Вы находитесь на странице: 1из 32

Authors Accepted Manuscript

Impact damaging of composites through online


monitoring and non-destructive evaluation with
infrared thermography

C. Meola, S. Boccardi, G.M. Carlomagno, N.D.


Boffa, F. Ricci, G. Simeoli, P. Russo
www.elsevier.com/locate/jndt

PII: S0963-8695(16)30132-3
DOI: http://dx.doi.org/10.1016/j.ndteint.2016.10.004
Reference: JNDT1800
To appear in: NDT and E International
Received date: 31 March 2016
Revised date: 12 October 2016
Accepted date: 20 October 2016
Cite this article as: C. Meola, S. Boccardi, G.M. Carlomagno, N.D. Boffa, F.
Ricci, G. Simeoli and P. Russo, Impact damaging of composites through online
monitoring and non-destructive evaluation with infrared thermography, NDT and
E International, http://dx.doi.org/10.1016/j.ndteint.2016.10.004
This is a PDF file of an unedited manuscript that has been accepted for
publication. As a service to our customers we are providing this early version of
the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting galley proof before it is published in its final citable form.
Please note that during the production process errors may be discovered which
could affect the content, and all legal disclaimers that apply to the journal pertain.
Impact damaging of composites through online monitoring and
non-destructive evaluation with infrared thermography

C. Meolaa*, S. Boccardia, G.M. Carlomagnoa, N.D. Boffaa, F. Riccia, G. Simeolib, P. Russoc


a
Department of Industrial Engineering - Aerospace Division, University of Naples Federico
II, Naples, Italy
b
CRdC Tecnologie s.c.r.l., Naples, Italy
c
Institute for Polymers, Composites and Biomaterials, National Council of Research,
Pozzuoli (Na), Italy
*
Corresponding author: Dr. Carosena Meola, Department of Industrial Engineering -
Aerospace Division, University of Naples Federico II, Via Claudio, 21, 80125 Napoli,
Italy. carmeola@unina.it

Abstract

The aim of this work is to highlight the help offered by infrared thermography in the

investigation of impact damaging of composites. In particular, infrared thermography is

herein used with a twofold function: monitoring of impact tests and non-destructive

evaluation of impacted specimens. Different types of composites are considered which

involve changing of either the matrix from a thermoset to a thermoplastic one with also

addition of a compatibilizing agent, or the reinforcement from carbon to glass. The

obtained results show a different behaviour under impact of the different materials with

fibres breakage only in thermoset matrix composites for the same impact energy. The

presence of the compatibilizing agent in the thermoplastic matrix prevents the material

from large deformation bringing it to behave more similar to a thermoset matrix based

material. Post-processing of thermal images allows evaluation of the overall impact-

affected zone.

Keywords
Infrared thermography, non-destructive evaluation, composite materials, thermoset matrix,
thermoplastic matrix, impact tests.

1. Introduction

Infrared thermography (IRT) allows to get a surface temperature map of any object, of even

complex geometry, in a remote and non-invasive way starting from the thermal energy

radiated by such an object in the infrared electromagnetic band of the used detector. The

resulting map can be exploited for different applications in different fields [1]. One main

field where IRT is gaining progressive popularity is non-destructive testing [2-4] in which

it has proven usefulness to evaluate different types of materials. Indeed, the use of infrared

thermography as non-destructive technique dates back to the beginning of the last century

[5], but at first, it appeared rather qualitative and not competitive with respect to other

methods. It was with the introduction of heat transfer basics in the interpretation of

thermographic images that infrared thermography received renewed attention leading to the

today in use thermographic techniques. Nowadays, infrared thermography is well

recognized as a technique for the inspection of composites for aerospace applications [6]

and is amongst the techniques covered by standards for qualification and certification of

personnel involved with non-destructive testing of materials (i.e., ISO 9712 [7]).

Indeed, infrared thermography is a non-destructive evaluation (NDE) technique which is

completely non-destructive, non-invasive and safe for both the inspected part and the

operator and fast at the same time. In fact, the inspection is performed in a remote manner

without any contact, which preserves the surface under inspection from any contamination

and the inspection can be performed far away from any environment dangerous for the
personnels safeguard. In addition, due to its 2D character, it allows getting in a fast way a

map of apparent temperature, temperature or phase values, depending on the thermographic

technique employed and on the desired information to get, of even large surfaces.

Till now, IRT has proved suitability to detect many types of defects in several materials. As

main requirement, NDE techniques, to be effective for composites, must be able to discover

barely visible impact damages. In fact, the vulnerability to impact damage is the main

weakness of composites especially of those based on a thermoset matrix. This because,

often, important damage may arise inside the material thickness without any perception

over the impacted side. However, any slight delamination, if remains undetected, may grow

in service leading to also catastrophic consequences. Besides, apart from discovering

delamination for discarding a part and/or for maintenance purposes, it is important to get

information on the material performance under load for design purposes. The best way to

get accomplished with the material performance is to monitor the material behaviour under

load; to this end, an infrared imaging device plays a unique role. In fact, it can be used to

visualize thermoelastic/plastic effects developing in a material during loading [8-11].

