You are on page 1of 13

Helicopter Rotor Health Monitoring

Using Adaptive Estimation


Jonathan Alkahe∗
Omri Rand†
Yaakov Oshman‡
Technion – Israel Institute of Technology
Department of Aerospace Engineering
Haifa 32000, Israel

Abstract

A new health and usage monitoring methodology for detection and identification of damage in
a helicopter rotor is presented. A full–scale rotor analysis in forward flight has been carried out
using a detailed model of the coupled blade–fuselage behavior. Several rotor component faults,
as well as local blade stiffness defects are considered. A set of Kalman filters is constructed,
where the calculated blade tip response, in addition to elastic modes, comprise a state vector.
In the proposed approach, each filter is based on the assumption that a particular fault has
occurred. The best fitting model, according to measurements taken from the truth model, is
determined in a probabilistic manner. In the numerical study used to demonstrate the perfor-
mance of the method, two sets of noisy measurements are generated. The first set is based on
blade tip sensors, and the second set consists of non–rotating hub loads. A Monte–Carlo anal-
ysis followed by a statistical experiment enable a comprehensive view of the statistical nature
of the results. A parametric study is presented and conclusions concerning the detectability of
damage in a helicopter rotor and the efficiency of the proposed method are drawn.

Introduction an increase in maintenance costs.


