Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Composite Structures 295 (2022) 115786

Contents lists available at ScienceDirect

Composite Structures
journal homepage: www.elsevier.com/locate/compstruct

Machine learning based thermal imaging damage detection in glass-epoxy


composite materials
Ali Sarhadi a , Rodrigo Q. Albuquerque b,c ,∗, Martin Demleitner b , Holger Ruckdäschel b,c , Martin
A. Eder a ,∗∗
a
Technical University of Denmark, Frederiksborgvej 399, Roskilde 4000, Denmark
b University of Bayreuth, Universitätsstrasse 30, Bayreuth 95447, Germany
c Neue Materialien Bayreuth GmbH, Gottlieb-Keim-Straße 60, Bayreuth 95448, Germany

ARTICLE INFO ABSTRACT

Keywords: Machine learning (ML) based fatigue damage detection from thermal imaging in glass-epoxy composites is
Machine learning an important component of remote structural health monitoring used for safety assessment and optimization
Thermal imaging of composite structures and components. However, accurate characterization of fatigue damage hotspots in
Damage detection
terms of size, shape, location, hysteretic heat, and local temperature deep inside the material using surface
3D thermal analysis
thermal images remains a challenge to date. This work aims at evaluating the theoretical accuracy level of
Composite material
hotspot characterization by training a ML model with artificially generated thermal images from 3D finite
element models with increasing complexity. Modelling the fatigue damage as an intrinsic heat source allowed
to significantly reduce the influence of thermal image noise and other uncertainties related to heat transfer. It
is shown that ML can indeed accurately recover the heat influx, depth, and geometry of the heat source from
the original thermal images of the composite materials with prediction accuracies in the range 85%–99%. The
effect of training set size and image resolution on the prediction error is also presented. The findings reported
in this work contribute to the advancement of accurate and efficient remote fatigue damage detection methods
for fibre composite materials.

1. Introduction detecting fatigue damage, where a distinct temperature signature, com-


monly referred to as a ‘‘hotspot’’, is known to form due to frictional
Fibre-reinforced composites are increasingly used for load-bearing processes [11–13]. However, the current challenge in using thermal
structures in various industries, from aerospace and automotive to sensors to detect damage in composites is the ability to accurately as-
rotor blades for wind turbines or sports applications, due to their high sess the severity, size, location through thickness, and hysteretic energy
specific strength and stiffness. However, defects arising during man- dissipation (growth stage) of a damage area deep within a normally
ufacture or during service life can significantly impair their excellent complex and intricate composite structure. As a result, thermal images
properties [1,2]. The anisotropic structure leads to complex damage of the surface rarely give a direct picture of the damage zone embedded
mechanisms such as delamination, cracks or fibre breakage, which deep within the material below the surface. This means that in the case
makes the detection and evaluation of defects in terms of location of thick laminates and/or laminates with complex stacking sequences,
and size a particular challenge, especially in the case of defects that sandwich panels, bonded joints, and other complex local geometric
lie far below the surface and make visual inspection impossible [3,4].
features, the thermal signature produced by the damaged region may
Here, non-destructive testing (NDT) techniques are essential to enable
be highly distorted, rendering direct damage assessment from surface
reliable operation and extend service life. Currently, methods such as
thermal images a nontrivial task.
X-ray tomography [5], ultrasound [6] or eddy current [7] are used.
Damage assessment using thermal images requires to solve an in-
Infrared thermography is particularly well suited because it offers
verse thermal conduction problem (IHCP), which is formulated by a
coupling agent-free measurement and fast scan rates that allow scan-
higher order partial differential equation (PDE) [14–16]. Here, the
ning of large components, which is not possible with the previously
mentioned methods [8–10]. In addition, thermal imaging is ideal for intensity and shape of the defect are related to the parameters in

∗ Corresponding author at: University of Bayreuth, Universitätsstrasse 30, Bayreuth 95447, Germany.
∗∗ Corresponding author at: Technical University of Denmark, Frederiksborgvej 399, Roskilde 4000, Denmark.
E-mail addresses: rodrigo.q.albuquerque@uni-bayreuth.de (R.Q. Albuquerque), maed@dtu.dk (M.A. Eder).

https://doi.org/10.1016/j.compstruct.2022.115786
Received 22 February 2022; Received in revised form 28 April 2022; Accepted 13 May 2022
Available online 21 May 2022
0263-8223/© 2022 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
A. Sarhadi et al. Composite Structures 295 (2022) 115786

