You are on page 1of 13

JOURNAL OF AIRCRAFT

Vol. 54, No. 5, September–October 2017

Scaling and Configuration Effects on Helicopter Rotor Hub


Interactional Aerodynamics

David Reich,∗ Steven Willits,† and Sven Schmitz‡


Pennsylvania State University, University Park, Pennsylvania 16802
DOI: 10.2514/1.C034250
A 1:17 scale model of a notional rotor hub of a large helicopter was tested in the Pennsylvania State University
Applied Research Laboratory 12 in. test-section water tunnel. Objectives of the experiment were to quantify the
effects of Reynolds number, advance ratio, and hub geometry configuration on the drag and shed wake of the rotor
hub. A range of flow conditions was tested, with hub-diameter-based Reynolds numbers ranging from 1.0 × 106 to
2.6 × 106 and advance ratios ranging from 0.2 to 0.6, as well as nonrotating cases. Five hub geometry configurations
were tested with various combinations of components including blade stubs, spiders, scissors, the swashplate, pitch
links, and beanie fairing. Measurements included the steady and unsteady hub drag and particle image velocimetry at
two downstream locations. Results include time-averaged and phase-averaged analyses of the unsteady drag and
Downloaded by 122.172.77.152 on December 15, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.C034250

wake velocity. A strong dependence of the steady and unsteady hub drag and wake on the advance ratio, Reynolds
number, and configuration was observed, demonstrating the importance of adequate Reynolds number scaling for
model helicopter rotor hub tests.

Nomenclature diameter, referred to as Rehub , are shown in Fig. 1 for three selected
A = hub frontal area, m2 helicopters at an advance ratio μ of 0.2. It is apparent that Rehub >
CD0 = fluctuating drag coefficient, D∕qA 3 × 106 for all configurations. While it is convenient to treat the hub
D = drag, N as a single bluff-body system at a high Reynolds number, it cannot be
q = freestream dynamic pressure, equal to 1∕2ρU2∞ assumed that either the near-field flow contributions to interactional
R = hub radius, m aerodynamics of the hub–pylon configuration drag or the long-age
RRotor = rotor radius, m wake is independent of Rehub .
Rehub = hub-diameter-based Reynolds number, equal to It is evident that the hub-component cross-sections experience
U∞ 2Rhub ∕ν subcritical and supercritical flow over the rotor azimuth. Past tests
U∞ = freestream velocity, m/s conducted by Keys and Rosenstein [3] further confirm that 1) rotation
u = local streamwise velocity, m/s does not significantly change the transition Reynolds numbers of hub
v = local spanwise velocity, m/s cross-sections and 2) transition from subcritical to supercritical flow
w = local vertical velocity, m/s occurs in less than 0.001 s, or 0.5% of one rotor rotation and,
X = streamwise coordinate axis therefore, without significant lag over the rotor azimuth.
Y = spanwise coordinate axis Furthermore, the effect of increasing advance ratio μ at constant
Z = vertical coordinate axis Rehub was observed to decrease the hub drag coefficient, which is
μ = advance ratio, equal to U∞ ∕2πΩRRotor consistent with the hub cross-sections operating in supercritical flow
ν = fluid kinematic viscosity, m2 ∕s over an extended range of azimuth angles. It must be noted, however,
ρ = fluid density, kg∕m3 that these results were obtained for a highly simplified two-bladed
Ω = hub rotation rate, Hz rotor hub consisting of a shaft and two circular hub shanks. Hence,
these results might not be transferrable to more complex hub
configurations typical of the helicopters shown in Fig. 1.
In addition to what was already mentioned [3], there have been
I. Introduction several studies throughout the past half-century that have alluded to
Rehub dependence [14]. Churchill and Harrington [4] first reported a
D RAG is a primary limiter of forward-flight speed and fuel
economy of the helicopter [1,2]. It is well known to the
rotorcraft community that the rotor hub assembly accounts for up to
Rehub dependence [4]. Montana [15] showed hub drag Rehub
dependence in the context of how sensitive the performance of
30% of the total vehicle parasite drag [1,3,4]. In addition to its adverse various hub fairing designs was to Rehub . Logan et al. [16] conducted
contribution to the total vehicle drag, unsteady flow separation from wind tunnel tests on two production rotor hub designs and asserted
the rotor hub is known to cause flow separation aft the engine pylon, Rehub dependence up to Rehub  1 × 106 . Felker [17] asserted hub
thus further increasing the vehicle drag [5–9]. Moreover, flow drag coefficient Rehub dependence for 1 × 106 < Rehub < 3 × 106 in
structures shed by the rotor hub assembly are known to persist far wind tunnel tests of the XH-59A rotor hub. Recently, Shenoy et al.
downstream and interact with the empennage, which may affect [18] computationally showed scaling effects on a rotor hub over a
handling qualities and cause other undesired effects such as tail wag large range of Reynolds numbers. Numerous other studies [19–22]
and tail shake [10–13]. The Reynolds numbers based on the hub (including but not limited to these) have examined Re dependence for
generic bluff-body shapes. These studies reinforce the fact that the
Received 7 October 2016; revision received 16 December 2016; accepted rotor hub, which is essentially a composition of bluff bodies, has a
for publication 4 February 2017; published online Open Access 8 May 2017. quantifiable dependence of drag and wake flows on the Reynolds
Copyright © 2017 by the American Institute of Aeronautics and Astronautics, number.
Inc. All rights reserved. All requests for copying and permission to reprint Therefore, the combined effects of Rehub and μ on the hub drag and
should be submitted to CCC at www.copyright.com; employ the ISSN 0021- the resultant near-field flow are important areas of research toward
8669 (print) or 1533-3868 (online) to initiate your request. See also AIAA enhanced physical understanding of rotor hub flows, the interactional
Rights and Permissions www.aiaa.org/randp.
*Ph.D. Candidate, Department of Aerospace Engineering. Student
aerodynamics among the various hub components, and the
Member AIAA. interaction of the hub flow as a whole with the engine pylon. The

Head, Fluids Machinery Department, Applied Research Laboratory. objective of this paper is to present a systematic experimental study of

Associate Professor, Department of Aerospace Engineering. Senior the combined effects of Rehub and μ on the time-averaged hub drag,
Member AIAA. the unsteady content of the hub drag force, and the nature of the flow
1692
REICH, WILLITS, AND SCHMITZ 1693

Fig. 1 Examples of hub-based Reynolds numbers Rehub at μ  0.2.

structures emanating from the various hub components. Several


variants of a representative rotor hub configuration are used for this
study. All experiments were conducted in the 12-in.-diam water
tunnel at the Applied Research Laboratory (ARL) at Pennsylvania
Downloaded by 122.172.77.152 on December 15, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.C034250

State University (PSU). The goal of this work is to suggest a


minimum Rehub , or Reynolds scale, that is required for model-scale
testing in order to obtain relevant results for a scaled application to
future large and high-speed helicopters.

