Thesis T Ooijevaar PDF

Download as pdf or txt
Download as pdf or txt
You are on page 1of 194

Vibration Based Structural

Health Monitoring of Composite


Skin-Stiffener Structures

Ted Ooijevaar
VIBRATION BASED STRUCTURAL HEALTH
MONITORING OF COMPOSITE SKIN-STIFFENER
STRUCTURES

Ted Ooijevaar
De promotiecommissie is als volgt samengesteld:

Voorzitter en secretaris:
prof.dr. G.P.M.R. Dewulf Universiteit Twente

Promotor:
prof.dr.ir. R. Akkerman Universiteit Twente
prof.dr.ir. T. Tinga Universiteit Twente

Leden (in alfabetische volgorde):


prof.dr.-ing. C. Boller Universität des Saarlandes
prof.dr.ir. F.J.A.M. van Houten Universiteit Twente
prof.dr.ir. P.M. Lugt Universiteit Twente
prof.dr. V. Michaud Ecole Polytechnique Fédérale de Lausanne

This research project was financially supported by the Eco-Design ITD within the
Clean Sky framework (grant agreement number CSJU-GAM-ED-2008-001).

Vibration based structural health monitoring of composite skin-stiffener structures


Ooijevaar, Theodorus Hendricus
PhD thesis, University of Twente, Enschede, The Netherlands
March 2014

ISBN 978-90-365-3624-0
DOI 10.3990/1.9789036536240
Copyright © 2014 by T.H. Ooijevaar, Enschede, The Netherlands
Printed by Ipskamp Drukkers B.V., Enschede, The Netherlands

Cover: photograph of the composite skin-stiffener structures used in the present research.
The curved lines represent the steady state velocity responses measured at the damaged area
of a pristine (upper curve) and impact damaged (lower curve) skin-stiffener structure.
VIBRATION BASED STRUCTURAL HEALTH
MONITORING OF COMPOSITE SKIN-STIFFENER
STRUCTURES

PROEFSCHRIFT

ter verkrijging van


de graad van doctor aan de Universiteit Twente,
op gezag van de rector magnificus,
prof.dr. H. Brinksma,
volgens besluit van het College voor Promoties
in het openbaar te verdedigen
op vrijdag 7 maart 2014 om 14:45 uur

door

Theodorus Hendricus Ooijevaar

geboren op 20 januari 1983

te Wognum
Dit proefschrift is goedgekeurd door de promotoren:

prof.dr.ir. R. Akkerman
prof.dr.ir. T. Tinga
Summary

Composite materials combine a high strength and stiffness with a relatively


low density. These materials can, however, exhibit complex types of damage,
like transverse cracks and delaminations. These damage scenarios can severely
influence the structural performance of a component. Periodic inspections
are required to ensure the integrity of a component during its life. The
current inspection methods are often time-consuming, costly and require the
components to be readily accessible. Vibration based structural health monitoring
(SHM) technologies propose a promising alternative and involve the continuous
monitoring of a structure by employing an integrated sensor system. These
methods are based on the concept that the dynamic behavior of a structure can
change if damage occurs.
Although many damage identification methods have been proposed in the
literature, there are still numerous difficulties in the practical application of these
approaches, especially to complex structures. The performance of a vibration
based damage identification approach is highly dependent on the actual design
of the structure and the damage scenario that is considered. This thesis focuses
on the identification of damage in advanced composite skin-stiffener structures.
The principle objective is to develop guidelines for the detection, localization and
characterization of damage in composite skin-stiffener structures based on changes
in the dynamic behavior.
A literature study supported by an analytical model showed that mode shape
curvatures combined with the modal strain energy damage index (MSE-DI)
algorithm are a potentially powerful damage feature and classifier for the
identification of damage in several advanced composite skin-stiffener structures. A
experimental set-up, including a shaker and laser-vibrometer, was used to measure
the dynamic responses. A linear dynamic system description is obtained by
applying experimental modal analysis. The vibration experiments demonstrated
the feasibility of the MSE-DI algorithm to detect, localize and roughly estimate the
size of barely visible impact damage (BVID) in advanced composite skin-stiffener
structures. It is concluded that the method is particularly effective for health
monitoring of skin-stiffener connections. The method remained inconclusive in the
case of pure skin related damage.
Experiments showed that damage at the skin-stiffener interface can introduce clear
nonlinear effects in the dynamic behavior of the structure. These nonlinear effects

i
ii

are attributed to the interaction between the skin and stiffener that occurs during
opening and closing motion of the damage. It is shown that linear damage
identification methods (e.g. modal domain methods) are feasible for low excitation
amplitudes, but the presence of nonlinear dynamic effects cannot remain silent for
higher amplitudes. The nonlinear dynamic effects can act as strong indicator of
damage, but can also be useful for characterization purposes.
The nonlinear dynamic effects introduced by the skin-stiffener damage urges
the development of nonlinear damage identification methods. A study on the
understanding and feasibility of using nonlinear vibro-acoustic modulations for
the detection, localization and characterization of impact damage in a composite T-
beam is presented. A time domain analysis at multiple spatial locations is used
to detect and localize impact damage in a skin-stiffener connection, based on
locally increased amplitude modulation effects. Analysis of the characteristics of
the nonlinear modulations opens the ability to characterize the nonlinear dynamic
behavior introduced by the damage at the skin-stiffener interface.
The work presented in this thesis showed that the relations between the
characteristics of the structure, the potential damage scenarios and the damage
identification method together define the performance of the vibration based
damage identification strategy. Therefore, it is concluded that the design of
a vibration based damage identification strategy is made-to-measure work and
requires a thorough physical understanding of the potential failure mechanisms,
the critical damage locations and their effect on the dynamic behavior. To aid in
this process, a scenario based procedure for the design of a damage identification
strategy is proposed. All findings presented in this thesis contribute to the
development of a design tool for research engineers, to assist the implementation
of structural health monitoring technology in safety-critical composite structures.
Samenvatting

Composiet materialen combineren een hoge sterkte en stijfheid met een relatief
lage dichtheid. Deze materialen kunnen echter complexe schadevormen vertonen,
zoals transversale scheuren en delaminaties. Deze schades kunnen de structurele
eigenschappen van een component aanzienlijk beïnvloeden. Periodieke inspecties
zijn noodzakelijk om de integriteit van een component gedurende zijn levensduur te
kunnen garanderen. De huidige inspectietechnieken zijn vaak tijdrovend, kostbaar
en vereisen dat de componenten gemakkelijk toegankelijk zijn. Het monitoren van
de structurele integriteit op basis van trillingen is een veelbelovend alternatief en
omvat het continue monitoren van een component middels een geïntegreerd sensor-
systeem. Deze technieken zijn gebaseerd op het concept dat het dynamische gedrag
van een component kan veranderen indien schade optreedt.
Hoewel er veel schade-identificatiemethodes zijn beschreven in de literatuur, zijn
er nog tal van moeilijkheden bij de praktische toepassing van deze methodes,
vooral voor complexe structuren. De prestaties van een schade-identificatiemethode
gebaseerd op trillingen zijn sterk afhankelijk van het ontwerp van een structuur
en de schade die wordt beschouwd. Dit proefschrift richt zich op de identificatie
van de schade in verstijfde composiet structuren. Het hoofddoel is om
ontwerprichtlijnen te ontwikkelen voor de detectie, lokalisatie en karakterisatie
van schade in verstijfde composiet structuren gebaseerd op veranderingen in het
dynamische gedrag.
Een literatuurstudie ondersteund door een analytisch model toonde aan dat de
kromming van trilvormen in combinatie met een modale schadeindex (MSE-DI)
algoritme een potentieel krachtige parameter en methode zijn voor de identificatie
van schade in verschillende verstijfde composiet structuren. Een experimentele
opstelling, inclusief een shaker en laser-vibrometer, is gebruikt om het dynamisch
gedrag te meten. Een lineair dynamische systeembeschrijving is verkregen
door het toepassen van een experimentele modaal analyse. De dynamische
metingen toonden aan dat het MSE-DI-algoritme in staat is om nauwelijks zichtbare
impactschade (BVID) in verstijfde composiet structuren te detecteren, te lokaliseren
en een schatting van de omvang te geven. Geconcludeerd is dat de methode
bijzonder effectief is voor het monitoren van de integriteit van de verbinding tussen
de huid en de verstijver. De methode bleef onbeslist in het geval dat de schade zich
puur in de huid van de structuur bevindt.
Experimenten toonden aan dat schade in de verbinding tussen de huid en

iii
iv

de verstijver duidelijke niet-lineaire effecten in het dynamische gedrag van een


component kan introduceren. Deze niet-lineaire effecten worden toegeschreven
aan de interactie tussen de huid en verstijver die optreedt tijdens de open-
en sluitbeweging van de schade. Er is aangetoond dat lineaire schade-
identificatiemethodes (bijvoorbeeld de modale domein methodes) geschikt zijn voor
lage excitatie amplitudes, maar dat de aanwezigheid van niet-lineaire dynamische
effecten niet verzwegen kan worden voor hogere amplitudes. De niet-lineaire
dynamische effecten kunnen fungeren als sterke aanwijzing voor schade, maar
kunnen ook nuttig zijn voor karakterisatie doeleinden.
De niet-lineaire dynamische effecten veroorzaakt door de schade in de verstijver,
spoort aan tot de ontwikkeling van niet-lineaire schade-identificatiemethodes.
Een studie gericht op het begrip en de toepasbaarheid van niet-lineaire vibro-
akoestische modulaties voor de detectie, lokalisatie en karakterisatie van impact-
schade in een composieten T-balk is uitgevoerd. Een analyse in het tijddomein
op verschillende locaties is gebruikt om schade in de verbinding tussen de huid
en de verstijver te detecteren en te lokaliseren op basis van lokaal verhoogde
amplitudemodulatie effecten. Analyse van de modulatie-eigenschappen geeft
mogelijkheden tot het karakteriseren van het niet-lineaire dynamische gedrag dat
is veroorzaakt door de schade in de verbinding tussen de verstijver en de huid van
de structuur.
Dit proefschrift laat zien dat de relatie tussen de eigenschappen van de structuur,
de potentiële schades en de schade-identificatiemethode samen de prestaties van
de op trillingen gebaseerde schade-identificatiestrategie bepalen. Daarom is
geconcludeerd dat het ontwerp van een strategie maatwerk is en dat het vereist
is dat er een grondig fysisch begrip is van de potentiële faalmechanismes, de
kritische schadelocaties en hun effect op het dynamische gedrag. Om te helpen
bij dit proces is er een procedure ontwikkeld voor het ontwerpen van een schade-
identificatiestrategie. Alle bevindingen gepresenteerd in dit proefschrift dragen
bij aan de ontwikkeling van een ontwerpgereedschap voor ingenieurs om de
implementatie van de schade-identificatietechnieken in composiet structuren te
bevorderen.
Contents

Summary i

Samenvatting iii

Nomenclature ix

1 Introduction 1
1.1 Background and motivation . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Composite structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2.1 Fiber reinforced plastics . . . . . . . . . . . . . . . . . . . . . . . 3
1.2.2 Damage types . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.2.3 Damage classification . . . . . . . . . . . . . . . . . . . . . . . . 5
1.3 Structural health monitoring . . . . . . . . . . . . . . . . . . . . . . . . 6
1.3.1 General . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.3.2 Classifications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.3.3 Techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.3.4 Major technology gaps . . . . . . . . . . . . . . . . . . . . . . . . 11
1.4 Objective and scope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.5 Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

2 Overview of vibration based damage identification methods 19


2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.2 Generalized description damaged system . . . . . . . . . . . . . . . . . 20
2.3 Literature overview vibration based methods . . . . . . . . . . . . . . . 22
2.4 Damage feature selection . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.4.1 Effect of damage on dynamic properties . . . . . . . . . . . . . 25
2.4.2 Information condensation . . . . . . . . . . . . . . . . . . . . . . 30
2.5 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

v
vi Contents

3 Vibration based structural health monitoring of a composite T-beam 39


3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.2 The modal strain energy damage index algorithm . . . . . . . . . . . . 41
3.3 Composite T-beam structure . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.4 Experimental analysis of a sub-structure . . . . . . . . . . . . . . . . . 44
3.4.1 Experimental outline . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.4.2 Set-up and vibration measurements . . . . . . . . . . . . . . . . 45
3.4.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

4 Damage identification in skin-stiffener structures based on curvatures 59


4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
4.2 Composite skin-stiffener structures . . . . . . . . . . . . . . . . . . . . . 61
4.3 Damage identification procedure . . . . . . . . . . . . . . . . . . . . . . 65
4.4 Experimental set-up and signal processing . . . . . . . . . . . . . . . . 69
4.5 Experimental results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
4.5.1 Two stiffener structure with damage scenario I and II . . . . . . 71
4.5.2 Three stiffener structure with damage scenario III . . . . . . . . 74
4.6 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
4.7 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

5 Nonlinear dynamic behavior of an impact damaged skin-stiffener structure 85


5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
5.2 Composite skin-stiffener structure . . . . . . . . . . . . . . . . . . . . . 87
5.3 Experimental work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
5.3.1 Set-up and signal processing . . . . . . . . . . . . . . . . . . . . 89
5.3.2 Initial global dynamic characterization . . . . . . . . . . . . . . 89
5.3.3 Harmonic waveform distortion . . . . . . . . . . . . . . . . . . . 92
5.3.4 Damage induced dynamic mechanisms . . . . . . . . . . . . . . 94
5.3.5 Spatial effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
5.3.6 Influence of excitation frequency . . . . . . . . . . . . . . . . . . 101
5.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
5.4.1 Underlying physical phenomena . . . . . . . . . . . . . . . . . . 102
5.4.2 Higher harmonics . . . . . . . . . . . . . . . . . . . . . . . . . . 104
5.5 Conclusions & future prospects . . . . . . . . . . . . . . . . . . . . . . . 106
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
Contents vii

6 Vibro-acoustic modulation based damage identification 109


6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
6.1.1 Background and motivation . . . . . . . . . . . . . . . . . . . . . 110
6.1.2 Vibro-acoustic modulation concept . . . . . . . . . . . . . . . . 110
6.1.3 Objective and outline . . . . . . . . . . . . . . . . . . . . . . . . 112
6.2 Theoretical description . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
6.2.1 Two-tone forced vibration of a nonlinear system . . . . . . . . . 113
6.2.2 Signal decomposition approach . . . . . . . . . . . . . . . . . . 115
6.2.3 Nonlinear response characteristics . . . . . . . . . . . . . . . . . 116
6.3 Experimental work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
6.3.1 Composite skin-stiffener structure . . . . . . . . . . . . . . . . . 118
6.3.2 Experimental set-up . . . . . . . . . . . . . . . . . . . . . . . . . 118
6.3.3 Experimental procedure . . . . . . . . . . . . . . . . . . . . . . . 120
6.4 Experimental results and discussion . . . . . . . . . . . . . . . . . . . . 122
6.4.1 Response decomposition . . . . . . . . . . . . . . . . . . . . . . 123
6.4.2 Spatial results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
6.4.3 Underlying dynamic behavior . . . . . . . . . . . . . . . . . . . 130
6.5 Conclusions & future prospects . . . . . . . . . . . . . . . . . . . . . . . 131
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133

7 Discussion 135
7.1 Design of an SHM strategy . . . . . . . . . . . . . . . . . . . . . . . . . 135
7.1.1 Scenario based design procedure . . . . . . . . . . . . . . . . . . 136
7.1.2 Combination of approaches . . . . . . . . . . . . . . . . . . . . . 140
7.2 Application of vibration based SHM . . . . . . . . . . . . . . . . . . . . 141
7.2.1 Integrated sensing . . . . . . . . . . . . . . . . . . . . . . . . . . 142
7.2.2 Operational and environmental effects . . . . . . . . . . . . . . 145
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145

8 Conclusions and recommendations 149


8.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
8.2 Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151

A Dynamics based nondestructive testing techniques 153

B Damage features and classifiers 157

C Hilbert transform 163

Dankwoord 165

Publications 169
Nomenclature

The symbols used in this thesis are classified into a Greek or a Roman symbol group.
Although some symbols can represent multiple quantities, its intended meaning
follows from the textual context. An overview of the most important abbreviations,
symbols and terminology used in the present thesis:

Greek symbols
α Extent of bilinearity [-]
βi Damage index of beam element i [-]
β ij Damage index of plate element ij [-]
(n)
γi Geometrical function of beam element i and mode n [1/m3 ]
(n)
γij Geometrical function of plate element ij and mode n [1/m2 ]
ǫ Control factor of the nonlinear function [-]
ζ (n) , ζ n Viscous damping of mode n [%]
θ Rotation angle [rad]
λn , λ∗n Poles (complex conjugate pair) of mode n [rad/s]
µ Mean value of the damage index β ij [-]
ν Poisson’s ratio [-]
ρ Density [kg/m3 ]
σn Damping factor of mode n [rad/s]
σ Standard deviation of the damage index β ij [-]
τ Shear stress [MPa]
φ Instantaneous phase [rad]
φp Phase of the pump excitation signal [rad]
φc Phase of the carrier excitation signal [rad]
χd Damage parameter (e.g. crack length, loss of stiffness)
χe Influence of the environmental and the operational
conditions (e.g. temperature, humidity)
ψn Mode shape vector of mode n [-]
ψmA , ψB Mode shape vector of case A and mode m, case B and mode n [-]
n
ω Frequency [rad/s]
ωdn Damped natural frequency of mode n [rad/s]
ωn Undamped natural frequency of mode n [rad/s]

ix
x Contents

ωp Pump excitation frequency [rad/s]


ωc Carrier excitation frequency [rad/s]
ωinst Instantaneous frequency [rad/s]

Roman symbols
Ac Carrier amplitude [m]
Asb1 Left sideband amplitude [m]
Asb2 right sideband amplitude [m]
Ainst Instantaneous amplitude (signal envelope) [m/s]
a Length of intact beam segment 1 [m]
ac Spanwise delamination location [m]
a Dimensionless length intact beam segment 1 [-]
ac Dimensionless spanwise delamination location [-]
b Length of delaminated beam segment [m]
b Dimensionless delamination length [-]
c Viscous damping coefficient [Ns/m]
c Length of intact beam segment 4 [m]
c Dimensionless length intact beam segment 4 [-]
D Flexural rigidity of a plate-like structure [Nm]
Dij Weighted average flexural rigidity of element ij of a plate-like [Nm]
structure
E Young’s modulus [N/m2 ]
Fop Operational load vector [N]
Ftest Test load vector [N]
Fa Excitation force [N]
Fp Pump excitation amplitude [N]
Fc Carrier excitation amplitude [N]
(n)
FB,i Local fractional modal strain energy of mode n and element i [-]
(n)
Fij Local fractional modal strain energy of mode n and element ij [-]
Fspring Spring force [N]
f Frequency [Hz]
fp Pump excitation frequency [Hz]
fc Carrier excitation frequency [Hz]
fa Harmonic excitation frequency [Hz]
f (n) , f n Natural frequency of the nth mode [Hz]
(n) (n)
fB , fT Natural frequency of the nth bending and torsion mode [Hz]
f inst Instantaneous frequency [Hz]
G Shear modulus [N/m2 ]
Gp , Gc Constants [-]
g Force vector of elastic forces, damping forces, etc. [N]
Contents xi

H Frequency response function matrix


HFv Frequency response function (mobility = [(m/s)/N)]
velocity/force)
HFx Frequency response function (admittance = [m/N]
displacement/force)
hin Transfer function between input signal u(t) and test loads
Ftest (t)
hout Transfer function between system response q(t) and
measured signal y(t)
hs Height of beam segment s = 1, 2, 3 and 4 [m]
h Beam height [m]
h2 Dimensionless transversal delamination location, h2 /h [-]
I Second moment of area [m4 ]
i Beam element number [-]
ij Plate element number [-]
J Torsion constant [m4 ]
k Stiffness [N/m]
l Length of the beam [m]
l x , ly Dimensions of the plate structure [m]
M Mass matrix [kg]
M Bending moment [Nm]
Ms Bending moment in beam segment s = 1, 2, 3 and 4 [Nm]
Ma Amplitude modulation [m/s]
Mf Frequency modulation [Hz]
MAC Modal assurance criterion [-]
m Mass [kg]
Ns Normal force in beam segment s = 1, 2, 3 and 4 [N]
Nmodes Number of modes [-]
Nx , Ny Number of elements in x- and y-direction [-]
n Mode number [-]
P Damage evolution function
Qn Modal scaling constant of mode n [-]
q Displacement vector [m]
qbp Bandpass filtered response [m]
Rn Residue matrix of mode n
S Fv Cross-power spectral density (force – velocity) [(Nm/s)/Hz]
S Fx Cross-power spectral density (force – displacement) [Nm/Hz]
S FF Auto-power spectral density (force – force) [N2 /Hz]
T Torque [Nm]
t Time [s]
t Thickness [m]
xii Contents

UT Elastic strain energy torsional deformations [Nm]


UB Elastic strain energy bending deformations [Nm]
(n)
UB Total modal strain energy of mode n [N/m]
(n)
UB,i Local modal strain energy of element i and mode n [N/m]
U (n) Total modal strain energy of mode n [N/m]
(n)
Uij Local modal strain energy of element ij and mode n [N/m]
u Input signal of the actuator
u Displacement [m]
u(n) Normalized mode shape of mode n [-]
V Voltage [V]
Vs Shear force in beam segment s = 1, 2, 3 and 4 [N]
vbp Bandpass filtered velocity response [m/s]
v Velocity response [m/s]
X Fourier spectra of the displacement signal, F ( x ) [m]
x Displacement signal [m]
x i −1 , x i Boundaries of element i in x-direction [m]
x, y, z Cartesian coordinates [m]
y Measured system response vector
y0 Baseline system response vector
∆y Deviation in the time domain system response
∆Y Deviation in the frequency or modal domain system response
y j −1 , y j Boundaries of element j in y-direction [m]
Zij Normalized damage index of element ij [-]
Zst Mechanical impedance of the structure [Ns/m]
z i −1 , z i Boundaries of element i in z-direction [m]
1, 2, 3 Material coordinate system [m]

Operators
∗ Complex conjugate
˜ Damaged case
T Transpose

Abbreviations
ADL Allowable damage limit
AE Acoustic emission
AU Acousto-ultrasonics
BVID Barely visible impact damage
CBM Condition based maintenance
CMIF Complex mode indicator function
EMA Experimental modal analysis
Contents xiii

EMI Electro-mechanical impedance


FIR Finite impulse response
FRF Frequency response function
ITD Integrated technology demonstrators
MAC Modal assurance criterion
MSE-DI Modal strain energy damage index algorithm
NDT Nondestructive testing
NLR National aerospace laboratories
ODS Operational deflection shape
PEKK Poly(ether-ketone-ketone)
RFP Rational fraction polynomial
SHM Structural health monitoring
SV Structural vibration
UL Ultimate load
UT Ultrasonic testing
VAM Vibro-acoustic modulation
VID Visible impact damage
Chapter 1

Introduction

1.1 Background and motivation

Composite materials combine a high strength and stiffness with a relatively low
density. This makes them extremely useful in applications where weight plays an
important role such as aircraft and wind turbines. These materials can, however,
exhibit unconventional and complex types of damage, like transverse cracks and
delaminations. These damage scenarios are often invisible, but can severely
influence the structural performance of a component, and hence tremendously
decrease its service life. Periodic inspections are required to ensure the integrity of a
component during its life. The current scheduled visual inspections are often time-
consuming, costly and require the components to be readily accessible, as shown in
Figure 1.1. Structural health monitoring (SHM) technologies propose a promising
alternative and involve the continuous monitoring of a structure by employing
a nondestructive testing (NDT) approach based on an integrated sensor system.
The output of this process is information regarding the ability of the structure to
perform its intended function in consideration of the applied loadings, aging and
degradation resulting from the operational environments.
Another motivation for the development of SHM technologies is that they
potentially can reduce the maintenance costs and increase the operational
availability of a system. Firstly, by making maintenance routines shorter and
more effective, given the actual physical condition of the component. Secondly, by
optimizing the maintenance intervals utilizing condition based maintenance (CBM)
routines rather than relying on the conventional time-based inspection intervals [1].
The use of structural health monitoring technologies will not only provide safety
benefits or enable new possibilities for maintenance concepts, but can also have a
significant influence on the design concepts. The change in design of lightweight
structures from a safe life to a damage tolerant design supported by a monitoring
system is, although far from reality, also considered as a potential weight-saving
benefit.

1
2 Chapter 1. Introduction

(a) Wind turbine inspection (b) Aircraft fuselage inspection (c) Bridge inspection
(photo courtesy of Spiegel (photo courtesy of Sandia (photo courtesy of Aspen
Online). National Laboratories, USA). Aerials, USA).

Figure 1.1 An overview of potential application fields for structural health monitoring techniques.

The research presented in this thesis was performed within the European research
program Clean Sky [2], which focuses on the development of breakthrough
technologies to significantly improve the environmental performances of airplanes
and air transport. Clean Sky comprises six Integrated Technology Demonstrators
(ITDs). The work presented here was part of the Eco-Design ITD, which
concentrates on green design and production, withdrawal and recycling of aircraft
by optimal use of raw materials and energy. The development of structural health
monitoring technologies is one of the objectives of this research program. Structural
components are usually replaced prior to their end of life, which is costly and
resource inefficient. A monitoring system can allow for an increase in the service
life of aircraft components and hence reduces the costs and long-term ecological
impact of these components.
To summarize, the development of structural health monitoring technologies for
composite structures is aiming to provide safety, cost saving (maintenance and
weight) as well as environmental benefits. However, the number of successful
practical applications of structural health monitoring technologies is still limited.
This is mainly due to the complexity of the composite components, the variety
of potential damage scenarios and the high performance demands of the damage
identification method. The work described in this thesis primarily concerns the
relation between those three aspects in order to achieve a higher level of maturity
of the structural health monitoring technologies. Background information about
composite structures and the typical damage scenarios is presented in the next
section, followed by a general introduction to structural health monitoring. The
1.2. Composite structures 3

knowledge contained in these sections is required to fully understand the theme of


this thesis.

1.2 Composite structures

1.2.1 Fiber reinforced plastics

A composite material is one in which two or more materials are combined to


obtain material properties that could not be obtained with the separate constituents.
Speaking about composite materials today often refers to combinations of a polymer
matrix and fiber reinforcement materials such as carbon or glass. This type of
composite is known as fiber reinforced plastic and is generally used in laminate
form. The fibers in these composites are used for their high strength and stiffness,
while the matrix transfers loads, binds the fibers together and protects them from
harsh environmental influences. Given the nature of the composites there are
numerous ways of combining matrix and fibers. Composite laminates also provide
the ability to stack the individual layers according to a desired lay-up. In this
way, every composite laminate can be tailored specifically for a certain application.
The scale of the fibers and the matrix or the scale of an individual layer are of
a lower level than the scale of the resulting laminate and structure as illustrated
in Figure 1.2. Composite materials are therefore multi-scale by nature [3]. The
mechanical properties at the structure or laminate level are the result of the material
characteristics at the lower levels.
Fiber reinforced plastics have some great advantages over more conventional
materials. They combine a high strength and stiffness with a low density. Fiber
reinforced plastics can also be formed into complex shapes, which allows for the
manufacturing of structures with complex geometries such as curved panels and
skin-stiffener structures. Other advantages include the high durability due to a
high resistance against corrosion, the possibility to tailor the material properties

Structure level

Macro Material levels

Meso
z Micro
x y
3
1 2

Figure 1.2 The multi-scale levels related to fiber reinforced plastics.


4 Chapter 1. Introduction

and the ability to reduce manufacturing costs by lowering the number of individual
components in an assembly. These strengths are why research in composites has
flourished over the past decades and why composites are more and more employed
over a variety of applications in, for example, the aerospace, automotive and marine
industry.

1.2.2 Damage types


Due to their complex nature, fiber reinforced plastics suffer from various damage
types unknown to homogeneous materials. These damage types can be examined
at different scales, according to the levels presented in Figure 1.2. Although they
originate at the material levels, they can be of profound importance for the integrity
of the structure.
One of the first damage mechanisms to occur is known as transverse (matrix)
cracking. This type of crack grows parallel to the fiber and in the thickness direction
of the laminate. Figure 1.3 illustrates a cracked laminate. In this case the crack
initiates at the free surface of the 0o layer, and grows in the thickness direction
until it meets the fibers of the 90o layer. Transverse cracks can be caused during
production by, for example, the difference in thermal expansion coefficient between
fiber and matrix or by in-service loading (e.g. impact). The small size makes them
generally hard to detect during inspections. The formation of transverse cracks
rarely means the total fracture of a laminate, as it does not affect the load carrying
capacity of the fibers. However, transverse cracks can influence the mechanical and
thermal properties of the laminate. Most importantly, this type of cracking forms a
trigger for further damage mechanisms.
Delamination is a damage type that generally is preceded by transverse cracking.
This damage type is a debonding between individual plies of a laminate. The crack
runs again in a plane parallel to the fibers, but at the interface between two layers.
Chronologically, it is recognized that a delamination mostly initiates from the tip of
a transverse crack. Figure 1.3 also shows a delamination at the interface between
the 0o and the 90o layer. Delaminations are hardly visible on the surface, since they

Transverse crack
Delamination

3
2
1

Figure 1.3 Schematic representation of transverse crack and delamination in a [0/90/0]s laminate.
1.2. Composite structures 5

Figure 1.4 Cross section of a glass-reinforced epoxy laminate with transverse cracks and delaminations
caused by an impact test.

are embedded within the composite structure. This makes them barely detectable
during, for example, visual inspections. Figure 1.4 shows a micrograph of a cross
section of a laminate with both transverse cracks and delaminations originating
from an impact test. Although delaminations do not lead to complete fracture, they
can seriously affect the thermal and mechanical properties of the laminate.

The type of damage that significantly decreases the load carrying capacity of a
laminate is fiber failure. Fiber related failure in a laminate is mostly accompanied
by matrix related damage like transverse cracks and delaminations. Typical failure
modes can involve local buckling of fibers, fiber breakage and fiber pull-out.

1.2.3 Damage classification

Structural components used in aircraft are usually designed according to the


damage tolerance principle. This principle implies that the structure needs to
function safely despite the presence of (minor) flaws. The severity of damage in
aircraft structures is classified into five categories according to [4]. This classification
is linked to the required residual strength and ranges from allowable damage,
category 1, up to very severe damage, category 5. The assessment of damage in
aircraft structures has historically relied on visual inspection methods to identify
damage. Category 1 is classified as barely visible impact damage (BVID) and
may remain undetected, while repair scenarios are required for the visible impact
damage (VID) of category 2 to 5. Structures containing BVID must sustain ultimate
load (UL) for the life of the aircraft structure. The dent depth is often used as
the damage metric to define BVID [5]. This criterion sets the lower bounds to the
identification capabilities of a structural health monitoring approach.
6 Chapter 1. Introduction

1.3 Structural health monitoring

1.3.1 General
Structural health monitoring is the multidisciplinary process of implementing a
strategy for damage identification in a way that nondestructive testing becomes
an integral part of the structure. This process involves the definition of potential
damage scenarios for a structure, the observation of the structure over a period of
time using periodically spaced measurements, the extraction of damage sensitive
parameters (features) from these measurements and the analysis of these features
to determine the current state of health of the structure (classification). The output
of this process is periodically updated information regarding the ability of the
structure to perform its intended function in consideration of the applied loadings,
aging and degradation resulting from the operational environments.
In contrast to conventional nondestructive testing techniques that are operated off-
line during maintenance, structural health monitoring techniques can be operated
off-line as well as on-line. On-line refers, in this case, to the monitoring during
operation of the system or structure. The structural health monitoring technique is
part of the on-board systems. Sensors are permanently attached (surface sensors)
or embedded (integrated sensors) in the structure. As a result, information on the
structural state is available at arbitrary times.
As stated by Farrar and Doebling [6], the process of structural health monitoring is
fundamentally one of statistical pattern recognition. A statistical pattern recognition
process aims to classify data (patterns) based on either a priori knowledge or on
information extracted from the patterns. The patterns to be classified are usually
groups of measurements or observations. A pattern recognition process covers all
stages of an investigation from problem formulation up to the interpretation of the
results. Figure 1.5 shows a simplified process, which consists of a diagnostic and a
prognostic part. The diagnostic analyses are used to estimate the current state of the
structure. The prognostic analysis evaluates the damage evolution and estimates
the residual service life [7]. The present thesis focuses on the diagnostic part of the
structural health monitoring process, which can be divided into a four-step process:

Diagnostics Prognostics
Representation Current Damage
Actuation Response pattern state evolution Decision
Feature pattern
Repair,
Sensor Feature Failure
Structure Classifier replacement,
system extractor probability residual life?
High
Level 1, 2, 3 Level 4 Low

Damage

Figure 1.5 The multidisciplinary structural health monitoring process.


1.3. Structural health monitoring 7

1. Operational evaluation
Operational evaluation answers questions regarding the implementation of
the structural health monitoring system, such as possible failure modes,
operational and environmental conditions and data acquisition related
limitations.
2. Data acquisition
This step defines the data acquisition in terms of the quantities to be
measured, the type and quantity of sensors to be used, the locations where
these sensors are to be placed and the hardware to be used. Moreover, it
defines the data fusion and cleansing, which is the determination of which
data is necessary and useful in the feature extraction process.
3. Feature extraction
This step in the structural health monitoring process receives the most
attention in the literature. Feature extraction is the process of identifying
damage sensitive parameters from measured data. These damage features are
defined in the time, frequency or modal domain. Information reduction and
condensation is also of concern for a large quantity of data, particularly if
comparisons of many measurements over the service life of the structure are
required.
4. Classification
The last step is concerned with the implementation of algorithms (e.g. neural
networks) that operate on the extracted features to distinguish between the
damaged and the undamaged structural state and to quantify the damage
state of the structure. Statistical methods are used to establish the feature’s
sensitivity to damage and to prevent false damage identification.

According to Doebling et al. [8], an ideal robust damage identification scheme


should be able to: detect damage at a very early stage, locate the damage within the
sensor resolution being used, provide some estimate of the extent or severity of the
damage and predict the remaining useful life of the structural component in which
damage has been identified, all independent from changes in the operational and
environmental conditions. The method should also be well suited to automation,
and should be independent of human judgement and ability.

1.3.2 Classifications
Damage identification methods can be classified in different ways. This section
summarizes the most important classifications used in this thesis.

Performance levels A performance based classification of the damage identification


methods was introduced by Rytter [9]. Rytter defined four levels of damage
8 Chapter 1. Introduction

identification:

• Level 1: Verification of the presence of damage in a structure.


• Level 2: Determination of the location of the damage.
• Level 3: Estimation of the extent / severity of the damage.
• Level 4: Prediction of the remaining service life of the structure.

Some researchers [10–12] included the determination of the type of damage


(characterization) as an additional step between level 2 and 3. Levels 1 to
3 are related to the damage diagnosis, while level 4 is concerned with the
damage prognosis. Higher levels generally represent an increasing degree of
complexity and a greater need for mathematical models. Generally, a level 4
prediction requires a fracture mechanics and fatigue life analysis based on structural
and damage models to predict the evolution of the damage [7].

Model and non-model based approach The second classification distinguishes


two approaches, namely model and non-model based damage identification methods.
In a non-model based method the results are compared with the results of
a reference measurement performed prior to setting the structure in service.
Deviances in the damage sensitive parameters are used to identify damage. In
a model based technique the response is compared with some form of model.
This can either be an analytical or a numerical (e.g. finite element) model.
Advantages of model based techniques are that these could well be extended to
provide information about the severity of the detected damage and can be used to
account for environmental or operational variations (e.g. temperature, boundary
conditions). On the contrary, it is rather difficult to obtain an accurate model
representation of complex (composite) structures. Moreover, the computational
costs can limit the applicability for in situ monitoring.

Local and global methods Damage identification techniques are usually classified
as local or global [13]. This classification is based on the relative size of the area that
can be inspected at once by the method with respect to the overall dimensions of
the structure. The local methods concentrate on a part of the structure and are
usually considered to be more sensitive than the global methods. They are capable
of detecting small damages such as cracks, but their application requires a prior
knowledge of the location of the damaged area. The global methods can analyze
a relatively large area at once, but the resolution is, however, rather limited. As a
consequence, only relatively severe damage cases can be identified.

Baseline and non-baseline One of the fundamental axioms of Structural Health


Monitoring proposed by Worden, et al. [11] reads that the assessment of damage
requires a comparison between two system states. The response of a structure
1.3. Structural health monitoring 9

Table 1.1 An overview of the most commonly used nondestructive testing (NDT) techniques.

Technique Ref. Inspection Inspection Structure


area mode accessibility
Electric, magnetic and
electromagnetic
Electrical conductivity [16, 17] Local/global Off-/on-line Not required
testing
Magnetic particle testing [18, 19] Local Off-line Required
Eddy current testing [18, 20] Local Off-line Required
Radiography (X-ray) [18, 21] Local Off-line Required
Infrared thermography [18, 22] Local/global Off-line Required
Mechanic, dynamic
(Quasi-) static [23] Local Off-/on-line Not required
Structural vibrations [24, 25] Local/global Off-/on-line Not required
and acoustics
Electro-mechanical [26, 27] Local/global Off-/on-line Not required
impedance
Acoustic emission [18, 28] Local/global On-line Not required
Acousto-ultrasonics [29, 30] Local/global Off-/on-line Not required
Ultrasonic testing [18] Local Off-line Required
Optical
Shearography [31, 32] Local Off-line Required
Visual inspection [5, 18] Local/global Off-line Required

measured at an earlier stage is usually utilized as a baseline to distinguish between


the damaged and undamaged state. For model based methods, this baseline can
also be obtained from a model (e.g. finite element model). Other researchers [14, 15]
also propose methods that do not require a baseline to classify the structure as
damaged or undamaged. This might be interpreted as not requiring a comparison
of system states. It can be argued that this discrepancy is a matter of terminology.
Non-baseline methods still compare two states, but instead of utilizing a baseline
measurement they rely on an assumed normal behavior (e.g. a smooth pattern or
a linear-elastic response) of the structure. The system is in this case classified as
damaged when the response deviates from the norm.

1.3.3 Techniques
A wide range of nondestructive testing techniques can be employed for damage
identification purposes. An overview of the most commonly used nondestructive
testing techniques and their characteristics is presented in Table 1.1. The majority
of these techniques can only be applied when the structure is not in operation
(‘off-line’) and readily accessible. Consequently, only a few of these techniques
are suitable to be applied in a health monitoring environment. As part of this
10 Chapter 1. Introduction

Table 1.2 An overview of the dynamics based nondestructive testing (NDT) technologies.

Technology Frequency Actuation Sensitivity Ease of data Applicability


range [Hz]1 approach to damage interpretation for SHM
Structural vibration 100 –104 active/   
and acoustics passive
Electro-mechanical 103 –105 active   
impedance
Acoustic emission 104 –106 passive   
Acousto-ultrasonics 104 –106 active   
Ultrasonic testing 105 –107 active   
1 Typical frequency range at which the technology is operating.

selection, the technologies based on electrical conductivity are generally limited to


conductive materials. The (quasi-) static techniques are of lower interest because
of a rather low sensitivity to damage compared to the dynamics based techniques.
The dynamics based techniques are applicable to a wide range of structures and are
therefore considered to be a promising group of technologies for structural health
monitoring.

The basic principles of the dynamics based techniques are described in Appendix A.
Each technique comprises a large amount of literature and is usually treated
as a different field of research. Table 1.2 provides a more detailed comparison
of their performances. The low frequency structural vibration (SV) and electro-
mechanical impedance (EMI) techniques primarily rely on standing wave patterns,
while the higher frequency acoustic emission (AE), acousto-ultrasonics (AU) and
ultrasonic testing (UT) utilize traveling wave characteristics. The former group of
methods provide data that is relatively easy to interpret. More complex structures
can be analyzed with these methods and a relatively large area can be explored at
once. The frequency range, and hence the resolution, is however limited [13]. As a
consequence, only relatively severe damage such as delaminations can be identified.
The latter group of methods are usually considered to be more sensitive. They
are capable of detecting small damage such as cracks [29]. For that reason, these
wave propagation based technologies are increasingly being explored for aircraft
applications [33, 34]. The downside is the more complex interpretation of the data,
in particular in case of non-flat or complex (composite) structures [35]. The rating
for the sensitivity is linked to the operational frequency range [11], while the other
aspects are ranked according to the available literature. It should be noted that these
ratings are rather subjective. The intention here is, however, to give an impression
of the relative strengths and weaknesses rather than to condemn techniques.
1.3. Structural health monitoring 11

1.3.4 Major technology gaps

Although many structural health monitoring techniques have been proposed in the
literature, there are still numerous difficulties in the practical application of these
approaches. The most important technical issues that need to be resolved before
structural health monitoring technologies can make the transition from a research
topic to actual practice are summarized below.

• Complex composite structures


The structural health monitoring technologies are extensively tested on
concrete and metallic structures. The applications to composite structures
are to a large extent limited to relatively simple composite beams and plates
with mainly well-defined or artificial damage scenarios. The complexity of
the components and the wide variety of potential damage scenarios hampers
the application of structural health monitoring to more complex composite
structures. Therefore, research should be focused on the application to
composite structures such as stiffened panels and torsion boxes, as well as
realistic damage scenarios.
• Selection damage feature and classifier
Damage identification aims to uniquely identify damage at an early stage with
a minimum of false positive results. For this purpose, an enormous amount
of damage features and (statistical) classifiers are addressed in the literature
with a varying level of success. None of the methods solves all problems in
all structures. The development and selection of damage sensitive features
and classifiers that provide a high detection probability without getting false
alarms is therefore one of the key challenges for structural health monitoring.
• High performance level
Current health monitoring approaches are often capable to detect (level 1)
and localize (level 2) damage, but are limited in their ability to estimate the
type or extent/severity (level 3) of the damage accurately. Damage severity
assessment is an important requirement for the analysis of the damage
evolution and the prediction of the remaining lifetime (level 4). The evolution
towards a high performance level is considered as an important step forward
in the development of autonomous monitoring of the integrity of structures.
• Integrated sensors and network
A structural health monitoring system requires an integrated sensor system.
The design and implementation of these systems involve numerous challenges.
These challenges range from the selection of the optimal position and number
of sensors and the monitoring of failure or debonding of a sensor to the data
transmission and the supply or harvesting of power. Consequently, a large
part of the research in the field of structural health monitoring is dedicated to
the development of sensor systems.
12 Chapter 1. Introduction

• Operational and environmental variability


A large obstacle for the practical application of structural health monitoring
technologies is the dependency of damage parameters on the operational
and environmental conditions, such as temperature, humidity, loads and
boundary conditions. Changes in these conditions can mask or magnify
the effects that are resulting from the damage. Methods should have the
ability to separate the damage related effects from those that are coming from
changes in environmental conditions. A wide variety of methods, comprising
statistical techniques and model based methods, are presented in the literature
to compensate for these variations, but confidence in these methods is lacking.

In addition to the technical issues described above, there are other nontechnical
issues that must be addressed before structural health monitoring technologies
can make the transition to actual application. These issues include, for example,
convincing operators, engineers and authorities of the potentials of the technology
as well as the certification of the technologies. More detailed discussions on this
topic are provided by Boller [36] and Farrar et al. [37].

1.4 Objective and scope


The development of a structural health monitoring strategy involves multidisciplinary
research challenges, as was shown in the previous section. Figure 1.6 schematically
illustrates the associated multidisciplinary framework. This framework comprises
four components (i.e. structure, damage identification method, damage scenario,
actuation and sensing technology). The characteristics of these components are
closely interconnected and together they define the performance of the structural
health monitoring strategy. Ideally, a strategy combines a high probability of
detection and a high performance level with a low number of false positives. The
success of a damage identification strategy is, however, dependent on the actual
structure and the damage scenario that is considered. The selection of the most
suitable approach is, therefore, far from straightforward and is finally a matter of
compromise. This gives rise to the development of a dedicated tool that can be
used to design a damage identification strategy depending on the type of structure
and the potential threats. Design recommendations and guidelines are required for
each scenario to assist in the development of such a tool.
This thesis is dedicated to the identification of damage in composite skin-stiffener
structures. Stiffened composite skins are a widely used engineering structure.
Besides the application in wind turbine blades, skin-stiffener structures are used
in nearly all aircraft wing and fuselage designs. Stiffeners are used to increase
the bending stiffness of the component without a severe weight penalty. A primary
failure mode for these structures is delamination damage at the connection between
1.4. Objective and scope 13

Sub-components Multidisciplinary framework

Damage
Structure identification
method
Strategy for
structural
health
monitoring

(Actuation)
Damage
& sensing
scenario
technology

Figure 1.6 The multidisciplinary framework for the design of a structural health monitoring system.

skin and stiffener. Impacts near these connections can lead to local skin-stringer
separation. This is a safety-critical failure mode, because it can significantly affect
the structural performance of the component while remaining invisible from the
outer surface. Skin-stiffener structures are therefore considered as a good candidate
for health monitoring.

