You are on page 1of 28

CHAPTER

HELICOPTER VIBRATION
REDUCTION
27 William A. Welsh
Sikorsky, A Lockheed Martin Company, Stratford, CT, United States

CHAPTER OUTLINE
1 Introduction ................................................................................................................................. 866
2 NextGen Vibration Levels .............................................................................................................. 866
3 Vibration Specifications ................................................................................................................ 866
4 Source of Helicopter Vibratory Loads ............................................................................................. 867
5 How Do Vibratory Loads Get Into the Fuselage? .............................................................................. 869
6 What Is Used for Vibration Control Now? ....................................................................................... 869
6.1 Why Not Isolation? ....................................................................................................... 870
6.2 The Venerable Frahm ................................................................................................... 871
6.3 Fuselage-Based Frahms ................................................................................................ 871
6.4 Rotor-Based Frahms .................................................................................................... 872
6.5 Frahms Are Heavy ........................................................................................................ 874
6.6 Active Vibration Control ................................................................................................ 874
6.7 Dynamic Antiresonant Vibration Isolator ......................................................................... 876
7 More Problems With Frahms ......................................................................................................... 878
8 Active Counter-Force .................................................................................................................... 879
8.1 Higher Harmonic Control .............................................................................................. 881
9 Individual Blade Control ................................................................................................................ 882
9.1 Hydraulic IBC .............................................................................................................. 883
9.2 Electrical IBC .............................................................................................................. 883
9.3 On-Blade Flaps ............................................................................................................ 885
10 The Path Forward ......................................................................................................................... 889
Acknowledgments .............................................................................................................................. 889
References ........................................................................................................................................ 889

Morphing Wing Technologies. https://doi.org/10.1016/B978-0-08-100964-2.00027-7


# 2018 Elsevier Ltd. All rights reserved.
865
866 CHAPTER 27 HELICOPTER VIBRATION REDUCTION

1 INTRODUCTION
Helicopter fuselage vibration degrades ride quality, causes crew fatigue, and damages components
necessitating expensive part replacement. Average cabin floor vibration levels in the 1960s were often
around 0.3 g resulting in a truly uncomfortable experience for crew and passengers and frequent
replacement of damaged parts [1]. With a consistent demand from the helicopter-user community,
levels have improved over the years as improved technology has been developed and applied. Now
levels are almost always between 0.1 and 0.2 g at most places in the cockpit and cabin; still not a
“Jet smooth ride,” but a significant improvement.
In this chapter, existing approaches to helicopter vibration reduction as well as new technologies
being pursued to achieve a jet-smooth ride are reviewed. (Note that the designations S-76 helicopter
and S-92 helicopter are registered trademarks of Sikorsky a Lockheed Martin Company.)

2 NextGen VIBRATION LEVELS


A good “NextGen” rule is that helicopter vibration levels should be less than 0.03–0.05 g throughout
the cockpit and cabin. These levels, would be virtually imperceptible [1] and, as shown in Fig. 1,
helicopter reliability would be significantly improved at these vibration amplitudes. Such low levels
are rarely achieved today in a consistent manner except in local areas in the helicopter cabin area
(you can always find somewhere where vibrations are low). In the following paragraphs, the current
state-of-the-art of helicopter vibration control will be summarized as well as some ideas on paths to
achieving NextGen levels.

3 VIBRATION SPECIFICATIONS
Several civil and government specifications define vibration limits quantitatively. For military projects,
the U.S. Army’s ADS-27 [2] vibration specification is sometimes used where the intent is to minimize
vibration experienced by the crew. In this context, the U.S. Army defines an “intrusion index” which is a

FIG. 1
Vibration reductions have potential to reduce helicopter operating cost.
4 SOURCE OF HELICOPTER VIBRATORY LOADS 867

parameter in units of velocity and is defined as a weighted square of the velocities in each of the orthog-
onal directions; vertical being weighted most importantly while lateral and longitudinal are weighted
successively less, reflecting the relative tolerance of humans vibration in each direction. For civil appli-
cations, the ISO-2631 specification [3] is commonly used to define the vibration levels of comfort at
various locations in an aircraft using a method not much different than that of ADS-27. The motive behind
both ADS-27 and ISO-2631, in part, is to define vibration levels that will allow the crew to perform their
tasks over some period without reduced proficiency. The FAA establishes the Federal Aviation Require-
ments or, more specifically, FAR29-251 [4] which states: “Each part of the rotorcraft must be free from
excessive vibration under each appropriate speed and power condition.” This somewhat vague guidance
makes sense because the FAA is more focused upon safe operation and less concerned about passenger
comfort or an advantage of one aircraft relative to its competitors.

4 SOURCE OF HELICOPTER VIBRATORY LOADS


A detailed description of the source of helicopter vibration is beyond the scope of the current summary
but this very rich subject is covered in detail in a few excellent books by Bramwell et al. [5], Johnson
[6], and Bielawa [7] in addition to many technical papers in the American Helicopter Society (AHS)
journal and proceedings. It is sufficient for our purposes here that the vibratory loads originate with the
main rotor and occur at the so-called blade passage frequency, that is, the rotor rotational speed times
the number of blades. Typically this frequency is between 15 and 25 Hz. The term “blade passage
frequency” is somewhat misleading because vibratory loads from the rotor are predominately transmit-
ted to the fuselage structurally via the main rotor hub, rather than by aerodynamic pressure loading
caused by blade passage over the fuselage. A secondary path is wake impingement on the vertical
and horizontal tail aerodynamic surfaces and on the tail rotor itself. Where do main rotor vibratory
loads come from? The vast change in apparent forward wind velocity over each blade as it progresses
from the advancing side, where the tip velocity is often near Mach-1 to the retreating side of the rotor
where much of the blade is stalled, causes sharp changes to loads versus blade azimuthal position
during rotation. The largest aerodynamic loads on the blade are in the so-called “flatwise” direction, i.e.,
perpendicular to the chord and it is this abrupt loading change in this direction versus azimuth that
causes the blade to vibrate at all harmonics of the rotor rotational speed. The usual nomenclature is
that loads and motions that occur at a frequency of once per revolution of the rotor are termed
“1P” loads and motions. Similarly, loads and motions that occur at the higher harmonics are termed
“2P, 3P, etc.” Luckily, and perhaps surprisingly to those unfamiliar with rotor dynamics, forces at
all frequencies except the blade passage frequency “cancel out” at the rotor hub (where the blades
are mounted). For example, for a 4-bladed helicopter, the only blade loads that eventually are trans-
mitted to the fuselage are 3P, 4P, and 5P loads but all three of these “rotating system” loads result in
only 4P loads on the “nonrotating” fuselage. In this case, the 3P blade loads originate on the rotating
blade and in the rotating system. Some simple phase plots can show that the sum of 3P loads from all 4
blades but phase lagged 90 degrees from each other, result in a single vector of 3P force (or moment)
that angularly travels in the same direction as the rotor rotation and progresses from 0 to 360 degrees at
three times per revolution in the rotating frame of reference. As a result of this forward progression in
the rotating frame, this identically sized force from each blade is transmitted to the fuselage as a
3P +1P ¼ 4P load. A similar phase plot would show that 2P loads on a 4-bladed rotor, “self-cancel.”
868 CHAPTER 27 HELICOPTER VIBRATION REDUCTION

This is a very fortunate outcome because the lower load harmonics are typically huge and would
otherwise result in potentially damaging loads and high vibration being transmitted into the fuselage.
Another nice “physics-fact” is that successive blade load harmonics become smaller with increasing
frequency, i.e., 5P loads are almost always smaller than 2P loads provided no blade resonances aggra-
vate the blade responses. Therefore, 7-bladed helicopters shake less than 2-bladed helicopters; for the
former case, the only loads transmitted to the fuselage originate from higher frequency but smaller 6, 7,
and 8P blade loads as opposed to much higher 1, 2, and 3P harmonic loads for the latter helicopter. The
blade passage loads, generically referred to as “Np” loads, where N is the number of blades, are trans-
mitted from the hub to the fuselage through the main rotor gearbox which is often rigidly attached to the
top of the fuselage to provide a reliable and robust connection; an important feature, since the very large
main rotor flight loads, i.e., the gross weight of the vehicle times any vehicle acceleration must pass
through this same connection.
The fixed system Np loads can be viewed as six independent loads, three forces and three
moments, as shown in Fig. 2. Any single one of these six loads can be large enough to cause unde-
sirable fuselage vibration but superimposed together, especially for helicopters with six or fewer
blades, some vibration mitigation is often required to make the vehicle practically useful. An alter-
native way of visualizing the vibratory hub loads is shown in Fig. 3. The three forces form a vector
whose tip orbits an ellipsoid centered at the rotor hub while the three moments form a similar but
differently shaped ellipsoid. Each ellipsoid changes shape vs. flight condition. In the discussion that
follows, the reader might find both ways of viewing the Np loads useful when considering the
methods for eliminating their effect. Suffice it say now that if a clever engineer could counter each
of these six independent vibratory loads, preferably at the rotor hub, the helicopter dynamicist’s task
would be over!