II. Experimental Methods


A. Facility and Test Rig
The experiment was conducted in the 12-in.-diam test-section
water tunnel located at PSU ARL. The water tunnel is schematically
shown in Fig. 2. The tunnel is powered by a 112 KW electric motor,
and empty test-section velocities range from 1 to 25 m∕s, and the
total pressure can be varied between 20 (3 psi) and 413 kPa (60 psi) Fig. 3 CAD model of the rotor hub test rig.
(absolute). Test-section freestream turbulence is less than 0.3%. The
circular test section, used in this experiment, is 30.48 cm (12 in.) in
diameter and 76.2 cm (30 in.) long. The test section can be the shaft entry into the tunnel, while maintaining compliance for load
interchanged with a rectangular test section, depending on the measurements. A fairing (NACA 0025) was used to shroud the
application. Windows provide optical access (camera and laser) on driveshaft in order to isolate the hub wake from vortex shedding off
three out of four sides of the test section (the top and sides are the driveshaft. The model rotor hub and driveshaft were tilted 5 deg
transparent, and the bottom is opaque). Flat-surfaced laser windows into the flow in reference to forward flight of a full-scale helicopter at
were used for optical measurements, while windows with a curved μ  0.2. Figure 4 shows one of the five hub configurations (C)
inner surface were used for drag measurements at higher Rehub due to installed in the tunnel.
their higher pressure rating. The model-scale rotor hub was based on a large commercial
The experimental test rig in the 12-in.-diam water tunnel is shown helicopter and was 6 in. in diameter. Simplifications to the complex
schematically in Fig. 3. The assembly consisted of an outer support geometry were made in order to obtain a canonical test case, while
structure of steel plates that were hung from the test-section flanges. maintaining the major geometric features that are responsible for the
A 3.7 kW (5 hp) electric motor (Weg W22) was mounted to a flat pressure and interference drag. The model was computer numeric
plate, which pivoted about a bar across the lower support structure. control machined out of 17-4 stainless steel. Each rotor hub
The motor and shaft were allowed to pivot in order to provide
compliance to measure drag with a single-component load cell
(Omega, LC305-300), which was placed in between the pivoting
plate and a plate that was fixed to the support structure. Radial and
axial bearings were placed on the upper shaft in order to isolate the
motor from radial and axial loads. The driveshaft penetrated the
tunnel floor through a flexible housing, which allowed for sealing of

Fig. 4 Hub and shaft fairing installed in the water tunnel (configuration
C). The view is from the side, angled toward the test-section inlet on the
Fig. 2 Schematic of the PSU ARL 12-in.-diam test-section water tunnel. right.
1694 REICH, WILLITS, AND SCHMITZ
Downloaded by 122.172.77.152 on December 15, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.C034250

Fig. 5 Tested hub configurations. The view is from the side, with assumed flow from left to right.

component was machined separately and assembled in varying 12-in.-diam water tunnel. The limiting factor in this facility is the onset
configurations throughout the experiment. The basic model of cavitation at high freestream speeds. This sets the maximum
geometry was an identical, scaled-down version of that used in a operating freestream speed at about 15 m∕s at an advance ratio of
previous study conducted in the PSU ARL 48-in.-diam test-section 0.2. Higher freestream speeds can be achieved at higher advance ratios,
water tunnel [23], with the addition of several features. As shown in since the relative velocity on the advancing tip is lower. At this
Fig. 5, these features include 1) the ability for component buildup, condition, the hub-diameter-based Reynolds number Rehub is
2) the ability to include pitch links, and 3) the ability to add a beanie 2.6 × 106 . This corresponds to approximately 33% of the full-scale
fairing on top of the rotor hub. The two beanies were of a generic Reynolds number for a large helicopter (e.g., Sikorsky S92 and
design (ellipsoid top and flat bottom) that had no previous claim to Eurocopter EC225) at sea level and about 45% of the full-scale
drag reduction effectiveness. The beanie in configuration A mounted Reynolds number at an altitude of 10,000 ft. For a small helicopter
directly to the upper spider. The beanie in configuration A-M was (e.g., Robinson R44), this corresponds to approximately 75% of the
offset slightly by four standoffs, while being slightly thinner to full-scale Reynolds number at sea level and 70% of the full-scale
maintain a projected frontal area equal to configuration A. Figure 5 Reynolds number at an altitude of 2000 ft. Also note that the values of
shows all five hub-component configurations that were tested. The Rehub achievable in the 12-in.-diam water tunnel are lower than that of
scissors also have the ability to be rotated between two different the 48 in. Garfield Thomas Water Tunnel, used in a previous study [23].
orientations 30 deg apart, though this was not considered in the
present study. Torque was transferred from the driveshaft to the hub
by use of two square keys. Finite-element analysis was conducted on C. Instrumentation and Measurement
the driveshaft and hub assembly to verify that no natural modes of the 1. Flow Conditions Monitoring
structure would be excited by the test conditions. Abaqus/complete Conditions including the water temperature, pressure, impeller
abaqus environment was used for modeling the structure as well as speed, impeller pitch, and test-section speed were measured
solving for the natural modes. The lowest natural frequency was continuously throughout the experiment. The test-section speed was
found to be 146 Hz, which is considerably higher than the highest computed using two measurements. The total pressure was measured
operating shaft speed of 22 Hz. Considering a strong 4/rev (88 Hz) by a Kiel probe placed upstream of the test section, and the static
excitation from the fluid loading on the rotating hub, the highest- pressure was measured by several static pressure taps on the test-
speed operating condition was deemed safe as far as structural section wall. The combination of the total and static pressures
considerations are concerned. allowed the fluid speed to be computed at each static pressure tap
location, which varied no more than 1.2% of the measured freestream
B. Reynolds Number Scaling at the upstream static pressure tap. The freestream speed was taken as
Water tunnels have an advantage compared to wind tunnels for the speed computed from the pressure tap closest to the center of the
Reynolds-scale testing since the kinematic viscosity of water is model rotor hub. The tunnel was pressurized to maintain 45 psi
nominally 1/15 that of ambient air. For this work, a large helicopter (static) in the test section in order to prevent cavitation on low-
flying at an advance ratio of 0.2 and forward-flight speed of 42 m∕s pressure regions of the hub.
(138 ft∕s, 82 kt) was taken as a reference. At these conditions, the
Mach number is less than or equal to 0.18 anywhere on the hub, thus
justifying the incompressible flow assumption. The full-scale rotor 2. Drag Measurements
speed was assumed to be 233 rpm with a main hub radius of about As mentioned previously, drag was measured with a 1334 N (300 lb)
15% of the blade radius. A 1:17 scaled model was considered for the capacity single-component button load cell (Omega LC305-300). The
REICH, WILLITS, AND SCHMITZ 1695

Fig. 7 PIV measurement locations and coordinate system.

60 mm X × 48 mm Z. The images were spatially calibrated with


a target consisting of a grid of dots spaced 2.5 mm apart.
Since the laser pulse repetition rate (15 Hz) was insufficient to fully
resolve the relevant temporal hub wake features, which contain
unsteady content greater than the Nyquist frequency of 7.5 Hz, the
PIV acquisition was configured to acquire phase-averaged data. This
Fig. 6 Application of oil-based paint for surface-flow visualization. involved configuring the PIV system to acquire 250 image pairs at a
specified hub azimuthal position using the shaft encoder as a
Downloaded by 122.172.77.152 on December 15, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.C034250