One class of damage detection methods in
The detection of damage as a part of self health which damage is seen as a change in the param-
and usage monitoring system (HUMS) in struc- eters of a structural model is based on modal
tural systems is an important contributor to their information.1–7 Typically, modal–based damage
safety, reliability and structural integrity. Early detection methods use a finite element model of
damage detection has the potential of reducing the system combined with experimental modal
life cycle costs and possibly increasing replace- data to determine damage location and extent.
ment time intervals. If damage is located and The effect of cracks on the natural frequencies
monitored, then components of the structure may of a cantilever beam is demonstrated in Refs. 3–
be replaced before some critical point is reached 4. These cracks were modeled using rotational
and a dangerous failure occurs. Particularly, the springs with equivalent stiffness. Since natural
components of a helicopter rotor are subjected to frequencies change very slightly as crack size and
high periodic loads and expected to perform un- location varies, the addition of noise, not treated
der harsh environmental conditions. These fac- in these studies, would significantly decrease the
tors, combined with the absence of redundant identification capability.
load paths, frequently resides in early replace- In Ref. 5 the changes in mode shape due to the
ment of structural components, therefore causing presence of structural damage was determined. A
∗ Graduate Student. finite element model with reductions of the mod-
† Associate Professor. ulus of elasticity in prescribed segments was im-
‡ Associate Professor.
plemented. It was shown that the elastic rotation
Presented at the American Helicopter Society 57th
Annual Forum, Washington, DC, May 9-11, 2001.
undergoes a step jump in value when crossing the
Copyright  c by the American Helicopter Society damage location, while the displacement param-
International, Inc. All rights reserved. eter takes a change in its slope.
An eigenstructure assignment technique for monitoring.
damage detection in rotating structures is
demonstrated in Ref. 6. The damage is simu- The HUMS described in Ref. 14 is composed
lated by a 10% loss of mass and stiffness in the of two major components. The first is the on–
damaged element. A test case with noise con- board system, which provides rotor trim and bal-
taminated mode shapes is also presented. An ex- ance, drive train monitoring and engine checks.
tension of this algorithm for rotating helicopter These functions are based on vibration measure-
blades accounting for hovering aerodynamics is ments. The second component is the ground sta-
presented in Ref. 7. The blade’s aerodynamics is tion, which provides diagnostic/prognostic and
incorporated as a damping term in the structural maintenance requirements. The validation of the
dynamics of the blade. The damage is shown to open system approach14 is described in Ref. 15.
be properly characterized when flapping modes Five modules from a diverse set of manufacturers
are used. However, the algorithm presented is compose the system.
based on a least–squares estimation procedure
which might be inaccurate in some cases. A comparison of aircraft usage based on the
Another approach for damage detection in original predicted flight spectrum and a HUMS
beam structures described in literature is using a measured aircraft usage is described in Ref. 16.
subspace rotation algorithm.2 This method views In this work a ground–based system is described
damage location and damage extent as two differ- and utilized. Overall, results indicate an increase
ent problems requiring two separate solutions. In in component life based on HUMS, although
this approach, damage is manifested as changes torsional fatigue load components resulted in a
in the mass, damping and stiffness matrices of the decreased life time. In addition, a small cost
structure. Strain sensors were used, therefore, benefit is also gained.
a method for extracting displacement frequency
responses from strain data was presented. This
A composite flexbeam as a part of a rotor sys-
study shows that higher–order vibration modes
tem is investigated in detail in Ref. 17. Piezoce-
are required to locate damage events. In addi-
ramic lead–zirconium–titanate (PZT) sensors are
tion, condensation methods can not be used to
placed on the top and bottom surfaces. Various
remove rotational degrees of freedom because of
damage detection methods are compared.
their coupling with translation degrees of free-
dom. In the present study, the damage detection
Several studies have been published concern- methodology is based on the multiple–model
ing damage detection in helicopters.8–11 In these adaptive estimation (MMAE) approach. This al-
works, a model of a rotor combined with a rigid gorithm, as opposed to other published methods,
fuselage is utilized to simulate typical main ro- treats process and measurement noise inherently,
tor components faults. The model results are and, therefore, is more suitable to account for
then inserted as inputs to a neural network in the modeling errors and noisy environment of a
order to complete the training stage. The net- helicopter rotor. In the MMAE method various
work’s detection capability is demonstrated us- rotor component faults and levels are considered,
ing another set of apriori calculated model results where each case is adequately represented by a
(validation data), and is tested in several cases finite element model. A Kalman filter is tuned
including noise corrupted inputs. In Ref. 12 a according to each model and the best fitting one
coupled elastic rotor–fuselage helicopter model is is determined in a probabilistic manner based on
presented and its results are compared with flight noisy measurements. Previous studies concern-
test data. The analysis is based on fuselage vibra- ing a rotating blade may be found in Ref. 18.
tion measurements. All of the faults considered Results have clearly demonstrated high damage
in this work were found to increase the 1/rev vi- detection and identification capability. In ad-
bration. The analysis was able to capture the dition, an extensive parametric study was car-
trend of vibration amplitude, however, the phas- ried out, giving insight on the influence of var-
ing of the vibration was not predicted accurately. ious parameters. Crack detection in helicopter
Moreover, a fault detection methodology is not main rotor elastic blades is discussed in Ref. 19.
described in the above study. Results indicate that detecting a local stiffness
The current status of HUMS is reviewed in defect, representing a crack, using blade tip re-
Refs. 13–17. The analysis of HUMS vibration sponse measurements, is very difficult due to the
diagnostic capabilities, described in Ref. 13 was small changes between the damaged and undam-
shown to be an effective maintenance tool. A aged models. However, a well designed statistical
wide range of gearbox, shafts, bearings and tail experiment enables high detection probabilities
rotor faults have been detected using vibration combined with low false alarm rates.
The Multiple–Model Adaptive Esti- Gaussian noise processes with covariances Q(t)
mation approach and R(t), respectively. a is a vector containing
all the uncertain dynamic parameters given in the
In various estimation problems, specifically in system model. F(a) is the system plant matrix,
damage detection cases, uncertain parameters ex- G(a) is the noise distribution matrix and C(a)
ist within the system model used for algorithm is the measurement matrix. Since a may assume
design. Typically, these parameters can undergo a continuous range of values over the space of al-
large jump changes. Such problems give rise to lowable parameters, it is necessary to discretize a
the need for the estimation of parameter values into a set of J vector values: a1 , a2 , ..., aJ . A mul-
simultaneously with estimation of state variables. tiple model adaptive estimator consists of J in-
One means of accomplishing this is the multi- dependent Kalman filters, in which the jth filter
ple model adaptive estimation (MMAE) tech- is constructed according to a specific parameter
nique.20, 21 The system is assumed to be ade- set aj . These J filters form a bank of elemen-
quately represented by a linear stochastic state tal filters which are processed in parallel. Each
model, with uncertain parameters affecting the elemental filter produces its own estimate of the
matrices defining the structure of the model or true state, denoted as x̂j (ti ), for the jth hypothe-
the noise distribution model. It is further as- sized value of a. The residuals of all J elemental
sumed that the parameters can take only dis- filters are then used to calculate the probabil-
crete values. In cases where continuous param- ity that a assumes the value aj at time ti , for
eter values are presented, representative discrete j = 1, 2, ..., J. This probability is called the “hy-
values have to be chosen throughout the contin- pothesis conditional probability” and is denoted
uous range of possible values. A Kalman filter is as pj (ti ). This conditional probability represents
then designed for each choice of parameter value, the validity of the jth filter’s system model at
resulting in a bank of K separate filters. Based time ti . The hypothesis conditional probabilities
on the residuals of each one of these K filters, the pj (ti ), j = 1, 2, ..., J, are calculated at each sample
conditional probabilities of each discrete parame- time ti , by the recursive equation:
ter value being “correct” (given the measurement
history up to that time) are evaluated iteratively. fz(t |a,Z(ti−1 )) (zi |aj , Zi−1 ) pj (ti−1 )
pj (ti ) = PJ i
This procedure is summarized in Fig. 1. k=1 fz(ti |a,Z(ti−1 )) (zi |aj , Zi−1 ) pk (ti−1 )
(3)
where Z(ti−1 ) is the measurement history from
^
the first sample time until sample time ti−1 ,
x
Filter of
r1
1
and the innovation probability density function
Fault 1
is given by
^2
x
Noisy
Filter of
r2 Model 1
Measurements
Fault 2
Probabilities Probabilities
fz(ti |a,Z(ti−1 )) (zi |aj , Zi−1 ) = S/2 1/2
Computation (2π)
|Ak (ti )|
š ›
1 T
^J ·exp − rk (ti ) A−1
k (ti ) rk (ti ) (4)
Filter of x 2
Undamaged rJ
Case
where S is the number of sensors. The kth filter
residual vector is:
Figure 1: The MMAE procedure
rk(ti ) = z(ti )−Hk (ti )x̂k (t−
i ) (5)