FIG. 2
Primary vibration path is the main gearbox which transmits six vibratory loads to the fuselage.
6 WHAT IS USED FOR VIBRATION CONTROL NOW? 869

FIG. 3
Vibratory loads and movements produced by a rotor follow an ellipsoid.

5 HOW DO VIBRATORY LOADS GET INTO THE FUSELAGE?


In virtually all helicopters the Np hub loads, both forces and moments along with more-or-less qua-
sisteady flight loads are directly transmitted to the fuselage through the main gearbox (MGB),
which is, itself, very rigid and, of course, strong enough to withstand very large low-cycle flight
loads and much smaller Np loads. But this nexus between the MGB and the helicopter fuselage is
often where the big differences between antivibration approaches of various helicopter models
become evident.

6 WHAT IS USED FOR VIBRATION CONTROL NOW?


The “fielded” approaches used for vibration control will be discussed first to be followed by approaches
being studied for the future. To clarify the current state-of-the-art, various classes of antivibration ap-
proaches are shown in Table 1. The author notes that this table is limited by his knowledge of what is
currently fielded. The entries that are, to the author’s knowledge, not fielded on any production aircraft
are labeled with the abbreviation “R&D” indicating that these have been studied in the research and
development context only. In Table 1, note the use of the word “Frahm” named after Hermann Frahm
who invented the “vibration absorber” in 1890 (US989958 A). A Frahm is a synonym for tuned vibra-
tion absorber; essentially a mass and spring tuned to the vibration forcing frequency. To tune a Frahm,
the spring is anchored to a virtually immovable table and the spring-mass assembly is tuned to the exact
forcing frequency that the engineer wants to suppress. Contradictory as it seems this process actually
produces an antiresonance or “bucket” at the forcing frequency when the device is mounted in the he-
licopter fuselage. The result is that the Frahm produces very low vibration at its connection point on the
fuselage, while the Frahm mass vibration amplitude is very large. Although beyond our scope for the
current purposes to go into the physics of how this device works, suffice it to say that Frahms, or some
variation of the idea, are used in a huge number of helicopters.
870 CHAPTER 27 HELICOPTER VIBRATION REDUCTION

Table 1 Overview of Vibration-Reducing Methods Used in the Helicopter Industry


Company Isolation Fuselage Suppression Rotor Suppression
Passive Active Frahm Active Passive Active

Sikorsky None Zero-Vibe UH-60L UH-60M UH-60L, HHC, IBC,


R&D M (Bifilar) active flap
R&D
Bell Main gearbox Active LIVE Bell-430 Bell-525 Bell-412 None
(MGB) R&D Bell-429 Bell-429 (pendular
isolation long (R&D) absorber)
springs, LIVE V-22 (pendular
for vertical absorber)
and roll
Boeing CH-47, floor None CH-47 CH-47 HHC and
isolation self- active flap
tuning R&D
absorber
Airbus H125/H130/ H135 (ARIS, H125 H22 H145 (out-of- Active flap
AS365 hydraulically H120 H130T2 plane Blue Pulse,
(barbeque actuated pendulum) IBC R&D
plate some DAVI H120 (in-plane
with principle) pendulum)
elastomer
mounts
instead of a
metallic
plate)
Tiger, NH-90
SARIB

6.1 WHY NOT ISOLATION?


The perceptive reader might wonder why simple elastomeric or metal isolation mounts, commonly
used for isolating foundations from industrial equipment vibration, are not used on helicopters to
completely isolate the fuselage from the MGB. The reason this is impractical is twofold. (1) Compared
to industrial equipment, the MGBs of helicopters are designed to be very light for a given power rating;
it is, after all, on a flying vehicle and every extra kilogram lessens the aircraft payload. (2) The Np
frequency is commonly low being between 15 and 25 Hz. For good isolation, the natural frequency
of the MGB on isolation mounts would need to be well below the Np frequency for all six possible
translations and rotations of the MGB; say 1/2 of the Np frequency, to provide good isolation from
the fuselage. This combination of low MGB mass and low required natural frequency, results in the
need for very soft isolator spring rates. Low mounting stiffness would allow very large MGB motions
under quasisteady flight loads. By themselves, large motions would not cause this approach to be im-
practical but one other fact ruins our hopes. In virtually all modern helicopters, gas-turbine engine(s),
mounted to the fuselage, drive the MGB via high-speed drive shaft(s) running at speeds that vary from
6 WHAT IS USED FOR VIBRATION CONTROL NOW? 871

6000 to 22,000 rpm. The MGB reduces this speed to a few hundred revolutions per minute to drive the
main rotor. Couplings on the ends of the high-speed drive shafts are almost always designed to handle
typical engine-to-MGB misalignments that occur due to installation tolerances but have limited ability
to withstand relatively large quasisteady motions necessarily allowed by MGB isolation. Hence, to the
author’s knowledge, pure MGB isolation has never been successfully implemented in all six degrees-
of-freedom on any helicopter.
Having discouraged the reader regarding six-degree-of-freedom isolation, we should not jump to
the conclusion that isolation is never attempted. The use of the so-called “DAVI” or Dynamic Anti-
resonant Vibration Isolator is a particularly interesting and widely used example which will be
described later.

6.2 THE VENERABLE FRAHM


As mentioned earlier, most large helicopters are designed with a very stiff connection between the MGB
and the fuselage. Aside from its obvious simplicity and weight efficiency, this approach also has the ben-
efit of simplifying the connection of the MGB to the engine(s), i.e., excessive relative deflections between
the MGB and engine are not a concern. Sikorsky and Boeing are two companies that generally follow this
approach. Direct connection has the disadvantage of enabling the Np loads to easily get into the fuselage.
Many companies attack this problem with a two-pronged strategy, by (1) countering the largest rotor
in-plane vibratory loads at the rotor head (where the blades are connected) with special rotating Frahms
of various styles and (2) suppressing the fuselage vibration caused by the remaining loads with Frahms in
the airframe. In the next sections, we will discuss fuselage-based Frahms followed by a description of
rotor-based Frahms.