triggering source. A minimum of one-half of a hub revolution


(180 deg) was acquired in 45 deg increments. The images were
load cell was placed between the pivoting plate and support bracket as recorded and processed using a commercial PIV software package
shown in Fig. 3. (DaVis 8, LaVision), which applied the image calibration and
Shaft rotational speed and position were measured with a through- computed the velocity vector field. The vectors were computed using
bore rotary encoder (Encoder Products Company 25T), which was standard cross-correlation methods with two passes. The inter-
mounted along the shaft as shown in Fig. 3. The encoder provided the rogation window was 32 × 32 pixels with 50% overlap. The final
rotation rate and position at a resolution of 1200 counts per vector spacing was 0.76 mm.
revolution, as well as a once-per-revolution pulse. This pulse was
used to determine the absolute shaft position as well as to trigger
D. Test Matrix
the particle image velocimetry (PIV) system to collect phase-
averaged data. The test conditions are summarized in Tables 1 and 2. The
The drag measurements were calibrated using an in situ process. Reynolds number Rehub was varied between 1.06 × 106 and 2.6 ×
A pulley was temporarily mounted in the tunnel directly downstream 106 and increased logarithmically in increments of logRe  0.077.
of the driveshaft. Calibration weights were then hung from the pulley The test-section velocity varied from 6.1 to 14.8 m∕s, and the rotor
to apply a known load to the shaft. Load cell output was measured for hub speed varied between 183 and 1333 rpm to maintain advance
each known load in order to obtain a calibration curve. Load was ratios of 0.2, 0.4, and 0.6. PIV data were acquired at two Reynolds
numbers, Rehub  1.06 × 106 and Rehub  1.8 × 106 , at advance
applied in an increasing and decreasing fashion, to quantify
ratio μ  0.2.
hysteresis effects, which were found to be negligible. Uncertainty
was calculated using the method detailed by Hufnagel [24]. The
difference between the known applied loads and the backcalculated
III. Results
loads, termed residual loads, was used to determine the measurement
uncertainty. The maximum residual load of 7.6 N (1.7 lb) was used A. Time-Averaged Drag
as the uncertainty for all drag measurements. Figure 8 shows the effects of Rehub on drag of the nonrotating
configuration B model rotor hub for four azimuth angles. There are
3. Surface-Flow Visualization noticeable differences in drag between different hub azimuthal
positions. This is due to the projected frontal area changing with the
To obtain a better qualitative understanding of the flow behavior
azimuthal position. The effect lessens as Rehub increases.
around the rotor hub, surface-flow visualization was performed using
Additionally, the sensitivity of drag on Rehub is affected by hub
oil-based paints. Standard art-grade oil pants were mixed with heavy-
azimuthal orientation. There also appears to be a drop in D∕q at
weight gear oil and then applied to the model with a foam paintbrush,
around Rehub  1.5 × 106 , with a subsequent leveling off. This is an
as shown in Fig. 6. The tunnel was filled and then ramped up to indicator of transition from sub- to supercritical flow and will be
Rehub  1.8 × 106 , while simultaneously adjusting the rotor hub discussed in further detail for the rotating cases.
rotational speed in order to maintain a constant advance ratio,
μ  0.2. The flow speed was later decreased in the same fashion.
Table 1 Drag measurement test matrix
4. Particle Image Velocimetry
Configurations μ Rehub
PIV measured the streamwise and vertical velocity components in
the double-frame, double-pulse mode at two planes downstream of A, B Nonrotating 1.06 × 106 to 2.6 × 106
the hub, nominally located at downstream distances of X  2Rhub A, B, C, D, A-M 0.2 1.06 × 106 to 2.6 × 106
A, B, C, D, A-M 0.4 1.06 × 106 to 2.6 × 106
and X  4Rhub , shown in Fig. 7. The PIV planes were illuminated A, B, C, D, A-M 0.6 1.06 × 106 to 2.6 × 106
with a light sheet that entered the flow from above the test section,
centered in the tunnel (y  0R), and aligned with the streamwise
direction. The sheet was formed with the beam of a pulsed
neodymium-doped yttrium aluminium garnet laser operating at Table 2 PIV measurement test matrix
532 nm wavelength. The maximum repetition rate of the laser was Configuration Location μ Rehub
15 Hz. The beam was formed into a sheet with the use of a cylindrical
lens placed above the test section, not directly on the laser. The light A X  2Rhub 0.2 1.06 × 106 , 1.8 × 106
A X  4Rhub 0.2 1.06 × 106 , 1.8 × 106
sheet was then angled into the test section via two plane mirrors. The B X  2Rhub 0.2 1.06 × 106 , 1.8 × 106
illuminated plane was imaged with a 1280 × 1024 pixel charge- B X  4Rhub 0.2 1.06 × 106 , 1.8 × 106
coupled device camera. The resulting nominal field of view was
1696 REICH, WILLITS, AND SCHMITZ

40
Rehub < 1.5 × 106 . In addition, there is no discernable difference in
35 drag between various advance ratios. This is an important result, since
it suggests that the rotation rate has little to no effect on the rotor hub
30 total drag. This again has implications on future design studies,
allowing the elimination of rotation rate variation in parametric
D/q [m 2 × 10-4 ]

25 studies of time-averaged drag. The unsteady drag and wake behavior,


20 as shown later, is not subject to this finding. The frequency and
strength of coherent and stochastic turbulent velocity fluctuations in
15 the wake are dependent on the rotation rate, which is shown later in
ψ =0 this paper. Figure 10a compares differences between configurations
10 ψ = 45 at one advance ratio, μ  0.2. All configurations containing a beanie
5
ψ = 90 fairing (A, A-M, and D) show slightly higher D∕q than
ψ = 135 configurations without a beanie fairing (B and C). The drag of
0 configuration C is virtually equal, within the uncertainty, to that of
1 1.5 2 2.5 3 configuration B, despite the added frontal area of the pitch links. The
6
Rehub [ × 10 ] pitch links add 2.5% to the maximum frontal area, within the
Fig. 8 Reynolds number effects, nonrotating hub, configuration B. uncertainty of the drag measurement, which is quantified to range
between 2.8% for the highest Rehub and 12% for the lowest Rehub . All
configurations enter the transitional region from sub- to supercritical
Figures 9 and 10 further examine the effects of Rehub on drag, flow at Rehub  1.5 × 106 . With reference to Table 3, this is
Downloaded by 122.172.77.152 on December 15, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.C034250

comparing first the advance ratio and then configurations. consistent with the supercritical Reynolds numbers reported in [19]
All configurations show relative Rehub independence for for two-dimensional circular and rectangular cylinders. To further

40 40 40
35 35 35
30 30 30
D/q [m × 10 ]

D/q [m2 × 10-4 ]

D/q [m2 × 10-4 ]


-4

25 25 25
20 20 20
2

15 15 15
10 10 10
µ = 0.2 µ = 0.2 µ = 0.2
5 µ = 0.4 5 µ = 0.4 5 µ = 0.4
µ = 0.6 µ = 0.6 µ = 0.6
0 0 0
1.0 1.5 2.0 2.5 3.0 1.0 1.5 2.0 2.5 3.0 1.0 1.5 2.0 2.5 3.0
Rehub [ ×10 6 ] Rehub [ ×10 6 ] Rehub [ ×10 6 ]

a) Configuration A b) Configuration A-M c) Configuration B

40 40
35 35
30 30
D/q [m2 × 10-4 ]

D/q [m2 × 10-4 ]

25 25
20 20
15 15
10 10 µ = 0.2
µ = 0.2
µ = 0.4 µ = 0.4
5 5
µ = 0.6 µ = 0.6
0 0
1.0 1.5 2.0 2.5 3.0 1.0 1.5 2.0 2.5 3.0
Rehub [ ×10 6 ] Rehub [ ×10 6 ]

d) Configuration C e) Configuration D
Fig. 9 Effect of advance ratio μ and Reynolds number Rehub on D∕q.