the depth of defects in CFRP in real time, achieving an accuracy of


Nomenclature 88%. The present work shows that it is possible to use a much simpler
𝜌 Material density (kg∕m3 ) regression model trained on artificially generated IR images based
𝑇 Temperature (K) on finite elements to predict not only the depth of defects of epoxy
thermoset systems, but also other three geometric parameters of the
𝑇0 Ambient reference temperature (K)
defects to achieve a prediction accuracy sometimes higher than 99%
𝑄̇ Total heat source influx (W)
(vide infra).
𝑞𝑉̇ Heat source influx per unit area (W∕m2 )
The novelty in this work is the training of the ML model with arti-
𝑞conv Convective heat surface flux per unit area ficially generated thermal images obtained from 3D numerical thermal
(W∕m2 ) analysis to perform the simultaneous prediction of four geometrical
𝑎∕𝑏 Semi-major/semi-minor axis of the elliptic parameters of defects in epoxy thermoset glass fiber reinforced (GFRP)
disc source (m) systems. This approach offers the advantage of well-controlled hotspot
𝑅 Disc perimeter radius (m) conditions, where each thermal image is precisely associated with a
ℎ𝑡𝑜𝑝 Film coefficient top surface (W/(m2 K)) priori defined heat transfer conditions, size, shape and position of the
ℎ𝑏𝑜𝑡 Film coefficient bottom surface (W/(m2 K)) heat source. As a result, thermal image noise and other uncertainties
𝑡 Time or thickness (s or m) are significantly reduced, allowing a largely unbiased assessment of the
𝑑 Distance of heat source from bottom sur- level of accuracy that can theoretically be achieved to determine key
face (m) hotspot parameters via ML. This work contributes to the advancement
𝑥, 𝑦, 𝑧 Cartesian coordinates (m) of thermal imaging-based structural health monitoring techniques that
can accurately and efficiently obtain information about the state of
𝑘𝑡𝑡 Coefficient of thermal conductivity in 𝑡-
fatigue damage in thick fibre/epoxy composites. Moreover, this work
direction (𝑡 = 𝑥, 𝑦 or 𝑧) (W/(m K))
opens up new opportunities for accelerated fatigue testing methods on
𝑐𝑝 Specific heat capacity (J/(kg K))
coupon and sub-component scale by ML aided adaptive load frequency
𝑦 True target property (m or W) control.
𝑦̂ Predicted target property (m or W)
𝑦̄ Average target property (m or W) 2. Methodology
𝑛 Number of samples (unitless)
2.1. Thermal analysis

Thermal analysis is conducted in order to phenomenologically


the source term of the heat conduction PDE (c.f. Eq. (1)). A popular model the heat transfer involved in fatigue damage accumulation
and widely used approach to solving IHCPs is by utilizing the finite processes with the purpose to create artificial thermal surface images
element method (FEM). The latter is used to compute the sensitivities under controlled ideal conditions. The heat transfer models presented
of the parameter space which can then be used to iterate towards in this work are described by the well-known general 3D heat equation
an inverse solution through an optimization procedure via e.g. the for anisotropic conductivity and internal heat source in its strong form
gradient descent method. Kien and Zhuang [17] proposed a deep La- as follows
grange method to solve structural optimization problems. The method ( )
𝜕𝑇 1 𝜕2 𝑇 𝜕2 𝑇 𝜕2 𝑇 𝑄̇
is capable of finding the minimum of the objective function which − 𝑘𝑥𝑥 + 𝑘𝑦𝑦 + 𝑘𝑧𝑧 = (1)
𝜕𝑡 𝑐𝑝 𝜌 𝜕𝑥2 𝜕𝑦2 𝜕𝑧2 𝑐𝑝 𝜌
could be applied to inverse problems. However, such approaches can
lead to a bottleneck because IHCPs are usually ill-posed (the solution The right hand side source term in Eq. (1) models the intrinsic heat
space is non-convex, there are many local minima) and computationally generated by a local defect e.g. a delamination or a crack located inside
demanding, and therefore time-consuming to solve even with high a material due to predominantly frictional effects under cyclic fatigue
computational power. Nanthakumar et al. [18] suggested to utilize loading conditions. Such a concentrated heat source causing a distinct
several experimental setups and to use improved regularization to temperature gradient that typically flares up in thermal images is com-
overcome these challenges. In that work, an algorithm was developed monly referred to as hotspot. The enthalpy increase of such a hotspot
to detect inclusions in a piezoelectric material by solving an inverse is closely associated with hysteretic energy dissipation which acts like
problem through the use of the extended finite element method. Other a heat pump. It was shown elsewhere [11] that typically the hysteretic
approaches rely on semi-analytical temperature field fitting optimiza- heat input in composite materials under high cycle fatigue conditions
tion such as the Trefftz method [19]. This method however, is available is low to moderate and constant, which eventually leads to thermal
only for a few specific cases which narrows the versatility. equilibrium. Therefore, with good approximation the problem can be
Recent advances in data science and machine learning (ML) have modelled in the steady state approximation in which case the transient
opened a new way to efficiently and accurately handle such inverse part represented by the left hand side, temporal partial term in Eq. (1)
problems and reduce computation time through more economical ap- can be neglected, i.e., 𝜕𝑇
𝜕𝑡
→ 0. The size of the hotspot source term was
proximations [20]. ML has already been used for photoacoustic to- chosen small compared to the geometrical dimensions of the model.
mography [21], geophysics [22], and heat conduction problems [23]. The hotspot temperatures in glass/epoxy composite materials typically
Here, elements of physics-based modelling complement ML models, range between 20 °C to 70 °C. In-plane hot spots parallel to the surface
improving understanding of the underlying physical processes and in- were assumed since defects in composites usually occur between plies
creasing prediction accuracy. Anitescu et al. [24] proposed an artificial in the form of debonding, delamination and microcracking.
neural network method capable of solving inverse boundary value Since radiative heat transfer is proportional to the fourth power of
problems. They adopted an adaptive collocation method that reduced the temperature, the effect of radiation can be considered insignificant
the computational effort. at these low temperatures. Instead, convection is the governing heat
Fang et al. [25] have reported the use of a relatively complex deep transfer mechanism in the present problem which prevails on both
learning model to detect the depth of artificially created defects in external surfaces that are in contact with the ambient air. The convec-
carbon fibre reinforced polymers (CFRP) with a prediction accuracy tive heat transfer boundary conditions assigned to both surfaces can be
( )
of about 95%. Saeed et al. [26] used a combination of convolutional written in the form 𝑞conv = ℎ 𝑇 − 𝑇0 . The in-plane dimensions of the
networks and deep feed-forward neural networks to detect and predict thin walled problem are chosen ten times the thickness dimension of