6.3 FUSELAGE-BASED FRAHMS


From the early days of the helicopter industry, fuselage and rotor-based Frahms have been used to
lessen vibration and are still in wide use today. The author’s search for the earliest patents using Frahms
on helicopters reached back to 1945 where pendular absorbers were used to lessen control vibrations in
helicopters (US2489342). The imperative for vibration suppression on helicopters is revealed by the
fact that this patent was granted only 5 years after the first flight of the Vought-Sikorsky VS-300;
widely considered the first demonstration of a practical helicopter. One of the earliest uses of Frahms
on production helicopters was in the Army’s Chinook CH-47B helicopter in the early 1960s. This
40,500 lbm helicopter with two 3-bladed rotors, incorporated five absorbers each weighing 90 lbm;
in total 1.1% of the aircraft gross weight. For the dynamics aficionado, the absorbers were arranged
with one in the nose, one each under the pilot and copilot and two in the aft pylon. In the CH-47C,
D, these fixed-tuned absorbers were replaced with three Self-Tuning Vibration Absorbers [8–10],
or “STVAs” each weighing 170 lbm. About 10 years later, three Frahms were used on the UH-60A
Black Hawk Helicopter. These absorbers were a clever design that used three nested leaf springs;
the airframe-end was connected to the fuselage with a “simple” support while the other side of the leaf
springs was cantilevered to the absorber mass. This approach was invented by John Marshall [11]. To
lessen the effect of airframe deformations on damping, these were later replaced by self-contained “box
frame” absorbers on the UH-60L as shown in Fig. 4. A perusal of this figure reveals three variations of
fuselage absorbers; two of the designs are box-frames and one design is a single leaf spring type. Use of
872 CHAPTER 27 HELICOPTER VIBRATION REDUCTION

Bifilar

Nose Vertical Stub Wing Forward Cabin Vertical


4P Absorber 4P Absorber 4P Absorber

FIG. 4
Bifilar and Frahms on a UH-60L Black Hawk.
Taken from W.A. Welsh, Evolution of active vibration control technology, in: Proceedings of the American Helicopter
Society, 4th Decennial Conference on Aeromechanics, January 21, 2004.

Frahms, in one form or another continues on currently produced helicopters but the fuselage-based
versions are slowly being replaced by active systems. A perusal of the weights given above will reveals
why absorbers are becoming obsolete; Frahms frequently make up from 1% to 2% of the helicopter
gross weight. Given the relentless drive for more vehicle productivity, they are a prime target for
weight reduction. The remaining Frahms in helicopters are on the helicopter rotors; a subject which
will be covered next.

6.4 ROTOR-BASED FRAHMS


Clearly, to eliminate the vibratory loads emanating from the rotor, one would like to invent devices that
produce loads with a similar ellipsoidal shape as previously discussed but in the “negative direction,”
then place these devices very near the source of rotor Np load generation, i.e., on the rotor head. Such
6 WHAT IS USED FOR VIBRATION CONTROL NOW? 873

counter-force devices capable of producing the negative ellipsoid do not yet exist but some approxi-
mations to these are in service. One example of a rotor-head mounted device is the bifilar. Many Si-
korsky aircraft have one or two bifilars like the model shown in Fig. 4, which was invented by William
Paul and Ken Mard of Sikorsky [12,13], in 1968 and 1969, respectively. A bifilar is a type of Frahm or
pendular absorber where the Frahm mechanical springs are replaced with a “centrifugal” spring. Since
the “spring” rate is now not fixed but rather, is proportional to rotor speed, the bifilar, as well as the
more conventional pendular absorber, has the very nice feature of automatically remaining tuned to Np
if the rotor speed changes. This can be seen by forming the natural frequency of a pendulum intended to
reduce 3P rotating system loads as follows with: L ¼ pendulum length, 3Ω ¼ rotor speed times the
harmonic to suppress, mpendular ¼ the pendulum mass, and g ¼ gravitational constant. Notice that
3Ω is squared but is also inside of the square root, i.e., the natural frequency increases proportionately
with rotor speed. Also, notice that the pendular mass term cancels, i.e., the natural frequency is not
sensitive to the pendular mass.
 1=2
ωn ¼ mpendular ∗ L ∗ g ∗ ð3ΩÞ2 =mpendular

Bifilars as defined by the original inventors suffer from the same problem that tends to plague pen-
dular absorbers, i.e., they tend to detune with increasing mass amplitude. To properly tune a conven-
tional bifilar and pendulum to the Np  1 or Np + 1 rotating system frequency, the c.g. of the somewhat
large pendular mass must only be a short distance from the pivot point resulting in a very short pen-
dular “arm.” But as the mass oscillates to larger amplitudes, with an increasing forward speed of the
helicopter, this large motion violates the small angle assumption resulting in a detuned natural fre-
quency and decreased effectiveness. This detuning effect was solved by John Madden [14] by adding
a special cycloidal shape to the bifilar bearings thus resulting in a natural frequency that is much less
sensitive to mass amplitudes. The author finds it interesting that the peculiar bearing design used to
retain the bifilar mass against centrifugal loading that was invented in 1969 also enabled the innova-
tion of using cycloid bearings invented in 1980. It is doubtful (in the author’s opinion) that the notion
of using cycloidal bearings would have even occurred if conventional roller or ball bearings continued
to be used to retain the mass as they are used in conventional pendular absorbers.
A bifilar (or pendulum) tuned to a single frequency creates a circular load (not an ellipse) that coun-
ters some of the in-plane rotor vibratory loads. Two bifilars, tuned to Np  1 and Np + 1 frequencies can
produce an ellipse in the plane of the rotor; but alas, not a tilted ellipsoid. Nevertheless, single or dual
bifilars are a good compromise because this arrangement cancels one or two in-plane hub loads, re-
spectively, while, regrettably, still allowing the vibratory vertical load, pitching and rolling moments
to “leak” into the fuselage. Unfortunately, no currently available single device can cancel all three
loads and three moments. The resulting residual fuselage vibration is, historically, lessened with
one or more Frahms mounted in the airframe.
As previously alluded to, another example of a rotor-head mounted device is the pendular absorber
which is used in the Bell 412 (Fig. 5) and the Boeing/Bell V-22 [15] (Fig. 6). These devices are simple
masses on the end of a short pendulum arm that are mounted to the helicopter rotor head. As the rotor
spins, the centrifugal force causes the mass to experience a virtual spring force in a manner very similar
to that of the bifilar. The mass and pendulum arm are adjusted to produce a Frahm antiresonance effect
either in or out of the rotor plane. For the case of the Bell 412, the pendular absorbers act
874 CHAPTER 27 HELICOPTER VIBRATION REDUCTION

FIG. 5
Bell 412 pendular absorbers.

perpendicularly to the rotor plane (Fig. 5) and are intended to suppress Np vibration in both the rotating
and fixed frames of reference while the V-22 pendular absorbers act in-plane (Fig. 6) and suppress
Np  1 vibration in the rotating frame of reference and Np in the fixed frame.

6.5 FRAHMS ARE HEAVY


Some weight estimates of Frahm-based systems can be gleaned from Fig. 4. The UH-60L system
weights, per this figure, make up about 1.5% of the helicopters gross weight; or about 300 lbm for
a 20,000 lb helicopter. The V-22 suffers from a smaller weight penalty as indicated by Rangacharyulu
et al. [15], each pendulum mass weighing 40 lbm (there are six of them). Adding this to the weight of
the fuselage based antivibration system, brings the total up to approximately 350 lbm. But the V-22 is a
40,000–65,000 lb aircraft so the Frahm weights are only 0.7% of the gross weight.

6.6 ACTIVE VIBRATION CONTROL


More recently, in the UH-60M, for example, fuselage mounted Frahms have been replaced with active
vibration systems where active vibration control (AVC) actuators are placed where the previous
Frahms resided. In this AVC system, the control computer acquires accelerations for 10 locations
in the fuselage and optimally commands the AVC antivibration actuators to produce vibratory
forces at Np to suppress these accelerations. The self-adaptive, real time, frequency domain controller
implemented in the UH-60M AVC system is derived from the seminal work cited by Taylor et al. [16].
In this reference, the control methods developed by J. Molusis and F. Farrar are described in detail. In
fact, the control algorithm, or derivatives of it, are probably being used in the vast majority of the
fielded AVC systems and is generally the algorithm used for rotor-based active systems to be discussed
later (IBC and HHC).
6 WHAT IS USED FOR VIBRATION CONTROL NOW? 875

(A)

(B)
FIG. 6
(A) V-22 hub pendulum absorbers. (B) V-22 hub pendulum absorbers.
(A) Taken from M.A. Rangacharyulu, Bell Helicopter Textron Inc., M.J. Moore, Boeing Helicopter Company, Flight vibration testing
of the V-22 tiltrotor aircraft, in: Proceedings of the 47th Annual Forum of the American Helicopter Society, Phoenix, AZ, May 6–8, 1991.
(B) Courtesy of the American Helicopter Society International.
876 CHAPTER 27 HELICOPTER VIBRATION REDUCTION

FIG. 7
Fuselage based active vibration system (AVC) on the Sikorsky S-92 helicopter.
Taken from W.A. Welsh, Evolution of active vibration control technology, in: Proceedings of the American Helicopter Society,
4th Decennial Conference on Aeromechanics, January 21, 2004.