40
1.4
35
1.2
30
D/q [m2 × 10- 4]

1
25
0.8
CD

20

15 A 0.6 A
B B
10 0.4 C
C
D D
5 0.2
A-M A-M
0 0
1 1.5 2 2.5 3 1 1.5 2 2.5 3
Re [× 10 6] Re [× 10 6]
hub hub
a) D/q b) Drag coefficient
Fig. 10 Comparison of the five configurations and the effect of Reynolds number Rehub on D∕q and CD (μ  0.2).
REICH, WILLITS, AND SCHMITZ 1697

Table 3 Span-averaged component Reynolds numbers (advancing side)


Rehub  2.6 × 106 Rehub  1.8 × 106 Rehub  1.06 × 106
Components μ  0.6 μ  0.2 μ  0.6 μ  0.2 μ  0.6 μ  0.2
Pitch links 0.37 × 105 0.64 × 105 0.27 × 105 0.45 × 105 0.17 × 105 0.30 × 105
Spiders 1.1 × 105 1.8 × 105 0.73 × 105 1.2 × 105 0.45 × 105 0.75 × 105
Scissors 1.1 × 105 1.9 × 105 0.80 × 105 1.3 × 105 0.47 × 105 0.76 × 105
Blade stubs 2.0 × 105 2.6 × 105 1.4 × 105 2.0 × 105 0.8 × 105 1.0 × 105

illustrate the local Reynolds number ranges in Table 3, the local blade Examining the effects of Rehub in Fig. 13, the overall drag levels
stub Reynolds number is plotted as a function of radius and azimuth are certainly affected, as would be expected from the previous
in Fig. 11 for Rehub  2.6 × 106 and three advance ratios (i.e., three discussion of time-averaged drag. As Rehub is increased, D∕q
rotor speeds). Flow is from left to right, and rotation is decreases, which is consistent with the time-averaged drag
counterclockwise (the hub viewed from above). Note that this is measurements in Figs. 9 and 10 and points to the fact that at least
the magnitude of the local component Reynolds number in the some of the hub components are operating in a flow regime that is
freestream direction, so reversed flow is not indicated, since the blade transitioning from sub- to supercritical local Reynolds numbers.
stubs are symmetrical from front to rear. It is apparent that with an The peak-to-peak amplitudes also decrease with an increasing
increasing advance ratio the contribution to the relative velocity Rehub , which is consistent with hub components operating in
decreases. For μ  0.6, the Reynolds number does not differ supercritical flow, thus exhibiting weaker fluctuations. Another
Downloaded by 122.172.77.152 on December 15, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.C034250

significantly between advancing and retreating blade stubs. As interesting feature is that the amplitude of the peaks is generally
mentioned earlier, both of the beanie fairing designs were generic, greater in configurations with a beanie fairing (A, A-M, and D) than
with no claim of a positive effect, i.e., reduced drag and reduced in configurations without a beanie fairing (B and C). Further
unsteadiness in the wake. Additionally, it should be noted that, explanation of this phenomenon can be gleaned from analysis of the
compared to drag coefficients in Fig. 10b, the behavior is reversed. In drag harmonics in the following paragraph.
fact, for configurations that include a beanie fairing, the drag Figures 14 and 15 show D∕q harmonics, which are computed by
coefficient is reduced. However, these configurations possess a taking the FFT of the phase-averaged D∕q shown in Figs. 12 and 13.
higher frontal area, increasing the overall drag on the hub. This Comparing advance ratios for all five configurations in Fig. 14, there
illustrates the importance of using D∕q (or, alternatively, using a are notable differences in the harmonic content between various
constant reference area for the drag coefficient) as a metric when advance ratios. For all configurations, the 4/rev D∕q fluctuations
comparing hub configurations. Note further that beanie fairings are decrease with an increasing advance ratio, which is expected.
sometimes used to deflect the hub wake to avoid undesirable flow However, contrary to expectations, the 4/rev harmonic is not always
interactions at certain flight conditions. the dominant frequency. In the μ  0.6 case, the 8/rev amplitude is, in
fact, greater than the 4/rev amplitude. This suggests a higher degree of
B. Time-Varying Drag interaction between the 4/rev shed structures for the μ  0.6 case. To
In this section, the unsteady drag is examined both in the rotor elaborate, the 8/rev frequency is likely the blade stub interacting with
phase domain and in the frequency domain. Phase-averaged data are the flow structure shed by the previous blade stub. Furthermore, at
shown, along with harmonics amplitude data, which are simply the higher advance ratio μ, the local blade stub Reynolds number is lower
fast Fourier transform (FFT) of the phase-averaged data. The phase- (see Table 3), which may change the behavior of the shed flow
averaged plots provide a physically intuitive representation of the structure, which will then interact differently with the trailing blade
drag, while harmonics plots show frequency content that may not be stub. Additionally, there is a notable 2/rev drag component in Fig. 14,
obvious in the phase-averaged plots. which is due to the scissors (the only geometric feature on the hub with
a passage frequency of twice per revolution). Examining the 2/rev drag
1. Phase-Averaged Drag harmonic, the amplitude is relatively constant with advance ratios
Figures 12 and 13 show phase-averaged drag over 180 deg of hub for configurations containing pitch links (A, A-M, and C). For
rotation. Only one-half of the hub rotation is shown due to the hub configurations that do not contain pitch links, the 2/rev amplitude is
geometry being symmetric about 180 deg. Comparing advance ratios much higher for μ  0.2 and decreases with an increasing μ. This is
for each configuration in Fig. 12, there are several features worthy of likely due to the pitch links interactional effect of breaking up the flow
discussion. In general, the amplitude of the drag oscillations is around the scissors (the only 2/rev geometric feature on the hub; see
inversely related to the advance ratio, which is expected due to the Fig. 5). This observation indirectly counters the argument made by
lower relative velocity fluctuations incurred by the hub components Dombroski and Egolf [25] that pitch links are not necessary for testing
at higher advance ratios (i.e., lower Ω for the same Rehub ). Another because they only add a small 4/rev content to the flow. The results here
notable feature is that the phase-averaged drag and harmonics are suggest that if the unsteady 2/rev component is of interest the pitch
rather similar between configuration B and configuration C, which is links must be included for their interactional effects on the flow over
in agreement with a previous study by Dombroski and Egolf [25]. the scissors. Examining the 6/rev drag harmonic, there is a difference in

Fig. 11 Effect of advance ratio μ on local blade stub Reynolds number, Rehub  2.6 × 106 . The flow is from left to right. From left to right: μ  0.2,
μ  0.4, μ  0.6.
1698 REICH, WILLITS, AND SCHMITZ

× 10-3 × 10-3 × 10-3


4 4 4
µ = 0.2 µ = 0.2 µ = 0.2
µ = 0.4 µ = 0.4 µ = 0.4
3 µ = 0.6 3 µ = 0.6 3 µ = 0.6
D/q [m2 ]

D/q [m2 ]

D/q [m2 ]
2 2 2

1 1 1

0 0 0
0 30 60 90 120 150 180 0 30 60 90 120 150 180 0 30 60 90 120 150 180
Ψ [deg] Ψ [deg] Ψ [deg]
a) Configuration A b) Configuration A-M c) Configuration B

-3 -3
× 10 × 10
4 4
µ = 0.2 µ = 0.2
µ = 0.4 µ = 0.4
3 µ = 0.6 3 µ = 0.6
D/q [m2 ]

D/q [m2 ]
2 2
Downloaded by 122.172.77.152 on December 15, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.C034250

1 1

0 0
0 30 60 90 120 150 180 0 30 60 90 120 150 180
Ψ [deg] Ψ [deg]
d) Configuration C e) Configuration D
Fig. 12 Advance ratio effects on phase-averaged D∕q for all five configurations, Rehub  2.6 × 106 .

