You are on page 1of 10

9TH INTERNATIONAL SYMPOSIUM ON FLOW VISUALISATION, 2000

SURFACE AND WAKE FLOW


CHARACTERISTICS OF A HOVERING
HELICOPTER ROTOR
J. G. Leishman1, P. B. Martin2, G. J. Pugliese3

Keywords: rotor, helicopter, tip vortex, liquid crystals, sublimation, oil flow, LDV,
PIV, laser-sheet

ABSTRACT
Surface flow visualization methods were examined for potential application to a hovering helicopter
rotor blade. These techniques included liquid crystals, chemical sublimation, and fluorescent oil
flow. Test cases involved both the identification of a laminar separation bubble and the effects of a
serrated boundary layer trip strip. The results were used to understand some of the boundary layer
physics associated with sub-scale helicopter rotor testing. In addition, the vortex system trailed
into the wake behind the rotor blades was studied using laser sheet smoke flow visualization. The
results were supported by quantitative measurements of the flow made using LDV and PIV.

1 INTRODUCTION
An improved understanding of the aerodynamics of helicopter rotors is one key to ad-
vanced rotor designs with better performance, lower vibratory loads, and less noise [1, 2].
The loads and overall performance of the helicopter rotor are highly dependent upon the
vortical wake structure below the rotor, and are primarily influenced by the strengths and
locations of the blade tip vortices. In addition, under some flight conditions the rotor wake
lies near the plane of the rotor resulting in a phenomenon called blade-vortex interaction.
This phenomenon can result in locally large unsteady airloads, increased rotor vibration
levels, and the generation of impulsive rotor noise. The highly three-dimensional nature of
the flow through and near a helicopter rotor means that measurements of the underlying
flow physics are difficult using any experimental technique. In particular, much is still to
be learned about the basic physics of tip vortex formation in terms of basic blade loading
parameters, as well as the subsequent viscous behavior of the tip vortices [2, 3].
Rotor flow field experiments have used a variety of diagnostic techniques such as hot
wire anemometry (HWA), laser Doppler velocimetry (LDV) and, more recently, particle
image velocimetry (PIV). Each technique has well-known advantages and disadvantages
in terms of its ability to measure the rotor flow field structures. Flow visualization is a
particularly important tool in helping to understand the physics of helicopter rotor wakes,
as well as to target the key flow structures for quantitative measurements. Many different

Author(s): 1 Professor. Department of Aerospace Engineering, Glenn L. Martin Institute of Tech-


nology, University of Maryland, College Park, Maryland 20742, USA
2 Graduate Research Assistant
3 Graduate Research Assistant

Corresponding author: J. G. Leishman (leishman@eng.umd.edu)

Paper number 189 189–1


J. G. Leishman, P. B. Martin, G. J. Pugliese

flow visualization methods have been applied to rotors; a concise survey describing these
applications is given in [4]. These methods include wake visualization by direct smoke
injection [5] and rotor blade boundary layer studies by chemical sublimation [6] and shear
sensitive liquid crystals [7].
The objectives of this work were to develop techniques for on-surface (blade) and
off-surface (wake) visualization of a sub-scale helicopter rotor operating in the hover
state. Sub-scale rotor testing usually involves operating a rotor blade at a lower chord
Reynolds number, with the attendant difficulties in extrapolating to full-scale. Surface
flow visualization was used to document the suspected presence of a substantial laminar
separation bubble along the entire blade radius. Because the local Reynolds number varies
along the blade radius, these methods were used to help size and position a boundary
layer trip strip that would attempt to replicate full-scale rotor Reynolds numbers. As
the boundary layers on each side of the blade detach into the wake, a vortex sheet flow
structure is formed. Another objective of this study was to visualize and attempt to
quantify the behaviors of the inboard vortex sheet trailed from the blade as well as the
concentrated vortex trailing from the blade tip.

2 EXPERIMENTAL PROCEDURES
A sub-scale helicopter rotor model with a teetering hub was manufactured. A water
cooled, variable frequency induction motor was used to drive the rotor. The rotor rig
was mounted in a flow conditioned test cell, which was located in a large high bay area.
The rotor was operated with one or two blades, the single-blade being balanced by a
counterweight. A tapered-tip blade planform was used for the present experiments, with
the 0.4 taper ratio starting at the blade 80% radial station (Fig. 1). The blade radius
was 410 mm, with a root chord of 44 mm and a tip chord of 17.6 mm. The blades used
a NACA 2415 airfoil section. The tip Mach number and Reynolds number were fixed at
0.3 and 120,000 respectively. The blade loading (ratio of rotor thrust to blade area) was
held nominally constant for all tests.