The AVC system, one example of which is shown in Fig. 7, has significant advantages over
fuselage based Frahms as shown by the author [17]; specifically, the AVC system can suppress vibra-
tion where the crew and passengers actually sit. The Frahms, being passive, can only suppress vibration
at their respective mounting locations. This can be a serious limitation; if there is no physical room to
mount Frahms near where the engineer would like to reduce vibration, they must be mounted at alter-
native locations, which are often less effective and less weight-efficient than desired. Thus, fuselage-
based AVC is more weight efficient, i.e., lighter AVC antivibration actuators reduce fuselage vibration
to the same or lower levels than Frahms systems with lighter and/or fewer AVC actuators (the heaviest
components in AVC systems).

6.7 DYNAMIC ANTIRESONANT VIBRATION ISOLATOR


As stated earlier, no successful six-degree-of-freedom isolation system has ever been implemented in
production helicopters. But very successful isolation systems do exist for eliminating the largest vibra-
tory loads that would otherwise be transmitted into the fuselage. For example, some helicopter
6 WHAT IS USED FOR VIBRATION CONTROL NOW? 877

companies, like Bell Helicopters [18] position very clever “isolation” devices called LIVE (Liquid In-
ertia Vibration Eliminator) devices between the main transmission and the fuselage to interrupt the
transmission of Np loads into the fuselage while allowing very reliable transfer of flight loads into
the fuselage; obviously, a key aspect of safe helicopter operation. This idea was first developed by
Kaman helicopter company and was called the “DAVI” or Dynamic Antiresonant Vibration Isolator
and was invented by William Flannelly from Kaman Corporation [19]. LIVE is a derivative of this
device, and is shown on a Bell-429 taken from Fig. 8 by Riedel [20]. It is placed in the main load path
between the helicopter main transmission and the airframe and produces a “notch” or antiresonance in
the load between the two. The DAVI was a purely mechanical device but subsequent refinements and
advancements of the basic concept has resulted in the LIVE that is equivalent from a physics standpoint
but replaces the DAVI’s oscillatory mass with a heavy liquid and replaces the mechanical lever and
bearings with pistons of differing areas. Presumably, this change improved reliability and certainly
resulted in a more elegant, sealed, and compact device. On many Bell Helicopters, LIVE devices
are used to lessen vertical and roll vibratory load transmission into the fuselage while passive isolation
is used to lessen the effect of vibratory pitching moments and, finally, Frahms are sometimes used to
lessen the residual vibration which still finds its way into the airframe.

Upper fluid reservoir

Outer cylinder
Airframe reaction
Inner cylinder

Transmission load Elastomeric elements

Tuning port
Airframe reaction
Lower fluid reservoir

Fp sin wt
up

mp b
ut
K
mt
uf a

mf

FIG. 8
Liquid Inertia Vibration Eliminator (LIVE) system.
Taken from K. Riedel, Vibration analysis and testing of Bell-429 helicopter, in: Proceedings of the American Helicopter
Annual Forum, Phoenix, AZ, May 11–13, 2010.
878 CHAPTER 27 HELICOPTER VIBRATION REDUCTION

Pivot

GB strut Support
MGB
Weight

Fitting Fuselage Membrane

SARIB leaf
Flapping mass

GB strut

Membrane
SARIB leaf

FIG. 9
Eurocopter’s (now Airbus) SARIB gearbox isolation.
From P. Konstanzer, B. Enenkl, Recent advances in Eurocopter’s passive and active vibration control, in: Proceedings of the
American Helicopter Society 64th Annual Forum, Montreal, Canada, April 29–May 1, 2008.

In addition to LIVE, there are quite a few other variants of the original DAVI concept implementing
innovations to adapt to the needs of a given helicopter. In the NH-90 Helicopter the so-called Système à
Resonateurs Integres dans les Barres or “SARIB,” shown in Fig. 9 taken from Konstanzer and Enenkl
[21] is such a variant. Another variant is the “IRIS” (Improved Rotor Isolation System), developed by
R. Desjardins and W.E. Hooper at Boeing [22,23].

7 MORE PROBLEMS WITH FRAHMS


As cited earlier, the venerable Frahm has been used (by the thousands) very successfully in heli-
copters for many years being distributed throughout the fuselage or, in their rotating-system ver-
sion, as bifilars or pendulum absorbers on the rotor head. From the 1960s to now they have enabled
8 ACTIVE COUNTER-FORCE 879

vibration levels of 0.15 g or even lower, but we need to improve upon this for NextGen helicopters.
Unfortunately, fuselage-mounted Frahms and AVC approaches both suffer from one common prob-
lem that ultimately limits their ability to reduce vibrations down to our target 0.03 g.
Both approaches produce an effect called “modal spillover” and to understand this effect, it helps
to view the fuselage dynamics from a “modal” and control system perspective. That is, fuselage
vibration can be viewed as the superimposed response of an infinite number of natural frequencies
and modes. Since the excitation is primarily due to rotor loads at Np, only a few of these modes
really respond with significant amplitude; typically six or fewer modes are important. Conse-
quently, the hopeful aircraft dynamicist/control engineer might conclude that these six modal de-
grees of freedom could be completely nullified by six “controls,” which in this discussion, are
represented by Frahms or AVC actuators distributed in the fuselage. Certainly, if six modes are
responding to the hub loads, it is necessary for any control system (Frahms or AVC actuators)
to have, at least, an equal number of controls to have any hope of nullifying the modal motion
but, unfortunately this is not a sufficient condition. Focusing upon Frahms for a moment, six de-
vices would surely produce virtually zero vibration but only at their six physical mounting loca-
tions. With some luck, low vibrations (0.1–0.15 g) can be obtained at locations other than at the
Frahms mounting points. Unfortunately, some locations in the airframe will experience much
higher levels to the point of making these seats the least favorite for passengers. This unfortunate
effect occurs because while reducing vibration of the modes excited by rotor Np loads, Frahms also
produce concentrated loads which excite modes that would not otherwise be excited by rotor Np
loads in the first place. This effect is called “modal spillover” and the consequence is that vibrations
everywhere in the fuselage cannot be reduced to the very low levels required for NextGen helicop-
ters. A caveat to this; given a large enough number of antivibration devices mounted in the
fuselage, i.e., Frahms or AVC actuators, NextGen vibrations can be achieved but experience dem-
onstrates that the resulting weight increase would significantly lessen aircraft payload.
So, distributing antivibration devices in the airframe is destined to fail as a NextGen antivibration
approach. But, following the same modal logic as previously indicated, it is also true that if antivibra-
tion loads are applied to the airframe in the same manner as the rotor Np loads, one would expect that
the identical modes (to rotor Np loads) would be “excited,” i.e., without modal spillover. The hoped-for
result, would be global vibration reduction, i.e., one where the entire helicopter stops vibrating thus
giving both passengers and equipment a smooth, damage-free ride.

