Sei sulla pagina 1di 18

Fluid Dynamics of Interacting Blade Tip

Vortices With a Ground Plane


Timothy E. Lee∗ J. Gordon Leishman† Manikandan Ramasamy‡

Department of Aerospace Engineering


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

Abstract Nomenclature
Experiments were conducted using flow visualization and A Rotor disk area
phase-resolved digital particle image velocimetry (DPIV) c Blade chord
to examine the fluid dynamic behavior of a rotor wake as CT Rotor thrust coefficient, = T /ρAΩ2 R2
it interacted with a horizontal ground plane. The experi- CP Rotor power coefficient, = P/ρAΩ3 R3
ments were conducted using a small rotor with a ground Nb Number of blades
plane that could be moved to as close as 25% of the ro- P Rotor power
tor radius to the rotor plane. Visualization of the rotor r Radial distance
wake and the wake/ground plane interactions was per- R Radius of blade
formed by illuminating a seeded flow using a Nd:YAG T Rotor thrust p
laser, phase-locked to the rotational frequency of the ro- vh Hover induced velocity, = T /2ρA
tor. Two-component DPIV measurements were obtained ρ Flow density
at four rotor heights off the ground. Measurements at sev- σ Rotor solidity, = Nb c/πR
eral wake ages were obtained to examine the dynamic fea- ψ Azimuth angle
tures of the wake/surface interaction process. The results Ω Rotational speed of the rotor
showed that the overall downwash induced by the rotor
becomes an unsteady radial wall jet defined between the
ground plane and the rotor wake boundary. The experi- Introduction
ments have shown that the individual tip vortices create
The performance of all types of rotorcraft will be affected
substantial unsteady shear stresses on the ground plane,
by the interference of the ground or any other boundary
in some cases exceeding twice the time-averaged values.
that may constrain the flow into the rotor or otherwise al-
Viscous processes such as vortex straining and turbulence
ter the natural development of the rotor wake. The issue of
appear to be significant factors in understanding fluid dy-
“ground effect” is of concern in actual flight operations as
namics of the resulting flow near the ground. At the higher
well as in the ground testing of rotors. This rotor/surface
rotor heights the vortex filaments diffuse under the action
interference effect has long been recognized, but the de-
of viscosity and turbulence before reaching the ground. At
tailed aerodynamics at the rotor or with the developing
the lower rotor heights, turbulence in the developing wall
flow near the ground are still not fully understood. There
jet quickly shears the coherent vortices, accelerating their
are many dependencies and interdependencies affecting
diffusion. Corresponding measurements of rotor perfor-
rotor flows in-ground-effect operation, least of all being
mance showed the expected reduction in rotor power re-
the effects of rotor thrust and proximity of the rotor tip
quired and an increase in rotor thrust for in-ground-effect
path plane to the ground. Interference effects can be ob-
operations.
tained both in hover and forward flight, but the effects are
∗ Graduate generally strongest in the hovering state.
Research Assistant. telee@umd.edu
† Minta From a performance perspective, it is well known that
Martin Professor. leishman@umd.edu
‡ Assistant Research Scientist. mani@umd.edu when a rotor operates in-ground-effect (IGE) its thrust
Presented at the 64th Annual Forum of the American Helicopter is increased for a given power supplied—see Küssner
Society, Montréal, Canada, April 29–May 1, 2008. 2008
c by (Ref. 1), Zbrozek (Ref. 2), Betz (Ref. 3), Knight and
the authors except where noted. Published by the AHS Interna- Hefner (Ref. 4), Cheeseman and Bennett (Ref. 5),
tional Inc. with permission. Fradenburgh (Refs. 6, 7), Stepniewski and Keys (Ref. 8)

1
Figure 1: Schematics of the general behavior of the Figure 2: The fluid dynamics of the rotor wake
rotor wake when operating in: (a) Hover/very-low for- in-ground-effect is important for understanding the
ward speed, (b) Low-speed forward flight. problem of brownout. (Image courtesy of Optical Air
Data Systems.)
and Prouty (Ref. 9). Hayden (Ref. 10) summarizes
flight test measurements using the standardized technique culation at lower advance ratios, making reliable free-air
of Lewis (Ref. 11). The results suggest significant effects measurements of rotor performance difficult or impossible
for rotor heights of one diameter off the ground, where even if suitable corrections could be derived.
power required can be reduced by up to 30% relative to Besides the interference effects of the ground on ro-
out-of-ground-effect operation at the same thrust (i.e., at tor performance, the velocity of the induced flow near
the same aircraft weight). Most of this power reduction is the ground may reach high enough thresholds to lift up
induced in nature, but there is also some small reduction in loose surface materials such as dirt and sand, resulting in
profile power because the blades operate at slightly lower a problem that is known as “brownout.” Brownout can
angles of attack to produce the same thrust. Measurements quickly obscure the pilots field of view, and has resulted
have also shown secondary dependencies on blade load- in several helicopter accidents. Figure 2 shows how a ra-
ing, blade aspect ratio, and blade twist. dially and vertically expanding dust cloud is produced by
Flow visualization of rotors operating IGE have been a helicopter landing over a sandy surface. The details of
performed by Fradenburgh (Ref. 7), Taylor (Ref. 12), the development of the dust cloud, however, is known to
Curtiss et al. (Refs. 14, 13), and Light and Norman differ substantially from rotorcraft to rotorcraft and with
(Ref. 15). Consider the schematics shown in Fig. 1— surface type and condition. Brownout is fundamentally a
because the ground must always be a streamline to the complex physical phenomenon because it can potentially
flow, the rotor wake slipstream tends to rapidly expand ra- be affected by several interdependent parameters, includ-
dially away from the rotor as it approaches the ground. ing but not limited to: rotor disk loading, blade loading,
This interference effect on the rotor wake development al- number and placement of rotors, number of blades, blade
ters the induced velocity in the plane of the rotor, and so twist, blade tip shape, and fuselage shape.
affects the rotor thrust and power. At low forward speeds, Ground effect problems have received considerable at-
a region of flow recirculation can be formed upstream of tention from a modeling perspective. Cheeseman and
the rotor, which can increase induced power requirements Bennett (Ref. 5) replaced the rotor by a simple source
relative to hover or free-air conditions. Ground effect is with a corresponding image source to simulate ground ef-
usually negligible for advance ratios greater than about fect. Knight and Hefner (Ref. 4) and Rossow (Ref. 18)
0.10 at any rotor height off the ground. used a similar approach but with a vortex cylinder model.
Ground interference effects are also apparent in wind More recent studies of the flows associated with rotors
tunnel tests on rotors, where the effects of the tunnel walls operating IGE have used both Lagrangian vortex track-
can alter the flow at the rotor. Ganzer and Rae (Ref. 16) ing methods and Eularian grid-based computational fluid
and Lehman and Besold (Ref. 17) have studied the ef- dynamic (CFD) methods. Vortex wake models have
fects experimentally. For tunnel dimensions that are at been based on potential flow theory, assuming vorticity
least twice the diameter of the rotor, the effects of the wind conservation even near the ground plane. Such invis-
tunnel walls are small for operating advance ratios greater cid models are described by DuWaldt (Ref. 19), Saberi
than 0.1. Generally, it must always be assumed that the and Maisel (Ref. 20), Quackenbush and Wachspress
effects of the tunnel walls will lead to some flow recir- (Ref. 21), Graber et al. (Ref. 22), Itoga et al. (Ref. 23),

