You are on page 1of 8

Interface Waves for SHM of Adhesively

Bonded Joints
H. KANNAJOSYULA, P. PUTHILIATH, C. J. LISSENDEN
and J. L. ROSE

ABSTRACT
Adhesive bonds are being increasingly used in many load bearing and safety critical
members or structures. The strength and reliability of adhesive bonds need to be
monitored for structural integrity. Ultrasonic guided waves offer potentially infinite
sets of inspection possibilities in terms of the frequency and phase velocity
combinations. In this paper, a Lamb wave ultrasonic nondestructive testing (NDT)
and structural health monitoring (SHM) technique for adhesively bonded metal
plates is considered. SHM of such structures for detection of defects in the adhesive
joint region can be more efficient if the wave structure of the Lamb wave mode is
such that the energy is concentrated in the adhesive region instead of in the
adherends. In this regard, the following method for determining such modes was
attempted. The adhesively bonded joint was modeled in two different ways. The
first model considers the bonded assembly i.e. the adherends (aluminum) and the
adhesive (epoxy) to be three layers having finite thickness in perfect contact. In the
second model the adhesive is considered to be a thin layer embedded between two
half spaces of adherend material. The analytical dispersion curves from both the
models were studied and it was observed that the regions where the dispersion
curves from the models coincide have predominant displacement in the adhesive
region. Parameters were extracted from analytical results with an aim to generate
single mode interface waves in lap joints. Finite element simulations were
performed to verify generation of single mode interface waves.

INTRODUCTION
Structural Health Monitoring (SHM) and ultrasonic Non-Destructive Evaluation

Dept. of Engineering Science and Mechanics; The Pennsylvania State University;


212 Earth-Engineering Sciences Building; University Park, PA 16802
Email: lissenden@psu.edu

247
(NDE) of adhesively bonded joints with overlaps like lap-joints and stringers have
received considerable attention in recent literature [1-17]. In these works many
approaches for ultrasonic NDE can be found. They involve use of bulk waves,
homogeneous and inhomogeneous Lamb waves, interface waves and plate waves.
Almost all the above works use transmission coefficients to characterize or test a
joint. The earliest work on the use of elastic interface waves can be found in [1].
Elastic interface waves as defined in [1] are waves which propagate in an elastic
interface layer bounded by two half spaces. These waves are different from
Stoneley waves in the sense that in the former case the physical restrictions on the
existence of these waves are much weaker. The Stoneley waves exist only when the
shear velocities of the materials forming the interface are similar. Whereas, elastic
interface waves exist even when the properties of the halfspace and the embedded
layer are distinct. In the limit when the embedded material has the same properties
as the halfspaces the interface wave degenerates to a bulk wave[1]. The dispersion
curves for propagating interface waves are bounded by the Rayleigh wave of the
configuration and the shear waves in the half spaces. In [1] these waves were used
for the monitoring the curing process of adhesive. In [2] it is demonstrated that
surface wave incident on one arm (adherend) of a lap joint results in an interface
wave of the kind described in [1].

THEORY
In this section we give a brief overview of formalism involved in modeling
interface and lamb waves. Wave propagation in a plate can be modeled analytically
using either the Helmholtz potential method or the partial wave technique [18].
Wave propagation in multilayered plates can be modeled using either the global
matrix method or the transfer matrix method [18, 19]. We use the global matrix
method along with the partial wave technique.
The governing equations for wave propagation in an isotropic elastic medium are as
follows.
Hooke's law:
τ = λ I∇ iu + µ ( ∇u + ∇uT ) (1)

where, u is the displacement vector; τ is the stress tensor; λ and µ are the Lame's
constants.
Navier equation:
∂ 2u (2)
( µ + λ ) ∇ ( ∇iu ) + µ∇ u = ρ 2
2

∂t

The solutions for the above equations for plane wave propagation in the x1
direction, can be expressed in terms of displacements as

248
x2

x1

Figure 1: A three layered model for aluminum plates bonded by an adhesive.

