Sie sind auf Seite 1von 19

Toward Understanding the Aerodynamic Efficiency of a Hovering Micro-Rotor

Manikandan Ramasamy∗ Bradley Johnson† J. Gordon Leishman‡

Department of Aerospace Engineering


Glenn L. Martin Institute of Technology
University of Maryland, College Park, MD 20742

Abstract Cl lift coefficient


CT rotor thrust coefficient, = T /ρAΩ2 R2
The velocity field of a micro-rotor was measured using CP rotor power coefficient, = P/ρAΩ3 R3
digital particle image velocimetry (DPIV) to help iden- D rotor blade sectional drag/unit span
tify the pertinent flow features and the various sources of DL disk loading, = T /A
potential aerodynamic losses. These flow measurements FM figure of merit
were complemented by balance measurements of rotor L rotor blade sectional lift/unit span
thrust and power. Several rotor blades were used with dif- PL power loading, = T /P
ferent planforms, twist distributions, and airfoil profiles. r radial distance
In each case, the operating blade tip Reynolds number r0 initial core radius of the tip vortex
was nominally 35,000. A unique aspect of the work was rc core radius of the tip vortex
the use of advanced DPIV correlation algorithms, which R radius of the blade
allowed for high-resolution flow measurements. This al- Rev vortex Reynolds number, = Γv /ν
lowed for the accurate estimation of the spanwise lift and t time
drag along the blade from sectional circulation measure- S perimeter of the loop integral
ments and wake velocity deficiency characteristics, re- T rotor thrust
spectively. Sharpening the leading- and trailing-edges of Vr radial velocity
the blade was shown to reduce the wake thickness and pro- Vθ swirl velocity
file losses, resulting in improved rotor performance. The Vz axial velocity
twisted blades also showed improved levels of rotor per- u, v, w velocities along x, y, z directions, respectively
formance. The evolutionary characteristics of the tip vor- x, y, z rotor coordinate system
tices, including the growth rate of the vortex cores and α Lamb’s constant, = 1.25643
the peak swirl velocities, were also measured. The vortex Γ circulation, = 2πrVθ
contained a circulation that was found to be about 80% Γb bound circulation
of the maximum sectional bound circulation, and showed Γc circulation at the core radius
slow dissipative characteristics with increasing wake age. Γv circulation at large distances
The circulation contained within the viscous core was δ ratio of apparent to actual kinematic viscosity
about 50% of the total circulation at early wake ages, and ζ wake age
continued to increase with wake age. ν kinematic viscosity
ρ air density
σ rotor solidity
Nomenclature ψ azimuthal position of blade
Ω rotational speed of the rotor
A rotor disk area 2-C two-component
c blade chord 3-C three-component
Cd drag coefficient 2-D two-dimensional
Cd0 minimum (profile) drag coefficient 3-D three-dimensional
∗ Assistant
Research Scientist. mani@glue.umd.edu
† MintaMartin Intern. bjo212@glue.umd.edu
‡ Minta Martin Professor. leishman@umd.edu Introduction
Presented at the American Helicopter Society International
Specialists Meeting on Unmanned Rotorcraft, January 23–25, While many types of manned aircraft play a role in the
2007, Phoenix, AZ. 2007
c by M. Ramasamy et al. Published military mission, military planners foresee a day in the
by the AHS International with permission. near future when swarms of micro air vehicles (or MAVs)

1
can be used to augment combat effectiveness and reduce with different rotor blade planforms. This includes ap-
casualties. A MAV has been defined variously as an air- proaches towards understanding methods of reducing the
craft having a maximum dimension of 15 cm (≈6 in) to as profile losses (by varying the airfoil shape, nose sharp-
much as 40 cm (≈16 in), with the requirement that they ness, and blade thickness distribution) and induced losses
are small and light enough so that they can be backpacked (by varying blade planform and twist shapes). The ob-
by single soldiers into the battlefield. MAVs have the ob- jective was to help further understand the flow field char-
vious advantage of small scale and correspondingly low acteristics of micro-rotors that, in turn, may help in the
radar cross-section and stealthiness. They can give good future validate mathematical models, and to develop de-
reconnaissance capability with rapid deployment and, in sign strategies to improve upon their performance. Simul-
many cases, real-time data acquisition capability such as taneous measurement of rotor performance and flow field
imaging. It is expected that in the future MAVs will take analysis is required to more clearly understand their aero-
on larger roles, and will support missions that are cur- dynamics. Also, it is necessary to estimate the sectional
rently performed by manned aircraft. performance across the blades, which is critical to develop
MAVs are usually required to hover, and previous strategies aimed at improving the aerodynamic efficiency
MAV designs have encompassed rotating-wings of vari- of micro-rotors. It is not yet clear whether conventional
ous types, including ducted and unducted rotor concepts. rotor design strategies can be used effectively to this end.
But existing designs fall well short of the desired pay-
load and endurance objectives because they have not yet
shown sufficiently good aerodynamic efficiency. Several Sources of Rotor Losses
experiments have clearly shown that their aerodynamic ef-
ficiency is substantially lower than their higher Reynolds Increasing the flight performance of a rotating-wing MAV
number rotor counterparts (Refs. 1–3). One main issue (i.e., their speed, range, endurance, etc.) directly depends
affecting efficiency is that rotating-wing MAVs operate at on improving their aerodynamic efficiency. One measure
blade chord Reynolds numbers that may be three orders of hovering performance is the figure of merit (FM ). Fig-
of magnitude lower than full-scale helicopters. This pro- ure 1 shows current performance levels of micro-rotors in
duces thick boundary layers on the blades, which results terms of their measured FM as well as the intended de-
in large values of profile drag. The resulting poor lift-to- sign efficiency goal of 0.8 or possibly higher. Figure 1
drag ratios of the blade sections (which may be 5 or lower) shows that for a given level of induced losses progres-
substantially decrease the overall operating efficiency of sively smaller improvements in the FM can be achieved
the rotor system. In concert with this effect, is the rela- by improving upon the efficiency of the airfoils through
3/2
tively higher induced power requirements of MAV-scale higher values of Cl /Cd (i.e., the lift-to drag) after a cer-
rotors, as well as higher rotational and turbulent losses in tain FM threshold is reached. Also, it shows that unless
3/2
the wakes downstream of the rotating blades. the the values of Cl /Cd can exceed at least 10, which
While a number of approaches have been examined to appears high for low Reynolds number airfoils (Ref. 5),
try to improve rotor performance (Refs. 2–5), only mod- there is no possibility of increasing FM much above 0.6.
est gains have been attained. One key to improved per- Induced and rotational losses must be reduced in unison
formance obviously lies within understanding the viscous with profile losses to produce any real improvement in
dominated wake structure and the need to reduce induced overall hovering efficiency. However, at low Reynolds
losses not just profile losses. In fact, there is reason to numbers profile and induced losses have likely interde-
believe that induced (wake) and profile (blade airfoils) ef- pendent influences, and improving one over the other is
fects are strongly interdependent at this small scale, and no guarantee that improved rotor efficiency levels will, in
that induced and viscous effects cannot be isolated and fact, be obtained.
examined separately to the same degree as they can on Even though FM can give one measure the relative hov-
full-scale rotor systems. For example, we should not for- ering performance of MAVs, it is the power loading that
get that the nature of the MAV rotor flow probably violates defines the overall hovering performance in terms of thrust
most, if not all, of the assumptions made in the develop- produced for the energy expended. The power loading is
ment of the fundamental theories of rotor analyses. What defined as
we lack for now are methods that allow for small-scale T CT
PL = = (1)
rotor design, and until such methods become established, P CP ΩR
MAVs will continue to demonstrate much lower levels of Rewiting Eq. 1 in terms of FM using simple momentum
performance capability than is desired. theory considerations gives
To this end, the objective of the present study was

to perform several performance (i.e., thrust and power) 2ρ FM
and high-fidelity flow field measurements of a micro-rotor PL = √ (2)
DL