More specifically, it has been demonstrated, by the research group at the University of

Naples Federico II [8], that visualizing the thermal signatures induced by impact (i.e.

monitoring the impact event) on a glass/epoxy composite makes possible to get information

which are useful for the material characterization, specifically for identifying the origin and

propagation of impact damage. This approach was firstly applied to thermoset-matrix based

materials [8, 9, 12-14] and later extended to the thermoplastic ones [10, 14, 15]. The latter,

thanks to their higher damage tolerance and interlaminar toughness, as well as many other

advantages over the thermoset ones (recyclability after life-cycle, reprocessing, faster
production processes, chemical and environmental resistance, reduced moisture absorption

and reduced costs), are becoming ever more attractive also to the aeronautical sector. As

well known, many of these materials react to the impact with a visible surface deformation,

a concavity on the impacted side and a quasi-conical protrusion on the rear one. Then, even

a low energy impact produces a visible sign, making, at first sight, as superfluous the use of

sophisticated non-destructive testing techniques. However, the main advantage of

polypropylene (PP) based laminates lies in the possibility to modify their interface strength

by adjusting the composition of the matrix [16]. Hence, to make the most of their features,

it is very important to understand the effects following a certain dosage of the matrix

ingredients. In other words, gaining information about the behaviour under impact of a

composite made of PP reinforced with glass fibres is useful, but it is even more useful to be

able to ascertain, in a fast and simple way, whether a small percentage of compatibilizing

agent, added to the PP matrix, may modify the behaviour of the material under impact. This

may help to decide the application use of a given new material, but it can also help to tailor

the desired material characteristics. In addition, it is certainly a great advantage to get

information on the material in a fast way by simply monitoring an impact event with an

infrared imaging device. This work would show as using infrared thermography may help

understanding the behaviour under impact of different composite materials involving

different types of matrix and/or reinforcement. In this sense, this work follows ref [14] with

the difference that Ref 14 was mainly devised towards thermo-elastic effects while this

work is mainly devised towards the thermoplastic ones. Then, particular attention is herein

given to post-processing of thermal images with the intent to define the limit between

sound and damaged material, to measure the damage size and the overall extension of
delamination. The ultimate aim is to get further information for a validation of the infrared

thermography monitoring method and to make it usable at the industrial level.

2. Materials and testing procedures

2.1. Description of specimens

Four different materials are considered for the preparation of specimens:

1. Thermoset matrix (epoxy resin) reinforced with unidirectional carbon fibres

oriented at 0, 90,45, or better the stacking sequence (0/45/90/-45)s includes 12

pre-impregnated plies with a resin content of 33%. The final thickness is 2.4 mm.

2. Thermoset matrix (epoxy resin) reinforced with unidirectional E-glass fibres

oriented at 0 and 90. The stacking sequence includes 8 plies [02, 902]s with an

overall thickness of 2.9 mm.

3. Thermoplastic matrix made of polypropylene (PP grade MA712 from Unipetrol

Czech Republic with MFI = 12 g/10 min) reinforced with weave woven glass fabric.

The stacking sequence includes 20 balanced glass fabric layers symmetrically

arranged with respect to the middle plane of the laminate [(0/90)10]s , with a target

thickness of 3 mm.

4. Thermoplastic matrix made of polypropylene (PP grade MA712 from Unipetrol

Czech Republic with MFI = 12 g/10 min) modified with the addition of a

percentage (2% by weight) of a common coupling agent (polypropylene grafted

maleic anhydride, PP-g-MA) that is commercialized under the trade name of

Polybond 3200 (MFI 115 g/10 min, 1 wt% maleic anhydride, from Chemtura). The
reinforcement is made with weave woven glass fabric as above described; the

overall thickness is 3 mm again.

All specimens were fabricated by the hand-lay-up technology. Specific details of the

investigated specimens in terms of code, dimensions and composition are summarized in

the following Table 1. Infrared thermography is used as both non-destructive testing

technique (lock-in thermography, LT) and online thermal monitoring technique during

impact tests.