The structural vibration based health monitoring approaches are considered in the
present work. These methods are based on the concept that the dynamic behavior
of a structure can change if damage occurs. The motivation is twofold: firstly,
because they do not require the structure to be readily accessible. Secondly, because
these low frequency methods provide data that is relatively easy to interpret.
This provides opportunities to analyze more complex structures, such as the skin-
stiffener structures. A drawback of the vibration based methods is the limited
sensitivity. The identification of barely visible impact damage, as discussed in
Section 1.2.3, sets the lower bound for the capabilities of the approach.

In summary, the objective of the research presented in this thesis is:

To develop guidelines for the detection, localization and characterization


of damage in composite skin-stiffener structures based on changes in the
dynamic behavior.

This work will contribute to the development of a design tool for research engineers,
to assist the implementation of structural health monitoring technology in safety-
critical composite structures.
14 Chapter 1. Introduction

Numerical

Loendersloot, Loendersloot,
et al. [38] et al. [39]
Linear
dynamic Experimental
behavior
2 & 3 stiffener
T-beam
structure
Chapter 3 Chapter 4
Vibration
Introduction Discussion Conclusions
methods
Chapter 1 Chapter 2 Chapter 7 Chapter 8
Nonlinear Vibro-acoustic
dynamics modulations
Nonlinear Chapter 5 Chapter 6
dynamic
behavior

Figure 1.7 Schematic overview of the thesis outline.

1.5 Outline
The outline of the thesis is schematically illustrated in Figure 1.7. The core
comprises five chapters. Chapters 3 to 6 are reproduced from research papers. As a
consequence, some of the essential details are repeated in the different chapters. The
author apologizes for any inconvenience caused by the chosen presentation. From
a more positive point of view, however, the reader is able to study any individual
chapter without having to miss out on any essential details.
Chapter 2 provides an overview of the vibration based damage identification
methods. The basic concept of these methods is explained based on a generalized
description of a damaged system. A literature study supported by an analytical
beam model are used to select a potentially powerful approach for the detection,
localization and characterization of damage in the skin-stiffener structure. Two
approaches are considered. The first approach utilizes mode shape curvatures and
assumes a linear dynamic behavior, while the second approach is focused on the
nonlinear dynamic effects that are introduced by the damage.
Chapter 3 is focused on the experimental feasibility of a vibration based damage
identification method to identify an artificial delamination at the skin-stiffener
connection of a composite T-beam. A force-vibration set-up, including a laser-
vibrometer system, is used to measure the dynamic behavior of the T-beam
experimentally. Both bending and torsion modes are considered in the analysis.
Special attention is paid to the effect of the number of measurement points.
Chapter 4 extends the work described in Chapter 3 to larger and more complex
skin-stiffener structures (i.e. a 2 stiffener structure and a 3 stiffener structure with
non-uniform skin thickness). Impact induced damage scenarios are considered.
References 15

The relation between the damage location, the structural design and the dynamic
behavior is investigated in order to extract recommendations for the effective
application of the methodology. Readers are referred to the book chapters of
Loendersloot et al. [38, 39] for numerical studies utilizing finite element models
of the structures that were used in Chapters 3 and 4.
The vibration based method used in Chapters 3 and 4 assumes a linear dynamic
behavior. The observation of potential nonlinear dynamic effects caused by
damage can also be a strong indicator of the damage. Chapter 5 describes a
study on the interaction of a low frequency vibration with skin-stiffener damage.
This interaction can yield dynamic phenomena that exhibit complicated nonlinear
behavior. Different phenomena are linked to measured waveforms with the help of
phase portraits.
The nonlinear effects introduced by skin-stiffener damage urges the development
of nonlinear damage identification methods. Chapter 6 concerns a study on the
understanding and feasibility of using nonlinear vibro-acoustic modulations for
the detection, localization and characterization of impact damage in a composite
T-beam. A time domain analysis of the vibro-acoustic modulation phenomena
is presented at multiple spatial locations. From a broader perspective, this work
intends to contribute to the development of enhanced methods for the identification
and characterization of damage in advanced composite structures.
The complete work is put into a broader perspective in Chapter 7. Additional
design recommendations and guidelines are extracted based on the work presented
in this thesis. The practical application of the methods is discussed. Finally,
Chapter 8 presents the important conclusions and provides the recommendations
for further research.

References
[1] T. Tinga. Application of physical failure models to enable usage and load based
maintenance. Reliability Engineering & System Safety, 95(10):1061–1075, 2010.
[2] Clean Sky website: http://www.cleansky.eu, visited on June 28th , 2013.
[3] R.L. Foye. Finite element analysis of the stiffness of fabric reinforced composites.
Technical report, NASA Contractor Report, 1992.
[4] Volume 3. Polymer matrix composites: material usage, design and analysis. In
Composite materials handbook CMH-17-3G, Chapter 12, page 952. SAE International, 2012.
[5] J. Baaran. Visual inspection of composite structures. Technical report, Institute of
Composite Structures and Adaptive Systems, DLR Braunschweig, 2009.
[6] C.R. Farrar, S.W. Doebling, and D.A. Nix. Vibration-based structural damage
identification. Philosophical Transactions of the Royal Society A: Mathematical, Physical and
Engineering Sciences, 359(1778):131–149, 2001.
[7] C.R. Farrar, H. Sohn, F.M. Hemez, M.C. Anderson, M.T. Bement, P.J. Cornwell, S.W.
Doebling, J.F. Schultze, N. Lieven, and A.N. Robertson. Damage prognosis: current
16 Chapter 1. Introduction

status and future needs. Technical report, Los Alamos National Laboratory, Los
Alamos, NM, 2001.
[8] S.W. Doebling, C.R. Farrar, M.B. Prime, and D.W. Shevitz. Damage identification and
health monitoring of structural and mechanical systems from changes in their
vibration characteristics: a literature review. Technical report, Los Alamos National
Laboratory, NM, USA, 1996.
[9] A. Rytter. Vibration based inspection of civil engineering structures. Ph.D. thesis, Aalborg
University, 1993.
[10] H. Sohn, C.R. Farrar, F.M. Hemez, D.D. Shunk, D. Stinemans, and B.R. Nadler. A
review of structural health monitoring literature: 1996-2001. Technical report, Los
Alamos National Laboratory, Los Alamos, NM, 2003.
[11] K. Worden, C.R. Farrar, G. Manson, and G. Park. The fundamental axioms of structural
health monitoring. Proceedings of the Royal Society A: Mathematical, Physical and
Engineering Sciences, 463(2082):1639–1664, 2007.
[12] K. Worden and J.M. Dulieu-Barton. An overview of intelligent fault detection in
systems and structures. Structural Health Monitoring, 3(1):85, 2004.
[13] C.-P. Fritzen and P. Kraemer. Self-diagnosis of smart structures based on dynamical
properties. Mechanical Systems and Signal Processing, 23(6):1830–1845, 2009.
[14] E.S. Sazonov, P. Klinkhachorn, U.B. Halabe, and H. V.S. GangaRao. Non-baseline
detection of small damages from changes in strain energy mode shapes. Nondestructive
Testing And Evaluation, 18(3-4):91–107, 2003.
[15] S. Park, C. Lee, and H. Sohn. Reference-free crack detection using transfer impedances.
Journal of Sound and Vibration, 329(12):2337–2348, 2010.
[16] J.C. Abry, S. Bochard, and A. Chateauminois. In situ detection of damage in CFRP
laminates by electrical resistance measurements. Composites Science and Technology,
59:925–935, 1999.
[17] R. Schueler, S.P. Joshi, and K. Schulte. Damage detection in CFRP by electrical
conductivity mapping. Composites Science and Technology, 61(6):921–930, 2001.
[18] C. Hellier. Handbook of nondestructive evaluation. McGraw-Hill, 2003.
[19] ISO 9934-1 Non-destructive testing – Magnetic particle testing – Part 1: general
principles.
[20] B.A. Auld and J.C. Moulder. Review of advances in quantitative eddy current
nondestructive evaluation. Journal of Nondestructive Evaluation, 18(1), 1999.
[21] P.J. Schilling, B.R. Karedla, A.K. Tatiparthi, M.A. Verges, and P.D. Herrington. X-ray
computed microtomography of internal damage in fiber reinforced polymer matrix
composites. Composites Science and Technology, 65(14):2071–2078, 2005.
[22] X.P.V. Maldague. Introduction to NDT by active infrared thermography. Materials
Evaluation, 60(9):1060–1073, 2002.
[23] A. Kesavan, S. John, and I. Herszberg. Strain-based structural health monitoring of
complex composite structures. Structural Health Monitoring, 7(3):203–213, 2008.
[24] E.P. Carden and P. Fanning. Vibration based condition monitoring: a review. Structural
Health Monitoring, 3(4):355–377, 2004.
[25] D. Montalvão, N.M.M. Maia, and A.M.R. Ribeiro. A review of vibration-based
structural health monitoring with special emphasis on composite materials. Shock and
Vibration Digest, 38(4):295–326, 2006.
[26] G. Park, H. Sohn, C.R. Farrar, and D.J. Inman. Overview of piezoelectric
impedance-based health monitoring and path forward. The Shock and Vibration Digest,
References 17

35(6):451–463, 2003.
[27] V. Gopal, M. Annamdas, and C.K. Soh. Application of electromechanical impedance
technique for engineering structures: review and future issues. Journal of Intelligent
Material Systems and Structures, 21(1):41–59, 2009.
[28] T. Kundu, S. Das, and K.V. Jata. Detection of the point of impact on a stiffened plate by
the acoustic emission technique. Smart Materials and Structures, 18(3):035006, 2009.
[29] A. Raghavan and C.E.S. Cesnik. Review of guided-wave structural health monitoring.
The Shock and Vibration Digest, 39(2):91–114, 2007.
[30] Z. Su, L. Ye, and Y. Lu. Guided lamb waves for identification of damage in composite
structures: a review. Journal of Sound and Vibration, 295(3-5):753–780, 2006.
[31] Y.Y. Hung and H.P. Ho. Shearography: an optical measurement technique and
applications. Materials Science and Engineering: R: Reports, 49(3):61–87, 2005.
[32] Y.Y. Hung. Applications of digital shearography for testing of composite structures.
Composites Part B: Engineering, 30(7):765–773, 1999.
[33] W.J. Staszewski, S. Mahzan, and R. Traynor. Health monitoring of aerospace composite
structures – Active and passive approach. Composites Science and Technology,
69(11-12):1678–1685, 2009.
[34] K. Diamanti and C. Soutis. Structural health monitoring techniques for aircraft
composite structures. Progress in Aerospace Sciences, 46(8):342–352, 2010.
[35] R.P. Dalton, P. Cawley, and M.J.S. Lowe. The potential of guided waves for monitoring
large areas of metallic aircraft fuselage structure. Journal of Nondestructive Evaluation,
20(1), 2001.
[36] C. Boller. Structural health monitoring – Its association and use. In W. Ostachowicz
and J.A. Güemes, editors, New Trends in Structural Health Monitoring, volume 542 of
CISM International Centre for Mechanical Sciences, Chapter 1, pages 1–79. Springer
Vienna, 2013.
[37] C.R. Farrar and K. Worden. An introduction to structural health monitoring.
Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering
Sciences, 365(1851):303–15, 2007.
[38] R. Loendersloot, T.H. Ooijevaar, L.L. Warnet, A. de Boer, and R. Akkerman. Vibration
based structural health monitoring and the modal strain energy damage index
algorithm applied to a composite T-beam. In C.M.A. Vasques and J.D. Rodrigues,
editors, Vibration and Structural Acoustics Analysis: Current Research and Related
Technologies, Chapter 6, pages 121–150. Springer, 2011.
[39] R. Loendersloot, T.H. Ooijevaar, A. de Boer, and R. Akkerman. Development of a
damage quantification model for composite skin-stiffener structures. In C. Boller and
H. Janocha, editors, New Trends in Smart Technologies, pages 99–108. Fraunhofer Verlag,
2013.
Chapter 2

Overview of vibration based damage


identification methods

Abstract
An overview of the vibration based damage identification methods
is presented in this chapter. The basic concept of these methods is
explained based on a generalized description of a damaged system.
One of the challenges in the development of a successful approach
is the selection of a damage identification method. A literature
study supported by an analytical model showed that mode shape
curvatures combined with the modal strain energy damage index
(MSE-DI) algorithm are a potentially powerful damage feature and
classifier. However, when severe nonlinear dynamic effects are
introduced by the damage, the modal domain based methods are
suffering under their linear system assumption. Alternative methods
in, for example, the time or frequency domain are in that case
required to obtain an adequate description of the nonlinear behavior
introduced by the damage.

19
20 Chapter 2. Overview of vibration based damage identification methods

2.1 Introduction

The basic concept of the vibration based damage identification methods is that
the dynamic behavior of a structure can change if damage occurs [1]. Damage in
a structure can alter the structural integrity and therefore the physical properties
like stiffness, mass and/or damping. The dynamic behavior of a structure is a
function of these physical properties and will therefore directly be affected by the
damage. The dynamic behavior can be described by time, frequency and modal
domain parameters. The changes in these parameters or properties derived from
these parameters are used as indicators of damage.
The structural vibration based approaches generally allow for a relatively easy
interpretation of the measured responses, have the ability to analyze complex
structures and do not require the structure to be readily accessible in order to
be able to identify damage [2]. Drawbacks are, however, the limited sensitivity
compared to higher frequency approaches and the number of required sensors in
case the standing wave patterns need to be described [3]. As a consequence, the
application of structural vibration based methods has to be tailored to the structure
and the expected damage cases to fully utilize the potential of this technology.
This chapter has two objectives. The first objective is to provide an overview of
the structural vibration based damage identification methods. For this purpose,
a fundamental description of the structural vibration based damage identification
problem is given, followed by a short literature overview of the damage features
and (statistical) classifiers that are commonly addressed. The second objective is to
select a promising damage identification method for the detection, localization and
characterization of skin-stiffener damage. The selection of an appropriate method
is often all but straightforward and finally a matter of compromise. To aid in
this process, two basic principles are discussed, namely the effect of the potential
damage case on the dynamic behavior and the consequences involved with the
information reduction in the signal processing. The former determines whether the
parameter is sensitive to the damage and is analyzed by considering an analytical
model of a delaminated composite beam. This model is a simplified representation
of the damage that is expected in skin-stiffener connections. The latter is explained
and demonstrated based on a mass-spring-damper system with a bilinear stiffness,
representing the potential opening and closing behavior of defects in a skin-stiffener
connection.

2.2 Generalized description damaged system

The dynamics of a general time-varying damaged structure can be described by the


coupled system of the nonlinear equation of motion and the nonlinear evolution of
2.2. Generalized description damaged system 21

the damage according to Fritzen et al. [3]:

M (χd , χe , t)q̈ + g (χd , χe , q, q̇, t) = Fop (t) + Ftest (t) , (2.1)


χ̇d = P (χd , χe , q, q̇, t) , (2.2)

where M is the mass matrix, g the force vector of elastic forces, damping forces,
etc. depending on the displacements q, the velocities q̇, the time t, a damage
parameter χd and a parameter χe that indicates the influence of the environmental
and operational conditions (e.g. temperature, humidity). The damage parameter
χd characterizes the damage scenario in terms of e.g. the crack length or the loss
of stiffness. Equation (2.1) describes the fundamentals of the damage diagnosis
process, whereas Equation (2.2) provides a starting point for a prognostic analysis
(e.g. damage evolution and residual service life estimation) [4]. The damage
evolution is usually considered as a rather slow process compared to the dynamic
behavior of the structure. The damage parameter χd is, therefore, assumed constant
during the short time span of a dynamic measurement. The forced excitation of the
system consists of the operational loads Fop (t) and the test loads Ftest (t). The latter
is absent in the case of a passive method. For the active approach, the test loads are
generated by an actuator (e.g. piezoelectric element, shaker). The transformation
of the input signal u(t) of the actuator to the forces and moments applied to the
structure can be described by:

Ftest (t) = hin (χd , χe , u(t)) . (2.3)

A similar transfer function can be defined for the response side of the system:

y(t) = hout (χd , χe , q, q̇, t) . (2.4)

The transfer function hout relates the state variables to the actually measured system
response y(t) (e.g. strains or voltage). The damage parameter χd in Equations (2.3)
and (2.4) can also account for any failure or debonding of the actuator or sensor.

Although the damage parameter χd provides the most direct description of the
damage, this parameter can usually not be measured directly. Alternatively, the
changes in the dynamic response related to this damage parameter is considered
in the damage identification process. For this purpose, the system response y(t) is
usually compared with an earlier obtained baseline response y0 . The associated
forward problem relates the deviations in the response ∆y(t) with the damage
characteristics χd and the environmental and operational conditions χe :

∆y = y − y0 = f (χd , χe ) . (2.5)

The key challenge for damage diagnosis is, however, to solve the inverse problem,
22 Chapter 2. Overview of vibration based damage identification methods

Selection of damage
Damage type feature & classifier
Environmental
Inverse problem influences
Transverse -1 χe Time
cracking
χ d = f ( Δy , χ e )
Δy(t)
Damage Changes in Changes in
Frequency
Delamination χd structural dynamic
ΔY(ω)
properties behavior
Fiber Forward problem Modal
failure Δy= f ( χ d , χ e ) ΔY(ωn)

Figure 2.1 Schematic illustration of the forward and inverse problem involved with vibration based damage
identification of composite structures.

which is defined as:

χd = f −1 (∆y, χe ) . (2.6)

The complexity involved with this problem comes from the non-uniqueness of the
solution and the usually incomplete response data due to a limited number of
measurement locations.
Figure 2.1 schematically illustrates the forward and inverse problem. They connect
the potential damage scenarios in composite materials (left side) with the wide
range of damage features (right side). Damage identification, first, aims to
maximize the correlation between the damage parameter and the selected damage
feature. Second, it is practically also desired to condense the information to
a lower dimensional parameter space. The deviations in the measured time
domain parameters ∆y(t) are therefore often replaced by parameters from the
frequency ∆Y (ω ) or modal ∆Y (ωn ) domain. For both purposes, the selection
of an appropriate damage feature is an important step in the vibration based
damage identification approach. Many additional aspects (e.g. sensing system,
operational conditions, computational costs, etc.) should be taken into account in
this process. The selection is also significantly complicated by the wide range of
potential damage scenarios and the large variety of damage features. None of the
approaches is, at the moment, able to solve all problems in all structures. The
selection of an appropriate approach is, therefore, often a matter of compromise.

2.3 Literature overview vibration based methods


The amount of literature that is focused on vibration based damage identification
methods is immense. This is clearly illustrated by the number of available review
papers [3, 5–9]. One of the most recent overviews is given by Fan et al. [9], but
2.3. Literature overview vibration based methods 23

Optical Dynamics Electromagnetic

Visual Structural Electro- Acoustic Acousto Ultrasonic Eddy current


Technique vibrations & Mechanical
inspection Emission Ultrasonics Testing testing
acoustics Impedance

Time Frequency Modal


domain domain domain

Damage Correlation Frequency Transmissibility Natural Mode Mode shape


response Higher
feature functions functions harmonics frequencies shapes curvatures
functions

Frequency Frequency Mode shape Gapped Modal strain Curvature


(Statistical) assurance energy
response function curvature smoothing damage factor
classifier criterion damage index squared method method damage index method

Figure 2.2 A concise overview of the wide range of damage identification approaches. The vibration based
methods are categorized according to their damage sensitive feature and (statistical) classifier.

can also be found in the books of Boller et al. [10] and Balageas et al. [11]. More
specifically, Montalvão et al. [1] and Zou et al. [2] provide reviews that are dedicated
to vibration based health monitoring of composite materials. The review by Li [12]
is specifically focused on strain based methods. Reviews only concentrating on
nonlinear damage identification methods are presented by Worden et al. [13] and
Jhang [14].
The vibration based damage identification methods can be categorized according
to their damage sensitive feature and (statistical) classifier, as schematically shown
in Figure 2.2. The earliest studies on this subject report on the relation between
natural frequency shift or modal damping changes and structural damage. Adams
and Cawley [15–18] were among the first to actively research this subject back
in the seventies. These modal domain methods comprise a large part of the
available literature. Nowadays, research is however also directed towards statistical
time series methods [19] (e.g. cross-correlations and auto-regressive moving
average representations) and advanced signal processing techniques (e.g. Wavelet
transforms [20] and Hilbert-Huang transforms [21]) to extract damage features in
the time and frequency domain. A comprehensive list of the damage features that
are used in the literature is presented in Table 2.1. Appendix B provides an overview
of the wide variety of classifiers associated with these features.
The comparison and assessment of the performance of the different methods is
rather difficult. Each method can have individual advantages and disadvantages
regarding the specific application, structure and damage case under investigation.
The review studies primarily list the available literature, but do not allow for
a detailed comparison of the performance of each method. It lies beyond the
scope of this chapter to discuss each method individually. Several experimental
and numerical comparison studies by different authors can be found in the
24 Chapter 2. Overview of vibration based damage identification methods

Table 2.1 An overview of the vibration based damage features categorized according to the time, frequency
and modal domain. An elaborate overview of the classifiers associated with these features is presented in
Appendix B.

Time domain Frequency domain Modal domain


• Time response / • Fourier / power spectra • Natural frequencies
waveform • Frequency response • Mode shapes
• Statistical time series function • Mode shape curvatures
analysis: • Frequency response • Modal damping
◦ Correlation functions function curvature • Dynamic stiffness
(non-parametric) • Mechanical impedance • Dynamic flexibility
◦ Autoregressive models • Transmissibility function • Updating methods
(parametric) • Antiresonances
• Time-frequency responses1 • Higher harmonics
(nonlinear)
• Modulations (nonlinear)
1 Time-frequency parameters are sometimes considered as a separate domain in the literature.

literature [9, 12, 22–27]. Although these studies are limited to a selection of
methods, structures and damage cases, they provide a general perception of the
strengths and weaknesses of the evaluated methods. In summary, the comparative
studies concluded that mode shape curvatures (damage feature) combined with
the modal strain energy damage index (MSE-DI) algorithm [28, 29] (classifier)
outperforms other frequently applied damage identification methods based on
natural frequencies [16, 30], mode shapes [31, 32], dynamic flexibility [33–35] and
other mode shape derivative based classifiers [36, 37]. Mode shape curvatures tend
to be a relatively sensitive damage feature [9]. Humar et al. [25] stated that, among
the evaluated methods, the MSE-DI formulation is most successful in detecting
and localizing damage, but is lacking in the quantification of the damage. These
observations are supported by Alvandi and Cremona [24]. They also concluded
that this approach showed a good stability regarding noisy signals. Drawbacks of
using mode shape curvatures are the number of required data points to accurately
describe the mode shapes and the numerical errors introduced with the calculation
of the second derivative [9, 27, 37].

2.4 Damage feature selection

As explained in Section 2.2, many aspects need to be considered in the process of


finding an appropriate damage feature and classifier. The present section discusses
two basic principles to aid in this process. Firstly, the effect of damage on the
structural properties is addressed. The skin-stiffener structure considered in this
thesis is expected to be primarily prone to delamination type of damage at the skin-
2.4. Damage feature selection 25

stiffener connection. An analytical model of a composite beam with a delamination


is analyzed to gain a more general understanding of how delamination damage can
affect the dynamic behavior of a beam-like structure. Special attention is paid to the
feasibility of modal curvatures, since they tend to be a strong indicator of damage
according to the literature. Secondly, determination of modal domain parameters
are accompanied by data condensation and reduction. The consequences involved
with this process are explained and demonstrated based on a mass-spring-damper
system with a bilinear stiffness. A delamination damage can potentially open and
close under an appropriate loading and can introduce nonlinear effects. These
nonlinear effects set requirements to the features that can be used for damage
characterization purposes.

2.4.1 Effect of damage on dynamic properties


Damage in a composite laminate can affect the material parameters in a highly
directional nature. This directional behavior sets the requirements for a vibration
based damage identification approach in two ways. First of all, for the selection of
the type and orientation of the excitation. A high sensitivity is only obtained when
the damaged area is stressed in the direction that shows the most severe change
in the structural properties. For example, a delamination in a plate structure will
barely affect the inplane stiffness properties, but is expected to be more pronounced
in the flexural behavior. Secondly, for the selection of the location, orientation and
type of sensors. They should be selected in such a way that the damage information
can be extracted from the structural response.

Delaminated composite beam

Several dynamic models have been developed for the analysis of the effect of
delaminations on dynamic behavior of beam structures. To overcome difficulties
associated with potential nonlinear dynamic effects introduced by a delamination,
many modeling attempts assume that the dynamic behavior of the structure
remains linear, i.e. the effect of the damage on the material parameters (e.g.
stiffness) is independent of the deformation. Della and Shu [38] categorize the linear
dynamic models of delaminated structures into two classes: the ‘free mode’ and the
‘constraint mode’ models. The ‘free mode’ models assume that the delaminated
layers deformed ‘freely’ without interacting with each other. On the other hand
the ‘constraint mode’ models assume that the delaminated layers are ‘constrained’
and will have identical transverse deformations. In addition, the delaminated
layers are assumed to be free to slide over each other in the axial direction. Both
representations show discrepancies with respect to what is expected in practice.
Nevertheless, the predicted frequencies based on the ’constrained model’ were
closer to the experimental values [39].
26 Chapter 2. Overview of vibration based damage identification methods

Clamped Integral section Delaminated section Integral section


side

1 2 4
3 Free side

a−c

V2, M2, N2 V2, M2, N2


V1, M1 x2 z2 V4, M4

x1 h3 x4
z1 − z4
x3 h
z3 −
h2
V3, M3, N3 V3, M3, N3
−a − −c
b
Figure 2.3 An analytical model of a beam with a single through-width delamination. The Vs represents the
shear force, Ms the bending moment and Ns the normal force for beam segment s = 1, 2, 3, 4. The dimensions
of the beam are indicated by a, b, c and height h. The delamination location is given by ac and h2 .

The first ‘constraint mode’ model was presented by Mujumdar and Suryanarayan [39].
This model determines the natural frequencies and mode shapes of an isotropic
beam containing a single through-width delamination in flexural vibration. The
delaminated beam model consists of three spanwise regions, a delaminated section
and two integral sections (see Figure 2.3). The delaminated section is composed of
two separate beams above and below the delamination plane and they are joined at
their ends to the integral beam segment. It is assumed that no gap exists between
the delaminated components and that they are free to slide, i.e. there is no friction
between the components. Each beam is treated as a Bernoulli-Euler beam. The
equations of motion are formulated for each beam segment. The dynamic solution
for the complete beam is obtained by imposing appropriate boundary conditions at
the ends of the two integral segments and continuity conditions at the delamination
junctions.
The basic theory is based on variations in the distribution of shear stress. Both
normal and shear stresses are developed in a beam that exhibits non-uniform
bending. In a beam without a delamination, the shear stress is distributed as
shown in Figure 2.4a. However, in case a delamination is present, the layers are
free to slide and the shear stress on the delaminated surface therefore equals zero
by definition. The shear stress distribution for this case is shown in Figure 2.4b.
This discontinuity in shear stress results in a decrease in the flexural rigidity of the
delaminated segment, which will affect the dynamic properties of the beam.
The two separate beams in the delaminated segment also give rise to coupling
between the axial and flexural deformation. Therefore, continuity of axial forces
2.4. Damage feature selection 27

τ τ N (a) N
N N
τ
(a) (b) (b)
Figure 2.4 Distribution of shear Figure 2.5 Delaminated layers without (a) and with (b)
stress τ in (a) a pristine and (b) a continuity conditions at the junctions between the delaminated
delaminated beam segment. and integral sections.

and displacements at the junctions between the integral and delaminated segments
is required. A normal force N is introduced, as indicated in Figure 2.5. These
normal forces compress one segment and stretch the other such that their ends lie
in the same plane. This internal loading results in an internal bending moment,
which partly compensates for the loss of flexural rigidity due to the new shear
stress distribution.
Grouve et al. [40] extended the analytical model of Mujumdar and Suryanarayan [39]
for laminated beams. Experimental validation showed an averaged difference
between theoretical and experimental resonance frequencies for the delaminated
beams equal to 3.3%. The differences between experimental and calculated
resonance are partly caused by the model assumptions made. The delaminated
segments are modeled to stay in contact during the vibration and slide without
friction. This is a requirement which cannot be satisfied in practice. Opening of the
delamination during vibration will result in a decrease of stiffness, while friction
between the segments will increase the flexural rigidity.

Analytical results

The model proposed by Grouve et al. [40] is used to analyze the effect of a
delamination on the dynamic behavior. A cantilever composite beam structure
(dimensions 300×20×2.0 mm) built from 16 unidirectional plies of carbon fiber
AS4D is considered in the analysis. The thermoplastic matrix is PEKK. The laminate
has a [0/90]4,s lay-up with the material properties as in Table 2.2. The dimensionless
delamination parameters (ac , b and h2 ) as defined in Figure 2.3 are used in the
investigation.

Table 2.2 The material properties of a unidirectional layer of carbon AS4D/PEKK. The material properties
are obtained from data provided by Fokker Aerostructures B.V., Hoogeveen, The Netherlands.

E1 E2 G12 ν12 ρ
[GPa] [GPa] [GPa] [-] [kg/m3 ]
133 10.5 5.56 0.3 1590
28 Chapter 2. Overview of vibration based damage identification methods

The effect of the delamination on the dynamic behavior of the composite beam
is a combination of two effects: one is the discontinuity in shear stress over the
beam thickness, the other is the stretching and compressing of the delaminated
beam segments. The first effect results in a reduction of flexural rigidity, while the
second effect tends to compensate for this. The weakening effect of a delamination
is only noticed in case a shear force is acting in the beam. This means that the
effect of a delamination vanishes if the delamination is located in regions where
the shear force approaches zero, but is the largest at locations where the shear
force is maximum. The non-uniform distribution of shear stress across the length
of the beam causes the effect to be dependent on the axial location ac of the
delamination. As the shear stress distribution differs for every vibration mode,
the delamination will also affect each mode differently. These effects are clearly
demonstrated in Figure 2.6, showing the second and third vibration mode shape
and its derivatives for a through-width delamination (ac = 0.5, b = 0.2, h2 = 0.5) in a
composite beam. Note that the shear force is a function of the third derivative of the
displacement. The delamination is located near a point where the third derivative
of the second mode approaches zero, but is large for the third mode. Consequently,
the delamination has a negligible effect on the second mode, while the third mode
is clearly affected. Figure 2.7 shows that a similar but opposite effect on the second
and third mode is observed when the delamination moves to the axial location
ac = 0.3.
The effect of the local delamination is more pronounced in the mode shape
curvatures than observed in the mode shapes themselves. The delamination
causes clear discontinuities in the mode shape curvature at the boundaries of the
delaminated segment. These discontinuities are different for each vibration mode
and reveal the presence, the location, as well as the size of the delamination.
The size of the discontinuity is described by the continuity equation for bending
moments at the junction between the integral and delaminated segment [39]. The
discontinuities are the largest when the delamination is located at a nodal point in
the mode shape and at the neutral plane. They vanish for locations near the surface
and at an antinode. The ’constrained mode’ model, however, does not incorporate
the potential opening of the delaminated segment in the results. The opening of the
delamination allows for additional curvature changes for delamination locations
where the shear related curvature effects are minimal.
Besides the literature studies discussed in Section 2.3, the analytical results also
showed that modal curvatures are an appropriate damage feature in the case of
delamination damage. The discontinuities caused at the delamination boundaries
are significantly more pronounced than, for example, deviations in mode shapes.
This makes curvatures an interesting candidate for damage identification. The
direct relation between curvature and strain for a beam in bending also opens
possibilities for integrated sensing technologies based on, for example, optical fiber
Bragg gratings [40] or piezoelectric strain sensors [41].
2.4. Damage feature selection 29

x delamination x delamination
z z

1 1
Mode 2 Mode 2
Norm. u(z) [−]

Norm. u(z) [−]


0 0

Mode 3
Mode 3
−1 −1
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Position z/L [−] Position z/L [−]
(b) Mode shape. (b) Mode shape.

1 1
Norm. du/dz [−]

Norm. du/dz [−]


Mode 2
Mode 2
0 0
Mode 3
−1 Mode 3
−1
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Position z/L [−] Position z/L [−]
(c) Mode shape, first derivative. (c) Mode shape, first derivative.
Norm. d u/dz [−]

Norm. d u/dz [−]

2 Mode 3
1
2

Mode 2
0 0
2

Mode 3 −1
−2 Mode 2
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Position z/L [−] Position z/L [−]
(d) Mode shape, second derivative. (d) Mode shape, second derivative.
4
Norm. d u/dz [−]

Norm. d u/dz [−]

2
3

Mode 2
2 0
Mode 3
3

0 −2 Mode 2 Mode 3

−2 −4
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Position z/L [−] Position z/L [−]
(e) Mode shape, third derivative. (e) Mode shape, third derivative.

Figure 2.6 The second and the third vibration Figure 2.7 The second and the third vibration
mode shape u(z) and the derivatives for a pristine mode shape u(z) and the derivatives for a pristine
(blue) and delaminated (red) composite beam (blue) and delaminated (red) composite beam
(ac = 0.5, b = 0.2, h2 = 0.5). (ac = 0.3, b = 0.2, h2 = 0.5).
30 Chapter 2. Overview of vibration based damage identification methods

Information
condensation
Modal domain Y(ωn)
Damage
temperature Frequency domain Y(ω) features
T
Time domain y(t)
displacement impedance
x Z
damping Damage parameter χd
Δζ
stiffness mass
force Δm
Δk
F damage
scenario strain
non- thermal
ε
voltage linearities ΔT velocity
U dimensions dx/dt
l
acceleration electrical
2
d x/dt
2 resistance
R

Figure 2.8 Damage features classified according to the amount of information condensation. The outer
circles represent a higher level of information condensation.

2.4.2 Information condensation

A second aspect that needs to be considered in the selection of a damage sensitive


parameter is the reduction and condensation of the measured information. The
challenge is to condense the damage related information to a lower dimensional
parameter space (e.g. modal domain) to obtain a robust indicator, while preserving
the correlation with the damage scenario. Figure 2.8 schematically illustrates the
condensation process. The damage parameter χd provides the best description
of the damage, but is often not directly measurable. The system response
parameters y(t) (e.g. strain, acceleration) are, therefore, acquired. These time
domain representations are related to the damage parameter χd , but often provide
an incomplete system description caused by a limited number of measurement
points. Moreover, the time series consists of large data sets, which makes it hard
to attribute measured deviations to the structural changes caused by the damage.
Consequently, the time series is often transferred to the frequency domain. The
frequency domain parameters (e.g. power spectral densities, frequency response
functions) are typically obtained under the assumption that the responses are
periodic. Depending on the time scale of the evolving damage, the system is also
assumed to be invariant during the measurement. Modal analysis will further
condense the damage information to a linear system description based on the
natural frequencies, modal damping factors and modal vectors. Subsequently,
slightly complex modal vectors are generally considered to be real normal, by
neglecting non-proportional damping effects. Moreover, the modal vectors are
2.4. Damage feature selection 31

Fspring(t)
αk x(t) αk

Ftest(t)
m
c x(t)
k
(1-α)k

Figure 2.9 A simplified representation of the opening and closing behavior of damage by a mass-spring-
damper system with a bilinear stiffness subjected to a dynamic loading Ftest (t). The force exerted by the
spring is given by: Fspring (t) = k( x ) x (t).

often scaled vectors without a physical quantity. Finally, for damage identification
purposes typically a selection of the most sensitive modes is considered to obtain
an indicator. The compressed subset of parameters must be able to describe the
damage scenario. Information condensation can result in a robust indicator that has
a strong correlation with the damage scenario. Each transformation is, however,
also accompanied by the loss of potentially valuable information caused by the
intermediate assumptions. These assumptions can inherently limit the applicability
of the selected damage features depending on the desired performance level, as
defined in Section 1.3.2.

Bilinear system representation

The consequences involved with the information condensation process for modal
parameters can easily be demonstrated by considering a mass-spring-damper
system. Damage can open and close under an appropriate dynamic loading. The
dynamic effects induced by the opening and closing motion are represented in a
simplified way by a mass-spring-damper system with a bilinear stiffness, depicted
in Figure 2.9. The governing dynamic equation reads:

m ẍ (t) + c ẋ (t) + k ( x ) x (t) = Ftest (t) , (2.7)

with:
(
αk x≥0
k( x) = (2.8)
k x<0

and where m represents the mass, c the damping, k ( x ) the stiffness, x (t) the
time-dependent displacement and Ftest (t) the (harmonic) force. The parameter α
describes the extent of the bilinearity. The system properties are chosen arbitrary
and equal: m = 1 kg, k = 105 N m-1 , c = 5 Ns m-1 , α = 0.5. The time response
32 Chapter 2. Overview of vibration based damage identification methods

−5
x 10
α=0.5 (bilinear) f 2f α=0.5 (bilinear)
2 a a
0
α=1.0 (linear) 10 α=1.0 (linear)
1

|X| [m]
3f
a 4f
x [m]

0 −2
a 5f
10 a
−1
−4
−2 10
10 10.1 10.2 10.3 10.4 0 50 100 150
Time [s] Frequency [Hz]
(a) Time responses x (t). (b) Magnitude of the Fourier spectra | X ( f )|.
−4
x 10
5
α=0.5 (bilinear) α=0.5 (bilinear)
α=1.0 (linear) 10
0
α=1.0 (linear)
|X| [m]
x [m]

0
−2
10

−5 10
−4
0 5 10 15 20 0 50 100 150
Time [s] Frequency [Hz]
(c) Time responses x (t). (d) Magnitude of the Fourier spectra | X ( f )|.

Figure 2.10 The time responses and the magnitudes of their Fourier transformed signals obtained for a (a,
b) single tone ( f a = 25 Hz) and a (c, d) linear swept harmonic ( f a = 1–150 Hz) excitation of a mass-spring-
damper system with a linear (α = 1.0) and a bilinear (α = 0.5) stiffness.

x (t) for a single tone harmonic excitation at frequency f a = 25 Hz is presented


in Figure 2.10(a) for a linear (α = 1.0) and bilinear (α = 0.5) system. The bilinear
stiffness affects the steady state time response. Their windowed Fourier spectra
X ( f ) are given in Figure 2.10(b). The spectrum for the system with the bilinear
stiffness reveals clear higher harmonic components. These higher harmonics are
integer multiples of the excitation frequency (n f a with n = 1, 2, 3, . . . ) and are
characteristic indicators of a nonlinear system response [42].
Dynamic analysis usually relies on broadband excitation signals to characterize a
dynamic system. The time and frequency domain responses for a linearly swept
harmonic excitation ( f a = 1–150 Hz) are presented in Figures 2.10(c) and 2.10(d).
The highest response amplitudes are obtained at the natural frequencies. The
bilinear system shows, besides a shift in natural frequency, clear scattering in the
frequency response compared to the smooth linear system response. The swept
excitation causes a different set of higher harmonics to be generated for each time
instance. Each set of harmonics has a different participation in the total response.
The scattering is consequently the result of calculating the Fourier transform over
2.4. Damage feature selection 33

the total response signal. The transformation of the time domain information to the
frequency domain significantly complicates, in this case, the interpretation of the
nonlinear response.

Transformation to the modal domain will further condense the system information.
Modal analysis essentially assumes a time invariant and a linear system. Several
modal analysis methods are available in the literature [43]. The method used in the
present work extracts the modal parameters by applying an experimental modal
analysis (EMA) procedure to the frequency response functions (FRFs) between the
excitation and the response location. These frequency response functions HFx (ω )
can be calculated in different ways. By assuming that mainly the response signal is
subject to noise, a least square estimate is given by [44]:

S Fx (ω )
HFx (ω ) = , (2.9)
S FF (ω )

with S FF (ω ) the auto-power spectral density of the excitation signal and S Fx (ω ) the
cross-power spectral density between the excitation and response signal.

When the frequency response functions have been calculated, the modal parameters
can be estimated. The experimental modal analysis procedure used in the present
work consists of two steps. In the first step, the number of modes, the damped
natural frequencies and the associated damping factors are determined from
the complete set of frequency response functions with the help of a singular
value decomposition. A detailed description of this process is presented in [45].
The second step concerns the estimation of the modal parameters by curve
fitting a mathematical modal-parameter model to all measured frequency response
functions. The modal model in partial fraction form for a multi-degree of freedom
system is given by [46, 47]:

Nmodes  
Rn R∗n
H (ω ) = ∑ + , (2.10)
n =1
jω − λn jω − λ∗n

where the frequency response function matrix H (ω ) is described in terms of a sum


of individual modes. The number of modes Nmodes is equal to one for the mass-
spring-damper system. The ∗ denotes the complex conjugate. During the curve
fitting step, the numerator and denominator of Equation (2.10) are written in a
polynomial form and are estimated by a least squared error fitting process, referred
to as the global polynomial method [48]. If these polynomials fit to the measured
frequency response functions then the modal parameters can be calculated. Each
mode n is represented by modal residues Rn and poles λn . The poles appear in
complex conjugate pairs given by:
q
λn , λ∗n = −σn ± jωdn = −ζ n ωn ± jωn 1 − ζ n2 , (2.11)
34 Chapter 2. Overview of vibration based damage identification methods

−60
α=0.5
Modal fit

|H | [dB]
−80
α=1.0
Modal fit

Fx
−100

−120

0 50 100 150
Frequency [Hz]

Figure 2.11 The curve fitted modal model and the magnitude of the original frequency response functions
between the harmonic sweep ( f a = 1–150 Hz) excitation and the displacement response of a mass-spring-
damper system with a linear (α = 1.0) and a bilinear (α = 0.5) stiffness.

where σn is the damping factor, ωdn is the damped natural frequency, ωn is the
undamped natural frequency and ζ n the relative damping of mode n. The residue
matrix Rn describes the mode strength and is related to the mode shape vector [49]:

Rn = Qn ψn ψnT , (2.12)

where Qn is a scaling constant and ψn is the mode shape vector for the nth mode.
The curve fitted mathematical model and the original frequency response function
for the linear (α = 1.0) and bilinear (α = 0.5) mass-spring-damper system are
presented in Figure 2.11. The frequency responses are reduced to a set of modal
parameters as shown in Table 2.3. The modal vectors are not presented here
because a single degree of freedom system is considered. An exact description
is obtained for the linear system. The modal representation of the bilinear system,
however, largely ignores the nonlinear effects and only provides a linear description
of the nonlinear system. The estimated natural frequency corresponds to a linear
system with a stiffness of k = 0.68 · 105 N m-1 (32% degradation). This stiffness is
essentially the result of a variable stiffness (k = 1.0 · 105 and αk = 0.5 · 105 N m-1 )
within each oscillation and therefore provides an incorrect characterization of the
damage. These results show that the modal parameter changes can be used to
indicate the presence of structural changes, but that the information is reduced

Table 2.3 Modal parameters estimated by curve fitting the modal model to the frequency response functions
of the linear (α = 1.0) and the nonlinear (α = 0.5) system.

Case Modal model Analytical


f n [Hz] ζ n [%] f n [Hz]
Linear (α = 1.0) 50.33 0.79 50.33
Bilinear (α = 0.5) 41.46 0.96 -
2.5. Concluding remarks 35

beyond the point where it allows for an adequate description of the damaged
system. This is primarily the case when damage causes severe nonlinear dynamic
effects. An adequate description is obtained if weakly nonlinear dynamic effects are
present.

2.5 Concluding remarks


An overview of vibration based structural health monitoring techniques is
presented in this chapter. A generalized description of a damaged system is used
to explain the basic concept of this technology. Damage identification aims to
maximize the correlation between the damage and the damage sensitive feature,
while it is also desired to condense the measured responses to a low dimensional
parameter space. Both aspects are considered in the selection of an appropriate
damage identification method for the skin-stiffener structure as considered in this
thesis.
The literature study in Section 2.3 revealed that mode shape curvatures combined
with the modal strain energy damage index algorithm are a potentially powerful
damage feature and classifier for damage identification. A basic analytical model of
a delaminated composite beam revealed that mode shape curvatures can show clear
discontinuities at the boundaries of the delaminated region in the case of flexural
vibrations. These discontinuities cause mode shape curvatures to be a strong
indicator for the presence, location and size of the delaminated region compared
to, for example, deviations in the mode shapes themselves. However, when severe
nonlinear dynamic effects are introduced by the damage, the modal domain based
methods suffer under their linear system assumption. Alternative methods in, for
example, the time or frequency domain are in that case required to obtain a more
adequate description of the nonlinear dynamic behavior introduced by the damage.