−3 −3 −3
x 10 x 10 x 10
5 5 5

4 4 4
D/q [m 2 ]

D/q [m 2 ]
D/q [m2 ]

3 3 3

2 2 2
6
Re = 1.8 × 10
6 Rehub = 1.8 × 106 Rehub = 1.8 × 10
hub
6
1 Re = 2.2 × 10 6
1 Re = 2.2 × 106 1 Rehub = 2.2 × 10
hub hub
6 6
Re = 2.6 × 10
6 Rehub = 2.6 × 10 Re = 2.6 × 10
hub hub

0 0 0
0 30 60 90 120 150 180 0 30 60 90 120 150 180 0 30 60 90 120 150 180
Ψ [deg] Ψ [deg] Ψ [deg]

a) Configuration A b) Configuration A-M c) Configuration B

−3 −3
x 10 x 10
5 5

4 4
D/q [m 2 ]
D/q [m 2 ]

3 3

2 2
Re = 1.8 × 106 Rehub = 1.8 × 10
6
hub
6
1 Rehub = 2.2 × 10 1 Rehub = 2.2 × 106
6
Rehub = 2.6 × 10 Re = 2.6 × 106
hub
0 0
0 30 60 90 120 150 180 0 30 60 90 120 150 180
Ψ [deg] Ψ [deg]

d) Configuration C e) Configuration D
Fig. 13 Reynolds number effects on phase-averaged D∕q for all five configurations, μ  0.2.

behavior between configurations with pitch links (A, A-M, and C) and pronounced in some, there is a decrease in the 4/rev D∕q amplitude
those without pitch links (B and D). For configurations with pitch links, with an increasing Reynolds number. This suggests that the blade
the 6/rev amplitude either decreases with the advance ratio (A-M) or stubs are entering the supercritical flow regime, where drag is
stays constant between μ  0.2 and μ  0.4, followed by a decrease decreasing with the Reynolds number. Also, contrary to its design
between μ  0.4 and μ  0.6. For configurations without pitch links, intention, the beanie fairing (configurations A, A-M, and D)
the 6/rev amplitude increases from μ  0.2 to μ  0.4, then decreases increases the 4/rev drag harmonic amplitude at the lower end of the
in amplitude from μ  0.4 to μ  0.6. Likely due to nonlinear Reynolds numbers shown. Interestingly, at the higher Reynolds
interactions, more detailed experiments and analysis are needed to numbers shown, the beanie fairing has the opposite effect, i.e.,
fully explain this phenomenon. reducing the 4/rev drag harmonic amplitude. This has important
Comparing Reynolds numbers Rehub in Figs. 13 and 15, there are implications for scaled-model testing, illustrating that proper
several notable trends. Across all configurations, although more Reynolds number scaling is essential in order to obtain behavior
REICH, WILLITS, AND SCHMITZ 1699

× 10-3 × 10-3 × 10-3


1 1 1
µ = 0.2 µ = 0.2 µ = 0.2
µ = 0.4 µ = 0.4 µ = 0.4
0.8 0.8 0.8
µ = 0.6 µ = 0.6 µ = 0.6
|FFT(D/q)| [m2]

|FFT(D/q)| [m2]

|FFT(D/q)| [m2]
0.6 0.6 0.6

0.4 0.4 0.4

0.2 0.2 0.2

0 0 0
0 2 4 6 8 10 12 14 16 0 2 4 6 8 10 12 14 16 0 2 4 6 8 10 12 14 16
f/Ω f/Ω f/Ω
a) Configuration A b) Configuration AM c) Configuration B

× 10-3 × 10-3
1 1
µ = 0.2 µ = 0.2
µ = 0.4 µ = 0.4
0.8 0.8
µ = 0.6 µ = 0.6
|FFT(D/q)| [m2]

|FFT(D/q)| [m2]
0.6 0.6
Downloaded by 122.172.77.152 on December 15, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.C034250

0.4 0.4

0.2 0.2

0 0
0 2 4 6 8 10 12 14 16 0 2 4 6 8 10 12 14 16
f/Ω f/Ω
d) Configuration C e) Configuration D
Fig. 14 Advance ratio effects on D∕q harmonics for all five configurations, Rehub  2.6 × 106 .

−3 −3 −3
x 10 x 10 x 10
2 2 2
6 6 6
Rehub = 1.8 × 10 Rehub = 1.8 × 10 ReD = 1.8 × 10
6 6 6
Rehub = 2.2 × 10 Rehub = 2.2 × 10 ReD = 2.2 × 10
1.5 1.5 1.5
6 6 6
|FFT(D/q)| [m ]

|FFT(D/q)| [m2]

|FFT(D/q)| [m ]

Rehub = 2.6 × 10 Rehub = 2.6 × 10 ReD = 2.6 × 10


2

1 1 1

0.5 0.5 0.5

0 0 0
0 2 4 6 8 10 12 14 16 0 2 4 6 8 10 12 14 16 0 2 4 6 8 10 12 14 16
f/Ω f/Ω f/Ω

a) Configuration A b) Configuration A-M c) Configuration B

−3 −3
x 10 x 10
2 2
6 6
Rehub = 1.8 × 10 Rehub = 1.8 × 10
Rehub = 2.2 × 106 Rehub = 2.2 × 106
1.5 1.5
|FFT(D/q)| [m ]

|FFT(D/q)| [m ]

6 6
Rehub = 2.6 × 10 Rehub = 2.6 × 10
2

1 1

0.5 0.5

0 0
0 2 4 6 8 10 12 14 16 0 2 4 6 8 10 12 14 16
f/Ω f/Ω
d) Configuration C e) Configuration D
Fig. 15 Reynolds number effects on D∕q harmonics for all five configurations, μ  0.2.

relevant to the full-scale aircraft. If below a critical Reynolds number, 2. Note on Vortex Shedding Effects
certain component designs can have the opposite effect that they do at It should also be noted that the effect of drag fluctuations due to
higher Reynolds numbers. Hence, improper Reynolds number vortex (Strouhal) shedding is small compared to the blade passage
scaling, among other shortcomings, in the design process has the frequency effects. Several studies [21,26–28] documented the
potential to cause suboptimal design decisions for the full-scale amplitude of the fluctuating drag coefficient CD0 of a nonrotating
aircraft, which may require costly design changes in the flight-testing circular cylinder with vortex shedding to be around 10% of the mean
phase. Keep in mind that the physical explanations are hypotheses at drag coefficient. When applied to the blade stubs (see Fig. 5) on the
this point and would need more detailed investigation of the near- nonrotating hub in this study at Rehub  2.6 × 106 , this leads to a
field flow phenomena to be fully understood. fluctuating drag amplitude of about D ≈ 0.8 N, or D∕q ≈ 7.3 × 10−6,
1700 REICH, WILLITS, AND SCHMITZ

which is well within the measurement uncertainty in this study. The which are expected due to radial pressure gradients associated with
effect is further diminished by the fact that, when rotating, the blade the rotating geometry. Lastly, number 3 suggests downwash due to
stub is only operating at this condition for a small portion of each hub the hub flow. This is important because, in real-world applications,
rotation. In addition, the frequencies associated with vortex shedding this is where the engine pylon would be located. It has been
from the hub components are extremely high compared to that of the documented that the rotor hub flow can cause premature separation
blade passage frequency. Considering the dimensions of the blade on the engine pylon [5,7,8,10,16]. It is not clear whether downwash
stub and a Strouhal number of a similar shape from [19], the in this case would be detrimental or beneficial to engine pylon flow.
frequency of this shedding at Rehub  2.6 × 106 and μ  0.2 occurs The downwash potentially could delay separation in the adverse
at 500 Hz, compared to the passage frequency of the blade stub pressure gradient region. On the other hand, the unsteady, separated
of 88 Hz. hub wake could cause premature separation. More detailed testing
would be required to definitively determine the effect of this rotor hub
C. Surface-Flow Visualization on an engine pylon. Figures 16b and 16c show flow patterns for
configuration B, nonrotating, at two different angles. Flow patterns
To obtain a better qualitative understanding of the flow physics,
for the nonrotating cases show differences from those of the rotating
surface-flow visualization was performed using oil-based paints.
case, particularly the angle of the stagnation and separation lines on
Figure 16 shows the result of this process for three cases. Figure 16a the upper surface of the blade stubs. Comparing rotating and
shows the surface flow of configuration A, rotating. There are three nonrotating cases, the location of the stagnation lines on the (what
notable features, which are denoted by numbers in the figure. would be) advancing blade stub of the nonrotating, 0 deg case closely
Number 1 shows a distinct line of concentrated paint on the resemble those of the rotating case. This suggests that the observed
advancing blade stub, indicating flow stagnation on the leading edge flow patterns from the rotating case are a representation heavily
of the blade stub. Number 2 shows radial flow lines on the blade stub,
Downloaded by 122.172.77.152 on December 15, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.C034250

weighted toward the advancing blade stub. Further investigation is


required to confirm this, including testing over a wider range of
nonrotating angles and a range of rotation rates.