2
A. Sarhadi et al. Composite Structures 295 (2022) 115786

Fig. 1. (a) Circular disc domain with 𝑅 ≫ 𝑡 with global coordinate system and location of point source on the 𝑧-axis and (b) axisymmetric representation of hatched rectangle
with thickness 𝑡, radius 𝑅 and heat transfer boundary conditions (Note: figure not to scale).

Table 1 2.1.2. Thermal model with elliptical heat source


Thermal modelling parameters common to all models presented in Section 2.
Fig. 2 depicts the geometry of the thermal model with the elliptical
𝑅 𝑡 𝜌 𝑐𝑝 ℎ𝑡𝑜𝑝 ℎ𝑏𝑜𝑡 𝑇0 disc heat source of zero thickness and the heat transfer boundary
m m kg∕m3 J/(kg K) W/(m2 K) W/(m2 K) K conditions.
0.5 0.05 1992.5 940.5 10.0 5.0 293.15 The numerical analysis procedure and discretization was the same
as previously described in Section 2.1.1. Orthotropic coefficients of
Table 2
thermal conductivity were adopted as provided in Table 2.
Coefficients of thermal conductivity with values resembling a generic glass/epoxy A total heat generation magnitude was randomly chosen from the
composite material in the global coordinate directions; in the orthotropic case the fibre range 2 < 𝑄̇ < 4 W and was uniformly distributed over the area of
direction is assumed to be aligned with the global x-direction whereas the thermal the elliptic disc centred on the 𝑧-axis at a randomly chosen distance
conductivity in the transverse and through-thickness direction was set equal.
within the range 0.005 < 𝑑 < 0.045 m (see red hatched ellipse and
Case 𝑘𝑥𝑥 𝑘𝑦𝑦 𝑘𝑧𝑧
red line in Fig. 2). Both, the semi-major and semi-minor parameters
– W/(m K) W/(m K) W/(m K) were randomly chosen from a range 3.0 × 10−2 < 𝑎 < 9.0 × 10−2 m
Isotropic 0.695 0.695 0.695 and 1.5 × 10−2 < 𝑎 < 4.5 × 10−2 m. The Comsol solver was run through
Orthotropic 0.805 0.374 0.374 a Matlab code which was used to change the parameters 𝑄, ̇ 𝑑, 𝑎 and
𝑏. The four parameters were randomly chosen between their limits in
such a way that unique parameter quartets were created in each run.
the model. Therefore, it can be assumed with reasonable approximation The generation of the artificial greyscale thermal images as well as the
that conductive heat transfer across the model boundary remote from temperature scale bar are analogous to Section 2.1.1.
the hotspot is small and has a negligible influence on the hotspot tem-
perature profile. Therefore, zero flux heat transfer boundary conditions 2.2. Machine learning modelling
were assigned to the through thickness perimeter surface.
Table 1 lists the geometrical properties of the circular disc mod- All models were created using Python libraries (e.g. Numpy,
els as well as the thermal properties and the ambient temperature Sklearn) and Jupyter notebook. All codes were written in Python 3.
consistently adopted throughout all cases.
2.2.1. Data treatment
2.1.1. Thermal model with point source The greyscale thermal images were imported, cropped, resized and
Fig. 1 depicts the geometry of the thermal model with the point heat converted into 2D arrays (116 × 115 pixels) with pixel intensity as
source and the heat transfer boundary conditions. elements. Light-saturated images with pixel intensities equal to 255
This geometry was discretized (in its weak form of Eq. (1)) within were removed from the dataset. The 2D arrays were flattened to
the commercially available finite element software package Comsol [27] become 1D arrays with dimension 13 340. Only every fifth element of
with around 1.9 × 105 ten-noded quadratic tetrahedral elements. Zero each 1D array was used to build the final data set, that contained 2668
flux conditions were assigned to the nodes pertaining to the through- features. Each feature was scaled to have zero mean and unit variance,
thickness perimeter of the circumferential boundary. Convective heat where scaling was first applied to the training set (80% of the data)
flux boundary conditions were assigned to either surface with different and then used to re-scale the features of the test set.
film coefficients (see Table 1). Two different cases were analysed
namely an isotropic case and an orthotropic case whose coefficients of 2.2.2. Model
thermal conductivity are provided in Table 2. The Kernel Ridge Regression (KRR [29]) model was trained to
The concentrated heat generation source magnitude was randomly predict the target properties 𝑄, ̇ 𝑑 and, when applicable, 𝑎 and 𝑏,
chosen from the range 2 < 𝑄̇ < 4 W and was assigned to a single node using the pixel intensities of thermal images as features. Although
located on the z-axis at a randomly chosen distance within the range image processing to predict target properties can be usually done using
0.005 < 𝑑 < 0.045 m (see red dot in Fig. 1). A generalized minimum convolutional neural networks, we have chosen KRR because it is
residual method (GMRES) was used to solve the linear problem. The much simpler and faster to implement and to have its hyperparameters
Comsol solver was run through a Matlab [28] code which was used to optimized, while being considerably precise, as will be shown in the
change the parameters 𝑄̇ and 𝑑. The two parameters were randomly results section. KRR combines the kernel trick [30] with a 𝐿2 regu-
chosen between their limits in such a way that unique parameter pairs larization term, which helps avoiding overfitting. This method has a
were created in each run. An artificial greyscale thermal image of the closed form that allows for efficient training and prediction for middle-
top-surface was created with a view vector direction along the negative sized datasets. In particular, the use of the radial basis function kernel
z-axis (see camera symbol in Fig. 1(a)). A fixed temperature contour means that the original features are expanded into an infinite number
range of 20 ≤ 𝑇𝑡𝑜𝑝 ≤ 70 °C was used for the thermal contour plots in all of terms, which are not possible anymore to be individually evaluated,
the runs. but whose related inner product needed for performing predictions can