References
[1] D. Montalvão, N.M.M. Maia, and A.M.R. Ribeiro. A review of vibration-based
structural health monitoring with special emphasis on composite materials. Shock and
Vibration Digest, 38(4):295–326, 2006.
[2] Y. Zou, L. Tong, and G.P. Steven. Vibration-based model-dependent damage
(delamination) identification and health monitoring for composite structures – A
review. Journal of Sound and Vibration, 230(2):357–378, 2000.
[3] C.-P. Fritzen and P. Kraemer. Self-diagnosis of smart structures based on dynamical
properties. Mechanical Systems and Signal Processing, 23(6):1830–1845, 2009.
[4] C.R. Farrar, H. Sohn, F.M. Hemez, M.C. Anderson, M.T. Bement, P.J. Cornwell, S.W.
Doebling, J.F. Schultze, N. Lieven, and A.N. Robertson. Damage prognosis: current
status and future needs. Technical report, Los Alamos National Laboratory, Los
Alamos, NM, 2001.
36 Chapter 2. Overview of vibration based damage identification methods

[5] S.W. Doebling, C.R. Farrar, M.B. Prime, and D.W. Shevitz. Damage identification and
health monitoring of structural and mechanical systems from changes in their
vibration characteristics: a literature review. Technical report, Los Alamos National
Laboratory, NM, USA, 1996.
[6] S.W. Doebling, C.R. Farrar, and M.B. Prime. A summary review of vibration-based
damage identification methods. The Shock and Vibration Digest, 30(2):91–105, 1998.
[7] H. Sohn, C.R. Farrar, F.M. Hemez, D.D. Shunk, D. Stinemans, and B.R. Nadler. A
review of structural health monitoring literature: 1996-2001. Technical report, Los
Alamos National Laboratory, Los Alamos, NM, 2003.
[8] E.P. Carden and P. Fanning. Vibration based condition monitoring: a review. Structural
Health Monitoring, 3(4):355–377, 2004.
[9] W. Fan and P. Qiao. Vibration-based damage identification methods: a review and
comparative study. Structural Health Monitoring, 10(1):83–111, 2011.
[10] C. Boller, F.-K. Chang, and Y. Fujino. Encyclopedia of structural health monitoring. John
Wiley & Sons, Ltd, Chichester, UK, 2009.
[11] D. Balageas, C.-P. Fritzen, and A. Güemes. Structural health monitoring. ISTE Ltd, 2006.
[12] Y.Y. Li. Hypersensitivity of strain-based indicators for structural damage identification:
a review. Mechanical Systems and Signal Processing, 24(3):653–664, 2010.
[13] K. Worden, C.R. Farrar, J. Haywood, and M. Todd. A review of nonlinear dynamics
applications to structural health monitoring. Structural Control and Health Monitoring,
15:540–567, 2008.
[14] K.-Y. Jhang. Nonlinear ultrasonic techniques for nondestructive assessment of micro
damage in material: a review. Precision Engineering and Manufacturing, 10(1):123–135,
2009.
[15] R.D. Adams, D. Walton, J.E. Flitcroft, and D. Short. Vibration testing as a
nondestructive test tool for composite materials. ASTM Special Technical Publication 580,
Composite Reliability, pages 159–175, 1975.
[16] P. Cawley and R.D. Adams. The location of defects in structures from measurements of
natural frequencies. Journal of Strain Analysis, 14(2):49–57, 1979.
[17] P. Cawley and R.D. Adams. A vibration technique for non-destructive testing of fibre
composite structures. Journal of Composite Materials, 13(2):161–175, 1979.
[18] P. Cawley and R.D. Adams. Improved frequency resolution from transient tests with
short record lengths. Journal of Sound and Vibration, 64(1):123–132, 1979.
[19] S.D. Fassois and J.S. Sakellariou. Time-series methods for fault detection and
identification in vibrating structures. Philosophical Transactions of the Royal Society A:
Mathematical, Physical and Engineering Sciences, 365(1851):411–48, 2007.
[20] M. Rucka and K. Wilde. Application of continuous wavelet transform in vibration
based damage detection method for beams and plates. Journal of Sound and Vibration,
297(3-5):536–550, 2006.
[21] H.F. Hu, W.J. Staszewski, N.Q. Hu, R.B. Jenal, and G.J. Qin. Crack detection using
nonlinear acoustics and piezoceramic transducers – Instantaneous amplitude and
frequency analysis. Smart Materials and Structures, 19(6):065017, 2010.
[22] C.R. Farrar and D.A. Jauregui. Comparative study of damage identification algorithms
applied to a bridge: I. Experiment. Smart Materials and Structures, 7:704–719, 1998.
[23] C.R. Farrar and D.A. Jauregui. Comparative study of damage identification algorithms
applied to a bridge: II. Numerical study. Smart Materials and Structures, 7:720–731, 1998.
[24] A. Alvandi and C. Cremona. Assessment of vibration-based damage identification
References 37

techniques. Journal of Sound and Vibration, 292(1-2):179–202, 2006.


[25] J. Humar, A. Bagchi, and H. Xu. Performance of vibration-based techniques for the
identification of structural damage. Structural Health Monitoring, 5(3):215–241, 2006.
[26] J.-M. Ndambi, J. Vantomme, and K. Harri. Damage assessment in reinforced concrete
beams using eigenfrequencies and mode shape derivatives. Engineering Structures,
24(4):501–515, 2002.
[27] A.P. Adewuyi, Z. Wu, and N.H.M. Kammrujaman Serker. Assessment of
vibration-based damage identification methods using displacement and distributed
strain measurements. Structural Health Monitoring, 8(6):443–461, 2009.
[28] N. Stubbs and C.R. Farrar. Field verification of a nondestructive damage localization
and severity estimation algorithm. In Proceedings of the 13th International Modal Analysis
Conference (IMAC XIII), pages 210–218, 1995.
[29] P.J. Cornwell, S.W. Doebling, and C.R. Farrar. Application of the strain energy damage
detection method to plate-like structures. Journal of Sound and Vibration, 224(2):359–374,
1999.
[30] O.S. Salawu. Detection of structural damage through changes in frequency: a review.
Engineering Structures, 19(9):718–723, 1997.
[31] E. Parloo, P. Guillaume, and M. van Overmeire. Damage assessment using mode shape
sensitivities. Mechanical Systems and Signal Processing, 17(3):499–518, 2003.
[32] L.M. Khoo, P.R. Mantena, and P. Jadhav. Structural damage assessment using vibration
modal analysis. Structural Health Monitoring, 3(2):177–194, 2004.
[33] A.K. Pandey and M. Biswas. Damage detection in structures using changes in
flexibility. Journal of Sound and Vibration, 169(1):3–17, 1994.
[34] T. Toksoy and A.E. Aktan. Bridge-condition assessment by modal flexibility.
Experimental Mechanics, 34(3):271–278, 1994.
[35] F.N. Catbas, M. Gul, and J.L. Burkett. Damage assessment using flexibility and
flexibility-based curvature for structural health monitoring. Smart Materials and
Structures, 17(1):1–12, 2008.
[36] A.K. Pandey, M. Biswas, and M.M. Samman. Damage detection from changes in
curvature mode shapes. Journal of Sound and Vibration, 145(2):321–332, 1991.
[37] M.M. Abdel Wahab and G. De Roeck. Damage detection in bridges using modal
curvatures: application to a real damage scenario. Journal of Sound and Vibration,
226(2):217–235, 1999.
[38] C.N. Della and D. Shu. Vibration of delaminated composite laminates: a review.
Applied Mechanics Reviews, 60(1-6):1–20, 2007.
[39] P.M. Mujumdar and S. Suryanarayan. Flexural vibrations of beams with delaminations.
Journal of Sound and Vibration, 125(3):441–461, 1988.
[40] W.J.B. Grouve, L.L. Warnet, A. de Boer, R. Akkerman, and J. Vlekken. Delamination
detection with fibre Bragg gratings based on dynamic behaviour. Composites Science
and Technology, 68(12):2418–2424, 2008.
[41] W. Lestari, P. Qiao, and S. Hanagud. Curvature mode shape-based damage assessment
of carbon/epoxy composite beams. Journal of Intelligent Material Systems and Structures,
18(3):189–208, 2007.
[42] N. Krohn, R. Stoessel, and G. Busse. Acoustic non-linearity for defect selective
imaging. Ultrasonics, 40(1-8):633–7, 2002.
[43] Á. Cunha and E. Caetano. Experimental modal analysis of civil engineering structures.
38 Chapter 2. Overview of vibration based damage identification methods

Sound and Vibration, pages 12–20, 2006.


[44] B.J. Schwarz and M.H. Richardson. Experimental modal analysis. In CSI Reliability
Week, Orlando, FL, pages 1–12, 1999.
[45] R.J. Allemang and D.L. Brown. A complete review of the complex mode indicator
function (CMIF) with applications. In Proceedings of International Conference on Noise and
Vibration Engineering (ISMA2006), pages 3209–3246. Leuven, 2006.
[46] M.H. Richardson and D.L. Formenti. Parameter estimation from frequency response
measurements using rational fraction polynomials. In Proceedings of the 1st International
Modal Analysis Conference (IMAC I), pages 1–15, 1982.
[47] M.H. Richardson and B.J. Schwarz. Modal parameter estimation from operating data.
Sound and Vibration, 37(1):28–39, 2003.
[48] M.H. Richardson and D.L. Formenti. Global curve fitting of frequency response
measurements using the rational fraction polynomial method. In 3rd International
Modal Analysis Conference, pages 1–8, 1985.
[49] B. Schwarz and M. Richardson. Scaling mode shapes obtained from operating data.
Sound and Vibration, 37(11):18–23, 2003.
Chapter 3

Vibration based structural health


monitoring of a composite T-beam1

Abstract
A vibration based damage identification method is investigated
experimentally for a 2.5-dimensional composite structure. The
dynamic response of an intact and a locally delaminated 16-layer
unidirectional carbon fiber PEKK reinforced T-beam is considered.
A force-vibration set-up, including a laser vibrometer system, is
employed to measure the dynamic behavior of the T-beam. The
modal strain energy damage index (MSE-DI) algorithm is applied
using the bending and torsion modes. Special attention is paid to the
effect of the number of measurement points required to detect and
localize the delamination accurately.

1 Reproduced from: T.H. Ooijevaar, R. Loendersloot, L.L. Warnet, A. Boer, R. Akkerman. Vibration
based structural health monitoring of a composite T-beam. Composite Structures, 92(9):2007-2015,
2010.

39
40 Chapter 3. Vibration based structural health monitoring of a composite T-beam

3.1 Introduction
Development of structural health monitoring (SHM) technologies for composite
based structural components for aircraft is one of the objectives of the European
research program Clean Sky, Eco-Design ITD. An increase of the part’s service life
reduces its cost and long-term ecological impact.
One of the key issues in composite structures is the early detection and localization
of damage. Often service induced damage does not involve visible plastic
deformation, but internal matrix related damage, like transverse cracks and
delaminations. Their detection imposes costly maintenance techniques. Vibration
based damage identification methods are promising as an alternative for the time-
consuming and costly nondestructive testing (NDT) methods currently available.
These methods also offer the potential to be used as a real-time health monitoring
system. The change of the dynamic properties is employed to identify damage such
as delaminations.
Grouve et al. [1] used a vibration based structural health monitoring technique
by monitoring shifts in natural frequencies to detect a delamination in a simple
composite beam structure. However, it is often complicated to localize damage
uniquely. Stubbs et al. [2] proposed a technique which is capable of detecting and
localizing damage in beam-like one-dimensional structures. They were the first
to introduce a damage identification method based on the observation that local
changes in the modal strain energy of the vibration modes of a structure are a
sensitive indicator of damage. Cornwell et al. [3] extended the method for plate-like,
hence two-dimensional, structures. Only recently, the step to three-dimensional
frame structures was made [4, 5].
Until now, the modal strain energy damage index (MSE-DI) algorithm is hardly
used on anisotropic, composite structures. Moreover, this methodology is mainly
applied to one-dimensional beam and two-dimensional plate-like structures with
only flexural vibrations. The question is whether the methodology is applicable for
more complex and larger structures with a minimum amount of sensors. A step in
this direction is to consider a 2.5-dimensional composite beam structure with a more
general deformation pattern, including bending and torsion, in order to extract
a maximum of damage related information from the measured vibration modes.
Duffey et al. [6] are the only ones who included torsion modes and vibrations
in their research. Torsion vibrations can become rather complex and tend to be
more difficult to measure. However, beside bending modes, torsion modes could
also provide damage related information. This could become advantageous, in
particular for larger structures, if the number of measurable vibration modes is
limited and the number of sensors is to be reduced. Therefore it is also important
to investigate the effect of the number of measurement points on the damage
identification results.
This chapter mainly presents the experimental work performed at the University
3.2. The modal strain energy damage index algorithm 41

of Twente. It reports on the application of the modal strain energy damage


index algorithm, applied to a 2.5-dimensional composite structure vibrating under
bending and torsion modes, in order to detect and localize a delamination. An
experimental set-up around an intact and delaminated T-beam is used for this
purpose. Use is made of an external measurement device for the characterization
of the dynamic behavior, in order to easily vary the amount of measurement points.
The results obtained in this study will set requirements on the methods and the
instrumentation necessary to apply the method in a structural health monitoring
environment.

3.2 The modal strain energy damage index algorithm

The strain energy of a vibration mode is referred to as the modal strain energy of
that mode. Consequently, the total modal strain energy is the sum of the modal
strain energy contributions of all modes considered. The modal strain energy is
calculated by linking the deformation of a structure to the strain. A distinction can
be made between axial, flexural and torsional deformation-strain relations [6]. Only
the bending and the torsional deformations are considered in the present work. The
mechanical relations read (see Figure 3.1 for the coordinate system used):

∂2 u x/y My/x ∂θ xy Tz
= , = , (3.1)
∂z2 EIy/x ∂z GJxy

with u the displacement, θ the rotation, M the bending moment, T the torque, EI
and GJ respectively the bending and torsion rigidity of the beam. The subscript
‘(x/y)’ refers either to x or y. The rotation angle for small rotations is defined
as: θ xy (z) = ∂uy ( x, z) ∂x. The strain energies for the bending and torsional
deformations are obtained by integrating the squared strains over the length l of
the structure [6]:

Z l
!2 Z l  2
1 ∂2 u x/y 1 ∂θ xy
UB = EIy/x dz , UT = GJxy dz . (3.2)
2 0 ∂z2 2 0 ∂z

Consider the structure to vibrate in the nth bending mode. The modal strain energy
of this bending mode is given by:
 (n)
2
(n) 1
Z l ∂2 u x/y
UB = EIy/x   dz , (3.3)
2 0 ∂z2
42 Chapter 3. Vibration based structural health monitoring of a composite T-beam

(n)
with u x/y (z) the normalized amplitude of the mode shape. Subsequently, the
structure is discretized in N elements in axial direction (i.e. z-coordinate). The
(n)
strain energy UB,i for the nth mode and associated with the ith element is then
given by:
 (n)
2
(n)
Z
1 zi ∂2 u x/y (n)
N
(n)
UB,i = EIy/x   dz with: UB = ∑ UB,i , (3.4)
2 z i −1 ∂z2 i =1

and zi−1 , zi the boundaries of element i of the discretized structure in z-direction.


Similar quantities can be defined for a damaged structure, using the normalized
mode shapes ũ(n) of the damaged structure. The local fractional strain energies, as
defined by Cornwell et al. [3], are:

(n) (n)
(n) UB,i (n) ŨB,i
FB,i = (n)
, F̃B,i = (n)
, (3.5)
UB ŨB

for the intact and damaged structure respectively. The fractional strain energy
remains relatively constant in the undamaged elements, under the assumption that
the damage is small and primarily located at element k [3, 7]:

(n) (n)
F̃B,i ≈ FB,i for i 6= k . (3.6)

This equation is rearranged and after substituting Equations (3.3), (3.4) and (3.5) it
can be derived that:
R zi  . 2 . R  . 2
f y/x ∂2 ũ(n) ∂z2 dz
EI
lf
EI y/x ∂ 2 (n)
ũ ∂z 2
dz
z i −1 x/y 0 x/y
1≈ R  . 2 . R  . 2 for i 6= k ,
zi 2 (n) 2 l 2 (n) 2
z EI y/x ∂ u x/y ∂z dz 0 EIy/x ∂ u x/y ∂z dz
i −1
(3.7)

The mean value theorem for integrals [8] is applied to the flexural rigidities EIy/x (z)
f y/x (z) in Equation (3.7) to obtain an average rigidity estimate for each
and EI
integral. For the integral over element i of the pristine case:
Z z  . 2 Z z  . 2
i (n) i (n)
EIy/x (z) ∂2 u x/y ∂z 2
dz = EIy/x (z0 ) ∂2 u x/y ∂z2 dz , (3.8)
z i −1 z i −1

where z0 is a point within the integration interval and EIy/x (z0 ) = ( EIy/x )i is the
weighted average flexural rigidity of element i. Similar estimates can be obtained for
the other integrals in Equation (3.7). Subsequently, a ratio of the flexural rigidities
is obtained by assuming that the change in the average value of the flexural rigidity
3.3. Composite T-beam structure 43

for the integrals over the entire length l of the beam is negligible for small damage:

R zi  (n)
. 2 .R 
(n)
. 2
(n)
.
∂2 ũ x/y ∂z2 ∂2 ũ x/y ∂z2
l
( EIy/x )i z i −1 dz 0 dz γ̃i γ̃(n)
≈ R  . 2 .R  . 2 = . for i 6= k ,
f y/x )i
( EI (n) (n) (n)
∂2 u x/y ∂z2 ∂2 u x/y ∂z2
zi l γi γ(n)
z i −1 dz 0 dz
(3.9)
(n)
where γi is the integral over element i and γ(n) the integral over the entire length
l. Stubbs et al. [2] and Cornwell et al. [3] used the right-hand side of Equation (3.9)
(n)
to define a local damage index β i . This indicator is approximately equal to 1
for the undamaged elements, i.e. i 6= k, but can change at the damaged element,
i.e. i = k. Finally, the information of each mode shape is combined into a single
damage index β i by using the definition proposed by Cornwell et al. [3], which is a
summation over the number of modes Nmodes considered:
,
Nmodes h . i Nmodes h . i
(n) (n) (n) (n)
β i = ∑ γ̃i γ̃ ∑ i γ γ . (3.10)
n =1 n =1

The derivation of the damage indicator for the torsional deformations follows
the same route. The starting point is the strain energy formulation for torsional
deformations as presented in Equation (3.2).

3.3 Composite T-beam structure


The structure investigated in this work is a composite T-shaped stiffener section.
This type of stiffener is frequently used in aerospace components to increase the
bending stiffness of the component without a severe weight penalty. Recently, a new
type of stiffener, depicted in Figure 3.1, was developed by Fokker Aerostructures,
in collaboration with the Dutch National Aerospace Laboratories (NLR). The
fabrication process for this type of stiffener is presented in [9]. This type of stiffener
connection is referred to as a T-joint.
The T-beam is built from 16 individual plies of unidirectional co-consolidated
carbon AS4D reinforced PEKK. A [0/90]4,s lay-up is used for all specimens, see
Figure 3.1. A PEKK injection molded filler is used as a connection. The lay-up was
selected to create a structure with a relatively high bending rigidity and a relatively
low torsional rigidity. The torsion modes will be more pronouncedly present, hence
easier to detect. As a result, it is more straightforward to assess the feasibility of
including the torsion modes in the analysis. The component is relatively long, so
as to obtain a sufficient deflection to be measured conveniently with the equipment
available at the laboratory of the University of Twente. The dimensions of the T-
44 Chapter 3. Vibration based structural health monitoring of a composite T-beam

Clamping area
2.2 mm 3
1
2 x
m
40 mm 3 60 m
Shear 1 2
center mm
1000

50 mm z
100 mm

Figure 3.1 The composite T-beam with a [0/90]4,s laminate lay-up. The global and material coordinate
systems are indicated.

beam are indicated in Figure 3.1.


A typical damage occurring to composite structures is delamination. The location
with the highest risk of failure of the structure is the injection molded thermoplastic
T-joint profile which connects the base to the stiffener. The aim of this research
is to identify and localize such damage, hence undamaged and damaged T-
beams are manufactured. The damaged T-beam contains a 100 mm long artificial
delamination, right under the T-joint. The location is indicated in Figure 3.2.
The delamination was created by inserting a 0.1 mm thick Polyimide film before
consolidating the beam in the autoclave.

3.4 Experimental analysis of a sub-structure

3.4.1 Experimental outline


Vibration measurements are performed on an intact and delaminated T-beam in
order to detect and localize the delamination and to validate the numerical model
and results [10]. The frequency response functions (FRFs) between the fixed

Side view
y Clamping area 100 mm Delaminated region

z
450 mm
60 1000 mm

Figure 3.2 A side view of the composite T-beam. The location and dimensions of the artificial delamination
at the T-joint are indicated.
3.4. Experimental analysis of a sub-structure 45

point of excitation and the measuring points along the T-beam are determined
using a laser vibrometer. The modal parameters, i.e. the natural frequencies, the
damping values and the mode shapes, are obtained from these frequency response
functions by using experimental modal analysis (EMA) [11]. Firstly, the effect of the
delamination on the natural frequencies and mode shapes is investigated. Secondly,
the modal strain energy damage index algorithm is evaluated using the mode shape
data from the intact and delaminated T-beam.

3.4.2 Set-up and vibration measurements

The complete dynamic set-up and data acquisition scheme used for the experimental
validation are shown in Figure 3.3. The T-beam is horizontally and vertically
clamped at one side employing the hydraulic clamps of an Instron 8516 Fatigue
system. For all tests a 6 kN vertical clamping force is combined with a clamping
pressure of 200 bar applied to the horizontal clamps. These clamping conditions
are obtained by analyzing the robustness of a frequency response plot to changes
in the clamping conditions. The repeatability of the experimental set-up has been
checked in Section 3.4.3. The T-beam was excited by a shaker with a stinger and
force transducer connected to a fixed point at the beam and a spring connected to
its support. A chirp excitation force has been applied to the structure. The laser
vibrometer sensor head is mounted on a traverse system, which has a horizontal
scanning range restricted to 590 mm. The laser vibrometer is used to measure the
velocities at an equally spaced measuring grid containing 3×30 points (L1/M1/R1
until L30/M30/R30, see Figure 3.4) within the range of the traverse system. These
grid lines were selected to be able to distinguish between torsion and bending
modes and to approximate the torsion angle. Two accelerometers, one at the point
of excitation and one at grid point R1, were used for validation of the responses
measured by the laser vibrometer. The frequency response functions between the
excitation force at the fixed excitation point and the velocities at all laser vibrometer
grid points are recorded by a Siglab system. A frequency range of 25–1025 Hz, with
a resolution of 0.313 Hz was selected. A measurement at each grid point consists of
20 averages, without overlap.

The modal parameters (i.e. natural frequency, mode shapes and damping values)
were obtained from the FRFs by using experimental modal analysis [11, 12]. A
Complex Mode Indicator Function (CMIF) together with a global curve fitting
procedure, based on a Rational Fraction Polynomial (RFP) function [13], was used to
estimate the poles and residue mode shape vectors. Slightly complex modal vectors
were considered to be real normal, by neglecting non-proportional damping effects.
The mode shapes from both T-beams were used for damage identification by the
modal strain energy damage index algorithm.
46 Chapter 3. Vibration based structural health monitoring of a composite T-beam

40
30
13.

Magnitude [dB]
20
10
0
-10
-20
0 200 400 600 800 1000 1200
Frequency [Hz]

12.

11. 10. 9.

3. F(t) a(t) a(t) v(t)

2.
4.
1.
6.

5.

8. 7.

No.Description Hardware No.Description Hardware


1. Force transducer B&K 8001 8. Traverse system Dantec
2. Shaker B&K 4809 9. Laser: controller Polytec OFV 3001
3. Spring 10. Conditioning amplifier B&K Nexus
4. Clamping device Instron 8516 11. Power amplifier B&K 2706
5. Accelerometer B&K 4517C-001 12. Data acquisition Siglab Model 20-42
6. Composite T-beam specimen 13. Computer, Siglab & ME’scope Analysis Software
7. Laser: sensor head Polytec OFV-303

Figure 3.3 Dynamic set-up and data acquisition systems used for the experimental investigation.

3.4.3 Results

Repeatability of the testing approach and set-up

The repeatability of the testing approach and experimental set-up is studied by


performing separate tests with constant conditions at distinct times. The specimen
was removed after each test. The resulting variations in the measured frequency
response plots, natural frequencies and mode shapes give an indication of the
repeatability of the testing method.
3.4. Experimental analysis of a sub-structure 47

Bottom view x
1000 mm 60
50 590 mm scanning range laser
R1 20 mm R30 204 mm
50
41 41

z
M1 M30

37
L1 L2 L30
3x30 measurement points Excitation point

Figure 3.4 A bottom view of the composite T-beam. The location of the excitation point and the 3×30
measurement points is indicated.

First of all the repeatability on the frequency response plot was tested. Three
tests were performed on each T-beam with the laser vibrometer positioned at grid
point R1. Figure 3.5 shows the frequency response plots of the intact T-beam. The
different curves compare well. The average standard deviation over all measured
peak values was 0.56%. Overall, the peak values of the torsion modes show a higher
standard deviation (average: 0.75%) than for the bending modes (average: 0.31%).
Even smaller standard deviations over the measured peak values (average 0.33%)
were found for the delaminated T-beam. The largest standard deviation in this
case was 0.61% for the 5th bending mode with an average frequency of 814.0 Hz.
The natural frequencies used for comparison of the intact and delaminated T-beam
are obtained from curve fitted data of the complete set of 3×30 FRF measurements
instead of a single FRF measurement. The standard deviations are of the same order
of magnitude.

The repeatability on the mode shapes was assessed in terms of the modal assurance
criterion (MAC). This criterion provides a measure for the degree of consistency
between two modal vectors. A value close to one indicates a good correspondence
between the modal vectors, whereas a value close to zero indicates that the modal

40
|HFv| [dB]

20
Case I
0 Case II
Case III
0 100 200 300 400 500 600 700 800 900 1000
Frequency [Hz]

Figure 3.5 The repeatability of the frequency response function based on three measurements performed at
grid point R1 of the pristine T-beam.
48 Chapter 3. Vibration based structural health monitoring of a composite T-beam

1 1
Case A Case A
0.5 Case B 0.5 Case B
uy (z) [−]

uy (z) [−]
Left Left
0 Center 0 Center
(n )

(n )
Right Right
−0.5 −0.5

−1 −1
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
z−coordinate [m] z−coordinate [m]
(a) 4th torsion mode shape (MAC = 0.999). (b) 7th bending mode shape (MAC = 0.995).

Figure 3.6 The repeatability of the (a) 4th torsion and (b) 7th bending mode shape of the delaminated
T-beam measured at the 3×30 grid points according to the left, center and right grid line.

vectors are not consistent. The MAC criterion is defined as [14]:


A T B ∗ 2
(ψm ) (ψn )
MAC (m, n) = A )T ( ψA ) ∗ ( ψB )T ( ψB ) ∗
, (3.11)
(ψm m n n

with ψm A the modal vector of mode m obtained at the first measurement (case A),
B
ψn the modal vector of mode n obtained at the second measurement (case B),
∗ the complex conjugate and T the transpose. Figures 3.6(a) and 3.6(b) show

the comparison of two mode shape measurements for the 4th torsion mode and
the 7th bending mode of the delaminated T-beam. The mode shapes of the two
measurements correspond well. Overall, this holds also for the other measured
bending and torsion mode shapes with, respectively, average MAC values of 0.960
and 0.975. Only a few mode shapes show slightly larger deviations. The reason for
these deviations can be various. Possible explanations are the effects caused by the
clamping, the spring tension of the shaker support, the positioning and alignment
of the set-up and several other noise-sensitive aspects.

Since all average standard deviations do not exceed the 1% and the average
MAC values of the delaminated T-beam exceed the 0.950, it is concluded that the
natural frequencies and mode shapes obtained at the experimental set-up satisfy
the demands in terms of repeatability. Results on the fabrication repeatability of
the intact and delaminated T-beams are not available yet. Equal process conditions,
materials and tools are used. The weight of the intact (518 grams) and delaminated
(514 grams) T-beam correspond well.
3.4. Experimental analysis of a sub-structure 49

60 Intact
Damaged
|H | [dB] 40

20
Fv

−20
0 100 200 300 400 500 600 700 800 900 1000
Frequency [Hz]

Figure 3.7 FRF comparison of the intact and delaminated T-beam at grid point R1.

Damage identification based on modal parameters

First of all the effect of the delamination on the natural frequencies and mode
shapes is investigated. In Figure 3.7 the frequency response plots of the intact
and delaminated T-beam at grid point R1 are compared. This figure indicates that
the natural frequencies of the delaminated T-beams are lower than for the intact
T-beam. Qualitatively this corresponds well with the literature [15, 16]. It can also
be observed that the difference in natural frequency is larger for higher modes.
Table 3.1 provides an overview of the natural frequencies f (n) and damping values
ζ (n) obtained from the curve fitted FRF data of the intact and delaminated T-beam.
The frequencies are grouped based on their dominating type of mode shape. The
bending modes in the yz-plane and torsion modes about the z-axis are respectively
referred by Byz and Tz . Coupling between some successive modes was found.
Some modes are experimentally not measured due to various reasons. The natural
frequency of the first torsion mode is smaller than the lowest frequency of 25 Hz
of the excitation signal. Other modes, like the 6th torsion mode, are not measured
because of the low response. In that case, the point of excitation almost coincides
with a point of zero amplitude in the mode shape. This overview also confirms
the shift in natural frequencies to lower values of the delaminated T-beam in the
case of the bending modes. Note that the natural frequencies of the torsion modes
hardly change. The delamination underneath the T-joint mainly affects the natural
frequencies of the bending modes and has less influence on the natural frequencies
of the torsion modes. Numerical results presented in [10] confirm this conclusion.
This observation can be explained by the fact that the delamination is located close
to the shear center of the cross section of the T-beam. The (viscous) damping
values ζ (n) are also determined (see Table 3.1) but no solid conclusions could be
drawn from these values. The statistic variations exceed the changes caused by the
delamination.
Each mode of the intact and delaminated T-beam, mentioned in Table 3.1, is related
to a mode shape. A selection of experimentally obtained mode shapes is shown
in Figures 3.8, 3.9 and 3.10. The mode shapes are determined for a section of the
50 Chapter 3. Vibration based structural health monitoring of a composite T-beam

Table 3.1 Experimentally obtained natural frequencies f (n) and damping values ζ (n) of mode n.

Intact T-beam Delaminated T-beam


Mode Bending Byz Torsion Tz Bending Byz Torsion Tz
f˜B [Hz] ζ̃ B [%] f˜T [Hz] ζ̃ T [%]
(n) (n) (n) (n) (n) (n) (n) (n)
n fB [Hz] ζ B [%] f T [Hz] ζ T [%]
1 37.2 0.272 - - 36.7 0.414 - -
2 217 0.194 65.5 1.014 2111 0.387 63.1 0.917
3 506 0.018 117.8 0.602 493 0.079 116.1 0.532
4 736 0.239 163.7 1.124 714 0.197 162.1 1.111
5 840 0.254 212 0.494 816 0.200 2111 0.387
6 - - - - 886 0.048 - -
7 934 0.131 351 0.210 920 0.198 348 0.095
8 994 0.172 - - 996 0.209 444 0.942
9 - - 616 0.726 - - 601 0.619
10 - - 699 0.124 - - 699 0.018
11 - - 826 0.256 - - 827 0.001
12 - - 971 0.273 - - 973 0.538
1 The mode shapes of these modes showed clear coupling between the bending and torsion mode.

T-beam due to the limited scanning range of the laser vibrometer. In most cases it
is possible to extrapolate the mode shapes, supported by the numerical model [10].

It is observed that the difference in mode shapes, caused by the delamination, is


less for the lowest frequencies. Mainly the mode shapes at higher frequencies are
affected, but every mode shape is affected differently by the delamination. The
(n)
4th and higher bending mode shapes ( f B > 700 Hz) clearly show the presence,
location and size of the delamination in the T-beam. The global behavior of the
T-beam is affected by the local delamination. This implies that measurements at the
left and right grid line, at a certain distance from the local delamination, can be used
to indicate the delamination. The 9th torsion mode shape (Figure 3.10) is nearly the
only torsion mode shape which is clearly affected by the delamination. The wave
of the mode shape nicely coincides with the position of the delamination. These
observations match well with the conclusions drawn from the natural frequencies.

Damage identification employing the MSE-DI algorithm

The mode shapes from the intact and delaminated T-beam are used for damage
identification by the modal strain energy damage index algorithm, presented in
Section 3.2. One of the advantages of this algorithm is that damage related
information of separate modes is combined to obtain a damage index distribution,
using Equation (3.10). The required derivatives of the mode shapes are obtained
after applying a cubic spline interpolation (piecewise function) to the data points of
3.4. Experimental analysis of a sub-structure 51

1 0 1 0
uy (z) [−]

~(n) (z) [−]


0 0
0.5 0.5
(n )

u y
−1 −1
−0.04 0.04 1 z−coordinate [m] −0.04 0.04 1 z−coordinate [m]

x−coordinate [m] x−coordinate [m]


uy (z) [−]

~(n) (z) [−]


1 1
0 0
(n)

u y
−1 −1
1 0.8 0.6 0.4 0.2 0 1 0.8 0.6 0.4 0.2 0
z−coordinate [m] z−coordinate [m]
(5) (5)
(a) Intact ( f B = 840 Hz). (b) Delaminated ( f˜B = 816 Hz).

Figure 3.8 Experimentally obtained 5th bending mode shapes (MAC = 0.852).

1 0 1 0
uy (z) [−]

~(n) (z) [−]

0 0
0.5 0.5
(n)

u y

−1 −1
−0.04 0.04 1 z−coordinate [m] −0.04 0.04 z−coordinate [m]
1
x−coordinate [m] x−coordinate [m]
uy (z) [−]

~(n) (z) [−]

1 1
0 0
(n)

u y

−1 −1
1 0.8 0.6 0.4 0.2 0 1 0.8 0.6 0.4 0.2 0
z−coordinate [m] z−coordinate [m]
(7) (7)
(a) Intact ( f B = 934 Hz). (b) Delaminated ( f˜B = 920 Hz).

Figure 3.9 Experimentally obtained 7th bending mode shapes (MAC = 0.833).
52 Chapter 3. Vibration based structural health monitoring of a composite T-beam

1 0 1 0
uy (z) [−]

~(n) (z) [−]


0 0
0.5 0.5
(n )

u y
−1 −1
−0.04 0.04 1 z−coordinate [m] −0.04 0.04 1 z−coordinate [m]

x−coordinate [m] x−coordinate [m]


uy (z) [−]

~(n) (z) [−]


1 1
0 0
(n )

u y
−1 −1
1 0.8 0.6 0.4 0.2 0 1 0.8 0.6 0.4 0.2 0
z−coordinate [m] z−coordinate [m]
(9) (9)
(a) Intact ( f T = 616 Hz). (b) Delaminated ( f˜T = 601 Hz).

Figure 3.10 Experimentally obtained 9th torsion mode shapes (MAC = 0.971).

the mode shapes. The interpolation is used to improve the calculation of the mode
shape derivatives at discrete points in the z-direction. In the analysis, the scanning
area is subdivided into 40 elements.
The results obtained by considering only the 4th , 5th , 7th and 8th bending dominated
mode shapes and the full set of 3×30 grid points, are shown in Figure 3.11. The
mode shapes of the lowest bending modes are not taken into account. These modes
show some coupling or other kind of distortions and contain less information about
the presence and location of the delamination. The data obtained at the center grid
line clearly identifies the presence, location and size of the delamination. This is
mainly due to the fact that the measuring points are close to the delamination. The
left and right grid line also show higher damage indices at the delaminated region,
apart from some fitting related artifacts. The ratio between the damage index of
the damaged area and the intact area is lower, but the location is still predicted
correctly. It was shown that the value of the damage index is a measure for the
sensitivity to identify damage at a certain distance from the measurement points.
The same analysis is performed by using less data points. Figure 3.12 shows the
results obtained by using 3×30, 3×15, 3×10, 3×8, 3×6 and 3×5 equally spaced
measurement points at each grid line to define the mode shape of the T-beam.
The bending modes incorporated into the analysis remains unchanged. The ratio
between the damage indices of the damaged area and the intact area is lower for a
decreasing number of data points. It is shown that the number of measurement
points affects the sensitivity to identify damage at a certain distance from the
3.4. Experimental analysis of a sub-structure 53

10 10
Damage Index β [−]

Damage Index β [−]


5 5

0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
z−coordinate [m] z−coordinate [m]
(a) Left grid line. (b) Right grid line.

10
Damage Index β [−]

0
0 0.2 0.4 0.6 0.8 1
z−coordinate [m]
(c) Center grid line.

Figure 3.11 Damage index β i distributions obtained for the bending dominated modes (4th , 5th , 7th and
8th ) defined by 3×30 grid points and using 40 MSE-DI elements.

measured points. The damage indices obtained from the center grid line of the
T-beam clearly indicate the presence, location and size of the delamination using 8
equally spaced data points within the scanning range. The left and right grid line
show higher damage indices at the delaminated region in case 10 equally spaced
data points were used. The location of the delamination is predicted correctly,
despite the lower absolute values of the damage index compared to the damage
indices at the center grid line. The difference between the results obtained from the
left and right grid line can be allocated to differences between mode shapes, and
therefore derivatives, at the left and right grid line.

The damage identification algorithm is also applied by incorporating torsion


dominated mode shapes. Therefore the first derivative of the torsion angle is
required. The normalized torsion angle at the shear center (see Figure 3.1) is
approximated using the mode shapes in y-direction at the left and right grid line
and the distance in x-direction between these two grid lines. Figure 3.13 shows the
results by incorporating the approximated torsion angle at the spline interpolated
54 Chapter 3. Vibration based structural health monitoring of a composite T-beam
Damage Index β [−]

Damage Index β [−]


10 10

5 5

0 0
0 0.2 0 0.2
0.4 0.4
0.6 0.6
0.8 0.8
1 z−coordinate [m] 1 z−coordinate [m]

(a) Left grid line. (b) Right grid line.

Data
Damage Index β [−]

points
10
30
15
10
5 8
6
0 5
0 0.2
0.4
0.6
0.8
1 z−coordinate [m]

(c) Center grid line.

Figure 3.12 Damage index β i distributions obtained for the bending dominated modes defined by 3×30,
3×15, 3×10, 3×8, 3×6 and 3×5 equally spaced grid points.

grid points of the 2nd , 3rd , 4th , 7th , 9th , 10th and 12th torsion mode shape. The
damage index based on torsion dominated mode shapes gives no indication of
presence, location and size of the delamination. This conclusion matches the
conclusions drawn based on the modal parameters. The explanation for this is
the small distance between delamination and shear center at the cross section of the
T-beam.

3.5 Conclusions

The results of the vibration experiments show that the modal strain energy damage
index algorithm can be used to identify a delamination in a composite T-beam
structure in the case of bending vibrations. Although measurements of changes
in natural frequencies of the bending modes only indicate the presence of the
delamination easily, the 4th and higher bending mode shapes and the damage
3.5. Conclusions 55

Damage Index β
4

0
0 0.2 0.4 0.6 0.8 1
z−coordinate [m]

Figure 3.13 Damage index β i distribution obtained for the torsion dominated modes (2nd , 3rd , 4th , 7th ,
9th , 10th and 12th ) defined by 30 grid points and using 40 MSE-DI elements.

index algorithm also predict the location and size of the defect. The damage
index distributions provide a more pronounced identification of the delamination
compared to the single mode shapes. The combination of the different mode shape
derivatives in the damage index algorithm makes the methodology sensitive, even
with a low number of measurement points.
The small distance between the location of the delamination and the shear center
of the cross section of the structure causes the torsion modes to be hardly affected
for this structure. Only a few mode shapes indicate the presence and location of
the delamination. The natural frequencies and damage index remain inconclusive,
as was shown by the measurements. No solid conclusions can be drawn in this
particular case. Perhaps there are other cases where torsional deformations will
have an influence on the damage identification results.
It was shown that measurements at a certain distance from the local delamination,
like the left and right grid line, can be used to identify the delamination. The
damage index is affected by the distance between the measured points and the
delamination.
A reduction in the number of measurement points affects the sensitivity to identify
damage at a certain distance from the measured points. The ratio between the
damage index of the damaged area and the intact area is lower, but the location is
still predicted correctly by using a minimum of eight equally spaced measurement
points within the limited scanning range at the center grid line.
A next step towards the practical application of vibration based structural health
monitoring is to use integrated sensor systems for the mode shape measurements
used in the damage identification process. For this purpose, the numerical model,
as proposed in [10] can be used to optimally place the different sensors. Moreover,
larger structures with multiple stiffeners should be investigated to further study the
capabilities of the vibration based damage identification approach.
56 Chapter 3. Vibration based structural health monitoring of a composite T-beam

Acknowledgements
The authors would like to acknowledge the support of Fokker Aerostructures B.V.,
Hoogeveen, The Netherlands, for the manufacturing of the composite T-beam
structures on which the experiments were performed. This work was carried out
in the framework of the European project Clean Sky, Eco-Design ITD (grant
agreement number CSJU-GAM-ED-2008-001).

References
[1] W.J.B. Grouve, L.L. Warnet, A. de Boer, R. Akkerman, and J. Vlekken. Delamination
detection with fibre Bragg gratings based on dynamic behaviour. Composites Science
and Technology, 68(12):2418–2424, 2008.
[2] N. Stubbs and C.R. Farrar. Field verification of a nondestructive damage localization
and severity estimation algorithm. In Proceedings of the 13th International Modal Analysis
Conference (IMAC XIII), pages 210–218, 1995.
[3] P.J. Cornwell, S.W. Doebling, and C.R. Farrar. Application of the strain energy damage
detection method to plate-like structures. Journal of Sound and Vibration, 224(2):359–374,
1999.
[4] H. Li, H. Yang, and S.-L.J. Hu. Modal strain energy decomposition method for damage
localization in 3D frame structures. Journal of Engineering Mechanics, 132(9), 2006.
[5] H. Li, H. Fang, and S.-L.J. Hu. Damage localization and severity estimate for
three-dimensional frame structures. Journal of Sound and Vibration, 301(3-5):481–494,
2007.
[6] T.A. Duffey, S.W. Doebling, C.R. Farrar, W.E. Baker, and W.H. Rhee. Vibration-based
damage identification in structures exhibiting axial and torsional response. Journal of
Vibration and Acoustics, 123(1):84, 2001.
[7] A. Alvandi and C. Cremona. Assessment of vibration-based damage identification
techniques. Journal of Sound and Vibration, 292(1-2):179–202, 2006.
[8] J. Stewart. Calculus: early transcendentals. 5th edition. Brooks/Cole - Thomson
Learning, 2003.
[9] A. Offringa, J. List, J. Teunissen, and H. Wiersma. Fiber reinforced thermoplastic butt
joint development. In Proceedings of the International SAMPE Conference, pages 1–16,
2008.
[10] R. Loendersloot, T.H. Ooijevaar, L.L. Warnet, A. de Boer, and R. Akkerman. Vibration
based structural health monitoring and the modal strain energy damage index
algorithm applied to a composite T-beam. In C.M.A. Vasques and J.D. Rodrigues,
editors, Vibration and Structural Acoustics Analysis: Current Research and Related
Technologies, Chapter 6, pages 121–150. Springer, 2011.
[11] B.J. Schwarz and M.H. Richardson. Experimental modal analysis. In CSI Reliability
Week, Orlando, FL, pages 1–12, 1999.
[12] M.H. Richardson and B.J. Schwarz. Modal parameter estimation from operating data.
Sound and Vibration, 37(1):28–39, 2003.
[13] M.H. Richardson and D.L. Formenti. Global curve fitting of frequency response
measurements using the rational fraction polynomial method. In 3rd International
Modal Analysis Conference, pages 1–8, 1985.
References 57

[14] R.J. Allemang. The modal assurance criterion – Twenty years of use and abuse. Journal
of Sound and Vibration, 1:14–21, 2003.
[15] P.M. Mujumdar and S. Suryanarayan. Flexural vibrations of beams with delaminations.
Journal of Sound and Vibration, 125(3):441–461, 1988.
[16] J. Lee. Free vibration analysis of delaminated composite Beams. Computers and
Structures, 74(2):121–129, 2000.
Chapter 4

Impact damage identification in


composite skin-stiffener structures
based on modal curvatures1

Abstract
The feasibility of a vibration based damage identification approach
for impact damage in two advanced composite skin-stiffener structures
is investigated in this chapter. Mode shape curvatures combined
with the modal strain energy damage index (MSE-DI) algorithm
are utilized to identify the damage. Special attention is paid to
the effective application of this vibration based methodology by
investigating the relation between the damage location, the structural
design and the dynamic behavior. The performance of a 1D and
2D formulation of the MSE-DI algorithm is compared for several
damage scenarios. Experiments demonstrated the capabilities of the
MSE-DI algorithm to detect, localize and roughly quantify the size
of barely visible impact damage in advanced composite structures.
It is concluded that the method is particularly effective for health
monitoring of skin-stiffener connections. The most effective results
were obtained by considering the 1D formulation in the direction of
the stiffeners for the stiffener mid-section and perpendicular to the
stiffeners for the stiffener run-out. The method remained inconclusive
in the case of pure skin related damage. The results obtained in this
study contribute to the development of guidelines for vibration based
structural health monitoring of advanced composite skin-stiffener
structures.