Table 1 Investigated specimens

Code Dimensions Composition (Matrix Reinforcement)


(mm)
CFRP 100x130x2.4 Unidirectional carbon fibres impregnated with epoxy resin
(0/45/90/-45)s
GFRP 100x130x2.9 Unidirectional glass fibres impregnated with epoxy resin
[02,902]s
PG 240x240x3.0 Neat PP - Woven glass fibres
PCG 240x240x3.0 Modified PP (2 wt% PP-g-MA) Woven glass fibres

2.2. Non-destructive testing

As shown in Fig. 1, the test setup includes the specimen, the infrared camera and a halogen

lamp (1 kW) for thermal stimulation of the specimen; in particular, one lamp is enough for

small specimens, while two are used for larger ones. The used infrared camera is the

SC6000 (Flir systems), which is equipped with a QWIP detector, working in the 8-9 m

infrared band, NEDT < 35mK, spatial resolution 640x512 pixels full frame, pixel size 25

m x 25 m and with a windowing option linked to frequency frame rate and temperature

range. The distance of both camera and lamp (lamps) to the specimen surface is varied

between 0.5 and 1 m with care put to the mutual orientations to avoid reflections.
The camera is equipped with the Lock-in module that drives the halogen lamp to generate a

sinusoidal thermal wave of selectable frequency f and the IRLock-In software (coupled

with the IR Lock-in package) for performing lock-in thermography analysis. The thermal

wave, delivered to the specimen surface, propagates inside the material and gets reflected

when it reaches zones where the heat propagation parameters change (in-homogeneities).

The reflected wave interacts with the surface wave producing an oscillating interference

pattern, which can be measured in terms of either temperature amplitude or phase angle ,

and represented as amplitude, or phase, images, respectively. The basic link of the thermal

diffusion length to the heating frequency f and to the mean material thermal diffusivity

coefficient is via the relationship:

(1)

The depth range for the amplitude image is given by , while the maximum depth p, which

can be reached for the phase image, corresponds to 1.8 . In general, it is preferable to

reduce data in terms of phase image because of its insensitivity to both non uniform heating

and local variations of emissivity coefficient, over the monitored surface. Hence, the

material thickness, which can be inspected, depends on the wave period (the longer the

period, the deeper the penetration) and on the material thermal diffusivity. According to Eq.

(1), the knowledge of the mean thermal diffusivity is fundamental to evaluate the depth at

which any detected anomaly is located, or to chose the frequency value to check the

material conditions at a given depth. To this end, the overall thermal diffusivity can be

evaluated with the lock-in technique itself [17], or with flash thermography [18].
Fig. 1 Test setup for lock-in thermography tests

2.3. Impact tests

Impact tests are carried out with a modified Charpy pendulum. In fact, Charpy impact tests

were originally developed for metals to evaluate the amount of energy to be provided to a

notched specimen until it was completely fractured during the impact. However, metals

exhibit a relatively simple tensile failure at the notch root whereas fibre reinforced

composites have a much more complex behaviour and it is hard to relate the absorbed

energy in a Charpy test to the energy absorbed by an actual component in a more realistic

structure. On the other hand the impact tests performed in the present work are mainly

intended to produce visible impact damage (VID) or barely visible impact damage (BVID)

on a plate-like structure supported by an appropriate fixture, rather than a complete fracture

of a specimen. Finally, the dimension of the impactor (12.7mm) and the windows (125mm

x 75mm) have been selected according to some considerations reported in [19] and some

constraints given by the Charpy pendulum frame. It is worth to recall the experimental set-

up used is well suited to monitor impact dynamics with the IR camera as sufficient room is

provided to the IR camera installation on the side opposite to the impacted surface as

shown in Fig. 2.
Fig. 2 Test setup for impact tests a) Charpy pendulum b) Specimen lodge and position of

the infrared camera

The impact energy E = 10 J is set by suitably adjusting the falling height of the Charpy arm.

The infrared camera is positioned at about 45 cm from the specimen surface to achieve a

spatial resolution of 4.3 pixels/mm and acquires sequences of thermal images during impact

tests at a frame rate of 96 Hz. To allow for a complete visualization of thermal effects

evolution with respect to the ambient temperature, the acquisition starts few seconds before

the impact and lasts for some time after. It is worth noting that care has been put to prevent

reflected radiation from the test-room. In particular, every bright part of the setup that could

lead to reflections on the viewed specimen was made opaque (in the infrared band of

course). In addition, a mobile screen has been used to cover the test apparatus including the

specimen lodge and the camera; this, owing to the specific test apparatus (Fig. 2b), is easy

to do.

3. Data analysis

Each sequence of thermal images acquired during impact tests is post-processed; in

particular, the first image (t = 0 s) of the sequence, i.e. the specimen surface temperature
(ambient) before impact, is subtracted to each subsequent image so as to generate a map of

temperature difference T:

(2)

i and j representing lines and columns of the surface temperature map.

3.1. On-line monitoring of impact tests

For an overview on the development of thermal phenomena under impact, some videos of

T images for each specimen, taken during impact at E = 10 J, are supplied for the online

version of the paper. Each video is identified with the name of the corresponding specimen

(i.e. CFRP, GFRP, PG, PCG). In the following, a sudden general description of each video

is given, while more details are supplied later on.