2.1 Blade Surface Visualization

To visualize the flow state on the blades, several methods of surface flow visualization were
studied. These included shear sensitive liquid crystals [8], chemical sublimation [9], and a
fluorescent oil film. In each case, trichloroethane was used as a solvent. The solvent was
applied to the blades by “dry” spraying a thin film of solvent/suspension solution using
an airbrush. The rotor was then operated at the required test condition for sufficient time
to allow the flow patterns to become established on the blade. The rotor was stopped,
and photographs were immediately taken of the resulting surface flow patterns.
Surface flow visualization was conducted transition-free and transition-fixed. In the
fixed transition case, a boundary layer trip in the form of serrated tape [10] was applied
to the blade surface. The serrated tape was applied at the 11% chord location gradually
tapering to the 14% chord location near the tip. This location was chosen after observing
laminar separation to occur near the 20% chord location over most of the blade. The
thickness of the trip was 0.3% chord with a chordwise extent of 6% chord. The serrations
were spaced at 0.64% of the blade radius.

2.2 Wake Visualization

The rotor tip vortex and wake structure were visualized by illuminating the seeded flow
with a strobed laser light sheet. The seed used for the visualization, as well as the PIV and

9th International Symposium on Flow Visualization, Heriot-Watt University, Edinburgh, 2000


Editors G M Carlomagno and I Grant 189–2
Surface and Wake Flow Characteristics of a Hovering Helicopter Rotor

PIV Image frames Blade-fixed Body-fixed


r/R z/R
Y
X

Motor Z

ζ = 119° Counter-weight

ζ = 75°

ζ = 217°
Teetering hub

ζ = 45° Ω

ζ = 12°
ζ = 7°
Blade passage Tip vortex filament ζ = 297°

Fig. 1 Schematic of experimental setup, showing rotor and PIV image frames in the
rotor wake.

LDV measurements, was atomized olive oil. By means of calibration, the mean seed parti-
cle size was 0.6±0.2 µm in terms of aerodynamic diameter. Phase-locked synchronization
of the light-sheet with the rotor phase was accomplished using an optical chopper with a
phase resolution of approximately ±0.1 degrees. The phase angle between the strobing
light sheet and the rotor blades could either be fixed or slipped. By slipping the phase, a
useful slow motion sequence was obtained.
The light scattering pattern of this size of particle is approximated by the Mie scat-
tering pattern, which is predominately forward scatter. To capture sufficient scattered
light, the camera optics were, therefore, optimized for low light levels and placed at an
angle to the illuminated plane. A high-speed telephoto lens with low dispersion glass
allowed zoom capability from 80 mm to 200 mm while maintaining an aperture setting
of f2.8 throughout the entire zoom range. A 35 mm SLR camera was used with high
speed 3200 ASA black and white film, push processed to approximately 6400 ASA. Video
images were obtained using the same telephoto lens and low light level CCD camera.
Accurate quantitative reconstruction of the rotor wake was then performed using a grid
based digital image processing technique. The images were corrected for the skew angle
between the light sheet plane and the camera by contour mapping a digitized grid. The
coordinates of the vortex cores and vortex sheets were then corrected using the coefficients
of the contour map.

2.3 PIV System

Phase-resolved stereoscopic PIV was used to acquire velocity field information in the rotor
wake. By interrogating the resulting image using various signal processing techniques, the
PIV images can provide a measure of the instantaneous three-dimensional velocity field.
Stereoscopic PIV measurements were made using a pair of frequency doubled, 35mJ/pulse
Nd:YAG lasers, an articulating light delivery arm, light-sheet optics, a pair of digital

9th International Symposium on Flow Visualization, Heriot-Watt University, Edinburgh, 2000


Editors G M Carlomagno and I Grant 189–3
J. G. Leishman, P. B. Martin, G. J. Pugliese

CCD cameras with 1K x 1K pixel resolution, a pair of high-speed frame grabbers, image
acquisition and analysis software, and a computer. A self contained mirror system in the
form of a fiber optic light arm transmitted the laser light from the mini-dual Nd:YAG
lasers to the rotor flow field.
The light sheet head was aligned with the rotor shaft to illuminate a radial plane
in the wake below the rotor. Light sheet optics, attached to the end of the articulating
light arm, produced vertical sheets with thickness of approximately 1 mm and height of
100 mm at the midpoint of the image area. The cameras and 60 mm focal length micro
lenses were attached to rotating mounts that were affixed to an optical rail. The angle
between the cameras was set to approximately 40◦ . The region of image overlap between
the two cameras, approximately 90 mm × 90 mm in area, was used to fix the cameras and
lenses to meet the Scheimpflug condition. This condition is used to record images with
the image sensor in the plane of best focus when the object plane and the lens principal
plane are not parallel. This arrangement allows a narrow depth of field and, therefore,
the minimum amount of illumination or laser energy per pulse.