2
and Griffiths and Leishman (Ref. 24). While reason-
able predictions of wake behavior in hover and in forward
flight has been demonstrated, more limited success has
been achieved in correctly predicting rotor performance
IGE when compared to experimental measurements. This
reflects the inherent complexity and strongly viscous na-
ture of the flow field for IGE operations. Reynolds Av-
eraged Navier–Stokes (RANS) approaches such as that
discussed by Moulton et al. (Ref. 25) seem promising
for predicting such problems. Vorticity transport models
seem to have met with reasonable success in predicting
both ground effect aerodynamics (Ref. 26) and brownout
(Ref. 27).
The present paper describes results from an ongoing ex-
perimental study using phase-resolved digital particle im-
age velocimetry (DPIV) to examine in detail the fluid dy-
namic behavior of the rotor wake with its embedded blade
tip vortices as it interacts with a ground plane. The work
was a precursor to three-component, dual-plane, two- Figure 3: Drawing of the rotor and test stand.
phase flow measurements of rotors operating over sur-
faces covered with loose sediment. Such an understanding
is a prerequisite to better predicting the induced veloc-
ity field and performance of rotors operating in-ground-
effect, and eventually to further the understanding the
problem of “brownout.” Predictive models for brownout
are still being developed, and they require better valida-
tion of rotor aerodynamics in-ground-effect and of the
two-phase flow interactions of vortices with the sediment
bed. Ultimately, the real mechanisms responsible for par-
ticulate uplift in a non-steady turbulent flow environment
must be understood. The present work is focused on un-
derstanding using DPIV the near-wall fluid behavior as the
tip vortices interact with the radially expanding turbulent
jet flow on the ground plane.

Description of the Experiment


Rotor Setup Figure 4: Experimental setup of the rotor and the
ground plane.
The experiment was conducted with a small, two-bladed
rotor in which the rotor tip-path-plane plane was mounted The ground plane was a circular disk that was three
parallel to a ground plane. The rotor blades were made times the diameter of the rotor, and was painted flat black
of composite carbon fiber, and had circular arc, cambered to minimize reflections from the laser sheet. The ground
airfoil sections. The radius of the blade was 86 mm, with plane was mounted on a mechanical traverse so that it
a uniform chord of 19 mm with no taper. No blade twist could be moved to any distance from the rotor plane, and
was used. The rotor had a solidity σ of 0.14. A drawing also set to any inclination. In all of the present experi-
the rotor is shown in Fig. 3, and photograph of the rotor ments, the rotor tip-path-plane was adjusted so that it was
and ground plane setup is shown in Fig. 4. The rotational perfectly parallel to the ground plane.
speed of the rotor was measured with a Hall-effect sensor The rotor was operated at a rotational frequency of be-
mounted on the rotor shaft, which provided a once-per- tween 35 Hz and 60 Hz, giving a maximum tip speed of
revolution trigger signal to the data acquisition system. 32.42 ms−1 . The nominal operating tip Mach number and
The data acquisition system ensured precise phase lock- Reynolds number based on chord were up to 0.095 and
ing to less than ±0.1 degrees. 42,170, respectively.