Starting from the CFRP, images initially appear with uniform colour, which corresponds to

T = 0. Then, a sudden variation occurs at the impact with appearance of some dark zones

due to thermo-elastic effect and of a white oblong structure, which accounts for temperature

rise caused by fibre breakage. More specifically, the hot stripe is composed of two tracts,

symmetrically spaced with respect to its centre which coincides with the tip of the hammer

nose. Fractions of a second later the colder zone disappears and the hot stripe warms up

consistently (T 20 K) accounting for large amount of energy dissipation. This means

that fibres break [8, 12] along their horizontal (longer side) direction while the remaining

surface is in tension under the pushing impact force. Going on to the successive images, the

negative T values tend to completely vanish whereas the hot zone tends to enlarge and the

maximum T value decreases.

Also in the GFRP video, images appear initially at uniform colour (T = 0); then, a sudden
variation occurs (at the impact) with appearance of some dark zones (thermo-elastic effect)

and of lighter (warm) lines with a central hot spot (heat dissipation). Going on to the

successive images, the negative T values tend to vanish while the hot spot becomes much

milder and a warmer area around it forms and grows up.

The two videos PG, and PCG are quite similar showing both at the impact first a central

almost circular dark region, which is replaced in time by a light warmer one. The main

differences are the lower reached temperature levels and that the circular central region is

smaller for the PCG specimen.

For a direct comparison between the different specimens, some T images taken 0.04 s

after impact at E = 10 J are reported in Fig.3. As a general comment, damage under impact

occurs along the fibres main directions. In addition, notwithstanding the same impact

energy, the different materials undergo different levels of damage. This is particularly

evident by comparing Fig. 3c and 3d which refer to two specimens perfectly equal in

thickness and type of fibres, the only difference being the presence of the compatibilizing

agent in the matrix of the PCG (Fig. 3d). In particular, a hot (white) oblong structure is

present over the CFRP specimen (Fig. 3a) accounting for fibres breakage along their

horizontal (longer side) direction. Such a structure is surrounded by a lighter area on top

and bottom and spiky ends on left and right sides; the lighter colour (i.e. slight temperature

rise) indicates the overall delamination.


T [K]

a) CFRP c) PG

b) GFRP d) PCG

Fig. 3 Comparison of T images, taken at t = 0.04 s, for E = 10 J; dimensions are in mm

The GFRP specimen shows a hot (white) spot surrounded by lighter vertical tracts

interspersed by darker zones. The hot spot indicates a local breakage, while the lighter

tracts help to identify fibres bundles; more specifically, fibres appear vertically as they are

on the external layer. It is possible to distinguish either fibres misalignment, or non-uniform

distribution of resin epoxy. In particular, it has to be observed that the hot spot engages

breakage of fibres over two tracts one of which (left side in Fig. 3b) appears interrupted

because of a zone rich of resin. It has been already demonstrated [12] that the presence of

manufacturing defects amplifies the weakness of the material to impact load.


In the other two specimens (Figs. 3c and d) the impact has caused a circular warmer zone

which is smaller for the PCG specimen (Fig. 3d) with respect to that of the PG one. This is

due to the presence of the compatibilizing agent which prevents large deformations [10].

To getting more information, the sequences of T images are post-processed in the Matlab

environment. For each specimen, first the maximum T values, Tmax, are extracted and

collected in Table 2. As shown, Tmax attains values close to, or higher than, 20 K for

CFRP and GFRP, while it does not exceed more than 2.6 K for the other two specimens.

This, considering that abrupt temperature rise is a symptom of fibres breakage and hot spots

coincide with the damage loci [8-9, 12], leads to infer that breakage occurred only in

thermoset matrix based specimens.

Table 2 Maximum T for the different specimens

Specimen Tmax (K)

CFRP 19.9

GFRP 24.7

PG 2.4

PCG 2.6

By comparing the four images in Fig. 3 one controversial aspect may arise regarding the

link between impact energy and temperature increase since the impact energy is the same

for the four specimens. One may challenge thickness differences between the specimens

(Table 1); in reality, looking carefully at Table 2, it is easy to exclude any liability in the
thickness variation. In fact, the CFRP specimen is thinner (0.5 mm less) than the GFRP

one, but it displays a lower T (4.8 K less) value. Instead, the GFRP specimen is thinner by

only 0.1 mm less than the PG one, but it displays a greater T (22.3 K more) value. In

addition, notwithstanding the same thickness, the PCG specimen displays a higher T (0.2

K more) with respect to the PG one.