u1 = {  Ae ( − k α x2 )
+ Be(
k α x2 )
   }
 − β Ce( − k β x2 ) − De( k β x2 )  exp {i (ωt − kx1 )}
(3)
u2 = i { α  Ae ( − k α x2 )
− Be ( k α x2 )   ( − k β x2 ) + De( k β x2 )
 − Ce
 } exp {i (ωt − kx )}
 1

where A, B, C and D are arbitrary constants which are to be solved from the
boundary conditions and α = 1 − c 2 / cL2 , β = 1 − c 2 / cT2 , while, k , cL and cT are
the wavenumber, bulk longitudinal velocity and shear velocity, respectively.
Consider a three layered plate (Configuration 1), infinite in the x1 direction as
shown in Figure 1. Each of the layers is assigned its own coordinate system. The
continuity and traction free boundary conditions at the adhesive/aluminum interface
and the free surfaces, respectively, result in a system of equations of the form (see
[18, 19])
A (ω , k ) X = 0
where, A is a 12 x 12 matrix and X is a 12 x 1 vector. For non-trivial solution of the
above equation, the following condition must be satisfied.
det(A ) = 0
which is the dispersion relation, and its solution gives the dispersion curves.
Instead of aluminum plates of finite thickness, consider the adhesive layer
embedded in between two half spaces made of aluminum (Configuration 2). For
bounded solutions, equation 3 is rewritten with B = D = 0 , for the aluminum
halfspaces. In addition the x2 coordinate axis for each halfspace is reoriented such
that x2 = 0 at the aluminum-epoxy interface. For the adhesive region equation 3 still
holds. For this case, the size of matrix A will be 8 x 8.

RESULTS
First, we present numerical calculation and comparison of dispersion curves for
configurations 1 and 2. It will be shown later that the dispersion curves for these
two cases coincide beyond a certain frequency range. The conditions under which
the dispersion curves coincide are studied. Second, we present results from finite
element simulations that support and build on the findings of the first subsection. In
particular excitability of these modes in more complex geometry like a lapjoint are
studied.

249
Figure 2. Dispersion curves of bonded aluminum plates and adhesive layer embedded between half spaces,
superimposed. Configuration 1: adhesively bonded aluminum plate, with adhesive layer thickness identical to
configuration 2; Configuration 2: epoxy layer embedded between aluminum halfspaces.

TABLE I MATERIAL PROPERTIES


ρ (g/cc) cL (km/s) cS (km/s)
Aluminum 2.7 6.230 3.130
Epoxy 1.180 2.380 1.136

Generation of single interface wave modes is anticipated to be useful for our


ongoing work on SHM of lapjoints and stringers. It is anticipated that wavenumber
frequency matching will enable excitability of single interface modes in the joint.
For this purpose dispersion curves of single plates (Configuration 3) are also
studied in the first subsection.

Analytical

As mentioned earlier, the aim of this section is to find the region and conditions in
which the dispersion curves for Configurations 1 and 2 coincide. The idea behind
this approach is that at higher frequency (lower wavelengths) the former
configuration would approximate the latter. Therefore it is anticipated that the
dispersion curves for both the curves should coincide wherever the short
wavelength condition is satisfied.
Figure 2 compares the dispersion curves of two configurations. The material
parameters used to generate the figure are given in Table I. The thicknesses of the
aluminum plate and the adhesive layer are 2 mm and 0.2 mm, respectively.

250
(a) (b)

(c) (d)
Figure 3. Comparison of dispersion curves for adhesively bonded plate with those of other configurations.
Dimensions of bonded plate (Configuration 2) corresponding to each subfigure are: (a) plate: 1mm, adhesive 0.1
mm; (b) plate: 1 mm, adhesive 0.2 mm; (c) plate: 2 mm, adhesive: 0.1 mm (d) plate: 2mm, adhesive: 0.2mm.
Configuration 1: adhesively bonded aluminum plate, with adhesive layer thickness identical to configuration 2;
Configuration 2: epoxy layer embedded between aluminum halfspaces; Configuration 3: bare aluminum plate
with thickness identical to aluminum plate in configuration 1.