3
A. Sarhadi et al. Composite Structures 295 (2022) 115786

Fig. 2. (a) Circular disc domain with 𝑅 ≫ 𝑡 with global coordinate system and location of zero thickness elliptical disc source centred on the 𝑧-axis with semi-major axis 𝑎 and
semi-minor axis 𝑏 and (b) axisymmetric representation of hatched rectangle with thickness 𝑡, radius 𝑅 and heat transfer boundary conditions (Note: figure not to scale).

2.2.3. Point source case


The thermal images generated using point defects were
pre-processed as already described. The hyperparameters 𝑘𝑒𝑟𝑛𝑒𝑙, 𝑎𝑙𝑝ℎ𝑎
and 𝑔𝑎𝑚𝑚𝑎 of the KRR method were optimized using a random grid
search method and an initial batch of 350 images corresponding to the
training set. During the hyperparameter optimization, the performance
of the trained model was evaluated using the leave-one-out cross-
validation (CV) technique with all images of the training set, as well
as the coefficient of determination (𝑅2 ) and the mean absolute error
(MAE), which are given by

1 ∑|
𝑛
MAE = 𝑦 − 𝑦̂𝑖 || (2)
𝑛 𝑖=1 | 𝑖
∑𝑛
(𝑦𝑖 − 𝑦̂𝑖 )2
̂ = 1 − ∑𝑖=1
𝑅2 (𝑦, 𝑦) 𝑛 (3)
̄2
𝑖=1 (𝑦𝑖 − 𝑦)
The mean absolute percentual error (MAPE) was also used to quan-
tify the accuracy of the models and is given by

100% ∑ || 𝑦𝑖 − 𝑦̂𝑖 ||
𝑛
MAPE = (4)
𝑛 𝑖=1 || 𝑦𝑖 ||

The leave-one-out CV technique starts with training the model with


all samples except only one and performing one single prediction with
that one-sample test set initially left out. This operation is performed
𝑛 times using every time a different sample of the dataset as the one-
sample test set. After this CV procedure, 𝑛 different models will have
been trained and evaluated using the same performance indicator (here,
MAE and 𝑅2 ). In order to evaluate the model with samples that were
never seen during the hyperparameter optimization, the trained model
was used to predict the target properties of the test set (88 samples).
The model performance was evaluated using the same CV technique
and the MAE and 𝑅2 parameters. The model was also evaluated by the
MAPE error.
Fig. 3. Steady state top surface temperature distribution for (a) point source with
isotropic thermal conductivity with 𝑄̇ = 3 W and 𝑑 = 35 mm, (b) point source with
orthotropic thermal conductivity with 𝑄̇ = 3 W and 𝑑 = 35 mm, (c) elliptical source
2.2.4. Elliptic disc source case
with orthotropic thermal conductivity with 𝑄̇ = 3.5 W, 𝑑 = 45 mm, 𝑎 = 60 mm, 𝑏 = 30 mm For the elliptic source case, 925 thermal images were used in the
and the corresponding radial temperature profiles of both surfaces for (d) point source dataset, where 740 images were randomly selected as the training set
— isotropic, (e) point source — orthotropic and (f) elliptical source — orthotropic. The and used to train the ML model already with optimized parameters to
defects are highlighted in red. ̇ 𝑑, 𝑎, and 𝑏 of the test set. The
predict the geometrical parameters 𝑄,
model was evaluated using MAE, 𝑅2 and MAPE.