2.4 LDV System

A three-component fiber optic based LDV system was used to make pointwise measure-
ments in the flow. The rotor tip vortex structures were of particular interest. A single
multi-line Argon-Ion laser beam was split into three pairs of beams, each of which mea-
sured a single component of velocity. A Bragg cell was used to produce the second shifted
beam of each pair. Frequency shifts on each channel were optimized for the range of the
bandpass filters and for the expected flow reversals. The laser beams were passed to the
transmitting optics by a set of single mode polarization preserving fiber optic cables. The
transmitting optics consisted of a pair of fiber optic probes. To reduce the probe volume
length, the crossing angle was increased using 2.6X beam expanders. Spatial coincidence
of the three probe volumes was ensured to within a 25µm radius using the beam align-
ment technique described by Martin et al. [11]. To further reduce the effective size of the
probe volume visible to the receiving optics, the off-axis backscatter technique was used.
Signal bursts were received by the optics and transmitted to a set of photomultiplier
tubes, where they were converted to analog signals. The signals were then digitized and
converted from Doppler frequencies into flow velocities. A rotary encoder enabled the
measured data to be tagged with blade azimuth information with a resolution of 0.1◦ . The
overall signal-to-noise ratio of the LDV measurements was improved by requiring both
spatial and temporal coincidence [11, 12]. The geometric (beam) coincidence determines
the spatial resolution of the probe volume. Measurements were also acquired within
a time coincidence window. The imposition of a temporal window ensured that for a
characteristic flow structure convecting past the probe, the phase-resolved measurements
were statistically correlated on all three channels.

3 RESULTS
3.1 Surface Flow Visualization

Figures 2 and 3 show representative images from the surface flow visualization study
using the liquid crystal and sublimation techniques, respectively. Results were obtained
transition-free and transition-fixed. The transition-free flow visualization confirmed the
suspected natural formation of a long laminar separation bubble at these relatively low
sub-scale Reynolds numbers, the behavior of which can significantly impact the rotor
performance at high thrust conditions. In the present tests, the blade pitch angles were

9th International Symposium on Flow Visualization, Heriot-Watt University, Edinburgh, 2000


Editors G M Carlomagno and I Grant 189–4
Surface and Wake Flow Characteristics of a Hovering Helicopter Rotor

(a) Transition free (b) Transition fixed

Bubble

Turbulent
boundary
layer

Trip
LS RL

LE
LE
TE TE

r /R r /R

Fig. 2 Shear sensitive liquid crystal surface flow visualization.

kept low enough to prevent stall. In the schematics next to each image are the interpreted
boundaries denoting laminar boundary layer separation (indicated by LS) and turbulent
boundary layer reattachment (indicated by RL).
While the liquid crystals respond to the surface shear stress exerted by the boundary
layer, they also respond to the centrifugal force on the rotating blade. Therefore, the
results in Fig. 2 show surface “streamlines” with a radially outward bias angle. This
angle depends on a balance of the surface shear stress and the centrifugal bias, the latter
being strongest near the blade tip. In addition, the liquid crystals respond with a color
change in regions where the shear stress changes magnitude or direction. Inside the
laminar separation bubble, where the surface shear is negligible, the surface streamlines
are directly radially outward. In spite of this, the liquid crystals clearly demarcate the
points of laminar separation and turbulent boundary layer reattachment. Notice that
the bubble varied in chordwise extent along the length of the blade. The bubble was
approximately 30% chord in length over the outer 50% of the blade radius. Inboard on
the blade, where the Reynolds number is very low, laminar separation occurred near the
leading edge with the formation of a relatively long bubble that gradually grew to 100%
of the chord near the 20% radial station.
Figure 3 shows the corresponding results obtained using the sublimation technique.
In this case, there is no centrifugal bias to the results. Again, the formation of a long
laminar separation bubble is evident. The sublimation technique has the disadvantage
of not clearly showing the point of laminar separation; the turbulent reattachment point
is, however, quite clearly defined. When sublimation shows a well defined line mark-
ing transition, then it was preceded by laminar separation. When sublimation shows a
ragged transition line, it was caused naturally by either Tollmien-Schlichting or crossflow

