Graphical Abstract Figure
Graphical Abstract Figure
Close modal

Abstract

For composite laminates, a rising R-curve is observed for their fracture toughness under Mode I stress, which is important for a comprehensive failure analysis of the materials. Since it is laborious to measure the R-curve due to its dependence on both the load and the crack extension, we put forward a novel compact tension specimen by modifying its geometry to eliminate the relation between fracture toughness and crack extension, so as to simplify the experimental process of the R-curve measurement by only recording the load history. Two machine-learning models were developed for the optimum sample design based on the finite element analysis of the effect of sample geometries on the R-curve. A simple neural network model was built for designing tapered specimen and a reinforcement learning model was created for further finding the best design from a broader design space. The results showed that, in contrast to the specimens with a tapered shape, which only ensure the independence between the R-curve and crack extension in the case of a small extension, the design provided by the reinforcement learning provides such independence across a wider range of crack length and an improved accuracy.

1 Introduction

Fiber-reinforced plastics (FRP) are widely used in aerospace engineering as structural materials. An important research direction in the field of the mechanics of composites is how to evaluate the damage initiation, damage evolution, and crack propagation of composite materials. A typical progressive damage feature is shown as the load is redistributed in other constituents or plies once failure occurs in a certain part of composite laminates. In this context, researchers have proposed the continuum damage model that has been proven very effective in revealing the failure mechanism in both static and dynamic loading cases [13]. While lots of the research effort has been taken on the criteria of damage initiation [46], the post-failure model has received relatively less attention. Among others, the most widely-used model is the fracture energy-based model. With a pre-assumed strain-softening law, the damage variable under certain stress conditions can be determined and stress components are then changed accordingly to describe the post-failure behavior [1,7,8]. While researchers have proposed different shapes of the strain-softening curve for considering various features of materials, such as the bi-linear model for brittle interfaces [9], the tri-linear model for fiber bridging [10], and the exponential model for nonlinear softening of materials [7], the area under these curves of damaged response is always equal to the fracture energy. It suggests that fracture toughness, both trans-laminar and inter-laminar, plays an important role in the damage modeling of composite laminates.

The fracture toughness, or the critical energy release rate (ERR) Gc can be thought of as the energy barrier of an additional infinitesimal increment of crack extension, which is an intrinsic property of material. While for perfectly brittle solids like ceramics, the value remains a constant. However, for ductile fracture, the crack grows in a way that leads to accelerated energy dissipation due to the increasing size of the plastic zone [11]. In FRP composites, the rising trend is also experimentally founded [1215]. Two mechanisms are ascribed, that are the increasing size of fracture process zone and the rougher fracture surface leading to the additional energy dissipation as the additional area of free surfaces in generated. Even though ASTM developed a standard practice (E1820-18) for determining R-curve to accommodate the widespread need for this type of data, the experiments are challenging to perform due to the requirement of exact capture of crack extension. The compliance method which indirectly predicts the crack length by measuring the continuously changing compliance of the specimen is also presented in the ASTM standard. However, multiple loading–unloading processes or multiple specimens are required for obtaining one curve. Some alternative ways have been brought out by researchers to avoid this difficulty of real-time measurement of crack extensions. The modified compliance calibration method [16] determines a compliance curve from the range of discrete crack lengths under linear-elastic FE modeling rather than loading–unloading sequences. The compliance combination method [17], in which a strain gauge is put on the top surface for independently obtain the compliance during the test, allows calculating the crack length. The double compliance method [18] utilized a two-dimensional (2D) elasticity model for a more accurate compliance-crack length relation and directly used data on load–displacement curve to calculate compliance. While these above methods are theoretically extracting the crack length from structural compliance, the J-integral-based method, on the other hand, seeks the solution for crack resistance from the full-field measurement. The contour integral can be calculated by integrating the 2D digital image correlation technique into standard fracture test [19]. Some researchers managed to find alternative test methods instead of the traditional double cantilever beam or compact tension (CT) tests. Size-effect law of the laminates, tested on cracked laminates, was used for deducing the R-curve based on a simple linear-elastic fracture mechanics (LEFM) model [20,21]. This approach has been effectively applied on both Mode I and Mode II loading cases. Similarly, machine learning-based model is proposed to indirectly predict the R-curve from the tensile strength of centre-cracked [22]. The connection between the strength value and R-curve was provided by a finite fracture mechanics model (FFM). These two methods completely eliminate the need to measure crack length, but rely on the accuracy of the size-effect law or FFM model. Meanwhile, data reduction is not straightforward compared to the standard tests.

The geometries of fracture test specimens can be an influencing factor for the fracture behavior. It has been widely studied that for CT tests, different geometric parameters leads to the different failure mechanisms in the loading process. Many researchers performed specimen design with the objective to determine the appropriate geometry to ensure crack extension for fracture toughness characterization before the specimen fails by any other damage mechanism. Typical geometries include the extended compact tension [23], the widened compact tension [24], the tapered compact tension (TCT) [24], the doubly-tapered compact tension (2TCT) [25], and the notched curved compact tension [26]. Among them, the configurations of TCT and 2TCT specimen, proposed with the aim of reducing the compressive stresses generated in the specific area of the specimen which might cause failure due to vertical compressive fiber breaking. This shows that the geometries can play an important role in regulating the stress distribution. Since the fracture toughness in LEFM is equivalent to the extent of stress concentration, it inspires us whether it is feasible to realize the independence of crack extension by modifying specimen geometries. In the field of self-healing polymer composite, a special fracture specimen with similar contour shape as that of the TCT called tapered double cantilever beam (TDCB) is widely used for evaluating the healing efficiency [27,28]. The theoretical analysis [29,30] shows that with the linear tapered shape, the compliance changes approximately linearly with the extending crack, thus resulting in a constant value of compliance rate change determined by the height of the cross section and consequently the no need of measuring crack length. However, such independence only works for a certain range of crack length, making it inappropriate for determining the whole R-curve involving a relatively wider range of crack extension. In this context, we are going to perform a throughout analysis of the geometries of CT specimens regarding to its effect on the R-curve measurement. We aim to provide an optimum design of the contour shape of CT specimens without restricting to the linear tapered shape. Since a vast design space may be explored by the indefinite shape, a machine learning algorithm combined with the finite element model (FEM), especially the reinforcement learning (RL), is employed as a search tool for finding the optimum design ensuring long-range independence of crack extension.

This paper is organized as follows. The finite element analysis was performed on both CT and TDCB specimens to identify the formulation of the problem and validate the numerical model in Sec. 2. In the following section, the effect of the geometries of specimens was analyzed in detail, with a particular emphasis on the contour shape of the tapered line and quadratic curve. In Sec. 4, a RL model was constructed for the best possible design of a CT specimen with a larger range of crack extension, after the disclosure of the geometric effect’s mechanism. Then the results and discussion about the optimized design are presented. The paper is concluded with some conclusions in Sec. 6.