1 Reproduced from: T.H. Ooijevaar, L.L. Warnet, R. Loendersloot, R. Akkerman and T. Tinga,
Impact damage identification in composite skin-stiffener structures based on modal curvatures.
In preparation for: Structural Control and Health Monitoring, 2014.

59
60 Chapter 4. Damage identification in skin-stiffener structures based on curvatures

4.1 Introduction
One of the key issues in composite structures is the early identification of damage.
Often service induced damage does not involve visible plastic deformation, but
internal matrix related damage, like transverse cracks and delaminations. Their
identification often imposes time-consuming and costly nondestructive inspection
techniques. Structural health monitoring (SHM) technologies propose to be a
promising alternative.
A wide range of technologies is currently employed for health monitoring
purposes [1]. The majority of these technologies utilizes local or global dynamic
characteristics [2, 3]. This classification is based on the relative size of the wave
length with respect to the overall structural dimensions. Global methods, typically
structural vibration based technologies, provide data that is relatively easy to
interpret, but the frequency range and hence the resolution are limited [3]. As a
consequence, only relatively severe damage can be identified. Typical applications
are found in civil and offshore structures [4–8]. The local methods, comprising
higher frequency wave propagation technologies, are capable of detecting small
damage such as cracks [9–13], but it is often more complicated to interpret the data,
in particular in the case of non-flat or complex structures. The selection of the best
technology for a SHM system is often a matter of compromise. None of the local or
global approaches is, at the moment, able to solve all problems in all structures [14].
The number of successful practical applications of the SHM technologies on
composite aircraft structures is limited compared to applications in civil and
offshore industry [14–17]. This is mainly due to the complexity of the components,
the variety of potential damage scenarios and the high demands on safety and
reliability of the SHM system. As a consequence, the unique identification
(presence, location and severity) of damage in realistic composite structures by
methodologies operating under field conditions is considered as one of the key
challenges for these technologies.
This chapter addresses an experimental feasibility study on the identification
of impact damage in two advanced (i.e. multiple stiffeners, non-uniform skin
thickness, large dimensions) composite skin-stiffener structures, representing
typical aircraft structures. The present study is an extension of earlier performed
research by the authors on a 1 m composite T-beam specimen with an artificial
delamination [18, 19]. A global vibration based method is considered. This
approach allows for a relatively easy and physical interpretation of the measured
responses, which is desired for the investigation of complex composite structures.
Mode shape curvatures combined with the modal strain energy damage index
(MSE-DI) algorithm [20, 21] are respectively utilized as damage feature and
classifier to detect and localize the naturally originated defects. Studies by
different authors [4, 22–26] showed that this mode shape curvature based approach
outperforms other frequently applied modal domain damage identification methods
4.2. Composite skin-stiffener structures 61

based on natural frequencies [27–29], mode shapes [30, 31], dynamic flexibility [32–
34] and other mode shape derivative based features [35, 36]. The MSE-DI algorithm
essentially utilizes the curvature changes that locally arise under dynamic loading
of damaged structures. The performance of a 1D and 2D formulation of the
MSE-DI algorithm is compared for several local damage scenarios. Up to now,
the feasibility of the method is mainly demonstrated on concrete and metallic
structures [4, 6, 24, 37]. Application to composite structures is limited to relatively
simple composite beams and plates [38–40] with mainly severe and well-defined
artificial damage, such as a saw cut. The present work aims to extend this to more
advanced composite structures and impact induced damage scenarios.
Although the low frequency technologies allow for a relatively easy interpretation
of the responses, their sensitivity to identify small and local damage is rather
limited [16, 23, 41]. Special attention is therefore paid to the effective application of
the vibration based methodology. The effectiveness of the methodology is directly
related to aspects such as the frequency level, the location, orientation, type and
size of the damage, the design of the structure, the location and number of sensors,
etc [18, 22]. This chapter addresses the relation between the damage location, the
structural design and the dynamic behavior in order to extract recommendations
for the effective application of the methodology. The results obtained in this
study contribute to the development of guidelines for vibration based structural
health monitoring of advanced composite skin-stiffener structures. Numerical
studies performed on the same skin-stiffener structures were presented in two book
chapters [42, 43].

4.2 Composite skin-stiffener structures


This research concentrates on carbon AS4D fiber reinforced thermoplastic (PEKK)
skin-stiffener structures. Stiffeners are frequently used in aerospace components to
increase the bending stiffness of the component without a severe weight penalty. A
new type of skin-stiffener connection was developed by Fokker Aerostructures [44].
Flat preforms are butt-joined to the composite skin in a co-consolidation process,
employing an injection molded thermoplastic filler for the connection. The filler is
made from PEKK and contains 20wt% short carbon fibers. This stiffener connection
concept allows for a low cost and simplified manufacturing of a large number of
stiffening ribs compared to conventional techniques. The fabrication process and
mechanical performances are discussed in more detail in [45].
The two composite skin-stiffener structures used in this research were produced by
Fokker Aerostructures according to the new joining concept. The first structure is
shown in Figure 4.1 and consists of a 400×282 mm composite plate with two T-
shaped stiffener sections. A quasi-isotropic lay-up [45/90/-45/0]2,s consisting of 16
unidirectional layers is used for both skin and stiffeners.
62 Chapter 4. Damage identification in skin-stiffener structures based on curvatures

2.16mm
Excitation 2.16mm
1 3
point 2
3
1 2
x Impact location: case II
y Damaged area: case II
z
Bottom view Side view
400 mm Damaged area: case II
38 Impact location: case II

70mm
#89 #99
31 31

~120 mm

α=45° Impact location: case I


36 34

Top-ply fibre

142mm
282mm
orientation

Excitation
y
point
10mm #01 x #11
8mm Measurement point 40mm

Figure 4.1 Three dimensional and bottom view of the composite skin-stiffener specimen with 2 butt-joint
stiffeners. The dimensions, the measurement points (dots) and the impact locations are indicated.

The second structure is a larger and more advanced three stiffener plate structure
(1304×456 mm) with a non-uniform skin thickness distribution, as shown in
Figure 4.2. It represents a more realistic aircraft structure. The plate consists of
a 16-layer quasi-isotropic mid-section and 44-layer and 30-layer end-sections. The
stiffeners are made of a 15-layer quasi-isotropic stack. The structure contains a Z-
shaped preform based stiffener run-out with a titanium insert. Offringa et al. [45]
stated that this run-out design allows the structure to withstand a 50 J impact on the
stiffener from the outside, satisfying airliner and business jet impact requirements.
Only limited damage should be visible at this impact level.

The skin-stiffener structures are vulnerable for impact damage. The connection
between skin and stiffener is considered as the location where damage due to
impact would have the largest effect on the integrity of the structure. A small
delamination at this connection can significantly reduce the local bending stiffness
and hence the performance of the skin-stiffener structure. This direct structural
effect along with the potential long-term delamination growth make the skin-
stiffener connection a good candidate for health monitoring.
4.2. Composite skin-stiffener structures 63

4 mm
t = 4.1
1 mm
t = 2.2
8 mm
t = 6.0

Z-shaped Stringer (3x)


preform with
titanium insert Rib (3x) connection holes
Impact location:
Damaged area: case III
case III
Bottom view
t = 4.14mm t = 2.21mm t = 6.08mm
1304mm
Section A-A
#175 45 A 242mm

92mm
67.5 67.5 67

#203

228mm
153 mm 150mm

456mm
50
Excitation
y
point #001
19mm x 25mm
26mm A Impact location: #029 26mm 35mm
Measurement point
case III Damaged area:
case III

Figure 4.2 Three dimensional and bottom view of the composite aircraft structure with three stiffeners,
a non-uniform skin thickness and the impact resistant stringer run-out design. The dimensions, the
measurement points (dots) and the impact location are indicated. The dotted vertical lines indicate the edges
of the transition zones between the sections with different thicknesses.

Three impact induced damage scenarios, summarized in Table 4.1, are considered
in this study. Naturally originated defects are obtained by applying local impact at
several locations from the outside (i.e. the side of the skin without the stiffeners)
of the structure with the help of a Dynatup 8250 falling weight impact machine. A
repeated impact of 10 J was used for the two stiffener structure, while a single 50 J
impact was applied in the case of the three stiffener structure. Visual and ultrasonic
inspections revealed the damage mechanisms occurring.

Figure 4.1 depicts the impact location and damaged area of damage scenario I
and II in the structure with two stiffeners. The first damage scenario, shown in
Figure 4.3(a), consists of local transverse cracks and a small delamination at the
skin in between both stiffeners. More severe damage is obtained for the second
damage case at the skin-stiffener connection of the same structure. This damage
consists of first-ply transverse cracks and a delamination between the first two
64 Chapter 4. Damage identification in skin-stiffener structures based on curvatures

Table 4.1 Overview of the impact induced damage scenarios.

Damage Structure Description Failure Dimensions


case mechanism (approx.)
I Two Impact damage at skin in between Local transverse 15×10 mm
stiffener both stiffeners due to repeated cracks and small
10 J impact, see Figure 4.3(a) delamination
II Two Impact damage at skin-stiffener Skin delamination Length:
stiffener connection due to repeated 10 J and interface ≈ 80 mm
impact, see Figure 4.3(b) failure
III Three Impact damage at stringer run-out Interface failure Length:
stiffener due to 50 J impact, see Figure 4.3(c) and delamination ≈ 90 mm

layers of the skin, as shown in Figure 4.3(b). The interface between the injection
molded filler and skin also failed over a significant length. The structure with three
stiffeners was impacted at a stiffener run-out, shown in Figure 4.2. This part of
the skin-stiffener connection is most vulnerable for failure under impact [45]. The
impact resulted in failure of the connection between the skin and the Z-shaped
stiffener run-out. Cracks are indicated in Figure 4.3(c). An ultrasonic inspection
also revealed delamination damage between the titanium insert and the skin.
The three damage scenarios considered show typical fiber reinforced material
related failure mechanisms. They were, except for a small dent, invisible from the
outside of the structure. The damage scenarios are referred to as barely visible
impact damage (BVID), according to the criteria used in aircraft industry [46].
Aircraft structures have historically relied on visual methods to identify BVID.
Presently, the dent depth is used as the damage metric to define BVID. The
largest dent depth measured for the two composite structures equaled 0.16 mm,
which is smaller than the 0.3 mm BVID threshold as stated in [46]. Structures
containing BVID must sustain ultimate load (UL) for the life of the aircraft structure.

Stiffener Z-shaped preform

Crack A-A A
Delamination and Skin Crack B-B
transverse cracks delamination A B B

(a) Damage case I. (b) Damage case II. (c) Damage case III.

Figure 4.3 Visible part of the impact damage introduced in the (a, b) two stiffener composite structure and
(c) three stiffener composite structure.
4.3. Damage identification procedure 65

This criterion sets the lower bounds for the identification capabilities of an SHM
approach.

4.3 Damage identification procedure


The diagnostic part of the damage identification procedure, according to Figure 4.4,
consists of two steps. In the first step the mode shape curvatures, referred to as
damage features, are extracted. This process comprises the estimation of the modal
parameters by utilizing modal analysis and the calculation of the second derivatives
of the mode shapes in x- and y-direction. Delaminations in, for example, a beam-
like structure under bending deformation theoretically cause a discontinuity in the
curvature at the boundaries of the damaged region [25]. These changes in curvature
between the pristine and damaged structure are the input for the second step. This
step comprises the damage classification process. The local change in the curvature
of a range of modes is combined to obtain a single indicator, the damage metric.
One of the first publications that adopted mode shape curvatures as a means to
identify structural damage is from Pandey et al. [35]. Since this paper, many
different curvature based damage metrics have been proposed, including the MSE-
DI algorithm. The MSE-DI algorithm has not only been used by several researchers,
but it has also appeared in a number of different variants [26, 43]. Modifications
have been made to improve the performance and robustness of the algorithm. It
lies beyond the scope of this article to address all variants that have appeared
throughout the years. The present work is focused on the original and most
frequently used formulation of the MSE-DI algorithm, since the validation of the
modified variants is limited. A 1D formulation of the MSE-DI algorithm was
introduced by Stubbs et al. [20], while Cornwell et al. [21] extended this approach
for the 2D case. Both formulations are employed in the present work. Only the
basics of the 2D formulation according to Cornwell et al. [21] are explained in this

Actuation Response Diagnostics


Fa(t) v(t) Representation
Structure Feature pattern Damage
pristine pattern 2 (n) 2 (n) metric
∂uz ∂uz
HFv(ω) 2 2
∂x , ∂y βij , Zij
Digital signal Feature Current
Classifier state
processing ~ extraction 2 ~(n) 2 ~(n)
~ HFv(ω) u
∂ z∂ z u
Fa(t) 2 2
Structure ∂x , ∂y
~
v(t)
damaged

Figure 4.4 A schematic representation of the damage identification procedure. The Fa (t) is the excitation
force, v(t) the velocity response, HFv (ω ) the frequency response function between an excitation and
(n)
measurement point, uz ( x, y) is the normalized mode shape of mode n, the tilde represents the damaged
case.
66 Chapter 4. Damage identification in skin-stiffener structures based on curvatures

y
(1,Ny)
ly (Nx,Ny)

yj
(i,j)
yj-1
z
(Nx,1)
(1,1)
0 x
0 xi-1 xi lx

Figure 4.5 A plate structure divided into Nx × Ny elements in x- and y-direction respectively.

section. The quasi-isotropic laminate properties are approximated by an isotropic


description. The derivation and assumptions of the 2D formulation are analogous
to the one used for the 1D formulation in x- or y-direction [18, 22].

The strain energy of a vibration mode is referred to as the modal strain energy
of that mode. The modal strain energy can be obtained by linking the flexural,
axial and/or shear modal deformation of a structure to the strain [47]. The present
work concentrates on the flexural deformations of the skin-stiffener structures. This
type of deformation is expected to be dominant for the vibration modes that are
excited by the shaker. The strain energy U, based on the normalized flexural mode
(n)
shape uz ( x, y) in z-direction, of each individual mode n of an isotropic vibrating
plate-like structure is represented by:

Z ly Z l x  !
(n) 2
!
(n) 2
(n) 1 ∂2 u z ∂2 u z
U = D ( x, y) +
2 0 0 ∂x2 ∂y2
(n)
! (n)
! (n)
!2 
∂2 u z ∂2 u z ∂2 u z
+ 2ν + 2 (1 − ν ) dxdy , (4.1)
∂x2 ∂y2 ∂x∂y

with D ( x, y) the flexural rigidity of the plate, ν the Poisson’s ratio, lx and ly the
dimensions of the plate structure in x- and y-direction respectively. The total modal
strain energy is approximated by the sum of Equation (4.1) over a limited set of
Nmodes modes, using the theory of modal superposition. The plate-like structure
can be discretized in Nx × Ny elements in x- and y-direction respectively, as shown
4.3. Damage identification procedure 67

in Figure 4.5. Each element ij contains a modal strain energy described by:

Z y Z x  (n)
!2 (n)
!2
(n) 1 j i ∂2 u z ∂2 u z
Uij = D ( x, y) +
2 y j −1 x i −1 ∂x2 ∂y2
(n)
! (n)
! (n)
!2 
∂2 u z ∂2 u z ∂2 u z
+ 2ν + 2 (1 − ν ) dxdy , (4.2)
∂x2 ∂y2 ∂x∂y

with:
Ny Nx
(n)
U (n) = ∑ ∑ Uij , (4.3)
j =1 i =1

and xi−1 , xi and y j−1 , y j the boundaries of element ij of the discretized structure
in x- and y-direction respectively. Similar quantities can be defined for a damaged
(n)
structure, using the normalized mode shapes ũz ( x, y) of the damaged structure.
Subsequently, the local fractional strain energies, as defined by Cornwell et al. [21],
are:
(n) (n)
(n)
Uij (n)
Ũij
Fij = , F̃ij = , (4.4)
U (n) Ũ (n)
for the pristine and damaged structure respectively. The fractional strain energy
remains relatively constant in the undamaged elements, under the assumption that
the damage is small and primarily located at element kl [21, 22]:

(n) (n)
F̃ij ≈ Fij for i, j 6= k, l . (4.5)

This equation is rearranged and by substituting Equations (4.1), (4.2) and (4.4) it
can be derived that:
R y j R xi .R l R
e (n) dxdy y lx e (n)
y j−1 xi−1 D w̃ 0 0 D w̃ dxdy
1≈ R . for i, j 6= k, l , (4.6)
y j R xi (n) dxdy
R ly R l x
(n) dxdy
y x D w 0 0 D w
j −1 i −1

where w(n) ( x, y) represents all terms between the square brackets in the integrand
of Equation (4.2), representing the information of the mode shape derivatives. The
mean value theorem for integrals [48] is applied to the stiffness terms D ( x, y) and
De ( x, y) in Equation (4.6) to obtain an average stiffness estimate for each double
integral. For the integral over element ij of the pristine case:
Z y Z x Z y Z x
j i j i
D ( x, y) w(n) ( x, y)dxdy = D ( x0 , y0 ) w(n) ( x, y)dxdy , (4.7)
y j −1 x i −1 y j −1 x i −1
68 Chapter 4. Damage identification in skin-stiffener structures based on curvatures

where x0 , y0 is a point within the integration intervals and D ( x0 , y0 ) = Dij is the


weighted average flexural rigidity of element ij. Similar estimates can be obtained
for the other integrals in Equation (4.6). Subsequently, a ratio of the flexural
rigidities is obtained by assuming that the change in the average value of the flexural
rigidity for the integrals over the entire plate dimensions l x and ly is negligible for
small damage:
R y j R xi .R l R .
(n) dxdy y lx (n) (n)
Dij y j −1 x i −1 w̃ 0 0 w̃ dxdy γ̃ ij γ̃(n)
≈ R . = . for i, j 6= k, l , (4.8)
e ij
D y j R xi R ly R l (n)
y x w(n) dxdy 0 0 x w(n) dxdy γij γ(n)
j −1 i −1

(n)
where γij is the integral of w(n) ( x, y) over element ij and γ(n) the integral of
w(n) ( x, y) over the entire dimensions l x and ly . Stubbs et al. [20], Cornwell et al. [21]
and Alvandi [22] used the right-hand side of Equation (4.8) to define a local damage
(n)
index β ij . This indicator is approximately equal to 1 for the undamaged elements,
i.e. i, j 6= k, l, but can change at the damaged element, i.e. i, j = k, l. Finally, the
information of each mode shape is combined into a single damage index β ij by
using the definition proposed by Cornwell et al. [21], which is a summation over
the number of modes Nmodes considered:
,
Nmodes h . i Nmodes h . i
(n) (n) (n)
β ij = ∑ γ̃ij γ̃ ∑ γij γ(n) . (4.9)
n =1 n =1

An overview of the most common alternative damage index formulations was


presented by Loendersloot et al. [43]. The indicators obtained by these formulations
are metrics for the detection and localization of damage but do not have a direct
physical quantity. The damage index β ij is usually normalized using the standard
deviation σ and the mean µ of the damage index β ij over all elements. This results
in a normalized damage index Zij [49]:

β ij − µ
Zij = . (4.10)
σ
This normalized indicator provides a statistical measure for the detection of outliers
in the damage index distributions, which are expected to be the damaged elements.
A strength of the MSE-DI algorithm is that it utilizes mode shape curvatures,
which showed to be a relatively sensitive damage feature compared to other modal
parameters [15, 35, 36]. The algorithm also combines damage related information
of a range of modes. This allows for a more robust damage index distribution,
according to Equation (4.9), because every mode is affected differently by the local
damage [18]. Weak points are the number of required data points to accurately
describe the mode shapes and the numerical errors introduced with the calculation
4.4. Experimental set-up and signal processing 69

of the second derivative [15, 36, 50, 51]. Although not extensively considered in
this chapter, many different publications address the problem of finding optimum
positions of the measurement points. Savananoz [52] particularly focused on
the estimation of the optimum spatial sampling interval for the calculation of
curvatures in beams. The present formulation of the MSE-DI algorithm also does
not allow to directly link the value of the damage index β ij to a damage severity
in terms of a stiffness loss ratio [22]. The ratio between the damage indices β ij at
the damaged and intact area showed to be a measure for the relative change in the
dynamic behavior of the structure due to the local damage [18]. The damage index
β ij is merely a qualitative and mathematical rather than a physical quantity [43].

4.4 Experimental set-up and signal processing


Vibration measurements were performed on both composite structures before and
after impact is applied. The complete dynamic set-up and data acquisition scheme
used for the experiments are presented in Figure 4.6. The structures were freely
suspended in vertical direction by two elastic wires in order to isolate the plate
from environmental vibrations.
A random excitation force was applied by a shaker. The shaker was connected by a
force transducer and stringer to driving point #01, according to Figures 4.1 and 4.2.
A laser vibrometer, mounted on a traverse system, measured the velocities along
an equidistant x × y measurement grid containing 11×9 and 29×7 points for the
two and three stiffener structure respectively. The mobility Frequency Response
Functions (FRFs) between the fixed point of excitation and the measurement points
were recorded by a Siglab data acquisition system. A frequency range of 50–2050 Hz
(resolution: 0.625 Hz) was selected for the two stiffener structure, while this was
50–1050 Hz (resolution: 0.3125 Hz) for the larger structure with three stiffeners.
These frequency ranges were selected based on the modal density. A measurement
at each grid point consisted of 20 windowed averages, with 50% overlap. The
modal parameters (i.e. natural frequency, mode shapes and damping values) were
obtained from the FRFs by using Experimental Modal Analysis [53, 54]. A Complex
Mode Indicator Function (CMIF) together with a global curve fitting procedure,
based on a Rational Fraction Polynomial (RFP) function [55], was used to estimate
the poles and residue mode shape vectors. Slightly complex modal vectors are
considered to be real normal, by neglecting non-proportional damping effects.
Successive measurements at distinct times showed sufficient repeatability of the
experimental set-up and testing approach. The natural frequencies measured for
the pristine two and three stiffener structure showed an average standard deviation
of respectively 0.20% and 0.05%. The largest standard deviation of the natural
frequencies equaled 0.95% and was measured for the two stiffener structure. The
residue mode shapes are assessed in terms of the modal assurance criterion (MAC).
70 Chapter 4. Damage identification in skin-stiffener structures based on curvatures

11.
40
30
12.

Magnitude [dB]
20
10
0

4. -10
-20
0 200 400 600 800 1000 1200

3. Frequency [Hz]

9.

10. 5.
F(t) 8.
7.
v(t)
6.
y
2. z
1.

No.Description Hardware No.Description Hardware


1. Force transducer B&K 8001 7. Traverse system x/y Dantec
2. Shaker B&K 4809 8. Laser: controller Polytec OFV 3001
3. Elastic wires 9. Conditioning amplifier B&K Nexus
4. Fixed frame 10. Power amplifier B&K 2706
5. Composite structure 11. Data acquisition Siglab Model 20-42
6. Laser: sensor head Polytec OFV-303 12. Computer, Siglab & ME’scope analysis software

Figure 4.6 Schematic representation of the experimental set-up and data acquisition systems used for the
dynamic measurements of both stiffened structures.

This criterion provides a measure for the degree of consistency between two modal
vectors. A value close to one indicates a good correspondence between the modal
vectors, whereas a value close to zero indicates that the modal vectors are not
consistent. The MAC criterion is defined as [56]:
A T B ∗ 2
(ψm ) (ψn )
MAC (m, n) = A )T ( ψA ) ∗ ( ψB )T ( ψB ) ∗
, (4.11)
(ψm m n n

with ψmA the modal vector of mode m obtained by a first measurement (case A),
B
ψn the modal vector of mode n obtained by a second measurement (case B), ∗ the
complex conjugate and T the transpose. Average MAC values of 0.997 and 0.996
were obtained for successive measurements at the two and three stiffener structure
respectively. The MAC values exceeded the 0.968 in all cases.
The residue mode shapes were subsequently cubic spline interpolated (piecewise
function) and normalized to their maximum amplitude. The interpolation is
used to improve the calculation of the second derivatives of the mode shapes at
4.5. Experimental results 71

discrete points in x- and y-direction. The mode shape curvatures from the pristine
and damaged structures were combined in the MSE-DI algorithm to investigate
the effect of the damage scenarios on the damage index distributions β ij or Zij ,
according to the equations presented in Section 4.3.

4.5 Experimental results


The results obtained for the 2D and the 1D formulation in x- and y-direction of
the MSE-DI algorithm are evaluated for the three damage scenarios. The damage
indices for the two and three stiffener structure are evaluated at an x ×y grid of
40×30 and 60×20 discrete elements respectively. All modes of the two and three
stiffener structure, identified within a frequency range of respectively 500–2000
Hz and 250–1000 Hz, are considered in the analysis. The matching pristine and
damaged mode shape combinations within these frequency ranges are found by
considering the highest MAC values. Figure 4.7 presents the calculated MAC
values for all mode combinations in the case of the three considered damage
scenarios. A minimum MAC value of 0.5 is adopted as the lower bound to
incorporate the mode combination in the analysis.

4.5.1 Two stiffener structure with damage scenario I and II


The damage index β ij distributions obtained for damage scenario I and II are
respectively shown in Figures 4.8 and 4.9. The MSE-DI algorithm remains
inconclusive for the first damage scenario where the damage was introduced at
the skin in between the two stiffeners. The results obtained for the 2D and both
1D formulations show damage index β ij distributions with all values close to one
(with standard deviations: σβ (2D) = 0.05, σβ (1Dx) = 0.15, σβ (1Dy) = 0.05). This indicates
a negligible effect caused by the damage. The failure mechanism, consisting of
local transverse cracks and a small delamination (15 × 10 mm), hardly affects the
structural stiffness or mass of the structure. As a result, the modal curvature based
MSE-DI algorithm could not detect and localize the impact damage at the composite
skin, for the frequency range and measurement point density considered in the
analysis. Note that the minor effect of the damage is also expressed by the MAC
values, as shown in Figure 4.7(a). They remain close to one for all matching mode
combinations.
Applying the MSE-DI algorithm to the damage at the skin-stiffener connection,
damage scenario II, shows more successful results. The skin-stiffener failure has a
larger effect on the global structural dynamic behavior, as indicated by the reduced
MAC values in Figure 4.7(b). Small damage at the skin-stiffener interface can
result in a significant local reduction of the bending stiffness of the structure. The
enhanced effect on the dynamic behavior is also depicted by the MSE-DI results
72 Chapter 4. Damage identification in skin-stiffener structures based on curvatures

1 1

1 1

MAC [−]
MAC [−]

0.5 0.5
0.5 0.5
0 0
5 5
10 20 10 15
15 15 10
20 10 0 15 5 0
5
Mode number Mode number Mode number Mode number
pristine structure damaged structure pristine structure damaged structure
(a) Two stiffener structure: damage scenario I. (b) Two stiffener structure: damage scenario II.
1

1
MAC [−]

0.5
0.5
0
5 25
10 20
15 15
20 10 0
5
Mode number 25 Mode number
pristine structure damaged structure
(c) Three stiffener structure: damage scenario III.

Figure 4.7 Comparison of the experimental mode shapes, in terms of MAC values, of the pristine and
impact damaged structures. The mode numbers correspond to the modes within the considered frequency
ranges, i.e. 500–2000 Hz and 250–1000 Hz for the two and three stiffener structure respectively.

of the 2D formulation, shown in Figure 4.9(a). The highest damage indices are
obtained at the damaged area, indicating the presence and location of the damage
correctly. However, the ratio between the damage index β ij of the damaged area
and the intact area is rather low. The relatively noisy damage index distribution also
tends to show a propensity for false damage predictions (indication of damage, but
no actual change in the structure). These false positives can be caused by aspects
such as inaccuracies in the mode shape data, misalignment of measurement points,
errors introduced by the numerical methods used in the algorithm, but also by
considering a small number of modes in the analysis.
The presence and location of the impact damage is predicted in a more pronounced
way by the 1D formulation in x-direction, as shown in Figure 4.9(b). The most
distinct damage indices are obtained by considering the measurement points on
top of the skin-stiffener connection. The damage index distribution also provides
a rough estimation of the dimensions of the damaged area. A delamination length
of roughly 105 mm is found. Earlier research by authors [18] showed that the
accuracy of this result highly depends on the number of measurement points and
4.5. Experimental results 73

2 2

Damage Index β [−]


Damage Index β [−]

1.8 1.8
5 5
1.6 1.6

1.4 0 1.4
0
0.2 0.3 1.2 0.2 0.3 1.2
0.2 0.2
0.1 0.1 0.1 0.1
0 0 x−coordinate [m] 1 0 0 x−coordinate [m] 1
y−coordinate [m] y−coordinate [m]
(a) 2D formulation. (a) 2D formulation.
5 5
m
5m
~10
Damage Index β [−]

Damage Index β [−]


4 4
5 5

3 3
0 0
0.3 2 0.3 2
0.2 0.2
0.2 0.2
0.1 0.1 0.1 0.1
0 0 x−coordinate [m] 1 0 0 x−coordinate [m] 1
y−coordinate [m] y−coordinate [m]
(b) 1D formulation in x-direction. (b) 1D formulation in x-direction.
5 5
Damage Index β [−]

Damage Index β [−]

4 4
5 5

3 3
0 0
0.3 2 0.3 2
0.2 0.2
0.2 0.2
0.1 0.1 0.1 0.1
0 0 x−coordinate [m] 1 0 0 x−coordinate [m] 1
y−coordinate [m] y−coordinate [m]
(c) 1D formulation in y-direction. (c) 1D formulation in y-direction.

Impact location and Impact location


damaged area Damaged area

Case I Case II
y y
x x

Figure 4.8 Damage index β ij distribution for Figure 4.9 Damage index β ij distribution for
damage scenario I in the two stiffener structure. All damage scenario II in the two stiffener structure.
modes within 500–2000 Hz and 40×30 MSE-DI All modes within 500–2000 Hz and 40×30 MSE-
elements are used in the analysis. DI elements are used in the analysis.
74 Chapter 4. Damage identification in skin-stiffener structures based on curvatures

the frequency range considered in the analysis. The real damage length (≈ 120 mm)
is underestimated by approximately 12% for the situation considered here. Despite
the fact that the MSE-DI algorithm does not allow for a physical interpretation of
the damage severity in terms of mechanical properties [22, 43], the elements with
the highest damage index values are predicted at the center of the damaged area.
These elements belong to the region within the damaged area with the most severe
change in the local dynamic behavior.
Figure 4.9(c) presents the results obtained by the 1D formulation in y-direction.
Slightly higher damage indices are predicted at the location of the skin-stiffener
damage. However, also several other potential damage locations are predicted,
hence showing a propensity for false damage predictions.
The results presented until now show that the sensitivity of the method depends on
the extent to which the dynamic behavior is affected by the local damage. This effect
is defined by the failure mechanism and the location of the damage with respect
to the structural design. Hence, the highest sensitivity is obtained for skin-stiffener
damage by considering the 1D formulation underneath and in the direction of the
stiffeners. This is caused by the fact that mainly the flexural stiffness in this direction
is affected by the skin-stiffener interface failure. The flexural stiffness perpendicular
to the stiffener is hardly affected by the interface damage. This directional behavior
is also clearly illustrated by comparing the normalized mode shape curvatures in
x- and y-direction of the pristine ( f n = 996.4 Hz) and damaged ( f n = 987.3 Hz)
structure in Figure 4.10. Locally higher amplitudes are found near the damaged
interface for the curvature in x-direction, while the curvature in y-direction seems
hardly affected. This difference in x- and y-direction also explains the reduced
sensitivity of the 2D formulation of the MSE-DI algorithm, since data obtained in
both directions is used in that case.

4.5.2 Three stiffener structure with damage scenario III


The distributions of the damage index β ij and normalized damage index Zij for
damage scenario III in the three stiffener structure are respectively shown in
Figures 4.11 and 4.12. The 2D formulation of the MSE-DI algorithm, Figure 4.11(a),
presents the highest indices at the impact damaged area. This indicates that the
damaged area is detected and localized by the 2D MSE-DI algorithm. The results
also allow for a very rough estimation of the geometry of the damaged area,
although the damage index values at the damaged area are relatively small.
The normalized damage index Zij , described by Equation (4.10), provides a
statistical measure for outliers caused by the local damage. It indicates by how
many standard deviations σ the damage index β ij in each element deviates from
the mean value µ. The damage index β ij tends to show a normal distribution with
values close to one for the undamaged regions. Elements with a Zij value larger
than 2 (95% confidence interval) are generally classified as damaged. Figure 4.12(a)
4.5. Experimental results 75

1 1
y−coordinate [m]

y−coordinate [m]
0.2 0.5 0.2 0.5

0 0
0.1 0.1
−0.5 −0.5

0 −1 0 −1
0 0.1 0.2 0.3 0 0.1 0.2 0.3
x−coordinate [m] x−coordinate [m]

(a) Normalized mode shape curvature [-] in (b) Normalized mode shape curvature [-] in
x-direction of the pristine structure. x-direction of the damaged structure.

1 1
y−coordinate [m]

0.2 0.5 y−coordinate [m] 0.2 0.5

0 0
0.1 0.1
−0.5 −0.5

0 −1 0 −1
0 0.1 0.2 0.3 0 0.1 0.2 0.3
x−coordinate [m] x−coordinate [m]

(c) Normalized mode shape curvature [-] in (d) Normalized mode shape curvature [-] in
y-direction of the pristine structure. y-direction of the damaged structure.

. .
(n) (n)
Figure 4.10 Normalized mode shape curvatures ∂2 uz ∂x2 and ∂2 uz ∂y2 of the (a, c) pristine
( f n = 996.4 Hz) and the (b, d) damaged ( f n = 987.3 Hz) two stiffener structure for damage scenario II.

shows for the 2D formulation that the damage index values Zij at the damaged
region deviate up to 9 times the standard deviation σ from the mean value µ,
indicating the significance of the obtained result. Comparison of Zij results between
different cases is often rather difficult due to its dependency on the number of
elements selected in the MSE-DI algorithm.
In contrast to the obtained result, it must be noticed that the modal parameters
themselves are hardly affected by the impact damage. The average change in the
natural frequency was ∆ f n = 0.06% (maximum: ∆ f n = 0.31%) within the considered
frequency range. The mode shapes within this range provide an average MAC
value of 0.992 (minimum: MAC = 0.930), as shown in Figure 4.7(c).
In the previous section it was shown that, in the case of skin-stiffener damage,
the best results were obtained by considering the 1D formulation of the MSE-DI
algorithm in the direction of the stiffeners. The opposite behavior is shown for
the impact damage at the stiffener run-out as considered here. Figures 4.11(b)
76 Chapter 4. Damage identification in skin-stiffener structures based on curvatures

2 10

1.8 8
Damage Index β [−]

Damage Index Z [−]


3 1.6 10 6
2 5
1.4 4
1 1.2 1.2
1 0 1
0 0.8 1.2 0.8 2
0.6 0.6
0.2 0.4 0.2 0.4
0.2 x−coordinate [m] 1 0.2 x−coordinate [m] 0
0 0 0 0
y−coordinate [m] y−coordinate
[m]
(a) 2D formulation. (a) 2D formulation.
2 10

1.8 Damage Index Z [−] 8


Damage Index β [−]

3 1.6 10 6
2 5
1.4 4
1 1.2 1.2
1 0 1
0 0.8 1.2 0.8 2
0.6 0.6
0.3 0.4 0.3 0.4
0.2 x−coordinate [m] 1 0.2 x−coordinate [m] 0
0 0 0 0
y−coordinate [m] y−coordinate [m]
(b) 1D formulation in x-direction. (b) 1D formulation in x-direction.
2 10

1.8 8
Damage Index Z [−]
Damage Index β [−]

3 1.6 10 6
2 5
1.4 4
1 1.2 1.2
0.9 0 0.9
0 1.2 2
0.6 0.6
0.2 0.3 1 0.2 0.3 0
0 0 x−coordinate [m] 0 0 x−coordinate [m]
y−coordinate [m] y−coordinate [m]
(c) 1D formulation in y-direction. (c) 1D formulation in y-direction.
Damaged area Damaged area
Impact location Impact location

m m
.1 m .1 m
Case III t=6 Case III t=6
m m
.2 m .2 m
y x t=2 y x t=2
m m
1m 1m
t = 4. t = 4.

Figure 4.11 Damage index β ij distribution for Figure 4.12 Normalized damage index Zij for
damage scenario III in the three stiffener structure. damage scenario III in the three stiffener structure.
All modes within 250–1000 Hz and 60×20 MSE- All modes within 250–1000 Hz and 60×20 MSE-
DI elements are used in the analysis. DI elements are used in the analysis.
4.5. Experimental results 77

y−coordinate [m]

y−coordinate [m]
t = 4.14 mm t = 2.21 mm t = 6.08 mm t = 4.14 mm t = 2.21 mm t = 6.08 mm
0.4 1 0.4 1

0.2 0 0.2 0

0 −1 0 −1
0 0.5 1 0 0.5 1
x−coordinate [m] x−coordinate [m]
(a) Local torsion ( f n = 538.2 Hz). (b) Local bending ( f n = 617.6 Hz).

(n)
Figure 4.13 Normalized mode shapes uz ( x, y) of the pristine structure with (a) local torsion
( f n = 538.2 Hz) and (b) local bending ( f n = 617.6 Hz) deformation around the central stiffener. Both
mode shapes show a wavelength transition in x-direction at the thicker sections of the structure.

and 4.11(c) show the damage index β ij distributions obtained by applying the MSE-
DI algorithm in respectively x- and y-direction. Figures 4.12(b) and 4.12(c) represent
the same directions in the case of the normalized damage index Zij . Several
false positives appear for the x-direction, while the y-direction clearly predicts the
presence and the location of the impact damage correctly.
The most likely explanation for the reduced effect in x-direction is threefold. Firstly,
the damage is located at a region that already shows a geometrical discontinuity in
x-direction (end of stiffened section) for the pristine case. This causes a transition in
the local deformation patterns and reduces the possibility for the run-out damage to
be expressed in the measured curvatures in this direction. Secondly, the damage is
located at the thickest part of the skin. Generally, the thickest skin section showed a
relatively large wavelength of the mode shape in x-direction, analogous to lower
vibration modes, compared to the wavelength in x-direction at the mid-section
and the wavelength in y-direction. The reduced sensitivity for the x-direction is
understandable, since lower vibration modes are in general less affected by local
damage compared to higher vibration modes [18]. The differences in wavelength
are clearly depicted in Figures 4.13(a) and 4.13(b). Thirdly, a substantial amount
of mode shapes show a local torsional deformation around the central stiffener in
x-direction, similar to Figure 4.13(a). An earlier study on a composite T-beam [18]
revealed that the extent of the influence depends on the type of local deformation
around the damaged area. The mode shape in x-direction will be hardly affected if
the deformation around the damage represents local torsion in the yz-plane, since
the damage is located close to the local rotational axis. Figure 4.14 schematically
demonstrates the difference between local torsion and local bending behavior
around a skin-stiffener connection. Any potential nonlinear effect (e.g. opening
and closing behavior of the delamination) in the measured structural response is
not considered in the MSE-DI analysis, since it relies on a linear structural model.
The modal data obtained from the random excitation reflects a linear and stationary
representation of the system.
The more pronounced effect of the damage in the MSE-DI results for the y-direction
78 Chapter 4. Damage identification in skin-stiffener structures based on curvatures

y y y y
z z Interface z z Interface
failure failure

(a) Pristine. (b) Damaged. (c) Pristine. (d) Damaged.

Figure 4.14 Schematic representation of (a, b) local torsion and (c, d) local bending deformation around a
skin-stiffener connection for a (a, c) pristine and (b, d) damaged structure.

can be explained by the widening of the stiffener towards its end, as shown in
Figure 4.3(c). The delamination underneath the titanium insert locally reduces the
bending stiffness in y-direction. Moreover, the wavelengths in y-direction are for
many mode shapes relatively short compared to the wavelengths in x-direction at
the damaged area. As explained before, this causes a larger sensitivity to changes
by local damage.

4.6 Discussion
The derivation of the damage identification algorithm, presented in Section 4.3,
relies on a number of assumptions and simplifications. One of the most important
simplifications concerns the 1D and 2D formulation of the algorithm. The 2D
formulation assumes an isotropic description of the plate flexural rigidity. On the
other hand, the 1D formulation assumes that the flexural rigidity in one direction is
dominant and that all twisting and Poisson related effects are neglectable. In reality,
the stiffeners cause a local anisotropic design of the structures. As a consequence,
the validity of the 2D formulation is essentially only justified at the unstiffened skin
sections, whereas the 1D formulation provides the most realistic representation
at the stiffened skin sections. The results obtained for the two stiffener structure
complies with this decomposition, since the 1D formulation outperforms the 2D
formulation at the stiffeners. The damage of the three stiffener structure showed a
different behavior due to the fact that the damage is located at the stiffener run-out.
The results presented in Section 4.5 demonstrated the capabilities of the modal
curvature based approach to identify damage in composite skin-stiffener structures.
Several researchers [16, 23], however, question the suitability of modal data for
damage identification due to a relatively low sensitivity. The present work showed
that a refinement of this statement is justified. First of all, there is the trade-
off between the sensitivity to damage of the algorithm and its noise rejection
capability. A high sensitivity is desired to identify damage at an early stage. On
the other hand, it also increases the sensitivity to uncertainties, hence potentially
causing false damage identifications. Secondly, the minimum required sensitivity
of an approach is generally defined by the allowable damage limit (ADL) of
4.7. Conclusions 79

the structural component. Visual inspection of aircraft components according to


the BVID criteria [46] allows for the identification of macroscopic damage rather
than microscopic defects. It is therefore not inherently relevant to require sub-
macroscopic damage identification capabilities for all components. Finally, the
specific application of the methodology is also of importance. The potential of
the method can be utilized by applying the method in an adequate way. Analyzing
the relation between the damage location, the structural design and the dynamic
behavior revealed that the current method tends to be primarily effective for health
monitoring of the skin-stiffener connections.
In view of application there are still a number of factors to consider. High
resolution monitoring requires a relatively large number of measurement points.
This contradicts with the practical application of a vibration based technology,
since the number of sensors is often limited. There is a need to optimize
the amount and location of the data points used to describe the mode shapes
accurately. The results obtained for the two stiffener structure showed that a
reduced measurement grid, that only contains measurement points at the skin-
stiffener connection (i.e. 2×11 measurement points), would still allow for a
successful identification of the skin-stiffener failure. This approach provides an
effective way of monitoring the most critical areas (i.e. the ‘hotspots’) of a skin-
stiffener structure. The practical application of the approach also requires an
integrated sensor system, since a laser vibrometer is not feasible in most situations.
Distributed strain sensing technologies based on, for example, piezoelectric or fiber
optic sensing technologies provide a potential solution [25, 29]. The direct relation
between strain and curvature could also be beneficial to overcome the numerical
errors involved with the calculation of the curvatures, as discussed in various
studies [15, 25, 50, 57]. Another factor to consider is the sensitivity of the dynamic
parameters to environmental changes. Constant environmental conditions were
assumed during the experimental investigation presented. The numerical models
discussed in [42, 43] could be employed to account for the effects of changing
environmental conditions.

4.7 Conclusions

The aim of this chapter was to experimentally investigate the feasibility of a


vibration based damage identification approach to identify impact damage in
advanced composite skin-stiffener structures. Mode shape curvature changes, that
locally arise under dynamic loading of damaged structures, combined with the
modal strain energy damage index (MSE-DI) algorithm were utilized to identify
the defects. Special attention was paid to the effective application of this vibration
based methodology by investigating the relation between the damage location, the
structural design and the dynamic behavior.
80 Chapter 4. Damage identification in skin-stiffener structures based on curvatures

Damage
identification
method

Damage
Structure
scenario

Figure 4.15 The multidisciplinary framework associated with the development of a structural health
monitoring approach.