The case can be simply settled by the mind to what happens during the impact. The

temperature increase is produced by the fraction of energy absorbed by the material. The

mechanism of energy absorption by the composite during impact depends on many factors

such as: impact velocity, geometrical parameters and material inherent characteristics (i.e.,

brittle, or ductile) [20-21]. For low-velocity impacts, the energy Ea absorbed by the

specimen is generally regarded as the sum of: membrane energy Em, bending energy Eb and

damage energy Ed.

(3)

The importance of each component depends on the material properties. In particular, for a

brittle material, Ed includes two terms, one accounting for fibre breakage Edb and the other

one for delamination Edd. Instead, for a ductile material, the energy is predominantly spent

in deformation and delamination perpendicular to the impactor axis.

Considering that most of the absorbed energy is dissipated as heat, it is apparent that the

sequence of thermal images, taken (at the rear side) with the infrared camera during the

impact event, can be usefully exploited for either a better understanding, or validation, of

previous hypothesis about the mechanisms of energy absorption. To this end, it is important

to take into account the material thermal properties and the involved heat transfer

mechanisms. As a general rule, the formation of micro-cracks is accompanied by small


dissipation of energy, which in turn gives rise to small temperature increase. Conversely, a

fibre breakage involves large amount of dissipation of energy with an abrupt local increase

of temperature [12].

3.1.1. Measurement of the damaged area

Often impact tests are carried out with the intent to assess the material performance for

design purposes; to this end, the resulting damage produced by an impact of given energy is

required. It is common practice to put the panel in the impact machine, perform an impact

of given energy, remove the panel, perform non-destructive tests and put it again in the

impact machine for another impact of different energy in another zone and so on. It is

generally unlikely to hit the imposed damage extension at the first impact; usually, many

attempts are necessary. Naturally, this procedure has some disadvantages: it is time

consuming and, sometimes, not very accurate. In fact, the most commonly in use NDT

techniques may fail to detect the actual delamination extent essentially because of two main

problems:

two delaminated surfaces tend to tightly adhere once the impactor moves away;

delamination propagates through tortuous pathways.

These problems may led to undetected delamination, or to underestimation of its real size

with of course negative consequences on the structure life.

Conversely, using infrared thermography, it is possible to recognize the type of damage

occurred, from the temperature rise and from the extension of the warm area, directly

during online monitoring with time saving. In the previous work [8], it has been shown that

the extension of the damaged area can be obtained by contouring the warm area. However,
this poses the problem of the minimum T to be assumed as boundary (Tb) between sound

and damaged material; some attempts have been made till now [15, 22], but this problem is

still not completely solved.

In the present work we drive attention towards the extension of the warm zone along

horizontal (x) and vertical directions (y). For different time instants, T profiles along x and

y for the CFRP specimen are reported in Fig. 4. A T image taken 0.031 s after impact is

shown in Fig. 4a to facilitate readability of T profiles; more specifically, T(x) of Figs. 4b

and 5a passes through y = 0, while T(y) of Figs. 4c and 5b passes through x = 0. From

these profiles it is possible to measure length (Fig. 4b) and width (Fig. 4c) of the warm

zone, which for convenience herein are referred to as diameters DH and DV. Basically, since

the hot zone lasts for some time (see the different curves), one could perform measurements

of DH and DV within at least 4 seconds. However, some problems may arise because of the

lateral thermal diffusion, for which, as the maximum T decreases, the warm area tends to

enlarge, especially along y (Fig. 4c). This effect may lead to an overestimation of the DV

value.

Since all curves overlap towards T = 0, the two graphs of Fig. 4 are shown again in Fig. 5

with magnification (smaller T scale) to facilitate discrimination between the different

curves. It has to be noticed that the blue curve attains the highest T value but presents also

negative values being affected by the thermo-elastic effects, still present for t = 0.031s, then

it seems more appropriate to refer to the second curve, which is no more affected by

thermo-elastic effects and not yet by the diffusive ones.


Fig. 4 T profiles for different time instants for the specimen CFRP a) T image taken

0.031 s after impact b) T along x; straight line for y = 0 in (a) c) T along y; straight line

for x = 0 in (a)
Fig. 5 T profiles for the specimen CFRP: magnification of Fig. 4 a) T along x; straight

line for y = 0 in (a) b) T along y; straight line for x = 0 in (a)

Attempting to measure DH along the x direction (Fig. 5a), it seems that, by excluding the

first blue and the last red, all curves appear well overlapped not only in the horizontal tracts

but also at the beginning of the lift at about 10 mm from the centre (x = 0). Owing to the

change of slope of the curves and assuming a Tb value close to zero, the distance of about
20 mm can be assumed as the overall DH value, or better, as the overall extension of

delamination along x. More complex appears the evaluation of DV along the y direction

because of the large data fluctuations (Fig. 5b). However, by discarding the first blue curve

the successive three ones (taken till 0.250 s) remain overlapped until the beginning of the

lift at about 7 mm from the centre (x = 0). Analogously, the distance of about 14 mm can

be assumed as the overall DV value, or better as the overall extension of delamination along

y. Of course, by varying the Tb value, it is possible to discriminate between more

important damage and slight delaminations. In particular, if Tb = 1 K, from Fig.5 it comes

up DH = 11 mm and DV = 3 mm which correspond to the dimensions of the cut (i.e. the

oblong structure in Fig. 3a).