2 Theoretical and Finite Element Method Analysis of TDCB Design

A common practice adopted in ASTM standard for measuring the fracture toughness and R-curve is the compact tension tests with the typical dimensions shown in Fig. 1. With a compact specimen at a force F(i), the fracture toughness can be calculated as [31]
(1)
where the dependence of the crack length is captured by the function of f(a/w). For standard CT specimen with specific size requirements for plastics are met, this function is given by ASTM E1820-18 [31]
(2)
The TDCB geometry as shown in Fig. 1 was designed to provide a constant value of the f(a/w) over a desired range of crack lengths. The energy release rate G can be presented in the form of Eq. (3) using the compliance method with dC/da denoting the derivative of the compliance with respect to crack length [16]. For isotropic material, the function of f(a/w) is related to the compliance by Eq. (4), with which the design of TDCB specimen can then be derived using the beam theory to calculate the structural compliance.
(3)
(4)
Various beam theories have been employed to approximate the height profile h. The compliance of the specimen is defined as twice the displacement of a single beam at the point of load application divided by the load. With this, the Euler Beam theory gives [18]
(5)
By considering the contributions from bending and shear deflections, the compliance of the beam determined by the simple shear-corrected beam theory (SBT) leads to a higher value of energy release rate (ERR) as Eq. (6) [28].
(6)
Here v is the Poisson’s ratio. Further studies account for the effects of beam root rotation, with a correction of the effective crack length is added which leads to [30]
(7)
It has to be noted that a constant h is assumed when using the above beam theories for deducing the derivative of compliance, while the h should be an unknown function of the distance x from the load-line, making it a bending problem of a beam with varied sections. Considering this, here we conduct a more precise analysis on the TDCB specimen. Taking the upper part of a simplified contour of TDCB as shown in Fig. 2 into consideration, beam theory is applied to two sections that connects at x=x0.
Fig. 1
The typical geometry of the standard CT specimen and the TDCB specimen
Fig. 1
The typical geometry of the standard CT specimen and the TDCB specimen
Close modal
Fig. 2
Upper part of a simplified contour of the TDCB specimen
Fig. 2
Upper part of a simplified contour of the TDCB specimen
Close modal
(i) For beam with a uniform section (0<x<x0), we have
(8)
(9)
where u1 is the vertical displacement and ϕ1 is the rotation of the Timoshenko beam.
(ii) For beam section with a linear height h(x)=h0+k(xx0) (x>x0)
(10)
(11)
The integration constants for u1 and u2 are determined by the following four equations of boundary conditions (BCs) and the conditions of continuity at x=x0:
(12)
(13)
(14)
(15)
Two corrections are added into the calculation of load-line displacement. One is the effect of tapered neutral axis. According to the schematic diagram of Fig. 2, with the incline angle of α/2 for the neutral axis in the tapered section and neglecting the shear deformation, the sectional rotation ϕ could be corrected as
(16)
Another correction is the consideration of the root rotation at x=a. It can be estimated using the elastic foundation method [30] where the un-cracked edge is assumed to be connected with a foundation whose stiffness is determined by the beam parameters as k´=2Et/h. Then, the rotation of the beam root can be determined by the bending problem with the distributed shear force of q(x)=k´u(x) as shown in Fig. 2. The solution leads to the approximate estimation of rotation at x=a by considering a uniform height ha
(17)
where
(18)
Then, the compliance can be determined as Eq. (19). Detailed descriptions of the function form can be found in the  Appendix.
(19)
To validate the above analysis and to further discover the optimum design of the geometries, the FEM was applied. The models for the standard CT and typical TDCB tests, of which the sizes were presented in Fig. 1, were established considering the symmetry. The applied force and BCs of the half models are illustrated in Fig. 3(a). A symmetric BC was applied on the un-cracked edge while free BC was set for cracked region. The virtual crack closure technique (VCCT) was utilized for calculating the ERR from the stress and displacement at the crack tip, of which the details can be found in Ref. [32]. Mesh sensitivity was studied on the standard CT model as shown in Fig. 3(b). A global mesh size of 1.0 was then employed to ensure both the accuracy and efficiency of ERR values provided by the FEM.
Fig. 3
Validation for the analysis on standard CT and typical TDCB tests: (a) finite element models, (b) data reduction formulation in ASTM E1820 and results by FEM with different mesh sizes on the CT specimen, and (c) results by different beam theories and by FEM on the TDCB specimen
Fig. 3
Validation for the analysis on standard CT and typical TDCB tests: (a) finite element models, (b) data reduction formulation in ASTM E1820 and results by FEM with different mesh sizes on the CT specimen, and (c) results by different beam theories and by FEM on the TDCB specimen
Close modal

The f(a/w) given by the FEM agrees well with that by Eq. (2) in ASTM E1820 for CT specimen, indicating the effectiveness of the FEM model for evaluating the data reduction of the Mode I fracture toughness. For the TDCB specimen, with FEM predictions as the reference, the results by different beam theories are compared in Fig. 3(c) including Euler beam theory of Eq. (5), SBT of Eq. (6), corrected effective length (CEF) of Eq. (7), and our analysis considering the varying height profile of Eq. (19). The results clearly show that while all the models provides values close to FEM when the crack is relatively small (0.2a/w0.5), only our model predicts a sudden rising trend with the increasing crack (0.5a/w0.8). This comparison suggests that the beam root rotation contributes more significantly to the load-line displacement when the crack approaches the boundary since the first term in Eq. (19) whch involves the Timoshenko beam theory decreases with longer crack. It can be found that the f(a/w) value is monotonically decreasing for Euler, SBT, and CEF analysis due to that the continuously increasing height h(x) is accounted rather than a uniform height h0. However, the way in which the height profile is considered is different from our model. In our model, h(x) is included initially in solving the bending governing equations. In Euler beam model of Eq. (5) and SBT model of Eq. (7), h is first assumed to be a constant when deriving the compliance, thus these two equations provide lower f(a/w) values as the uniformly changing height is accounted instead of the real-tapered shape. For the ASTM CT specimen, the f(a/w) value continues to increase since the K should obviously increase with growing crack. The TDCB specimen was designed to counter the increasing trend by making the specimen stiffer with a changing section. However, the analysis by Euler and SBT beam theory would overestimate the counter effect since they consider a uniformly changing section instead of the real tapered shape.

For the CEF model where beam root rotation is introduced as an extended crack length, results demonstrate that the correction is good enough for smaller crack but fails in accuracy when crack extends near the boundary. In our model, the correction of beam rotation is dependent on the un-cracked ligament (wa), which gives rise to the prediction of larger beam root rotation as the remaining ligament goes small. Even though our model shows a same trend as the FEM reference, there is still discrepancy especially for the part after the sudden increase. Such error might come from the elastic foundation model for calculating the beam root rotation where the stiffness of the foundation and the height profile were simplified to reduce the complexity. In the following Fig. 4, the contribution of the two terms in Eq. (19) is shown. The first term refers to Timoshenko beam theory, while the second term relates to the rotation of an elastic foundation. It can be found that the contribution of the term of rotation is negligible for small cracks but significant for large cracks. As a result, the simplification of not taking the varied section into account while considering the rotation of elastic foundation would only affect the part of curve after sudden increase. The simplification would lead to larger increase as the uniformly changing section causes a more significant change of the rotation. This can be validated by comparing the two curves of the FEM and Eq. (19) in Fig. 3(c). Since the main point of the paper is to investigate the shape effect of the specimen on the scenarios of large crack extensions, therefore the FEM results by VCCT technique is utilized as our references.