2
Figure 1: Rotor FM variations and goals in terms in-
duced losses and blade section aerodynamic efficiency. Figure 3: Relationship between rotor power loading
and FM versus blade loading coefficient.

its normal working state. Rewriting Eq. 1 gives


CT
PL =
ΩR(CP0 +CPi +CPo )
CT
= 3/2
! (3)
C σCd0
ΩR κ √T + +CPo
2 8

In the first instance, it is convenient to group the other in-


duced and rotational losses into an average induced power
factor κ. Differentiating Eq. 3 with respect to CT and
equating it to zero results in a maximum PL at a FM of
2/(3κ). But κ is always greater than one because there
Figure 2: Power loading versus effective disk loading.
are always non-ideal induced losses, and so the maximum
power loading for the rotor system will be achieved when
It is apparent from Eq. 2 that the power loading improves when the FM is smaller than 2/3, as shown in Fig. 3. In
with an increase in FM and a reduction in disk loading. this particular case, κ is assumed to be 1.8, which is con-
Figure 2 shows the values of power loading versus equiv- sistent with the measured levels of performance (as given
alent disk loading for different scale hover-capable air- by the data points) for a micro-rotor. Because other losses
craft and biological mechanisms, which also operate over are involved, this is a value that is considerably higher
a wide range of Reynolds numbers based on blade or wing than those typically found with rotors operating at higher
chord. Notice the logarithmic scale on both axes. The tip chord Reynolds numbers.
fundamental momentum theory given by Eq. 2 seems to It will be seen from Fig. 3 that the required blade load-
hold good at all scales, including insects and humming- ing coefficient at the most efficient or “optimum” hover-
birds, as it should bearing in mind it is simply a state- ing condition is extremely low, and this requires a large
ment of the theoretical ideal limit as given by Newton’s rotor disk area. Because of overall maximum allowable
2nd Law. Clearly the key to improved hovering efficiency rotor size and weight constraints for a MAV, the rotor must
(best power loading) is to operate at a low disk loading generally operate at a higher disk loadings and CT values
and at as high a figure of merit as possible at that disk than its optimum value. However, Fig. 3 also shows that
loading. the power loading curve is relatively flat after the best op-
It is convenient to idealize total rotor losses into the erating point is reached, suggesting that the performance
constituents of induced (i), profile (0), and “other” (o) losses associated with operating at slightly higher disk
losses losses. The other losses have their source in the loadings are modest. Nevertheless, the design point for
rotational and turbulent flow inside the rotor wake, and the most efficient rotor operation is clear.
are difficult to quantify a priori because these are inter- Even though Fig. 1 assumes that effect of the profile and
dependent functions of profile and induced losses. On a induced losses on the FM of the rotor system operate as in-
rotor operating at full-scale Reynolds numbers, rotational dependent parameters, these two effects actually have in-
and turbulent losses are low when the rotor is operating in terdependent influences, especially at low blade Reynolds

3
Tip vortex
ζ = 10 deg

Tip vortex
ζ = 190 deg
Tip
Turbulent vortices
vortex sheet

Rotor shaft

Figure 4: Representative laser light sheet flow visualization image in the flow field of a rotating-wing MAV.

Rotor blade

Tip
vortices
Merging
boundary
layers
(vortex sheet)

Figure 5: A close up view of the flow field immediately behind the blade revealing the presence of thick turbulent
wake sheet with strong streamwise vorticity.

numbers. For example, κ will vary with CT , and rotational


and turbulent losses become a function of blade section
characteristics (profile losses).
There is evidence to support this hypothesis. Figure 4
shows a laser sheet flow visualization image of the flow
field of a hovering micro-rotor (Ref. 1). Several flow fea-
tures can be identified. One distinctive feature in the wake
of a micro-rotor is that the thicker boundary layers from
the top and bottom surfaces of the blades merge to result
in a relatively thick, turbulent wake sheets.
This can be more clearly seen in Fig. 5, which is the
image of the wake immediately behind the rotor blade.
Clearly organized streamwise vorticity (wake sheets) that Figure 6: Variation of two-dimensional profile drag co-
result from the merging of the thick boundary layers from efficient versus chord Reynolds number.
top and bottom surface of the blade can be noticed. These
wake sheets are a source of rotational and turbulent losses. Experimental Setup
It is known that an increase in the boundary layer thick-
ness increases profile drag at the lower chord Reynolds The current experiments involved the following: 1. Ro-
numbers below 50,000—see Fig. 6. Notice from Fig. 4 tor thrust and torque measurements using a mass balance
that at the outer edges of the sheets are rolled-up tip vor- and a torque sensor; 2. Two-component (2-C) DPIV mea-
tices, which seem to have relatively large viscous cores surements for estimating the sectional lift and drag com-
compared to the rotor dimensions. ponents along the blade span; 3. Three-component (3-C)

4
DPIV for measuring the overall rotor flow field.
Eight blade shapes were used in the study. All of the
blades were made of composite carbon fiber, and had cam-
bered, circular arc airfoil sections. The blades included:
1. Baseline rectangular blade, as shown in Fig. 7a.
2. Baseline blade with sharpened leading edge (SLE).
3. Baseline blade with sharpened leading edge and
trailing-edge (SLT).
4. Tapered blade (Type-1), as shown in Fig. 7b.
5. Tapered blade (Type-2), as shown in Fig. 7c.
6. Inverse tapered blade, as shown in Fig. 7d.
7. −9◦ linearly twisted blade.
8. −24◦ linearly twisted blade.
The rotor had a radius of 86 mm and was operated in a
two-bladed configuration. The baseline blade had a uni-
form chord of 19 mm, giving a blade aspect ratio of 3.7
and it had no twist or taper. The general dimensions of the
other blades are shown in Fig. 7. Each rotor system had
an effective thrust weighted solidity, σe , of 0.14 as given
by
  
σt − σr 1 3
σe = (σr − σt )y1 3 + − + y1 3 − y1 4 + σt
1 − y1 4 4
(4)
where the suffixes r and t represent the values at the root
and tip of the blade, respectively, and y1 represents the
point at which the taper starts. For the sharpened leading- Figure 7: Schematics showing the different rotor blade
and trailing-edge blades, the blades were sharpened from configuration used for the measurements.
the 15% and 85% chord locations by linearly reducing
their thickness towards the leading- and trailing-edges, re-
spectively.