3
Loads Balance
The rotor stand (Fig. 3) was mounted on a micro mass
balance capable of measuring the thrust with a precision
of ±0.1 grams. A linear reaction sensor was used to mea-
sure torque. The torque cell was separately calibrated by
using precision weights applied though a known arm. A
digital inclinometer was used to set the pitch angle of the
blades at the 75% span location.
At each rotational speed, torque and thrust data were
collected over a 5 second period. The rotor was also tested
at all conditions without the blades attached to the hub
to determine tares. These tares were subtracted from the
thrust and torque measurements taken with the blades at-
tached.
To guarantee measurement precision and repeatability Figure 5: Schematic explaining the laser sheet and
of the experiment, multiple tests were made at the same camera setup for flow visualization and DPIV.
pitch angles and rotational speeds. The measured thrust
was found to be within the instrument precision of the blade azimuth angles. The ground plane was positioned at
mass balance, while the measured torque had an uncer- several ground plane distances (z/R), typically z/R = 4.0,
tainty of ±3% of the mean value. The final balance data 3.0, 2.5, 2.0, 1.75, 1.5, 1.25, 1.0, 0.9, 0.8, 0.75, 0.7, 0.6,
were ensemble averaged, and used to find the thrust and 0.5, 0.4, and 0.25. These conditions are referred to in this
power coefficients. paper as the in-ground-effect (IGE) results. Baseline tests
were conducted without the ground plane present, which
are referred to as the out-of-ground effect (OGE) results.
Flow Visualization Not all types of measurements were made at all ground
plane heights.
For the flow visualization studies, a thermally produced
mineral oil fog was used as a flow tracer material. The av-
erage size of the tracer particles was about 0.20 microns in Digital Particle Image Velocimetry
diameter, and were determined to be small enough not to
have any significant tracking errors in the flow (Ref. 28). Two-component flow measurements were made using a 2
Judicious adjustment of the flow rate from seeder and the mega-pixel CCD camera for the DPIV imaging. A 532 nm
location of the outlet nozzle was required to introduce band pass filter prevented the camera from receiving any
concentrations of tracer particles at the locations needed to reflections other than the reflections from the seed par-
clearly identify the key flow structures such as the tip vor- ticles. A relatively small region of focus (100 mm-by-
tices and vortex sheets, as well as the developing boundary 100 mm) was chosen, which required a 60 mm lens with
layer flow on the ground. an f-stop of 2.8. The interrogation window size was 16-
The illumination of the flow was performed using a by-16 pixels with a 50% overlap. Measurements could be
Nd:YAG laser, the pulse frequency being phased synchro- made to within a few millimeters of the ground surface, at
nized with the rotational frequency of the rotor. An artic- which point laser reflections made it difficult to perform
ulated optical arm transmitted the laser light to the region successful cross-correlations with the DPIV images.
of interest. A 500 mm cylindrical lens at the end of the DPIV imaging requires a calibration process to incor-
arm converted the beam to a planar laser light sheet that porate the registration of the camera and its mapping from
was approximately 1 mm thick at its waist. Flow visu- the object plane onto the image plane to correct for distor-
alization images were taken with a Nikon 6.1 mega-pixel tions from variable magnification across the image. For
camera with 60 mm focal length lens. The camera was the present system, the single camera was mapped in the
positioned so that it was exactly perpendicular to the laser usual way for any PIV system. A nonlinear mapping func-
light sheet—see Fig. 5. tion was created from images of a calibration target. This
Many high-resolution images were taken at each rotor precision calibration target was made from regular grid of
operating state; only images showing all of the key fea- white dots on a black anodized background. A fiducial
tures of the wake and the interactions with the ground reference point on the target defined the origin for all the
plane were retained for further analysis. The phase of calibration images.
the trigger signal from the rotor could be altered to cap- For the digital particle image velocimetry (DPIV) ex-
ture images of the wake at different ages. i.e., at different periments, the entire test area was uniformly seeded with

4
the tracer particles before each sequence of measure- Spurious flow velocity vectors in the results were re-
ments. This is in contrast to the flow visualization ex- moved using standard deviation filter, followed by median
periments, which required concentrating tracer particles filter, and finally by mean filter based on 5-by-5 neighbor-
in regions of the flow needed to clearly identify the tip ing nodes. The spurious vectors were then replaced with
vortices, vortex sheets, and shear layers. interpolated vectors. Care was taken to minimize the num-
The correlation algorithm used for the DPIV was based ber of interpolated vectors. 100 DPIV image pairs were
on a deformation grid correlation method (Refs. 33, 34). acquired at each ground plane distance and for each wake
This procedure is similar to the more common recursive age. Any image that had spurious vectors of more than
technique, however, in this method the second window is 1% of the total was eliminated. Typically, this reduced the
deformed in that it is both sheared and translated instead total number of instantaneous velocity maps to between
of just undergoing a simple translation—see Fig. 6. The 93 and 95. The results were then simple phase-averaged
procedure starts with the correlation of an interrogation before further processing.
window of a defined size (say, 64-by-64), which is the
first iteration. After the mean displacement of that region
is estimated, the interrogation window of the displaced Results and Discussion
image is moved by integer pixel values for better corre-
lation in the second iteration. This third iteration starts by Rotor Performance Measurements
moving the interrogation window of the displaced image
by sub-pixel values based on the displacement estimated Measurements of rotor performance (thrust and power)
from second iteration. Following this, the interrogation were made for different rotor heights off the ground. Both
window is sheared twice (for integer and sub-pixel val- the thrust and power changed during these tests because
ues) based on the velocity magnitudes from the neighbor- they were made at a fixed blade pitch. From the power
ing nodes before performing a fourth and fifth iteration, polars of CT versus CP , the results were interpolated to
respectively. find the power required at a constant thrust, as well as the
Once the flow velocities were estimated after these five thrust produced at a constant power.
iterations, the window was split into four equal windows Figure 7 shows the relationship between the thrust pro-
(of size 32-by-32). These windows were moved by the av- duced in-ground-effect, TIGE , to that for out-of-ground-
erage displacement estimated from the final iteration (us- effect operation, TOGE , for a constant power. There was
ing a window size of 64-by-64) before starting the first little effect on thrust for heights of more than three ro-
iteration at this resolution. This procedure was continued tor radii, which is an expected result for any rotor based
until the resolution required to properly resolve the flow on all prior measurements of rotor performance IGE. For
field was reached. The second interrogation window was heights below one diameter, the rotor thrust increased
deformed until the particles remain at the same location more rapidly, rising to as high as 1.4TOGE at z/R =
after the correlation. This technique, especially the in- 0.25R. There was some small dependency on power set-
troduction of shear, has been shown (Ref. 34) to be the ting within the range that could be tested, but the results
most appropriate for measuring the high velocity gradi- were practically the same.
ents found inside rotor wake flows, and especially at the
near-wall flow region for a rotor operating IGE.

Figure 7: Measurements of rotor thrust at a constant


power show a substantial increase in thrust for opera-
tion near to the ground plane.
Figure 6: DPIV image processing technique.

5
OGE Operation

Figure 9 shows a representative flow visualization image


of the rotor wake for OGE operation. All of the basic fea-
tures of a hovering rotor wake are present here, including
the blade tip vortices and the turbulent vortex sheets. The
signature of the tip vortices appear as spiral bands of tracer
particles with clear voids at the vortex cores resulting from
the action of centrifugal and Coriolis accelerations on the
particles.
These vortices then convect along the slipstream bound-
ary that separate the higher velocities inside the rotor
wake from the external quiescent flow. Because a two
bladed rotor was used, successive tip vortices below the
Figure 8: Measurements of rotor power at a constant rotor are 180 degrees of age apart. For OGE operation,
thrust show a rapid decrease in power for operation the tip vortices remained distinct (with clear tracer par-
near to the ground plane. ticle voids) until they were about three rotor revolutions
old. As expected, the was some evidence of aperiodicity
Figure 8 shows the relationship between PIGE and POGE (i.e., deviations from the identical flow state from rotor
for constant rotor thrust. Nearly a 40% reduction in power revolution-to-revolution) in the rotor wake, this being an
was achieved at z/R = 0.25R. There was some evidence inherent characteristic of all types of helicoidal vortical
that the results were dependent on the operating state of wakes (Ref. 29).
the rotor, with the higher thrusts (and hence higher blade The presence of the turbulent sheets trailing from the
loadings) producing a slightly more significant effect on
rotor IGE performance.