Fig. 4
The contribution of Timoshenko beam theory and rotation of the elastic foundation in Eq. (19)
Fig. 4
The contribution of Timoshenko beam theory and rotation of the elastic foundation in Eq. (19)
Close modal

3 Geometric Effect in Mode I R-Curve Measurement

It can be inferred from Eq. (1) that, if the specimen is properly designed to keep a constant value of f(a/w), then the measurement of fracture toughness, especially the crack resistance curve would be easy to implement without the need of capturing the crack lengths. Given that the specimen’s susceptibility to fracture will rise as the crack propagates, it is obvious that this increasing trend on the f(a/w) curve is unavoidable. However, if the specimen can be designed to be stiffener as crack extends, it would be possible to weaken the trend. To this end, in this section, we will conduct analysis on the geometric effect of specimen, or effect of the height profile, on the f(a/w) to discuss the optimum design of specimens for Mode I R-curve measurement.

3.1 Tapered Design.

Effect of the tapered shape of TDCB specimen on f(a/w) was investigated here. On the basis of the geometry of standard CT specimen where h0=30mm and w=30mm, several different tapered shapes were represented by the varied height of the right-hand end l. The results are shown in Fig. 5(a) for a total of five different l/h0 from 0 to 2. Compared to the standard CT specimen, where the ratio l/h0=1, such tapered design by only varying l at the right-hand end did not make much difference except for the shape of l/h0=0. When the l/h0<1, the f(a/w) value is higher than the standard CT specimen since the decreasing height leads to a larger value of compliance. But for l/h0>1, the f(a/w) is nearly the same. Since the main focus of the design is to ensure a constant f(a/w), the ratio of maximum and minimum values of f(a/w) with a/w in the range of [0.3,0.6] is extracted to evaluate the performance. The closer the ratio to 1, the better the design. As illustrated in the bar plot of Fig. 5(a), all the max/min ratios are over 2 for l/h0=1, 1.5, and 2 and up to 2.9 for the case of l/h0=0, indicating the tapered design can hardly meet the need of constant f(a/w) value when h0=30mm.

Fig. 5
Effect of the tapered shape of TDCB specimen on f(a/w). Results for the shapes of (a) h0=30 with varying l to h0 ratios and (b) h0=15 with varying l to h0 ratios.
Fig. 5
Effect of the tapered shape of TDCB specimen on f(a/w). Results for the shapes of (a) h0=30 with varying l to h0 ratios and (b) h0=15 with varying l to h0 ratios.
Close modal

Then, we changed the h0 to 15 mm as shown in Fig. 5(b), which is similar to the recommended TDCB shape. The same trend can be found that the function values become higher as the l/h0 increases. While for both cases the height at the right-hand l make slight difference with l/h0 greater than 1.0, a substantial influence is shown with the change of the h0. Compared to the shape of h0=30mm, the max/min ratio of f function values is significantly reduced to 1.5. We further compared the two cases of h0=15mm and h0=30mm with a same l=30. The two curves agrees well with a/w>0.6, but shows a clear discrepancy when a/w<0.6 as the initial f values are higher for h0=15mm, thus resulting in a smaller max/min ratio.

Therefore, two hints can be concluded for making the geometry design:

  1. The h0 should be small to ensure a higher initial value of f(a/w).

  2. A positive slope of the tapered shape is needed for the postpone the increasing trend of the f(a/w) thus making it remains small values at the desired range of crack extension.

The above analysis on the tapered shape design suggests that the combination of h0 and l should be delicately designed to ensure the lowest value of max/min ratio. Therefore, we built a simple neural network model to assist in finding the best tapered shape for TDCB specimen. The model framework is presented in Fig. 6(a). A neural network model was trained to predict the max/min ratio from two inputs, which were the two positions of point A and B for representing the tapered shape. A Monte-Carlo method was utilized to randomly generate the training dataset with yA uniformly chosen from [10,30] and yB uniformly chosen from [10,90], both were integers. The yA and yB were then imported into FEM script to automatically calculate the corresponding max/min ratios for 0.3<a/w<0.6, which serves as outputs in the dataset. A total of 160 training data was generated which was 10% of the total possible 1600 combinations of yA and yB in the given range. The neural network consists of one input layer, one output layer, and three hidden layers. The number of neurons in each layer is 4, 6, and 4 respectively. A activation function of relu is used. After using the MinMax scaler for data pre-processing, a weighted mean square error (W-MSE) is defined as loss function to increase the accuracy when predicting a relatively low value of the max/min ratio, which is the key metrics for our optimization problem [33]. Here, yi is the predicted value and yi^ is the target value.
(20)
The training performance of this neural network is shown in Fig. 6(b). The loss values of the weighted MSE in training and test dataset with increasing epochs suggest that the training process was good as the loss instantly dropped and converged. The training process was ended at 1000 epochs to prevent the over-fitting. R-square score for training and test regressions were 0.978 and 0.982 respectively, which shows good accuracy of the ML model’s predictions on the Max/Min ratio. With this well-trained model, it is thus easy to explore the whole design space of yA and yB using brutal-force search to find the optimum tapered shape. The results were shown in Fig. 6(c). The best shape given by the ML model is yA=10 and yB=21 with the Max/Min ratio is 1.11. Since the training dataset only covers 10% of the design space, we then extracted the best shape from the training data, which is yA=11 and yB=31 with the Max/Min ratio is 1.17. This demonstrates that the model is able to learn from the limited data and help find the optimized shape in the unseen data. It is noted that the shape provided by ML model is better than TDCB shape whose Max/Min ratio is 1.13. The model was then utilized to generate the contour plot of the Max/Min ratio over the whole design space. It can be found from the contour plot that this ratio is mainly yA-dominant as the relatively small values of the ratio locates at the region where yA is small. Meanwhile, the contour line suggests that there is a threshold value of yB for each yA. With the increasing yB, the ratio value gradually decreased, and then converged to a certain value after yB exceeded this threshold. These two conclusions are consistent with the two hints obtained before from direct FEM analysis.
Fig. 6
An neural network model for exploring the design space of tapered shape: (a) model framework, (b) training and regression performance, and (c) the optimized results of the tapered design provided by the model using brutal-force search
Fig. 6
An neural network model for exploring the design space of tapered shape: (a) model framework, (b) training and regression performance, and (c) the optimized results of the tapered design provided by the model using brutal-force search
Close modal

3.2 Quadratic Curve.

Despite that the tapered shape is able to provide a good design for the crack size of 0.3<a/w<0.5 compared to the standard CT specimen, the f function will still rise significantly for a/w>0.5. Inspired by the two hints concluded by the simple tapered design, we further explored the shape of quadratic curve for considering the more possible shapes with small h0 and increasing height profile.