9th International Symposium on Flow Visualization, Heriot-Watt University, Edinburgh, 2000


Editors G M Carlomagno and I Grant 189–5
J. G. Leishman, P. B. Martin, G. J. Pugliese

(a) Transition free (b) Transition fixed

LS RL

Bubble
Turbulent
boundary
layer

Tape
Trip

LE
TE

TE
LE

r /R r /R

Fig. 3 Chemical sublimation surface flow visualization.

instabilities [9].
As previously mentioned, one of the problems with sub-scale helicopter rotor testing is
the relatively low Reynolds numbers on the blade. These may be less than 105 compared to
those found at full-scale, which may be up to two orders of magnitude higher. In addition,
on a rotor blade the Reynolds number varies linearly from a maximum at the blade tip to
a fairly small value inboard. While the aerodynamic problems associated with sub-scale
rotor testing have been widely acknowledged, no standardized techniques for simulating
full-scale rotor Reynolds numbers have been established. The calibration tape in Fig. 3(a)
shows how forcing transition eliminates the bubble from forming. In the present case, it
is shown in Figs. 2(b) and 3(b) that the full span application of the serrated trip strip
causes a clean transition to a turbulent boundary layer and eliminates the large separation
bubble normally found with the sub-scale rotor. The sublimation picture on the tripped
sub-scale rotor appears substantially similar to the full-scale sublimation results shown
in the hovering rotor tests of Boatwright [13].

3.2 Rotor Wake Visualization

A representative example of the visualized wake below the rotor is shown in Fig. 4. In
this case, the laser light-sheet was positioned in a radial plane through the rotor hub and
perpendicular to the rotational plane of the rotor. In this image, the rotor blade has
passed through the illuminating plane shown in Fig. 1, and the resulting flow structures
are about 75 degrees old relative to the original blade position. The rotor hub is to the
right of the image and the blade tip has previously passed near to the left-hand side of
the image.
Features to note in this image are the concentrated vortex that has trailed from the

9th International Symposium on Flow Visualization, Heriot-Watt University, Edinburgh, 2000


Editors G M Carlomagno and I Grant 189–6
Surface and Wake Flow Characteristics of a Hovering Helicopter Rotor

Vortex core

Vortex sheet

Fig. 4 Example of laser sheet flow visualization of the blade tip vortex and vortex sheet
at wake-age of 75 degrees.

blade tip and also the inboard turbulent wake, the latter being normally referred to
as a vortex sheet. The tip vortex itself is characterized by a very definite seed void,
which results from the centrifugal forces generated by high swirl velocities surrounding
the vortex core. In proximity to the vortex core is an apparent line of discontinuity in
the local streaklines. This is the vortex sheet and is the signature of the inboard viscous
wake trailed from the blade trailing edge. Notice that the vortex sheet feeds directly into
the tip vortex core; the tip vortex itself is a tightly rolled-up portion of the original vortex
sheet. In the case of two-bladed rotors, the vortex sheet generated by one blade interacts
with both the tip vortex from the same blade and also the tip vortex from the other blade
– see Martin et al. [5]. Therefore, the development of the vortex sheets and vortex cores
become highly interdependent flow features as the rotor wake evolves.

3.3 PIV and LDV Measurements

Phase-resolved stereoscopic particle image velocimetry (PIV) measurements were made


in the rotor wake. The results were also compared with phase-resolved coincident three-
dimensional laser Doppler velocimetry (LDV) measurements when using the same seeding
medium. PIV measurements were acquired for several wake ages during one rotor revo-
lution. LDV measurements were made along a radial grid through the vortex core at the
same approximate wake ages as the PIV measurements. This allowed a direct compari-
son of the induced velocity profiles across the vortex core when using both measurement
techniques. The relatively small vortex core size and high tangential velocities produced
by the small tip chord of the rotor blade provided a particularly good challenge to the
spatial resolution capabilities of both the PIV and LDV techniques.