Flow Visualization

Flow visualization was conducted for both rotor out-of-


ground-effect (OGE) and rotor in-ground-effect (IGE) op-
eration. Digital images of the rotor wake and surface flow
were taken for different rotor heights and at different blade
phase angles. Judicious seeding of the flow was required
to produce the best images that most clearly identified the
key flow features, including the blade tip vortices and tur-
bulent surface wall jet. The operating conditions for the
rotor at the different heights off the ground are given in
Table 1.

z/R CT /σ CP /σ Test conditions


∞ (OGE) 0.0917 0.01569 Flow viz.
2.5 0.0916 0.01582 Flow viz.
2 0.0946 0.01607 Flow viz.
1.5 0.0997 0.01629 Flow viz., DPIV
1.25 0.1030 0.01636 Flow viz.
1 0.1067 0.01646 Flow viz., DPIV
0.75 0.1099 0.01637 Flow viz.
0.5 0.1153 0.01636 Flow viz., DPIV
0.25 0.1329 0.01679 Flow viz., DPIV

Table 1: The operating conditions for the rotor at the


different heights off the ground. Figure 9: Representative flow visualization of the rotor
wake for hover out-of-ground-effect.

6
inboard parts of the blade can also be seen in Fig. 9, these can be directly compared with each other. Notice that the
being relatively thick because of the lower chord Reynolds wake distortion becomes particularly pronounced at the
number at the blades. These vortex sheets convect axially lower rotor heights, where the wake quickly expands radi-
more quickly than the blade tip vortices, and so interact ally outward within the first rotor revolution after its for-
with the older tip vortices as they age in the flow. The mation.
inner parts of the vortex sheets remain relatively distinct For the highest height of z/R = 2.5R, the initial devel-
to two or three rotor revolutions. The wake further down- opment of the wake was found to be similar to the OGE
stream becomes filled with the folded and merged vortex case—compare Fig. 11(a) with Fig. 9. However, the tip
sheets, and the tip vortices have spun down under the ac- vortices were noted to be axially little closer together than
tion of viscosity and turbulence. As a result, the far wake for the OGE case, and exhibited some evidence of pairing
and becomes relatively turbulent overall and begins to ex- or bundling, which is a common characteristic of wakes
pand slowly, eventually becoming a turbulent jet-like flow with reduced helicoidal pitch and more closely spaced
containing fairly diffused zones of residual vorticity. vortices.
By the time that the far wake approaches the ground
plane it had already become relatively turbulent and the
IGE Operation
tip vortices have diffused under the action of viscosity and
In all cases, it was found that the presence of the ground turbulence. The flow near the ground, therefore, becomes
plane significantly altered the geometry of the developing a developing turbulent wall jet that is relatively free of co-
rotor wake compared to the OGE case (Fig. 9), the wake herent vortical flow structures but was notably unsteady
expanding (stretching) significantly radially outward as it (aperiodic). This jet expands radially away from the rotor,
approaches the ground surface. A simple schematic of the and exhibits the features of both a boundary layer and a
general flow developments is shown in Fig. 10. On the free shear layer. Quantification of the resulting flow veloc-
ground plane, the flow becomes a wall jet with a bound- ities and the boundary layer profiles was measured using
ary layer close to the surface and a developing shear flow using DPIV, which is discussed later.
region above with significant turbulent mixing. Figure 11(b) shows flow visualization results for z/R =
Figure 11 shows flow visualization results for eight ro- 2.0R. In this case, the older blade tip vortices (i.e., 3 rotor
tor heights above the ground plane. The images were revolutions old) approach the ground plane before show-
taken for the same blade phase and are scaled in both the ing any evidence of spinning down and diffusing. As they
vertical and horizontal dimensions so that all the images interact with the ground, the vortices undergo a stretching
process (Refs. 30, 31), which results in vorticity intensifi-
cation that counters normal diffusive behavior, at least for
a while. As a result, in this case the tracer particle void
signatures of the tip vortices were apparent in the flow vi-
sualization to about five rotor revolutions. The initial de-
velopment of the surface wall jet, therefore, is different in
that it now contains significant embedded vorticity from
the tip vortices and, to a lesser extent, additional vorticity
contributions from the vortex sheets.
Figure 11(c) shows results for z/R = 1.5R. These re-
sults are similar to the z/R = 2.0R case, but the tip vor-
tices were only visible up to about four rotor revolutions
old. There were now significantly large eddies present in
the wall surface flow. Lowering the height to z/R = 1.25R
produced significant pairing of the older vortices as they
approached the wall—see Fig. 11(d). This pairing was
periodic, the vortices spinning around each other and un-
dergoing some evidence of viscous merging as they were
sheared in the developing wall flow.
The vortex sheets trailed from the inner parts of the
blades were also more evident near the wall in this case.
The vortex sheets were seen to remain embedded inside
the developing boundary layer, at least initially. The re-
Figure 10: General development of the flow below the sulting wall jet for this condition seemed to have signifi-
rotor in ground effect operation. cantly higher levels of vorticity with a variety of eddies at

7
(a) Rotor height z/R = 2.5 (b) Rotor height z/R = 2.0

(c) Rotor height z/R = 1.5 (d) Rotor height z/R = 1.25

(e) Rotor height z/R = 1.0 (f) Rotor height z/R = 0.75

(g) Rotor height z/R = 0.5 (h) Rotor height z/R = 0.25
Figure 11: Flow visualization of the rotor wake for several rotor heights off the ground plane: (a) z/R = 2.5, (b) z/R
= 2.0, (c) z/R = 1.5, (d) z/R = 1.25, (e) z/R = 1.0, (f) z/R = 0.75, (g) z/R = 0.5, and (h) z/R = 0.25.