D. Particle Image Velocimetry


1. Panoramas
Figures 17–19 show phase-averaged vertical w velocity contours,
which have been combined to create a panorama. Eight phase-
averaged frames, at streamwise locations shown in Fig. 7, with a hub
azimuthal position ranging from 0 to 180 deg, were stitched together
to provide a pseudospatially resolved view of the flow over one-half
of a hub rotation. The image stitching procedure was based on
Taylor’s hypothesis, which states that if the freestream speed is much
greater than the velocity fluctuations then one can assume that
velocity variations at a point are only due to the passage of a
nonvarying pattern of turbulence over that location [29]. Taking the
mean flow velocity in the wake and the time between phase-averaged
frames, the offset is calculated, and then the images are overlapped to
create the panorama. These figures illustrate both 4/rev (positive
magnitude) and 2/rev (negative magnitude) flow disturbances being
shed from the hub. The difference in magnitude between the two blue
areas is the source of the 2/rev fluctuation. As discussed in [23],
considering their phase and orientation on the hub with respect to
each other, these disturbances can be attributed to the blade stubs/
spiders and scissors, respectively. This is consistent with the drag
measurements, as well as measurements previously reported for
Rehub  4.9 × 106 in [23]. It is apparent that the flow disturbances
are dissipated with downstream distance, which will be quantified
with harmonics in the next section. When examining Rehub effects
(i.e., comparing Figs. 17 and 18), there are qualitatively only small
differences in the wake. All of the major flow features are common
between the two Rehub . Figure 19 shows panoramas for two locations
downstream of hub configuration A. There are clear differences
between the panoramas of configuration A and B. Configuration A
has substantially stronger 4/rev and 2/rev flow disturbances in the
wake. The other components that were added in configuration Awere
the beanie fairing and pitch links. As discussed previously, the beanie
fairing could be the cause of the higher velocity fluctuations in the
wake. The beanie is likely creating a low-pressure region over
the upper surface, which drives more flow (that would otherwise
go under the hub, interacting with the scissors) upward and
over the blade stubs, strengthening the 4/rev flow disturbance.
This observation in the hub wake between Figs. 18 and 19 is
consistent with the drag harmonics shown in Figs. 15a and 15c
for Rehub  1.8 × 106.

2. Average Velocity
Figure 20 shows mean u- and w-velocity profiles extracted from
PIV data of this study, as well as a previous study [23]. All profiles are
Fig. 16 Surface-flow patterns visualized with oil-based paints. for configuration B. Shown are distributions of mean values along
REICH, WILLITS, AND SCHMITZ 1701

Fig. 17 Panorama of vertical w velocity contours for configuration B, Rehub  1.06 × 106 , μ  0.2, representing 180 deg of hub rotation from left to
right. Dots on the right-hand side show vertical locations of velocity probes for harmonics analysis.
Downloaded by 122.172.77.152 on December 15, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.C034250

Fig. 18 Panorama of vertical w velocity contours for configuration B, Rehub  1.8 × 106 , μ  0.2, representing 180 deg of hub rotation from left to right.
Dots on the right-hand side show vertical locations of velocity probes for harmonics analysis.

Fig. 19 Panorama of vertical w velocity contours for configuration A, Rehub  1.8 × 106 , μ  0.2, representing 180 deg of hub rotation from left to right.
Dots on the right-hand side show vertical locations of velocity probes for harmonics analysis.

with shaded regions representing the root-mean-square fluctuations of Rehub on the wake profile that is consistent with load-cell data
about the mean due to the passing flow structures (note that these are presented in Fig. 9. These wake profiles compare with measured
not measurement uncertainties); see also PIV panoramas in Figs. 17 drag coefficients, with a higher velocity deficit consistently
and 18. Although the vertical locations of the PIV planes are not the correlating to a higher drag (or D∕q) across the range of Rehub tested in
same, there is some overlap for comparison. Examining u-velocity this test as well as in [23]. The velocity deficit decreases for the first
(i.e., streamwise velocity) profiles in Fig. 20a, there is a clear effect three Rehub , followed by a slight increase at the highest Rehub , which is
1702 REICH, WILLITS, AND SCHMITZ

Fig. 20 Velocity profiles from PIV data at X  2Rhub for four different Rehub , μ  0.2 (configuration B). Shaded areas denote the root-mean-square of
the fluctuating velocity component about the mean.
Downloaded by 122.172.77.152 on December 15, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.C034250

Fig. 21 Comparison of vertical velocity harmonics downstream of


configuration B at μ  0.2 for four different values of Rehub .

consistent with the drag coefficient trends shown in this study and [23].
Examining the w velocity (i.e., vertical velocity) in Fig. 20b, there are
smaller differences in the w velocity across the range of Rehub than
observed for the u velocity. This suggests that, although the mean
streamwise velocity component is highly dependent on Rehub , the
vertical velocity component is not significantly affected by Rehub.
However, the harmonic content, shown in Figs. 21 and 22, shows a
clear Rehub dependence, with 6/rev content becoming stronger at
Rehub  1.8 × 106 , when the flow transitions from the sub- to
supercritical regime.

3. Harmonics
Although azimuthal resolution is limited to 4.5 deg increments
(equivalent to 80/rev), relevant harmonic content can be obtained
from phase-averaged PIV data. Note that the harmonics were not
extracted from the panorama data but from values at a fixed point in
space across 40 phase-averaged vector fields with azimuthal
resolution of 4.5 deg. Figure 21 compares PIV harmonics at two
probe locations and four Reynolds numbers at X  2Rhub (near
wake). The Rehub  2.45 × 106 and Rehub  4.9 × 106 data are from
a previous test in the 48-in.-diam test-section Garfield Thomas Water
Tunnel at PSU ARL, detailed in [23]. Several trends are observed in
the harmonic content in the wake flow. In general, the 2/rev content is
fairly consistent with an increasing Rehub , while 4/rev and 6/rev
content increases with an increasing Rehub . Note that these data are Fig. 22 PIV vertical velocity harmonics for both configurations
only available for configuration B. This further suggests that high (Config) and both downstream measurement locations, μ  0.2,
Reynolds numbers are needed to accurately test drag and wake Rehub  1.8 × 106 .
REICH, WILLITS, AND SCHMITZ 1703