T profiles along x and y for the GFRP specimen, at different time instants, are reported in

Fig. 6a and b, respectively. As already described, the specimen GFRP displays sudden at

the impact (Fig.3b) a hot spot of high temperature increase (T 25 K) surrounded by

intermittent distribution of warmer tracts over the entire viewed surface; this accounts for

local breakage and wide impact-affected zone. Both graphs are presented in a magnification

fashion to allow discrimination of the different curves plotted for the different time instants.

By setting Tb = 1 K, it results that DH 4 mm and DV 3 mm which correspond to the

dimensions of the hot spot (i.e. fibre breakage). Again, from Fig. 6a, by discarding the first

blue curve, which is still affected by thermo-elastic effects, and searching for any change of

curve slope (low Tb), it can be assumed DH 21.5 mm with a distance from the hot spot

centre of about 12.5 mm on the left and of about 9 mm on the right. Going to the next graph

in Fig. 6b, the different curves do not have any straight tract but depart sudden with a slope;
this because, due to the presence of fabrication defects [12], almost the entire viewed

surface shows some impact effects. For a more detailed discussion see the next section 3.2.

Fig. 6 T profiles for the specimen GFRP a) T along x b) T along y

The warm area for specimens with a thermoplastic matrix attains an almost circular shape,

so it can be assumed DH = DV and then only one diameter can be evaluated from profiles

along the single direction x. In Fig. 7, T profiles along x for PG (Fig. 7a) and PCG (Fig.

7b) specimens are, respectively, compared. For Tb = 1 K, DH is about 12 mm for the PG

specimen and about 8 mm for the PCG one; instead for the same reasoning as for the
specimen CFRP, the overall deformation extends for a diameter of about 20 mm for PCG

specimen and greater than 28 mm for the PG specimen (measurements not reliable on the

right side of Fig. 7a). The values of DH and DV of the different specimens are collected in

Table 3 for two Tb values.

Fig. 7 T profiles along x for different time instants for specimens PG and PCG a)
PG specimen b) PCG specimen
Table 3 Measurement of warm area diameters for the different specimens

Specimen DH (mm) DH (mm) DV (mm) DV (mm)


Tb 0 Tb = 1K Tb 0 Tb = 1K
CFRP 20 11 14 3
GFRP 21.5 4 3
PG >28 12 >28 12
PCG 20 8 20 8

3.2. Non-destructive evaluation

The scope of the inspection is to find the impact damage shape and size for a comparison

with the warm stain visualized during online monitoring. To this end, tests are carried out

by varying the heating frequency from 1 Hz down to 0.05 Hz, which is a reasonable range

based on experience and on an estimation of the average thermal diffusivity with the lock-

in technique itself [17]. Of course, to relate the detected damage to the depth, the exact

thermal diffusivity value (Eq. 1) must be known; this requires specific tests considering that

the damage caused by the impact affects the local thermal diffusivity value. However, this

is outside the purpose of this work, which is mostly on the extension in plane rather than in

depth of the damage.

Some phase images, taken on the specimens PG and PCG, impacted at E = 10 J, are shown

in the following figures 8 and 9, respectively. Specimens are viewed from the impacted

side. As can be seen, for the specimen PCG (Fig. 9) at f = 0.53 Hz, an almost circular dark

zone, which enlarges by decreasing the heating frequency to 0.15 Hz, is present. Such a

variation of area accounts for the damage progression through the specimen thickness. The
damaged area assumes a different shape for the specimen PG (Fig. 8) showing a long

branch along the horizontal direction and a shorter one in the vertical direction.

Two observations can be made. One is that the presence of the two branches accounts for a

larger damaged zone for the PG specimen with respect to the PCG one. This evidence, as

already observed during on line monitoring of impact tests, indicates a less extensive

deformation because of the presence of the compatibilizing agent in the matrix. The other

point regards the different length of the two branches for the PG specimen; this is mainly

due to the distribution of the clamping force during impact [23].