made at rotational frequencies of 20, 30, 40, 50 and 60


Thrust and Torque Measurements
Hz, and at collective pitch angles (measured at 75% ra-
The rotor stand was rigidly attached to a linear reaction dius) of 8◦ , 12◦ , 16◦ , and 20◦ . At each rotational speed,
torque sensor, which in turn was mounted on a micro mass torque and thrust data were collected over a 5 second pe-
balance capable of measuring the rotor thrust at a pre- riod. The final data was ensemble averaged and used to
cision of ±0.1 grams—see Fig. 8. The torque cell was find the the coefficients of thrust and power and the FM
calibrated by using weights applied at a precisely known of each rotor at the different pitch angles and blade tip
arm. The analog signal from the torque sensor was ap- Reynolds numbers. The rotor was also tested at all con-
propriately signal-conditioned and read into a digital data ditions without the blades attached to the hub to deter-
acquisition system. The rotational speed of the rotor was mine tares. These tares were subtracted from the thrust
measured with a Hall-effect sensor mounted adjacent to and torque measurements taken with the blades attached.
the rotor shaft. To guarantee measurement precision and repeatability of
A digital inclinometer was used to precisely set the the experiment, multiple tests were made on each rotor at
pitch angle of the blades at 75% span location. The ro- the same pitch angles and rotational speeds. The measured
tor system was positioned to thrust vertically downwards thrust was found to be within the instrument precision of
(wake directed upward) to prevent ground effect interfer- the mass balance, while the measured torque had an un-
ence. The measurement of rotor thrust and torque were certainty of ±3% of the mean value.

5
Y

r X

ND-Yag laser sheet

DPIV camera

(a) 2-C DPIV


ND-YAG Laser sheet

Figure 8: Schematic showing the experimental setup


used in measuring the performance of the micro-rotor.

Flow Field Measurements DPIV Cameras

The flow field analysis using the 2-C DPIV was limited
to five blade configurations: (1) The baseline blade; (2)
(b) 3-C DPIV
The baseline blade with sharp leading- and trailing-edges;
(3) The tapered blade (Type-1); (4) The inverse tapered Figure 9: Experimental set for 2-component and 3-
blade; (5) The −9◦ linearly twisted blade. The evolution- component PIV measurements.
ary characteristics of the wake and tip vortices were mea-
sured using 3-C DPIV only for the baseline case, although grabber, and DPIV analysis software. For the 3-C mea-
other measurements are in progress. For all the flow field surements, two digital CCD cameras with 2 mega-pixel
analyses, the rotor was operated at 50 Hz, with a tip speed resolution were placed at an angle to satisfy the Scheim-
of 27 m/s. The nominal operating tip Mach number and plfug criteria—see Fig. 9b. The lasers could be fired at
Reynolds number based on chord were 0.082 and 35,000, any blade phase angle around the azimuth, enabling flow
respectively. field measurements to be made at any required wake age.
The flow at the rotor was seeded with a thermally pro- The two lasers were fired with a pulse separation time of
duced mineral oil fog. The average size of the seed parti- 2µs; this corresponded to less than 0.02◦ of blade motion.
cles were between 0.20 to 0.22 microns in diameter. For
the PIV experiments, the entire test area was uniformly Image Processing
seeded before each sequence of measurements. For the
flow visualization, judicious adjustment of the seeder was For the 2-C DPIV measurements, a simple adaptive grid
required to introduce concentrations of fog at the locations was used to correlate the images to obtain the velocity vec-
needed to clearly identify specific types of flow structures. tors. The details of this correlation technique have been
The PIV system included dual Nd:YAG lasers that were explained in Ref. 8. The entire region of focus was di-
operated in phase synchronization with the rotor, the op- vided in to 64× 64 nodes. An interrogation window of
tical arm to transmit the laser light into the region of in- 24 pixels on either side with a 50% overlap was used for
terrogation, a digital CCD camera with 1 mega-pixel res- correlation. The number of allowed interpolated vectors
olution placed orthogonally to the laser light sheet for 2- was restricted to less than 1% of the total number vectors
C measurements (see Fig. 9a), a high-speed digital frame in any one image.

6
Figure 10: Schematic explaining the DPIV image processing technique used in the current study.

For the 3-C DPIV measurements, a new recursive wake flows. Detailed analysis on the uncertainty associ-
technique called deformation grid correlation was used ated with the process has also been made in Ref. 15.
(Ref. 14). This procedure is similar to that of the sim-
ple recursive technique, however, the second window is
deformed in that it is both sheared and translated instead Results
of just using the simple translation—see Fig. 10. The pro-
cedure starts with the correlation of an interrogation win- The observed results have been analyzed in the following
dow of a defined size (say, 64-by-64), which is the first categories: (1) Performance of the micro-rotor estimated
iteration. After the mean displacement of that region is from mass balance and torque sensors; (2) Sectional lift
estimated, the interrogation window of the displaced im- and drag estimation from the flow field analysis; (3) Evo-
age is moved by integer pixel values for better correla- lutionary characteristics of the tip vortices.
tion in the second iteration. This third iteration starts by
moving the interrogation window of the displaced image Performance
by sub-pixel values based on the displacement estimated
from second iteration. Following this, the interrogation Figures 11a through 11c show the performance spolars
window is sheared twice (for integer and sub-pixel val- for the various blade configurations. All the plots include
ues) based on the velocity magnitudes from the neighbor- measurements obtained at the four collective pitch angles
ing nodes before performing a fourth and fifth iteration, and five rotational speeds. The theoretical results were
respectively. made using the simple momentum theory and by estimat-
ing average values of κ and Cd0 in a least-squares sense
Once the velocity is estimated after these five itera- from the measured data. It should be understood that in
tions, the window is split into four equal windows (of this context momentum theory may not be applicable be-
size 32 × 32). These windows are moved by the average cause of its assumptions, especially at these low operating
displacement estimated from the final iteration (using a Reynolds numbers. However, the results can still be used
window size of 64 × 64) before starting the first iteration as a baseline for comparative analysis; this is the purpose
at this resolution. This procedure can be continued until here and it is not meant to be an endorsement of the appli-
the resolution required to properly resolve the flow field cability of this theory.
is reached. The second interrogation window is deformed Figure 11a shows the measurement made for three rect-
until the particles remain at the same location after the angular blade planforms: (1) The simple baseline; (2) The
correlation. This technique, especially the introduction of baseline with sharpened leading edge (SLE); (3) The base-
shear, has been shown in Ref. 15 to be appropriate for line with sharpened leading- and trailing-edges (SLT). It
measuring the high velocity gradients found inside rotor is apparent that the modified baseline cases (i.e., the SLE