Experiments on two advanced composite skin-stiffener structures demonstrated the


capability of a modal curvatures based MSE-DI algorithm to detect and localize
barely visible impact damage at the skin-stiffener connections. The method allows
for a rough estimation of the geometry of the damaged area, even in a stiffened
structure with non-uniform skin thickness distribution. Each of the individual
formulations (1D in x-direction and 1D in y-direction) can be beneficial, depending
on the damage location with respect to the structural design.
The 1D formulation of the MSE-DI outperforms the 2D formulation in terms of
sensitivity, due to the anisotropic design of the specimen. The most effective results
were obtained by considering the 1D formulation underneath and in the direction
of the stiffeners for damage at the stiffener mid-section and perpendicular to the
stiffeners for damage at the stiffener run-out. These results are explained by the
location of the damage with respect to the structural design and to what extent this
affects the measured dynamic behavior in terms of curvature in x- and y-direction.
Impact damage at the skin in between stiffeners could not be identified by
considering mode shape curvatures and the MSE-DI method. The local failure
mechanism hardly affects the structural stiffness or mass. The methodology
requires a substantial change in bending stiffness or mass. This limits the effective
application of this method to health monitoring of skin-stiffener connections.
The method can be utilized as an approach for ‘hotspot’ monitoring by only
considering the measurement points underneath the skin-stiffener connections.
Failure primarily located in the skin tends to require a different SHM approach.
A combination of several damage identification technologies, operating at different
frequency levels, can be a potentially strong solution.
Finally, it can be concluded that SHM is a design related problem (see Figure 4.15).
It was shown that the design of a structure and the position of the damage affect
the effectiveness of the methodology. The development, selection and application
References 81

of damage identification algorithms is made-to-measure work and requires a


thorough physical understanding of the critical damage locations, the potential
failure mechanisms and their effect on the dynamic behavior. Therefore, this
process has to become an integral part of the structural design process. A scenario
based approach, based on the results obtained in this study, contributes to the
development of guidelines for effective health monitoring of advanced composite
skin-stiffener structures.

Acknowledgments
The authors kindly acknowledge the support of Fokker Aerostructures B.V.,
Hoogeveen, The Netherlands, for manufacturing the composite panels used in this
research. This work is carried out in the framework of the European project Clean
Sky, Eco-Design ITD (grant agreement number CSJU-GAM-ED-2008-001).

References
[1] C. Boller, F.-K. Chang, and Y. Fujino. Encyclopedia of structural health monitoring. John
Wiley & Sons, Ltd, Chichester, UK, 2009.
[2] K. Diamanti and C. Soutis. Structural health monitoring techniques for aircraft
composite structures. Progress in Aerospace Sciences, 46(8):342–352, 2010.
[3] C.-P. Fritzen and P. Kraemer. Self-diagnosis of smart structures based on dynamical
properties. Mechanical Systems and Signal Processing, 23(6):1830–1845, 2009.
[4] C.R. Farrar and D.A. Jauregui. Comparative study of damage identification algorithms
applied to a bridge: I. Experiment. Smart Materials and Structures, 7:704–719, 1998.
[5] C.R. Farrar and D.A. Jauregui. Comparative study of damage identification algorithms
applied to a bridge: II. Numerical study. Smart Materials and Structures, 7:720–731, 1998.
[6] J.J. Lee and C.B. Yun. Damage diagnosis of steel girder bridges using ambient
vibration data. Engineering Structures, 28(6):912–925, 2006.
[7] C.-P. Fritzen. Vibration-based structural health monitoring – Concepts and
applications. Key Engineering Materials, 293-294:3–18, 2005.
[8] N. Stubbs, S. Park, C. Sikorsky, and S. Choi. A global non-destructive damage
assessment methodology for civil engineering structures. International Journal of
Systems Science, 31(11):1361–1373, 2000.
[9] Z. Su, L. Ye, and Y. Lu. Guided lamb waves for identification of damage in composite
structures: a review. Journal of Sound and Vibration, 295(3-5):753–780, 2006.
[10] A. Raghavan and C.E.S. Cesnik. Review of guided-wave structural health monitoring.
The Shock and Vibration Digest, 39(2):91–114, 2007.
[11] J.-B. Ihn and F.-K. Chang. Pitch-catch active sensing methods in structural health
monitoring for aircraft structures. Structural Health Monitoring, 7(1):5–19, 2008.
[12] W.J. Staszewski, S. Mahzan, and R. Traynor. Health monitoring of aerospace composite
structures – Active and passive approach. Composites Science and Technology,
69(11-12):1678–1685, 2009.
82 Chapter 4. Damage identification in skin-stiffener structures based on curvatures

[13] G. Park and D.J. Inman. Structural health monitoring using piezoelectric impedance
measurements. Philosophical Transactions of the Royal Society A: Mathematical, Physical and
Engineering Sciences, 365(1851):373–392, 2007.
[14] D. Montalvão, N.M.M. Maia, and A.M.R. Ribeiro. A review of vibration-based
structural health monitoring with special emphasis on composite materials. Shock and
Vibration Digest, 38(4):295–326, 2006.
[15] W. Fan and P. Qiao. Vibration-based damage identification methods: a review and
comparative study. Structural Health Monitoring, 10(1):83–111, 2011.
[16] E.P. Carden and P. Fanning. Vibration based condition monitoring: a review. Structural
Health Monitoring, 3(4):355–377, 2004.
[17] M. Dilena and A. Morassi. Dynamic testing of a damaged bridge. Mechanical Systems
and Signal Processing, pages 1–23, 2011.
[18] T.H. Ooijevaar, R. Loendersloot, L.L. Warnet, A. de Boer, and R. Akkerman. Vibration
based structural health monitoring of a composite T-beam. Composite Structures,
92(9):2007–2015, 2010.
[19] R. Loendersloot, T.H. Ooijevaar, L.L. Warnet, A. de Boer, and R. Akkerman. Vibration
based structural health monitoring and the modal strain energy damage index
algorithm applied to a composite T-beam. In C.M.A. Vasques and J.D. Rodrigues,
editors, Vibration and Structural Acoustics Analysis: Current Research and Related
Technologies, Chapter 6, pages 121–150. Springer, 2011.
[20] N. Stubbs and C.R. Farrar. Field verification of a nondestructive damage localization
and severity estimation algorithm. In Proceedings of the 13th International Modal Analysis
Conference (IMAC XIII), pages 210–218, 1995.
[21] P.J. Cornwell, S.W. Doebling, and C.R. Farrar. Application of the strain energy damage
detection method to plate-like structures. Journal of Sound and Vibration, 224(2):359–374,
1999.
[22] A. Alvandi and C. Cremona. Assessment of vibration-based damage identification
techniques. Journal of Sound and Vibration, 292(1-2):179–202, 2006.
[23] J. Humar, A. Bagchi, and H. Xu. Performance of vibration-based techniques for the
identification of structural damage. Structural Health Monitoring, 5(3):215–241, 2006.
[24] J.-M. Ndambi, J. Vantomme, and K. Harri. Damage assessment in reinforced concrete
beams using eigenfrequencies and mode shape derivatives. Engineering Structures,
24(4):501–515, 2002.
[25] W. Lestari, P. Qiao, and S. Hanagud. Curvature mode shape-based damage assessment
of carbon/epoxy composite beams. Journal of Intelligent Material Systems and Structures,
18(3):189–208, 2007.
[26] Y.Y. Li. Hypersensitivity of strain-based indicators for structural damage identification:
a review. Mechanical Systems and Signal Processing, 24(3):653–664, 2010.
[27] P. Cawley and R.D. Adams. The location of defects in structures from measurements of
natural frequencies. Journal of Strain Analysis, 14(2):49–57, 1979.
[28] O.S. Salawu. Detection of structural damage through changes in frequency: a review.
Engineering Structures, 19(9):718–723, 1997.
[29] W.J.B. Grouve, L.L. Warnet, A. de Boer, R. Akkerman, and J. Vlekken. Delamination
detection with fibre Bragg gratings based on dynamic behaviour. Composites Science
and Technology, 68(12):2418–2424, 2008.
[30] E. Parloo, P. Guillaume, and M. van Overmeire. Damage assessment using mode shape
sensitivities. Mechanical Systems and Signal Processing, 17(3):499–518, 2003.
References 83

[31] L.M. Khoo, P.R. Mantena, and P. Jadhav. Structural damage assessment using vibration
modal analysis. Structural Health Monitoring, 3(2):177–194, 2004.
[32] A.K. Pandey and M. Biswas. Damage detection in structures using changes in
flexibility. Journal of Sound and Vibration, 169(1):3–17, 1994.
[33] T. Toksoy and A.E. Aktan. Bridge-condition assessment by modal flexibility.
Experimental Mechanics, 34(3):271–278, 1994.
[34] F.N. Catbas, M. Gul, and J.L. Burkett. Damage assessment using flexibility and
flexibility-based curvature for structural health monitoring. Smart Materials and
Structures, 17(1):1–12, 2008.
[35] A.K. Pandey, M. Biswas, and M.M. Samman. Damage detection from changes in
curvature mode shapes. Journal of Sound and Vibration, 145(2):321–332, 1991.
[36] M.M. Abdel Wahab and G. De Roeck. Damage detection in bridges using modal
curvatures: application to a real damage scenario. Journal of Sound and Vibration,
226(2):217–235, 1999.
[37] H. Hu and C. Wu. Development of scanning damage index for the damage detection
of plate structures using modal strain energy method. Mechanical Systems and Signal
Processing, 23(2):274–287, 2009.
[38] P. Qiao, K. Lu, W. Lestari, and J. Wang. Curvature mode shape-based damage
detection in composite laminated plates. Composite Structures, 80(3):409–428, 2007.
[39] H. Hu and J. Wang. Damage detection of a woven fabric composite laminate using a
modal strain energy method. Engineering Structures, 31(5):1042–1055, 2009.
[40] A.A. Cury, C.C. Borges, and F.S. Barbosa. A two-step technique for damage assessment
using numerical and experimental vibration data. Structural Health Monitoring,
10(4):417–428, 2010.
[41] S. Banerjee, F. Ricci, E. Monaco, and A. Mal. A wave propagation and vibration-based
approach for damage identification in structural components. Journal of Sound and
Vibration, 322(1-2):167–183, 2009.
[42] R. Loendersloot, T.H. Ooijevaar, L.L. Warnet, A. de Boer, and R. Akkerman. Vibration
based structural health monitoring of a composite plate with stiffeners. In Proceedings
of International Conference on Noise and Vibration Engineering (ISMA2010), pages 909–924,
2010.
[43] R. Loendersloot, T.H. Ooijevaar, A. de Boer, and R. Akkerman. Development of a
damage quantification model for composite skin-stiffener structures. In C. Boller and
H. Janocha, editors, New Trends in Smart Technologies, pages 99–108. Fraunhofer Verlag,
2013.
[44] A. Offringa, J. List, J. Teunissen, and H. Wiersma. Fiber reinforced thermoplastic butt
joint development. In Proceedings of the International SAMPE Conference, pages 1–16,
2008.
[45] A. Offringa, J.W. van Ingen, and A. Buitenhuis. Butt-joined, thermoplastic
stiffened-skin concept development. SAMPE Journal, 48(2):7–15, 2012.
[46] J. Baaran. Visual inspection of composite structures. Technical report, Institute of
Composite Structures and Adaptive Systems, DLR Braunschweig, 2009.
[47] T.A. Duffey, S.W. Doebling, C.R. Farrar, W.E. Baker, and W.H. Rhee. Vibration-based
damage identification in structures exhibiting axial and torsional response. Journal of
Vibration and Acoustics, 123(1):84, 2001.
[48] J. Stewart. Calculus: early transcendentals. 5th edition. Brooks/Cole - Thomson
84 Chapter 4. Damage identification in skin-stiffener structures based on curvatures

Learning, 2003.
[49] R.E. Shiffler. Maximum Z scores and outliers. The American Statistician, 42(1):79–80,
1988.
[50] A.P. Adewuyi, Z. Wu, and N.H.M. Kammrujaman Serker. Assessment of
vibration-based damage identification methods using displacement and distributed
strain measurements. Structural Health Monitoring, 8(6):443–461, 2009.
[51] B. Samali, J. Li, F.C. Choi, and K. Crews. Application of the damage index method for
plate-like structures to timber bridges. Structural Control and Health Monitoring, pages
1–23, 2009.
[52] E.S. Sazonov and P. Klinkhachorn. Optimal spatial sampling interval for damage
detection by curvature or strain energy mode shapes. Journal of Sound and Vibration,
285(4-5):783–801, 2005.
[53] B.J. Schwarz and M.H. Richardson. Experimental modal analysis. In CSI Reliability
Week, Orlando, FL, pages 1–12, 1999.
[54] M.H. Richardson and B.J. Schwarz. Modal parameter estimation from operating data.
Sound and Vibration, 37(1):28–39, 2003.
[55] M.H. Richardson and D.L. Formenti. Global curve fitting of frequency response
measurements using the rational fraction polynomial method. In 3rd International
Modal Analysis Conference, pages 1–8, 1985.
[56] R.J. Allemang. The modal assurance criterion – Twenty years of use and abuse. Journal
of Sound and Vibration, 1:14–21, 2003.
[57] C.S. Hamey, W. Lestari, P. Qiao, and G. Song. Experimental damage identification of
carbon/epoxy composite beams using curvature mode shapes. Structural Health
Monitoring, 3(4):333–353, 2004.
Chapter 5

Nonlinear dynamic behavior of an


impact damaged composite
skin-stiffener structure1

Abstract
One of the key issues in composite structures for aircraft applications
is the early identification of damage. Often, service induced damage
does not involve visible plastic deformation, but internal matrix
related damage. A wide range of technologies, comprising global
vibration and local wave propagation methods, can be employed
for health monitoring purposes. Traditional modal analysis based
methods are linear methods. The effectiveness of these methods
is sometimes limited since they rely on a stationary and linear
description of the system. The nonlinear interaction between a
low frequency wave field and a local impact induced damage in
a composite skin-stiffener structure is experimentally demonstrated
in this work. The different mechanisms linked to the distorted
waveforms are separated with the help of phase portraits. The
harmonic waveform distortions are concentrated at the damaged
region and increased for higher excitation amplitudes. It is shown
that linear damage identification methods are feasible for low
excitation amplitudes, but that the presence of nonlinear dynamic
effects cannot remain silent for higher amplitudes. Analyzing the
damage induced nonlinear effects can provide useful information
about the current state of the structure.

1 Reproduced from: T.H. Ooijevaar, M.D. Rogge, R. Loendersloot, L.L. Warnet, R. Akkerman,
T. Tinga, Nonlinear dynamic behavior of an impact damaged composite skin-stiffener structure.
In preparation for: Composite Science and Technology, 2014.

85
86 Chapter 5. Nonlinear dynamic behavior of an impact damaged skin-stiffener structure

5.1 Introduction
Composite skin-stiffener structures are a typical aerospace structural component
used to increase the bending stiffness of a component without a severe weight
penalty [1]. These skin-stiffener structures are considered to be safety-critical
components. Small defects at the skin-stiffener connection, caused by for example
impact, can significantly affect the performance of such a component. Often, the
damage does not involve visible plastic deformations, but is barely visible and
internal matrix related. This makes skin-stiffener connections a suitable candidate
for health monitoring.
A wide range of technologies, comprising global vibration and local wave
propagation methods, can be employed to monitor the integrity of structural
components [2–6]. The selection of the best technology is often a matter of
compromise. Traditional low frequency modal analysis based damage identification
methods are linear methods [3, 7]. Damage sensitive features are extracted by
fitting a mathematical model to the responses measured. Despite various successful
applications [8, 9], the effectiveness of these methods is sometimes limited since
they rely on a stationary and linear description of the system. A more realistic
damage description requires a physical understanding of the effect of the damage
on the dynamic behavior of the structure. The interaction between damage and an
acoustic wave field can yield dynamic phenomena that exhibit complicated material
(e.g. nonlinear elasticity) and geometrical (e.g. contact and friction) nonlinear
behavior [10, 11]. The symptoms related to these nonlinear phenomena can be
utilized for nonlinear system characterization and damage identification purposes.
The most commonly used nonlinear feature is probably the generation of higher
harmonics in the structural response. Selective imaging of these higher harmonics
was used by Krohn et al. [12] and Solodov et al. [13] to localize and characterize
small defects. Alternative nonlinear features are the generated sub-harmonics,
DC offsets, wave form distortions, frequency shifts, coherence functions, signal
modulations, etc. [3, 14]. Various researchers [15–17] state the potentially higher
sensitivity involved with these characteristics over the classical linear methods.
However, difficulties concern the experimental evidence and understanding of the
physical mechanisms related to the nonlinearities combined with the diversity of the
nonlinear mechanisms originating at different scales of the composite component
(micro-, meso- and macroscopic) [10]. Similar nonlinear effects can be a result
of different mechanisms, hence hampering the separation of the mechanisms.
Nonlinear elasticity and contact nonlinearities can, for example, both yield higher
harmonics in a dynamic response [14].
Researchers try to obtain a better understanding of the mechanisms involved with
the help of modeling. Della and Shu [18] reviewed models for the vibration of
delaminated structures, while Friswell and Penny [19] did the same for crack-like
damage. The complexity of the nonlinear effects introduced when non-bonded
5.2. Composite skin-stiffener structure 87

surfaces come into contact is one of the difficulties encountered in obtaining a


realistic model representation. Solodov et al. [11] describe two of the nonlinear
mechanisms involved with contact, the clapping and the friction mechanism.
The clapping nonlinearity is assumed to be concerned with the lack of stiffness
symmetry across the damage interface. One of the studies focusing on this
delamination induced mechanism is presented by Müller et al. [20, 21]. They
experimentally analyzed the vibro-impacting motion, including energy dissipation
effects, along the contact interface of the delamination to establish a finite element
description. As in many other investigations, their study is limited to an artificial
delamination, in this case in a simple aluminium beam.
This chapter focuses on a composite skin-stiffener-like structure featuring impact
induced damage. An experimental time domain analysis of the nonlinear dynamic
behavior introduced by the damage is presented. The impact induced damage
is located at the connection between skin and stiffener and mainly consists of
delamination damage. The objective is to gain an understanding of the interaction
between a dynamic wave field and the local skin-stiffener failure in order to support
the development of damage identification approaches that are able to characterize
skin-stiffener damage. A low frequency steady state harmonic shaker excitation
was used. Responses were measured by a laser vibrometer at multiple points
underneath the skin-stiffener connection. The dynamic mechanisms related to the
nonlinear dynamic effects in the harmonic waveforms are identified with the help
of phase portraits. Special attention is paid to the spatial effects and the influence of
the excitation parameters. The study presented in this chapter can lead to guidelines
for the applicability of the linear damage identification methods, but can also trigger
the development of enhanced damage identification methodologies.

5.2 Composite skin-stiffener structure


Thermoplastic composites provide specific opportunities due to their relatively high
toughness and unique processing capabilities [1]. The ambition for automated
manufacturing of large quantities of large composite structures has fueled the
development of new cost-reducing joining concepts. Recently, a new butt-joint
skin-stiffener concept was developed by Fokker Aerostructures [1]. This simplified
manufacturing concept uses preforms, which are connected by an injection molded
filler in a co-consolidation process.
The structure investigated in this study is a thermoplastic skin-stiffener section
made by Fokker Aerostructures according to their new joining concept. The
structure is schematically illustrated in Figure 5.1. Both the skin and the stiffener
are built from 16 individual plies of unidirectional carbon AS4D fiber reinforced
thermoplastic (PEKK) material with a [90/0]4,s lay-up. The filler is made from
PEKK and contains 20 wt% short carbon fibers.
88 Chapter 5. Nonlinear dynamic behavior of an impact damaged skin-stiffener structure

Bottom view
100 mm
25 25 25

20
Piezoelectric
diaphragm 2
1 3
2
32 Impact
40 mm

butt joint 1 location

360 mm
y
x Impact
z Node line X1
2.2

location
Piezoelectric Shaker
Point (x,y) =

180 mm
diaphragm 1 connection
(25,120) mm

y
x

20
Node line Y1 Y2 Y3

Figure 5.1 Three dimensional and bottom view of the composite skin-stiffener structure with a butt-joint
stiffener. The dimensions, the measurement points (dots), the position of the shaker and the impact location
are indicated.

The damage scenario analyzed in this study comprises an impact damage located
at the connection between skin and stiffener. This connection is considered as a
safety-critical area, because damage at this connection can have major implications
for the integrity of the entire component. Damage was introduced by utilizing
a falling weight impact device and applying a repeated impact up to 15 J. The
ultrasonic C-scan in Figure 5.2 reveals a relatively complex, but typical fiber
reinforced material related damage scenario. The damage consists of delamination
at the interface between the skin and stiffener accompanied by a limited amount
of damage between the first and second ply of the skin. Local delaminations were
also introduced underneath one of the supports that was used during the impact
testing.

5.3 Experimental work

The experimental process used to investigate the interaction between the dynamic
wave field and the local damage consists of two steps. The global dynamic behavior
of the pristine and damaged skin-stiffener structure is characterized in the first step.
These dynamic characteristics are used as an input for the second step to study the
nonlinear dynamic behavior of the damaged structure in detail. The current section
5.3. Experimental work 89

Supports (4×)
y
during impact
x
Stiffener 10 mm
and filler

79 mm

Impact Cracked
75 mm
location Skin
filler
Point (x,y) = delamination
(25,120) mm
85 mm

y First-ply Delamination Impact


x failure under stiffener location

Figure 5.2 Ultrasonic C-scan of the impact damaged skin-stiffener structure showing a complex
combination of failure mechanisms near the skin-stiffener interface.

addresses both steps, but starts with an introduction of the experimental set-up
used in this work.

5.3.1 Set-up and signal processing

The set-up and data acquisition systems used for all experiments are schematically
illustrated in Figure 5.3. Responses were measured before and after the impact
damage was introduced in the composite skin-stiffener structure. The structure
was freely suspended by an elastic wire, glued to the butt-joint, in order to isolate
the structure from environmental vibrations. An electromechanical shaker was
connected by a stringer and a force transducer to a corner of the structure. This
position is selected in order to excite a large number of vibrational modes. A
laser vibrometer, mounted on an x/y traverse system, measured the velocities at
different points at the skin of the structure. The measurement points are ordered
according to three lines in y-direction (‘Y1’, ‘Y2’, ‘Y3’) and one line in x-direction
(‘X1’), as shown in Figure 5.1. The location of line ‘X1’ matches with the center of
the damaged area. A data acquisition system, controlled by a Labview application,
was utilized to simultaneously send an excitation signal to the shaker and acquire
the force and velocity responses of the structure.

5.3.2 Initial global dynamic characterization

In the first set of experiments, the global dynamic behavior of the structure is
determined in terms of natural frequencies and the frequency dependent deflection
patterns, referred to as operational deflection shapes (ODS). An excitation signal
90 Chapter 5. Nonlinear dynamic behavior of an impact damaged skin-stiffener structure

11.
x 10
4 12.

Acceleration [mm/s ]
2
5

−5

4. 0.8 0.801 0.802


Time [s]
0.803 0.804

3.
10. 9.

5. 8.
7.
F(t)
v(t)
6.
y
2. z
1.

No. Description Hardware No. Description Hardware


1. Force transducer PCB 208C02 7. Traverse system x / y
2. Shaker B&K 4809 8. Laser: controller Polytec OFV-5000
3. Wired suspension 9. Condition amplifier PCB Model 480D06
4. Fixed frame 10. Power amplifier Dynakit Mark III 60W
5. Composite structure 11. Data acquisition NI PCI-6110E
6. Laser: sensor head Polytec OFV-505 12. Computer with Labview DAQ software

Figure 5.3 Experimental set-up.

composed of a linear sweep between 150 and 3050 Hz was sent to the shaker. The
force response was measured at the fixed excitation point. The velocities were
acquired at 51 (3×17) points, matching the node lines in y-direction according
to Figure 5.1. All signals were acquired for 2.62 seconds (217 samples) at a rate
of 50 kHz. A measurement at each point was repeated 10 times. The acquired
time responses were subsequently windowed, Fourier transformed and averaged to
obtain the power spectral densities. The auto-power spectral density S Fi Fi (ω ) and
cross-power spectral density S Fi v j (ω ) are used to calculate the mobility frequency
response functions HFi v j (ω ) between the fixed point of excitation i and the roving
measurement points j according to [22]:

S Fi v j (ω )
HFi v j (ω ) = (5.1)
S Fi Fi (ω )

Figure 5.4 shows the magnitude of the total set of frequency response functions
(FRFs) of the pristine structure. Sharp peaks correspond to the natural frequencies
of the structure.
The operational deflection shapes of the structure are extracted at the natural
5.3. Experimental work 91

2
10

|H | [(mm/s)/N] 0
Fv 10

−2
10

500 1000 1500 2000 2500 3000


Frequency [Hz]

Figure 5.4 The magnitude of the frequency response functions for all 51 (3 × 17) measurement points of
the pristine structure.

frequencies with the help of peak picking. At these frequencies, the complex
operational deflection shapes become predominantly real valued and are a close
approximation of the associated mode shapes [22]. Bending and torsion dominant
operational deflection shapes are distinguished by considering the dominant real
part of the operational deflection shapes, as shown in Figure 5.5. Table 5.1 provides
an overview of the natural frequencies of the first six bending and torsion shapes
for the pristine and damaged structure.
The present work concentrates on bending dominant operational deflection shapes
in the yz-plane. Earlier research [8] showed that skin-stiffener damage hardly affects

20 4
0.3 0.3 0.3 0.3
30
2 15
0.25 20 0.25 0.25 0.25
10 2
y−coordinate [m]

y−coordinate [m]

y−coordinate [m]

y−coordinate [m]

10
0.2 0.2 0 0.2 5 0.2
0
0.15 0.15 0.15 0 0.15 0
−10
−2 −5
0.1 −20 0.1 0.1 0.1
−10 −2
−30
0.05 0.05 −4 0.05 0.05
−40 −15
0 0 0 0 −4
0 0.05 0 0.05 0 0.05 0 0.05
x−coordinate [m] x−coordinate [m] x−coordinate [m] x−coordinate [m]
(a) ℜ( HFv ) in (b) ℑ( HFv ) in (c) ℜ( HFv ) in (d) ℑ( HFv ) in
[(mm/s)/N] at [(mm/s)/N] at [(mm/s)/N] at [(mm/s)/N] at
(4) (4) (4) (4)
f T = 662 Hz. f T = 662 Hz. f B = 1456 Hz. f B = 1456 Hz.

Figure 5.5 Real (a,c) and imaginary (b,d) part of the operational deflection shapes of the pristine structure
at the natural frequencies of the 4th torsion mode (a,b) and the 4th bending mode (c,d).
92 Chapter 5. Nonlinear dynamic behavior of an impact damaged skin-stiffener structure

Table 5.1 Natural frequencies of the 1st till the 6th bending and torsion modes of the pristine and damaged
(indicated by a tilde) structure.

Mode Bending Byz Torsion Ty


[Hz] f˜ [Hz] [Hz] f˜T [Hz]
(n) (n) (n) (n)
n fB B fT
1 956 949 - -
2 1040 10761 228 221
3 1241 1215 441 414
4 1456 1455 662 6641
5 1839 1833 1149 1119
6 2328 23401 1606 1586
1 A shaker replacement in between the measurements at the
pristine and damaged structure caused some frequencies
to increase slightly (while they were expected to decrease).
The type of deflection shape remained unchanged.

torsional deformation shapes around the y-axis in these kinds of structures due to
the small distance between the damage and the rotational axis.

5.3.3 Harmonic waveform distortion


In the second set of experiments, the interaction between the damage and the
dynamic deformation of the structure is studied by applying a single tone harmonic
excitation signal. The frequency of this signal corresponds to one of the bending
frequencies according to Table 5.1 and is varied in strength. The force and velocity
responses were simultaneously acquired at a sample rate of 1 MHz for 1.05 seconds
(220 samples). No averaging was applied. Only the steady state response is
considered in the analysis because it is independent of the initial conditions.
The structure vibrates at steady state after the transient (start-up) responses have
disappeared. A period of 0.8 seconds was revealed to be sufficiently long for these
transient effects to become negligible.
Figures 5.6(a) and 5.7(a) show steady state velocity responses obtained for,
respectively, the pristine and damaged structure by a forced excitation at a
(4) (4)
frequency matching the 4th bending mode ( f = 1456 Hz, f˜ = 1455 Hz) and
B B
5 different excitation amplitudes. The velocity responses are measured at location
(x,y) = (0.025,0.120) m according to the coordinate system presented in Figure 5.1.
This point is located at the skin near the skin-stiffener connection and matches the
damaged region in the case of the damaged structure. A low pass filter (zero-
phase finite impulse response) is used to remove high frequency measurement
noise from the responses. A cutoff frequency of 50 kHz is selected to maintain
the dominant frequency components. The accelerations, shown in Figures 5.6(b)
and 5.7(b), are calculated from the measured velocity responses by a central
difference approximation.
5.3. Experimental work 93

20
Velocity [mm/s]

Velocity [mm/s]
5
10
0 0
−10
−5
−20

0.8 0.801 0.802 0.803 0.804 0.8 0.801 0.802 0.803 0.804
Time [s] Time [s]
(a) Velocity. (a) Velocity.
4 5
x 10 x 10
4
Acceleration [mm/s ]

Acceleration [mm/s ]
2

2
5
2

0 0

−2
−5
−4
0.8 0.801 0.802 0.803 0.804 0.8 0.801 0.802 0.803 0.804
Time [s] Time [s]
(b) Acceleration. (b) Acceleration.

Figure 5.6 Velocity and acceleration responses Figure 5.7 Velocity and acceleration responses
measured at location (x,y) = (0.025,0.120) m of the measured at location (x,y) = (0.025,0.120) m of the
pristine structure for a forced excitation at the 4th damaged structure for a forced excitation at the 4th
(4) (4)
bending mode ( f B = 1456 Hz) and five excitation bending mode ( f˜B = 1455 Hz) and five excitation
amplitude levels. amplitude levels.

The velocity and acceleration responses of the pristine structure show nearly purely
harmonic oscillations for all excitation amplitudes, while the waveforms are clearly
distorted in the damaged case. The amount of harmonic waveform distortion
is most pronounced in the acceleration response and increases for higher shaker
excitation amplitudes.
The velocity and acceleration responses are combined to obtain the phase portraits
in Figure 5.8. Each trajectory represents a different excitation amplitude. The
concentric circles around an equilibrium for the pristine structure indicate that the
motion is periodic, stable and purely harmonic. The damaged structure also shows
periodic and stable behavior, but the pure harmonic motion is distorted by the skin-
stiffener damage. The motion approaches nearly fundamental harmonic behavior
for the lowest excitation amplitude, while the harmonic distortion increases with
increasing excitation amplitude. Higher harmonics are also generated in the case
of the higher excitation amplitudes. The damage causes the motion of the skin at
94 Chapter 5. Nonlinear dynamic behavior of an impact damaged skin-stiffener structure

4 5
x 10 x 10
4
Acceleration [mm/s ]

Acceleration [mm/s ]
2

2
5
2

0 0

−2
−5
−4
−5 0 5 −20 0 20
Velocity [mm/s] Velocity [mm/s]
(a) Pristine structure. (b) Damaged structure.

Figure 5.8 Phase portraits measured at location (x,y) = (0.025,0.120) m of the pristine and damaged
(4) (4)
structure for a forced excitation at the 4th bending mode ( f B = 1456 Hz, f˜B = 1455 Hz) and five excitation
amplitude levels. Each trajectory represents a different excitation amplitude.

the damaged region to behave nonlinearly for the bending mode considered. The
nonlinearities mainly occur at one side of the phase portrait while the rest of the
motion keeps behaving rather linearly with respect to the excitation amplitude.

5.3.4 Damage induced dynamic mechanisms

The skin at the damaged region behaves dynamically nonlinearly for the bending
mode considered. The physical mechanisms related to these nonlinear effects are
still unknown. The phase portrait revealed that the nonlinearities mainly occurred
at one side of the phase portrait, but did not link these nonlinear effects to physical
mechanisms. Therefore, the displacement, velocity and acceleration signals are
analyzed in more detail.
The skin-stiffener damage, although fairly complex, locally separates the skin from
the stiffener. This separation will potentially allow the damage to open under a
forced excitation at a bending dominant operational deflection shape. The relative
motion between skin and stiffener at the damaged area is required to verify the
opening of the delamination. The current set-up did not allow to directly measure
the deformation of the stiffener and to calculate the relative motion between the
skin and the stiffener. Alternatively, the motion of the stiffener at the damaged
region is approximated by analyzing the motion of the skin at multiple spatial
points underneath the skin-stiffener connection, i.e. 34 points at node line ‘Y2’
according to Figure 5.1. The motion of the stiffener at the undamaged interface
regions is equal to the motion of the skin, since both are connected. The high
flexural rigidity of the stiffener controls the skin motion at the undamaged regions.
At the damaged region, the skin and stiffener behave more independently. It is
assumed that the changes in the flexural motion of the stiffener due to the damage
5.3. Experimental work 95

y
x Damaged region
1 2 3 4 5 6 8 10 14 18 22 26 28 29 30 31 32 33 34

Original region Interpolated region Original region


Node (x,y) = (25,120) mm

Figure 5.9 Interpolated nodes to approximate the motion of the stiffener at the damaged region.

are minor compared to the changes in the motion of the skin. This is justified by the
fact that the stiffener has a significantly higher flexural rigidity (approximately 150
times) than the skin. As a result, the motion of the stiffener at the damaged region
is approximated by spatially interpolating the motion of the skin at the undamaged
regions on both sides of the damaged region. Figure 5.9 schematically shows the
interpolated region within the node set. A piecewise cubic spline interpolation,
based on nodes 1–5 and 26–34, is utilized to estimate the velocity and acceleration
patterns as a function of the position and the time for nodes 6–25.
Figure 5.10 shows the spatial velocity and acceleration distributions of the skin and
(4)
stiffener for the forced excitation at the 4th bending mode ( f˜ = 1455 Hz) and B
the highest excitation amplitude. The three snapshots match with three different
time instances according to the phase portrait presented in Figure 5.11. Despite the
fact that the interpolation only provides an approximation of the stiffener motion,
the snapshots reveal a significant local difference in the velocity and acceleration
amplitudes between the skin and the stiffener. A difference in speed between the
skin and the stiffener causes the delamination to open or close.
These physical mechanisms are linked to sections in the phase portrait with the
help of the associated displacement, velocity and acceleration waveforms for the
skin at the damaged region, as shown in Figure 5.11. A relative estimation of the
skin displacement is obtained by time integrating the velocity response. The skin
at location (x,y) = (0.025,0.120) m showed a 5 µm peak-to-peak displacement for
the highest excitation amplitude considered. A detailed analysis of the waveforms
makes it possible to distinguish three damage related dynamic mechanisms within
each oscillation:

Opening phase | A
The first mechanism, assigned to section A, represents the movement of
the skin towards the laser vibrometer. In this phase, the velocity of
the skin is positive, while the acceleration is getting negative up to the
point where the skin obtains its maximum positive displacement. The
snapshot in Figure 5.10(a) shows that the skin at the damaged region reaches
significantly higher velocities than those of the stiffener. This positive
difference indicates the opening of the delamination at the skin-stiffener
96 Chapter 5. Nonlinear dynamic behavior of an impact damaged skin-stiffener structure

5
x 10
4
Skin Skin

Acceleration [mm/s ]
2
20
Stiffener Stiffener
Velocity [mm/s]

2
10
0 0
−10
−2
−20 t = 0.80008 s t = 0.80008 s
−4
0 0.1 0.2 0.3 0 0.1 0.2 0.3
y−coordinate [m] y−coordinate [m]
(a) Velocity at t = t0 . (b) Acceleration at t = t0 .
5
x 10
4
Skin Acceleration [mm/s ]
2 Skin
20
Stiffener Stiffener
Velocity [mm/s]

2
10
0 0
−10
−2
−20 t = 0.80021 s t = 0.80021 s
−4
0 0.1 0.2 0.3 0 0.1 0.2 0.3
y−coordinate [m] y−coordinate [m]
(c) Velocity at t = t1 . (d) Acceleration at t = t1 .
5
x 10
4
Skin Skin
Acceleration [mm/s ]
2

20
Stiffener Stiffener
Velocity [mm/s]

2
10
0 0
−10
−2
−20 t = 0.80031 s t = 0.80031 s
−4
0 0.1 0.2 0.3 0 0.1 0.2 0.3
y−coordinate [m] y−coordinate [m]
(e) Velocity at t = t2 . (f) Acceleration at t = t2 .

y y
x x

0 0.1 0.2 0.3 0 0.1 0.2 0.3


Node line Y2 y−coordinate [m] Node line Y2 y−coordinate [m]
(g) C-scan of the damaged structure. (h) C-scan of the damaged structure.

Figure 5.10 Snapshot of the skin and (approximated) stiffener velocity and acceleration distribution at
(4)
node line ‘Y2’ for the forced excitation at f˜B = 1455 Hz and the highest excitation amplitude. The three
time instances match the time stamps presented in the phase portrait in Figure 5.11.
5.3. Experimental work 97

−3
x 10
Displacement [mm] 4 t1 t1

2 t2 t0 t2 t0

Composite
0 B C A B C A
Node line structure
Y2
t3 t3
Damaged area
0.7995 0.8 0.8005 0.801
Shaker Laser vibrometer
t0 t0
Velocity [mm/s]

20 v(t)
y
10 x
0
t1 t1 Fz z
t3 t3
−10 Force
transducer
−20
t2 t2
0.7995 0.8 0.8005 0.801
5 5
x 10 x 10
Acceleration [mm/s ]

Acceleration [mm/s ]
4 4
2

C
2 2
t3 t3 t3
t2 t0 t2 t0 t0
0 0 t2
B A
−2 −2
t1 t1 t1
−4 −4
0.7995 0.8 0.8005 0.801 −20 0 20
Time [s] Velocity [mm/s]

Figure 5.11 The displacement, velocity and acceleration responses measured at location
(x,y) = (0.025,0.120) m of the damaged structure for a single tone harmonic excitation at the 4th
(4)
bending mode ( f˜B = 1455 Hz) and the highest excitation amplitude. The capitals refer to the opening (A),
closing (B) and contact (C) phase of the skin stiffener failure. The time stamps t0 – t3 correspond to the
zero-crossings of the velocity and acceleration responses.

interface. The approximated flexural motion of the stiffener is smaller than


the motion of the much more compliant skin at the damaged region.
Closing phase | B
After the skin reached its maximum positive displacement towards the laser
vibrometer at t = t1 , the velocity of the skin starts to become negative. The
approximated velocity of the stiffener remained slightly positive according to
the snapshots given in Figure 5.10(c) and 5.10(e). The skin moves away from
the laser vibrometer, while the stiffener is moving (although with a relatively
low velocity) towards the skin. This part of the trajectory is referred to as the
98 Chapter 5. Nonlinear dynamic behavior of an impact damaged skin-stiffener structure

‘closing phase’ and is indicated by section B in Figure 5.11.


Contact phase | C
The last mechanism is related to the longest time interval where the most
severe distortions of the harmonic waveform are found. These distortions
are larger for the higher excitation amplitudes and only occur when the skin
and the stiffener are moving towards each other. This mechanism suddenly
generates higher harmonics, as clearly revealed in the acceleration signal.
At the same time, the velocity of the skin starts to decrease, i.e. slowing
down the motion of the skin when it approaches the stiffener. As a result,
the displacement waveform is somewhat rounded near a local minimum
compared to the sharpened peak around the positive displacement extrema.
The slowing down of the skin and the suddenly generated higher harmonic
distortions suggest that the skin at the damaged region is interacting with the
underlying structural interface. A clapping type of behavior can occur when
the skin interacts with the stiffener. It is observed that the velocity difference
between skin and stiffener causes a smooth change in the displacement
waveform. The interaction does not result in a hard impact-like contact
behavior as was found by Müller et al. [20, 21] for an artificial delamination
in a slender aluminium beam.

The opening, closing and contact mechanisms are repeated periodically. Analyzing
the different phases of this periodic motion revealed that the nonlinear dynamic
behavior of the skin is mainly introduced when the skin and stiffener are interacting
with each other. The motion of the skin at the damaged region is barely distorted
during the opening and closing phase.

5.3.5 Spatial effects


A spatial analysis is performed to study the influence of the damage on the dynamic
behavior of the skin-stiffener in more detail. The spatial dependency of the phase
portrait is analyzed first, followed by a study of the influence of the damage on
the local amplitude and phase. The time responses of the skin are considered at
multiple points according to the node lines presented in Figure 5.1.

Phase portrait

Figure 5.12 shows five phase portraits, each containing five trajectories, at different
positions along the damaged skin-stiffener interface for a single tone excitation at
the 4th bending mode, together with a snapshot of the velocities at t = t2 . The most
severe asymmetric distortions of the phase portraits are obtained for the skin at the
damaged region as illustrated by Figure 5.12(c). The motion of the skin is barely
affected at the undamaged regions and only shows some small distortions for the
5.3. Experimental work 99

C-scan of the structure

20

Velocity [mm/s]
10
0
−10
−20 t = 0.80031 s
0 0.1 0.2 0.3
y−coordinate [m]
Acceleration [mm/s ]
2

5 5 5 5 5
x 10 (a) x 10 (b) x 10 (c) x 10 (d) x 10 (e)
4 4 4 4 4
2 2 2 2 2
0 0 0 0 0
−2 −2 −2 −2 −2
−4 −4 −4 −4 −4
−20 0 20 −20 0 20 −20 0 20 −20 0 20 −20 0 20
Velocity [mm/s] Velocity [mm/s] Velocity [mm/s] Velocity [mm/s] Velocity [mm/s]

Figure 5.12 Phase portraits measured at different positions at node line ‘Y2’ of the damaged structure for a
(4)
forced excitation at f˜B = 1455 Hz and five excitation amplitude levels. Each trajectory represents a different
excitation amplitude.

highest excitation amplitudes. The damage interface acts as a nonlinear diffuser in


an otherwise predominantly linear system. The distortions reduce for positions at
a larger distance from the damaged region.

Amplitude and phase

Figures 5.13 and 5.14 show the velocity and acceleration time responses at node
line ‘Y2’ of the pristine and damage structure when excited in the 4th bending
mode. Note that the amplitudes are indicated by different color scales. The velocity
and acceleration amplitudes gradually vary around the nodal point at y = 0.23 m.
All points at the skin-stiffener connection move exactly in phase or out of phase
by 180 degrees for the pristine case. The responses change considerably in the
damaged case. The velocity distribution in Figure 5.14(a) shows a local increase in
the negative and positive amplitudes at the damaged region, while the acceleration
only shows a locally increased negative amplitude. The positive amplitude remains
in the same order of magnitude for all points at node line ‘Y2’. The interaction
between skin and stiffener restricts the acceleration of the skin to locally increase to
higher positive values.
The motion of the skin at the damaged region also shows a local phase shift with
100 Chapter 5. Nonlinear dynamic behavior of an impact damaged skin-stiffener structure

0.802 0.802
10 20
0.8015 5 0.8015
10

Time [s]
Time [s]

0.801 0 0.801 0
−5 −10
0.8005 0.8005
−10 −20
0.8 0.8
0 0.1 0.2 0.3 0 0.1 0.2 0.3
y−coordinate [m] y−coordinate [m]
(a) Velocity [mm/s]. (a) Velocity [mm/s].
5 ‘Closing’ ‘Distortion’ 5
x 10 x 10
0.802 0.802
1
0.8015 0.5 0.8015
0
Time [s]
Time [s]

0.801 0 0.801
−2
0.8005 −0.5 0.8005
−1
0.8 0.8 −4
0 0.1 0.2 0.3 0 0.1 0.2 0.3
y−coordinate [m] ‘Opening’ y−coordinate [m]

(b) Acceleration [mm/s2 ]. (b) Acceleration [mm/s2 ].

Node line Y2 Node line Y2


y y
x x

0 0.1 0.2 0.3 0 0.1 0.2 0.3


y−coordinate [m] y−coordinate [m]
(c) C-scan of the pristine structure. (c) C-scan of the damaged structure.

Figure 5.13 Velocity and acceleration responses Figure 5.14 Velocity and acceleration responses
measured at node line ‘Y2’ of the pristine structure measured at node line ‘Y2’ of the damaged structure
(4) (4)
for a forced excitation at f B = 1456 Hz and the for a forced excitation at f˜B = 1455 Hz and the
highest excitation amplitude. highest excitation amplitude.

respect to the undamaged regions of node line ‘Y2’. This phase shift is a result of
two effects. Firstly, the skin and the stiffener behave relatively independently at the
damaged region underneath the stiffener, while the bending motion of the skin at
the undamaged regions is controlled by the high flexural rigidity of the stiffener.
Secondly, the damage is located at a region where the motion of the skin at both
sides of the stiffener shows an antinode. These high amplitudes on both sides of the
stiffener are clearly illustrated by the operational deflection shape in Figure 5.5(c).
They drive the motion of the skin at the damaged region underneath the stiffener.
5.3. Experimental work 101

Skin−stiffener interface
0.802
200
0.8015 100

Time [s]
0.801 0

−100
0.8005
−200
0.8
C-scan of the structure First-ply failure

Measurement
points at
node line X1 0 0.025 0.05
x−coordinate [m]

Figure 5.15 Velocity response [mm/s] measured at node line ‘X1’ of the damaged structure for a forced
(4)
excitation at f˜B = 1455 Hz. The skin at the damaged skin-stiffener connection moves nearly in phase with
the motion of the skin on both sides of the stiffener.