Fig. 8 Phase images of the specimen PG impacted at E = 10 J a) f = 0.53 Hz b) f = 0.36 Hz

c) f = 0.15 Hz

Fig. 9 Phase images of the specimen PCG impacted at E = 10 J a) f = 0.53 Hzb) f = 0.36 Hz
c) f = 0.15 Hz
At last, two phase images, taken from the rear surface (opposite to the impacted one) and

showing the largest detectable damage, of the two specimens CFRP and GFRP are reported

in Fig. 10. As can be seen, the damage develops along the fibres direction. In fact, a long

horizontal stripe with smears along the fibres at 45 is visible for the CFRP specimen;

instead, an oblong structure with its longer side along the vertical direction is visualized for

the GFRP specimen. For both specimens it is also possible to see the presence of other

signs, like fibres orientation, presence of porosity and further details.

From phase images, by knowing the spatial resolution of the used instrument (infrared

detector and lens) it is possible to measure the size of the damaged area; for present tests

the spatial resolution is in the range 3 - 3.7 pixels/mm. Quantitative measurements are again

expressed in terms of the two diameters DH and DV (along horizontal and vertical

directions) and collected in Table 4. Considering the values for the CFRP specimen and

going back to Fig. 5 and Table 3 it is possible to recognize that the damage detected with

LT corresponds to the warmest zone, which is to say, to the zone with more important

damage. It has to be noticed that slight delaminations, far from the impact point, can be

visualized only during the impact while do not affect enough the phase angle which is

visualized through NDT.


a) CFRP b) GFRP
Fig.10 Phase images showing the largest detectable damage for thermoset matrix

specimens, taken from the rear side at f = 0.05 Hz

Table 4 Measurement of damage size for the different specimens

Specimen DH (mm) DV (mm)


CFRP 14 11
GFRP 7 13
PG 30 15
PCG 10 10

Particular considerations deserve the GFRP specimen since the discussion made in the

previous section 3.1.1 was not exhaustive. In fact, a reliable estimation of the overall

delaminated area was not possible from graphs in Fig. 6b. Luckily enough, being the

material translucent, after impact a local permanent faded area is visible to the naked eye in

transparency, which corresponds to the damaged area and can be assumed as a reference.

Such visible area well matches the damaged area detected by LT (Fig.10b) and the warmer

area after 1 s. For a direct comparison, both the visible image and the T image, taken 1 s

after impact, are shown in Fig. 11.


a) T image for t = 1 s b) visible image
Fig. 11 T and visible images of the GFRP specimen
4. Conclusions

In the present work, the impact damage of different types of composites are considered

which involves changing of either the matrix from a thermoset to a thermoplastic one, with

also addition of a compatibilizing agent, or the reinforcement from carbon to glass. Every

tested specimen is subjected to an evaluation with lock-in thermography and online thermal

monitoring during impact.

The obtained results show that:

the thermal signature caused by the impact supplies information about the impact

damaging of the tested composites. In particular, the abrupt rise of temperature

bears witness for fibres breakage, while lower temperature variations indicate either

delamination and/or plastic deformation;

non-destructive evaluation with lock-in thermography is able to detect impact

damage for the different types of composites. There is a general agreement between

what detected by lock-in thermography and what visualized through online thermal

monitoring.

lock-in thermography fails to appraise the whole extension of the delaminated zone.

In fact, slight delaminations get confused with the background scene, since similar

small variations of the phase angle may be induced by local material non-

uniformities.

the temperature profiles given by direct online thermal monitoring of the impacted

zone allow to better recognize the type of damage occurred, from the temperature
rise and from the extension of the warm area.

From the practical point of view, the obtained results bear witness for the possibility to use

infrared thermography for fast assessment of the behaviour under impact of composites

directly through online monitoring of impact tests. However, the two diameters

measurement procedure herein proposed is only a fast preliminary approach. The next step

is towards the measurement of both impact damaged area and overall extension of impact-

affected area. This, of course, will be performed by considering a more vast variety of

materials specimens.