observed with laser-Doppler velocimetry measurements by Berry in


the wake of a full rotor and fuselage model in a wind tunnel test [30].
Associated with this is an increase in 6/rev content in the wake of
configuration A in Fig 22, which does not occur for configuration B.
As discussed earlier with respect to drag harmonics, for the current
hub design, the pitch links act as an interactional aerodynamic
element between the scissors (2/rev) and the blade stubs (4/rev),
modulating the latter and thus adding a 6/rev content.
The PIV harmonics may be compared to the drag harmonics
presented earlier, recalling that the PIV data only represent a plane in
the flow and do not capture the entire three-dimensional wake. The
higher 4/rev velocity fluctuations for configuration A are consistent
with the higher 4/rev drag fluctuations in the drag harmonics and
phase-averaged data.
To further the discussion of the increased 4/rev magnitude due to
the beanie fairing, Fig. 23 shows the magnitude of the 4/rev velocity
fluctuation as a function of vertical location for configurations A and
B. It is apparent that the increase in 4/rev velocity fluctuations occurs
primarily in the physical location associated with the beanie fairing.
There is a slight increase in the 4/rev velocity fluctuation due to the
scissors; however, the bulk of the increase occurs directly
Downloaded by 122.172.77.152 on December 15, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.C034250

downstream of the beanie fairing and blade stubs. This further


supports the previous claim that the beanie is accelerating flow over
Fig. 23 Comparison of the magnitude of the 4/rev vertical w velocity
the blade stubs, increasing the 4/rev drag and velocity fluctuations.
fluctuations at X  2Rhub . Configuration silhouettes are shown in
background for reference.

IV. Conclusions
characteristics for model-scale rotor hubs. The apparent increase In summary, effects of the Reynolds number, advance ratio, and
in 4/rev content with an increasing Rehub is attributed to the blade geometry configuration of a model-scale rotor hub were quantified in
stubs operating in increasingly supercritical flow; see also Table 3. the Pennsylvania State University Applied Research Laboratory
This 4/rev content interacting with the 2/rev content caused by the 12-in.-diameter test-section water tunnel. Time-averaged and time-
scissors, which shows much less Rehub dependence in the center of dependent D∕q were examined over a range of Reynolds numbers
the hub, results in a 6/rev signal with growing strength as Rehub and advance ratios for five hub-component configurations. Planar
increases, as shown in Fig. 21. Note that the opposite is true for the particle image velocimetry was conducted at two downstream
8/rev content. This process appears to be initiated at Rehub  locations, for two geometry configurations, and for two values of
1.8 × 106 when the hub flow transitions from the sub- to Rehub . Analysis was conducted on the drag and particle image
supercritical flow regime. A strong 6/rev disturbance that velocimetry data in both the time and frequency domains for
maintains strength in the long-age wake can have adverse effects maximum insight into the fundamental physics of the hub flow.
on empennage aerodynamics [25]. The effect of Rehub on drag is evidenced by the following:
Figure 22 shows harmonics of vertical velocity at five probe 1) Nonrotating and rotating data show a clear transition from sub-
locations taken from the PIV data. Probes are aligned vertically with a to supercritical flow around Rehub  1.5 × 106 . For the hub
spacing of 0.1R. The first item of note is that the 4/rev velocity configurations considered in this work, this appears to be the
fluctuation magnitudes are larger in the wake of configuration A than minimum Rehub for a consistent wake behavior.
in the wake of configurations B. As mentioned earlier in the drag 2) Time-averaged D∕q is configuration dependent, with a
section, this was unexpected, considering the beanie fairing should notional, generic beanie fairing causing increased D∕q at Rehub <
not add additional per-revolution content due to its shape. As 1.8 × 106 over the unfaired beanie. Pitch links do not significantly
mentioned before, it is theorized that the beanie, which was only affect time-averaged D∕q, or the general trend of D∕q with Rehub .
notional in design (i.e., elliptical in shape and not optimized), is This is attributed to the small frontal area of pitch links when
bringing flow upward and over the blade stubs, causing an increased compared to other hub components for the hubs considered in
interaction of the incoming flow with the blade stubs and hence a this work.
higher 4/rev flow disturbance in the wake. This effect is further The effect of advance ratio μ on drag is the following. Advance
illustrated in Fig. 23. The magnitude of the 4/rev vertical velocity ratio μ has no measureable effect on the time-averaged D∕q of any
fluctuation is plotted as a function of the vertical location for configuration tested. However, the advance ratio certainly affects the
configurations A and B in the near wake. It is clearly evident that the time-varying D∕q. An increasing advance ratio causes a decrease in
4/rev velocity fluctuation is significantly stronger for configuration A the amplitude of the unsteady D∕q harmonics up to 6/rev for all
than configuration B at the vertical location of the beanie fairing. This configurations and Reynolds numbers tested, as would be expected.
is further evidence that the beanie is accelerating the flow over the The effect of the configuration on drag is the following:
blade stubs and therefore strengthening the 4/rev wake velocity 1) Pitch links play an imporant role as an element causing
fluctuation. This is consistent with the drag measurements. It is interactional aerodynamics between 2/rev, 4/rev, and 6/rev flow
important to keep in mind that the drag measurements show a trend of features of the rotor hub.
a decreasing 4/rev harmonic strength with an increasing Reynolds 2) A beanie fairing can cause an increase in overall D∕q as well as
number. Therefore, the beanie might be beneficial at higher Reynolds the amplitude of unsteady D∕q harmonics. However, with an
numbers than what were considered in this study. There is also an increasing Reynolds number, this effect is diminished. At Reynolds
increase in the 4/rev magnitude at the lowest vertical position due to numbers higher than what were tested in this study, the beanie may
the addition of the pitch links. While noticeable, this difference is actually have a positive effect on the unsteady D∕q of the rotor hub.
considerably smaller than the increase due to the beanie. Additionally, the beanie fairing may have other benefits such as
Also of particular interest is the fact that, for configuration A, at redirecting the flow away from a control surface for a particular
certain vertical locations the 2/rev velocity fluctuation amplitude aircraft configuration.
increases from X  2Rhub to X  4Rhub , which is contrary to the The effect of Rehub and the configuration on wake is the following:
expected decay of unsteadiness with a downstream location in a 1) The rotor hub time-averaged wake deficit is Reynolds number
wake. This behavior is not, however, unprecedented, as it was dependent, showing trends consistent with drag measurements.
1704 REICH, WILLITS, AND SCHMITZ

2) The rotor hub wake unsteady content is configuration and Rehub [10] Roesch, P., and Dequin, A., “Experimental Research on Helicopter
dependent, with the beanie fairing causing increased 4/rev velocity Fuselage and Rotor Hub Wake Turbulence,” Journal of the American
fluctuations over that of the unfaired hub at the Reynolds numbers Helicopter Society, Vol. 30, No. 1, 1985, pp. 43–51.
tested. However, the drag observations in the previous conclusion [11] Hoffman, J., “The Relationship Between Rotorcraft Drag and Stability
and Control,” American Helicopter Society International 31st Annual
suggest a reduced negative impact and possibly a benefit of the beanie
National Forum, American Helicopter Soc., Washington, D.C.,
on the unsteady wake at Reynolds numbers approaching full scale, May 1975, pp. 8.1–8.3.
assuming that primary drag contributors operate in supercritical flow. [12] Prouty, R., and Amer, K., “The YAH-64 Empennage and Tail Rotor—A
3) The rotor hub wake unsteady content is dependent on Rehub , Technical History,” American Helicopter Society International 38th
indicating high Reynolds-scale testing is necessary for accurate Annual National Forum, American Helicopter Soc., Washington, D.C.,
measurement of the wake. 1982, pp. 247–261.
4) Ultimately, full-scale Rehub is required to obtain accurate drag [13] Sheridan, P., and Smith, R., “Interactional Aerodynamics—A New
measurements in reference to a specific production design. This may Challenge to Helicopter Technology,” Journal of the American
or may not be the case for wake measurements, however. Given wake Helicopter Society, Vol. 25, No. 1, 1980, pp. 3–21.
data at fractions on the order of 1/10 to 1/2 of the full-scale Rehub , it [14] Reich, D., Shenoy, R., Smith, M., and Schmitz, S., “A Review of 60
may be a possible to extrapolate data to full scale. Years of Rotor Hub Drag and Wake Physics: 1954–2014,” Journal of the
American Helicopter Society, Vol. 61, No. 2, 2016, pp. 1–17.
5) It should be noted that the unsteady wake is problematic only
doi:10.4050/JAHS.61.022007
when it combines with other factors including structural modes of the [15] Montana, P., “Experimental Investigation of Three Rotor Hub Fairing
airframe, rotor downwash, and engine exhaust at a specific point in Shapes,” David Taylor Naval Ship Research and Development Center
the flight envelope. While those factors are beyond the scope of this TR ASED 333, 1975.
study, the results presented here are useful for a basic understanding [16] Logan, A., Prouty, R., and Clark, D., “Wind Tunnel Tests of Large- and
of rotor hub flow physics.
Downloaded by 122.172.77.152 on December 15, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.C034250