9th International Symposium on Flow Visualization, Heriot-Watt University, Edinburgh, 2000


Editors G M Carlomagno and I Grant 189–7
J. G. Leishman, P. B. Martin, G. J. Pugliese

-0.05

0
V / ΩR
Axial station, z/R

0.14
0.12
0.10
0.08
0.05 0.06
0.04
0.02
0.00
0.1

0.15
1 0.95 0.9 0.85 0.8
Radial station, r/R

Fig. 5 PIV vector field of tip vortex and vortex sheet at a wake-age of 75 degrees.

Representative results showing the velocity vector field as measured with PIV are
given in Fig. 5 for a wake age of approximately 75◦ (see Fig. 1). The characteristic
signature of a tip vortex is apparent. Notice, however, that the radial velocity field is
not symmetric with respect to an axis centered at the vortex core. This is because of
two reasons. First, the tip vortex is convecting through the flow; in this case downward
and to the left in the image. This means that the flow field velocities are highest on the
advancing (right hand) side of the vortex flow. Second, the tip vortex convects along
the boundary of a jet-like rotor wake. Inside this wake the velocities are relatively high;
outside the wake boundary the flow is essentially quiescent. Notice also in Fig. 5 the
signature of the inner vortex sheet. Further PIV results documenting the structure of the
tip vortex as it convects through the flow below the rotor are given by Martin et al. [11].
Figure 6 shows representative tangential velocity profiles through the core of the rotor
tip vortex. These PIV and LDV results are for a wake age of approximately 75◦ . It was not
possible to obtain LDV and PIV results at exactly the same wake ages, so the LDV results
in Fig. 6 are shown at two wake ages just before and just after the PIV measurements.
While the LDV measurements clearly capture the steep velocity gradients around the
vortex core, it is apparent that the PIV results fail to resolve these velocity gradients.
Notice that away from the vortex core, both sets of measurements show good agreement.
Further LDV and PIV results documenting the tip vortex properties at different wake
ages are given in [14]. The high spatial resolution of the LDV system allowed the tangential
velocity profiles and viscous core dimensions to be estimated as a function of wake age.
The LDV results suggest a logarithmic core growth trend consistent with a Lamb-like
model of the vortex. The present results suggest that while the PIV used here is capable
of phase-resolved measurements in the rotor wake, there are several problems that require

9th International Symposium on Flow Visualization, Heriot-Watt University, Edinburgh, 2000


Editors G M Carlomagno and I Grant 189–8
Surface and Wake Flow Characteristics of a Hovering Helicopter Rotor

1 0.95 0.9 0.85


-0.3 -0.3

LDV, ζ ≈ 65°
-0.2 LDV, ζ ≈ 83° -0.2
PIV, ζ ≈ 75°

-0.1 -0.1

LDV
V θ / ΩR

0 0

PIV
0.1 0.1

0.2 0.2

0.3 0.3
1 0.95 0.9 0.85
Radial station, r/R

Fig. 6 Comparison of the tangential velocity profile in the tip vortex as measured by
PIV and LDV at a wake-age of 75 degrees.

resolution. First, the vortical nature of the flow results in the entrainment of seed particles
into concentrated bands. This nonuniform seeding was found to be a source of difficulty
in correlating stereo images from both cameras. Second, the presence of the seed void
in the region of the vortex core was not captured by PIV. Results produced by the PIV
image analysis software across regions of the flow where there is clearly no seed indicate
a problem with either the temporal or spatial resolution. In effect, the seed void is not
frozen on the grid, and so the velocity field surrounding the vortex core is not accurately
resolved. Third, there is a loss in spatial resolution because of the phase-resolved ensemble
frame averaging of PIV results. Despite these limitations, the PIV results have provided
reasonable quantitative insight into the flow physics inside the rotor wake.

4 CONCLUSIONS
Liquid crystal and sublimation surface flow visualization methods were used to examine
the surface flow on a sub-scale hovering helicopter rotor blade. The purpose of the ex-
periments was to help understand some of the boundary layer physics associated with
sub-scale rotor testing. The results have identified a substantial laminar separation bub-
ble on the blade, which was expected at the sub-scale Reynolds numbers of these tests.
A serrated boundary layer trip strip placed near the leading edge of the blade was shown
to successfully trip the boundary layer and prevent laminar separation.
The rotor wake and tip vortex system was studied using laser sheet smoke flow visu-
alization. The interdependence of the trailed vortex sheet and the tip vortex was noted.
Quantitative measurements of the flow field and tip vortex structure were made using
LDV and PIV. The PIV results were found to give good quantitative insight into the

9th International Symposium on Flow Visualization, Heriot-Watt University, Edinburgh, 2000


Editors G M Carlomagno and I Grant 189–9
J. G. Leishman, P. B. Martin, G. J. Pugliese

overall flow physics near the rotor. However, compared to that possible using LDV, it
was found that because of inhomogeneous seeding and other problems the present PIV
system was generally unable to resolve the steep velocity gradients present in the tip
vortices.