where x̂k (t−


i ) is the kth filter predicted state es-
Following the development presented in
timate. The kth filter–computed residual covari-
Refs. 21,22, consider the system model described
ance matrix, Ak (ti ) is calculated by
by the first–order, linear, stochastic differential
state equation of the form:
Ak (ti ) = Hk (ti )Pk (t− T
i )Hk (ti )+Rk (ti ) (6)
ẋ(t) = F(a)x(t)+G(a)w(t) (1)
where Pk (ti− ) is the kth filter prediction er-
with noisy measurements described by ror covariance. The residual of the jth filter
plays a major role in determining pj (ti ). As
z(ti ) = C(a)x(ti )+v(ti ) (2) is evident from (3), the filter with the smallest
value of rTj (ti )A−1
j (ti )rj (ti ) assumes the largest
where x(t) is the system state vector and z(t) conditional hypothesis probability. Thus, this
is the measurement vector. It is assumed that algorithm is consistent with the intuition that
w(t) and v(t) are independent, zero–mean, white the residuals of a well–matched filter should be
smaller (relative to the filter’s internally com- probability of the jth model. Then
puted residual covariance, Aj ) than the residuals P
pj (t)
of a mismatched filter. To allow the estimator to qj = PJ t∈tP
d
(7)
adapt to the changing parameter value, the hy- k=1 t∈td pk (t)
pothesis conditional probabilities are artificially
bounded below by a small number (0.0005). This Let qmax be defined as
ensures preventing any of them from converging qmax , max {qj } (8)
to zero, which would make it very difficult for j∈{1,2,...,J}
them to change significantly in response to a sub-
The model associated with qmax is said to de-
sequent change in true parameter value.
scribe the true damaged behavior in the best
manner.
In order to provide a comprehensive view of
Structural Model
the statistical nature of the results, a Monte–
Carlo analysis24 is carried out, where the same
The full–scale rotor analysis has been carried out
damaged case is repeated. For each one of the
using the software package RAPID/Plus,23 which
tested cases, the Monte–Carlo procedure contin-
is capable of modeling general rotorcraft config-
ues until no change occurs in the model found to
urations, conventional helicopters and tilt-rotors.
be the most fitting, and its identification prob-
RAPID/Plus may handle nonuniform and dissim-
ability, or a limit of 150 runs is reached. This
ilar blades and is therefore suitable to the cur-
procedure results in estimated probability val-
rent task. Both rigid and elastic blade analyses
ues for the single–run process. When the true
are possible. Blade elasticity is modeled using
case is the case of no damage, the single–run
a built-in modal based analysis for structurally
false alarm rate, designated by pFA , is calculated.
pretwisted spars. This analysis enables including
Single–shot detection and identification proba-
the blade’s axial, lead–lag, flap and twist elas-
bilities result for each one of the cases where a
tic deformations, designated by u, v, w and φ,
fault occurs. These probabilities are designated
respectively. The model also includes fully artic-
by pD and pI respectively.
ulated blades with arbitrary pitch, flap and lag
offsets, root springs and dampers, and a detailed
control system mechanism (swashplate, elastic Statistical Experiment
pitch links, pitch horn, etc.), which enables a The overall false alarm and detection probabili-
study of faults in these components. In this work ties can be controlled using a well designed sta-
it is assumed that only one fault occurs. For tistical experiment. Let H0 and H1 denote the
the case of rigid articulated blades, the follow- hypotheses of “no damage” and “damage”, re-
ing component faults are investigated: pitch–link spectively. Assume that N is the total number
damage, lag damper defect, friction in pitch bear- of single–run repetitions and n is the number of
ing and moisture absorption. The case of a local runs in which damage was detected. A decision
damage in hingeless elastic blades is also consid- criterion is defined as follows
ered.
H1
The damage detection algorithm consists of 5
n T nD (9)
different models running in parallel – 4 damaged
H0
rotor models along with the baseline undamaged
one. where nD is a predetermined threshold. When
the true undamaged case is not detected, a false
alarm results (Type I error). The false alarm
Damage Identification Logic probability is calculated using the following ex-
pression
The fault detection and identification (FDI) pro-
cess is carried out during 5 rotor revolutions (ap- XN ’ “
N i
proximately 1 sec) in which all the 5 models run PFA , P (H1 |H0 ) = pFA (1−pFA )N −i
i
in parallel. Detection based on a 5–rev. time in- i=n D