8
various scales. The resulting flow suggested the develop- Time-Averaged Flow Field Results
ment of a marginally thicker turbulent wall jet and shear
flow as it expanded away from the rotor. Time-averaged velocity field results are shown in Fig. 12
in the form of flow vectors on a color-contoured back-
Lowering the rotor to z/R = 1.0R off the ground showed ground for the four rotor heights off the ground. Only
that the older tip vortices were now being significantly every other flow vector has been plotted in these cases
stretched as they came into proximity with the ground— to prevent congestion. Because the rotor thrust varies
see Fig. 11(e). In this case, the older tip vortices were ac- for all four IGE cases, the results shown here have been
tually stretched to about twice their initial length, which normalized by the momentum theory factor for the av-
obviously results in considerable vorticity intensification. eragep induced velocity, i.e., by dividing by the factor
This stretching process decreases the viscous core size of vh = T /2ρA. This approach allows the measurements
the vortices and increases their peak swirl velocities rela- to be better compared on the basis of constant rotor thrust.
tive to those at the previous rotor revolution.
In all four cases, it is obvious that the rotor induced flow
The reduction of the size of the seed void to values close is forced to expand radially outward as a wall jet as pre-
to those seen after the first rotor revolution suggests that viously shown with the flow visualization. The quiescent
the stretching effect (which conserves circulation) is in- flow (blue zones) outside of the rotor wake boundary is
deed a powerful mechanism in prolonging the life of the apparent. For z/R = 1.5, as shown in Fig. 12(a), it can be
tip vortices during their entrainment into the surface wall seen that the flow velocities within the wake boundary are
jet. However, it should be noted that the seed void does relatively high (about twice as large than the value of vh ,
not represent the dimensions of the viscous core but sim- which is expected based on momentum theory considera-
ply an equilibrium position for the rotating seed particles. tions). These high velocities reach out to about r/R=1.5 at
Nevertheless, the presence of a seed void suggests high this rotor height, and thereafter drop quickly. The thick-
vorticity and is evidence that the tip vortex is still present ness of the wall jet (including the shear layer) is about
in the flow. The signatures of the vortex sheets can also 0.2R, and the thickness slowly increases as the wake ex-
be seen in the resulting sheared flow zone at the ground pands radially to distances beyond r/R = 1.5.
surface flow, perhaps as far away as two rotor radii. For the z/R = 1.0 rotor height case, as shown in
The remaining three rotor heights of z/R = 0.75R, 0.5R Fig. 12(b), the overall flow was found to be similar to
and 0.25R—see Figs. 11(f), (g) and (h), respectively—are the previous IGE case, but now a much higher local-
relatively low rotor heights that would encountered with ized flow velocity was produced at the wall for radial
a conventional helicopter, but nevertheless provide a use- stations between r/R=1.2 to 1.8. The highest flow ve-
ful set of results in helping to understand the fundamen- locities exceeded 2vh and occurred at about r/R = 1.5.
tal flow physics of rotor IGE operations. These cases all Again, the surface velocities dropped off quickly beyond
resulted in relatively low helical spacing between succes- r/R = 1.6 and the wall jet expanded. Continuity consider-
sive turns of the rotor wake, and so significant tip vor- ations would dictate that the centerline velocities should
tex pairing was produced as the wake began to interact decrease inversely with distance in the fully developed
with the ground plane. The tip vortices are significantly wall jet far away from the rotor.
stretched soon after their formation, and appeared to dif- For the two lowest rotor heights of z/R = 0.5 and 0.25,
fuse their vorticity at relatively young wake ages as they as shown in Fig. 12(c) and (d), respectively, the aver-
interacted with the ground. The resulting flow near the age flow velocities might be expected to increase because
ground seemed relatively free of vorticity but was clearly of the closer proximity of the rotor to the ground plane.
very turbulent. However, interestingly enough the results show that the
velocities remain relatively low (for a nominally constant
rotor thrust), the local peak velocity values actually de-
creasing from the results at the higher rotor height of z/R
= 1.0. For z/R = 0.5 there are two zones where the local
DPIV Flow Field Measurements velocities exceeded 1.5vh , one at about r/R = 1.25 and the
other at about r/R = 1.85.
DPIV measurements were made for four rotor heights off These results reflect the overall nonlinear effects of the
the ground, namely z/R = 1.5, 1.0, 0.5 and 0.25. The ground on the flow field, where the resulting outflow from
DPIV results were post-processed to obtain both phase- the rotor at the ground plane (i.e., the groundwash) de-
averaged and time-averaged velocity vectors, and the data pends on the influence of the average rotor induced down-
were also processed in terms of velocity profiles measured wash combined with the local effects of the blade tip vor-
normal to the ground plane at different radial distances tices and the vortex sheets. Much of the local increase
from the rotational axis of rotor. in flow velocity near the ground seen in Fig. 12(b) is be-

9
(a) Rotor height z/R = 1.5 (b) Rotor height z/R = 1.0

(c) Rotor height z/R = 0.5 (d) Rotor height z/R = 0.25

Figure 12: DPIV results of the time-averaged velocity field through a radial plane for different rotor heights off the
ground: (a) Rotor height at z/R = 1.5, (b) Rotor height at z/R = 1.0, (c) Rotor height at z/R = 0.5, (d ) Rotor height
at z/R = 0.25.