be precisely calculated via the kernel trick. Other linear and non-linear 2.2.5. Influence of the training set and image resolution
ML models (Gradient Boosting Regression, Support Vector Machine and The influence of the size of the training set, as well as of the
Lasso) were also tested but their performances were slightly worse than image resolution on the model’s accuracy were also investigated to
KRR (see the Supporting Information, SI). provide more insight on the ML model. For the influence of the size

4
A. Sarhadi et al. Composite Structures 295 (2022) 115786

Fig. 4. Comparison between the predicted and true parameters for the test set using the trained ML model and point source case. The accuracies were 99 and 96% for 𝑄̇ and 𝑑,
respectively.

of the training set, ML models were built using a variable number 3.3. Predictions for the elliptical disc source case
of training samples (𝑁) from the training set while predictions were
always performed on the same test set. The MAE and 𝑅2 parameters The inclusion of two additional parameters as target properties,
were monitored for different 𝑁 values, both for the point source and namely 𝑎 and 𝑏, slightly decreased the predictive accuracy of the ML
elliptic source cases. The influence of the image resolution on the model for 𝑄̇ and 𝑑 (see the SI). The initially chosen ranges used in
parameters MAE and 𝑅2 was also investigated for both source cases the Comsol/FEM simulations for 𝑎 and 𝑏 were relatively small (𝑎, 𝑏:
by monitoring these parameters for different image resolutions, which 5 × 10−3 m to 1.5 × 10−2 m), which resulted in a very noisy prediction
were in the range 20 × 20–165 × 165 pixels using the full training set of 𝑎 and 𝑏, as well as relatively small accuracies: 75 and 82% for 𝑎
to train the model and the test set to perform predictions. and 𝑏, respectively. It seems that the larger value of 𝑘𝑥𝑥 compared to
𝑘𝑦𝑦 worsens the prediction accuracy of 𝑎 compared to 𝑏. The larger 𝑘𝑥𝑥
leads to a smaller variation between neighbouring pixels, decreasing
3. Results and discussion the sensitivity of the ML model towards the prediction of 𝑎. Moreover,
smaller ranges of 𝑎, 𝑏 worsened MAPE and 𝑅2 as compared with larger
3.1. Artificial thermal images for point and elliptical disc source ranges.
In order to investigate the improvement of the predictions, a new
Fig. 3 shows representative examples of the three different cases set of wider and individually different ranges was then chosen (𝑎:
investigated. Fig. 3(a) shows the axisymmetric temperature distribution 3 × 10−2 m to 9.0 × 10−2 m and 𝑏: 1.5 × 10−2 m to 4.5 × 10−2 m) to generate
due to a point source in the isotropic case. Conversely, Fig. 3(b) shows the thermal images. The increase of the size of the ellipse improved the
accuracy of the ML predictions considerably in the case of 𝑎 and 𝑏, as
that the surface temperature due to a point source in an orthotropic
can be seen in Fig. 5.
material appears elongated due to the higher thermal conductivity in
The prediction accuracy was in the range 85%–98%, and the lowest
𝑥-direction. The elongated temperature distribution shown in Fig. 3(c)
accuracy found, which corresponded to the depth parameter 𝑑 (accu-
is a combination of both, the elliptic shape of the source and the
racy = 85%) was comparable to the best accuracy reported in another
anisotropic heat conductivity. The pronounced temperature peak seen
related prediction reported in the literature [26]. The use of wider
in Fig. 3(d,e) is the result of a singular heat generation intensity
ranges for 𝑎 and 𝑏 to generate the thermal images seems to be the best
at the point source. On the other hand, the elliptical heat source is
compromise for achieving good accuracy for the prediction of all four
not singular which results in a lower heat generation intensity and parameters. However, this approach penalized the excellent accuracy
a more bell-shaped temperature profile. Comparison of Fig. 3(b) and found for 𝑄̇ (99%) and mainly for 𝑑 (94%) that were obtained using
(c) demonstrates the difficulty to predict the defect geometry since narrower ranges. For applications where the depth and the temperature
both the anisotropic heat conductivity and the heat source shape itself of the defect are the most important properties to be predicted, it is
contribute to the elongated surface temperature distribution observed recommended to train a ML model using thermal images generated
in both images. adopting narrower ranges for 𝑎 and 𝑏.