terval, designated as tp , is termed, in this study, (10)


single–run detection process. The decision logic When true damage is not detected, a missed de-
is based on calculating the fitness probability for tection (or type II error) occurs. This probability
each of the models over an inspection time in- is calculated as follows
terval (designated as td ), which includes only the PMD = 1 − PD , P (H0 |H1 ) =
last revolution. Let pj (ti ) denote the hypothe- D −1 ’
nX “
sis conditional probability of the jth model at a N
(1 − pD )i pN
D
−i
(11)
discrete time ti (3), and qj refer to the fitness i=1
i
A detector operating characteristics (DOC) plot, aged cases under consideration, an example of
showing PD vs PFA can be constructed. Using measurement differences is presented in Fig. 2.
this plot, the total number of runs and the thresh- The tip flap displacement difference is shown in
old level can be tuned to achieve an overall low Fig. 2(a), and Fig. 2(b) shows the lead–lag dis-
false alarm rate combined with a high detection placement differences. The differences are calcu-
probability. lated for all damaged models with respect to the
baseline undamaged tip response. Clearly, pitch–
Rotor component FDI results link damage results in a relatively large response
difference and is, therefore, expected to be eas-
Tip response measurements ily detectable. However, the lag damper damage
causes hardly any difference in the observables,
Consider a typical full–scale articulated 2 bladed suggesting detection difficulties.
fixed shaft rotor with rigid blades, whose prop- Figs. 3–7 show the single–run probability con-
erties are listed in Table 1. Forward flight con- tours, estimated using a Monte–Carlo analysis
ditions (µ = 0.3) were simulated including trim consisting of a maximum of 150 runs. Figs. 3(a),
commands. 3(b) present the detection and identification
probability contours for a pitch link fault, respec-
Table 1: Rotor properties. tively. The relatively high probability values in-
dicate that this fault is easily detected and iden-
Radius 7.1 m tified. Moreover, a low noise level combined with
Chord 0.76 m a high damage level produce better detection and
Blade twist angle −9 deg identification results, as can be expected from
Baseline pitch link stiffness 106 N/m beforehand. An interesting phenomenon is re-
Baseline lead–lag damper 3000 N·m·s vealed where the detection probability decreases
Baseline pitch damper 10 N·m·s as the noise level increases, until a certain point
Mass per unit length 20 Kg/m where the former starts to increase. The reason
Angular velocity 34 rad/s is that for high noise levels, the differences be-
tween the models are masked by the noise, ren-
dering similar probabilities (around 0.2) to all the
models. Therefore, theoretically, the detection
A parametric study was carried out, investigat-
probability, consists of the probability sum of all
ing the influence of various damage levels, for the
4 damaged models, tends to rise to a value of
faults mentioned in the previous section, com-
approximately 0.8. This phenomenon is demon-
bined with several noise levels. As mentioned
strated in Figs. 4(a), 5(a) and Fig. 6(a) as well.
above, only one fault may occur in one of the
Figs. 4(a), 4(b) shows the probability contours
blades. Five damage levels were examined (ex-
for the lag damper damage. Clearly, this dam-
cept for the lag damper fault which has only 3
age case is less detectable. As seen from the fig-
damage levels), along with five noise levels. The
ures, the detection and identification probabili-
noise levels presented in the following figures are
ties are not sensitive to the damage level. This is
dimensional, where displacement noise is mea-
because relatively small lag damping values ex-
sured in meters and pitch angle noise is measured
ist even in the undamaged case. Figs. 5(a), 5(b)
in radians. The measurements taken are tip re-
shows the probability contours for the pitch fric-
sponse of the two blades (tip flap and lead–lag
tion case. As shown, this fault is also relatively
displacements along with the tip pitch angle).
easy to detect and identify. Figs. 6(a), 6(b) shows
Pitch link damage is manifested as a stiffness
the probability contours for the moisture absorp-
reduction caused by a crack. The damage spec-
tion case. In this case, the probability range
trum includes five levels which vary from 10%
is well covered. Fig. 7 presents the false alarm
stiffness reduction to 30%. Lag damper damage
rate contours. Since all 5 models are included, at
is modeled by reducing the damping coefficient by
each damage level all 4 faulty models assume the
50%, 60% and 68% (lead–lag stability margin).
corresponding damage value. For example dam-
Pitch bearing friction is modeled by an increase
age level 1 means 10% pitch link stiffness reduc-
in the pitch damping coefficient at the blade root.
tion, 50% lag damping reduction, pitch friction
Five levels are considered: an increase by a fac-
increase of a factor of 4 and 2% mass increase
tor of 4 up to a factor of 8. Moisture absorption
due to moisture absorption. For high noise lev-
(for the case of a composite blade) is manifested
els the false alarm approaches 0.8, as theoretical
by an increase in the blade mass per unit length,
analysis predicts.
from 2% mass increase up to 4% (five levels).
Since the performance of any FDI algorithm A statistical experiment composed of several
is based on differences between the various dam- repetitions of a single–run process enables some
0.15 0.01
Pitch link Pitch link
Lag damper Lag damper
Pitch friction 0 Pitch friction
0.1
Moisture Moisture
−0.01
0.05
−0.02