cause of the persistence of the tip vortices to relatively old number of blades (i.e., the number of vortices reaching
wake ages as a direct result of vortex stretching effects— the ground per unit time).
see Figs. 11d and (e).
At the higher rotor heights the vortex filaments spin Phase-Averaged Flow Field Results
down and diffuse under the action of viscosity and tur-
bulence before they can ever reach the ground; at the Phase-averaged velocity field results are shown in Fig. 13
lower rotor heights the turbulence in the developing wall for a rotor height of z/R = 1.0R. These images represent
jet quickly shears the vortices, which causes them to lose a snapshot of the flow field for different blade positions.
their coherent flow structures within only 1.5 rotor revolu- From the sequence of images, it can be seen how the tip
tions of wake age. It is these two competing mechanisms vortices convect along the slipstream boundary and into
that suggest that for any rotor there will be some interme- the developing boundary layer flow at the ground plane.
diate rotor height off the ground where the groundwash Notice that distinguishable vorticity exists even after
velocities will reach a maximum. This height will depend the vortices become several rotor revolution old and have
on the average downwash (a thrust and/or disk loading ef- begun to interact with the developing turbulent flow on
fect) produced by the rotor, the strength of the tip vortices the ground. It is apparent that the action of the wall jet
(a blade loading effect for a constant tip speed), and the velocity, as well as the turbulent wake sheets and other

10
(a) Phase angle = 0◦ (b) Phase angle = 30◦

(c) Phase angle = 60◦ (d) Phase angle = 90◦

(e) Phase angle = 120◦ (f) Phase angle = 150◦


Figure 13: Phase-averaged velocity field for z/R = 1.0: (a) Phase angle = 0◦ ,
(b) Phase angle = 30◦ , (c) Phase angle =
◦ ◦ ◦ ◦
60 , (d) Phase angle = 90 , (e) Phase angle = 120 , and (f) Phase angle = 150 .

turbulence in the developing flow, begins to shear the vor- velocities near the ground with considerable fluctuations
tices. The vorticity contained in the vortices, therefore, as each vortex sweeps by.
quickly becomes more diffused. Nevertheless, it is appar- As previously mentioned, at higher rotor heights the
ent that closer to the rotor the vortices induce high local vortex filaments undergo diffusion before they reach the

11
(a) Rotor height z/R = 1.5 (b) Rotor height z/R = 1.0

(c) Rotor height z/R = 0.5 (d) Rotor height z/R = 0.25
Figure 14: Phase-averaged velocity field for different rotor heights off the ground at a phase angle of 0◦ : (a) Rotor
height at z/R = 1.5, (b) Rotor height at z/R = 1.0, (c) Rotor height at z/R = 0.5, (d ) Rotor height at z/R = 0.25.

(a) Phase angle = 0◦ (b) Phase angle = 30◦

(c) Phase angle = 60◦ (d) Phase angle = 90◦

(e) Phase angle = 120◦ (f) Phase angle = 150◦

Figure 15: Detail of the tip vortex pairing and shearing in the developing flow near the ground flow.

12
ground, so their effect on the development of the wall jet The boundary layer region is relatively thin, typically
is relatively small. As the rotor height from the ground only a few millimeters or less than 5% of rotor radius. The
plane is lowered however, the tip vortices begin to reach small thickness of the boundary layer poses some chal-
the ground before they spin down and so begin to induce lenges for the DPIV because of the need for a very fine
higher local mean surface velocities as well as higher ve- spatial resolution near the surface while also minimizing
locity fluctuations in the flow—see the sequence in Fig. 14 laser reflections from the surface. The results show that
and compare with the time-averaged results in Fig. 12. measurements are feasible well into the boundary layer
If the vortices reach the ground, then they are stretched and that its profile can be properly resolved using DPIV.
radially, which increases their core vorticity and induced Directly below the rotor, as shown in Fig. 17(a), the ra-
velocities—see Figs. 14(a) and (b). The persistence of the dial flow has already become established, and by r/R =
vortices to four or five rotor revolutions has already been 1.0 (Fig. 17(b)) the wall jet has already become well de-
noted in Fig. 11(b). veloped. Notice that the highest surface flow velocities are
Figures 14(c) and (d) show that at the lower rotor obtained for a rotor height of z/R = 1.0. It is these veloc-
heights off the ground, the tip vortices are sheared as they ities and velocity gradients that are responsible for shear-
enter the wall jet. This shearing action quickly diffuses ing the flow in the tip vortices. Moving radially outwards
vorticity, and the signatures of the vortices are only appar- (Fig. 17(c) through (f)), the wall jet becomes thinner and
ent for up to two rotor revolutions. Nevertheless, at the reaches its peak velocity at about r/R = 1.5. Thereafter,
lowest rotor height the wall jet is seen to contain signifi- the peak velocity decreases and the boundary layer be-
cant fluctuating velocities and turbulence. comes slightly thicker.
The process is detailed in Fig. 15, which shows DPIV
results at the ground plane for a series of increasing times.
Notice that the vortices begin to pair and merge, but are Phase-Averaged Surface Velocity Profiles
also sheared in the developing flow. The merging results Phase-averaged velocity profiles for a rotor height of z/R
in some locally high induced velocities on the surface, but = 1.0 are shown in Fig. 18. Notice in these cases, that the
these are also diminished by the shearing action. The net signatures of the convecting tip vortices are apparent, at
effect is a zone of residual vorticity within the developing least up to about r/R = 1.25. The largest unsteady fluctua-
wall jet that produces significant velocities at the wall. tions near the ground plane occur at r/R = 1.0, where there
are local excursions of about 50% from the time-averaged
Time-Averaged Surface Velocity Profiles values. The swirl velocity induced by the tip vortex inside
the wall jet is in the same direction as the groundwash ve-
The DPIV results were further processed to extract time- locity and results in overall augmentation of the local flow
averaged radial velocity profiles as a function of distance velocities near the surface.
normal to the ground surface for several rotor heights off This mechanism of how the wake interacts with the
the ground. Measurements at six radial stations at r/R = ground is important because it will affect the wall shear
0.8, 1.0, 1.25, 1.5, 1.75 and 2.0 were examined. It was stress and so, ultimately, the “brownout” problem alluded
found that developing jet-like flow near the ground takes to in the introduction of this paper through the changes to
on the features of both a turbulent boundary layer and a the local thresholds that will govern the uplift of particles.
shear region, as shown in the schematic in Fig. 16. Underestimation of the overall decay rate of a blade tip
vortex, for instance, will result in increased entrainment
of loose surface sediment, and more extended regions of
sediment lift-up. The mechanisms of shear and turbulent
diffusion, however, will likely increase the mixing of any
particles near the ground, again affecting the number of
particles that are lofted into the flow. The shear fluctu-
ations on the ground clearly become much smaller with
increasing radial distance from the rotor.
Figure 19 shows the near-wall phase-averaged velocity
profiles for a rotor height of z/R = 1.0. In this case, re-
sults are for the zone within 0.4 rotor radius (i.e., 34.4 mm
or 1.35 inches) off the ground plane. The flow could be
measured successfully down to about 0.015 radius (i.e.,
Figure 16: The jet-like flow at the ground exibits the 1.29 mm or 0.051 inches) from the ground plane, at which
features of a thin turbulent boundary layer at the sur- point surface reflections of the laser light (even from a
face and a shear region above. matt-black surface) made it difficult to perform successful