3.2. Predictions for the point source case 3.4. Influence of the training set and image resolution

The influence of the number of training samples on the accuracy of


The trained kernel ridge model had the hyperparameters optimized
the predictions is shown in Fig. 6. Note that the predicted parameters
using the training set, evaluated via the leave-one-out cross validation
on the y-axes of Fig. 6 are normalized for comparison reasons (their
(see SI for more details), from where the final hyperparameters were
absolute values are provided in the SI).
obtained: 𝑘𝑒𝑟𝑛𝑒𝑙 = ‘‘rbf’’, 𝑔𝑎𝑚𝑚𝑎 = 0.0041, 𝑎𝑙𝑝ℎ𝑎 = 0.00009. The model For the point source case (Fig. 6(a,b), solid lines) the model im-
trained with the whole training set was used to predict 𝑄̇ and 𝑑 for proves considerably up to about 100 training samples and basically
the thermal images of the test set. For this simple point source case, stabilizes the performance from about 250 training samples. Adding
the agreement between the predicted and true target properties was more samples to the training set from this point improves the model
excellent (see Fig. 4). more slowly. Both the MAE and 𝑅2 trends could be nicely reproduced
The MAPE error of both predictions were (in %): 0.65 (𝑄) ̇ and by a fitting of the type 𝑦 = 𝑐1 𝑥𝑐2 +𝑐3 (dashed lines in Fig. 6(a,b)), where
3.59 (𝑑). The corresponding model’s accuracy is larger than 99% for 𝑐1 , 𝑐2 and 𝑐3 are adjustable parameters. For the elliptic source case
the prediction of 𝑄̇ and still very high (96.4%) for the prediction of (Fig. 6(c,d)), MAE and 𝑅2 vary very little when the size of the training
𝑑. In addition, the predicted and true target properties were highly set is close to 740 training samples, and the largest improvement occurs
correlated, as concluded from 𝑅2 values very close to 1. up to 100 training samples. Fig. 6 reveals that the chosen sizes of

5
A. Sarhadi et al. Composite Structures 295 (2022) 115786

Fig. 5. Comparison between the predicted and true parameters for the test set using the trained ML model and elliptic source case. The accuracy values (in %) for these predictions
̇ 𝑑, 𝑎, and 𝑏, respectively.
were 98, 85, 94, and 90 for 𝑄,

Fig. 6. Influence of the number of training samples (𝑁) on the models’ performance for the point source (subplots a and b) and elliptic disc source (subplots c and d) cases in
an orthotropic material. MAE and 𝑅2 were normalized for comparison reasons and curve fittings were added for the point source case (dashed lines).

the training sets used in the predictions shown in Figs. 4 and 5 were related to the width and height of the image, which were the same.
appropriate. The trends are much noisier than those found in Fig. 6, especially for
The influence of the image resolution on the model’s performance the elliptic source case. MAE and 𝑅2 already stabilize from resolutions
is shown in Fig. 7 for the point source (subplots a and b) and elliptic of about 50 × 50 pixels in the point source case (Fig. 7(a,b)), but the
source (subplots c and d) cases. The resolution shown in the x-axes is same parameters stabilize with considerably higher resolutions of about

6
A. Sarhadi et al. Composite Structures 295 (2022) 115786

Fig. 7. Influence of image resolution (in pixels) for the point source (subplots a and b) and elliptic disc source (subplots c and d) cases in an orthotropic material. MAE and 𝑅2
were normalized for comparison reasons.

130 × 130 pixels in the elliptic source case (Fig. 7(c,d)), although from The proposed ML method is able to overcome the inherent com-
about 50 × 50 pixels a very large improvement of the model’s accuracy plexity of IHCPs resulting from the anisotropic heat transfer behaviour,
is also observed for the latter case. asymmetric thermal boundary conditions, and non-uniformly shaped
The trends shown in Fig. 7 were noisier at the low-resolution hotspots located deep inside a glass/epoxy composite structure. Pre-
regions. This is due to random mismatches between the pixels extracted diction is possible with temperature information from only one side
to train the ML model (i.e., every fifth pixel, see Section 2.2.1) and of the damaged plate. It should be noted, however, that the method
pixels carrying some information other than the zero intensity corre- presented was applied specifically to high cycle fatigue conditions
sponding to the black background. One fifth of all extracted pixels where quasi-stationary conditions prevail. Such heat transfer conditions
of images exhibiting close but different resolutions (e.g. 30 × 30 are less complex than in thermographic damage detection, depend-
ing significantly on time evolution. While the transient response in
pixels versus 32 × 32 pixels) might contain a very different number
thermographic damage detection requires higher order neural network
of non-black pixels that indeed contribute to the improvement of the
models, the steady-state response of a fatigue damage hotspot allows
corresponding ML models.
the use of a simpler ML model, which provides higher efficiency in such
applications.
4. General remarks In all cases studied, the predictive performance of the ML algorithm
was found to follow a power function of the size of the training set. The
The ML model proved itself quite robust, where as little as 100 accuracy improved significantly up to about 100 training images, after
thermal images suffice for training to get a good performance. This which the improvement increased more slowly and asymptotically. This
suggests that building a ML model based on experimentally measured comparatively small training set proves the efficiency of the method.
IR images of defects in epoxy composites is feasible, especially if one Doubling the number of independent parameters from two to four
uses data augmentation also including artificially generated images. significantly expands the solution space, which affects the prediction
However, this ML model also has some limitations. For instance, accuracy of the elliptical disc source model compared to the point
no defect detection can be done to identify its exact location in the source. The prediction accuracy improved as the size of the elliptical
thermal image, and only one defect is expected to be present in the heat source dimensions increased. For a small mean relative heat source
same image. The geometry of real defects of composite materials can size of 𝑎𝑚𝑒𝑎𝑛 ∕𝑅 = 2%, the prediction was poor, while a larger relative
be very different from a simple ellipse. Therefore, the current ML model mean heat source size of 𝑎𝑚𝑒𝑎𝑛 ∕𝑅 = 9% gave a much more accurate
prediction. This means that a smaller size of the heat source results in
could be further extended to address these problems, which is the
lower variation of the surface pixel temperature values in the images
subject of a future investigation.
and accordingly makes it more challenging for the ML algorithm to
predict the parameter. By the same argument, the lower prediction
5. Conclusion accuracy of parameter 𝑏 compared to parameter 𝑎 for the larger set
of ranges is due to the smaller range of the former, even though 𝑘𝑥𝑥 is
In this work, it was shown that ML can be used to predict four key more than twice as large as 𝑘𝑦𝑦 .
physical and geometric parameters of a fatigue damage hotspot from Higher image resolutions resulted in better performance of the ML
surface thermal images, namely heat source intensity, depth, and size. model and the observed trend was more choppy in low resolution