Small-Scale Rotor Hubs and Pylons,” U.S. Army Aviation Research and
Development Command TR-80-D-21, 1981.
Acknowledgments [17] Felker, F., “An Experimental Investigation of Hub Drag on the XH-
59A,” AIAA Applied Aerodynamics Conference, AIAA Paper 1985-
This work was supported under Vertical Lift Research Center of 4065-302, 1985.
Excellence task 1.2 at the Pennsylvania State University, funded by [18] Shenoy, R., Holmes, M., Smith, M., and Komerath, N., “Scaling
the U.S. Army, the U.S. Navy, and NASA. Tom Maier, Mahendra Evaluations on the Drag of a Hub System,” Journal of the American
Bhagwat (Army Aeroflightdynamics Directorate), and Judah Helicopter Society, Vol. 58, No. 3, 2013, pp. 1–13.
Milgram (Office of Naval Research) are the technical Points Of doi:10.4050/JAHS.58.032002
Contact. This research is partially funded by the U.S. Government [19] Delany, N., and Sorensen, N., “Low-Speed Drag of Cylinders of Various
under agreement number W911W6-11-2-0011. The authors would Shapes,” NACA TR NACA TN-3038, 1953.
[20] Roshko, A., “Experiments on the Flow Past a Circular Cylinder at Very
also like to thank Leonard Metkowski for invaluable assistance in High Reynolds Number,” Journal of Fluid Mechanics, Vol. 10, No. 3,
setup and running experiments, as well as Michael McPhail and 1961, pp. 345–356.
Grant Dowell for their invaluable assistance in performing the doi:10.1017/S0022112061000950
particle image velocimetry measurements. The views and [21] Batham, J., “Pressure Distributions on Circular Cylinders at Critical
conclusions contained in this document are those of the authors Reynolds Numbers,” Journal of Fluid Mechanics, Vol. 57, No. 2, 1973,
and should not be interpreted as representing the official policies, pp. 209–228.
either expressed or implied, of the U.S. Government. doi:10.1017/S0022112073001114
[22] Roshko, A., “Perspectives on Bluff Body Aerodynamics,” Journal of
Wind Engineering and Industrial Aerodynamics, Vol. 49, Nos. 1–3,
References 1993, pp. 79–100.
[1] Williams, R., and Montana, P., “A Comprehensive Plan for Helicopter doi:10.1016/0167-6105(93)90007-B
Drag Reduction,” American Helicopter Society Symposium on [23] Reich, D., Elbing, B., Berezin, C., and Schmitz, S., “Water Tunnel Flow
Helicopter Aerodynamic Efficiency, National Symposium on Helicopter Diagnostics of Wake Structures Downstream of a Helicopter Rotor
Aerodynamic Efficiency, Hartford, Connecticut, March 1975; also Hub,” Journal of the American Helicopter Society, Vol. 59, No. 3, 2014,
Proceedings (A75-38340 18-05), American Helicopter Soc. Inc. Paper pp. 1–12.
13.1-13.26, New York, 1975. doi:10.4050/JAHS.59.032001
[2] Keys, C., and Wiesner, R., “Guidelines for Reducing Helicopter Parasite [24] Hufnagel, K., “Force and Moment Measurement: Steady and Quasi-
Drag,” Journal of the American Helicopter Society, Vol. 20, No. 1, 1975, Steady Measurement,” Handbook of Experimental Fluid Mechanics,
pp. 31–40. edited by C. Tropea, A. Yarin, and J. Foss, Springer, Würzbur, Germany,
doi:10.4050/JAHS.20.31 2007, pp. 592–593.
[3] Keys, C. N., and Rosenstein, H., “Summary of Rotor Hub Drag Data,” [25] Dombroski, M., and Egolf, A., “Drag Prediction of Two Production
NASA TR NASA CR-152080, 1978. Rotor Hub Geometries,” American Helicopter Society International
[4] Churchill, G., and Harrington, R., “Parasite-Drag Measurements of Five 68th Annual National Forum, American Helicopter Soc., Washington,
Helicopter Rotor Hubs,” NASA TM 1-21-59L, 1959. D.C., 2012, pp. 1–9.
[5] Young, L., Graham, D., and Stroub, R., “Experimental Investigation of [26] Fung, Y., “Fluctuating Lift and Drag Acting on a Cylinder in a Flow at
Rotorcraft Hub and Shaft Fairing Drag Reduction,” Journal of Aircraft, Supercritical Reynolds Numbers,” Journal of Aerospace Sciences,
Vol. 24, No. 12, 1987, pp. 861–867. Vol. 27, No. 11, 1960, pp. 901–813.
doi:10.2514/3.45530 [27] Richter, A., and Naudascher, E., “Fluctuating Forces on a Rigid Circular
[6] Stroub, R., Young, L., Graham, D., and Louie, A., “Investigation of Cylinder in Confined Flow,” Journal of Fluid Mechanics, Vol. 78, No. 3,
Generic Hub Fairing and Pylon Shapes to Reduce Hub Drag,” NASA 1976, pp. 561–576.
TM-10008, 1987. doi:10.1017/S0022112076002607
[7] Graham, D., Sung, D., Young, L., Louie, A., and Stroub, R., “Helicopter [28] Liskoski, D., “Nominally Two Dimensional Flow About a Normal Flat
Hub Fairing and Pylon Interference Drag,” NASA TM-101052, 1989. Plate,” Ph.D. Thesis, California Inst. of Technology, Pasadena, CA,
[8] Martin, D., Mort, R., Squires, P., and Young, L., “Hub and Pylon Fairing 1993.
Integration for Helicopter Drag Reduction,” American Helicopter [29] Taylor, G., “The Spectrum of Turbulence,” Proceedings of the Royal
Society International 47th Annual National Forum, Phoenix, AZ, Society of London, Series A: Mathematical and Physical Sciences,
May 1991; also Proceedings, Vol. 2, American Helicopter Soc. Paper Vol. 164, No. 919, 1938, pp. 476–490.
A92-14326 03-01, Alexandria, VA, 1991, pp. 897–912. doi:10.1098/rspa.1938.0032
[9] Martin, D. M., Mort, R. W., Young, L. A., and Squires, P. K., [30] Berry, J., “Unsteady Velocity Measurements Taken Behind a
“Experimental Investigation of Advanced Hub and Pylon Fairing Model Helicopter Rotor Hub in Forward Flight,” NASA TM-4738,
Configurations to Reduce Helicopter Drag,” NASA TM-4540, 1993. 1997.

You might also like