References

[1] C. Meola, G.M. Carlomagno Recent advances in the use of infrared thermography.
Meas Sci Technol, 2004; 15: R27-R58.
[2] G. Giorleo, C. Meola, Comparison between pulsed and modulated thermography
in glass-epoxy laminates, NDT&E International, 2002; 35: 287-292.
[3] G.M. Carlomagno, C. Meola, Comparison between thermographic techniques for
frescoes NDT, NDT&E International, 2002; 35: 559-565.
[4] C. Ibarra-Castanedo, J.R. Tarpani, X.P. Maldague. Nondestructive testing with
infrared thermography, European J Physics, 2013; 34: S91-S109.
[5] V.P. Vavilov. Thermal non destructive testing: short history and state-of-art Proc.
QIRT 92 (Paris) ed D. Balageas, G. Busse and G.M. Carlomagno (Paris: EETI
editions), 1992, p. 17993.
[6] Standard Practice for Infrared Flash Thermography of Composite panels and
Repair Patches Used in Aerospace Applications, ASTM E2582-07.
[7] ISO 9712:2012 Non-destructive testing (currently under revision).
[8] C. Meola, G.M. Carlomagno, Impact damage in GFRP: new insights with Infrared
Thermography, Composites Part A, 2010; 41: 1839-1847.
[9] C. Meola, G.M. Carlomagno and F. Ricci, Monitoring of impact damage in
Carbon Fibre Reinforced Polymers, QIRT 2012, Napoli, June 11-14, 2012, paper
n. 374, pp.8.
[10] C. Meola, G.M. Carlomagno, S. Boccardi, G. Simeoli, D. Acierno and P. Russo,
Infrared Thermography to Monitor Thermoplastic-matrix Composites Under Load,
Proc. 11th ECNDT, Prague, Czech Republic, 6-10 October 2014, ISBN 978-80-
214-5018-9.
[11] S. Boccardi, G.M. Carlomagno, C. Bonavolont, M. Valentino, and C. Meola,
Infrared thermography to monitor Glare under cyclic bending tests with
correction of camera noise, Proc. QIRT 2014, Bordeaux, France, 7-11 July 2014,
paper 215.
[12] C. Meola, G.M. Carlomagno, Infrared thermography to evaluate impact damage in
glass/epoxy with manufacturing defects, Int J Impact Engineer, 2014; 67: 1-11.
[13] C. Meola, S. Boccardi, G. M. Carlomagno, N.D. Boffa, E. Monaco, F. Ricci,
Nondestructive evaluation of carbon fibre reinforced composites with infrared
thermography and ultrasonics, Composite Structures, 2015; 134: 845-853.
[14] C. Meola, S. Boccardi, N.D. Boffa, F. Ricci, G. Simeoli, P. Russo, G.M.
Carlomagno, New perspectives on impact damaging of thermoset- and
thermoplastic-matrix composites from thermographic images, Composite
Structures, 2016; 152: 746-754.
[15] S. Boccardi, G.M. Carlomagno, C. Meola, P. Russo, G. Simeoli Monitoring
impact damaging of thermoplastic composites, Journal of Physics: Conference
Series 658 (2015) (XXII AIVELA Annual Meeting) 012005 doi:10.1088/1742-
6596/658/1/012005.
[16] G. Simeoli, D. Acierno, C. Meola, L. Sorrentino, S. Iannace, P. Russo, The role of
interface strength in the low velocity impact behavior of PP/glass fiber laminates,
Composites B, 2014; 62: 88-96.
[17] C. Meola, G.M. Carlomagno, A. Squillace, G. Giorleo Non-destructive control of
industrial materials by means of lock-in thermography, Meas Sci Technol, 2002;
13: 1583-1590.
[18] C. Meola, C. Toscano Flash Thermography to Evaluate Porosity in Carbon Fiber
Reinforced Polymer (CFRPs), Materials, 2014; 7: 1483-1501.
[19] Compiled by Aircraft Energy efficiency ACEE Composites Project Office.
Standard Tests for Toughened Composites, NASA Reference Publication 1092,
July 1983.
[20] S. Abrate Impact on laminated composite materials, Applied Mechanics Review,
1991; 44: 155-190.
[21] T.J. Kang, C. Kim Impact Energy Absorption Mechanism of Largely Deformable
Composites with Different Reinforcing Structures, Fibers and Polymers, 2000; 1:
45-54.
[22] S. Boccardi, G.M. Carlomagno, G. Simeoli, P. Russo, C. Meola Evaluation of
impact affected areas of glass fibres thermoplastic composites from thermographic
images, Measurement Science and Technology, 2016; 7: 075602 (12pp).
[23] S. Boccardi, G.M. Carlomagno, C. Meola, P. Russo and G. Simeoli, Infrared
thermography to evaluate impact damaging of thermoplastic composites, Proc.
QIRT 2014, Bordeaux, France, 7-11 July 2014. paper 214.

Highlights
Monitoring of impact events with infrared thermography is advantageous.
Infrared detectors sense even slighter heat dissipated through low impact energy.
Post-processing of thermographic images supply information which help understand
impact damage mechanisms.
The visualized thermal signatures allow evaluation of the overall impact-affected
area.
Lock-in thermography is able to detect the most severe impact damage.
Legends of videos
CFRP A sequence of thermal images acquired at 96 Hz during impact at E = 10 J of the
specimen CFRP.

GFRP A sequence of thermal images acquired at 96 Hz during impact at E = 10 J of the


specimen GFRP.

PG A sequence of thermal images acquired at 96 Hz during impact at E = 10 J of the


specimen PG.

PCG A sequence of thermal images acquired at 96 Hz during impact at E = 10 J of the


specimen PCG.
Graphical Abstract

Вам также может понравиться