, α=11o+5osin(ωt) 0.8


Forcing slot at 30% chord.
0.6 0.6 Baseline, 11 deg. mean
M = 0.35 Baseline Baseline, 13 deg. mean
0.4 M = 0.35, F+(1), Cµ(1) 0.4 Baseline, 15 deg. mean
Baseline, 17 deg. mean
M = 0.4 Baseline
0.2 0.2 Baseline, 19 deg. mean
M = 0.4, F+(1), Cµ(1) Quasi-Static Baseline
0 0
0 5 10 15 20 25 0 5 10 15 20 25
α (deg) α (deg)
0 2
VR7: M=0.3, k = 0.05,
-0.02 (b) 1.8 α=αm+5osin(ωt) (b)
Forcing slot at 30%
-0.04 1.6 chord.

-0.06 1.4

-0.08 1.2
Cm Cl
-0.1 1

-0.12 VR7: k = 0.05, α=11o+5osin(ωt) 0.8


Forcing slot at 30% chord.
-0.14 0.6 F+(3), Cµ(5), 11 deg. mean
M = 0.35 Baseline F+(3), Cµ(5), 13 deg. mean
-0.16 M = 0.35, F+(1), Cµ(1) 0.4 F+(3), Cµ(5), 15 deg. mean
M = 0.4 Baseline F+(3), Cµ(5), 17 deg. mean
-0.18 0.2 F+(3), Cµ(5), 19 deg. mean
M = 0.4, F+(1), Cµ(1)
Quasi-Static Baseline
-0.2 0
0 5 10 15 20 25 0 5 10 15 20 25
α (deg) α (deg)
Figure 10. Unsteady lift (a) and moment (b) Figure 11. Unsteady lift coefficients for baseline (a)
coefficients vs. angle of attack at Ma = 0.35 and Ma = and excitation (b) cases for airfoil pitch oscillations
0.4; cases with equivalent excitation momentum about five different mean angles.
coefficients compared to baseline cases.

12
American Institute of Aeronautics and Astronautics
0 2.4
VR7: M=0.1, k = 0.05,
-0.02
(a) 2.2 α=17o+5osin(ωt)
Forcing slot at 30%
(a)
2 chord.
-0.04
1.8
-0.06
1.6
-0.08
1.4
Cm Cl
-0.1 1.2

-0.12 VR7: M=0.3, k = 0.05, 1


α=αm+5osin(ωt)
-0.14 0.8
Baseline, 11 deg. mean
Baseline, 13 deg. mean 0.6
-0.16 Baseline
Baseline, 15 deg. mean
Baseline, 17 deg. mean
0.4 F+(4), Cµ(8)
-0.18 Baseline, 19 deg. mean F+(5), Cµ(8)
0.2
Quasi-Static Baseline Baseline, Quasi-Static
-0.2
0
0 5 10 15 20 25
0 5 10 15 20 25
α (deg) α (deg)
0 0

-0.02 (b) -0.02 (b)


-0.04
-0.04
-0.06
-0.06
-0.08
-0.08
-0.1
Cm Cm
-0.1 -0.12
VR7: M=0.3, k = 0.05,
α=αm+5osin(ωt) -0.14
-0.12 VR7: M=0.1, k = 0.05,
Forcing slot at 30%
chord. α=17o+5osin(ωt)
-0.14 -0.16
Forcing slot at 30% chord.
F+(3), Cµ(5), 11 deg. mean -0.18
-0.16 F+(3), Cµ(5), 13 deg. mean Baseline
F+(3), Cµ(5), 15 deg. mean
-0.2 F+(4), Cµ(8)
F+(3), Cµ(5), 17 deg. mean
-0.18 F+(3), Cµ(5), 19 deg. mean F+(5), Cµ(8)
Quasi-Static Baseline
-0.22
Baseline, Quasi-Static
-0.2
-0.24
0 5 10 15 20 25
0 5 10 15 20 25
α (deg) α (deg)
Figure 12. Unsteady moment coefficients for baseline Figure 13. Unsteady lift (a) and moment (b)
(a) and excitation (b) cases for airfoil pitch oscillations coefficients vs. angle of attack for deep stall, low Mach
about five different mean angles. number, and high excitation amplitude. (Baseline and
quasi-static cases shown for comparison.)

13
American Institute of Aeronautics and Astronautics
1.25 0.16
Baseline: Cd=0.0729
0.14 F+(1), Cµ(5): Cd=0.0719
F+(3), Cµ(5): Cd=0.0704
0.12 VR7: M=0.3, k = 0.05, α=11o+5osin(ωt)

u/U(1-u/U) [time-mean]
Forcing slot at 30% chord.
1
0.1

U/Uh 0.08

0.06
0.75 VR7: M = 0.3,
k = 0.05,
α = 15o+5osin(ωt) 0.04
Baseline bottom top
Forcing slot at 30%
chord. F+(1), Cµ(5) side side
0.02
α = 12o: pitch-down F+(3), Cµ(5)

0.5 0
-1.2 -0.8 -0.4 0 0.4 0.8 1.2 -1 -0.5 0 0.5 1
y/c y/c
Figure 14. Normalized wake velocity profiles for Figure 15. Integrated time-mean momentum deficit
baseline and two excitation cases at a single angle of plotted across the wake for baseline and two excitation
attack during the downstroke. cases; total drag values for each case shown in legend.
0.5
Baseline
0.45
F+(1), Cµ(5)
0.4 F+(3), Cµ(5)

0.35 VR7: M=0.3, k = 0.05, α=11o+5osin(ωt)


Forcing slot at 30% chord.
0.3

0.25
Cd
0.2

0.15

0.1

0.05

-0.05
12 13 14 15 16 17
α (deg)
Figure 16. Instantaneous drag coefficient vs. angle of attack for baseline and two excitation cases.