7
A. Sarhadi et al. Composite Structures 295 (2022) 115786

regions, which is due to larger variations in the number of non-black [9] Chulkov AO, Nesteruk DA, Vavilov VP, Shagdirov B, Omar M, Siddiqui AO,
pixels effectively used to train the model. Prasad YLVD. Automated procedure for detecting and characterizing defects in
gfrp composite by using thermal nondestructive testing. Infrared Phys Technol
Future work will address arbitrarily shaped heat sources present at
2021. http://dx.doi.org/10.1016/j.infrared.2021.103675.
different depths of the composite material by utilizing convolutional [10] Marani Roberto, Palumbo Davide, Galietti Umberto, D’Orazio Tiziana. Deep
neural networks. learning for defect characterization in composite laminates inspected by
step-heating thermography. Opt Lasers Eng 2021. http://dx.doi.org/10.1016/j.
CRediT authorship contribution statement optlaseng.2021.106679.
[11] Eder Martin A, Sarhadi Ali, Chen Xiao. A novel and robust method to quantify
fatigue damage in fibre composite materials using thermal imaging analysis. Int
Ali Sarhadi: Writing – original draft, Preparation of the dataset, J Fatigue 2021;150:106326. http://dx.doi.org/10.1016/j.ijfatigue.2021.106326.
Comsol simulations. Rodrigo Q. Albuquerque: ML modelling, Pa- [12] Gornet Laurent, Wesphal Ophélie, Burtin Christian, Bailleul Jean Luc, Rozy-
per writing and correction, Discussion. Martin Demleitner: Paper cki Patrick, Stainier Laurent. Rapid determination of the high cycle fatigue
limit curve of carbon fiber epoxy matrix composite laminates by thermography
correction, Discussion. Holger Ruckdäschel: Paper correction, Discus-
methodology: Tests and finite element simulations. In: Procedia Engineering.
sion. Martin A. Eder: FEM simulations, Paper writing and correction, 2013, http://dx.doi.org/10.1016/j.proeng.2013.12.123.
Discussion. [13] Peyrac Catherine, Jollivet Thomas, Leray Nolwenn, Lefebvre Fabien, West-
phal Ophélie, Gornet Laurent. Self-heating method for fatigue limit determination
Declaration of competing interest on thermoplastic composites. In: Procedia Engineering. 2015, http://dx.doi.org/
10.1016/j.proeng.2015.12.639.
[14] Chen Aizhou, Li Yuan, Yan Bo. Computational inverse methods of heat source
The authors declare that they have no known competing finan- in fatigue damage problems. In: AIP Conference Proceedings. 2018, http://dx.
cial interests or personal relationships that could have appeared to doi.org/10.1063/1.5033668.
influence the work reported in this paper. [15] Peng Tishun, Liu Yongming. 3D crack-like damage imaging using a novel inverse
heat conduction framework. Int J Heat Mass Transfer 2016. http://dx.doi.org/
10.1016/j.ijheatmasstransfer.2016.06.018.
Acknowledgements [16] Muramatsu M, Nakasumi S, Harada Y, Suzuki T. Application of the inverse heat
conduction analysis to the evaluation of defects in carbonfiber-reinforced plastics.
DTU Wind Energy is grateful for the financial support of this publi- Mech Compos Mater 2015. http://dx.doi.org/10.1007/s11029-015-9458-y.
cation by VILLUM FONDEN with grant number 36050. Furthermore, [17] Kien DN, Zhuang X. A deep neural network-based algorithm for solving structural
optimization. J Zhejiang Univ Sci A 2021;22(8):609–20. http://dx.doi.org/10.
DTU Wind Energy acknowledges the financial support through the
1631/jzus.A2000380.
EUDP project ReliaBlade with grant number 64018-0068. [18] Nanthakumar SS, Lahmer T, Zhuang X, Zi G, Rabczuk T. Detection of
material interfaces using a regularized level set method in piezoelectric struc-
Appendix A. Supplementary data tures. Inverse Problems Sci Eng 2016;24(1):153–76. http://dx.doi.org/10.1080/
17415977.2015.1017485.
[19] Ciałkowski Michał J, Grysa Krzysztof. Trefftz method in solving the inverse
Supplementary material related to this article can be found on- problems. J. Inverse Ill-Posed Prob 2010. http://dx.doi.org/10.1515/JIIP.2010.
line at https://doi.org/10.1016/j.compstruct.2022.115786. More de- 027.