As a consequence, the skin-stiffener damage opens and closes nearly in phase with
the motion of the skin on both sides of the stiffener, as demonstrated by the velocity
time response at node line ‘X1’ in Figure 5.15.

5.3.6 Influence of excitation frequency


Whether, when and to what extent the opening, closing and contact mechanisms
take place depends on the relative motion between the skin and the stiffener at
the damaged region. This relative motion is defined by the local deformation
patterns and is therefore damage location, excitation amplitude and frequency
dependent. The damage location could not be adjusted. The influence of the
excitation amplitude was briefly discussed in Section 5.3.3, leaving the influence
of the excitation frequency to be addressed in the current section.
When the structure is excited at the 5th bending mode, the damaged region matches
a nodal point in the deformation of the skin, hence resulting in minor distortions.
An excitation frequency equal to the 6th bending mode shows a clearly distorted
motion of the skin at the damaged region in Figure 5.16. The phase portraits show
some aberrant behavior compared to the phase portrait of the previously discussed
4th bending mode. A smaller amount of high frequency scattering is generated
during the contact phase, yielding an overall smoother phase portrait. Moreover,
the phase portrait reveals wrinkles around the zero crossings of the acceleration.
Analysis of the velocity and acceleration distributions at multiple points along the
skin-stiffener connection in time (i.e. Figures 5.13, 5.14, 5.17 and 5.18) revealed a
possible explanation for these differences. The skin at the damaged region shows
a ‘clapping’ type of behavior for the 4th bending mode, while the interaction in the
102 Chapter 5. Nonlinear dynamic behavior of an impact damaged skin-stiffener structure

5 6
x 10 Wrinkles x 10
Acceleration [mm/s ]
2

Acceleration [mm/s ]
1

2
1
0.5
0 0
−0.5 Wrinkles
−1
−1

−5 0 5 −50 0 50
Velocity [mm/s] Velocity [mm/s]
(a) Pristine. (b) Damaged.

Figure 5.16 Phase portraits measured at location (x,y) = (0.025,0.120) m of the pristine and damaged
(6) (6)
structure for a forced excitation at the 6th bending mode ( f B = 2328 Hz, f˜B = 2340 Hz) and five amplitude
levels. Each trajectory represents a different excitation amplitude.

case of the 6th bending mode is better described by a smoother ‘rolling’ motion.
This ‘rolling’ motion is demonstrated by the acceleration distribution of the skin in
Figure 5.18(b). The diagonal lines in the zoomed inset indicate traveling acceleration
waves from left to right at the delaminated region during the closing phase. This
propagating effect matches the wrinkles around the zero crossing of the acceleration
in the phase portrait.
The smooth ‘rolling’ motion is attributed to the position of the damage with respect
to the nodes and antinodes in the motion of the skin and the stiffener. Figure 5.19
shows that the damaged region is centered with respect to the nodal points on both
sides (l1 ≈ l2 ) in case of the 4th bending mode. In contrast, the 6th bending mode
shows an off-centered position (l1 < l2 ) of the damaged region, causing the rolling
motion in the acceleration of the skin.

5.4 Discussion
The present section provides a discussion of two aspects. Firstly, the underlying
physical phenomena associated with the nonlinear dynamic behavior are discussed.
Secondly, the higher harmonic components are addressed in more detail.

5.4.1 Underlying physical phenomena


A smooth change in the calculated displacement waveform demonstrated that the
interaction between skin and stiffener does not result in a hard impact-like contact
behavior as was found by Müller et al. [20, 21] for an artificial delamination in an
aluminum beam. However, a slowing down of the skin motion combined with the
5.4. Discussion 103

0.801 0.801
100
10
50
5

Time [s]
Time [s]

0.8005 0 0.8005 0
−5 −50
−10
−100
0.8 0.8
0 0.1 0.2 0.3 0 0.1 0.2 0.3
y−coordinate [m] y−coordinate [m]
(a) Velocity [mm/s]. (a) Velocity [mm/s].
5 6
x 10 x 10
0.801 2 0.801
1
1
Time [s]
Time [s]

0
0.8005 0 0.8005

−1 −1

0.8 −2 0.8 −2
0 0.1 0.2 0.3 0 0.1 0.2 0.3
y−coordinate [m] y−coordinate [m]

(b) Acceleration [mm/s2 ]. (b) Acceleration [mm/s2 ].

Node line Y2 Node line Y2


y y
x x

0 0.1 0.2 0.3 0 0.1 0.2 0.3


y−coordinate [m] y−coordinate [m]
(c) C-scan of the pristine structure. (c) C-scan of the damaged structure.

Figure 5.17 Velocity and acceleration responses Figure 5.18 Velocity and acceleration responses
measured at node line ‘Y2’ of the pristine structure measured at node line ‘Y2’ of the damaged structure
(6) (6)
for a forced excitation at f B = 2328 Hz and the for a forced excitation at f˜B = 2340 Hz and the
highest excitation amplitude. highest excitation amplitude.

generation of higher harmonic acceleration signals suggests a nonlinear interaction


between the skin and the underlying stiffener. This interaction causes the phase
portraits to distort at one side for the excitation amplitudes and frequencies
considered (see Figures 5.8(b), 5.11, 5.12(c), 5.16(b)).
Various physical phenomena can cause the measured nonlinear waveform distortions.
Although it was shown in Section 5.3.4 that the harmonic distortions at the damaged
region are mainly introduced when the skin and stiffener approach each other,
the experiments did not allow for a more detailed and unique separation of the
104 Chapter 5. Nonlinear dynamic behavior of an impact damaged skin-stiffener structure

100
Node Node Node Node
20
Velocity [mm/s]

l1 l2 l1 l2

Velocity [mm/s]
50
10
0 0
−10
−50
−20 t = 0.80031 s t = 0.80045 s
−100
C-scan of the structure C-scan of the structure

0 0.1 0.2 0.3 0 0.1 0.2 0.3


y−coordinate [m] y−coordinate [m]
(4) (6)
(a) f˜B = 1455 Hz. (b) f˜B = 2340 Hz.

Figure 5.19 Snapshots at time instance t = t2 of the velocity distributions at node line ‘Y2’ of the damaged
structure for the forced excitation at the 4th and 6th bending mode and five amplitude levels. The position of
the damaged region with respect to the associated nodal points is indicated by l 1 and l 2 .

underlying physical phenomena. The most obvious and commonly addressed


phenomenon is the nonlinearly varying stiffness across the damage interface [13].
Contact is just one of the phenomena. Lateral stress-stiffening of the skin and fiber
bridging across the damage interface could, although in the opposite direction, also
lead to asymmetric stiffness effects. The opening and closing motion of the skin-
stiffener damage also allows for ’breathing’ phenomena in the damaged zone, hence
potentially leading to viscous effects at the interface. Other potential mechanisms
involved are frictional, hysteretic or thermoelastic phenomena [23], all resulting in
energy dissipation.
A model representation of the damage between the skin and stiffener could help
to distinguish the most dominant underlying nonlinear phenomena. The number
of models presented in the literature [18, 19] illustrates, however, the complexity of
describing a realistic damage scenario.

5.4.2 Higher harmonics


The generated higher harmonics measured at the skin near the damaged skin-
stiffener interface are clearly presented by the Fourier transformed signals in
Figure 5.20 for the highest excitation amplitude. Analogous to the phase portraits,
the response obtained for an excitation at the 4th bending mode shows more severe
higher harmonic scattering than measured for the 6th bending mode. However,
this representation fails to unveil the timing of the waveform distortions. Time
responses composed of a variable number of higher harmonics are obtained by
applying a zero-phase finite impulse response (FIR) low pass filter with a high filter
5.4. Discussion 105

100 100
f f
2f 2f
Velocity [mm/s]

Velocity [mm/s]
4f
3f 3f 5f
1 1
4f
5f

0.01 0.01

0.0001 0.0001
0 0.5 1 1.5 2 0 0.5 1 1.5 2
Frequency [Hz] 4 Frequency [Hz] 4
x 10 x 10
(4) (6)
(a) f˜B = 1455 Hz. (b) f˜B = 2340 Hz.

Figure 5.20 Fourier transforms, indicating multiple higher harmonics, of the responses measured at
location (x,y) = (0.025,0.120) m of the damaged structure for a single tone harmonic excitation at the (a) 4th
and (b) 6th bending mode and the highest excitation amplitude.

order to the original time response. The associated phase portraits are shown in
Figure 5.21. The differences between the original and the filtered phase portraits
give an impression of the information that is discarded when the higher harmonics
are not included in the analysis. The largest differences are obtained for the
acceleration responses. This is caused by the fact that the majority of the distortions
is introduced around the zero-crossing of the velocity response.
An FRF extracted for the damaged structure is dominated by fundamental
harmonic components, but also contains higher harmonic information for a range of
excitation frequencies. A classical FRF based operational deflection shape analysis,
as was shown in Section 5.3.2, only considers the amplitude and phase at one
selected frequency. The higher harmonic information associated with this frequency

5 6
x 10 x 10
4
Acceleration [mm/s ]

Acceleration [mm/s ]
2

2 1

Original
0 0
Original f, 2f, 3f, 4f
f, 2f, 3f f, 2f, 3f
−2 −1
f, 2f f, 2f
−4 f f
−20 0 20 −50 0 50
Velocity [mm/s] Velocity [mm/s]
(4) (6)
(a) f˜B = 1455 Hz. (b) f˜B = 2340 Hz.

Figure 5.21 Phase portraits of the original time response and several low pass filtered time responses
comprising a selection of harmonics. The responses are measured at location (x,y) = (0.025,0.120) m for a
single tone harmonic excitation at the (a) 4th and (b) 6th bending mode and the highest excitation amplitude.
106 Chapter 5. Nonlinear dynamic behavior of an impact damaged skin-stiffener structure

is therefore not included in the operational deflection shape. Modal analysis based
damage detection approaches are subject to similar issues. They rely on a stationary
and linear representation of the system as was shown in Chapter 2. Nonlinear
effects in the structural response are not identified, while they can provide very
useful information about the health of the structure as shown in the present study.

5.5 Conclusions & future prospects


The interaction between damage and an acoustic wave field can cause dynamic
phenomena that exhibit complicated material and geometrical nonlinear behavior.
The work presented has investigated this nonlinear dynamic interaction for impact
induced damage in a composite skin-stiffener structure.
Analysis of the steady state responses resulting from a single tone harmonic
excitation at a number of predetermined natural frequencies revealed that, besides
amplitude and phase related effects, skin-stiffener damage can distort the harmonic
motion of the skin at the damaged region. A phase portrait representation between
velocity and acceleration at the damaged region showed that the motion approaches
nearly fundamental harmonic behavior for the lowest excitation amplitude, while
the harmonic distortion increases for higher excitation amplitudes.
The analysis of the periodic motion revealed three principle mechanisms within
each oscillation: opening of the damage, closing of the damage and a section where
the skin and stiffener are interacting with each other. The nonlinear dynamic
behavior of the skin at the damaged region is mainly introduced during the last
phase. Clear higher harmonics are generated, while the skin at the damaged region
keeps behaving almost purely harmonic during the opening and closing phase. It is
observed that the skin and stiffener are interacting, but that this interaction does not
result in a hard impact-like behavior. The results show that the damage interface
acts as a nonlinear diffuser, but the results did not allow for a unique separation of
the underlying physical phenomena.
It is concluded that the interaction between skin and stiffener depends on the
specific operational deflection shape of both components. The (‘clapping’ or
‘rolling’) interaction is an important parameter in the way the harmonic response
of the system is distorted. A spatial analysis revealed that the opening and closing
motion of the damage is mainly driven by the motion of the skin at both sides of
the stiffener.
The analysis of the results and the discussion demonstrated the complexity of the
interaction between damage and a harmonic wave field in a skin-stiffener structure
for a specific set of modes. From a more general point of view, the damage induced
nonlinear effects can provide very useful information about the current state of
the structure. Linear damage identification methods are feasible for low excitation
References 107

amplitudes, but the presence of nonlinear effects cannot remain silent for higher
amplitudes.
Additional research is required on the separation of the underlying physical
nonlinear phenomena. A model representation of the skin-stiffener damage can
help to distinguish the most dominant physical phenomena and to provide a
quantification of the experimental observations. Future research will also focus
on methods to embed the nonlinear dynamics in a structural health monitoring
method.

Acknowledgments
This material is based on work supported by National Aeronautics and Space
Administration, Langley Research Center under Research Cooperative Agreement
No. NNL09AA00A awarded to the National Institute of Aerospace.
The authors kindly acknowledge the support of Fokker Aerostructures B.V.,
Hoogeveen, The Netherlands, for manufacturing the composite structure used in
this research. This work is funded by the European research project Clean Sky,
Eco-Design ITD (grant agreement number CSJU-GAM-ED-2008-001).

References
[1] A. Offringa, J.W. van Ingen, and A. Buitenhuis. Butt-joined, thermoplastic
stiffened-skin concept development. SAMPE Journal, 48(2):7–15, 2012.
[2] C. Boller, F.-K. Chang, and Y. Fujino. Encyclopedia of structural health monitoring. John
Wiley & Sons, Ltd, Chichester, UK, 2009.
[3] K. Worden, C.R. Farrar, J. Haywood, and M. Todd. A review of nonlinear dynamics
applications to structural health monitoring. Structural Control and Health Monitoring,
15:540–567, 2008.
[4] D. Montalvão, N.M.M. Maia, and A.M.R. Ribeiro. A review of vibration-based
structural health monitoring with special emphasis on composite materials. Shock and
Vibration Digest, 38(4):295–326, 2006.
[5] K. Diamanti and C. Soutis. Structural health monitoring techniques for aircraft
composite structures. Progress in Aerospace Sciences, 46(8):342–352, 2010.
[6] A. Raghavan and C.E.S. Cesnik. Review of guided-wave structural health monitoring.
The Shock and Vibration Digest, 39(2):91–114, 2007.
[7] M.H. Richardson and B.J. Schwarz. Modal parameter estimation from operating data.
Sound and Vibration, 37(1):28–39, 2003.
[8] T.H. Ooijevaar, R. Loendersloot, L.L. Warnet, A. de Boer, and R. Akkerman. Vibration
based structural health monitoring of a composite T-beam. Composite Structures,
92(9):2007–2015, 2010.
[9] M. Dilena and A. Morassi. Dynamic testing of a damaged bridge. Mechanical Systems
and Signal Processing, pages 1–23, 2011.
108 Chapter 5. Nonlinear dynamic behavior of an impact damaged skin-stiffener structure

[10] A. Klepka, W.J. Staszewski, R.B. Jenal, M. Szwedo, T. Uhl, and J. Iwaniec. Nonlinear
acoustics for fatigue crack detection - Experimental investigations of vibro-acoustic
wave modulations. Structural Health Monitoring, 11(2):197–211, 2011.
[11] I. Solodov, N. Krohn, and G. Busse. CAN: an example of nonclassical acoustic
nonlinearity in solids. Ultrasonics, 40:621–625, 2002.
[12] N. Krohn, R. Stoessel, and G. Busse. Acoustic non-linearity for defect selective
imaging. Ultrasonics, 40(1-8):633–7, 2002.
[13] I. Solodov, D. Döring, and G. Busse. New opportunities for NDT using non-linear
interaction of elastic waves with defects. Journal of Mechanical Engineering,
57(03):169–182, 2011.
[14] K.-Y. Jhang. Nonlinear ultrasonic techniques for nondestructive assessment of micro
damage in material: a review. Precision Engineering and Manufacturing, 10(1):123–135,
2009.
[15] P.B. Nagy. Fatigue damage assessment by nonlinear ultrasonic materials
characterization. Ultrasonics, 36(1-5):375–381, 1998.
[16] K.E.-A. Van Den Abeele, P.A. Johnson, and A. Sutin. Nonlinear elastic wave
spectroscopy (NEWS) techniques to discern material damage, part I: nonlinear wave
modulation spectroscopy (NWMS). Research in Nondestructive Evaluation, 12:17–30,
2000.
[17] K.E.-A. Van Den Abeele, J. Carmeliet, J.A. Ten Cate, and P.A. Johnson. Nonlinear
elastic wave spectroscopy (NEWS) techniques to discern material damage, part II:
single-mode nonlinear resonance acoustic spectroscopy. Research in Nondestructive
Evaluation, 12:31–42, 2000.
[18] C.N. Della and D. Shu. Vibration of delaminated composite laminates: a review.
Applied Mechanics Reviews, 60(1-6):1–20, 2007.
[19] M.I. Friswell and J.E.T. Penny. Crack modeling for structural health monitoring.
Structural Health Monitoring, 1(2):139–148, 2002.
[20] I. Müller. Clapping in delaminated sandwich-beams due to forced oscillations.
Computational Mechanics, 39(2):113–126, 2005.
[21] I. Müller, A. Konyukhov, P. Vielsack, and K. Schweizerhof. Parameter estimation for
finite element analyses of stationary oscillations of a vibro-impacting system.
Engineering Structures, 27(2):191–201, 2005.
[22] M.H. Richardson. Is it a mode shape, or an operating deflection shape? Sound and
Vibration, pages 1–11, 1997.
[23] C.J. Pye and R.D. Adams. Detection of damage in fibre reinforced plastics using
thermal fields generated during resonant vibration. NDT International, 14(3):111–118,
1981.
Chapter 6

Vibro-acoustic modulation based


damage identification in a composite
skin-stiffener structure1

Abstract
Vibro-acoustic modulation based damage identification relies on
the modulation of a high-frequency carrier signal by an intenser
low-frequency vibration signal due to damage induced structural
nonlinearities. A time domain analysis of the vibro-acoustic
modulation phenomena was presented at multiple spatial locations
in an impact damaged composite skin-stiffener structure. The
instantaneous amplitude and frequency of the carrier velocity
response were extracted to analyze the intermodulations between
the two excitation signals. Increased amplitude modulations at the
damaged region revealed the presence, location and length of the
skin-stiffener damage. The damage hardly modulated the frequency
of the carrier response. This difference in behavior was attributed
to the nonlinear skin-stiffener interaction introduced by the periodic
opening and closing of the damage, according to earlier research
by authors on the same structure. A parametric study showed
that the amplitude and phase of the amplitude modulation are
dependent on the selected carrier excitation frequency, and hence
the high frequency wave field that is introduced. The present work
demonstrates the potential, but also the complexity of the vibro-
acoustic modulation based damage identification approach.

1 Reproduced from: T.H. Ooijevaar, M.D. Rogge, R. Loendersloot, L.L. Warnet, R. Akkerman,
T. Tinga, Vibro-acoustic modulation based damage identification in a composite skin-stiffener
structure. In preparation for: Structural Health Monitoring, 2014.

109
110 Chapter 6. Vibro-acoustic modulation based damage identification

6.1 Introduction

6.1.1 Background and motivation


Composite materials can exhibit complex types of damage, like transverse cracks
and delaminations. These damage scenarios can severely influence the structural
performance of a component, and hence tremendously decrease its service life.
Periodic inspections are required to ensure the integrity of a component during its
lifetime. A wide range of technologies can be employed for damage identification
purposes [1, 2]. Some of these technologies utilize the change in structural
dynamic characteristics as an indicator for damage. The low-frequency structural
vibration based approaches generally allow for a relatively easy interpretation of
the measured responses, have the ability to analyze complex structures and do
not necessarily require the structure to be readily accessible [3]. Drawbacks are,
however, the limited sensitivity compared to higher frequency approaches and
the number of required sensors in case the standing wave patterns need to be
described [4].
Traditional vibration based methods often rely on linear system descriptions [5],
while more recent work also features the nonlinear dynamic effects introduced
by local defects [6, 7]. Despite various successful applications of classical linear
methods [8], researchers state the potential benefits in terms of sensitivity [9–11] and
environmental robustness [12, 13] involved in the monitoring of nonlinear dynamic
effects. Frequently used nonlinear features for the identification of damage are the
generated sub-/higher harmonics in the structural response, waveform distortions,
frequency shifts as a function of the excitation amplitude, coherence functions,
etc. [5, 7]. All these methods rely on the principle that the level of nonlinearity
in the acoustic response of materials containing structural damage is greater than
in materials with no structural damage.

6.1.2 Vibro-acoustic modulation concept


A recently introduced nonlinear approach that has been shown to be sensitive
to the severity of damage in geometrically complex structures is the nonlinear
vibro-acoustic modulation (VAM) method [10, 14, 15]. This approach relies on the
modulation of a high-frequency ultrasonic wave (‘carrier’) by a more intense low-
frequency vibration (‘pump’). This modulation occurs in the presence of structural
nonlinearities, as schematically illustrated in Figure 6.1. Both excitation signals
are applied to the structure simultaneously. The pump signal with frequency f p
excites the structure and any nonlinearity, while the more sensitive carrier signal
at a frequency f c is used to analyze the potential intermodulation effects. A
linear structural response, illustrated in Figure 6.1(b), will be the combination of
its response to each of the signals individually. The associated Fourier spectrum
6.1. Introduction 111

(a)

Force [N]
Pump and carrier 0
excitation signals. f = 2 kHz
p
f = 50 kHz
c
−5
1 1.001 1.002
Time [s]
Linear Nonlinear
system Fp(t) + Fc(t) system
Linear response vl(t) vnl(t) Nonlinear response
Velocity v [mm/s]

Velocity v [mm/s]
4
2 2
(b) (c )
Superposition 0 0 Modulated
of responses. −2 response.
−2
−4
1 1.001 1.002 1 1.001 1.002
Time [s] Time [s]
Velocity |V| [mm/s]

Velocity |V| [mm/s]

3 f 3 f
p p (e)
2 2 Fundamental
(d) harmonics fp and fc,
Fundamental fc 2fp fc
1 1 higher harmonics
harmonics fc-fp fc+fp
3fp and sidebands.
fp and fc.
0 0
0 2 4 6 8 0 2 4 6 8
Frequency [Hz] 4 Frequency [Hz] 4
x 10 x 10
1
[mm/s]

Velocity vbp [mm/s]

1
(g)
(f) Bandpass Bandpass Modulated
bp

Carrier 0 filtered 0 filtered carrier


Velocity v

response. response.
−1
−1
1 1.001 1.002 1 1.001 1.002
Time [s] Time [s]

Figure 6.1 A schematic illustration of the vibro-acoustic modulation concept. Two harmonic excitation
signals (a) are simultaneously applied to the system. A linear system response (b, d, f) will be the superposition
of its response to each of the signals individually. The high-frequency system response (c) is modulated (e, g)
by the low-frequency pump excitation in case nonlinearities are present.

presented in Figure 6.1(d) will therefore only exhibit the two fundamental frequency
components. In the presence of nonlinearities, the system response does not
consist anymore of the linear superposition of the fundamental responses. The
two signals interact in such a way that the carrier signal is modulated by the
pump signal in amplitude, frequency, or both amplitude and frequency (see
Figure 6.1(c) ). The Fourier spectrum of the response signal reveals additional
components, higher harmonics (n f p and n f c , with n = 1, 2, 3, . . .) and sidebands
( f c + n f p and f c − n f p , with n = 1, 2, 3, . . .) around the high-frequency component,
as shown in Figure 6.1(e).
112 Chapter 6. Vibro-acoustic modulation based damage identification

Despite the fact that many studies showed that the amount of modulation is
correlated to the severity of the damage [16], the underlying physical phenomena
of the modulations are still not well understood. The main difficulty concerns the
diversity of the nonlinear phenomena originating at several material length scales
(micro-, meso- and macroscopic) combined with the fact that similar nonlinear
effects in an acoustic response can be a result of different physical phenomena.
This consequently hampers the separation of these underlying phenomena. Initially,
researchers claimed that the modulations are created by changes in the contact area
at the damage interface caused by the pump signal [17]. Studies by Zaitsev et al. [15]
and Klepka et al. [14], however, revealed that modulation can occur even when the
cracks are not being opened fully by the pump signal. They concluded that energy
dissipation based phenomena at the damaged area constitute the major mechanism
behind the modulations.
Regardless of the exact physical phenomena that cause the nonlinear modulations,
it has been demonstrated that a vibro-acoustic modulation based approach can be
employed to detect defects in a wide variety of structures ranging from aluminum
beams and plates [18, 19], bolt connections [20], to geometrically more complex
structures such as metallic automotive as well as aircraft subcomponents [10, 21].
The amount of work focused on composite materials is, however, rather limited [16,
22].
The majority of current vibro-acoustic modulation based investigations rely on
Fourier transformed responses obtained at one or a few measurement points [16,
22–24]. The sideband amplitudes of the carrier frequency are monitored to
detect and track the progression of damage. Limitations of this approach are,
firstly, the fact that the sidebands are represented by sharp peaks, while the
frequency resolution is often limited. Consequently, it is rather difficult to
obtain an accurate estimation of the amplitude of the sidebands. Secondly, the
sideband amplitudes can be the combined result of two phenomena, amplitude
and frequency modulations [25]. Analyzing only the sideband amplitudes is
insufficient to separate these effects and hence complicates the understanding of
the modulation phenomena. Alternative methods are required to overcome these
difficulties. Nearly all studies regarding vibro-acoustic modulation are limited to
one or a few measurement points. The spatial dependency of the modulations
has hardly been investigated. Consequently, these studies are mainly focused on
detection and quantification of defects, rather than on localizing them.

6.1.3 Objective and outline


The objective of this research is to analyze whether the vibro-acoustic modulation
method can be utilized to detect, localize and characterize impact damage
in a composite skin-stiffener structure. The work further aims to provide a
better understanding of the modulation phenomena by identifying the dominant
6.2. Theoretical description 113

parameters involved in the modulations. For these purposes, a time domain


analysis of the vibro-acoustic modulation phenomena at multiple spatial locations
is presented. This approach is an alternative to the commonly considered sideband
amplitudes at a single measurement point. The presented work intends, from a
broader perspective, to contribute to the development of enhanced methods for the
identification and characterization of damage in advanced composite structures.
This chapter starts with a theoretical description of the vibro-acoustic modulation
effect in the next section. This is followed by a description of the experimental
set-up and procedures in Section 6.3. The present work concentrates on the same
composite specimen and set-up as was utilized in earlier research by the present
authors [26]. Only the damaged structure was considered in the analysis. The
low-frequency pump excitation was applied by a shaker, while a piezoelectric
diaphragm was used to simultaneously apply a high-frequency carrier excitation.
A laser vibrometer was utilized to measure the steady state velocity responses at
multiple spatial locations on the structure. The modulated response signals were
decomposed to extract the instantaneous amplitude and frequency of the carrier
signal. These instantaneous characteristics were used to analyze the vibro-acoustic
modulation phenomena in more detail. The influence of the carrier excitation
signal on the measured modulations is then discussed, based on a parametric study.
Finally, the conclusions and recommendations for future work are presented.

6.2 Theoretical description

The current section provides a more thorough description of the wave modulation
phenomena. First, a basic theoretical description of the vibro-acoustic modulation
phenomena is presented. The development of the sideband components is
demonstrated based on a nonlinear single degree of freedom system. This is
followed by a description of the signal decomposition approach that is used to
extract the instantaneous characteristics of the modulated carrier response. The
last part essentially combines the two previous parts. A single degree of freedom
system is numerically solved for four different types of nonlinearities. The signal
decomposition approach is used to qualitatively analyze the system responses.

6.2.1 Two-tone forced vibration of a nonlinear system

A generalized quasi-harmonic nonlinear system [27] can be described by a


differential equation consisting of two parts, one part containing linear terms and a
second part containing nonlinear terms:

q̈ + ω02 q = −ǫ f (q, q̇) , (6.1)


114 Chapter 6. Vibro-acoustic modulation based damage identification

where q(t) is the displacement, ω0 the natural frequency, f (q, q̇) is a nonlinear
function. This nonlinear function is controlled by the parameter ǫ. For a weakly
nonlinear system, the parameter ǫ is considered relatively small such that the
solution of this system can be approximated by utilizing a perturbation technique
based on a power series [27]:

q(t) = q0 (t) + ǫq1 (t) + ǫ2 q2 (t) + . . . . (6.2)

An undamped system containing a quadratic nonlinearity subjected to a two-tone


forced excitation is considered in this study to illustrate the modulation effects in
the steady state response. A similar derivation can be followed for other types of
nonlinearities [27]. The quadratic nonlinear system is described by:

q̈ + ω02 q = −ǫq2 + Fp cos(ωp t + φp ) + Fc cos(ωc t + φc ) , (6.3)

with Fp and Fc the amplitudes, ωp = 2π f p and ωc = 2π f c the frequencies in radians


per second (where: ωp ≪ ωc ) and φp and φc the phases of respectively the pump
and carrier excitation signals. Assuming a solution in the form of Equation (6.2)
and separating orders of magnitude yields the following set of equations:

O(ǫ0 ) : q̈0 + ω02 q0 = Fp cos(ωp t + φp ) + Fc cos(ωc t + φc ) , (6.4)


1
O(ǫ ) : q̈1 + ω02 q1 = −q20 , (6.5)
2
O(ǫ ) : q̈2 + ω02 q2 = −2q0 q1 , (6.6)
... : ...

which can be solved recursively. The steady state solution of Equation (6.4) is:

q0 (t) = Gp cos(ωp t + φp ) + Gc cos(ωc t + φc ) , (6.7)

in which:
Fp Fc
Gp = and Gc = .
ω02 − ωp2 ω02 − ωc2

Substituting q0 (t) into Equation (6.5) and rearranging gives:

q̈1 + ω02 q1 = − 21 Gp cos(2ωp t + 2φp ) − 12 Gc cos(2ωc t + 2φc )


− Gp Gc cos((ωc − ωp )t + φc − φp ) − 21 Gp2
− Gp Gc cos((ωc + ωp )t + φc + φp ) − 21 Gc2 . (6.8)

Following from the nature of the excitation, the response q1 (t) consists of a linear
combination of harmonic components with frequencies equal to: 2ωp , 2ωc , ωc − ωp ,
ωc + ωp and a constant term. The fundamental harmonic frequency components
6.2. Theoretical description 115

ωp and ωc are added in the case of a first-order approximation of the solution q(t)
according to Equation (6.2). The carrier component ωc and its associated sidebands
(i.e. ωc − ωp , ωc + ωp ) are utilized for damage identification purposes. The system
response is therefore analyzed in a narrow frequency band around the carrier
frequency ωc . Consequently, the first-order steady state solution of q(t) reduces
to the narrow band response qbp (t) given by:

carrier lower sideband


z }| { z }| {
qbp (t) = Ac cos(ωc t + φc ) + Asb1 cos((ωc − ωp )t + φc − φp )
+ Asb2 cos((ωc + ωp )t + φc + φp ) , (6.9)
| {z }
upper sideband

in which:
−ǫGp Gc −ǫGp Gc
Ac = Gc , Asb1 = and Asb2 = .
ω02 (ωc
− ωp ) 2 ω02 (ωc
+ ωp ) 2

This analytical solution demonstrates the intermodulation of the carrier response


signal with the pump signal. The carrier response is only modulated in amplitude
and not in frequency in this case. The response of a system containing another type
of nonlinearity may show not only different combinations of harmonics, amplitudes
and phase angles, but can also exhibit both amplitude and frequency modulation
effects. A linear system (i.e. ǫ = 0) does not show amplitude or frequency
modulation effects. The response will, in that case, only consist of harmonic
components with frequencies equal to ωp and ωc .

6.2.2 Signal decomposition approach

The time domain analysis utilized in the present work is schematically illustrated
in Figure 6.2. This approach consists of two steps. In the first step, a zero-phase
bandpass FIR (i.e. finite impulse response) filter in a narrow frequency band
around the carrier frequency is applied to separate the carrier response and its
potential sidebands from the rest of the system response. Subsequently, the Hilbert
transform is applied to this filtered response vbp (t) to extract the instantaneous
amplitude Ainst (t) (i.e. the signal envelope) and frequency f inst (t). A description
of the Hilbert transform is presented in Appendix C. The instantaneous amplitude
and frequency reveal whether amplitude and/or frequency modulation effects are
present. These instantaneous characteristics are constant in the case of a purely
linear system response, but start to oscillate when nonlinearities are present, as
was illustrated in Figure 6.1. The peak-to-peak values of the oscillations in the
instantaneous amplitude and instantaneous frequency, represented by Ma and Mf
respectively, are used as a measure for the amount of modulation.
116 Chapter 6. Vibro-acoustic modulation based damage identification

Signal decomposition
Modulated Instantaneous Amplitude/
Actuation velocity Filtered velocity amplitude/ frequency
(2x single-tone) response response frequency modulation
Fp(t) v(t) vbp(t) Ainst(t) Ma
Bandpass Hilbert
Structure
Fc(t) filter, fc transform finst(t) Mf

Figure 6.2 Time domain signal analysis process of the response signal v(t) containing a bandpass filter
around the carrier frequency f c and the Hilbert transform to extract the instantaneous amplitude Ainst (t)
and frequency f inst (t). The peak-to-peak values of these instantaneous characteristics are used as a measure
for the amount of amplitude modulation Ma and frequency modulation Mf .

6.2.3 Nonlinear response characteristics


Numerical simulations were performed to qualitatively analyze the effect of
different types of nonlinearities on the steady state system response of a single
degree of freedom system. Response characteristics can be associated with the
different system representations. A damped nonlinear system (mass m = 1 kg,
stiffness k = 105 N/m and damping coefficient c = 20 Ns/m) was solved for
a forced excitation by two single-tone excitation signals (Fp = 10 N, Fc = 2 N,
ωp = 2π f p = 40π rad/s, ωc = 2π f c = 2000π rad/s, φp = φc = 0 rad). Four systems
with different types of nonlinearities were considered: quadratic and cubic with
respect to displacement q(t) and quadratic and cubic with respect to velocity q̇(t).
The nonlinear factor ǫ was used to control the strength of the nonlinear function,
according to Equation (6.1). The initial conditions were q(t) = 0 m and q̇(t) = 0 m/s
for all cases. Only the steady state system response was considered in this analysis.
The Fourier transformed velocity responses of a quadratic and cubic nonlinear
system with respect to displacement q(t) are shown in Figure 6.3. Although
both nonlinear responses show multiple higher harmonic components and carrier
sidebands, there is a notable difference in the arrangement of these additional
harmonics. The response of the system with the quadratic nonlinearity exhibits
higher harmonics and sidebands that deviate n f p (with n = 1, 2, 3, . . .) from the
fundamental frequencies f p and f c , while this is 2n f p in the case of the cubic
nonlinear system. This observation complies with the statements given in [10, 28]
and can be verified by utilizing the perturbation approach shown in Section 6.2.1.
The time domain decomposition approach, as described in Section 6.2.2, is used
to calculate the instantaneous amplitude and frequency of the carrier response
signal. The results obtained for the four different nonlinear systems are shown
in Figure 6.4. Each case contains three response signals corresponding to three
different values for nonlinearity factor ǫ. The systems containing a nonlinearity
in terms of displacement q(t) (see Figures 6.4(a) and 6.4(b) ) only show amplitude
modulated responses, while systems with a velocity q̇(t) related nonlinear term
(see Figures 6.4(c) and 6.4(d) ) exhibit both amplitude and frequency modulations.
6.2. Theoretical description 117

fp 2fp fp
fc 3fp fc
3fp 5fp
−5 −5

Velocity [m/s]
fc-nfp f +nf fc-2nfp f +2nf
Velocity [m/s]

10 c p 10 c p

−10 −10
10 10

0 200 400 600 800 1000 1200 0 200 400 600 800 1000 1200
Frequency [Hz] Frequency [Hz]

(a) Quadratic nonlinearity: f (q2 ). (b) Cubic nonlinearity: f (q3 )

Figure 6.3 Fourier transformed response signals for a system containing a (a) quadratic and (b) cubic
nonlinearity with respect to displacement q(t) showing sidebands at f c ± n f p and f c ± 2n f p (with
n = 1, 2, 3, . . .) respectively.

This result is analogous to the fact that the response of an undamped system is
in phase with its excitation, whereas the response of a viscously damped system
can exhibit a phase deviation. A nonlinearity in terms of velocity can consequently
result in varying phase deviations. Frequency modulations are obtained because
the instantaneous frequency is calculated as the time derivative of the instantaneous
phase as shown in Appendix C. Note that the frequency of the modulation effect
is different for the quadratic and cubic nonlinear systems. The responses of
the systems containing a quadratic nonlinearity show a dominant superimposed

−4 −4 −4 −4
x 10 x 10 x 10 x 10
3.2 3.195
Ainst (t) [m/s]

3.195 3.192
3.19 3.19
3.191
3.19
3.18
15 15.1 15.2 15 15.1 15.2 15 15.1 15.2 15 15.1 15.2
Time [s] Time [s] Time [s] Time [s]

1001 1001 1001 1001


finst(t) [Hz]

1000 1000 1000 1000

999 999 999 999


15 15.1 15.2 15 15.1 15.2 15 15.1 15.2 15 15.1 15.2
Time [s] Time [s] Time [s] Time [s]
2 3 .2 .3
(a) f( q ) (b) f( q ) (c) f ( q ) (d) f ( q )

Figure 6.4 Instantaneous amplitude Ainst (t) (top) and frequency f inst (t) (bottom) variations obtained
for a system containing four different types of nonlinearity: (a) quadratic and (b) cubic with respect to
displacement q(t) and (c) quadratic and (d) cubic with respect to velocity q̇(t). Each case contains three
response signals corresponding to three different values for nonlinearity factor ǫ.
118 Chapter 6. Vibro-acoustic modulation based damage identification

frequency equal to f p , while this is 2 f p for the cubic nonlinear systems. This
inherently corresponds to the frequency differences between the sidebands as
shown in Figure 6.3.

6.3 Experimental work


The current section introduces the experimental work. The composite skin-stiffener
structure will be presented in the first section, followed by a description of the
experimental set-up. The last subsection addresses the two-step experimental
procedure used to measure the vibro-acoustic modulation behavior. Only the
responses measured at the damaged structure are utilized in the analysis.

6.3.1 Composite skin-stiffener structure


The structure investigated in this study is a thermoplastic skin-stiffener section
made by Fokker Aerostructures according to the joining concept explained in [29].
Preformed skin and stiffener laminates are connected by an injection molded filler
in a co-consolidation process. The structure is schematically illustrated in Figure 6.5.
Both the skin and the stiffener are built from 16 individual plies of unidirectional
carbon AS4D fiber reinforced thermoplastic (PEKK) material with a [90/0]4,s lay-up.
The filler is made from PEKK and contains 20% short carbon fibers.
The damage analyzed in this study is located at the connection between skin and
stiffener. This connection is considered as a safety-critical area due to the high
importance of the connection for the structural integrity of the component. Damage
was introduced by utilizing a falling weight impact device and applying a repeated
impact up to 15 J. The ultrasonic C-scan in Figure 6.6 reveals damage consisting
of a delamination at the interface between the skin and stiffener accompanied by
a limited amount of damage between the first and second ply of the skin. Local
delaminations were also introduced underneath one of the supports that was used
during the impact testing.

6.3.2 Experimental set-up


The set-up and data acquisition systems used for all experiments are schematically
illustrated in Figure 6.7. The composite structure was freely suspended by an elastic
wire, glued to the butt joint, in order to isolate the structure from environmental
vibrations. An electromechanical shaker was connected by a stringer and a
force transducer to a corner of the structure. The shaker was used to introduce
the low-frequency pump waves, while a piezoelectric diaphragm was glued at
another corner of the structure to introduce the high-frequency carrier waves. A
6.3. Experimental work 119

Bottom view
100 mm
25 25 25

20
Piezoelectric
diaphragm 2
1 3
2
32 Impact
40 mm

butt joint 1 location

360 mm
y
x Impact
z
2.2

location
Piezoelectric Shaker
Point (x,y) =

180 mm
diaphragm 1 connection
(25,120) mm

y
x

20
Node line Y1 Y2 Y3

Figure 6.5 Three dimensional and bottom view of the composite skin-stiffener structure with a butt-joint
stiffener. The dimensions, the measurement points (dots) and the impact location are indicated.

Supports (4×)
y
during impact
x
Stiffener 10 mm
and filler

79 mm

Impact Cracked
75 mm

location Skin
filler
Point (x,y) = delamination
(25,120) mm
85 mm

y First-ply Delamination Impact


x failure under stiffener location

Figure 6.6 Ultrasonic C-scan of the impact damaged skin-stiffener structure showing a complex
combination of failure mechanisms near the skin-stiffener interface.
120 Chapter 6. Vibro-acoustic modulation based damage identification

13. 14.
5

Velocity [mm/s]
0

5. −5
0.8 0.802
Time [s]
0.804 0.806

12. 10. 4.
MODEL 240L POWER AMPLIFIER

11.
6. 9.
F(t) 8.

7. v(t)
1. y
3. z
2.

No. Description Hardware No. Description Hardware


1. Piezoelectric disk Murata 7BB-12-9 8. Traverse system x / y
2. Force transducer PCB 208C02 9. Laser: controller Polytec OFV-5000
3. Shaker B&K 4809 10. Condition amplifier PCB Model 480D06
4. Wired suspension 11. Power amplifier Dynakit Mark III 60W
5. Fixed frame 12. Power amplifier ENI model 240L 40W
6. Composite structure 13. Data acquisition NI PCI-6110E
7. Laser: sensor head Polytec OFV-505 14. Computer with Labview DAQ software

Figure 6.7 The experimental set-up.

laser vibrometer, mounted on a x/y traverse system, measured the velocities at


different points at the skin of the structure. The measurement points are ordered
according to three node lines in y-direction (‘Y1’, ‘Y2’, ‘Y3’), as shown Figure 6.5.
A data acquisition system, controlled by a Labview application, was utilized to
simultaneously send the excitation signals and to acquire the force and velocity
responses.

6.3.3 Experimental procedure

The experimental process consisted of two steps: a global dynamic characterization


of the skin-stiffener structure followed by the actual vibro-acoustic modulation
experiments. The first step is used to extract the interesting vibrational modes of
the structure. The frequencies of these modes are input for the second step to force
the structure to vibrate in a distinct deformation pattern during the vibro-acoustic
modulation experiment. Both steps are described in the present subsection.
6.3. Experimental work 121

Initial global dynamic characterization

In the first set of experiments, the global dynamic behavior of the structure is
determined in terms of natural frequencies and the frequency dependent deflection
patterns, referred to as operational deflection shapes (ODS). An excitation signal
composed of a linear sweep between 150 and 3050 Hz was sent to the shaker. The
force response was measured at the fixed excitation point. The velocities were
acquired at 51 (3×17) points, matching the node lines in Figure 6.5. All signals
were sent and acquired for 2.62 seconds at a rate of 50 kHz (217 samples). A
measurement at each point was repeated 10 times. These time responses were
subsequently windowed, Fourier transformed and averaged to obtain the power
spectral densities. The cross-power spectral density S Fi v j (ω ) is divided by the auto-
power spectral density of the input force S Fi Fi (ω ) to obtain the mobility frequency
response functions HFi v j (ω ) between the fixed point of excitation i and the roving
measurement points j according to [30].

Figure 6.8 shows the magnitude of the total set of frequency response functions of
the damaged structure. Sharp peaks correspond to the natural frequencies of the
structure. The bending and torsion dominant operational deflection shapes of the
structure are extracted at the natural frequencies with the help of peak picking.
The natural frequencies of the first six bending and torsion dominant operational
deflection shapes of the damaged structure are presented in Table 6.1. Note that
the first torsion mode was not measured. The natural frequency of this mode is
lower than the lowest frequency used in the excitation. Earlier research [8] showed
that bending dominant deformations in the yz-plane are more affected by skin-
stiffener damage than torsional deformations around the y-axis. The present work
is, therefore, focused on the vibro-acoustic modulation effects introduced by the
bending dominant operational deflection shapes.

2
10
|HFv| [(mm/s)/N]

0
10

−2
10

500 1000 1500 2000 2500 3000


Frequency [Hz]

Figure 6.8 The magnitude of the frequency response functions for all 51 (3 × 17) measurement points of
the damaged structure.
122 Chapter 6. Vibro-acoustic modulation based damage identification

Table 6.1 Natural frequencies of the first six bending and torsion modes of the damaged (indicated by a
tilde) structure.