8 ACTIVE COUNTER-FORCE
With the advent of rare earth magnets and high speed computer control, brushless electric motors have
enabled some exciting developments in AVC. For example, it is now within the realm of possibility that
all six hub loads comprising the elliptical figures previously described can be more or less directly
suppressed by applying “equal and opposite” counter forces. This is still an area of active research
as can be seen in Fig. 10, which was taken from Andrews [24]. In this Zero-Vibe system, active devices
supplied by LORD Corporation are grouped around the main transmission in a manner that they can
supply countering loads to every one of the six degrees-of-freedom. The system is comprised of a Hub
Mounted Vibration Suppressor (HMVS) which is an active analog of two bifilars, i.e., it can create any
880 CHAPTER 27 HELICOPTER VIBRATION REDUCTION

(A)

(B)
FIG. 10
(A) Sikorsky’s Zero-Vibe system. (B) LORD Circular Force Generators and Integrated Control Unit.
(A) Taken from J. Andrews, W.A. Welsh, R. Altieri, J. DiOttavio, Ground and flight testing of a hub mounted vibration suppression
system, in: Proceedings of the American Helicopter Society 70th Annual Forum, Montreal, Quebec, Canada, May 20–22, 2014.
(B) Reprinted by permission of LORD Corporation.

elliptical in-plane counter-force to that produced by the main rotor. Other AVC actuators grouped
around the main transmission supply the other counter-force and counter-moment loads. The Zero-
Vibe intent is to avoid the modal spillover effects described in the Frahm discussion by applying
counter-force loads at virtually the same location as does the main rotor thus exciting, or in this case,
8 ACTIVE COUNTER-FORCE 881

suppressing the same modal responses. A clarification is in order here; the newest actuators are still too
large to literally place all six of them on the main rotor hub where the blade loads are applied, but this
does not cause a problem because the main transmission, including the main rotor hub together, are a
virtual rigid body. So, these six actuators can be placed either on the rotor hub or, alternatively, on or
near the gearbox without suffering from being somewhat remote from the main rotor head. In fact, as a
practical matter, actuators that produce a single directional counter-force must be placed a reasonable
distance apart to act through a “moment arm” to produce a counter-moment thus necessitating being on
opposite sides of the physically large part of the rigid body, i.e., the main transmission. A relatively
recent addition to the helicopter world is the so-called “Circular Force Generators” (CFGs) by LORD
Corporation shown in Fig. 10B. These novel devices produce a “circular” force of adjustable magni-
tude and phase. By adjusting these two parameters for six CFGs, it is possible to completely nullify the
motion of a rigid body albeit the reader might find it hard to visualize how six whirling vectors, each
with a controlled magnitude and phase, can add up to nullify the six hub loads (but it does work).

8.1 HIGHER HARMONIC CONTROL


A dream of helicopter dynamicists is to oscillate the main rotor blades in a manner to reduce the Np
loads transmitted into the fuselage. The idea promises to achieve vibration reduction at the source (i.e.,
the rotor) and with low weight penalty since no extra actuators need be added to the aircrafts empty
weight.
The first active rotor approach, tested in the 1960s (Warnicke and Drees [25]) on a 2-bladed UH-1
relied upon oscillating the existing main rotor blades through the rotor control system; a method known
as Higher Harmonic Control (HHC). Typically, helicopters have three hydraulic servos that move the
swashplate which, in turn, steers the main rotor according to the input from the pilot’s flight controls.
Normally, these servos move slowly in response to pilot inputs but, with HHC, the servos are oscillated
at Np. The resulting rotor blade oscillations modify the Np vibratory loads that are transmitted to the
fuselage. This approach has also been tried in many wind tunnel tests, for example by Shaw et al. [26]
and some flight testing by Wood et al. [27] in 1985 and Straub and Byrnes [28] in 1986 on a 4-bladed
OH-6A. In 1986, Miao et al. [29,30] tested HHC on an S-76 helicopter. This system is shown in Fig. 11.
Again, the story is not an entirely happy one. In these references, it is shown that HHC is able to reduce
vibrations in the fuselage but not to NextGen levels. Also, in some cases [29], at least for helicopters
with articulated rotors as in this reference, the reduction came at the price of an extravagant expenditure
of power in the form of hydraulic flow rates at high pressure (typically 3000 psi). This is a technical
roadblock that has yet to be overcome and is probably responsible for the fact that no commercial or
military helicopter has been equipped with HHC even 50 years after the approach was first attempted.
What could overcome this roadblock? It is interesting to note that very little net energy is needed to
oscillate a rotor blade. It surely takes considerable work to increase the pitch of a rotor blade because of
the powerful “propeller” returning movement exerted by the spinning blade but, at least ideally, the
blade will gladly return this work when pitch is reduced to zero. Unfortunately, because of their op-
erating principle, typical hydraulic systems expend the same amount of work on both the up-stroke and
down-stroke although the latter is, ideally, unnecessary. Perhaps a regenerative hydraulic system would
overcome this problem. Another, more fundamental limitation of HHC is that, on a conventional
helicopter, there are only three main rotor servos. Having only three controls necessarily limits the abi-
lity of HHC alone to suppress three hub loads or three modes or vibration at three locations. Clearly, for
882 CHAPTER 27 HELICOPTER VIBRATION REDUCTION

FIG. 11
Higher Harmonic Control (HHC) system.
Taken from W. Miao, S.B.R. Kottapalli, M.M. Frye, Flight demonstration of higher harmonic control (HHC) on S-76, in: Proceedings of the
42nd Annual Forum of the American Helicopter Society, Washington, DC, June 1986.

NextGen aircraft, we must suppress all six hub loads or hope that the helicopter only has three signi-
ficant hub loads in the first place; not usually a good assumption. One potential solution to this
controllability deficit is to use HHC to counter three of the hub loads, say vertical forces and pitching
and rolling moments and augment HHC with some other actuation technology to attack the remaining
loads, i.e., in-plane rotor loads: Fx, Fy, and Yaw moment: Mz. For example, the dual bifilar arrange-
ment mentioned earlier could counter Fx and Fy. Another possibility is to augment HHC with a modern
analog to the bifilar; the so-called Hub Mounted Vibration Suppressor [24] (HMVS) to be discussed
later. The reader interested in further information regarding active rotor control should see the article by
Dr. U. Arnold in the current book. Additionally, see Jacklin [31], for an excellent discussion and list of
references regarding HHC and the next subject, IBC.

9 INDIVIDUAL BLADE CONTROL


An alternative approach to suppressing vibratory loads closer to the source was examined by Ham and
Quackenbush [32] in the 1980s who proposed so-called “IBC” or Individual Blade Control for stall and
vibration alleviation. The idea behind IBC, at least for vibration control, is to oscillate the entire blade
at Np  1, Np, and Np + 1 frequencies in the rotating frame of reference thus producing Np in the
nonrotating frame of reference, i.e., the fuselage. IBC has an advantage over HHC because the blade
9 INDIVIDUAL BLADE CONTROL 883

pitches can also be oscillated at Np  2 and Np + 2 which some researchers believe will improve rotor
forward flight performance, i.e., L/De. This approach has, sporadically, seen about 40 years of exper-
imentation in wind tunnels. See Jacklin et al. [33] for a BO-105 rotor IBC test, Jacklin et al. [34]
(Fig. 12A) for a UH-60 rotor IBC test, at the Ames 80  120 wind tunnel. In the latter of these tests,
the IBC actuators were designed and produced by ZF Luftfahrttechnik and are shown in Fig. 12B from
Norman [35], and some results are given in Fig. 13 for a similar test at the Ames 40  80 wind tunnel
taken from the same reference. Also, see IBC flight testing in Schimke et al. [36] on a BO-105 and Furst
[37] for flight testing on a CH-53G. For readers who might want a more in-depth review of IBC tech-
nology and history, the author suggests the excellent paper by one of the pioneers in this area,
P. Friedmann [38].

9.1 HYDRAULIC IBC


The approach in the testing [33,36,37] was to replace the usually rigid blade pitch control rods with
hydraulic actuators; the latter is needed because the blade exhibits powerful propeller moments that
require actuators with high-force capability. The motions, i.e., blade pitch inputs are also quite large,
i.e., on the order of one degree; thus, eliminating any piezo-electric approach. The antivibration
performance seen in the aforementioned testing was quite good albeit not to the desired NextGen stan-
dard of 0.03 g throughout the aircraft. Despite the partial success seen in wind tunnel and flight testing,
the long history of work on this concept has shown that IBC, at least in the manner it was implemented,
is impractical. The most intractable roadblock, which HHC did not have, is the need to get hydraulic
power into the rotating system through some sort of hydraulic slip ring. Some researchers have had
good luck with hydraulic slip rings during extended test programs but the authors experience has
not been good and it is his opinion that no practical hydraulic slip ring has been invented which
can withstand the combination of high pressure and high seal speeds and, at the same time, is reason-
ably reliable over a long period. IBC also suffers from the same potential for high power usage and
controllability problems as does HHC. The power usage can be solved, just as for HHC, by regeneration
and the controllability issue would yield to the same augmentation suggested for HHC, i.e., add more
antivibration actuators to nullify the remaining hub loads.