tails about model performances and general trends are available as PDF. [20] Adler Jonas, Öktem Ozan. Solving ill-posed inverse problems using iterative deep
neural networks. 2017, http://dx.doi.org/10.1088/1361-6420/aa9581,
[21] Antholzer Stephan, Haltmeier Markus, Schwab Johannes. Deep learning for
photoacoustic tomography from sparse data. 2019, http://dx.doi.org/10.1080/
References 17415977.2018.1518444,
[22] Vamaraju Janaki, Sen Mrinal K. Unsupervised physics-based neural networks
[1] Persson Erik, Eriksson Ingvar, Zackrisson Leif. Effects of hole machining defects for seismic migration. Interpretation 2019. http://dx.doi.org/10.1190/INT-2018-
on strength and fatigue life of composite laminates. Composites A 1997. http: 0230.1.
//dx.doi.org/10.1016/S1359-835X(96)00106-6. [23] Deng S, Hwang Y. Applying neural networks to the solution of forward and
[2] Hörrmann Susanne, Adumitroaie Adi, Viechtbauer Christoph, Schagerl Martin. inverse heat conduction problems. Int J Heat Mass Transfer 2006. http://dx.doi.
The effect of fiber waviness on the fatigue life of CFRP materials. Int J Fatigue org/10.1016/j.ijheatmasstransfer.2006.06.009.
2016. http://dx.doi.org/10.1016/j.ijfatigue.2016.04.029. [24] Anitescu Cosmin, Atroshchenko Elena, Alajlan Naif, Rabczuk Timon. Artificial
[3] Bard S, Demleitner M, Häublein M, Altstädt V. Fracture behaviour of prepreg neural network methods for the solution of second order boundary value prob-
laminates studied by in-situ sem mechanical tests. In: Procedia Structural lems. Comput Mater Continua 2019;59(1):345–59. http://dx.doi.org/10.32604/
Integrity, Vol. 13. 2018, http://dx.doi.org/10.1016/j.prostr.2018.12.299. cmc.2019.06641, URL http://www.techscience.com/cmc/v59n1/27936.
[4] Bakis Gökhan, Wendel Jan Felix, Zeiler Rico, Aksit Alper, Häublein Markus, [25] Fang Qiang, Maldague Xavier. A method of defect depth estimation for simulated
Demleitner Martin, Benra Jan, Forero Stefan, Schütz Walter, Altstädt Volker. infrared thermography data with deep learning. Appl Sci 2020;10(19):6819.
Mechanical properties of the carbon nanotube modified epoxy–carbon fiber http://dx.doi.org/10.3390/app10196819.
unidirectional prepreg laminates. Polymers 2021. http://dx.doi.org/10.3390/ [26] Saeed Numan, King Nelson, Said Zafar, Omar Mohammed A. Automatic defects
polym13050770. detection in CFRP thermograms, using convolutional neural networks and
[5] Jolly Mr, Prabhakar A, Sturzu B, Hollstein K, Singh R, Thomas S, Foote P, transfer learning. Infrared Phys Technol 2019;102:103048. http://dx.doi.org/
Shaw A. Review of non-destructive testing (NDT) techniques and their appli- 10.1016/j.infrared.2019.103048, URL https://www.sciencedirect.com/science/
cability to thick walled composites. In: Procedia CIRP. 2015, http://dx.doi.org/ article/pii/S1350449519303135.
10.1016/j.procir.2015.07.043. [27] COMSOL Inc. Comsol. 2020, URL http://www.comsol.com/products/
[6] Kempf M, Skrabala O, Altstädt V. Acoustic emission analysis for characterisation multiphysics/.
of damage mechanisms in fibre reinforced thermosetting polyurethane and epoxy. [28] MATLAB 2021. Documentation, version 7.10.0 (R2020a). Natick, Massachusetts:
Composites B 2014. http://dx.doi.org/10.1016/j.compositesb.2013.08.080. The MathWorks Inc.; 2021.
[7] Koyama Kiyoshi, Hoshikawa Hiroshi, Kojima Gouki. Eddy current nondestructive [29] Vovk V. Empirical Inference. Springer; 2013, p. 105–16. http://dx.doi.org/10.
testing for carbon fiber-reinforced composites. J. Pressure Vessel Technology, 1007/978-3-642-41136-6.
Transactions of the ASME 2013. http://dx.doi.org/10.1115/1.4023253. [30] Hofmann T, Schölkopf B, Smola AJ. Kernel methods in machine learning. Ann
[8] Bang Hyun Tae, Park Solmoi, Jeon Haemin. Defect identification in composite Statist 2008;36(3):1171–220. http://dx.doi.org/10.1214/009053607000000677.
materials via thermography and deep learning techniques. Compos Struct 2020.
http://dx.doi.org/10.1016/j.compstruct.2020.112405.

You might also like