13
(a) r/R = 0.8 (b) r/R = 1.0

(c) r/R = 1.25 (d) r/R = 1.5

(e) r/R = 1.75 (f) r/R = 2.0

Figure 17: Time-averaged velocity profiles measured using DPIV at various radial distances from the rotor: (a) r/R
= 0.8, (b) r/R = 1.0, (c) r/R = 1.25, (d) r/R = 1.5, (e) r/R = 1.75, (f) r/R = 2.0.

cross-correlations on the DPIV images. Conclusions


Experiments have been conducted using flow visualiza-
Notice that the profiles near the ground show the fea- tion and phase-resolved digital particle image velocime-
tures of a wall jet, with the boundary layer flow being rel- try (DPIV) to examine the fluid dynamic behavior of
atively thin. The boundary layer thickness was estimated a rotor wake, with its embedded blade tip vortices and
in each case by using the location where the surface ve- vortex sheets, as it interacted with a horizontal ground
locity reached its peak, the results of which are shown in plane. Two-component DPIV measurements were ob-
Fig. 20. Notice that the boundary layer becomes thinner tained at four rotor heights off the ground, and allowed
moving away from the rotor. While the initial thickness the quantification of the flow within a few millimeters of
depends on the rotor height off the ground, the final thick- the ground surface. Measurements at several wake ages
ness is the same in each case. were also obtained to examine the dynamic features of the

14
(a) r/R = 0.8 (b) r/R = 1.0

(c) r/R = 1.25 (d) r/R = 1.5

(e) r/R = 1.75 (f) r/R = 2.0

Figure 18: Phase-averaged velocity profiles measured using PIV at various radial distances for a rotor height of z/R
= 1.0: (a) r/R = 0.8, (b) r/R = 1.0, (c) r/R = 1.25, (d) r/R = 1.5, (e) r/R = 1.75, (f) r/R = 2.0.

wake/surface interaction process. wash induced by the rotor is forced to escape as a jet
The following conclusions have been drawn from the between two boundaries defined by the ground plane
work: and the wake boundary, the latter which at least ini-
tially is defined by the path of the tip vortices. As
1. Measurements of rotor performance showed the a re-
the vortices age in the flow, the wake boundary be-
duction in rotor power required for a constant thrust
comes much less distinct as it mixes with the devel-
and an increase in the rotor thrust for a constant
oping flow on the ground and produces a shear zone.
power during in-ground-effect operations. These re-
sults are typical for any hovering rotor operating in
3. DPIV measurements of the flow were successfully
ground effect, and do not appear to be affected sub-
made near the ground plane, and down to about 1.5%
stantially by scale.
of the rotor radius. The resulting flow velocity at the
2. Flow visualization suggests that the overall down- ground plane closely resembled the form of a classic

15
(a) r/R = 0.8 (b) r/R = 1.0

(c) r/R = 1.25 (d) r/R = 1.5

(e) r/R = 1.75 (f) r/R = 2.0

Figure 19: Near-wall phase-averaged velocity profiles at various radial distances for a rotor height of z/R = 1.0: (a)
r/R = 0.8, (b) r/R = 1.0, (c) r/R = 1.25, (d) r/R = 1.5, (e) r/R = 1.75, (f) r/R = 2.0.

wall jet, which shows the characteristics of a turbu- tices were also sheared in the developing wall jet.
lent boundary layer very near the wall and a shear The residual velocity field was still sufficient to in-
flow away from the wall. duce significant fluctuating shear on the wall.
4. At the higher rotor heights the vortex filaments spin 6. The competing flow mechanisms (diffusion, stretch-
down (diffuse) under the action of viscosity and tur- ing and shearing) suggest that there will be some in-
bulence before they can reach the ground. At the termediate rotor height off the ground where the sur-
lower rotor heights the turbulence in the develop- face flow (groundwash) velocities will reach a maxi-
ing wall jet quickly shears the vortices, which causes mum. This height will depend on the average down-
them to diffuse quickly. wash (a disk loading effect) produced by the rotor,
5. In some cases, pairing and merging of the blade tip the strength of the tip vortices (a blade loading ef-
vortices occurred near the ground. However, the vor- fect), and the number of blades.

16
9 Prouty, R. W., “Ground Effect and the Helicopter,”

AIAA Paper 85-4034, AIAA/AHS/ASEE Aircraft Design


Systems and Operations Meeting, Colorado Springs, CO,
October 14–16, 1985.
10 Hayden, J. S., “The Effect of the Ground on He-
licopter Hovering Power Required,” 32th Annual Na-
tional V/STOL Forum of the American Helicopter Soci-
ety, Washington DC, May 10–12, 1976.
11 Lewis, R. B., “Army Helicopter Performance Trends,”
Journal of the American Helicopter Society, Vol. 17, (2),
1972, pp. 15–23.
12 Taylor, M. K., “A Balsa-Dust Technique for Air-Flow
Figure 20: Measurements of the radial development Visualization and Its Application to Flow through Model
of the boundary layer thickness as a function of rotor Helicopter Rotors in Static Thrust,” NACA Technical
height off the ground. Note 2220, 1950.
Acknowledgments 13 Curtiss,H. C., Erdman, W., and Sun, M., “Ground Ef-
fect Aerodynamics,” Vertica, 11 (1/2), 1987, pp. 29–42.
This research was partly supported by the Army Research
Office (ARO) under grant W911NF0610394 and partly 14 Curtiss, H. C., Sun, M., Putman, W. F., and Hanker,
by the Multi-University Research Initiative under Grant E. J., “Rotor Aerodynamics in Ground Effect at Low Ad-
W911NF0410176. Additional support for the first author vance Ratios,” Journal of the American Helicopter Soci-
was provided by the Minta Martin Aeronautical Fund. ety, Vol. 29, (1), 1984, pp. 48–55.
15 Light,J. S. and Norman, T., “Tip Vortex Geometry of
References a Hovering Helicopter Rotor in Ground Effect,” 45th An-
nual Forum of the American Helicopter Society, Boston,
1 Küssner,H. G., “Helicopter Problems,” NACA TM MA, May 22–24, 1989.
827, January 1937. 16 Ganzer, V. M. and Rae, W. H., “An Experimental Inves-