Next, excitation of single interface modes in a complex geometry is explored.


Excitation of a single interface mode is preferable because it permits a simple
waveform analysis. In order to excite a single mode, wavenumber matching
between the three configurations 1, 2, 3 were considered. It was anticipated that this
procedure might yield wavenumber frequency combinations that result in optimal
energy transfer in addition to generation of a single mode wave in the joint region.
The following can be deduced from Figure 3, which shows dispersion curves for
different layer thicknesses. For a fixed adhesive thickness (0.1 mm in Figs. 3a and
3c and 0.2 mm in Figs. 3b and 3d), the modes for Configurations 1 and 2 are more
likely to coincide as the thickness of the aluminum plates increase; i.e., in the limit
as the plate thickness becomes infinite, the dispersion curves for Configuration 1
reduce to those of Configuration 2.

Finite Element Simulation

Finite element modeling was used to examine the excitability of interface waves
in irregularly shaped geometries like lap joints. A few cases where wavenumber
matching was obtained are considered for numerical simulation (see Table II). We

251
wish to verify if indeed single modes can be excited using these selected
wavenumber-frequency combinations.
TABLE II CASES CONSIDERED FOR NUMERICAL EXPERIMENTS
Case 1 Case 2
Aluminum thickness (mm) 2.0 2.0
Epoxy thickness (mm) 0.1 0.2
Frequency (MHz) 3.5 5.25
Phase velocity (km/s) 2.910 2.921

(a)

(b)

(c)
Figure 4. (a) Schematic of the lap joint sample with the comb loading. The rectangle with broken boundary
marks the observation window. Snapshots of waveform for (b) Case 1 and (c) Case 2.

In both the numerical simulations, a guided wave mode is generated in the left
flank of the adhesive joint using a wavelength spaced comb loading corresponding
to the phase velocity and frequency combination shown in Table II. For the Case 1,
A0 mode propagating along the aluminum flank of the joint is incident on the
bondline. In Case 2, Rayleigh wave is incident on the bondline.
As seen from Figure 4b, for Case 1 an interface wave along with a Rayleigh like
wave are present in the joint region. This is because the guided wave modes for all
the three cases considered in Figure 3 do not lie exactly at one point. Still a strong
interface wave can be observed at the bondline at this comparatively lower
frequency. From Figure 4c, it can be clearly seen that the wave propagating is an
interface wave. A faster but weaker mode can also be observed.

252
CONCLUSIONS
Interface waves were modeled using two approaches and compared. The
comparison showed that, depending on the size of the structure and the wavelength
under consideration, the dispersion curves for both the cases can be identical. For a
fixed adhesive thickness, the modes in the two cases are more likely to coincide as
the thickness of the plates increase-- in limit the three layered plate reduces to an
adhesive layer embedded between half spaces. Finite element modeling was used
to examine the excitability of this kind of wave in irregularly shaped geometries
like lap joints. Wave number matching was used to achieve single mode excitation,
which is often desirable in our ongoing work towards SHM of bond joints. The
FEA results show that when the Rayleigh wave is generated on one of the arms of
the lap joint, the result is predominantly an interface wave. When the input wave is
a Lamb wave -- specifically, the fundamental antisymmetric mode -- the resulting
waveform in the joint region appeared to be an interface wave and in addition a
Rayleigh type wave, which propagated along one of the free surfaces.
Future work will involve experimentation. Planned future modeling work will
incorporate viscoelasticity and anisotropy in the adhesive and adherend regions.