5 ACKNOWLEDGMENTS
This work was supported by the National Rotorcraft Technology Center (NRTC) under
grant NCC 2944. The authors appreciate the contributions of Steven Anderson and
Darrell Havir of TSI, Inc. to this work.

REFERENCES

[1] McCroskey, W. J., “Vortex Wakes of Rotorcraft,” Paper 95–0530, 33rd AIAA Aerospace
Sciences Meeting and Exhibit, Reno, NV, Jan. 9–12, 1995.
[2] Leishman, J. G. and Bagai, A., “Challenges in Understanding the Vortex Dynamics of He-
licopter Rotor Wakes,” AIAA Journal, Vol. 36, No. 7, July 1998, pp. 1130–1140. See also
Paper 96–1957, 27th AIAA Fluid Dynamics Conference, New Orleans, LA, June 18–20, 1996.
[3] Lorber, P. F., Stauter, R. C., Hass, R. J., Torok, M. S., and Kohlhepp, F. W., “Techniques
for Comprehensive Measurement of Model Helicopter Rotor Aerodynamics,” 50th Forum of
the American Helicopter Society, Washington D.C., May 11–14, 1994.
[4] Leishman, J. G., “Techniques for Flow Visualization of Helicopter Rotor Wakes,” 8th Inter-
national Symposium on Flow Visualization, Sorrento, Italy, 1998.
[5] Martin, P. B., Bhagwat, M. J., and Leishman, J. G., “Strobed Laser–Sheet Visualization
of a Helicopter Rotor Wake,” 2nd Pacific Symposium on Flow Visualization and Image
Processing, Honolulu, Hawaii, 1999. To be published in The Journal of Flow Visualization
and Image Processing.
[6] McCroskey, W. J., “Measurements of Boundary Layer Transition, Separation and Streamline
Direction on Rotating Blades,” NASA TN D–6321, pp. 1–35, 1971.
[7] Reda, D. C., Smith, R. W., Bryant, T. C., and Schluter, L. L., “Liquid Crystals for Sur-
face Shear Stress Visualization on Wind Turbine Airfoils,” Windpower: The International
Renewable Energy Conference, SAND–88–0521C, pp. 156–165, Hawaii, 1988.
[8] Britcher, C. P., “Experimental Aerodynamics,” Technical Notes, Old Dominion University,
1996.
[9] Obara, C. J., “Sublimating Chemical Technique for Boundary–Layer Flow Visualization in
Flight Testing,” Journal of Aircraft, Vol. 25, No. 6, pp. 493–498, 1988.
[10] Lyon, C., Selig, M., and Broeren, A., “Boundary Layer Trips on Airfoils at Low Reynolds
Numbers,” AIAA paper 97–0511, 35th Aerospace Sciences Meeting and Exhibit, Reno, NV,
January 6–10, 1997.
[11] Martin, P. B., Pugliese, G. J., and Leishman, J. G., “Laser Doppler Velocimetry Uncertainty
Analysis For Rotor Blade Tip Vortex Measurements,” AIAA CP 2000–0263, 38th Aerospace
Sciences Meeting and Exhibit, Reno, NV, 2000.
[12] Barrett, R. V. and Swales, C., “Realisation of the Full Potential of the Laser Doppler
Anemometer in the Analysis of Complex Flows,” Aeronautical Journal, Vol. 102, No. 1016,
1998, pp. 313–320.
[13] Boatwright, D. W., “Three–Dimensional Measurements of the Velocity in the Near Flow
Field of a Full–Scale Hovering Rotor,” EIRS–ASE–74–4, Mississippi State University, 1974.
[14] Martin, P. B., Leishman, J. G., Pugliese, G. J., and Anderson, S. L., “Stereoscopic PIV
Measurements in the Wake of a Hovering Rotor,” Proceedings of the 56th Annual Forum of
the American Helicopter Society, Virginia Beach, VA, May 2–4, 2000.

9th International Symposium on Flow Visualization, Heriot-Watt University, Edinburgh, 2000


Editors G M Carlomagno and I Grant 189–10

You might also like