2 Zbrozek, tigation of the Effect of Wind Tunnel Walls on the Aero-


J., “Ground Effect on the Lifting Rotor,”
dynamic Performance of a Helicopter Rotor,” NASA TN
British ARC R & M 2347, 1947.
D-415, 1960.
3 Betz,
A., “The Ground Effect on Lifting Propellers,” 17 Lehman, A. F. and Besold, J. A., “Test Section Size
NACA TM 836, 1937.
Influence on Model Helicopter Rotor Performance,” US-
4 Knight, AAVLABS TR 71-6, 1971.
M. and Hefner, R. A., “Analysis of Ground Ef-
fect on the Lifting Airscrew,” NACA TN 835, 1941. 18 Rossow, V. J., “Effect of Ground and/or Ceiling Planes
5 Cheeseman, I. C. and Bennett, W. E., “The Effect of the on Thrust of Rotors in Hover,” NASA Technical Memo-
randum 86754, 1985.
Ground on a Helicopter Rotor in Forward Flight,” ARC R
& M 3021, 1955. 19 DuWaldt, F. A., “Wakes of Lifting Propellers (Rotors)

in-ground-effect,” Cornell Aeronautical Laboratory, Re-


6 Fradenburgh, E. A., “The Helicopter and the Ground port CAL No. BB-1665-S-3, November 1966.
Effect Machine,” Journal of the American Helicopter So-
ciety, Vol. 5, (4), 1960, pp. 26–28. 20 Saberi,H. A. and Maisel, M. D., “A Free-Wake Rotor
Analysis Including Ground Effect,” 43rd Annual Forum
7 Fradenburgh, E. A., “Aerodynamic Factors Influencing of the American Helicopter Society, St. Louis, MO, May
Overall Hover Performance,” AGARD-CP-111, 1972. 18–20, 1987.
8 Stepniewski, W. Z. and Keys, C. N., Rotary-Wing Aero- 21 Quackenbush, T. R. and Wachspress, D. A., “Enhance-
dynamics, Dover Publications, New York, NY. Part II, ments to a New Free Wake Hover Analysis,” NASA CR
Chapters 2 and 3, 1984. 177523, April 1989.

17
22 Graber, A., Rosen, A., and Seginer, A., “An Investi- 33 Scarano, F., “Iterative Image Deformation Methods

gation of a Hovering Rotor in Ground Effect,” Proceed- in PIV,” Measurement Science and Technology, Vol. 13,
ings of the 16th European Rotorcraft Forum, Glasgow, 2002, pp. R1–R19.
September 18–20, 1990. 34 Ramasamy, M., and Leishman, J. G., Benchmarking
23 Itoga,N., Nagashima, T., Iboshi, N., Kawakami, S., Particle Image Velocimetry with Laser Doppler Velocime-
Prasad, J. V. R., and Peters, D. A., “Numerical Anal- try for Rotor Wake Measurements, AIAA Journal, Vol. 45,
ysis of Ground Effect for a Lifting Rotor Hovering at (11), November 2007, pp. 2622–2633.
Close Proximity to Inclined Flat Surface,” American He-
licopter Society Specialists’ Meeting on Advanced Heli-
copter Technology and Disaster Relief, Gifu, Japan, April
21–23, 1999.
24 Griffiths,
D. A., Ananthan, S., and Leishman, J. G.,
“Predictions of Rotor Performance in Ground Effect Us-
ing a Free-Vortex Wake Model, Journal of the American
Helicopter Society, Vol. 49, (3), October 2005.
25 Moulton,M., O’Malley, J. A., and Ranjagopalan, G.,
“Rotorwash Prediction Using an Applied Computational
Fluid Dynamics Tool,” 60th Annual Form of the American
Helicopter Society, Baltimore, MD, June 7–10, 2004.
26 Brown, R. and Whitehouse, G., “Modelling Rotor
Wakes in Ground Effect,” Journal of the American He-
licopter Society, Vol. 49, (3), October 2004, pp. 238–249.
27 Philips,
C., and Brown, R., “Eulerian Simulation of the
Fluid Dynamics of Helicopter Brownout,” 64th Annual
Forum Proceedings of the American Helicopter Society
International, Montréal Canada, April 29–May 1, 2008.
28 Leishman, J. G., “On Seed Particle Dynamics in
Tip Vortex Flows,” Journal of Aircraft, Vol. 33, (4),
July/August 1996, pp. 823–825.
29 Bhagwat, M. J. and Leishman, J. G., “Stability Analy-
sis of Helicopter Rotor Wakes in Axial Flight, Journal of
the American Helicopter Society, Vol. 45, (3), July 2000,
pp. 165–178.
30 Ananthan, S., Leishman, J. G., and Ramasamy, M.,
“The Role of Filament Stretching in the Free-Vortex Mod-
eling of Rotor Wakes,” American Helicopter Society 58th
Annual National Forum, Montréal, Canada, June 11–13,
2002.
31 Ramasamy, M., and Leishman, J. G., “The Interdepen-
dence of Straining and Viscous Diffusion Effetcs on Vor-
ticity in Rotor Flow Fields,” American Helicopter Society
59th Annual National Forum Phoenix, Arizona, May 6–8
2003.
32 Leishman,J. G., “On the Aperiodicity of Helicopter
Rotor Wakes, Experiments in Fluids, Vol. 25, 1998,
pp. 352–361.

18

Potrebbero piacerti anche