Mode Bending Byz Torsion Ty


f˜ [Hz] f˜ [Hz]
(n) (n)
n B T
1 949 -
2 1076 221
3 1215 414
4 1455 664
5 1833 1119
6 2340 1586

Vibro-acoustic modulation experiments

The second set of experiments are the actual vibro-acoustic modulation measurements.
Two single-tone harmonic signals with different amplitude and frequency were
simultaneously sent to the shaker and the piezoelectric diaphragm. The pump
excitation signal introduced by the shaker had a frequency f p corresponding to
one of the bending frequencies listed in Table 6.1 and was varied in strength. The
piezoelectric diaphragm was subjected to a weaker 50 kHz ultrasonic excitation
signal. The amplitude of this carrier signal Ac was kept constant. The carrier
frequency f c is chosen in such a way that the carrier sideband components are well
separated from the higher harmonics of the pump frequency, but that the high-
frequency wave field can still be described by a limited number of measurement
points. The shaker excitation force and the velocity response were simultaneously
acquired at a sample rate of 1 MHz for 1.05 seconds (220 samples). No averaging was
applied. Only the steady state response is considered in the analysis. The structure
vibrates at steady state after the transient (start-up) response has disappeared.
A period of 0.8 seconds was revealed to be sufficiently long for these transient
effects to become negligible. The captured responses are processed by the time
domain signal decomposition approach as explained in Section 6.2.2. The results are
presented in the next section and are used to analyze the modulation phenomena
in the composite skin-stiffener structure in detail.

6.4 Experimental results and discussion

The experimental results are presented in the current section. The results obtained
for a single measurement point are discussed first, followed by the results obtained
for multiple points underneath the stiffener. Possible explanations for the measured
results are formulated and discussed. The effect of the underlying dynamic
behavior on the measured modulations is addressed in the last subsection.
6.4. Experimental results and discussion 123

6.4.1 Response decomposition

The steady state velocity response v(t) of the damaged structure caused by two
single-tone harmonic excitation signals is shown in Figure 6.9(a). The excitation
consists of an intense pump excitation at the 4th bending frequency ( f p = 1455 Hz)
and a weaker carrier excitation ( f c = 50 kHz). The velocity response v(t) is measured
at location (x,y) = (25,120) mm according to the coordinate system presented in
Figure 6.5. This point is located at the skin near the skin-stiffener connection and
coincides with the damaged region. Low and high frequency components can be
distinguished in the velocity response. The amplitude of the Fourier spectrum of
this response, i.e. |V ( f )|, is shown in Figure 6.9(b) and reveals higher harmonic
components n f p and multiple carrier sidebands f c ± n f p (with n = 1, 2, 3, . . .). These
additional components are indicative of a nonlinear response.
The low frequency part of the response matches the nonlinear response that was

20
fp
1
20 fc
15
|V(f)| [mm/s]
v(t) [mm/s]

0.5
0 10
0
fc 4 5 6
5 x 10
4
−20

0
0.8 0.801 0.802 0.803 0.804 0 2 4 6 8 10
Time [s] Frequency [Hz] 4
x 10
(a) Original velocity response v(t). (b) Fourier spectrum of the original velocity
response v(t).

5
v (t) [mm/s]

0
bp

−5
0.8 0.801 0.802 0.803 0.804
Time [s]
(c) Bandpass filtered velocity response vbp (t).

Figure 6.9 The (a) original and (c) bandpass filtered (40-60 kHz) velocity response measured at location
(x,y) = (25,120) mm of the damaged structure for two single-tone excitation signals with f p = 1455 Hz and
f c = 50 kHz and the highest shaker excitation amplitude. The Fourier spectrum of the original response (b)
shows the higher harmonic components n f p and carrier sidebands f c ± n f p (with n = 1, 2, 3, . . .).
124 Chapter 6. Vibro-acoustic modulation based damage identification

measured and analyzed in earlier research by authors [26] on the same specimen
using a single-tone harmonic shaker excitation. The distorted harmonic response
of the skin in that study was linked to the opening, closing and contact phase of
the skin-stiffener damage. In the present work, however, the high frequency of the
response is utilized. The carrier response and its dominant sideband components
are separated from the rest of the response by applying a bandpass filter within a
f c ± 10 kHz frequency range. The resulting narrow band velocity response vbp (t)
is depicted in Figure 6.9(c). The oscillating signal envelope clearly indicates that
amplitude modulation effects are present.
The nonlinear modulation effects in the bandpass filtered velocity response are
extracted by utilizing the Hilbert transform, as was described in Section 6.2.2. The
periodic behavior of the instantaneous amplitude Ainst (t) and frequency f inst (t)
variations in Figures 6.10(a) and 6.10(b) shows that both amplitude and frequency
modulation effects are present. The amount of modulation, represented by the

4
x 10
6 5.1
Ma
5.5 Mf
(t) [mm/s]

5.05
(t) [Hz]

5
5
4.5
inst
inst

4.95
A

3.5 4.9
0.8 0.801 0.802 0.803 0.804 0.8 0.801 0.802 0.803 0.804
Time [s] Time [s]
(a) Instantaneous amplitude Ainst (t). (b) Instantaneous frequency f inst (t).

0.8 250
1455 Hz
(f)| [mm/s]

200 1455 Hz
0.6
(f)| [Hz]

150
0.4
100
inst
inst

|f

0.2
|A

50

0 0
0 2000 4000 6000 8000 10000 0 2000 4000 6000 8000 10000
Frequency [Hz] Frequency [Hz]

(c) Fourier spectrum of the instantaneous (d) Fourier spectrum of the instantaneous
amplitude Ainst (t). frequency f inst (t).

Figure 6.10 The (a) instantaneous amplitude Ainst (t) and (b) instantaneous frequency f inst (t) of the
bandpass filtered velocity response vbp (t) measured at location (x,y) = (25,120) mm of the damaged structure
for five shaker excitation amplitudes. The associated Fourier spectra (c, d), illustrated for the highest shaker
excitation amplitude, reveal a dominant frequency equal to f p = 1455 Hz.
6.4. Experimental results and discussion 125

peak-to-peak value, increases for higher amplitude levels of the pump excitation.
The dominant frequency of the amplitude and frequency modulation phenomena
matches the pump excitation frequency f p = 1455 Hz, as shown in Figures 6.10(c)
and 6.10(d). According to the numerical results presented in Section 6.2.3, this
indicates that the nonlinearity in the structure tends to be more a quadratic rather
than a cubic type of nonlinearity.

6.4.2 Spatial results


The same vibro-acoustic measurement is performed at multiple locations at node
line ‘Y2’ and for multiple pump excitation frequencies. The velocity distribution
of the damaged structure when excited by a 1455 Hz pump excitation and a
50 kHz carrier excitation is shown in Figure 6.11(a). The local changes in amplitude
and phase of the low frequency part of the response are caused by the damaged
skin-stiffener interface, as was shown in earlier research by authors on the same
specimen [26]. The velocity distribution obtained after applying the f c ± 10 kHz
bandpass filter is depicted in Figure 6.11(b). The local higher amplitudes within
region ‘I’ and ‘III’ nicely correspond to the location and geometry of the skin-
stiffener damage. Note that modulations in the amplitude are visible at, for
example, y = 0.15 m. The lower amplitudes at the intermediate region ‘II’ are due to
the fact that the corresponding measurement points are located at a region where

I II III
0.802 0.802
10
20
Time [s]

Time [s]

0.801 0 0.801 0

−20 −10
0.8 0.8
0 0.1 0.2 0.3 0 0.1 0.2 0.3
y−coordinate [m] y−coordinate [m]
(a) Original velocity response v(t) [mm/s]. (b) Bandpass filtered velocity response
vbp (t) [mm/s].
y y
x x

0 0.1 0.2 0.3 0 0.1 0.2 0.3


Node line Y2 y−coordinate [m] Node line Y2 y−coordinate [m]

Figure 6.11 The (a) original and (b) bandpass filtered (40–60 kHz) velocity response measured at node
line ‘Y2’ of the damaged structure for two single-tone harmonic excitation signals with f p = 1455 Hz and
f c = 50 kHz and the highest shaker excitation amplitude.
126 Chapter 6. Vibro-acoustic modulation based damage identification

I II III (x,y) = (25,120) mm

x-coordinate [m]
y

(a) x

0 0.05 0.1 0.15 0.2 0.25 0.3


20
Velocity [mm/s]

t = 0.801005 s

(b) 0

−20
0 0.05 0.1 0.15 0.2 0.25 0.3

Velocity [mm/s]
0.8011
10
Time [s]

(c ) 0.8010 0

-10
0.8009
0 0.05 0.1 0.15 0.2 0.25 0.3
y-coordinate [m]

Figure 6.12 Comparison of the (a) damage location and geometry with the (b, c) bandpass filtered velocity
distribution vbp (t) of the skin at node line ‘Y2’.

the interface between the skin and the injection molded filler is not delaminated
over the entire width of the filler. A more detailed comparison between the damage
geometry and the bandpass filtered velocity response is presented in Figure 6.12.
The instantaneous amplitude and frequency of the bandpass filtered velocity
distribution are subsequently extracted according to the procedure described in
Section 6.2.2. The peak-to-peak values of the oscillations in these instantaneous
characteristics, represented by Ma and Mf , are utilized as a measure for the amount
of modulation. Figures 6.13(a) and 6.13(b) show the results obtained for a pump
excitation frequency equal to the 4th bending mode ( f p = 1455 Hz) and three pump
excitation amplitudes Fp . Figures 6.13(c) and 6.13(d) show the same results in the
case of the 6th bending mode ( f p = 2340 Hz). Although the structure exhibits
both amplitude and frequency modulation effects over the entire length of the
structure, increased amplitude modulation effects are measured at the damaged
area. The damage seems to hardly modulate the frequency of the carrier signal.
Figure 6.15 provides a closer look to the bandpass filtered time responses measured
at three successive points (i.e. y = 115 mm, y = 120 mm, and y = 125 mm) at the
damaged region. The dominant frequency of the amplitude modulation patterns
again matches the pump excitation frequency f p . These figures also reveal that
the phase of the amplitude modulation can significantly vary between the different
measurement points.
The frequency modulation in Figures 6.13(b) and 6.13(d) shows several higher
peaks. These peaks are attributed to a local low amplitude of the fundamental
6.4. Experimental results and discussion 127

3
x 10
6 4
Fp= 0.9 N Fp= 0.9 N
Fp= 5.4 N Fp= 5.4 N
Ma [mm/s]

Mf [Hz]
Fp= 9.4 N Fp= 9.4 N
2
2

0 0
0 0.1 0.2 0.3 0 0.1 0.2 0.3
y−coordinate [m] y−coordinate [m]
(a) Amplitude modulation for f p = 1455 Hz. (b) Frequency modulation for f p = 1455 Hz.
4
x 10
5 6
Fp= 0.9 N Fp= 0.9 N
4
Fp= 5.4 N Fp= 5.4 N
Ma [mm/s]

4
3 Fp= 9.4 N Mf [Hz] Fp= 9.4 N

2
2
1

0 0
0 0.1 0.2 0.3 0 0.1 0.2 0.3
y−coordinate y−coordinate [m]
(c) Amplitude modulation for f p = 2340 Hz. (d) Frequency modulation for f p = 2340 Hz.

y y
x x

0 0.1 0.2 0.3 0 0.1 0.2 0.3


Node line Y2 y−coordinate [m] Node line Y2 y−coordinate [m]

Figure 6.13 The (a, c) amplitude and (b, d) frequency modulation distributions of the carrier response
measured at node line ‘Y2’ of the damaged structure caused by two single-tone harmonic excitation signals
for three pump excitation amplitudes Fp . The results for f p = 1455 Hz are shown in (a, b), while (c, d) show
the results for f p = 2340 Hz. The carrier frequency f c was 50 kHz for all cases.

carrier response (e.g. near a nodal point) combined with a relatively large amount
of amplitude modulation, almost leading to over-modulation effects [31]. For
example, the peak at y = 250 mm in Figure 6.13(d). Figure 6.14 shows that
the associated bandpass filtered velocity response is almost fully modulated in
amplitude. Consequently, poor estimations of the instantaneous frequency are
obtained at the time instances where the envelope of the response approaches zero.

The underlying physical phenomena associated with wave modulations are


generally not well understood by researchers [19, 25]. Although the theory
128 Chapter 6. Vibro-acoustic modulation based damage identification

1
5
( t) [mm/s]

(t) [mm/s]
0.5
0 0
(x,y) = (25,125) mm
−0.5

bp
bp

−5 (x,y) = (25,120) mm

v
v

−1 (x,y) = (25,250) mm (x,y) = (25,115) mm

0.8 0.8005 0.801 0.8015 0.802 0.8 0.8005 0.801 0.8015 0.802
Time [s] Time [s]

Figure 6.14 Severe amplitude modulations of Figure 6.15 Phase differences in the amplitude
the bandpass filtered (40-60 kHz) velocity response modulations of the bandpass filtered (40-60 kHz)
measured at location (x,y) = (25,250) mm of the velocity response measured at three successive
damaged structure for two single-tone harmonic points on the damaged area for two single-tone
excitation signals with f p = 2340 Hz and harmonic excitation signals with f p = 2340 Hz
f c = 50 kHz and the highest shaker excitation and f c = 50 kHz and the highest shaker excitation
amplitude. amplitude.

(b) Open state


(a) Ainst(t) I II
y Stiffener
1 Ma z
I Skin
vbp [mm/s]

0.5
v(t) Damaged
0
(c) Closed state area
−0.5
−1 y
Carrier
II z v(t) wave
0 0.5 1 1.5 2
Time [s] −3 field
x 10

Figure 6.16 A simplified and schematic explanation of the (a) carrier modulation principle introduced by
the periodic opening and closing of the damage under an intense low-frequency pump excitation. In the (b)
open state the skin is free to vibrate, whereas the carrier amplitudes are compressed in the (c) closed situation.

presented in Section 6.2 provides an understanding of the relevant aspects involved,


finding a physical explanation for the measured modulation behavior is still rather
difficult. It was demonstrated by the present authors in [26] that the skin-stiffener
damage can open and close under a low frequency excitation, but also that the skin
can start to behave nonlinearly when the skin and stiffener are approaching each
other. The same excitation frequencies and amplitude were used for the pump
excitation in the vibro-acoustic experiments. Consequently, the nonlinear skin-
stiffener interaction is considered as the most likely reason why the modulation
effects develop.
Based on this finding, a possible explanation is formulated for the increased
6.4. Experimental results and discussion 129

amplitude modulations and the hardly modulated frequency of the carrier signal
at the damaged region. The applied low-frequency pump excitation signal will
change the damage interface conditions. In an ideal situation, the skin-stiffener
damage is completely opened and closed by the flexural vibrations of the skin, as
schematically illustrated in Figure 6.16. During the open state (b), the skin at the
damaged region is free to vibrate under the simultaneously applied high-frequency
carrier wave field. The amplitude of this wave field is, however, expected be to
smaller during the closed state (c). As a result, the high-frequency carrier wave
predominantly experiences periodic modulations of the amplitude, while the effects
on the frequency are expected to be minor.

The most obvious description of an opening and closing defect is given by a


varying stiffness across the damage interface [28]. There are two observations that
correspond to the behavior that is expected in the case of the periodic opening and
closing motion of the damage. Firstly, the fact that the frequency of the modulations
corresponds to the pump excitation frequency. The numerical results presented in
Section 6.2.3 revealed that this can indicate that the nonlinearity in the structure
tends to be more a quadratic type of nonlinearity, which is also expected for the
periodic opening and closing of a defect. Secondly, the numerical results also
suggest that the dominant amplitude modulation effects measured at the damaged
region indicate that the nonlinear dynamic behavior of the skin at the damaged
region is dominated by a nonlinearity in terms of displacement rather than velocity.
Both observations support the proposed explanation for the increased amplitude
modulations according to Figure 6.16. Note that the observations based on the
numerical results should be treated with some care; firstly, because the theoretical
analysis was limited to a single degree of freedom system. Secondly, because only
a few distinct types of nonlinearities were analyzed. Combinations of different
nonlinear terms can potentially result in the same nonlinear response.

Following the explanation illustrated in Figure 6.16, the amplitude modulations


are expected to be in phase with the low frequency part of the response. This
behavior is confirmed for the theoretical description of a single degree of freedom
system, as presented in Section 6.2. The experimental results, however, revealed
that the phase of the amplitude modulation can significantly vary between the
different measurement points. The complex geometry of the damage is expected to
play a role in the timing of the modulations to develop. The damage, depicted in
Figure 6.6, is geometrically complex, covers multiple levels and involves not only
the interface between skin and filler, but also the skin and filler itself. This is likely
to be one of the reasons, but may not be the only aspect, to explain the variations in
the phase. The next section addresses, therefore, a brief parametric study to analyze
the effect of the carrier excitation signal on the measured modulations.
130 Chapter 6. Vibro-acoustic modulation based damage identification

6.4.3 Underlying dynamic behavior


The variations in the phase of the amplitude modulation between the measurement
points at node line ‘Y2’ are studied in more detail by a parametric study. For this
purpose, the influence of the carrier excitation signal on the modulation behavior
was analyzed. Figures 6.17(a) and 6.17(b) show the effect of, respectively, the
carrier excitation amplitude Fc and the frequency f c on the instantaneous amplitude
Ainst (t) of the bandpass filtered response. The responses were measured at location
(x,y) = (25,120) mm of the damaged structure for a pump excitation frequency equal
to f p = 1455 Hz. An increase in the amplitude of the carrier excitation causes larger
amplitude modulations, whereas the phase of the modulation remains unchanged.
On the other hand, a change in the carrier frequency, shown in Figure 6.17(b),
affects both the amplitude as well as the phase of the amplitude modulation.
Even a small shift in the carrier frequency can have a significant effect on the
modulation behavior. These results imply that the amplitude and phase of the
amplitude modulation are highly dependent on the high frequency wave field that
is introduced by the selection of the carrier excitation frequency.
Yoder et al. [12] found that there is a strong correlation between the amplitude
of the carrier sidebands and the magnitude of the underlying spectral response
of the damaged structure. They state that the amount of modulation is directly
related to the high frequency underlying dynamic behavior of the structure. A
sideband will, for example, increase in amplitude if the frequency of the sideband
coincides with a resonance frequency of the structure. Although for a single degree
of freedom system, the theoretical description presented in Section 6.2 supports this
observation. The sideband amplitudes in Equation (6.9) are a direct function of the
natural frequency ω0 and will increase when the carrier frequency approaches this
resonance frequency.

6 10
(t) [mm/s]

(t) [mm/s]

4
5 f = 48 kHz
c
F
inst

inst

2 c f = 50 kHz
c
A

f = 52 kHz
c
0 0
0.8 0.802 0.804 0.806 0.8 0.802 0.804 0.806
Time [s] Time [s]
(a) Variable Fc = 0.01–0.07 V, f c = 50 kHz. (b) Variable f c = 48/50/52 kHz, Fc = 0.07 V.

Figure 6.17 The effect of the carrier excitation amplitude Fc and frequency f c on the instantaneous
amplitude Ainst (t). The responses are measured at location (x,y) = (25,120) mm of the damaged structure
for a pump excitation frequency equal to f p = 1455 Hz.
6.5. Conclusions & future prospects 131

|H | [(mm/s)/V]
1
(a)
0.1

hf
(x,y) = (25,120) mm
0.01

100
∠H [ ]
o

(b)
hf

0
−100
Velocity [mm/s]

f =48 kHz f =50 kHz f =52 kHz


10 c c c

(c) 1
0.1
0.01
0.001
4.4 4.6 4.8 5 5.2 5.4 5.6
Frequency [Hz] 4
x 10

Figure 6.18 The (a) magnitude and (b) phase of the frequency response functions within the 44–56 kHz
range measured at location (x,y) = (15,120) mm on node line ‘Y2’ of the damaged structure. The high
modal density causes the sidebands of the (c) Fourier spectrum of the modulated velocity response at the same
location to coincide with different modes, and hence introduces amplitude and phase deviations between the
left sideband, the carrier and right sideband components. The Fourier spectra associated with three carrier
excitation frequencies f c and a pump excitation frequency f p equal to 1455 Hz are presented.

In the experimental situation, the high modal density causes the frequencies of the
carrier and its sideband components to coincide with different modes, as illustrated
in Figure 6.18. This can lead to an unequal distribution of sideband amplitudes
around the carrier as was shown in Figure 6.9(b), but can also cause phase
deviations between the different harmonic components. Consequently, the spectral
effects introduced by the underlying dynamic behavior affect the amplitude and
phase of the amplitude modulation for each selected carrier frequency, as shown in
Figure 6.17(b), but also for each measurement point, as was shown in Figure 6.15.
The variations in the phase of the amplitude modulation distributions shown in the
previous section are inherently expected to be a result of the underlying spectral
response of the damaged structure, which are intensified by the high modal density
and the spatial variability of the complex wave field.

6.5 Conclusions & future prospects

The objective of this research was to analyze whether the vibro-acoustic modulation
method can be utilized to detect, localize and characterize impact damage
in a composite skin-stiffener structure. The work further aimed to obtain a
better understanding of the modulation phenomena. Vibro-acoustic modulation
132 Chapter 6. Vibro-acoustic modulation based damage identification

experiments were performed by simultaneously applying a low-frequency pump


signal and a higher frequency carrier excitation signal. The instantaneous
amplitude and frequency of the carrier velocity responses were extracted to analyze
the nonlinear intermodulations between the responses of these two excitation
signals in the time domain at multiple spatial locations.
Increased amplitude modulations at the damaged region revealed the presence,
location and length of the skin-stiffener damage. The damage hardly modulated the
frequency of the carrier response. This difference in behavior was attributed to the
nonlinear skin-stiffener interaction introduced by the periodic opening and closing
of the damage according to earlier research by authors on the same structure.
The dominant frequency of the amplitude modulation patterns matched the pump
excitation frequency, whereas the phase of the modulation behavior varied between
the different measurement points.
A parametric study shows that the amplitude and phase of the amplitude
modulation are dependent on the selection of the carrier excitation frequency, and
hence the high frequency wave field that is introduced. The high modal density
causes the frequency of each harmonic component of the modulated response (i.e.
carrier and sidebands) to coincide with a different dynamic system behavior. The
modulations in the carrier response signal will consequently vary in amplitude
and phase. The phase variations in the amplitude modulation distributions are
therefore expected to be the combined result of the high modal density and its
spatial variability.
The present study demonstrates the potential of the vibro-acoustic modulation
based damage identification approach in the time domain. A traditional analysis
purely based on sideband amplitudes in the frequency domain does not allow for
a separation between amplitude and frequency modulation effects. Moreover, the
sharp sideband peaks combined with a frequency resolution make it difficult to
obtain an accurate estimation of the amplitude. This inherently complicates the
understanding of the measured modulation phenomena.
Additional research is recommended on the selection of the ultrasonic carrier
frequency and its effect on the modulation phenomena. Methods to normalize
the modulated response signals could help to account for the spatial variations
in the ultrasonic wave field. Moreover, swept excitation signals as well as
propagating carrier signals (e.g. tone burst excitation signal) could enhance the
damage identification capabilities and the practical applicability of the vibro-
acoustic modulation method for real-time monitoring.

Acknowledgments
This material is based on work supported by National Aeronautics and Space
Administration, Langley Research Center under Research Cooperative Agreement
References 133

No. NNL09AA00A awarded to the National Institute of Aerospace.


The authors kindly acknowledge the support of Fokker Aerostructures B.V.,
Hoogeveen, The Netherlands, for manufacturing the composite structure used in
this research. This work is funded by the European research project Clean Sky,
Eco-Design ITD (grant agreement number CSJU-GAM-ED-2008-001).

References
[1] C. Boller, F.-K. Chang, and Y. Fujino. Encyclopedia of structural health monitoring. John
Wiley & Sons, Ltd, Chichester, UK, 2009.
[2] K. Diamanti and C. Soutis. Structural health monitoring techniques for aircraft
composite structures. Progress in Aerospace Sciences, 46(8):342–352, 2010.
[3] Y. Zou, L. Tong, and G.P. Steven. Vibration-based model-dependent damage
(delamination) identification and health monitoring for composite structures – A
review. Journal of Sound and Vibration, 230(2):357–378, 2000.
[4] C.-P. Fritzen and P. Kraemer. Self-diagnosis of smart structures based on dynamical
properties. Mechanical Systems and Signal Processing, 23(6):1830–1845, 2009.
[5] K. Worden, C.R. Farrar, J. Haywood, and M. Todd. A review of nonlinear dynamics
applications to structural health monitoring. Structural Control and Health Monitoring,
15:540–567, 2008.
[6] N. Krohn, R. Stoessel, and G. Busse. Acoustic non-linearity for defect selective
imaging. Ultrasonics, 40(1-8):633–7, 2002.
[7] K.-Y. Jhang. Nonlinear ultrasonic techniques for nondestructive assessment of micro
damage in material: a review. Precision Engineering and Manufacturing, 10(1):123–135,
2009.
[8] T.H. Ooijevaar, R. Loendersloot, L.L. Warnet, A. de Boer, and R. Akkerman. Vibration
based structural health monitoring of a composite T-beam. Composite Structures,
92(9):2007–2015, 2010.
[9] P.B. Nagy. Fatigue damage assessment by nonlinear ultrasonic materials
characterization. Ultrasonics, 36(1-5):375–381, 1998.
[10] K.E.-A. Van Den Abeele, P.A. Johnson, and A. Sutin. Nonlinear elastic wave
spectroscopy (NEWS) techniques to discern material damage, part I: nonlinear wave
modulation spectroscopy (NWMS). Research in Nondestructive Evaluation, 12:17–30,
2000.
[11] K.E.-A. Van Den Abeele, J. Carmeliet, J.A. Ten Cate, and P.A. Johnson. Nonlinear
elastic wave spectroscopy (NEWS) techniques to discern material damage, part II:
single-mode nonlinear resonance acoustic spectroscopy. Research in Nondestructive
Evaluation, 12:31–42, 2000.
[12] N.C. Yoder and D.E. Adams. Vibro-acoustic modulation utilizing a swept probing
signal for robust crack detection. Structural Health Monitoring, 9(3):257–267, 2010.
[13] S. Vanlanduit, E. Parloo, and P. Guillaume. Combined damage detection techniques.
Journal of Sound and Vibration, 266(4):815–831, 2003.
[14] A. Klepka, W.J. Staszewski, R.B. Jenal, M. Szwedo, T. Uhl, and J. Iwaniec. Nonlinear
acoustics for fatigue crack detection - Experimental investigations of vibro-acoustic
wave modulations. Structural Health Monitoring, 11(2):197–211, 2011.
134 Chapter 6. Vibro-acoustic modulation based damage identification

[15] V. Zaitsev and P. Sas. Nonlinear response of a weakly damaged metal sample: a
dissipative modulation mechanism of vibro-acoustic interaction. Journal of Vibration and
Control, 6(6):803–822, 2000.
[16] F. Aymerich and W.J. Staszewski. Impact damage detection in composite laminates
using nonlinear acoustics. Composites Part A: Applied Science and Manufacturing,
41(9):1084–1092, 2010.
[17] D. Donskoy, A. Sutin, and A. Ekimov. Nonlinear acoustic interaction on contact
interfaces and its use for nondestructive testing. NDT & E International, 34:231–238,
2001.
[18] D. Dutta, H. Sohn, K.A. Harries, and P. Rizzo. A nonlinear acoustic technique for crack
detection in metallic structures. Structural Health Monitoring, 8(3):251–262, 2009.
[19] Z. Parsons and W.J. Staszewski. Nonlinear acoustics with low-profile piezoceramic
excitation for crack detection in metallic structures. Smart Materials and Structures,
15(4):1110–1118, 2006.
[20] F. Amerini and M. Meo. Structural health monitoring of bolted joints using linear and
nonlinear acoustic/ultrasound methods. Structural Health Monitoring, 10(6):659–672,
2011.
[21] A. Zagrai, D. Donskoy, A. Chudnovsky, and E. Golovin. Micro- and macroscale
damage detection using the nonlinear acoustic vibro-modulation technique. Research in
Nondestructive Evaluation, 19(2):104–128, 2008.
[22] N.A. Chrysochoidis, A.K. Barouni, and D.A. Saravanos. Delamination detection in
composites using wave modulation spectroscopy with a novel active nonlinear
acousto-ultrasonic piezoelectric sensor. Journal of Intelligent Material Systems and
Structures, 22(18):2193–2206, 2011.
[23] M. Meo and G. Zumpano. Nonlinear elastic wave spectroscopy identification of impact
damage on a sandwich plate. Composite Structures, 71(3-4):469–474, 2005.
[24] U. Polimeno and M. Meo. Detecting barely visible impact damage detection on aircraft
composites structures. Composite Structures, 91(4):398–402, 2009.
[25] H.F. Hu, W.J. Staszewski, N.Q. Hu, R.B. Jenal, and G.J. Qin. Crack detection using
nonlinear acoustics and piezoceramic transducers – Instantaneous amplitude and
frequency analysis. Smart Materials and Structures, 19(6):065017, 2010.
[26] T.H. Ooijevaar, M.D. Rogge, R. Loendersloot, L.L. Warnet, R. Akkerman, and T. Tinga.
Nonlinear dynamic behavior of an impact damaged composite skin-stiffener structure.
In preparation for: Composite Science and Technology, 2014.
[27] L. Meirovitch. Fundamentals of vibrations. McGraw-Hill Higher Education, New York,
NY, 2001.
[28] I. Solodov, D. Döring, and G. Busse. New opportunities for NDT using non-linear
interaction of elastic waves with defects. Journal of Mechanical Engineering,
57(03):169–182, 2011.
[29] A. Offringa, J.W. van Ingen, and A. Buitenhuis. Butt-joined, thermoplastic
stiffened-skin concept development. SAMPE Journal, 48(2):7–15, 2012.
[30] M.H. Richardson. Is it a mode shape, or an operating deflection shape? Sound and
Vibration, pages 1–11, 1997.
[31] Q. Li and L. Atlas. Over-modulated AM-FM decomposition. In Franklin T. Luk, editor,
Proceedings of SPIE Volume 5559, Advanced Signal Processing Algorithms, Architectures, and
Implementations XIV, pages 172–183, 2004.
Chapter 7

Discussion

The work presented in previous chapters concerned the identification of damage in advanced
composite skin-stiffener structures by employing changes in the dynamic behavior. This
chapter broadens the discussions from the previous chapters with respect to the objective of
this research. Firstly, the design of a vibration based structural health monitoring approach
to identify damage in composite skin-stiffener structures is discussed. The results obtained in
previous chapters are combined to extract additional recommendations for the development
of a structural health monitoring approach. Secondly, a more general discussion on the
practical application of the vibration based methodologies is presented.

7.1 Design of an SHM strategy

A damage identification strategy aims to provide a high probability of detection


with a low number of false positive indications. Moreover, a high performance level,
according to the classification presented in Section 1.3.2, is usually desired. A large
number of damage identification methods are proposed in the literature, as was
shown in Chapter 2. These damage identification methods are often developed from
the standpoint of the algorithm rather than from the damage and the structure. The
performance of a damage identification approach is, however, highly dependent on
the actual structure and the damage scenario that is considered. The work presented
in this thesis was therefore focused on a more holistic approach by incorporating the
relations between the characteristics of the structure, the damage scenarios and the
damage identification method. Figure 7.1 schematically illustrates the associated
framework, which was introduced in Section 1.4. The characteristics of these three
components together define the performance of the structural health monitoring
strategy.

135
136 Chapter 7. Discussion

Damage
identification
method

Damage
Structure
scenario

Figure 7.1 The multidisciplinary framework for the design of a structural health monitoring strategy.

7.1.1 Scenario based design procedure

The implementation of the framework shown in Figure 7.1 is application specific.


Finding the most suitable damage identification method for a specific structure
and damage scenario is therefore far from straightforward and is often a matter
of compromise. This gives rise to the development of a tool that can be used to
design a damage identification strategy depending on the type of structure and the
potential threats. The results presented in this thesis provide an input for such a
tool dedicated to composite skin-stiffener structures. A scenario based procedure
for the design of a damage identification strategy is presented in Figure 7.2 and
comprises four steps.

Step 1: Potential damage scenarios

The first step is illustrated in Figure 7.2(b) and concerns the analysis of the potential
failure mechanisms and the potential damage scenarios of the structure that is
under investigation. Knowledge about the loading conditions, historic failure
data or structural models can provide useful information for this step. This step
inherently encourages the development of a structural health monitoring approach
to become an integral part of the design process of a component. This structural
design process can provide information about, for example, the safety-critical
damage locations or estimates of the damage tolerance of the component. Both
aspects set the requirements for the performance of the structural health monitoring
approach.
The need for repair of aircraft components is commonly assessed based on visual
inspection according to the BVID criterion [1]. Visual inspection allows for the
identification of macroscopic damage rather than microscopic defects. Three impact
induced damage scenarios were considered for the skin-stiffener structures that
7.1. Design of an SHM strategy 137

Sub-components
(a)

Structure

Step 1 Skin-stiffener Skin


Analysis potential (b)
failure mechanisms
and potential
damage scenarios. Stiffener run-out
Damage
scenario

Step 2 τ
Shear Open/close (c)
Understanding mode Delamination
mode
effect of damage on τ
dynamic behavior. τ
Delamination Stiffener Skin

Structural
dynamics
Delaminated area
2
Acceleration [-]

1
d u/dz [−]

Step 3 Mode Nonlinear


2

Selection damage shape 0 dynamic 0


(d)
curvature behavior
2

identification
-1
method. Pristine
−2
0 0.2 0.4 0.6 0.8 1 -1 0 1
Damaged Position z/L [−] Velocity [-]

Damage
identification
method
Level 1 Level 2 Level 3
Step 4 detection localization severity/extent
Performance (e)
Method A
classification for
each method. Method B
...

Structural health
monitoring strategy
for sub-component

Figure 7.2 The scenario based design procedure of a vibration based structural health monitoring strategy.
138 Chapter 7. Discussion

were used in the present work, i.e. damage at the skin, damage at the skin-
stiffener connection and damage at the stiffener run-out. These damage scenarios
are all classified as barely visible impact damage based on the dent depth. This
means that they may remain undetected according to the BVID criterion. The
skin related damage is indeed considered to be less safety-critical and is therefore
of lower importance. However, the damage at the skin-stiffener connection and
the skin-stiffener run-out can significantly affect the performance of the structure.
An evaluation of the potential consequences involved in the damage scenarios is
therefore required in the first step of the design procedure in order to assess the
importance of detection.

Step 2: Effect of the damage on the dynamic behavior

The second step is focused on the relation between the damage scenario and the
dynamic behavior of the component, as depicted in Figure 7.2(c). This step aims to
obtain a physical understanding of the effect of the potential damage scenarios
on the dynamic behavior of the component. Both analytical and experimental
analyses were used in the present research to investigate the effect of the damage
on the dynamic behavior. The damage at the skin-stiffener connection is usually
dominated by a delamination type of damage. The analysis of a delaminated beam
model in Chapter 2 revealed that the effect of a delamination is more pronounced in
the mode shape curvatures than in the mode shapes themselves. The mode shape
curvatures can show clear discontinuities at the boundaries of a delamination in the
case of flexural vibrations. The two driving phenomena behind these discontinuities
in the mode shape curvatures are, firstly, the change in shear stress distribution (see
Figure 7.3(a) ), and secondly, the potential opening and closing of the delaminated
segments (see Figure 7.3(b) ). The work presented in Chapter 5 experimentally
confirmed the opening and closing behavior and revealed severe nonlinear dynamic
effects associated with this motion. The study in Chapter 5 also showed that the
opening and closing motion of delaminated segments is magnified by the motion of
the skin at both sides of the stiffener, as schematically illustrated in Figure 7.3(c) ).
These geometrical effects can, on the one hand, result in larger discontinuities in
the mode shape curvature during the opening phase. On the other hand, they can
increase the nonlinear effects that are introduced during the contact phase.

Step 3: Selection damage identification method

The third step of the scenario based design procedure concentrates on the selection
of a suitable damage identification method, i.e. damage features and classifiers.
This step is schematically illustrated in Figure 7.2(d) and is an important step
in the process to obtain an effective damage identification strategy. The damage
identification method should match the observed changes in the dynamic behavior.
7.1. Design of an SHM strategy 139

x x x
z τ z y

τ
δ
τ δ

(a) Shear mode (b) Open/close mode (c) Geometrical effects

Figure 7.3 Dynamic phenomena related to delamination type of damage in skin-stiffener structures.

Moreover, the consequences involved with information reduction in the signal


processing need to be considered. Chapter 2 provides an elaborate description
of this process.
The physical understanding of how the potential damage scenarios affect the
dynamic behavior has led to the selection of two potentially strong indicators for
skin-stiffener damage, i.e. mode shape curvatures and the nonlinear dynamic
effects. Considering two different appearances of the damage in the dynamic
behavior of the structure is a potentially strong solution. Section 7.1.2 provides
a more elaborate discussion on the potential benefits involved in combining several
damage features (e.g. natural frequencies, modal damping, mode shape curvatures,
nonlinear behavior) in the damage identification process. The analysis of the
nonlinear dynamic effects set the requirements to the damage features that can be
used for damage characterization purposes. A linear dynamic system description,
as obtained by modal analysis based methods, is feasible for low excitation
amplitudes, while time domain or frequency domain methods are required for an
adequate description of the nonlinear dynamic effects introduced by the damage in
the case of the higher excitation amplitudes.

Step 4: Performance classification

The last step is a classification of the performance of each method in terms of its
detection, localization and characterization capabilities, as depicted by Figure 7.2(e).
This classification can be used to define a structural health monitoring strategy
according to the desired performance level.
The work presented in Chapters 3 and 4 experimentally demonstrated the feasibility
of using mode shape curvatures combined with the modal strain energy damage
index algorithm to detect, localize and roughly estimate the size of barely visible
impact damage in advanced composite skin-stiffener structures. The results showed
that the effective application of the method is limited to health monitoring of
140 Chapter 7. Discussion

skin-stiffener connections. Failure primarily located in the skin tends to require


a different damage identification approach. The nonlinear dynamic effects caused
by the skin-stiffener damage were analyzed as an alternative to the mode shape
curvature based method. Studies dedicated to nonlinear damage identification
methods, such as shown in Chapter 6, are therefore desired. The vibro-acoustic
modulation phenomenon can be utilized as an indicator for damage, but also
provides a first step towards a method that allows for the characterization of
the nonlinear dynamic behavior caused by the skin-stiffener damage. For some
applications linear damage identification techniques give adequate results and
nonlinear methods do not, while in other cases the opposite is true. To circumvent
this problem a combination of multiple damage identification approaches is usually
required.

7.1.2 Combination of approaches

A combination of several damage features and classifiers (see Figure 7.4(a))


or combining several damage identification techniques (see Figure 7.4(b)) can
potentially be strong structural health monitoring solutions to take advantage of
the strengths of each approach. Firstly, because they can increase the probability
of detection of a damage scenario by the diagnostic system. Secondly, they can
provide a system with a higher level of performance, according to the categorization
proposed in Section 1.3.2. Finally, a combination of approaches can also allow
for the identification of a wider range of damage scenarios. This is particularly
relevant for composite materials, considering the large diversity of potential damage
scenarios. For example, Chapter 4 revealed that the mode shape curvature based
approach worked well for impact damage at the skin-stiffener connection, but
was not able to identify the damage at the skin in between the two stiffeners.
The detection of defects primarily located in the skin tends to require a different
technique. A combination of a wave propagation based technique (i.e. acousto-
ultrasonics) to monitor the skin and the curvature based approach to monitor the
skin-stiffener connection utilizing a complementary sensor set can potentially be a
strong solution (see Figure 7.5).
Another way to benefit from the strengths of different methods is to apply them
in a more sequential manner, where one provides input for another. Typical
examples are the combination of a passive and active technique or a global and
local technique. In the former case, a passive technique, such as acoustic emission,
can serve as a trigger for the execution of an active approach, such as acousto-
ultrasonics. In the latter case, a global approach could detect the presence and
location of a structural anomaly, whereas the local method subsequently is used for
a more detailed localization and an estimation of the severity of the defect. Both
approaches would help to minimize the system requirements (e.g. computation
time, data transfer, power consumption) of the structural health monitoring system.
7.2. Application of vibration based SHM 141

Diagnostics

Statistical
classifier I
Feature
Actuation Response extractor I
Statistical
classifier II
Sensor Current state
Structure
system of the structure
Feature Statistical
extractor II classifier
Damage

(a) Combining different damage sensitive features (e.g. natural frequency, mode shape curvatures,
nonlinear dynamic behavior) and classifiers.

Technique I
Actuation Sensor Feature Statistical
system extractor classifier

Current state
Structure Technique II of the structure
Actuation
Sensor Feature Statistical
system extractor classifier

Damage
(b) Combining different damage identification techniques (e.g. structural vibration based methods,
acoustic emission, acousto-ultrasonics).

Figure 7.4 A schematic illustration of the diagnostic part of a structural health monitoring process showing
two ways to combine multiple damage identification methods: (a) for a combination of different damage
sensitive features and classifiers.

7.2 Application of vibration based SHM

The results presented in this thesis contribute towards obtaining a higher level of
maturity in the structural health monitoring technologies. However, several issues
need to be considered before field application of the structural health monitoring
technologies is possible. Five major technology gaps were described in Section 1.3.4:

• Complex composite structures


• Selection damage feature and classifier
• High performance level
• Integrated sensors and network
• Operational and environmental variability
142 Chapter 7. Discussion

Sensor
Skin-stiffener Skin
damage damage

y z x
Actuator

Figure 7.5 The skin-stiffener structure that was analyzed in Chapter 4. A combination of damage
identification techniques (e.g. structural vibration based methods, acousto-ultrasonics) is required to identify
the skin and the skin-stiffener interface damage.

The first three challenges were addressed in the present research. The extension
to more complex composite structures was the main focus of the work presented
in Chapters 3 and 4. Chapters 5 and 6 provide a first step towards a higher
performance level by characterizing the nonlinear dynamic behavior introduced by
the damage. The selection of appropriate damage features and classifiers was one
of the main topics addressed in this thesis. The scenario based design procedure,
as described in the previous section, showed itself to be an important outcome
and provides a useful guideline for the development of a suitable structural health
monitoring solution for safety-critical structures. The remaining two challenges, i.e.
the integrated sensing and the environmental effects, are discussed in the remainder
of this chapter.