0
∆ w [m]

∆ v [m]
−0.03

−0.05 −0.04

−0.05
−0.1
−0.06
−0.15
−0.07

−0.2 −0.08
0 45 90 135 180 225 270 315 360 0 45 90 135 180 225 270 315 360
ψ ψ

(a) Tip flap differences ∆w (b) Tip lead–lag differences ∆v

Figure 2: Blade tip response differences

96 0.5
0. 7 0.6
0.98
0.7
0.9
0.9
9 0.8
−3 −3
0.9 2
0.93
0.9 4
5

10 10 0.9
0.9

0.91
Noise Level [m]

Noise Level [m]

−4 −4
10 10

10 15 20 25 30 10 15 20 25 30
Pitch−Link Stiffness Reduction [%] Pitch−Link Stiffness Reduction [%]

(a) Detection results (b) Identification results

Figure 3: Pitch link damage results


0.8
0.7
0.6

−3 0.5 −3
0.4

0.3

10 10
0.1
Noise Level [m]

Noise Level [m]

0.2

0.3
0.5 0.4
0.6 0.5
0.7 0.6
0.7
−4 0.8 −4
10 10 0.8
0.9
0.9

50 55 60 65 50 55 60 65
Lag Damping Reduction [%] Lag Damping Reduction [%]

(a) Detection results (b) Identification results

Figure 4: Lag damper damage results


0.94

2
−3 −3 0.9

0.86

0.88

0.92

2
4
0. 6
88
0.9
10 10

9
0.8
0.8
0.8

0.
0.94 0.94
0.96 0.96
Noise Level [m]

Noise Level [m]


0.98 0.98

−4 −4
10 10

4 5 6 7 8 4 5 6 7 8
Pitch Friction Increase Factor Pitch Friction Increase Factor

(a) Detection results (b) Identification results

Figure 5: Pitch friction results


0.1
0.885 0.2
0. 0.75 0.3
0.4
0.7
0.5
0.65

−3 −3
10 10 0.6
0.7
0.75 0.7
Noise Level [m]

Noise Level [m]

0.8
0.8
0.85
0.9
0.9
0.95

−4 −4
10 10

2 2.5 3 3.5 4 2 2.5 3 3.5 4


Moisture Absorption [%mass] Moisture Absorption [%mass]

(a) Detection results (b) Identification results

Figure 6: Moisture absorption results

control on the overall false alarm and detection


0.7
0.6
probabilities. The results of such an experiment
0.5 depend on pFA and pD , as discussed in the pre-
−3 0.4
vious section. As an example, the case of lag
10
0.3 damper damage (50% lag damping reduction) is
considered. In order to achieve reasonable single–
Noise Level [m]

run false alarm and detection probabilities, a


0.2 very low noise level is permitted, as previously
demonstrated. Fig. 8 shows two examples: In
0.1 the first case, shown in Fig. 8(a), a low noise
−4
10 level (σ = 10−4 m) is considered, and the single–
run probabilities, pFA and pD , assume the values
0.1 and 0.8 respectively. Clearly, the number of
repetitions (N) needed for a low false alarm and
1
Pitch link 10%
Lag damper 50% 2 3 4 5 a high detection rates is not large (N ≈ 5). For
Pitch friction *4
Moisture 2%
Damage Intensity a higher noise level, where pFA and pD assume the
unacceptable values 0.4 and 0.5 respectively, as
Figure 7: False alarm rate demonstrated in Fig. 8(b), a much larger number
1 1
nD
3 89
0.9 0.9 nD

0.8 nD=2 0.8 46

0.7
0.7 24
0.6
0.6

D
PD

0.5

P
0.5 nD=10
0.4
0.4
0.3 N=20
0.3 N=2 N=50
0.2
pFA=0.1 N=3 pFA=0.4 N=100
0.2 N=5 0.1 N=200
p =0.8 pD =0.5
D N=10 N=400
0.1 0
0 0.1 0.2 0.3 0.4
P0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5
PFA
0.6 0.7 0.8 0.9 1
FA

(a) Experiment results for a noise level of σ = 10−4 m (b) Experiment results for a noise level of σ = 10−3 m