The shapes with quadratic curves and the corresponding results are presented in Fig. 7. A total of seven curved shapes were considered and determined by the yC of the controlling point. The value of yC for the tapered shape is 15.5. Then, two convex curves and five concave curves were included here. Interestingly, the convex and concave shapes showed distinct feature on the f(a/w) curves. At the beginning of the crack extension where 0.3<a/w<0.5, the convex shapes (yC>15) close to the tapered shape were clearly better as the increasing height prevents the ERR from rising. A slight increasing trend can be found for tapered shape as shown in Fig. 5(c), which was alleviated by the convex shapes. However, when the crack continues to propagate where a/w>0.5, the concave shape is preferred as smaller variance was found. It can be inferred from the theoretical analysis of TDCB that the majority of the compliance is contributed by the root rotation. Therefore, the smaller variance on the f(a/w) might be due to the suppression of the root rotation by the concave–curved shape at the ends. The f(a/w) values for the concave shape of yC=11 keeps stable for cracks at a/w=0.65 while significant increase has been found early at a/w=0.55 for convex shapes.

Fig. 7
The schematics of the CT specimens with a curved shape controlling by yC and the corresponding results of f(a/w)
Fig. 7
The schematics of the CT specimens with a curved shape controlling by yC and the corresponding results of f(a/w)
Close modal

Compared to the Max/Min ratio of 1.12 for the tapered shape, the minimum value of this ratio is 1.09 for yC=16 for the region of 0.3<a/w<0.6, suggesting that this quadratic curved design can achieve better performance than simple tapered shape with the specific range of crack lengths.

3.3 Combination of the Convex and Concave Curve Designs.

Since the convex and concave quadratic curved shapes are preferred for the small and large crack extensions, respectively, in this section, the combined shape design of the convex and concave curve was investigated in order to take both of the two advantages.

Six different shapes by combining the convex and concave curves were compared. As presented in Fig. 8(a), the shapes were determined using the spline approach containing four controlling points, i.e., (35,yD), (40,yE), (52.5,yF), and (65,yG). The four variables of the y-axis coordinates were chosen to ensure a convex-to-concave transition on the height profile. The results of f(a/w) of the six shapes are plotted with two curves of convex (yC=16) and concave (yC=11) as the references. It can be seen that, with the chosen parameters, the combing design strategy is able to generate intermediate curves of f(a/w) that lie between the typical curves of convex and concave designs. The results of the six shapes are clearly divided into two groups. For yD=11 and yE=11.5, which shows a small curvature for the initial convex shape, the function curves exhibits a slightly higher initial value and a rising phase. Compared to the fully convex shape denoted by the black dash line in the figure, the incorporation of concave shape suppresses further increasing and make the curve maintain around a certain value. For yD=13 and yE=14, which shows a relatively large curvature for the initial convex shape, a different type of curve path was presented. No obvious initial rising is found and the curve starts to go up as a/w is over 0.5. As the slope brought by the convex shape is higher which is controlled by yF and yG, the rise in the curve can be hindered more significantly. Among these six shapes, the smallest Max/Min ratio is 1.13 from the Spine(13,14,24,40) for 0.3<a/w<0.6. However, the smallest Max/Min ratio is 1.23 from the Spine(11,11.5,24,40) if a/w<0.65 was considered. Therefore, it suggests that by choosing the appropriate shapes of the spline, a balance might be obtained between the two groups which may lead to a more consistent curve throughout the larger range of crack extension. With such minds, we then manage to find one spline design with better performance. By using a trial and error method, the shape of Spline(11,12,29,50) was found to show both of the advantages of convex and concave designs, which are the stability at the onset of crack extension and a postponed increasing starting point of a/w>0.6. With this shape, the Max/Min ratios are reduced to 1.06 for 0.3<a/w<0.6 and 1.15 for a/w<0.65.

Fig. 8
The schematics of the CT specimens (a) with the spline shape by combining the convex and concave curves and (b) with the combined shape of concave and tapered line, and the corresponding results of f(a/w)
Fig. 8
The schematics of the CT specimens (a) with the spline shape by combining the convex and concave curves and (b) with the combined shape of concave and tapered line, and the corresponding results of f(a/w)
Close modal

In the process of iteratively modifying the shapes of the spline using the four y-axis coordinates of controlling points to search for better results, it is observed that there is no significant improvement when continues to increase the yE and yG which are supposed to lower the Max/Min ratio for a/w<0.65. This may due to that the adopted spline generating method only causes a tiny change on the shape around the controlling points, as shown by the enlarged figure in Fig. 8(a). Therefore, we adopted a convex curve and tapered line combination design as illustrated in Fig. 8(b). It is shown that the tapered line between controlling points (40, yE) and (65,yG) is more favorable than the concave curve. Compared to the curved design, the influence of the yG was more remarkable for yG at smaller values. With the same convex shape at 30<x<40, the tapered design of yG=40 offers the Max/Min ratios of 1.03 for 0.3<a/w<0.6 and 1.12 for a/w<0.65, which is better than the concave design of yG=50. The smaller value of yG is very important considering the material consume and fixture adaptability of the CT specimen.

4 Reinforcement Learning Model for Layout Design

The above analysis indicates that the best design for a constant f(a/w) value is possibly hidden in the huge design space with irregular shape. Among the strategies for solving the inverse problems in design without investigating the entire design space such as heuristic algorithms [33], gradient-based topology optimization [34], and generative networks-based schemes [35,36], the RL algorithms do not rely on the prior knowledge nor large amount of initialization samples, are not likely to yield an local minimum. The RL method can select the actions that maximize the profit by iteratively interacting with the operation environment. In this section, we proposed a novel reinforcement learning system for the layout design of CT specimen based on off-policy Q-learning structure with the FEM as simulation environment.

The major challenge by using the RL on solving the inverse design problem is the proper expression of the problem by the Marov decision process and the action-state update strategies. To demonstrate our proposed RL based design system, the architecture of the model is shown in Fig. 9. The key components in the proposed intelligent design systems include:

  1. State and action. The design problem is converted to the Marov decision process in which the agent would go through seven states and is trained to execute one action from a pool of four possible actions. At the beginning, the design domain is assigned (e.g. w=51, y0=10 as illustrated in the figure) and then the seven states refer to the seven controlling points where x=35,40,45,50,55,60,65. The action is to determine the y-axis coordinate at each state by introducing a variable of m denoting the y increment. The action space is m choosing from [1,2,4,8] with yi+1=yi+m to ensure the increasing trend on the height profile. The shape design in one training episode can then be identified with the completion of this sequential drawing process.