14
American Institute of Aeronautics and Astronautics
-8 -8
(a) Baseline
(b)
-7 -7 F+(2), Cµ(1), slot x/c = 0.3
CP*
F+(2), Cµ(1), slot x/c = 0.5
-6 Baseline -6
F+(3), Cµ(5), slot x/c = 0.3 VR7: M = 0.4, k = 0.05,
-5 F+(2), Cµ(5), slot x/c = 0.5 -5 α = 11o+5osin(ωt)
CP CP α = 15o: pitch-up
VR7: M = 0.3, k = 0.05,
-4 -4
α = 11o+5osin(ωt)
α = 15o: pitch-up CP*
-3 -3

-2 -2

-1 -1

0 0

1 1
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x/c x/c
Figure 17. Chordwise pressure coefficient distributions for baseline and two excitation cases at a single angle of
attack during the upstroke for two Mach numbers; Ma = 0.3 (a) and Ma = 0.4 (b). Critical pressure coefficient values
and excitation slot locations (vertical dashed lines) are also shown.

VR-7 Airfoil: M inf= 0.20, R e= 1.43 m illion


2.00

1.80

1.60
Lift Coefficient, CL

1.40

1.20
ba s e line (IIT - e x pt.)
w ith c ontr ol (IIT - e x pt.)
1.00 fully turbule nt
tr: U0 5 _ L0 5
tr: U0 5 _ L1 5
tr: U1 0 _ L1 0
0.80 tr: U2 0 _ L2 0
w ith c ontr ol CFD: U0 5 -L1 5

0.60
9 10 11 12 13 14 15 16 17 18
An g le o f Atta c k (d e g s .)

Figure 18. Predicted and measured sectional lift coefficients for the VR-7 airfoil in the
absence and the presence of synthetic jet flow control: effects of laminar-turbulent transition.

15
American Institute of Aeronautics and Astronautics
2.50

2.00

Lift Coefficient, CL
1.50

1.00

0.50 experimental
predicted

0.00
0.0 5.0 10.0 15.0 20.0 25.0 30.0
Angle of Attack (degrees)
1.20

1.00 experimental
Drag Coefficient, CD

predicted
0.80

0.60

0.40

0.20

0.00

-0.20
0.0 5.0 10.0 15.0 20.0 25.0 30.0
Angle of Attack (degrees)

0.10

0.00
Pitching Moment Coefficient, Cm

-0.10

-0.20

-0.30

-0.40
experimental
-0.50 predicted

-0.60

-0.70
0.0 5.0 10.0 15.0 20.0 25.0 30.0
Angle of Attack (degrees)

Figure 19. Predicted and measured lift, pressure drag, and moment coefficients for the
baseline VR-7 airfoil during pitch oscillation.

16
American Institute of Aeronautics and Astronautics
V R -7 Airfo il: M in f=0 .1 0 , R e =7 1 4 ,0 0 0 , k =0 .0 5
1 .8 0
Ex p t. (b a se lin e )
1 .7 0 Ex p t. (w ith co n tro l)
C F D (b a se lin e )
1 .6 0
C F D (w ith co n tro l)

1 .5 0
Lift Coefficient, CL

1 .4 0

1 .3 0

1 .2 0

1 .1 0

1 .0 0

0 .9 0

0 .8 0
6 8 10 12 14 16
A n g le o f Atta c k (d e g s .)

Figure 20. Predicted and measured unsteady lift characteristics for the VR-7 airfoil during
pitch oscillation and a free stream Mach number of 0.10. Also shown are the measured and the
predicted values for the controlled airfoil.

2.1
b a se lin e (e x p t.)
1.9 w ith co n tro l (e x p t.)
b a se lin e (p re d icte d )
1.7 w ith co n tro l (p re d icte d )
Lift Coefficient, CL

1.5

1.3

1.1

0.9

0.7

0.5
5 7 9 11 13 15 17

A n g le o f A tta c k (d e g re e s )
Figure 21. Predicted and measured unsteady lift characteristics for the VR-7 airfoil during
pitch oscillation and a free stream Mach number of 0.30. Also shown are the measured and the
predicted values for the controlled airfoil.

17
American Institute of Aeronautics and Astronautics
1.4

1.3
Cl max / Cl max baseline

1.2

1.1

0.9
k = 0.05, 9+/-5 deg. k = 0.05, 10+/-5 deg.
k = 0.05, 11+/-5 deg. k = 0.05, 13+/-5 deg.
0.8
k = 0.05, 15+/-5 deg. k = 0.05, 17+/-5 deg.
k = 0.05, 19+/-5 deg. k = 0.1, 11+/-5 deg.
0.7
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45
Mach No.
Figure 22. Effects of forcing on maximum lift for various Mach numbers and mean angles of attack; data includes
the full range of momentum coefficients tested.

1.4

1.3
Cl max / Cl max baseline

1.2

1.1

1
M=0.1, 30% slot, k=0.05 M=0.1, 50% slot, k=0.05
0.9 M=0.3, 30% slot, k=0.05 M=0.3, 50% slot, k=0.05
M=0.35, 30% slot, k=0.05 M=0.35, 50% slot, k=0.05
M=0.4, 30% slot, k=0.05 M=0.4, 50% slot, k=0.05
0.8
M=0.1, 50% slot, k=0.1 M=0.3, 30% slot, k=0.1
M=0.3, 50% slot, k=0.1 19+/-5 deg. cases
0.7
Cmu (% )

Figure 23. Effects of forcing on maximum lift for various Mach numbers vs. momentum coefficient demonstrating
Mach-number independence of this type of active flow control; data includes the full range of mean angles of attack
tested with the maximum (19°) separately identified.

18
American Institute of Aeronautics and Astronautics

You might also like