Figure 8: Detector operating characteristics examples

of runs is needed (N ≈ 400). associated with the highest probability is indeed


Let the required overall false alarm and detec- the correct one. Moreover, during the decision
tion rates assume acceptable representative val- time interval td , the probability results are quite
ues of 0.05 and 0.95, respectively. Using the decisive, suggesting that a Monte–Carlo analysis
above statistical experiment, the number of rep- is not needed. The high probability values (≈ 1)
etitions (N) can also be presented as contours, indicates that the statistical experiment is also
for each fault under consideration. The contour unnecessary.
lines in Fig. 9 constitute the smallest value of N
required to meet those rates.
Detection of a blade local
stiffness defect based on
Non–rotating hub load measure-
hub load measurements
ments
Consider the same rotor as in the previous sec- Consider a hingeless fixed shaft rotor with elas-
tion. In this case, only one damage level is con- tic blades, with the properties listed in Table 1.
sidered. A 5% stiffness reduction for the pitch The blade beamwise and chordwise stiffnesses
link damage, 68% lag damping reduction, pitch and torsional rigidity are: EIb = 7.6 · 104 N m2 ,
friction which results in a pitch damping increase EIc = 12·105 N m2 , GJ = 8.6·104 N m2 , respec-
of factor 1.5 and 0.5% mass increase for the mois- tively. The blade local damage, discussed in this
ture absorption case. Note that all these damage section, is a crack at a specific location along
levels are lower that the levels presented in the the span. This crack is simulated by reducing
previous section. Moreover, a much higher mea- the stiffnesses (bending stiffnesses and torsional
surement noise level is induced, with standard rigidity) at a particular finite element in the blade
deviations of 1 [KN] and 1 [KN-m] for the force structural model, as illustrated in Fig. 12.
and moment measurements, respectively.
Two load differences are shown in Fig. 10 as an
example. Fig. 10(a) demonstrates the normal-
ized hub force differences in the lateral direc- Deformed
tion (∆Fy ), and Fig. 10(b) shows the normalized φ
pitching moment differences (∆My ). The mois-
ture absorption case clearly results in larger dif-
w
ferences, indicating high detectability. The other
damage cases produce load differences bounded Und
efor
med
by ±1σ of the noise.
u v
The single–run identification probability time Damaged section
histories, for the various damage cases, are shown
in Fig. 11. These results indicate that the model Figure 12: A damaged elastic blade
50
400

20
−3 −3
10 10 10
Noise Level [m]

Noise Level [m]


5

3
200

100
50
−4 −4 20
10 10

P =0.05 PFA=0.05
FA
P =0.95 PD =0.95
D
10 15 20 25 30 50 55 60 65
Pitch−Link Stiffness Reduction [%] Lag Damping Reduction [%]

(a) Pitch link fault (b) Lag damper fault

40
20 0
0
20 10
0
−3 −3 50
10 10 10
20
Noise Level [m]

Noise Level [m]

3 10

5
3
−4 −4
10 10

PFA=0.05 PFA=0.05
PD =0.95 PD =0.95

4 5 6 7 8 2 2.5 3 3.5 4
Pitch Friction Increase Factor Moisture Absorption [%mass]

(c) Pitch friction fault (d) Moisture absorption fault

Figure 9: The required number of runs (N)

0.08 0.2
Pitch link Pitch link
3σ Lag damper Lag damper
0.06 Pitch friction 0.15 Pitch friction
Moisture Moisture

0.04 0.1

σ
0.02 0.05 σ
∆ My/Q
∆ Fy/W

0 0

−0.02 −0.05
−σ
−σ
−0.04 −0.1
−3σ
−0.06 −0.15
−3σ
−0.08 −0.2
0 90 180 270 360 0 90 180 270 360
ψ ψ

(a) Normalized lateral force differences ∆Fy /W (b) Normalized pitching moment differences ∆My /Q

Figure 10: Non–rotating hub load differences


1 1
No damage No damage
Pitch link Pitch link
0.9 Lag damper 0.9 Lag damper
td td
Pitch friction Pitch friction
0.8 Moisture 0.8 Moisture

0.7 0.7
Probability

Probability
0.6 0.6

0.5 0.5

0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1

0 0
0 1 2 3 4 5 0 1 2 3 4 5
Revs. Revs.

(a) No damage case (b) Pitch link fault

1 1
No damage No damage
Pitch link Pitch link
0.9 Lag damper 0.9 Lag damper
td td
Pitch friction Pitch friction
0.8 Moisture 0.8 Moisture

0.7 0.7
Probability

Probability

0.6 0.6

0.5 0.5

0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1

0 0
0 1 2 3 4 5 0 1 2 3 4 5
Revs. Revs.

(c) Lag damper fault (d) Pitch friction fault

1
No damage
Pitch link
0.9 Lag damper
td
Pitch friction
0.8 Moisture

0.7
Probability

0.6

0.5

0.4

0.3

0.2

0.1

0
0 1 2 3 4 5
Revs.