  2. Q-learning algorithm. The core unit of the RL is the decision maker which gives Q-values representing the expected total rewards of taking the corresponding action. By searching the position of maximum Q-value, the next action for the new state can be determined. The Q-values comprise a tabular collection of data which is updated after each training episode to approach the feedback from the FEM simulation. In this work, the simplest form of 7×4 Q-table is used with states as rows and actions as columns. The values are updated via off-policy based on the following Bellman equation.
    (21)
    where Q(s,a) is the current Q-value, Qnew(s,a) is the updated Q-value, α is the learning rate, R(s,a) is the reward, γ is a number between [0,1] and is used to discount the reward as the step progresses, valuing rewards received earlier higher than those received later. Besides, an epsilon-greedy exploration strategy is used to balance exploration and exploitation. During training, the agent selects a random action with probability ε and selects the action with the highest Q-value with probability (1 - ε).
  3. Reward function. The Q-values are updated based on the rewards received for the action taken, even if it is not the optimal action according to the current policy. For our problem, the reward function R(s,a) is calculated by the Min/Max ratio of f(a/w) value from the FEM simulation to accommodate to the goal of maximize the cumulative reward.

Fig. 9
Framework of the reinforcement learning model for the design of compact tension specimens
Fig. 9
Framework of the reinforcement learning model for the design of compact tension specimens
Close modal

5 Result and Discussion

5.1 Training Performance and Parameters Tuning.

The whole training process was written in a python script and executed in the abaqus environment as the reward was generated by the post-processing of the finite element analysis. It is also feasible to train a surrogate machine learning model to replace this numerical analysis. However, the time efficiency would be a problem as a large dataset is normally required to ensure the accuracy of the surrogate model. In the current study, a pre-processing script was also included to automatically build finite element models, apply boundary conditions, and run solvers according to the shape design provided by the trajectory in RL model. Therefore, the training process contains loops of trajectory chosen according to policy, finite element analysis, and Q-value update. The training process ends after a given number of episodes.

The whole design process shown in Fig. 9 has a sparse reward problem. The reward is obtained after all the seven steps are finished instead of a prompt feedback. Therefore, the parameters of the RL model were studied to adapt to an effective and efficient training process. The parameters in the RL model include γ for discount factor, ε for greedy policy, and α for learning rates. A total of four cases were investigated with the parameters shown in Table 1.

Table 1

Parameters adopted in the cases of training of RL model

CaseDiscount factor γGreedy probability εLearning rates α
10.80.80.1
21.00.80.1
31.00.80.2
41.00.90.2
CaseDiscount factor γGreedy probability εLearning rates α
10.80.80.1
21.00.80.1
31.00.80.2
41.00.90.2

The training performance of each case is shown by the curve of reward-episode relation, as shown in Fig. 10. Unlike other optimization method like the genetic algorithm, the RL model does not monotonically improve the design quality with more sample numbers. Some perturbations can be observed in the curves due to the ε-greedy policy adopted in the training process, leading to a randomness in choosing actions at each step. It can be seen that the best performance was found from the case2 and case3 where the averaged reward reaches it historical best with the training episode increases. The best reward can be up to 0.97 after 180 samples in case3. This sample number is very impressive as 180 is only 1% of all the 16,384 possible action trajectories of the proposed Markov decision process, showing the high efficiency of the potential of RL model in designing structures. However, it is also observed that the choice of the parameters of RL model significantly impact the performance.

Fig. 10
The change of the averaged reward with the increasing episodes in the training process of each case. The shaded area indicates the corresponding standard deviation.
Fig. 10
The change of the averaged reward with the increasing episodes in the training process of each case. The shaded area indicates the corresponding standard deviation.
Close modal

The discount factor γ, as presented in the Bellman equation of Eq. (21), allows to model the agent behavior with respect to future rewards. Generally, when the discount factor is low, the agent becomes greedy and choose actions that yield the maximum immediate reward. This is because, when the cumulative reward is calculated, future rewards will have a limited impact due to their heavy discounting. Besides, it also impacts the convergence of the algorithms. In problems of finite horizons (or time-steps), discount factor can be safely set to one. However, in problems of infinite horizons, the discount factor must be set lower than 1, otherwise convergence is not guaranteed. It can be seen that with the discount factor of 0.8 in case1, the model fails to discover better designs in the given training episodes. By adopting the lower value of discount factor, the Q-values for at the first two actions were one order of magnitude smaller compared to that of other cases. This may lead to the improper shape design at the beginning stage, which significantly impacts the reward according to previous analysis.

The ε in the greedy policy marks the tradeoff between exploration and exploitation. The higher the value of ε, the more possible that we chosen actions according to the current policy. This causes issue in exploration as we can get stuck easily at a local optima. Here, we only discussed the influence of the fixed value of ε. However, some adaptive greedy policies have been found to make a better balance at both the initial and subsequent training episodes. Comparative study between case3 and case4 shows that the ε may also greatly influence the training performance. Higher value of ε resulted in smaller deviation but failed to discover the better design even tough the RL model somehow has found several good design at the episode of 10.

The learning rate is the magnitude at which a leap is taken in determining optimal policy. In terms of simple Q-learning, this refers to how much the Q value is updated with each step. When training an RL agent, the learning process should have a visible impact on some metric of problem solving. If the learning rate is set too high, the periodic evaluation score will fluctuate. However, the Q value will change significantly after each optimizer step, which may prevent it from converge to an optimum. Therefore, a trail-and error study on the learning rate should be performed with attention paid on the changes of the reward. In our study, RL model with the learning rate of 0.1 can find optimal design faster than that with learning rate of 0.2. It has to be noted that the chosen parameters are by no means optimal. However, they are tuned within a certain range based on their mathematical meaning in the RL algorithm.

5.2 Optimum Shape Design.

The final proposed shape of CT specimen via the RL model and the corresponding f(a/w) curve are shown in Fig. 11. According to the finalized Q-table, the positions of the controlling points of the optimum shape are y=12,13,17,25,33,41,49, for x=35,40,45,50,55,60,65 respectively. It can be observed that the performance of the RL-designed specimen is better than the best output of the combined curve and line design in the last section. For the crack extension ranges in 0.3<a/w<0.6, the Max/Min ratio is 1.02. The error is less than 2% by replacing the f(a/w) with a constant number of 15.6 in calculating the fracture toughness in Eq. (1). This improved performance may comes from the enlarged design space that were considered in the RL model. Some common features can be found as the slope of the height profile is always positive and gradually increase from the initial low value. However, the RL model plots the transition area more thoroughly than the previous approach, which did not display. The existing design framework cannot prevent the growing trend in f(a/w) for large cracks a/w>0.6, implying that the RL-designed shape is only effective for cracks a/w<0.6.

Fig. 11
The optimum shape design provided by the RL model and the corresponding results of f(a/w) curve
Fig. 11
The optimum shape design provided by the RL model and the corresponding results of f(a/w) curve
Close modal

5.3 Validation by Virtual Testing.

To validate the proposed design of CT specimen for the measurement of Mode I fracture toughness, especially with the effect of R-curve, we built a virtual testing model in which the crack extensions were simulated via cohesive zone model. Two FEM models were created for the standard CT specimen and the proposed design respectively. The deformed models are shown in Fig. 12. The only difference of the model compared to the previous FEM analysis in Sec. 2 is that the whole specimen was modeled instead of the half considering symmetry. Then the properties of fracture toughness were imported by defining the cohesive contact behavior in the interaction module of abaqus.