7
(a) Rectangular blade configurations
(a) Rectangular blade configurations

(b) Tapered blade configurations


(b) Tapered blade configurations

(c) Twisted blade configurations


Figure 11: Power polars for the various rotor blade
(c) Twisted blade configurations
configurations.
Figure 12: Figure of merit for the various rotor blade
and SLT blades) perform better than the baseline. This configurations.
means that the amount of power consumed to produce the
same amount of thrust is smaller for the modified baseline in Fig. 12a. It can be seen that the measured FM was
cases. It can be seen that CP0 , which represents the min- improved by about 10% for the modified baseline cases.
imum power consumed (essentially at non-lifting or zero Because all the three blades have identical rectangular
CT ) is also smaller for the modified baseline cases. This tip shapes and were operated under identical conditions,
suggests that reducing the profile losses by sharpening the the induced losses among these blades should also be
leading-edge (or both the leading- and trailing-edges) is fairly similar, unless the modified boundary layer (from
effective for increasing rotor performance. the sharpening of the blades) altered the roll-up and evolu-
The improved performance in terms of FM are shown tionary characteristics of the wake and tip vortices. There

8
was some evidence of this, although only preliminary re-
sults have been measured thus far. Between the SLE and
SLT cases, the SLT blade showed a slightly better perfor-
mance than for the SLE blade.
Figure 11b shows the performance characteristics of
the tapered blade configurations. One immediate observa-
tion is that the inverse taper configuration performs poorly
when compared with the conventional tapered blades. In-
verse taper was designed to increase the operating tip
Reynolds number, which initially was felt that it may help
to reduce the profile losses (see Fig. 6). However, it is
apparent that the value of CP0 is actually higher for the
inverse tapered blade. On the other hand, at the same
pitch angle and at the same rotor tip speed, the inverse (a) Rectangular blade configurations
tapered blade produced significantly higher thrust than for
the other two blades.
Even though no substantial difference in the polars
could be observed between the two linearly tapered
blades, plotting their FM revealed some differences. Be-
cause the Taper-2 blade configuration produces substan-
tially lower thrust at lower pitch angles for the same
power, the FM values are also lower. Also, there is no
substantial increase in the lift characteristics of the tapered
blade when compared with the baseline case. This is con-
sistent with the observation made in Ref. 4.
Both of the twisted blades exhibited better performance
compared with the baseline (untwisted) case. It can be
seen that the twisted blades produced slightly higher val- (b) Tapered blade configurations
ues of CT for the same power when compared to the base-
line case. This is expected because the inflow distribu-
tion across the twisted blade should be more uniform (see
later), resulting in lower induced losses and better rotor
performance.
Figures 13a through 13c show the measured power
loading for all the rotor blade configurations with respect
to disk loading. It can be seen that the operating disk
loadings are substantially lower than for large scale rotors.
Consequently, their power loading is much better. This is
expected based on Eq. 1. However, hovering efficiencies
(the FM) are still relatively low at these disk loadings. Be-
cause rotor power varies with the cube power of rpm while
thrust increases by the square of rpm, operating the rotor (c) Twisted blade configurations
at a lower rotational speed and at higher pitch angles will Figure 13: Variation of measured power loading ver-
reduce the profile losses (Ref. 4). sus disk loading.

Bound Circulation & Sectional Lift micro-scale rotors with such small, thin blades, estimating
Measuring the sectional lift across the blade span is es- the spanwise lift distribution using pressure sensors is not
sential to understand the aerodynamic performance of the possible. In such cases where direct measurement of pres-
rotating-wing system. This is because the formation (roll- sure is not possible, the concept of “bound” circulation
up) and evolutional characteristics of tip vortices directly has been used successfully (Refs. 22–24, 17).
depends upon the distribution of bound circulation across The concept of estimating the sectional lift per unit span
the wing span (Refs. 16, 17). The distribution of lift (from the bound circulation) at a specified span location
can be measured using static pressure taps. However, for can be obtained from the Kutta—Joukowski theorem, as

9
Figure 15a shows the results of this closed loop in-
tegration process for three representative spanwise loca-
tions. It can be seen that the estimated circulation us-
ing the primary integral loop is smaller. However, with
an increase in the integration path length, the sectional
circulation increases significantly before asymptoting to
a constant value. Similar results were seen at all span-
wise locations, except for very close to the blade tip. The
blade lift loading can then be determined from this asymp-
totic value of sectional circulation through the application
of the Kutta—Joukouski theorem (Eq. 5). The measured
peak bound circulation and the associated lift value per
unit span are given in Table 1.
Figure 14: Integration around the blade to estimate Figures 15b to 15f show the estimated lift distribution
the sectional “bound” circulation at a spanwise loca- across the blade span for the five different blade configu-
tion. rations. The figures also include the distribution of drag,
the estimation of which is detailed in a later section of this
given by paper. It can be seen that the lift is nearly linearly dis-
L = ρ(Ωr)Γb (5) tributed in the baseline case and is also heavily tip loaded.
The value of sectional lift can be seen to reach its peak
The challenges involved in estimating a unique value value at approximately the 88.5% span location. It should
of circulation that can be related to the sectional lift of the be appreciated that the experiment was performed with a
blade are discussed in Ref. 17. Care must be taken to en- finite spatial resolution, especially along the spanwise di-
sure that the circulation contained within the integration rection of the blade. As a result, the spanwise location at
loop contains only the “bound” part. The shed and trailed which the maximum value of sectional lift was observed
vorticity from the wake must be avoided, which otherwise, may only be approximate. The reduction in sectional lift
would result in an incorrect estimation of the circulation. observed near the tip is from the tip vortex, which induces
In the case of a fixed-wing, this process is relatively eas- a downwash velocity that reduces the effective angle of
ier because of the simpler downstream wake structure, ex- attack to nearly zero.
cept near the wing tip. In the case of rotating-wings, the The lift distribution across the blade for the SLT cases
presence of the returning wake makes the estimation of appears similar to that of the simple baseline case. How-
“bound” circulation much more difficult. As suggested by ever, the point of maximum lift appears slightly closer to-
Taylor (Ref. 25), which was later confirmed by Bhagwat wards the blade tip. The baseline blade with sharpened
& Leishman (Ref. 17), the integration loops must be mod- leading edge (SLE) and the baseline blade with the SLT
ified to prevent the inclusion of any shed vorticity; this can were designed to reduce the profile losses. However, as
be achieved by keeping the appropriate branches of the in- mentioned earlier, profile and induced losses are likely
tegration contours perpendicular to the streamwise axis of to be interdependent at this scale, so reducing one is no
the shed wake. guarantee of better rotor performance As expected, the
The schematic in Fig. 14 shows the procedure followed overall thrust produced by the SLT blade (obtained from
in estimating the sectional bound circulation along the the integration of the spanwise distribution of lift) ap-
blade. The first loop completely encloses the rotor blade, pears slightly higher than for the baseline case—see Ta-
i.e., the diagonal of this primary integration loop connects ble 2. The table also includes the performance values mea-
the leading- and trailing-edge of the blade section. In the sured by the balance for comparison; good correlation was
second step, the length of the integration loop is increased found between the two techniques.
by ∆x along all four sides. Subsequent integral loops in-
The lift distribution for the tapered blade was different
crease their path length in a similar manner by increasing
from all other blades because it had a double lift peak near
∆x on all four sides from the previous integral loop. Esti-
the tip. This is likely a result of a secondary trailed vortex
mates of the bound circulation, Γb were then made by nu-
(other than the tip vortex), which results from the sudden
merically evaluating the closed loop integral of the mea-
change in the circulation gradient at the point where the
sured velocity field around a counterclockwise contour,
taper starts. This secondary vortex has the same direction
and fully encompassing the blade section without extra-
of rotation as that of the conventional tip vortex. As a re-
neous vorticity, i.e., by using
sult, there is an upwash from a trailed vortex (between the
I point of origination and the blade tip) which counters the
Γb = V. ds = ∑ UT i (∆x) +WT i (∆z) (6) downwash induced by tip vortex. This limits the reduction
c ¯