(e) Moisture absorption fault

Figure 11: Single–run identification probability time history


0.02
x/R=0.15
In this study, the damage intensity is modeled x/R=0.35
x/R=0.55
by a 5% stiffness reduction. Hub loads are mea- 0.015
x/R=0.75
sured with standard deviations of 1 [KN] and 1
[KN-m] for the force and moment measurements, 0.01

respectively. Four filters are constructed, based


0.005
on 4 models in which the damage location x, mea-

∆ Fy/W
sured from the hub, assumes 4 discrete normal-
0
ized values: x/R = 0.15, 0.35, 0.55, 0.75. The 5th
filter simulates the undamaged case. The true
−0.005
damage, also simulated as a 5% stiffnesses reduc-
tion, is supposed to occur at a discrete location −0.01
x/R = 0.1, 0.2, 0.3 . . . 0.9. (Note that the true lo-
cation is never identical to one of the modeled −0.015
0 90 180 270 360
locations). ψ
Fig. 13 gives an example of the damaged case load
differences with respect to the undamaged loads. (a) Normalized lateral force differences ∆Fy /W
As can be expected from beforehand, the damage
located near the root results in the largest differ- 0.08
x/R=0.15
ence. As the damage location approaches the tip, x/R=0.35
the differences diminish. 0.06 x/R=0.55
x/R=0.75
Table 2 summarizes the single–run results,
0.04
which indicate good identification capability. In
most cases, the model in which the damage loca- 0.02
tion is closest to the true location, indeed receives
∆ My/Q

the highest fitness probability. The case where 0


the true damage occurs near the tip (x/R = 0.9,
−0.02
marked in red) results in a missed–detection. As
can be expected, this case constitutes a very dif- −0.04
ficult problem and requires special treatment.
−0.06