9.2 ELECTRICAL IBC


Faced with the IBC and HHC road blocks, the research community has appeared to shift their focus to
electrically powered IBC and on-blade devices to produce the same result, i.e., reduce fuselage vibra-
tions by modifying Np loads transmitted by the rotor to the fuselage. Electrical IBC solves two of the
problems exhibited by the hydraulic version: it eliminates the need for a hydraulic slip ring and also has
the potential for lessening the power required. Electrical IBC actuators require an electrical slip ring
but, unlike hydraulic slip rings, electrical slip rings are a proven technology, after all, many helicopters
have de-icing systems that require large amounts of electrical power to be transmitted to the blades in
order to heat them and preclude dangerous ice buildup. Also, as mentioned earlier, hydraulic actuators
require high pressure fluid to both pitch the blade up but also to allow the blade to pitch down. The latter
power requirement is unnecessary from a purely physics-based viewpoint because the blade propeller
moments will cause the blade to return to flat-pitch with very little real need for any work to be
expended. Electrical IBC solves this problem because they have the potential for “regeneration.”
That is, after work is done by the electrical actuator on the blade to pitch it up, the blade can then
884 CHAPTER 27 HELICOPTER VIBRATION REDUCTION

(A) UH-60A rotor with IBC


actuators

Pitch horn
Rod end

Actuator
piston
Main
Safety cylinder
cylinder housing
housing
Servo valve

Filling
value Hydraulic
fitting

LVDT Electrical
housing connector
Swashplate
UH-60A rotor with IBC Rod end
(B) actuators
IBC actuators
developed by ZF
Luftfahrttechnik
FIG. 12
(A) Individual Blade Control (IBC) system testing on the UH-60A rotor in the Ames 80  120 wind tunnel. (B) IBC
actuators designed and produced by ZFL.
(A) Taken from S.A. Jacklin, A. Haber, G. de Simone, T.R. Norman, C. Kitaplioglu, P. Shinoda, Full-scale wind tunnel test

of an individual blade control system for a Uh-60 helicopter, in: Proceedings of the American Helicopter 58th Annual Forum, Montreal,
Canada, June 11–13, 2002.
9 INDIVIDUAL BLADE CONTROL 885

Normal force Normal force


Axial force Axial force
Side force Side force
Pitch moment Pitch moment
Amplitude

Amplitude
Roll moment Roll moment
Yaw moment Yaw moment

0 1 2 3 4 5 6 7 8 9 10 11 12 0 1 2 3 4 5 6 7 8 9 10 11 12
N/rev harmonic N/rev harmonic
Vibration spectrum at 46 kts Vibration spectrum at 46 kts with 1.0 degree of 3/rev
IBC applied at 315 degree phase angle.
FIG. 13
IBC results in NFAC wind tunnel.
Taken from S.A. Jacklin, A. Haber, G. de Simone, T.R. Norman, C. Kitaplioglu, P. Shinoda, Full-scale wind tunnel test of
an individual blade control system for a Uh-60 helicopter, in: Proceedings of the American Helicopter 58th Annual Forum,
Montreal, Canada, June 11–13, 2002.

“back-drive” the actuator turning it into a virtual generator and thus recovering some of the power
expended on the up-stroke. Various attempts at electrical IBC have been and still are being researched.
Although no flight testing of this approach has yet been done, some notable research is being pursued,
primarily by ZF Luftfahrttechnik GmbH [39,40].

9.3 ON-BLADE FLAPS


Another approach to electrical IBC is to activate a discrete on-blade trailing edge flap that morphs the
blade instead of pitching the entire blade. One beauty of this method is that the actuators can be very
small and light as compared to those required to pitch the entire blade. In fact, the Kaman company has
always used trailing edge flaps which they call “servo-tabs” to provide primary pitch control for their
helicopters. That is, they use the servo tabs to actually steer the helicopter in maneuvering flight. The
servo-tab adds a second beauty to the on-blade flap approach because it is a so-called, balanced
device, i.e., the servo-tab is hinged behind its leading edge as shown in Fig. 14. This clever arrange-
ment, which is well-known to fixed wing airplane designers, enable servo-tabs to be actuated with very
little effort, not only because they represent a small portion of the blade chord but also because they are
subjected to only small aerodynamic pitching movements themselves. The latter is because the hinge is
placed at the tab’s quarter chord where the summation of the lift loads on the tab acts to produce pitch-
ing moments that are typically small. In fact, on Kaman’s helicopters, the rotor is controlled via various
mechanical control linkages, by the pilot’s and copilot’s efforts without the hydraulic boosting that is
virtually universally used in the remainder of the helicopter world. On the negative side, that portion of
the servo-tab which is ahead of its hinge, protrudes into the high-speed air traveling over the surface of
the blade which, undoubtedly, introduces a drag penalty especially at very high speeds. For on-blade
886 CHAPTER 27 HELICOPTER VIBRATION REDUCTION

Mechanical
connection
between c/4 of tab
sections

Servo-tab
Main blade section

FIG. 14
Typical servo-tab.

active flaps intended for antivibration purposes, the servo-tab idea is often borrowed to lessen the
required actuation effort. Typically, on-blade flaps used for antivibration are pitched only about
5 degrees or less at the three frequencies we have discussed before, i.e., Np  1, Np, and Np + 1. When
the tab is pitched in the trailing-edge-down direction, for example, an upward load is exerted on the aft
part of the chord which thus causes a pitching movement on that portion of the blade, which in turn,
twists the blade nose-downward. The loss of lift thus causes the blade to “dive” downward. In the case
of vibratory inputs, the blade thus oscillates in flap and generates the vibratory Np hub loads that can
be part of the anti-load ellipse discussed earlier. Of course, the pitch stiffness, i.e., torsional stiffness of
the blade must be low enough to allow this sort of deformation but this is the case on most helicopters.
The new, commanded pitching of the blade at the three aforementioned frequencies is then modulated
in phase and amplitude by an active controller to reduce fuselage Np vibrations. Much of the early whirl
test and wind-tunnel work on this approach was performed Lemnios at Kaman [41,42]. Straub et al.
[43,44] performed extensive whirl and wind tunnel testing using piezo-electrically powered trailing
edge flaps as shown in Fig. 15. For the first time, flight testing of piezo-electrically controlled trailing
edge flaps was performed by Konstanzer and Enenkl [21]. The configuration, for this test, called “Blue
Pulse,” is shown in Fig. 16. The latter two researchers used mechanically amplified piezo-electric
actuators, aided by balanced servo-tabs. Lorber et al. [45] (Fig. 17) tested a nonbalanced-tab configu-
ration using higher authority, on-blade actuators powered by brushless electric motors whose rotational
motion was converted to linear motion by means of ball-screws. The nice feature of this design is that
the aerodynamic drag penalty of the servo tab was eliminated and the authority of the actuator was
much larger than that of the piezo-electric actuator thus opening the potential for primary flight control
as well as vibration control. All the testing previously cited demonstrated very good vibration reduc-
tions albeit still limited by the same controllability problem exhibited by HHC and IBC. Of course, the
same solutions to compensate for this shortcoming are equally applicable to blade morphing. One other
nice feature of a discrete on-blade flap is that the entire mechanism including flap and actuator can be
made accessible for routine maintenance. As the reader might surmise, the vibratory environment on a
helicopter blade is hostile; steady acceleration levels can be almost 1000 g and vibratory levels can
easily be 10–20 g. Clearly, any concept that requires maintenance due to failures caused by this
high-vibration environment must be made accessible for replacement and/or repair. The on-blade flap
equipment in the concept previously described, can fulfill this requirement by being placed under var-
ious sorts of access hatches designed into the blade.
9 INDIVIDUAL BLADE CONTROL 887

SMART Blade, flap, actuator cross-section

SMART rotor blade on whirl tower SMART rotor in the NFAC 40-by 80-foot
wind tunnel (looking upstream)
FIG. 15
Boeing on-blade flap whirl and wind tunnel test (Refs. [43, 44]).