10
Blade Span Maximum sectional Maximum sectional
configuration location (y/R) bound circulation, (Γb /ΩRc) lift (lbs/ft)
Baseline 0.884 0.355 0.36282
Baseline–sharp l.e. & t.e. 0.912 0.313 0.32672
Taper 0.855 0.411 0.40242
Inverse taper 0.912 0.3255 0.33974
Twist (−9◦ linear) 0.825 0.336 0.31763
Table 1: Measured values of maximum sectional bound circulation, maximum lift per unit span, and their corre-
sponding span locations: θ = 12◦ , ΩR = 27 m/s.

Blade configuration CT CP CT CP0 CPi CP


3/2 √
(Technique (Mass (Torque (Bound (Wake CT / 2 CP0 +CPi
used) balance) sensor) circulation) integral)
Baseline 0.015992 0.0030651 0.014678 0.0021507 0.0012574 0.0034081
Baseline - SLT 0.017223 0.0024482 0.015316 0.0016715 0.0013402 0.0030118
Taper 0.014586 0.0024443 0.013748 0.0020448 0.0011399 0.0031847
Inverse taper 0.016702 0.0036304 0.011909 0.0019163 0.0009189 0.0028352
Twist (−9◦ linear) 0.014814 0.0025662 0.016683 0.0018024 0.0015238 0.0033261
Table 2: Comparison of the coefficient of thrust and torque measured using different techniques: θ = 12◦ , ΩR = 27
m/s.
in angle of attack induced by the tip vortex. Consequently, across the entire blade (Ref. 26). The more the devia-
the lift produced at the blade span locations between the tion from uniform values of inflow, the poorer the aero-
two vortices shows higher values than for any other blade. dynamic performance would be. In the experiments, the
However, the overall integrated lift values produced by the value of inflow was measured just above the 1/4-chord of
tapered blades is still lower than for the baseline case. the blade. It can be seen from Fig. 16b, that as expected
The intent of the blade with inverse taper was to op- the twisted blades have more uniform inflow across the
erate the tip region at higher chord Reynolds numbers. entire blade span, while the profile losses are similar in
Even though the maximum thrust produced by the in- value to the baseline blade case (see Fig. 11c).
verse taper blade was found to be higher than any other The results for SLT blade were found to be similar to
blade, partially satisfying the intended goal of producing the baseline case. As a result, the induced losses can be
more thrust, the thrust estimated from the bound circula- assumed to be almost similar to that for the simple base-
tion concept seems to be substantially lower. Unlike the line case as it has essentially a similar linear inflow—see
tapered blade, the blade with inverse taper did not show Fig. 16a. However, sharpening the leading- and trailing-
a double lift peak. The presence of a secondary trailed edges of the blades significantly reduced the profile losses.
vortex with the tapered blade could have been identified
by making measurements in the spanwise plane, however,
this is left for future work.
Wake Integration & Profile Drag
The linearly twisted blade shows a lift distribution simi- Unlike the difficulties in the estimation of the sectional
lar to that of baseline case. Here, as mentioned earlier, the lift distribution across the blades using the bound circula-
intent of the twisted blade is to try to produce a more uni- tion approach, measuring the distribution of profile drag is
form inflow distribution across the disk, and thereby help slightly easier (Refs. 27–31). The momentum deficit ap-
to minimize induced power losses. This can be achieved proach, defined by Betz (Ref. 16), can be used to estimate
by operating the inboard part of the blade at higher angles the sectional drag by comparing the momentum upstream
of attack. Again, even though the lift distribution appears and downstream of the airfoil (wing section) (Ref. 30).
similar to that of baseline case, the overall thrust produced Wu et al. (Ref. 32) and Hackett & Sugavanam (Ref. 33)
by the twisted blade is higher than for any other blade further developed this equation for finite wings to separate
configuration—see Table 2. out the components of drag (induced and profile). A fur-
The importance of the inflow spanwise distribution can ther explanation of this approach is also given by McAlis-
be better understood by reverting back to simple rotor the- ter (Ref. 34).
ory, which suggests that a rotating-wing will have maxi- In the case of two-dimensional airfoils (infinite wings),
mum aerodynamic performance if the inflow is uniform there is no induced drag because of the absence of a trailed

11
(a) Inverse taper (b) Baseline

(c) Baseline–sharp l.e.& t.e. (d) Taper

(e) Twist (f) Inverse taper


Figure 15: Performance measurement for various rotor blade configurations (solid line – lift, dashed line – drag).

wake and tip vortices. For finite wings, the tip vortices upstream and downstream of the blade section gives
are the predominant flow features and they primarily in- Z ∞ Z ∞
duce two components of velocity: 1. A spanwise velocity D= (p1 + ρU1 2 )dS − (p2 + ρU2 2 )dS (7)
−∞ −∞
from tip of the wing towards the root (VT ); 2. A veloc-
ity component normal to the wing resulting from the swirl By introducing total pressure, which is given by
velocity component of the tip vortex (WT ). This can be 1
P = p + ρ(U 2 +V 2 +W 2 ) (8)
better understood from Fig. 17. The resulting induced 2
velocity field inclines the lift vector aft, which upon re- the momentum balance in Eq. 7 (for an incompressible
solving introduces a component of drag (induced drag). flow) reduces to
By tilting the lift vector back, i.e, by having VT 1 = VT 2 Z ∞
1
Z∞
and WT 1 = WT 2 , it is possible to measure the profile drag D = (P1 − P2 )dS + ρ (UT 1 2 −UT 2 2 )dS(9)
−∞ 2 −∞
alone.
1
Z ∞
In reference to the rotating blade, a momentum balance − ρ ((VT 1 2 +WT 1 2 ) − (VT 2 2 +WT 2 2 ))dS
(10)
2 −∞

12
Figure 17: Difference between 2-D and 3-D flow showing the source of induced drag.

(a) Baseline & twisted blade configurations

Figure 18: Measurement of profile drag per unit


span—velocity deficit behind the blade at the 94%
span location.
ing
1
Z ∞ Z ∞
D= (P1 − P2 )dS + ρ (UT 1 2 −UT 2 2 )dS (11)
−∞ 2 −∞
R ∞
Here, the integral −∞ (P1 − P2 )dS can be assumed to be
negligible because of there is a very small reduction of
pressure inside the rotor wake downstream of the blade.
Using continuity gives

(b) Tapered blade configurations


ρUT 1 = ρUT 2 (12)

and substituting this in Eq. 11 gives the expression for


Figure 16: Measured inflow distribution across the
drag as
span for various blade configurations. 1
Z ∞
D= ρ UT (UT 1 −UT 2 )dS (13)
where the subscripts 1 and 2 denote the upstream and 2 −∞ 2
downstream locations relative to the blade section, re- Here UT 1 − UT 2 is the decrease in flow velocity, which
spectively. The first two integrals represent the profile when multiplied by the mass flux ρUT 2 , gives the decre-
drag, and the third integral represents the induced drag ment in momentum per unit time in the drag direction.
(Ref. 34). Figure 18 shows the reduction in flow velocity behind
In the present study, the profile drag was estimated us- the rotor blade obtained from the measurements and the

13
Figure 20: Example of phase-averaged velocity mea-
surements in the flow field of the micro-rotor.
ious blade sections for the different blades. It is apparent
that the baseline blade with the SLT configuration and the
twisted blades have a better L/D ratio than any of the other
blades. Consequently, the overall performance of the ro-
tors with these blades is also better.