Table 2: Identification results. −0.08


0 90 180 270 360
ψ
True damage Identified Identification
location location probability (b) Normalized pitching moment differences ∆My /Q
No damage No damage 0.99
0.1 0.15 0.99 Figure 13: Non–rotating hub load differences
0.2 0.15 0.99
0.3 0.35 0.91
0.4 0.35 0.89
0.5 0.55 0.63 to simulate various rotor faults. The proposed
0.6 0.55 0.99 method also enables sensors of various types to
0.7 0.75 0.99 be implemented.
0.8 0.75 0.97
0.9 No Damage 0.90
The algorithm presented constitutes a new ap-
proach towards damage detection in mechanical
systems. Contrary to other model–based dam-
age detection methods in helicopter rotors, such
Conclusions as methods based on neural networks, this ap-
proach requires no training stage. Moreover, this
A model–based damage detection algorithm for algorithm treats measurement and process noise
a helicopter rotor, using an adaptive estimation inherently, accounting for the noisy rotor envi-
technique, incorporating noisy measurements, is ronment and modeling uncertainties. Combined
presented. The coupled blade-fuselage dynamic with the model–based feature, the proposed al-
behavior is introduced using the RAPID/Plus gorithm eliminates the need for a training stage
package where finite element model of the blade and enables a wide range of flight regimes. The
is utilized. A set of Kalman filters is constructed damage detection capability is tested in various
cases. In general, for low noise levels, this ap- Journal of Sound and Vibration, Vol. 183,
proach produces excellent results. The illustra- No. 5, 1995, pp. 857–871.
tive rotor faults considered in this work demon-
5
strate several levels of detectability. Pitch link Yuen, M. M. F., “A Numerical Study of the
damage and pitch bearing friction were found to Eigenparameters of a Damaged Cantilever,”
be highly detectable, for both sets of measure- Journal of Sound and Vibration, Vol. 103,
ments considered. This emerges from the rela- No. 3, 1985, pp. 301–310.
tively high response sensitivity to pitch changes. 6
On the contrary, lag damper damage was hard Kiddy, J. and Pines, D., “Eigenstructure As-
to detect using blade tip response measurements, signment Technique for Damage Detection in
since small damping values are present. Lag Rotating Structure,” AIAA Journal , Vol. 36,
damping changes, in this case, have a small effect No. 9, September 1998, pp. 1680–1685.
on the overall rotor response. The detectability of 7
Kiddy, J. and Pines, D., “The Effects of Aero-
this type of damage was shown to increase signifi-
dynamic Damping on Damage Detection in
cantly using hub load measurements. Blade mois-
Helicopter Main Rotor Blades,” 54th Annual
ture absorption detectability is also quite high. In
Forum of the American Helicopter Society,
cases of poor single–run performance, the statis-
Montreal, Canada, May 25–27 1999.
tical experiment suggested, enables sustaining a
low false alarm rate combined with a high detec- 8
Ganguli, R. and Chopra, I., “Simulation of
tion capability. The case of a blade local damage Helicopter Rotor-System Structural Damage,
was also investigated. It was shown that, using Blade Mistracking, Friction and Freeplay,”
noisy hub load measurements, this approach is Journal of the American Helicopter Society,
capable of detecting, and, in most cases, correctly Vol. 35, No. 4, July 1998, pp. 591–597.
locating small stiffness changes along the span.
The significant advantage of the proposed ap- 9
Stevens, P. W., Hall, D. L., and Smith, E. C.,
proach arises from the filtering process enabling “A Multidisciplinary Research Approach to
probabilistic determination based on relatively Rotorcraft Health and Usage Monitoring,”
little information. Introducing an external ex- 52nd Annual Forum of the American Heli-
citation in this case would probably increase copter Society, Washington, D.C., June 4–6
damage detectability since the transient sys- 1996.
tem response will provide additional information.
Moreover, other sensor types like fuselage vibra- 10
Hass, D. J. and Schaefer, C. G., “Emerging
tion measurements (in trimmed flight) are ex- Technologies for Rotor System Health Moni-
pected to enhance detection capability. These toring,” 52nd Annual Forum of the American
topics are currently under investigation. Helicopter Society, Washington, D.C., June
4–6 1996.
References 11
Ganguli, R., Chopra, I., and Hass, D. J., “A
1
Linder, D. K. and Kirby, G., “Location and Phisics Based Model for Rotor System Health
Estimation of Damage in a Beam Using Iden- Monitoring,” 22nd European Rotorcraft Fo-
tification Algorithms,” AI AA/AS M E Adap- rum, Brighton, UK, September 1996.
tive Structures Forum, Washington, D.C., 12
Yang, M., Chopra, I., and Hass, D. J., “Rotor
April 21–22 1994.
Sysytem Health Monitoring Using Coupled
2
Kahl, K. and Sirkis, J. S., “Damage Detection Rotor–Fuselage Vibration Analysis,” 56th An-
in Beam Structures Using Subspace Rotation nual Forum of the American Helicopter Soci-
Algorithm with Strain Data,” AIAA Journal , ety, Virginia beach, Virginia, May 2–4 2000.
Vol. 34, No. 12, December 1996, pp. 2609– 13
2614. Larden, B. D., “An Analysis of HUMS Vi-
bration Diagnostic Capabilities,” 53rd An-
3
Ostachowicz, W. M. and Krawczuk, M., nual Forum of the American Helicopter So-
“Analysis of the Effect of Cracks on the Nat- ciety, Virginia beach, Virginia, April 29–May
ural Frequencies of a Cantilever Beam,” Jour- 1 1997.
nal of Sound and Vibration, Vol. 150, No. 2,
14
1991, pp. 191–201. Hess, R., Chaffee, M., Page, R., and Rose-
berry, K., “Realizing an Expandable Open
4
Sundermeyer, J. N. and Weaver, R. L., “On HUMS Architecture,” 56th Annual Forum
Crack Identification and Characterization In of the American Helicopter Society, Virginia
a Beam by Non–Linear Vibration Analysis,” beach, Virginia, May 2–4 2000.
15
Hass, D. J., Baker, T., Sspracklen, D., and
Schaefer, C., “Joint Advanced Health and
Usage Monitoring System (JAHUMS) Ad-
vanced Concept Technology Demonstration
(ACTD),” 56th Annual Forum of the Amer-
ican Helicopter Society, Virginia beach, Vir-
ginia, May 2–4 2000.
16
Robeson, E., “MH–47E Structural Usage
Monitoring System (SUMS) Fleet Demon-
stration Results,” 56th Annual Forum of the
American Helicopter Society, Virginia beach,
Virginia, May 2–4 2000.
17
Ghoshal, A., Harrison, J., Sundaresan, M. J.,
Schulz, M. J., and Pai, P. F., “Toward Devel-
opement of an Intelligent Rotor System,” 56th
Annual Forum of the American Helicopter So-
ciety, Virginia beach, Virginia, May 2–4 2000.
18
Alkahe, J., Rand, O., and Oshman, Y., “Dam-
age Detection in a Rotating Helicopter Blade
Using an Adaptive Estimator,” AI AA Guid-
ance, Navigation and Control Conference,
Denver, CO, August 14–17 2000.
19
Alkahe, J., Rand, O., and Oshman, Y.,
“Damage Detection in a Helicopter Rotor by
Adaptive Estimation,” 26th European Rotor-
craft Forum, The Hague, The Netherlands,
September 26–29 2000.
20
Magill, D. T., “Optimal Adaptive Estima-
tion of Sampled Stochastic Processes,” IEEE
Transactions on Automatic Control , Vol. AC–
10, No. 4, October 1965, pp. 434–439.
21
Maybeck, P. S., Stochastic Models, Estima-
tion, and Control , Vol. 2, Academic Press,
NY, 1982.
22
Stepaniac, M. J. and Maybeck, P. S.,
“MMAE–based Control Redistribution Ap-
plied to the VISTA F–16,” IEEE Transactions
on Aerospace and Electronic Systems, Vol. 34,
No. 4, October 1998, pp. 1249–1259.
23
Rand, O., “RAPID/Plus: Rotorcraft Analy-
sis for Preliminary Design + Aeroelasticity:
User Manual,” Tech. rep., Haifa, Israel, Jan-
uary 1999.
24
Freund, J. E., Modern Elementary Statistics,
Prentice–Hall, New Jersey, 6th ed., 1984.

You might also like