Close-up of deflected flap

BK117 helicopter with Blue PulseTM active flap rotor Flap unit assembly
FIG. 16
Eurocopter (now Airbus) active flap.
From P. Konstanzer, B. Enenkl, Recent advances in Eurocopter’s passive and active vibration control, in: Proceedings of the American
Helicopter Society 64th Annual Forum, Montreal, Canada, April 29–May 1, 2008.
888 CHAPTER 27 HELICOPTER VIBRATION REDUCTION

FIG. 17
Full authority active flap with electromechanical actuator (Ref. [46]).

Previously, we discussed discrete on-blade devices but some researchers have studied and tested
more integrated morphing designs called “Active Twist Rotors” or ATRs. Cesnik [46] and Wilbur
and coauthors [47], for example, have highly integrated piezo-electric actuators directly into the spar
of the helicopter blade. This enables the blade twist to be varied via application of high-voltage to the
embedded devices. ATRs have been wind-tunnel tested and have produced good vibration reductions,
with limitations caused by the deflection capability of the piezo-electric material and again with the
limitation caused by a limited number of control of degrees of freedom. The engineer might also con-
sider that such highly integrated designs must also demand a high degree of reliability; it would be a
very expensive failure if the piezo-electric device fails after years of service as it cannot be repaired or
replaced, thus requiring the replacement of the entire blade and the failed blade to be discarded or
re-built.
The results for both trailing edge flaps and active twist rotors are fairly impressive. So why are
on-blade devices not universally adopted despite decades of development? The answer is complicated
by the extreme challenges that must be overcome to make an on-blade “machine” work reliably in such
a high-g environment but, the author considers that there is a more serious roadblock. Recall the pre-
vious discussion regarding the six independent loads or six degrees-of-freedom that must be eliminated
to guarantee virtually zero vibration. IBC, HHC and on-blade flaps are all limited to essentially, three
control degrees-of-freedom at the Np frequency. This can be visualized by imagining the commands in
any of these active control systems to be one of three types (1) collective, where all the controls produce
an in-phase or umbrella-type motion of the rotor disk, (2) pitch, where the controls produce a vibratory
pitching of the rotor disk producing Np pitching moments, and (3) roll, where the controls yield rolling
of the disk with resulting Np rolling moment. Three controls are not enough to guarantee complete
suppression consequently all the rotor control methods must be augmented by additional anti-vibe de-
vices to make up for the deficit. It is a “tough sell” to program managers if the technologist needs a lot of
money to develop such a system and, at the same time, needs additional money for a second system.
Consequently, the author’s opinion is that this need for two systems is probably the primary reason why
rotor-based methods have not been productionized.
REFERENCES 889

10 THE PATH FORWARD


In this article, we have described some of the history of helicopter anti-vibration methods along with
some that have been under development for many years and some approaches that are relatively recent.
Which approach will make it to production is still a matter of speculation but some educated predictions
can be made and some approaches can probably be eliminated. It seems certain that hydraulic HHC and
IBC will not be making a resurgence anytime soon unless for very special applications; they just require
too many inventions, i.e., better slip rings, better seals, and some ability to regenerate power expended.
Thus, by elimination, electrical systems seem likely to be the wave of the future. How about on-blade
devices vs. fuselage based devices? As we have suggested, on-blade devices suffer from an inability to
do the entire task; they can reduce three degrees-of-freedom but not the six required to truly nullify all
the vibratory loads created by the main rotor. The author considers that electrical IBC and electrical
on-blade flaps have a real potential for “making the cut” to production. This notion is bolstered by the
potential of such systems to compensate for the inevitable blade-to-blade differences that crop up in
real aircraft. Such on-rotor systems, when augmented by three or more fuselage based antivibration
actuators to produce controllability of six degrees-of-freedom seems a viable approach. Finally, the
counter-force idea of having six controllable actuators on or near the gearbox is very alluring. Such
a system is relatively simple and has solid physics-based reasons why it will get helicopter vibrations
to near-zero levels. Current antivibration actuators seem just adequate for this job but there is a chal-
lenge here also. Antivibration actuators are heavy. Can the industry invent an actuator that is more
weight efficient? Current units produce about 300 N of output load per kilogram of dead weight.
The industry should pick up the challenge to double this value. If such a device were available, it could
be used in the on-blade + fuselage systems as well as in the all-fuselage systems.

ACKNOWLEDGMENTS
The author thanks Robert Blackwell, Sikorsky Helicopter Co. (ret.), Lawrence Eastman, Sikorsky Helicopter Co.
(ret.) and Christopher Sutton of Sikorsky a Lockheed Martin Company, Brahmananda Panda of Boeing, Troy
Schank and Michael Smith of Bell Helicopter A Textron Company, Paul Cranga of Airbus Helicopters, Mark Jolly
of Lord Corporation, Dr. Uwe Arnold of ZF Luftfahrttechnik and Michael Hirschberg of the American Helicopter
Society International for their invaluable assistance in editing this article. The author also thanks their respective
companies for kindly giving permission to reproduce the information from their excellent technical publications.
All opinions, errors and omissions are solely the author’s responsibility.

REFERENCES
[1] A.C. Veca, Vibration Effects on Helicopter Reliability and Maintainability, USAAMRDL Technical Report
73-11, April 1973.
[2] Aeronautical Design Standard, Standard Practice, Requirements for Rotorcraft Vibration Specifications,
Modeling and Testing, ADS-27A-SP, May 2, 2006.
[3] International Standard, ISO-2631-1, Mechanical vibration and shock – Evaluation of human exposure to
whole-body vibration.
[4] U.S. Federal Aviation Regulation, FAR-29.251.
890 CHAPTER 27 HELICOPTER VIBRATION REDUCTION