Figure 19: Measured lift-to-drag ratio for various Tip Vortex Measurements
blade configurations.
Understanding the formation and evolutionary character-
process outlined above. The drag per unit span measured istics of the blade tip vortices is critical. This is because
at different span locations for the various blades is shown of their influence on the induced inflow and overall per-
in Figs. 15. One immediate observation is that the drag is formance of any rotating wing system, irrespective of the
approximately one-fifth of the estimated lift at that loca- size or the operating Reynolds number (Refs. 1, 6–8).
tion for all of the blades. This is consistent with the obser- In the case of MAVs, this is more critical because the tip
vations made by Laitone (Ref. 28) and Mueller (Ref. 29) vortices are larger in size relative to the dimensions of the
for low Reynolds number wing and airfoil flows. The sec- rotor blade (Ref. 1). Figure 15 clearly demonstrates the
tional drag is also linearly distributed from the root of the importance of the tip vortices, where the loss of circula-
blade to until approximately 85% of the blade radius, and tion (or lift) near the tip of the blade is primarily a result
the drag then falls off rapidly towards the tip because of of the downwash velocity induced by the tip vortices.
the influence of the tip vortex.
To understand the relationship between the tip vor-
Operating the blade sections at or close to their best
3/2 tex and the rotor performance, DPIV measurements were
values of Cl /Cd is known improve the aerodynamic ef- made in the wake of the baseline blade to measure all three
ficiency of a rotor system. This can be understood by writ- components of the flow velocities. The acquired instanta-
ing the FM in terms of Cl /Cd , i.e, neous velocity maps were conditionally phase-averaged to
1 correct for the effects of aperiodicity in the flow (Ref. 35).
FM = !−1 (14) The resulting mean velocity across the region of interest
3/2
2.6 CL is shown in Fig. 20.
κ+ √
σ CD Betz (Ref. 16) developed a mathematical model to
relate the roll-up of the tip vortices with the maximum
after assuming that the overall average lift coefficient sectional bound circulation found across the span of the
CL = Cl = 6(CT /σ) and all the blade sections operate at blade. Assuming that the tip vortex is responsible for the
the same sectional drag coefficient Cd = CD . Figures 19a loss of circulation near the tip, it was hypothesized that the
and b show the sectional lift-to-drag ratio obtained at var- circulation contained within the vortex should be equal to

14
Figure 21: Ratio of total vortex circulation in the tip (a) Growth of the tip vortex core
vortex to the maximum measured bound circulation.

the maximum sectional bound circulation, i.e., all the vor-


ticity beyond the point of maximum sectional circulation
should roll up into the tip vortex. In the present study,
the maximum sectional bound circulation for the baseline
blade was measured to be 0.1820 m2 s−1 .
Figure 21 shows the circulation contained within the to-
tal vortex normalized by the peak sectional bound circu-
lation at various wake ages. Two key observations should
be made here: (1) the peak value of this ratio, and (2) the
wake age (or time) at which this ratio reached its peak
value. The maximum value of this ratio can be seen to be
approximately 0.8. This means that only 80% of vortic-
ity beyond the point of maximum circulation has rolled (b) Reduction in the peak swirl velocity with time
up in to the tip vortex, which is smaller than that pre- Figure 22: Measured sizes of the tip vortex cores and
dicted by the Betz model. Also, it can be observed that the their associated peak swirl velocities at various wake
peak value is attained at 60◦ wake age, i.e., the tip vortex ages.
roll up is complete over a time that is substantially longer
than for rotor blades operated at higher Reynolds numbers both horizontal and vertical cuts through the vortex core
(Refs. 6, 17). The observed increase in roll-up time was have been used instead of horizontal cuts only in the case
also found to reflect in the evolutionary properties of tip of Ref. 1. Also, the spatial resolution is much higher in
vortices—see Fig. 22—and is consistent with the observa- the present study compared with the other experiment.
tion made in Ref. 1. The ratio of total vortex circulation Figure 23 is also plotted with an theoretical solution to
to the maximum sectional bound circulation can be seen the N–S equations as obtained by Iversen (Ref. 36). A
to decrease slowly with time, suggesting the dissipation of larger value of the circulation ratio basically means that
energy through turbulent mechanisms. the diffusion of vorticity will be very slow similar to those
The ratio of circulation contained within the viscous found in a laminar flow at very low vortex Reynolds num-
vortex core to that of the total vortex circulation and its bers. Despite operating at a lower vortex Reynolds num-
variation with Reynolds number is given in Fig. 23. It ber, the value of this ratio for the current set of measure-
can be seen that approximately 50–70% of the total vor- ments is similar to those found at higher vortex Reynolds
tex circulation is contained within the vortex core. This numbers. This implies then that the growth properties of
figure also includes measurements from another (but sim- tip vortices should also be similar to that of tip vortices
ilar) micro-rotor experiment (Ref. 1), as well as measure- generated at higher vortex Reynolds numbers.
ments obtained from rotors operated at higher Reynolds
numbers. The small difference between the current mea- Core size and Peak Swirl Velocity
surements and those shown in Ref. 1 can be attributed to
the difference in the experimental and data processing pro- The variation of the measured core size (half the distance
cedures used to estimate the core and total vortex circula- between the swirl velocity peaks) with respect to wake age
tion. In the present study, mean values estimated from (or time) is given in Table 3. The table lists the dimensions

15
Wake age rc /c Vtip /ΩR rc /c Vtip /ΩR rc /c Vtip /ΩR
ζ H-cut H-cut V-cut V-cut Mean Mean
6◦ 0.0804 0.2888 0.1088 0.2303 0.0946 0.2596
30◦ 0.0520 0.3327 0.05216 0.3106 0.0521 0.3163
60◦ 0.0804 0.2735 0.0994 0.2264 0.0899 0.2500
90◦ 0.1041 0.2316 0.1041 0.2095 0.1041 0.2205
120◦ 0.0946 0.2114 0.1183 0.2055 0.1065 0.2085
150◦ 0.1135 0.2043 0.1278 0.1940 0.1207 0.1991
180◦ 0.0994 0.2180 0.0899 0.2102 0.0946 0.2141
210◦ 0.1372 0.1963 0.1325 0.1700 0.1349 0.1831
240◦ 0.1183 0.2163 0.1325 0.1748 0.1254 0.1956
Table 3: Evolutionary characteristics of the tip vortices: (H-cut – Horizontal cut, V-cut – Vertical cut).