Fig. 12
The deformed FEM models of (a) ASTM compact tension and (b) RL-designed specimens for virtual Mode I fracture toughness tests at the crack extension of 3 mm. Deformation scale factor is 10.
Fig. 12
The deformed FEM models of (a) ASTM compact tension and (b) RL-designed specimens for virtual Mode I fracture toughness tests at the crack extension of 3 mm. Deformation scale factor is 10.
Close modal

Figure 13 presents the results for the fracture process of materials with a constant critical fracture energy. In this model, the base material was an isotropic material with E=10 GPa and μ=0.3. The cohesive contact behavior was defined with the damage initiation being 60 MPa and critical ERR being 0.33N/mm, referring to the values from Ref. [3]. A displacement control was utilized, with a total of 1 mm applied to each of the loading points over a one-second step period. The FEM models were solved by the abaqus/standard solver. To ensure the stability and the convergence of the crack propagation process, a damage stabilization factor of 1×105 was assigned. The creep dissipated energy was kept to a tiny proportion (<5%) of the energy dissipated by cohesive damage, indicating that the damping factor value was chosen properly. The load–displacement curve at the load-line was shown in Fig. 13(a). For the ASTM standard CT specimen, a progressive decrease in load was observed load following the commencement of the crack propagation, which is the typical pheromone of the Mode I fracture of brittle material. This decrease in load was counterbalanced by the growing values of f(a/w) in Eq. (2), thus making a relatively constant critical ERR calculated by the data reduction method in ASTM E1820 standard. The critical ERR values for CT specimen, as shown in Fig. 13(b), were in good agreement with the material input of 0.33N/mm, indicates the effectiveness of the FEM model as the virtual testing method. The higher values at the initial stage were resulted from the damage stabilization. For the specimen designed by RL, the load-displacement curve shows that first, the structural compliance and failure load is different from that in CT specimen. The load remains nearly unchanged after the onset of crack extension at a lower load value. More importantly, by adopting the constant f(a/w) value of 16, the RL-designed specimen is able to provide as good results of the critical ERR as that by ASTM standard CT specimen. The constant f(a/w) means that the measurement of crack extension is avoid and the data reduction process is simplified. Moreover, the trend of crack resistance curve can be simply observed from the load history.

Fig. 13
Virtual test results of material with a constant critical fracture toughness: (a) load–displacement curve and (b) the calculated fracture toughness with the propagating crack
Fig. 13
Virtual test results of material with a constant critical fracture toughness: (a) load–displacement curve and (b) the calculated fracture toughness with the propagating crack
Close modal
Figure 14 presents the results for the fracture process of materials with a typical growing R-curve feature of critical fracture energy. The input crack resistance curve is illustrated in Fig. 14(b) which starts from 0.2N/mm, grows to 0.5N/mm, and the shape is determined by Eq. (22) (m0 = 0.2, m1 = 0.5, m2 = –1). To incorporate this varying ERR, the contact area between the upper and lower parts of the specimen was discretized into 20 regions. The relevant ERR for each region was determined using Eq. (22) according to their positions. A same feature of the progressive growth and gradual convergence can be found in the load–displacement curve of RL-designed specimen. Therefore, by only recording the load, it is easy to find out the starting and propogated values of the R-curve, i.e., m0 and m1 in Eq. (22). Then, knowing only one data point of crack extension for deriving the m2, the whole R-curve can be determined. However, for the standard CT specimen, such information cannot be obtained from the load history and multiple data points of crack extension are required for generating one complete R-curve. The calculated crack resistance curves of the two specimens are shown in Fig. 14(b). Both the results agree well with the material input. Note that the result of the CT specimen was plotted with multiple data points of crack extension from FEM modeling. The result of RL-designed specimen fitted by Eq. (22) with one data point (ERR value of 0.39N/mm at the crack extension of 1.5 mm) was also presented. It shows that by adopting the shape designed by RL model, one data point during the crack propagating process can provide sufficient accuracy compared to the ASTM CT tests.
(22)
From the above comparative analysis of the virtual testing, two advantages can be claimed by the RL-designed specimen while maintaining similar accuracy as the ASTM CT specimen: The trend of the fracture toughness with the extending crack can be directly read from the load history. By adding just one data point of crack extension, the entire R-curve can be determined. The crack tip measurement can be completely avoid if a constant fracture toughness was assumed.
Fig. 14
Virtual test results of material with R-curve of growing critical fracture toughness: (a) load–displacement curve and (b) the calculated fracture toughness with the propagating crack
Fig. 14
Virtual test results of material with R-curve of growing critical fracture toughness: (a) load–displacement curve and (b) the calculated fracture toughness with the propagating crack
Close modal

6 Conclusion

In this paper, we have proposed a novel geometric design of compact tension specimen to simplify the measurement of Mode I fracture toughness. The data reduction of the current commonly used CT specimen requires the concurrent determination of both loads and crack extensions, which greatly increases the complexity of experimental operations. We revisited the widely-used ASTM compact tension and TDCB specimen, and aimed to avoid the measurement of the crack lengths through the shape design of the Mode I specimen.

First, by performing a theoretical analysis on the deformation of the specimen, it is claimed that the former analysis on the TDCB design fails to provide accurate fracture toughness especially when the crack propagates to a large size (a/w>0.5). The error mainly comes from the neglect of the beam root rotation and the varying height profile.

Second, by taking a throughout study on the geometric design, the independence of the fracture toughness on the crack extension is minimized with the height profile outputted by the machine learning model. It is validated by the virtual testing method that the proposed design shows better performance compared to the CT specimen in ASTM standard and the widely-used TDCB specimen in the literature.

Finally, with the proposed specimen, the experiments for determining the crack resistance curve can be done in the following way:

  1. Cutting the specimen into our proposed shape shown in Fig. 11.

  2. Pin loading for stable Mode I fracture.

  3. Recording the load history. The history of critical fracture toughness can then be calculated by replacing f(a/w) in Eq. (1) with a constant value of 15.5.

  4. If the curve of crack resistance versus crack extensions was desired, capture one (Δa(i), K(i)) at any crack extension. Then, the complete R-curve can be fitted by Kinitial, Kpropagated, and (Δa(i), K(i)) with the form of Eq. (22).

Fracture toughness, especially R-curve, is an important parameter for the evaluation of damage tolerance of composites. Deeper research has been conducted on the analysis of extreme ambient conditions, loading rate dependency, fatigue, and aging. This brings challenge to the experimental measurement of fracture toughness as it is not an easy task even under normal experimental conditions. The current work simplifies the test procedure without introducing any new equipment or techniques, which would provide insights into the experiments in the aforementioned situations. The work also shows the great potential of the machine learning approaches for the structural design with the ability to efficiently explore the vast design space while avoiding the problem of dropping into local optimum. Our ongoing further work extends the current ML-based design system on the problem of Mode II fracture toughness measurement.