7.2.1 Integrated sensing


The studies presented in this thesis rely on laser vibrometer measurements. These
measurement systems are less suitable for on-line health monitoring in, for example,
aircraft. An on-line health monitoring system requires permanently attached or
embedded sensor systems.
Sensor systems based on, for example, optical fiber Bragg gratings (FBGs) [2], can
have a positive influence on the development of an applicable health monitoring
system. Fiber Bragg gratings possess some advantages over conventional sensing
systems. The ability of multiplexing, i.e. having multiple strain sensors on one
fiber, offers great potential. A network of gratings provides opportunities to map
the strain field in a structure during vibration. Other advantages can be found
in the capabilities to withstand harsh environments. The practical application of
fiber Bragg gratings also encounters some restrictions, like the rather low signal-
7.2. Application of vibration based SHM 143

to-noise ratio as was found by Grouve et al. [3]. Despite this limitation, successful
applications are found for low-frequency vibration sensing [4, 5] as well as high-
frequency sensing of propagating waves [6]. New developments of optical fiber
distributed sensing systems provide the ability for strain sensing along the optical
fiber with an extremely high spatial resolution [7]. A low sample rate, however,
limits the current application of this technology to quasi-static measurements.
Alternatively, piezoelectric elements are widely used for actuation and sensing
purposes in health monitoring [8–10]. These piezoelectric elements have the ability
to convert mechanical energy into electrical energy, and vice versa [11]. They utilize
the direct piezoelectric effect to sense structural deformations and the inverse effect
to actuate the structure. One of the main advantages of piezoelectric elements is
that they can be used over a broad frequency range. This makes them applicable for
nearly all dynamics based damage identification techniques, i.e. structural vibration
based methods [12], electromechanical-impedance based techniques [13], acoustic-
emission [14] as well as acousto-ultrasonics [8]. Other advantages include the cost
effectiveness, as well as the ability to harvest energy [15].
Distributed strain sensing based on fiber optic or piezoelectric sensors is potentially
advantageous for the mode shape curvature based approach. The direct relation
between bending strain and curvature can be beneficial to overcome the numerical
errors involved in the calculation of the second derivative of the displacement mode
shapes. This issue is addressed by several studies. Lestari et al. [16] and Qiao et
al. [17] used MSE-DI algorithm in combination with strain mode shapes measured
by PVDF film sensors in order to identify damage in a composite specimen.
Adewuyi et al. [18] performed similar experiments utilizing long-gage optical fiber
Bragg grating sensors.
An important challenge regarding the practical application of the vibration based
damage identification techniques is the number and the position of the sensors.
The vibration based methods utilize standing wave patterns. Consequently, high
spatial resolution monitoring requires a relatively dense sensor network to avoid
spatial aliasing (except for the natural frequencies and modal damping based
methods). This contradicts the practical application of a vibration based technology,
since the number of sensors is often limited. The optimal sensor placement has
received considerable attention in the literature [19–21]. Savananoz [22] particularly
addresses the optimum spatial sampling interval for damage detection based on
curvatures.
There are some other ways to deal with the sensor issue. The first and most
straightforward approach is to utilize new sensing technologies. Optical fiber
distributed sensing systems as well as other distributed sensor technologies are
under development [7]. The number of successful practical applications of these
emerging technologies is, however, limited. Secondly, and the more feasible, is to
monitor only the most safety-critical locations of the structure, which is referred
to as ‘hotspot’ monitoring. These critical locations can follow from historic failure
144 Chapter 7. Discussion

Damage Index β [−]


4
5
3

2
0
0.2 0.3
0.2 1
0.1 0.1 x−coordinate [m]
y−coordinate [m] 0 0
Skin-stiffener
damage

Measurement
y z x points

Figure 7.6 Damage index β ij distributions (1D formulation in x-direction) of the MSE-DI algorithm
obtained by utilizing only the measurement points underneath the stiffeners of the skin-stiffener structure
that was analyzed in Chapter 4.

data or from the structural design process. Chapter 4 showed that the effective
application of the mode shape curvature based approach is limited to health
monitoring of the skin-stiffener connections. Moreover, defects located at the skin
are considered to be less safety-critical. Consequently, the number of measurement
points can be reduced to only the measurement points underneath the skin-stiffener
connection, as shown in Figure 7.6. This solution can provide an effective way
of monitoring the most critical areas of a skin-stiffener structure. However, this
requires good insight into the failure behavior and mechanisms of the structure
and the most safety-critical damage locations. Finally, utilizing techniques based on
propagating waves can also allow for a reduced number of sensors to cover large
areas. This is also where the vibro-acoustic modulation based method, studied in
Chapter 6, can potentially be advantageous. Although the current work is focused
on the feasibility and the understanding of the approach in the case of the steady
state behavior, it opens the ability to use pulsed carrier excitation signals traveling
between an actuator and sensor, as was shown by Kazakov et al. [23]. In that case,
the approach essentially corresponds to the more conventional acousto-ultrasonics
techniques (pitch-catch or pulse-echo) [8, 24, 25] supplemented with an additional
lower-frequency pump source in order to produce the modulations.
References 145

7.2.2 Operational and environmental effects

A general problem concerning structural vibration based health monitoring consists


of the sensitivity of dynamic parameters to changes in the operational (ambient
loading) and environmental (temperature or humidity) conditions [26]. Damage
induced changes in the dynamic properties can be masked or magnified by the
effects that are introduced due to variations in, for example, the temperature.
Methods should therefore have the ability to separate the dynamic effects caused by
the operational and environmental conditions from the changes caused by damage.
A wide variety of methods, comprising model based methods and statistical
techniques, is presented in the literature to compensate for these variations
[27–29]. Constant environmental conditions were assumed during the experimental
investigation presented. The effect of change has not been investigated. However,
the numerical models presented by Loendersloot et al. [30, 31] can be extended to
account for the effects of changing environmental conditions.

References

[1] J. Baaran. Visual inspection of composite structures. Technical report, Institute of


Composite Structures and Adaptive Systems, DLR Braunschweig, 2009.
[2] D. Balageas, C.-P. Fritzen, and A. Güemes. Structural health monitoring. ISTE Ltd, 2006.
[3] W.J.B. Grouve, L.L. Warnet, A. de Boer, R. Akkerman, and J. Vlekken. Delamination
detection with fibre Bragg gratings based on dynamic behaviour. Composites Science
and Technology, 68(12):2418–2424, 2008.
[4] J. Frieden, J. Cugnoni, J. Botsis, T. Gmür, and D. Ćorić. High-speed internal strain
measurements in composite structures under dynamic load using embedded FBG
sensors. Composite Structures, 92(8):1905–1912, 2010.
[5] M.D. Todd, G.A. Johnson, and S.T. Vohra. Deployment of a fiber Bragg grating-based
measurement system in a structural health monitoring application. Smart Materials and
Structures, 10:534–539, 2001.
[6] N. Takeda, Y. Okabe, J. Kuwahara, S. Kojima, and T. Ogisu. Development of smart
composite structures with small-diameter fiber Bragg grating sensors for damage
detection: quantitative evaluation of delamination length in CFRP laminates using
Lamb wave sensing. Composites Science and Technology, 65(15-16):2575–2587, 2005.
[7] A. Guemes, A. Fernández-López, and B. Soller. Optical fiber distributed sensing –
Physical principles and applications. Structural Health Monitoring, 9(3):233–245, 2010.
[8] A. Raghavan and C.E.S. Cesnik. Review of guided-wave structural health monitoring.
The Shock and Vibration Digest, 39(2):91–114, 2007.
[9] V. Giurgiutiu, A. Zagrai, and J. Jing Bao. Piezoelectric wafer embedded active sensors
for aging aircraft structural health monitoring. Structural Health Monitoring, 1(1):41–61,
2002.
[10] V. Giurgiutiu. Structural health monitoring: with piezoelectric wafer active sensors. Elsevier,
2008.
146 Chapter 7. Discussion

[11] J. Sirohi and I. Chopra. Fundamental understanding of piezoelectric strain sensors.


Journal of Intelligent Material Systems and Structures, 11(4):246–257, 2000.
[12] W. Lestari and P. Qiao. Damage detection of fiber-reinforced polymer honeycomb
sandwich beams. Composite Structures, 67(3):365–373, 2005.
[13] V. Giurgiutiu and A. Zagrai. Damage detection in thin plates and aerospace structures
with the electro-mechanical impedance method. Structural Health Monitoring,
4(2):99–118, 2005.
[14] T. Kundu, S. Das, and K.V. Jata. Detection of the point of impact on a stiffened plate by
the acoustic emission technique. Smart Materials and Structures, 18(3):035006, 2009.
[15] H.A. Sodano, D.J. Inman, and G. Park. A review of power harvesting from vibration
using piezoelectric materials. The Shock and Vibration Digest, 36(3):197–205, 2004.
[16] W. Lestari, P. Qiao, and S. Hanagud. Curvature mode shape-based damage assessment
of carbon/epoxy composite beams. Journal of Intelligent Material Systems and Structures,
18(3):189–208, 2007.
[17] P. Qiao, W. Lestari, M.G. Shah, and J. Wang. Dynamics-based damage detection of
composite laminated beams using contact and noncontact measurement systems.
Journal of Composite Materials, 41(10):1217–1252, 2007.
[18] A.P. Adewuyi, Z. Wu, and N.H.M. Kammrujaman Serker. Assessment of
vibration-based damage identification methods using displacement and distributed
strain measurements. Structural Health Monitoring, 8(6):443–461, 2009.
[19] E.B. Flynn and M.D. Todd. Optimal placement of piezoelectric actuators and sensors
for detecting damage in plate structures. Journal of Intelligent Material Systems and
Structures, 21(3):265–274, 2009.
[20] K. Worden and A.P. Burrows. Optimal sensor placement for fault detection.
Engineering Structures, 23(8):885–901, 2001.
[21] H.Y. Guo, L. Zhang, L.L. Zhang, and J.X. Zhou. Optimal placement of sensors for
structural health monitoring using improved genetic algorithms. Smart Materials and
Structures, 13(3):528–534, 2004.
[22] E.S. Sazonov and P. Klinkhachorn. Optimal spatial sampling interval for damage
detection by curvature or strain energy mode shapes. Journal of Sound and Vibration,
285(4-5):783–801, 2005.
[23] V.V. Kazakov, A. Sutin, and P.A. Johnson. Sensitive imaging of an elastic nonlinear
wave-scattering source in a solid. Applied Physics Letters, 81(4):646, 2002.
[24] Z. Su, L. Ye, and Y. Lu. Guided lamb waves for identification of damage in composite
structures: a review. Journal of Sound and Vibration, 295(3-5):753–780, 2006.
[25] J.-B. Ihn and F.-K. Chang. Pitch-catch active sensing methods in structural health
monitoring for aircraft structures. Structural Health Monitoring, 7(1):5–19, 2008.
[26] H. Sohn. Effects of environmental and operational variability on structural health
monitoring. Philosophical Transactions of the Royal Society A: Mathematical, Physical and
Engineering Sciences, 365(1851):539–60, 2007.
[27] A. Deraemaeker, E. Reynders, G. Deroeck, and J. Kullaa. Vibration-based structural
health monitoring using output-only measurements under changing environment.
Mechanical Systems and Signal Processing, 22(1):34–56, 2008.
[28] A. Yan, G. Kerschen, P. Deboe, and J. Golinval. Structural damage diagnosis under
varying environmental conditions – Part I: a linear analysis. Mechanical Systems and
Signal Processing, 19(4):847–864, 2005.
[29] E. Figueiredo, G. Park, C.R. Farrar, K. Worden, and J. Figueiras. Machine learning
References 147

algorithms for damage detection under operational and environmental variability.


Structural Health Monitoring, 10(6):559–572, 2010.
[30] R. Loendersloot, T.H. Ooijevaar, L.L. Warnet, A. de Boer, and R. Akkerman. Vibration
based structural health monitoring and the modal strain energy damage index
algorithm applied to a composite T-beam. In C.M.A. Vasques and J.D. Rodrigues,
editors, Vibration and Structural Acoustics Analysis: Current Research and Related
Technologies, Chapter 6, pages 121–150. Springer, 2011.
[31] R. Loendersloot, T.H. Ooijevaar, A. de Boer, and R. Akkerman. Development of a
damage quantification model for composite skin-stiffener structures. In C. Boller and
H. Janocha, editors, New Trends in Smart Technologies, pages 99–108. Fraunhofer Verlag,
2013.
Chapter 8

Conclusions and recommendations

This thesis focuses on the development of design recommendations and guidelines


for the detection, localization and characterization of damage in advanced
composite skin-stiffener structures, employing changes in the dynamic behavior.
These guidelines contribute to the development of a design tool for research
engineers, to assist the implementation of structural health monitoring technology
in safety-critical composite structures.

8.1 Conclusions

The major conclusions of this work with respect to the objective of this thesis are
presented below.

General
• Vibration based structural health monitoring is a design related problem.
The development of a damage identification strategy is made-to-measure
work and requires a thorough physical understanding of the potential failure
mechanisms, the critical damage locations and their effect on the dynamic
behavior.
• A combination of several damage identification techniques, damage features
and classifiers is required to identify the wide variety of damage scenarios
that can occur in a composite skin-stiffener structure.

Linear dynamic behavior


• Mode shape curvatures combined with the modal strain energy damage index
(MSE-DI) algorithm can be used to detect and localize barely visible impact

149
150 Chapter 8. Conclusions and recommendations

damage (BVID) at the skin-stiffener connection of advanced composite skin-


stiffener structures. The method can also roughly estimate the geometry of
the damaged area.
• Damage at the skin-stiffener connection can be detected and localized with
more confidence by employing the MSE-DI algorithm than by considering
individual mode shapes or mode shape curvatures, since damage related
information of separate mode shape curvatures is combined into a single
indicator.
• The effective application of the MSE-DI method is highly dependent on the
damage location with respect to the structural design. The most effective
results were obtained by considering the 1D MSE-DI formulation underneath
and in the direction of the stiffeners for damage at the stiffener mid-section
and perpendicular to the stiffeners for damage at the stiffener run-out. Impact
damage at the skin in between the stiffeners could not be identified. This
limits the effective application of the method to health monitoring of skin-
stiffener connections.
• Damage at the skin-stiffener interface mainly affects the bending modes. The
modes with torsional deformations around the stiffener are hardly affected
since the damage is located close to the local rotational axis.
• The number of measurement points and their location with respect to the
damage affect the sensitivity to identify damage at the skin-stiffener interface.
A reduced number of measurement points or a larger distance between the
measurement points and the damage (perpendicular to stiffener) lowers the
ratio between the damage index values of the damaged area and the pristine
area.

Nonlinear dynamic behavior


• Damage at the skin-stiffener interface can introduce clear nonlinear effects in
the dynamic behavior of the structure. These nonlinear effects are attributed
to the interaction between the skin and stiffener that occurs during opening
and closing motion of the damage.
• The opening and closing motion of the damage at the skin-stiffener connection
is mainly driven by the motion of the skin at both sides of the stiffener.
• A linear system description obtained by modal analysis based damage
identification methods is feasible for low excitation amplitudes, but suffers
under its linear system assumption for the higher amplitudes. Nonlinear
methods are, in that case, required to obtain a more adequate description
of the nonlinear dynamic behavior introduced by the skin-stiffener damage.
• Vibro-acoustic modulation based damage identification allows for the detection
8.2. Recommendations 151

and localization of impact induced damage at the skin-stiffener connection,


based on locally increased amplitude modulation effects. The damage
hardly modulated the frequency of the carrier response. Analysis of the
characteristics of the nonlinear modulations opens the ability to characterize
the nonlinear dynamic behavior introduced by the damage at the skin-stiffener
interface.
• A time domain analysis of the vibro-acoustic modulation phenomena at
multiple spatial locations revealed that the nonlinear modulation phenomena
in the dynamic response is highly dependent on the underlying dynamic
behavior of the structure around the carrier excitation frequency.

8.2 Recommendations
With respect to the work presented in this thesis further research is recommended
on the following aspects:

• More research is required on the influence of changes in the operational and


environmental conditions.

• A next step towards the practical application of the presented methodologies


for vibration based structural health monitoring is to use integrated sensor
systems. For this purpose, a numerical model can be utilized to optimally
place the sensors.

• A combination of multiple damage identification techniques, damage features


and classifiers can be beneficial to the overall damage identification performance.
The current approach showed that it was primarily successful in the
identification of damage at the skin-stiffer connection. The extension to
alternative techniques operating on a complementary sensor set can enable
the development of a complete structural health monitoring solution.

• Other types of composite structures, such as sandwich panels and torsion


boxes, should be investigated for the development of a design tool. A design
tool can assist the implementation of structural health monitoring technology
in safety-critical composite structures.

From a broader perspective, the technology gaps, as described in Section 1.3.4,


provide a good impression of what direction further research is required to take.
Appendix A

Dynamics based nondestructive


testing techniques

The present appendix provides a description of the most commonly used


nondestructive testing technologies that utilize dynamic principles to identify
damage. The structural vibrations (SV) and electro-mechanical impedance (EMI)
techniques primarily rely on standing wave patterns, while the acoustic emission (AE),
acousto-ultrasonics (AU) and ultrasonic testing (UT) approaches utilize traveling
wave characteristics.

Structural vibrations and acoustics (SV)


This group of techniques utilizes the changes in the structural dynamic behavior
(e.g. natural frequencies, damping, modes of vibration) of structures caused by
damage, as schematically illustrated in Figure A.1(a). Damage in a structure can
significantly alter the structural integrity and therefore the physical properties like
stiffness, mass and/or damping. The dynamic behavior of a structure is a function
of these physical properties and will therefore directly be affected by damage. The
low-frequency vibration based technology provides data that is relatively easy to
interpret. More complex structures can be analyzed with these methods and a
relatively large area can be explored at once. The frequency range, and hence the
resolution, is however limited [1]. As a consequence, only relatively severe damage
such as delaminations can be identified.

Electro-mechanical impedance (EMI)


The electro-mechanical impedance technique is essentially a structural vibration
technique, but is usually considered to be a separate technique. This approach uses
changes in the mechanical impedance of the structure to identify damage [2]. The
mechanical domain and electrical domain are connected by means of a piezoelectric

153
154 Appendix A. Dynamics based nondestructive testing techniques

Velocity
response v(t)
Frequency
Response
(a) Structural vibration

|Hij| [dB]
and acoustics 50
pristine
F(t) 0 damaged
Excitation
0 500 1000
force Frequency [Hz]

I(t)

Mechanical
Electrical k
(b) Electro-mechanical

PZT
V(t) m
impedance
c
Actuation Structural
& sensing impedance

’Passive’
approach Loading
F(t)
(c) Acoustic emission
Damage
or Plate
impact
event Sensor

’Active’
approach
(d) Acousto-ultrasonics
Damage Plate

Actuator Sensor

Coupling
medium Transducer

(e) Ultrasonic testing


h

Damage Structure

Figure A.1 Schematic representation of the most commonly used nondestructive testing techniques that
utilize structural dynamic principles to identify damage.
155

element that is attached to the structure (see Figure A.1(b) ). An excitation signal is
applied to the piezoelectric element. The applied voltage V (t) and the current I (t)
that flows through the element are measured simultaneously to determine the
mechanical impedance Zst . This approach is low cost and easy to apply, but is
generally lacking in the physical interpretation of the measured deviations [3].

Acoustic emission (AE)


The acoustic emission technique utilizes the transient stress waves generated by a
local source [4]. The variety of sources comprises, for example, an actively growing
defect in a structure under operational loading, a mechanical impact, but also a
fluid or gas leakage. The emitted stress waves propagate through the surface and
are recorded by a network of sensors, shown in Figure A.1(c). Once processed, the
signals allow the source to be located by using a triangulation algorithm [5]. Unlike
most other techniques, Acoustic Emission does not require an active excitation.
Other advantages include the fast inspection using a limited number of integrated
sensors as well as the ability to discern between developing and stagnant defects. A
drawback of Acoustic Emission is that a continuous operation of the system needs
to be guaranteed. One of the key difficulties is to discern the signals from other
environmental noise while the structure is in operation.

Guided wave acousto-ultrasonics (AU)


The acoustic-ultrasonic technique essentially combines the acoustic emission
technique with an active ultrasonic excitation [6]. Two configurations are possible.
In the pitch-catch configuration (see Figure A.1(d) ) a narrow bandwidth pulse
signal is sent across the specimen, while a sensor at another location receives
the signal. The propagating signal will be affected by intermediate disruptions
caused by damage. For the pulse-echo configuration, a sensor collocated with the
actuator is used to listen to the echoes of the pulse coming from damage induced
discontinuities. Despite a high sensitivity and a small number of required sensors,
the complex interpretation of the wave forms significantly impedes the application
of this technique to complex structures [6, 7].

Ultrasonic testing (UT)


Ultrasonic testing is, in many respects, similar to the acousto-ultrasonics technique.
The method also utilizes high frequency ultrasonic waves to characterize a
specimen. The main difference lies in the traveling direction of the ultrasonic
waves, which is usually normal to the surface of the specimen (see Figure A.1(e) ).
Consequently, ultrasonic testing is primarily a local scanning method and therefore
156 Appendix A. Dynamics based nondestructive testing techniques

requires access to the structure. Common formats to present the responses are
known as A-scan, B-scan and C-scan presentations [4]. Advantages of this approach
include the high sensitivity to structural discontinuities and the high accuracy
to determine the position, size and shape of subsurface defects. The required
accessibility of the surface and the time-consuming scanning process makes this
technique less suitable for health monitoring applications.

References
[1] C.-P. Fritzen and P. Kraemer. Self-diagnosis of smart structures based on dynamical
properties. Mechanical Systems and Signal Processing, 23(6):1830–1845, 2009.
[2] C. Liang, F.P. Sun, and C.A. Rogers. An impedance method for dynamic analysis of
active material systems. Journal of Intelligent Material Systems and Structures,
116(4):323–334, 1994.
[3] W. Yan and W.Q. Chen. Structural health monitoring using high-frequency
electromechanical impedance signatures. Advances in Civil Engineering, 2010:1–11, 2010.
[4] C. Hellier. Handbook of nondestructive evaluation. McGraw-Hill, 2003.
[5] T. Kundu, S. Das, and K.V. Jata. Detection of the point of impact on a stiffened plate by
the acoustic emission technique. Smart Materials and Structures, 18(3):035006, 2009.
[6] A. Raghavan and C.E.S. Cesnik. Review of guided-wave structural health monitoring.
The Shock and Vibration Digest, 39(2):91–114, 2007.
[7] R.P. Dalton, P. Cawley, and M.J.S. Lowe. The potential of guided waves for monitoring
large areas of metallic aircraft fuselage structure. Journal of Nondestructive Evaluation,
20(1), 2001.
Appendix B

Damage features and classifiers

A structural health monitoring technique typically consists of the selection of


an appropriate damage sensitive feature and the subsequent condensation and
classification of the information with the help of a damage metric. The following
table provides a general overview of the damage features and the (statistical)
classifiers that are commonly utilized for vibration based damage identification
purposes. Although the overview is far from complete, it gives an impression of the
versatility of the damage identification methods that are proposed in the literature.

Time domain
Time response / waveform
Statistical time series analysis
COR Correlation functions (non-parametric) Fassois 2007 [1]
AR Autoregressive models (parametric) Fassois 2007 [1]
Time-frequency analysis
WA Wavelet transform based methods Rucka 2006 [2]
HT/HHT Hilbert(-Huang) transform based methods Hu 2010 [3]

Frequency domain
Fourier / power spectra
Frequency response function / operational deflection shapes
FRFCH Frequency response function change Kessler 2002 [4]
FRFDI Frequency response function damage index Banerjee 2009 [5]
FRFSHP Frequency response function shape method Liu 2009 [6]
DRQ Detection and relative damage quantification Sampaio 2009 [7]
indicator
FRAC Frequency response assurance criterion Nefske 1996 [8]
FDAC Frequency domain assurance criterion Pascual 1997 [9]

157
158 Appendix B. Damage features and classifiers

RVAC Response vector assurance criterion Sampaio 2009 [7]


ODSC Operational deflection shape change Pascual 1999 [10]
GSC Global shape correlation Zang 2007 [11]
GAC Global amplitude correlation Zang 2007 [11]
Frequency response function curvature
FRFC Frequency response curvature Sampaio 1999 [12]
SFRFDI Strain frequency response damage index Maia 2003 [13]
Mechanical impedance (Z)
RMSD Root mean square deviation Park 2003 [14]
MAPD Mean absolute percentage deviation Giurgiutiu 2005 [15]
CCD Correlation coefficient deviation Giurgiutiu 2005 [15]
Transmissibility function (T)
TDI Transmissibility damage index Johnson 2002 [16]
Antiresonances
AFN Antiresonance shift Dilena 2004 [17]
Higher harmonics (nonlinear effect)
HH Higher harmonic imaging Krohn 2002 [18]
Modulations (nonlinear effect)
VAM Vibro-acoustic modulations Donskoy 1998 [19]

Modal domain
Natural frequencies
CAC Cawley-Adams criteria Cawley 1979 [20]
DLAC Damage location assurance criteria Messina 1998 [21]
MDLAC Multiple damage location assurance criteria Messina 1998 [21]
Mode shape
MS Mode shape amplitude change Ho 2003 [13]
MSS Squared mode shape slope change Ho 2003 [13]
RMS Relative mode shape change Khoo 2004 [22]
MAC Modal assurance criteria Parloo 2003 [23]
IMAC Inverse modal assurance criteria Allemang 2003 [24]
COMAC Co-ordinate modal assurance criteria Parloo 2003 [23]
FD Fractal dimension method Hadjileontiadis 2005 [25]
GFD Generalized fractal dimension Wang 2007 [26]
MSRC Mode shape rotation change Abdo 2002 [27]
Mode shape curvature
MSC Mode shape curvature method Pandey 1991 [28]
MSC Mode shape curvature squared method Maia 2003 [13]
CDF Curvature damage factor method Wahab 1999 [29]
References 159

GSM Gapped smoothing method Ratcliffe 1997 [30]


MSEDI Modal strain energy damage index method Stubbs 1995 [31]
MSECR Modal strain energy change ratio Shi 1998 [32]
Modal damping
SDC Specific damping capacity Guild 1981 [33]
Dynamic stiffness (combination of modal parameters)
ESC Effective stiffness change Khoo 2004 [22]
Dynamic flexibility (combination of modal parameters)
MFL Modal flexibility method Pandey 1994 [34]
MFLC Modal flexibility curvature method Lu 2002 [35]
ULS Uniform load surface method Zhang 1998 [36]
ULSC Uniform load surface curvature Wu 2004 [37]
MCI Modal compliance index method Choi 2005 [38]
Updating methods
Other methods
MF Modal filters Deraemaeker 2006 [39]
MD Modal peak density Chrysochoidis 2004 [40]

References

[1] S.D. Fassois and J.S. Sakellariou. Time-series methods for fault detection and
identification in vibrating structures. Philosophical Transactions of the Royal Society A:
Mathematical, Physical and Engineering Sciences, 365(1851):411–48, 2007.
[2] M. Rucka and K. Wilde. Application of continuous wavelet transform in vibration
based damage detection method for beams and plates. Journal of Sound and Vibration,
297(3-5):536–550, 2006.
[3] H.F. Hu, W.J. Staszewski, N.Q. Hu, R.B. Jenal, and G.J. Qin. Crack detection using
nonlinear acoustics and piezoceramic transducers – Instantaneous amplitude and
frequency analysis. Smart Materials and Structures, 19(6):065017, 2010.
[4] S.S. Kessler, S.M. Spearing, M.J. Atalla, C.E.S. Cesnik, and C. Soutis. Damage detection
in composite materials using frequency response methods. Composites Part B:
Engineering, 33(1):87–95, 2002.
[5] S. Banerjee, F. Ricci, E. Monaco, and A. Mal. A wave propagation and vibration-based
approach for damage identification in structural components. Journal of Sound and
Vibration, 322(1-2):167–183, 2009.
[6] X. Liu, N.A.J. Lieven, and P.J. Escamilla-Ambrosio. Frequency response function
shape-based methods for structural damage localisation. Mechanical Systems and Signal
Processing, 23(4):1243–1259, 2009.
[7] R. Sampaio and N. Maia. A Simple Correlation Factor as an Effective Tool for
Detecting Damage. In Modeling, Simulation and Control of Nonlinear Engineering
Dynamical Systems, pages 233–242. Springer, 2009.
160 Appendix B. Damage features and classifiers

[8] DJ Nefske and SH Sung. Correlation of a coarse-mesh finite element model using
structural system identification and a frequency response assurance criterion. In
Proceedings of 13th International Modal Analysis Conference, pages 597–602, 1996.
[9] R. Pascual, J.C. Golinval, and M. Razeto. A frequency domain correlation technique for
model correlation and updating. In Proceedings of the 15th International Modal Analysis
Conference (IMAC XV), pages 587–592, 1997.
[10] R. Pascual, J.C. Golinval, and M. Razeto. On-line damage assessment using operating
deflection shapes. In Proceedings of the 17th International Modal Analysis Conference
(IMAC XVII), pages 238–243, 1999.
[11] C. Zang, M.I. Friswell, and M. Imregun. Structural health monitoring and damage
assessment using frequency response correlation criteria. Engineering Mechanics,
133(September):981–993, 2007.
[12] R.P.C. Sampaio, N.M.M. Maia, and J.M.M. Silva. Damage detection using the
frequency-response-function curvature method. Journal of Sound and Vibration,
226(5):1029–1042, 1999.
[13] N. Maia, J.M.M. Silva, E.A.M. Almas, and R.P.C. Sampaio. Damage detection in
structures: from mode shape to frequency response function methods. Mechanical
Systems and Signal Processing, 17(3):489–498, 2003.
[14] G. Park, H. Sohn, C.R. Farrar, and D.J. Inman. Overview of piezoelectric
impedance-based health monitoring and path forward. The Shock and Vibration Digest,
35(6):451–463, 2003.
[15] V. Giurgiutiu and A. Zagrai. Damage detection in thin plates and aerospace structures
with the electro-mechanical impedance method. Structural Health Monitoring,
4(2):99–118, 2005.
[16] T.J. Johnson and D.E. Adams. Transmissibility as a differential indicator of structural
damage. Journal of Vibration and Acoustics, 124(4):634, 2002.
[17] M. Dilena and A. Morassi. The use of antiresonances for crack detection in beams.
Journal of Sound and Vibration, 276(1-2):195–214, 2004.
[18] N. Krohn, R. Stoessel, and G. Busse. Acoustic non-linearity for defect selective
imaging. Ultrasonics, 40(1-8):633–7, 2002.
[19] D.M. Donskoy and A.M. Sutin. Vibro-acoustic modulation nondestructive evaluation
technique. Journal of Intelligent Material Systems and Structures, 9(9):765–771, 1998.
[20] P. Cawley and R.D. Adams. The location of defects in structures from measurements of
natural frequencies. Journal of Strain Analysis, 14(2):49–57, 1979.
[21] A. Messina, E.J. Williams, and T. Contursi. Structural damage detection by a sensitivity
and statistical-based method. Journal of Sound and Vibration, 216(5):791–808, 1998.
[22] L.M. Khoo, P.R. Mantena, and P. Jadhav. Structural damage assessment using vibration
modal analysis. Structural Health Monitoring, 3(2):177–194, 2004.
[23] E. Parloo, P. Guillaume, and M. van Overmeire. Damage assessment using mode shape
sensitivities. Mechanical Systems and Signal Processing, 17(3):499–518, 2003.
[24] R.J. Allemang. The modal assurance criterion – Twenty years of use and abuse. Journal
of Sound and Vibration, 1:14–21, 2003.
[25] L.J. Hadjileontiadis, E. Douka, and A. Trochidis. Fractal dimension analysis for crack
identification in beam structures. Mechanical Systems and Signal Processing,
19(3):659–674, 2005.
[26] J. Wang and P. Qiao. Improved damage detection for beam-type structures using a
uniform load surface. Structural Health Monitoring, 6(2):99–110, 2007.
References 161

[27] M.A.-B. Abdo and M. Hori. A numerical study of structural damage detection using
changes in the rotation of mode shapes. Journal of Sound and vibration, 251(2):227–239,
2002.
[28] A.K. Pandey, M. Biswas, and M.M. Samman. Damage detection from changes in
curvature mode shapes. Journal of Sound and Vibration, 145(2):321–332, 1991.
[29] M.M. Abdel Wahab and G. De Roeck. Damage detection in bridges using modal
curvatures: application to a real damage scenario. Journal of Sound and Vibration,
226(2):217–235, 1999.
[30] C. Ratcliffe. Damage detection using a modified laplacian operator on mode shape
data. Journal of Sound and Vibration, 204(3):505–517, 1997.
[31] N. Stubbs and C.R. Farrar. Field verification of a nondestructive damage localization
and severity estimation algorithm. In Proceedings of the 13th International Modal Analysis
Conference (IMAC XIII), pages 210–218, 1995.
[32] Z. Shi and S.S. Law. Structural damage localization from modal strain energy change.
Journal of Sound and Vibration, 218(5):825–844, 1998.
[33] F.J. Guild and R.D. Adams. The detection of cracks in damaged composite materials.
Journal of Physics D: Applied Physics, 12(1):5–42, 1981.
[34] A.K. Pandey and M. Biswas. Damage detection in structures using changes in
flexibility. Journal of Sound and Vibration, 169(1):3–17, 1994.
[35] Q. Lu, G. Ren, and Y. Zhao. Multiple damage location with flexibility curvature and
relative frequency change for beam structures. Journal of Sound and Vibration,
253(5):1101–1114, 2002.
[36] Z. Zhang and A.E. Aktan. Application of modal flexibility and its derivatives in
structural identification. Research in Nondestructive Evaluation, 10(1):43–61, 1998.
[37] D. Wu and S.S. Law. Damage localization in plate structures from uniform load surface
curvature. Journal of Sound and Vibration, 276(1-2):227–244, 2004.
[38] S. Choi, S. Park, S. Yoon, and N. Stubbs. Nondestructive damage identification in plate
structures using changes in modal compliance. NDT & E International, 38(7):529–540,
2005.
[39] A. Deraemaeker and A. Preumont. Vibration based damage detection using large array
sensors and spatial filters. Mechanical Systems and Signal Processing, 20(7):1615–1630,
2006.
[40] N.A. Chrysochoidis and D.A. Saravanos. Assessing the effects of delamination on the
damped dynamic response of composite beams with piezoelectric actuators and
sensors. Smart Materials and Structures, 13:733–742, 2004.
Appendix C

Hilbert transform

The Hilbert transform H[ x (t)] of a real-valued function x (t) extending from −∞ <
t < ∞ is a real-valued function x̂ (t) defined by [1]:
Z +∞
1 x (τ )
H[ x (t)] = x̂ (t) = dτ . (C.1)
π −∞ t−τ

The associated analytical signal z(t) yields:

z(t) = x (t) + i x̂ (t) = Ainst (t)eiφ(t) , (C.2)

where Ainst (t) is the envelope signal and φ(t) the instantaneous phase signal. The
envelope Ainst (t) is extracted from the original time signal x (t) and the Hilbert
transformed signal x̂ (t) by:
q
Ainst (t) = x2 (t) + x̂2 (t) . (C.3)

The associated instantaneous phase φ(t) is described by:

x̂ (t)
φ(t) = arctan = ℑ ln z(t) , (C.4)
x (t)

and can be used to calculate the instantaneous frequency ωinst (t) [2]:

dφ(t) d 1 dz(t)
ωinst (t) = = ℑ ln z(t) = ℑ . (C.5)
dt dt z dt

References
[1] M. Feldman. Hilbert transform in vibration analysis. Mechanical Systems and Signal
Processing, 25(3):735–802, 2011.
[2] J.F. Claerbout. Earth soundings analysis: processing versus inversion. Blackwell Scientific
Publications, 2004.

163
Dankwoord

De bel heeft geklonken, de laatste ronde is ingegaan en de finishlijn komt in het


zicht. Ik kijk naar de laatste rit van de Olympische 5 km langebaanschaatsen in
Sochi, Rusland. Net zoals voor de meeste sporters die in actie komen tijdens de
Olympische spelen is dit proefschrift ook het resultaat van een 4-jarige inspanning.
In de afgelopen jaren heb ik geleerd dat er meerdere overeenkomsten zijn tussen
het nastreven van sportieve ambities en het verrichten van een promotieonderzoek.
De weg naar het beoogde resultaat vereist de nodige toewijding en wordt
nimmer beschreven door een lineair verband tussen de inspanning en de gewenste
progressie. Bovendien kan het gewenste eindresultaat niet bereikt worden zonder
de hulp van een goede begeleiding. De totstandkoming van dit proefschrift was
daarom niet mogelijk zonder de ondersteuning en het vertrouwen van veel mensen.
Allereerst wil ik graag Remko Akkerman en André de Boer bedanken. Jullie hebben
het mogelijk gemaakt dat ik na mijn afstuderen verder heb kunnen werken aan
dit onderzoek. De vrijheid die ik heb gekregen om invulling te geven aan het
onderzoek heb ik als uitermate plezierig ervaren. Jullie scherpe en pragmatische
benadering hebben mij geholpen om dit proefschrift te voltooien. Bovendien wil ik
jullie bedanken voor het ondersteunen van de leerzame onderzoekssamenwerking
met de Nondestructive Evaluation Sciences Branch van NASA Langley Research
Center. Tevens wil ik Tiedo Tinga bedanken. Jouw heldere kijk op het onderwerp
en feedback na het lezen van het proefschrift hebben mij geholpen dit boekje te
verbeteren. Veel succes met het uitbouwen van de nieuwe leerstoel gericht op
’Dynamics Based Maintenance’.
Mijn dagelijkse begeleiders Richard Loendersloot en Laurent Warnet verdienen
eigenlijk meer dan een bijzondere vermelding. Ik heb genoten om ook samen met
jullie dit onderzoek vanaf de grond af aan op te bouwen. Ik wil jullie bedanken voor
de vele waardevolle discussies over het onderzoek, maar ook over alles wat daar
buiten viel. Onze samenwerking heeft geholpen om een eigen visie te ontwikkelen
over waar het onderzoek op dit vakgebied naar toe zou moeten. Jullie enthousiasme
was een inspiratie om het onderzoek door te zetten.
My research is financially supported by the European research program Clean Sky
and was part of the Eco-Design Integrated Technology Demonstrator. Within the
framework of this project I would like to thank all industrial partners. In particular,
I appreciate the support of Nava Sela from the Israel Aerospace Industries (IAI) by
leading the work package.

165
In the third year of my PhD research, I received the opportunity to conduct research
at the Nondestructive Evaluation and Sciences Branch (NESB) of NASA Langley
Research Center in Hampton, Virginia, USA. My special thanks goes to Matthew
Rogge, who provided me this opportunity. It was an honor to collaborate with you
for a period of approximately 7 months. I have learned a lot from your views and
approaches regarding the challenges I have faced during that time. Also the support
of Cara Leckey, James Ratcliffe and Wade Jackson was really helpful to make this
experience a success.
Het onderzoek dat gepresenteerd is in dit proefschrift was onmogelijk zonder de
onvoorwaardelijke steun van Fokker Aerostructures. Alle constructies die in dit
proefschrift zijn gebruikt, zijn door jullie geproduceerd en beschikbaar gesteld. In
het bijzonder wil ik Jaap Willem van Ingen bedanken. Zelfs als ik proefstukken
wilde gebruiken voor experimenten in de Verenigde Staten kon dat op korte termijn
gerealiseerd worden. Geweldig!
Frank Grooteman van het Nationaal Lucht- en Ruimtevaartlaboratorium (NLR) wil
ik bedanken voor de samenwerking op het gebied van met name optische sensor
technologie. De uitdagingen waar we voor gestaan heb ik als uitermate leerzaam
ervaren. Helaas is maar een klein deel van het werk in dit boekje terecht gekomen.
Hierbij wil ik tevens de onlangs gestarte promovendus Jason Hwang (tevens NLR)
veel succes wensen met het vervolgonderzoek.
Ik kan dit proefschrift niet afsluiten zonder alle (oud-)collega’s binnen de
vakgroepen Toegepaste Mechanica (TM) en Productie Technologie (PT) te bedanken.
De talloze discussies hebben bijgedragen aan de realisatie van dit werk. In het
bijzonder wil ik mijn kantoorgenoten Arnoud van der Stelt, Emiel Drenth en Shaojie
Liu bedanken. Ik heb veel baat gehad bij de inhoudelijke discussies. Ook onze
gesprekken over van alles en nog wat waren uitermate plezierig. Arnoud, wanneer
gaan we een broodjeszaak openen...?
De experimenten die ik heb verricht zouden waarschijnlijk een mislukking zijn
geworden zonder de technische ondersteuning van Bert Wolbert, Bert Vos, Gert-
Jan Nevenzel en Laura Vargas Llona. Wouter Grouve wil ik graag bedanken voor
het beschikbaar stellen van zijn thesis template. Tevens bedank ik afstudeerder
Matthijs Oomen (inmiddels promovendi) voor de prettige samenwerking. Speciale
dank gaat uit naar de secretaresses Debbie Zimmerman van Woesik en Belinda
Bruinink. Jullie zijn van onschatbare waarde om het reilen en zeilen binnen de
vakgroepen op rolletjes te laten verlopen.
Gedurende de eerste jaren van mijn promotieonderzoek heb ik het geluk gehad dat
ik mijn sportieve ambities nog redelijk heb kunnen nastreven. De balans tussen het
sporten enerzijds en het promoveren anderzijds heeft me in die tijd in veel opzichten
geholpen. De diverse hoogtepunten waren niet mogelijk geweest zonder de steun
van mijn teamgenoten. Aanvullend daarop wil ik de (schaats)trainingsgroep
bedanken voor de opbeurende woorden als ik weer eens een uurtje kwam ’harken’
op mijn inline-skates of schaatsen.
Tot slot bedank ik mijn familie en vrienden. Jullie waren nooit te beroerd om mijn
’geneuzel’ aan te horen op de momenten dat het nodig was.

Ted Ooijevaar
Publications

Journal articles

1. T.H. Ooijevaar, R. Loendersloot, L.L. Warnet, A. de Boer, R. Akkerman.


Vibration based structural health monitoring of a composite T-beam. Composite
Structures, 92(9):2007-2015, 2010 (Chapter 3 of this thesis).
2. T.H. Ooijevaar, L.L. Warnet, R. Loendersloot, R. Akkerman and T. Tinga.
Impact damage identification in composite skin-stiffener structures based
on modal curvatures. In preparation for: Structural Control and Health
Monitoring, 2014 (Chapter 4 of this thesis).
3. T.H. Ooijevaar, M.D. Rogge, R. Loendersloot, L.L. Warnet, R. Akkerman,
T. Tinga. Nonlinear dynamic behavior of an impact damaged composite skin-
stiffener structure. In preparation for: Composite Science and Technology, 2014
(Chapter 5 of this thesis).
4. T.H. Ooijevaar, M.D. Rogge, R. Loendersloot, L.L. Warnet, R. Akkerman,
T. Tinga. Vibro-acoustic modulation based damage identification in a
composite skin-stiffener structure. In preparation for: Structural Health
Monitoring, 2014 (Chapter 6 of this thesis).

Book chapters

1. R. Loendersloot, T.H. Ooijevaar, A. de Boer, R. Akkerman. Development of


a damage quantification model for composite skin-stiffener structures. In:
New Trends in Smart Technologies. Fraunhofer Verlag, Stuttgart, pp. 99-108,
ISBN 978-38-396-0577-6, 2013.
2. R. Loendersloot, T.H. Ooijevaar, L.L. Warnet, A. de Boer, R. Akkerman.
Vibration based structural health monitoring and the modal strain energy
damage index algorithm applied to a composite T-beam. In: Vibration and
Structural Acoustics Analysis: Current Research and Related Technologies. Springer
Netherlands, Chapter 6, pp. 121-150, ISBN 978-94-007-1702-2, 2011.

169
Conference proceedings
1. T.H. Ooijevaar, R. Loendersloot, M.D. Rogge, R. Akkerman and T. Tinga.
Vibro-acoustic modulation based damage identification in a composite skin-
stiffener structure. Submitted to: 7th European Workshop on Structural Health
Monitoring (EWSHM 2014), Nantes, France, 2014.
2. T.H. Ooijevaar, M.D. Rogge, R. Loendersloot, L.L. Warnet, R. Akkerman
and A. de Boer. Nonlinear dynamic behavior of impact damage in a
composite skin-stiffener structure. 9th International Workshop on Structural
Health Monitoring (IWSHM 2013), Stanford, CA, USA, 2013.
3. T.H. Ooijevaar, L.L. Warnet, R. Loendersloot, R. Akkerman, A. de Boer.
Vibration based damage identification in a composite T-beam utilising low
cost integrated actuators and sensors. 6th European Workshop on Structural
Health Monitoring (EWSHM 2012), Dresden, Germany, 2012.
4. R. Loendersloot, T.H. Ooijevaar, A. de Boer, R. Akkerman. Development of
a damage quantification model for composite skin-stiffener structures. 5th
ECCOMAS Thematic Conference on Smart Structures and Materials (SMART’11),
Saarbrucken, Germany, 2011.
5. T.H. Ooijevaar, R. Loendersloot, L.L. Warnet, R. Akkerman, A. de Boer.
Structural health monitoring of an advanced composite aircraft structure
using a modal approach. 8th International Workshop on Structural Health
Monitoring (IWSHM 2011), Stanford, CA, USA, 2011.
6. T.H. Ooijevaar, F.P. Grooteman, L.L. Warnet, R. Loendersloot, R. Akkerman,
A. de Boer. Dynamic characterisation of a damaged composite structure
with stiffeners employing fibre Bragg gratings. 5th International Conference
on Composites Testing and Model Simulation (CompTest 2011), Lausanne,
Switzerland, 2011.
7. R. Loendersloot, T.H. Ooijevaar, L.L. Warnet, A. de Boer, R. Akkerman.
Vibration based structural health monitoring of a composite plate with
stiffeners. International Conference on Noise and Vibration Engineering, ISMA2010,
Leuven, Belgium, 2010.
8. T.H. Ooijevaar, L.L. Warnet, R. Loendersloot, R. Akkerman, A. de Boer.
Vibration based structural health monitoring of a composite plate structure
with multiple stiffeners. 5th European Workshop on Structural Health Monitoring
(EWSHM 2010), Sorrento, Italy, 2010.
9. R. Loendersloot, T.H. Ooijevaar, L.L. Warnet, R. Akkerman, A. de Boer.
Vibration based structural health monitoring in fibre reinforced composites
employing the modal strain energy method. 3rd International Conference on
Integrity, Reliability and Failure (IRF2009), Porto, Portugal, 2009.
10. T.H. Ooijevaar, R. Loendersloot, L.L. Warnet, A. de Boer, R. Akkerman.
Vibration based structural health monitoring in fibre reinforced composites
employing the modal strain energy method. 15th International Conference on
Composite Structures (ICCS15), Porto, Portugal, 2009.
ISBN: 978-90-365-3624-0

You might also like