Figure 24: Measured non-dimensionalized velocities in


Figure 23: Ratio of circulation contained within the the tip vortex at ζ = 60◦ wake age.
vortex core to the total vortex circulation.

of the vortex cores and the estimated peak swirl velocities tionary properties of tip vortices is identifying the char-
that were obtained by making two orthogonal (horizon- acteristic variables. Usually, core size of the tip vortex
tal and vertical) cuts through the vortex, along with their is non-dimensionalized using a characteristic length such
mean values. It can be observed that the vortex core is as chord or radius of the blade. Swirl velocity is normal-
not axisymmetric because the peak swirl velocity occurs ized using the tip speed of the rotor. However, analyzing
at different radial distances between the two cuts. This the evolutionary properties of tip vortices also requires a
suggests that a complete characterization of the tip vor- characteristic time scale. The use of absolute time is not
tex properties is not possible using a measurement grid appropriate because the diffusion of vorticity (and hence
that passes along a horizontal or vertical line through the the growth of the vortex core) is not only a function of
center of the vortex, as is usually assumed in most laser time, but also a function vortex Reynolds number (or to-
Doppler velocimeter (LDV) measurements. tal vortex circulation). As a result, vortex circulation must
Figure 24 shows the measured mean swirl velocities be taken into account to properly understand the growth
across the tip vortex at 60◦ wake age. By convention, the properties of the blade tip vortices.
velocities and the radial distances were normalized by the In the present study, following Refs 36 and 1, the non-
tip speed of the rotor and the blade radius. Notice that the dimensionalized mean core size and the peak swirl veloc-
peak swirl velocity is not symmetrical on either side of the ity were plotted in Fig. 22 against non-dimensionalized
vortex, mainly because of its non-zero convection veloc- time. It can be seen from Fig. 22 that the core size ini-
ity through the flow. The maximum average peak swirl tially decreases before starting to increase, which is a di-
velocity was estimated to be 25% of the tip speed of the rect result of the action of viscous and turbulent diffusion.
rotor. The difference between the two peaks on each side The initial reduction in the core size of the tip vortex, as
of the core can be assumed to be its average convection mentioned earlier, can be attributed to its delayed roll up
velocity, which is also a measure of the induced velocity (Ref. 1) at these low Reynolds numbers. Also, the nor-
inside the slipstream boundary of the rotor wake. malized initial core size of the tip vortex was found to be
One of the complexities in understanding the evolu- larger than those found on larger scale rotors (Ref. 7). In

16
the present study, the initial core size, r0 , was found to be 3. DPIV was found to be a powerful technique for in-
approximately 5% of the mean blade chord. terrogating the flow near the rotor. Good correla-
The curve fit in Fig. 22a is obtained using a variation of tions were found between the balance measurements
Squire’s core growth model, which is given by against the estimated thrust and torque values based
r on integration of section properties. Estimating the
ζ lift and drag distributions along the blade span using
rc = r02 + 4ανδ (15) flow field measurements with the circulation and mo-

mentum deficiency techniques gave good insight into
where α is Lamb’s constant and δ is the ratio of appar- the performance of the rotor system.
ent to actual viscosity. A value of δ = 6.4 seems to fit the
growth rate of the vortex core estimated from the measure- 4. The tip vortex seems to plays a more important role
ments. This observation is consistent with Fig. 23, which in influencing the performance of a MAV-scale ro-
suggested that the rate of core growth of the tip vortices tor because of its larger overall size relative to the
should be similar to those observed using larger scale ro- radius and chord dimensions of the rotor. The roll-
tors, despite operating at lower vortex Reynolds numbers. up process of the tip vortex was measured to take a
The measured peak swirl velocity for increasing time is relatively longer time when compared to tip vortex
shown in Fig. 22b. This, as expected, can be seen to de- developments that have been measured higher vortex
crease inversely with t 1/2 to conserve circulation. Reynolds numbers.

5. Despite the much lower vortex Reynolds numbers


found on MAV-scale rotors, the tip vortices exhibited
Conclusions growth properties (viscous and turbulent diffusion)
similar to that observed at higher Reynolds numbers.
To identify and measure the various sources of profile But the evolutionary characteristics of tip vortices,
and induced losses, measurements of the performance and when analyzed in the appropriate time scales, con-
the wake characteristics of a small MAV-scale or “micro- firmed a higher growth rate than that would be ex-
rotor” was performed. Several rotor blade configurations pected based on the vortex Reynolds number. About
have been tested over a range of pitch angles and rota- 50% of the total vortex circulation was found to be
tional speed combinations. Several new observations have inside the viscous vortex core, the circulation which
been found that help in understanding the nature of such increased with time.
small-scale rotor performance. The following are conclu-
sions derived from this study:
Acknowledgments
1. Conventional rotor design strategies for reducing in-
duced and profile losses by altering the blade shape This research was supported, in part, by the Multi-
seem applicable to MAV-scale rotor systems. The ef- University Research Initiative under Grant ARMY
fects of blade twist were shown to reduce losses and W911NF0410176. Dr. Tom Doligalski is the technical
improve performance. It is not yet clear, however, if monitor. The authors also wish to acknowledge Tyler
there is an optimum blade twist that can help to max- Huismann for his contributions in this work.
imize rotor efficiency, although linear twist seems to
be a good start.
References
2. Reducing profile losses through blade section (airfoil 1 Ramasamy, M., Leishman, J. G., and Lee, T. E..,
“Flow
shape) changes (albeit non-traditional changes) is
Field of a Rotating Wing MAV,” American Helicopter So-
an effective means of improving rotor performance.
ciety 62th Annual National Forum Proceedings, Phoenix,
In the present case, modified rectangular blades
AZ, May 7–9, 2006.
with circular-arc airfoil sections that have sharpened
leading- and trailing-edges showed improved perfor- 2 Bohorquez, F., Samuel, P., Sirohi, J., Pines, D., Rudd,
mance compared to the other rotor blades that were L., and Perel, R., “Design Analysis and Hover Perfor-
tested. The primary issues for the future are in bet- mance of a Rotating Wing Micro Air Vehicle,” The Jour-
ter understanding the airfoil shapes that can give best nal of American Helicopter Society, Vol. 48, No. 2, April
aerodynamic performance in the rotating-wing flow 2003, pp. 80–90.
environment at low chord Reynolds numbers below
50,000. To this end, further spanwise loads measure- 3 Hein, B., and Chopra, I., “Hover Performance of a Mi-
ments are necessary. cro Air Vehicle: Rotor at Low Reynolds Number,” AHS