Acknowledgment

This work was supported in part by the Project of Hetao Shenzhen-Hong Kong Science and Technology Innovation Cooperation Zone (HZQB-KCZYB-2020083), the Department of Science and Technology of Guangdong Province (Project #: 2022A0505030023), and the international partnership program of the Chinese Academy of Sciences (Grant No. 025GJHZ2022103FN).

Conflict of Interest

There are no conflicts of interest.

Data Availability Statement

The datasets generated and supporting the findings of this article are obtainable from the corresponding author upon reasonable request.

Appendix: TDCB Compliance Analysis

With the quadratic differential equations of Eqs. (8)(11) for the two sections of TDCB, the form of vertical displacement can be obtained
(A1)
and
(A2)
The four integration constants A1, A2, B1, and B2 are then solved by the four equations of Eqs. (12)(15) as
  • The continuity of displacement u2(x=x0)=u1(x=x0)
    (A3)
  • The continuity of rotation ϕ1(x=x0)=ϕ2(x=x0)
    (A4)
  • The boundary condition of displacement u2(x=a)=0
    (A5)
  • The boundary condition of rotation ϕ2(x=a)=0
    (A6)
Thus, with the above four equations, the vertical displacement at the load point can be calculated as
(A7)
where
(A8)
Then derivative of the contribution of the beam deflection is given by
(A9)
The correction of the tapered neutral axis is simply introduced by multiplying Eq. (A9) with (1/(1+k2)+1)/2. The governing differential equation for the beam deflection considering a elastic foundation with stiffness k´ is
(A10)
where H(x)=0 for x<a and H(x)=1 for x>a. The solution to this equations is obtained by considering, (i) the continuity of the deflection and rotation at x=a; (ii) the shear force and free end conditions at the left- and right-hand ends respectively. The results can be written as
(A11)
where
(A12)
and
(A13)
The beam root rotation (at x=0 in Eq. (A11)) can then be estimated as shown in Eq. (17). The derivative
(A14)
with
(A15)
By incorporating Eqs. (A9), (A14), and (A15) into Eq. (19), the derivative of compliance of TDCB specimen can be calculated, and thus the f(a/w) function can be determined by Eq. (4).