[5] A.R.S. Bramwell, D. Balmford, G. Done, Bramwell’s Helicopter Dynamics, second ed., AIAA (American
Institute of Aeronautics & Ast); second ed., 2001.
[6] W. Johnson, Helicopter Theory, Revised ed., Dover Publications Inc., Mineola, NY, 1994.
[7] R.L. Bielawa, Rotary Wing Structural Dynamics and Aeroelasticity, second ed., American Institute of
Aeronautics & Astronautics, Reston, VA, January 1, 2006.
[8] J.J. O’Leary, in: Reduction in vibration of the CH-47C helicopter using self-tuning vibration absorbers, Paper
Proceedings of the Shock and Vibration Symposium, December 1969, pp. 191–202. (Bulletin No. 40, Part 5).
[9] R. Gabel, P. Lang, in: CH-47D airframe vibration reduction through airframe stiffening, Proceedings of the
American Helicopter Society 51st Annual Forum, Fort Worth, TX, May 9–11, 1995.
[10] D. Richard Teal, D. McCorvey, in: Active vibration suppression for the CH-47D, Mallory, Proceedings of the
American Helicopter Society 53rd Annual Forum, 1997.
[11] J. Marshall II, Tuned Spring Mass Vibration Absorber, U.S. Patent 4230291, October 28, 1980.
[12] W.F. Paul, K.C. Mard, Vibration Damped Helicopter Rotor, U.S. Patent 3540809, 1970.
[13] W.F. Paul, in: Development and evaluation of the main rotor bifilar absorber, Proceedings of the American
Helicopter Society Annual Forum, May 14–16, 1969.
[14] J. Madden, Constant Frequency Bifilar Vibration Absorber, U.S. Patent 4218187, 1980.
[15] M.A. Rangacharyulu, Bell Helicopter Textron Inc., M.J. Moore, in: Flight vibration testing of the V-22 til-
trotor aircraft, Boeing Helicopter Company. Proceedings of the 47th Annual Forum of the American Heli-
copter Society, Phoenix, AZ, May 6–8, 1991.
[16] R.B. Taylor, P.E. Zwicke, P. Gold, W. Miao, Analytical Design And Evaluation Of An Active Control
System For Helicopter Vibration Reduction And Gust Response Alleviation, July, 1980 (Prepared under
Contract No. NAS2-10121 by United Technologies Research Center East Hartford, Connecticut and Sikor-
sky Aircraft for the Ames Research Center, NASA).
[17] W.A. Welsh, in: Evolution of active vibration control technology, Proceedings of the American Helicopter
Society, 4th Decennial Conference on Aeromechanics, January 21, 2004.
[18] D.R. Halwes, in: LIVE—Liquid Inertial Vibration Eliminator, Proceedings of the American Helicopter So-
ciety 36th Annual Forum, Washington, DC, May 1980.
[19] W.G. Flannelly, in: The dynamic anti-resonant vibration absorber, Proceedings of the American Helicopter
Annual Forum, 1966.
[20] K. Riedel, in: Vibration analysis and testing of Bell-429 helicopter, Proceedings of the American Helicopter
Annual Forum, Phoenix, AZ, May 11–13, 2010.
[21] P. Konstanzer, B. Enenkl, in: Recent advances in eurocopter’s passive and active vibration control, Proceed-
ings of the American Helicopter Society 64th Annual Forum, Montreal, Canada, April 29–May 1, 2008.
[22] R.A. Desjardins, W.E. Hooper, in: Rotor isolation of the hingeless rotor BO-105 and YUH-61 helicopters,
Proceedings of the Second European Rotorcraft and Powered Lift Aircraft Forum, September 1976.
[23] R. Desjardins, W.E. Hooper, Antiresonant rotor isolation for vibration reduction, J. Am. Helicopter Soc.
25 (3) (1980) 46–55.
[24] J. Andrews, W.A. Welsh, R. Altieri, J. DiOttavio, in: Ground and flight testing of a hub mounted vibration
suppression system, Proceedings of the American Helicopter Society 70th Annual Forum, Montreal, Quebec,
Canada, May 20–22, 2014.
[25] R.K. Wernicke, J.M. Drees, in: Second harmonic control, Proceedings of the 19th Annual National Forum of
the American Helicopter Society, Washington, DC, May 1963.
[26] J. Shaw, N. Albion, E.J. Hanker, R.S. Teal, in: Higher harmonic control: wind tunnel demonstration of fully
effective vibratory hub force suppression, Proceedings of the 41st Annual Forum of the American Helicopter
Society, Fort Worth, TX, May 1985.
[27] E.R. Wood, R.W. Powers, J.H. Cline, C.E. Hammond, On developing and flight testing a higher harmonic
control system, J. Am. Helicopter Soc. 30 (1) (1985) 3–20(18).
REFERENCES 891

[28] F.K. Straub, E.V. Byrnes Jr, Application of Higher Harmonic Blade Feathering on the OH-6A Helicopter For
Vibration Reduction, NASA-CR-4031, 1986.
[29] W. Miao, S.B.R. Kottapalli, M.M. Frye, in: Flight demonstration of higher harmonic control (HHC) on S-76,
Proceedings of the 42nd Annual Forum of the American Helicopter Society, Washington, DC, June 1986.
[30] J. O’Leary, S. Kottapalli, M. Davis, in: Adaptation of a modern medium helicopter (Sikorsky S-76) to higher
harmonic control, Proceedings of the 40th Annual Forum of the American Helicopter Society, 1984.
[31] S.A. Jacklin, in: Second test of a helicopter individual blade control system in the NASA AMES 40-by
80-foot wind tunnel, Proceedings of the 2nd International Aeromechanics Specialists’ Conference, Bridge-
port, CT, October 11–13, 1995.
[32] N.D. Ham, T.R. Quackenbush, in: A simple system for helicopter individual blade control and its application
to stall induced vibration alleviation, Proceedings of the National Specialists’ Meeting on Helicopter Vibra-
tion, Sponsored by the American Helicopter Society, Northeast Region, Hartford, CT, November 2–4, 1981.
[33] S.A. Jacklin, A. Blais, D. Teves, R. Kube, in: Reduction of helicopter BVI noise, vibration and power con-
sumption through individual blade control, Proceedings of the American Helicopter 51st Annual Forum, Fort
Worth, TX, May 9–11, 1995.
[34] S.A. Jacklin, A. Haber, G. de Simone, T.R. Norman, C. Kitaplioglu, P. Shinoda, in: Full-scale wind tunnel test
of an individual blade control system for a UH-60 helicopter, Proceedings of the American Helicopter 58th
Annual Forum, Montreal, Canada, June 11–13, 2002.
[35] T.R. Norman, C. Theodore, P. Shinoda, D. Fuerst, U.T.P. Arnold, S. Makinen, P. Lorber, J. O’Neill, in: Full-
scale wind tunnel test of a UH-60 individual blade control system for performance improvement and vibra-
tion, loads, and noise control, Proceedings of the American Helicopter 65th Annual Forum, Grapevine, TX,
2009.
[36] D. Schimke, U.T.P. Arnold, R. Kube, in: Individual blade root control demonstration evaluation of recent
flight tests, Proceedings of the American Helicopter Society 54th Annual Forum, Washington, DC, May
20–22, 1998.
[37] D. F€urst, C. Keßler, T. Auspitzer, M. M€uller, A. Hausberg, H. Witte, in: Closed loop IBC-system and flight
test results on the CH-53G helicopter, Proceedings of the American Helicopter Annual Forum, Baltimore,
2004.
[38] P.P. Friedmann, On blade control of rotor vibration, noise, and performance: just around the corner? The 33rd
Nikolsky honorary lecture, J. Am. Helicopter Soc. 59 (4) (2014) 1–37.
[39] U.T.P. Arnold, D. Fuerst, T. Neuheuser, R. Bartels, in: Development of an integrated electrical swashplate-
less primary and individual blade control system, Cheeseman Award Paper Invited for Presentation at the
American Helicopter Society International 63rd Annual Forum, Virginia Beach, VA, May 1–3, 2007.
[40] D. Fuerst, A. Hausberg, T. Neuheuser, in: Experimental verification of an electro-mechanical-actuator for a
swashplateless primary and individual helicopter blade control system, Proceedings of the American Heli-
copter Society 64th Annual Forum, Montreal, Canada, April 29–May 1, 2008.
[41] A.Z. Lemnios, H.E. Howes. Wind Tunnel Investigation of the Controllable Twist Rotor Performance and
Dynamic Behavior, JSAAMRDL-TR 77-10, June 1970.
[42] A.Z. Lemnios, A.F. Smith. An Analytical Evaluation of the Controllable Twist Rotor Performance and
Dynamic Behavior, USAAMRDL-TR-72-16, 1972.
[43] F.K. Straub, et al., in: Development and whirl tower test of the SMART active flap rotor, SPIE Conference on
Smart Materials and Structures, San Diego, 2004.
[44] F.K. Straub, V.R. Anand, T.S. Birchette, B.H. Lau, in: Wind tunnel test of the SMART active flap rotor,
Proceedings of the American Helicopter Society 65th Annual Forum, Grapvine, TX, May 27–29, 2009.
[45] P.F. Lorber, B. Hein, J. Oneill, B. Isabella, J. Andrews, M. Brigley, J. Wong, P. LeMasurier, B.E. Wake, in:
Whirl and wind tunnel testing of a sikorsky active flap demonstration rotor, Proceedings of the American
Helicopter Society 67th Annual Forum, Virginia Beach, VA, May 3–5, 2011.
892 CHAPTER 27 HELICOPTER VIBRATION REDUCTION

[46] C.E.S. Cesnik, S.J. Shin, W.K. Wilkie, M.L. Wilbur, P.H. Mirick, in: Modeling, design, and testing of the
NASA/Army/MIT active twist rotor prototype blade, Proceedings of the American Helicopter Society 55th
Annual Forum, Montreal, Canada, May 25–27, 1999.
[47] D. Fogarty, M.L. Wilbur, M.K. Sekula, D.D. Boyd Jr., in: Prediction of BVI noise for an active twist rotor
using a loosely coupled CFD/CSD method and comparison to experimental data, Proceedings of the Amer-
ican Helicopter Society 68th Annual Forum, Fort Worth, Texas, May 1–3, 2012.

You might also like