17
International Specialists’ Meeting on Unmanned Rotor- 15 Ramasamy, M., and Leishman, J. G., “Benchmarking

craft: Design, Control and Testing, Phoenix, AZ, January PIV with LDV for Rotor Wake Vortex Flows,” AIAA pa-
2005. per 2006-3479, 24th Applied Aerodynamics conference
Proceedings, San Francisco, CA, June 6–8, 2006.
4 Gray, N., Kim, J., and Koratkar, N., “Aerodynamic
16 Betz,
A., “Behavior of Vortex Systems,” NACA
Design Considerations for Micro-Rotorcraft,” American
Helicopter Society 58th Annual National Forum Proceed- TM 713, 1935.
ings, Montreal, Canada, June 11–13, 2002. 17 Bhagwat, M., and Leishman, J. G., “Measurements
5 Kim, J., and Koratkar, N., “Effect of Surface Tempera- of Bound and Wake Circulation on a Helicopter Rotor,”
ture and Heat Transfer on the Aerodynamic Performance Journal of Aircraft, Vol. 37, No. 2, pp. 227–234, March
of Micro-Rotorcraft” American Helicopter Society 59th 2000.
Annual National Forum Proceedings, Phoenix, AZ, May 18 Fage,A., and Simmons, L. F., “An Investigat ion of the
6–8, 2003.
Air-Flow Pattern in the Wake an Aerofoil of Finite Span,
6 Martin, P. B., Pugliese, G., and Leishman, J. G., “High Philosophical Transactions of the Royal Society, Series A,
Resolution Trailing Vortex Measurements in the Wake of Vol . 225, 1925, pp. 303–330, 1925.
a Hovering Rotor,” Journal of the American Helicopter 19 Bryant, L. W., and Williams, D. H., “An Investiga-
Society, Vol. 49, No. 1, January 2004, pp. 39–52.
tion of the Flow of the Flow of Air Around an Aerofoil
7 Ramasamy, M., and Leishman, J. G., “Interdependence of Infinite Span,” Philosophical Transactions of the Royal
of Diffusion and Straining of Helicopter Blade Tip Vor- Society, Series A, Vol . 225, pp. 199–237, 1925.
tices,” Journal of Aircraft, Vol. 41, No. 5, September 20 McAlister,
K. W., and Takahashi, R. K., “NACA
2004, pp. 1014–1024. 0015 Wing Pressure and Trailing Vortex Measurements,”
8 McAlister, NASA TR 3151 AVSCOM TR 91-A-003, November
K., “Rotor Wake Development During the
1991.
First Revolution,” Journal of the American Helicopter So-
ciety, Vol. 49, No. 4, October 2004, pp. 371–390. 21 Kitapliouglu,
C., and Caradonna, F. W., “Aerodynam-
9 Yu., ics and Acoustics of Blade-Vortex Interaction using Inde-
Y. H., and Tung, C., “The HART-II Test: Ro-
pendantly Generated Vortex,” American Helicopter So-
tor Wakes and Aeroacoustics with Higher Harmonic Pitch
ciety Aeromechanics Specialists Meeting, January 19–21,
Control (HHC) Inputs,” American Helicopter Society
San Francisco, CA, 1994
58th Annual National Forum, Montréal, Canada, June 11–
13, 2002. 22 Orloff,K. L., “The Spanwise Lift Dist ribut ion on a
10 Mahalingam,
Wing from Flowfield Velocity Survey,” Proceedings of the
R., and Komerath, N. M., “Measure- AIAA 11th Fluid and Plasma Conference, Seattle, WA,
ments of the Near Wake of a Rotor in Forward Flight,” July 1978.
AIAA Paper No. 98-0692, 36th Aerospace Sciences
Meeting & Exhibit, Reno, NV, January 12–15, 1998. 23 Felker,
F. F., Piziali, R. A., and Gall, J. K., “Spanwi
se Loading Distribution and Wake Velocity Surveys of a
11 Tung, C., Pucci, S. L., Caradonna, F. X., and Morse,
Semi-Span Wing,” NASA TM 84213 USAARADCOM
H. A., “The Structure of Trailing Vortices Generated TR 82-A-1, Feb. 1982.
by Model Helicopter Rotor Blades,” NASA TM 81316,
1981. 24 Desabrais,
K. J., and Johar i, H., “Direct Circulation
Measurement of a Wing Tip Vortex Using Ultrasound,”
12 Cook, C. V.,
“The Structure of the Rotor Blade Tip Vor- Proceedings of 36th Aerospace Sciences Meet ing and Ex-
tex,” Paper 3, Aerodynamics of Rotary Wings, AGARD hibit, AIAA Paper 98-0609, Reno, NV, 1998.
CP-111, September 13–15, 1972.
25 Taylor, G. I., “Note on the Connection Between the
13 Leishman,J. G., “Seed Particle Dynamics in Tip Vor- Lift on an Aerofoil in a Wind Tunnel and the Circulation
tex Flows,” Journal of Aircraft, Vol. 33, No. 4, 1996, Round It , Philosophical Transactions of t he Royal Soci-
pp. 823–825. ety, Series A, Vol . 225, pp. 238–245, 1925.
14 Scarano,F., “Iterative Image Deformation Methods 26 Leishman, J. G., Principles of Helicopter Aerody-
in PIV,” Measurement Science and Technology, Vol. 12, namics, Cambridge University Press, New York, 2000,
2002, pp. R1–R19. Chap. 3, pp. 96–98.

18
27 Schmitz, F. W., “Aerodynamics of the Model Air-

plane,” Redstone Arsenal Translation, N70-39001, RSIC-


721, Nov. 22, 1967
28 Laitone, E. V., “ Wind Tunnel Tests of Wings at
Reynolds Numbers Below 70,000,” Experiments in Flu-
ids, Vol. 23, pp. 405–409, 1997.
29 Mueller, T. J., and Batill, S. M., Experimental Stud-

ies of Separation on a Two-Dimensional Airfoil at Low


Reynolds Numbers,” Journal of Aircraft, Vol. 20, pp. 457–
463, 1982.
30 Kusunose,K., “Drag Prediction Based on a Wake-
Integral Method,” 16th Applied Aerodynamics Confer-
ence, Albuquerque, NM, June 15–18, 1998. pp. 501–514.
31 Maskell,E. C., “Progress Towards a Method of Mea-
surement of the Components of Drag of a Wing of Fi-
nite Span,” Royal Aeronautical Establishment, RAE-TR-
72232, 1979
32 Wu, J. C., Hackett, J. E., and Liley, D. E., “A Gener-
alized Wake-Integral Approach for Drag Determination in
Three-Dimensional Flows,” American Institute of Aero-
nautics and Astronautics, 17th Aerospace Sciences Meet-
ing, New Orleans, La, 15–17 January 1979.
33 Hackett,
J. E., and Sugavanam, A., “Recent Develop-
ments in Three-Dimensional Wake Analysis,” AGARD
Report 723, Aircraft Drag Prediction and Reduction,
1985.
34 McAlister,
K. W., Schuler, C. A., Branum, L., and Wu,
J. C., “3-D Wake Measurements Near a Hovering Rotor
for Determining Profile and Induced Drag,” NASA TP
3577 ATCOM TR 95-A-006, August 1995.
35 Leishman, J. G., “Measurements of the Aperiodic
Wake of a Hovering Rotor,” Experiments in Fluids,
Vol. 25, 1998, pp. 352–361.
36 Iversen,
J. D., “Correlation of Turbulent Trailing Vor-
tex Decay Data,” Journal of Aircraft, Vol. 13, (5), May
1976, pp. 338–342.

19

Das könnte Ihnen auch gefallen