References

1.
Pinho
,
S.
,
Iannucci
,
L.
, and
Robinson
,
P.
,
2006
, “
Physically-Based Failure Models and Criteria for Laminated Fibre-Reinforced Composites With Emphasis on Fibre Kinking: Part I: Development
,”
Compos. Part A: Appl. Sci. Manuf.
,
37
(
1
), pp.
63
73
.
2.
Shor
,
O.
, and
Vaziri
,
R.
,
2017
, “
Application of the Local Cohesive Zone Method to Numerical Simulation of Composite Structures Under Impact Loading
,”
Int. J. Impact Eng.
,
104
, pp.
127
149
.
3.
Qiu
,
C.
,
Guan
,
Z.
, and
Du
,
S.
,
2022
, “
An Improved Characteristic Length Method for Predicting the Single Bolt Joint Bearing Strength Considering Secondary Bending Effect
,”
Mech. Adv. Mater. Struct.
,
29
(
10
), pp.
1405
1417
.
4.
Li
,
S.
,
Sitnikova
,
E.
,
Liang
,
Y.
, and
Kaddour
,
A.-S.
,
2017
, “
The Tsai-Wu Failure Criterion Rationalised in the Context of UD Composites
,”
Compos. Part A: Appl. Sci. Manuf.
,
102
, pp.
207
217
.
5.
Huang
,
Z.-M.
,
2004
, “
A Bridging Model Prediction of the Ultimate Strength of Composite Laminates Subjected to Biaxial Loads
,”
Compos. Sci. Technol.
,
64
(
3–4
), pp.
395
448
.
6.
Wiegand
,
J.
,
Petrinic
,
N.
, and
Elliott
,
B.
,
2008
, “
An Algorithm for Determination of the Fracture Angle for the Three-Dimensional Puck Matrix Failure Criterion for UD Composites
,”
Compos. Sci. Technol.
,
68
(
12
), pp.
2511
2517
.
7.
Reinoso
,
J.
,
Catalanotti
,
G.
,
Blázquez
,
A.
,
Areias
,
P.
,
Camanho
,
P.
, and
París
,
F.
,
2017
, “
A Consistent Anisotropic Damage Model for Laminated Fiber-Reinforced Composites Using the 3D-Version of the Puck Failure Criterion
,”
Int. J. Solids. Struct.
,
126
, pp.
37
53
.
8.
Park
,
K.
, and
Paulino
,
G. H.
,
2011
, “
Cohesive Zone Models: a Critical Review of Traction-Separation Relationships Across Fracture Surfaces
,”
ASME Appl. Mech. Rev.
,
64
(
6
), p.
060802
.
9.
Alfano
,
G.
, and
Crisfield
,
M.
,
2001
, “
Finite Element Interface Models for the Delamination Analysis of Laminated Composites: Mechanical and Computational Issues
,”
Int. J. Numer. Methods Eng.
,
50
(
7
), pp.
1701
1736
.
10.
Gong
,
Y.
,
Hou
,
Y.
,
Zhao
,
L.
,
Li
,
W.
,
Zhang
,
J.
, and
Hu
,
N.
,
2020
, “
A Modified Mode I Cohesive Zone Model for the Delamination Growth in DCB Laminates With the Effect of Fiber Bridging
,”
Int. J. Mech. Sci.
,
176
, p.
105514
.
11.
Grein
,
C.
,
Béguelin
,
P.
, and
Kausch
,
H.-H.
,
2003
, “A New Way for Polymer Characterisation Using a Combined Approach LEFM–Plastic Zone Corrected LEFM,”
European Structural Integrity Society
,
B. R. K.
Blackman
,
A.
Pavan
, and
J. G.
Williams
, eds., Vol.
32
,
Elsevier
,
Les Diablerets, Switzerland
, pp.
129
141
.
12.
Hou
,
F.
, and
Hong
,
S.
,
2014
, “
Characterization of R-Curve Behavior of Translaminar Crack Growth in Cross-Ply Composite Laminates Using Digital Image Correlation
,”
Eng. Fract. Mech.
,
117
, pp.
51
70
.
13.
Xu
,
X.
,
Wisnom
,
M. R.
, and
Hallett
,
S. R.
,
2019
, “
Deducing the R-Curve for Trans-Laminar Fracture From a Virtual Over-Height Compact Tension (OCT) Test
,”
Compos. Part A: Appl. Sci. Manuf.
,
118
, pp.
162
170
.
14.
Xu
,
X.
,
Sun
,
X.
, and
Wisnom
,
M. R.
,
2021
, “
Initial R-Curves for Trans-Laminar Fracture of Quasi-Isotropic Carbon/Epoxy Laminates From Specimens With Increasing Size
,”
Compos. Sci. Technol.
,
216
, p.
109077
.
15.
Xu
,
X.
,
Takeda
,
S.-I.
, and
Wisnom
,
M. R.
,
2023
, “
Investigation of Fracture Process Zone Development in Quasi-Isotropic Carbon/Epoxy Laminates Using In Situ and Ex Situ X-ray Computed Tomography
,”
Compos. Part A: Appl. Sci. Manuf.
,
166
, p.
107395
.
16.
Laffan
,
M.
,
Pinho
,
S.
,
Robinson
,
P.
, and
Iannucci
,
L.
,
2010
, “
Measurement of the In Situ Ply Fracture Toughness Associated With Mode I Fibre Tensile Failure in FRP. Part I: Data Reduction
,”
Compos. Sci. Technol.
,
70
(
4
), pp.
606
613
.
17.
Yoshihara
,
H.
, and
Kawamura
,
T.
,
2006
, “
Mode I Fracture Toughness Estimation of Wood by DCB Test
,”
Compos. Part A: Appl. Sci. Manuf.
,
37
(
11
), pp.
2105
2113
.
18.
Xu
,
W.
, and
Guo
,
Z.
,
2018
, “
A Simple Method for Determining the Mode I Interlaminar Fracture Toughness of Composite Without Measuring the Growing Crack Length
,”
Eng. Fract. Mech.
,
191
, pp.
476
485
.
19.
Cidade
,
R. A.
,
Castro
,
D. S.
,
Castrodeza
,
E. M.
,
Kuhn
,
P.
,
Catalanotti
,
G.
,
Xavier
,
J.
, and
Camanho
,
P. P.
,
2019
, “
Determination of Mode I Dynamic Fracture Toughness of IM7-8552 Composites by Digital Image Correlation and Machine Learning
,”
Compos. Struct.
,
210
, pp.
707
714
.
20.
Arteiro
,
A.
,
Hayati
,
M.
, and
Camanho
,
P.
,
2014
, “
Determination of the Mode I Crack Resistance Curve of Polymer Composites Using the Size-Effect Law
,”
Eng. Fract. Mech.
,
118
, pp.
49
65
.
21.
Catalanotti
,
G.
, and
Xavier
,
J.
,
2015
, “
Measurement of the Mode II Intralaminar Fracture Toughness and R-Curve of Polymer Composites Using a Modified Iosipescu Specimen and the Size Effect Law
,”
Eng. Fract. Mech.
,
138
, pp.
202
214
.
22.
Qiu
,
C.
,
Han
,
Y.
,
Shanmugam
,
L.
,
Guan
,
Z.
,
Zhang
,
Z.
,
Du
,
S.
, and
Yang
,
J.
,
2021
, “
Machine Learning-Based Prediction of the Translaminar R-Curve of Composites From Simple Tensile Test of Pre-Cracked Samples
,”
J. Micromechan. Mole. Phys.
,
6
(
1
), p.
2050017
.
23.
Laffan
,
M.
,
Pinho
,
S.
,
Robinson
,
P.
, and
McMillan
,
A.
,
2012
, “
Translaminar Fracture Toughness Testing of Composites: A Review
,”
Polym. Test.
,
31
(
3
), pp.
481
489
.
24.
Cher
,
W.
,
2006
, “Development of a Test to Measure Energy Absorption in Fibre Failure Modes of Woven Composite Laminates,” Technical Report, Final Year Project Report.
25.
Blanco
,
N.
,
Trias
,
D.
,
Pinho
,
S.
, and
Robinson
,
P.
,
2014
, “
Intralaminar Fracture Toughness Characterisation of Woven Composite Laminates. Part I: Design and Analysis of a Compact Tension (CT) Specimen
,”
Eng. Fract. Mech.
,
131
, pp.
349
360
.
26.
Katafiasz
,
T.
,
Iannucci
,
L.
, and
Greenhalgh
,
E.
,
2019
, “
Development of a Novel Compact Tension Specimen to Mitigate Premature Compression and Buckling Failure Modes Within Fibre Hybrid Epoxy Composites
,”
Compos. Struct.
,
207
, pp.
93
107
.
27.
Brown
,
E. N.
,
Sottos
,
N. R.
, and
White
,
S. R.
,
2002
, “
Fracture Testing of a Self-Healing Polymer Composite
,”
Exp. Mech.
,
42
, pp.
372
379
.
28.
Brown
,
E.
,
2011
, “
Use of the Tapered Double-Cantilever Beam Geometry for Fracture Toughness Measurements and Its Application to the Quantification of Self-Healing
,”
J. Strain Anal. Eng. Des.
,
46
(
3
), pp.
167
186
.
29.
Qiao
,
P.
,
Wang
,
J.
, and
Davalos
,
J. F.
,
2003
, “
Tapered Beam on Elastic Foundation Model for Compliance Rate Change of TDCB Specimen
,”
Eng. Fract. Mech.
,
70
(
2
), pp.
339
353
.
30.
Blackman
,
B.
,
Hadavinia
,
H.
,
Kinloch
,
A.
,
Paraschi
,
M.
, and
Williams
,
J.
,
2003
, “
The Calculation of Adhesive Fracture Energies in Mode I: Revisiting the Tapered Double Cantilever Beam (TDCB) Test
,”
Eng. Fract. Mech.
,
70
(
2
), pp.
233
248
.
31.
ASTM E1820-18
,
2018
. “Standard Test Method for Measurement of Fracture Toughness,” Standard, American Society for Testing and Materials.
32.
Qiu
,
C.
,
Guan
,
Z.
,
Du
,
S.
,
Li
,
Z.
, and
Huang
,
Y.
,
2019
, “
Prediction of the Net-Tension Strength of Single Bolt Joint Using the 0° Ply Strength and Fracture Toughness
,”
Compos. Part B: Eng.
,
161
, pp.
386
394
.
33.
Qiu
,
C.
,
Han
,
Y.
,
Shanmugam
,
L.
,
Jiang
,
F.
,
Guan
,
Z.
,
Du
,
S.
, and
Yang
,
J.
,
2022
, “
An Even-Load-Distribution Design for Composite Bolted Joints Using a Novel Circuit Model and Neural Network
,”
Compos. Struct.
,
279
, p.
114709
.
34.
Qian
,
C.
, and
Ye
,
W.
,
2021
, “
Accelerating Gradient-Based Topology Optimization Design With Dual-Model Artificial Neural Networks
,”
Struct. Multidiscipl. Optim.
,
63
, pp.
1687
1707
.
35.
Qiu
,
C.
,
Han
,
Y.
,
Shanmugam
,
L.
,
Zhao
,
Y.
,
Dong
,
S.
,
Du
,
S.
, and
Yang
,
J.
,
2022
, “
A Deep Learning-Based Composite Design Strategy for Efficient Selection of Material and Layup Sequences From a Given Database
,”
Compos. Sci. Technol.
,
230
, p.
109154
.
36.
Mao
,
Y.
,
He
,
Q.
, and
Zhao
,
X.
,
2020
, “
Designing Complex Architectured Materials With Generative Adversarial Networks
,”
Sci. Adv.
,
6
(
17
), p.
eaaz4169
.