ACKNOWLEDGEMENTS
We gratefully acknowledge the support of the NASA Aircraft Aging and Durability
Project under cooperative agreement number NNX07AB41A

REFERENCES
1. Rokhlin S., Hefets M., and Rosen M. 1980. “An elastic interface waves guided by a thin film
between two solids,” J. Appl. Phys, 51(7): 3579-3582.
2. Turner T. M., Claus R. O., and Ocheltree S. L. 1981. “Pulse Echo Interface Wave
Characterization of Bolted Plates,” Southeastcon '. Conference Proceedings, IEEE: 484-486.
3. Mal A. K., Xu P., and Bar-Cohen Y. 1990. “Leaky Lamb Waves for the Ultrasonic
Nondestructive Evaluation of Adhesive Bonds,” J. Eng. Mat and Tech., 122(3): 255-260.
4. Pilarski A., Rose J. L. and Balasubramaniam K. 1990. “The angular and frequency
characteristics of reflectivity from a solid layer embedded between two solids with imperfect
boundary conditions,” J. Acoust. Soc. Am., 87(2): 532-542.
5. Datta S. K., and Xu P. C. 1990. “Guided Waves in a Bonded Plate: A Parametric Study,” J.
Appl. Phys., 67(11): 6779-6786.
6. Rokhlin S. I. 1991. “Lamb wave interaction with lap-shear adhesive joints: Theory and
experiment,” J. Acoust. Soc. Am., 89: 2758-2765.
7. Parikh O. K., and Achenbach J. D. 1992. “Analysis of Nonlinearly Viscoelastic Behavior of
Adhesive Bonds,” J. of Nondest. Eval., 11(3/4): 221-226.
8. Nagy P B, and Adler L. 1989. “Nondestructive evaluation of Adhesive Joints by Guided
Waves,” J. Appl. Phys, 66(10): 4658-4663.
9. Rokhlin S. I. Larantyev A. I., and Li B. 1993., “Ultrasonic Evaluation of Environmental
Durability of Adhesive Joints,” Res Nondestr Eval, 5: 95-105.
10. Lowe M. J. S., and Cawley P. 1994. “The Applicability of Plate Wave Techniques for the
Inspection of Adhesive and Diffusion Bonded Joints,” J Nondest. Eval., 13(4): 185-200.
11. Rokhlin S. I., and Larantyev A. I. 1994. “Models for ultrasonic characterization of interfaces in
adhesive joints of environmental degradation, J. Appl Phys., 76(8): 4643-4650.
12. Lowe M. J. S., Challis R. E., and Chan C. W. 2000. “The transmission of Lamb waves across
adhesively bonded lap joints,” J. Acoust. Soc. Am., 107: 1333-1345.

253
13. Heller K., Jacobs L. J., and Qu J. 2000. “Characterization of adhesive bond properties using
Lamb waves,” NDT&E International, 33: 555-563.
14. Rose J. L. 2000. “Guided Wave Nuances for Ultrasonic Non-Destructive Evaluation,” IEEE
Trans. Ultrasonics, Ferroelectrics, and Frequency Control, 47(3): 575-583.
15. di Scalea F. L., and Piervincenzo R. 2004. “Propagation of ultrasonic guided waves in lap-shear
adhesive joints: Case of incident a0 Lamb wave,” J. Acoust. Soc. Am., 115(1): 146-156.
16. Matt H., Bartoli I., and di Scalea F. L. 2005. “Ultrasonic guided wave monitoring of composite
wing skin-to-spar bonded joints in aerospace structures,” J. Acoust. Soc. Am. 118: 2240-2252.
17. di Scalea F. L., Matt H., Bartolli I., Coccia S., Park G., Farrar C. 2007. “Health Monitoring of
UAV Wing Skin-to-spar Joints using Guided Waves and Macro Fiber Composite Transducers,”
J. Intell. Matl. Struct., 18: 373-388.
18. Rose J. L. 1999. Ultrasonic Waves in Solid Media, Cambridge University Press, Cambridge,
UK.
19. Lowe M. J. S. 1995. “Matrix Techniques for Modeling Ultrasonic Waves in Multilayered
Media,” IEEE Trans. on Ultrasonics, Ferroelectrics, and Frequency Control, 42(4): 525-542.

254

You might also like