Zum Hauptinhalt springen

Development and Numerical Implementation of a Modified Mixed-Mode Traction-Separation Law for the Simulation of Interlaminar Fracture of Co-Consolidated Thermoplastic Laminates Considering the Effect of Fiber Bridging.

Sioutis, I ; Tserpes, K
In: Materials (Basel, Switzerland), Jg. 15 (2022-07-22), Heft 15
Online academicJournal

Development and Numerical Implementation of a Modified Mixed-Mode Traction–Separation Law for the Simulation of Interlaminar Fracture of Co-Consolidated Thermoplastic Laminates Considering the Effect of Fiber Bridging 

In the present work, a numerical model based on the cohesive zone modeling (CZM) approach has been developed to simulate mixed-mode fracture of co-consolidated low melt polyaryletherketone thermoplastic laminates by considering fiber bridging. A modified traction separation law of a tri-linear form has been developed by superimposing the bi-linear behaviors of the matrix and fibers. Initially, the data from mode I (DCB) and mode II (ENF) fracture toughness tests were used to construct the R-curves of the joints in the opening and sliding directions. The constructed curves were incorporated into the numerical models employing a user-defined material subroutine developed in the LS-Dyna finite element (FE) code. A numerical method was used to extract the fiber bridging law directly from the simulation results, thus eliminating the need for the continuous monitoring of crack opening displacement during testing. The final cohesive model was implemented via two identical FE models to simulate the fracture of a Single-Lap-Shear specimen, in which a considerable amount of fiber bridging was observed on the fracture area. The numerical results showed that the developed model presented improved accuracy in comparison to the CZM with the bi-linear traction–separation law (T–SL) in terms of the predicted strength of the joint.

Keywords: thermoplastics; co-consolidated joints; fracture toughness; cohesive zone modeling; finite element analysis

1. Introduction

Polymeric matrix composite materials have become the norm for the aerospace industry in the past few decades. Various factors led to this transition over their metallic counterparts, such as their high strength-to-weight ratio and application dependent properties' designation [[1]]. The multi-material composition of fiber reinforced polymers has led to the emergence of numerous mechanical phenomena, inducing complexity in the study of their fracture behavior.

Fiber bridging constitutes such a phenomenon that has been under investigation and impacts the delamination behavior of composite materials. Generally, fiber bridging is defined as the separation of fibers or fiber bundles during the debonding of adjacent plies, thus increasing the interlaminar fracture toughness. The severity of this phenomenon is attributed by factors such as fiber orientation and matrix–fiber interface strength [[2]]. Numerous works have been published, focusing on both the definition of the mechanical behavior and the usage of fiber bridging in favor of the structure's integrity through thickness reinforcement. Available studies include analytical approaches and experimental works on fiber bridging. Budiansky et al. [[3]], Hutchinson et al. [[4]], Cox et al. [[5]] and Bao et al. [[6]] have developed micromechanical models in order to describe fiber bridging. Other efforts in the development of geometry independent bridging laws have been conducted by Sørensen and Jacobsen [[7]] and other researchers [[8], [10], [12], [14]]. Other works also include extensive experimental and numerical efforts to accurately predict and model the fiber bridging mechanism [[15], [17]].

The optimization and integration of the mechanism in question for composite structures can be achieved through numerical simulations. However, the most commonly used bilinear cohesive zone modeling in composite joints cannot accurately model the fracture toughness increment due to the bridging. This is due to the fact that the T–SL in both loading directions only represents the mechanical separation effect of the matrix interface. Efforts have been made towards the integration of the fiber bridging phenomenon in numerical models. Afshar et al. [[19]] have developed a numerical model through the extended finite element method (XFEM) by simulating the effect of bridging in the interface through non-linear springs.

The most common numerical approaches found in the literature are based on a modified cohesive zone modeling technique, where the effect of both the matrix and fibers are superimposed in a final multilinear T–SL. Such a technique can be found in the work of Heidari-Rarani et al. [[20]], where the superposition of two CZMs, representing the matrix interface and fiber bridging, respectively, is performed. The resulting T–SL follows a tri-linear form emerging from the simultaneous effect of the two bi-linear CZMs peaking at the same displacement jump.

Gong et al. [[21]] presented a model in the same principle as described above; however, in this case the two bi-linear CZMs did not have the same peak. The bridging part of the T–SL was initiated when the interface failed completely, which also abides to the physical behavior of the phenomenon. According to the above study, the proposed model [[22]] was experimentally validated and simplification techniques for more straight forward law extraction were presented [[23]].

Although a considerable amount of work has been conducted on the numerical simulation of the fiber bridging phenomenon, the applications are limited only on thermoset matrix composites to the authors' knowledge. Additionally, in the case of thermoplastic laminates, where joining techniques such as welding are available, there is potential for further study on joints, not only for delamination. Considering that, the present work focused on the development of such a model, incorporating fiber bridging effects in low melt PAEK matrix laminates loaded in mixed mode conditions. This paper includes the analytical, experimental, and numerical procedure for the determination of a modified cohesive zone model. The general objective of the work was to simulate the bridging behavior present in previously conducted tests on mixed-mode Single-Lap-Shear (SLS) specimens, where the simple bilinear law did not manage to accurately approach the results.

In comparison to previously conducted studies, this work incorporated the application and further validation of the modified traction–separation law for fiber bridging simulation in joined thermoplastic composites. A complete experimental characterization of the mixed mode bridging behavior was performed. The development of two new computational tools contributed both in the more straightforward experimental results' extraction and the more practical implementation of the model in question. Overall, the developed model adequately predicted the experimental results and was proved efficient in simulating the fiber bridging phenomenon.

2. Theoretical: The Modified Traction–Separation Law

The incorporation of the fiber bridging mechanism into the cohesive material model was achieved through the employment of a tri-linear T–SL, recently proposed by Gong et al. [[21]]. This section briefly describes the principle and theory of this cohesive zone model.

The representation of the mechanical behavior of the interface is described by a tri-linear T–SL (OABC), which accrues from the superposition of two simple bi-linear CZMs (Figure 1). For the representation of the matrix interface separation, the bi-linear part (ODE) is employed. For the 0δ0 displacement jump of the cohesive, linear elastic behavior is prescribed with K1 stiffness, followed by the degradation part (DE) of the law. In the same principle, the effect of fiber bridging is modeled with another bi-linear law (OBC), where the onset of its degradation coincides with the total failure of the matrix interface. Finally, by superimposing the described components, a tri-linear traction–separation law (OABC) is obtained [[21]].

The tri-linear CZM is described by the following equation:

(1) σ=(1d)Dij0d

where the initial stiffness tensor is defined as

(2) Dij0=δ¯ijK0

and the global damage variable is divided as follows:

(3) d={0,δ0(1KABK0)(1δ0δ),δ0<δδb1KBCK0(1δfδ),δb<δδf1,δδf

where K0 , KAB and KBC stand for the (OA), (AB) and (BC) lines' tangents, respectively. Other crucial parameters required for the composition of the T–SL are the initial fracture toughness Gini , the propagation fracture toughness Gprop , the interfacial strength σ0 , the bridging stress σb , as well as the stiffnesses K1 (or KOD ) and K2 (or KOB ). In the case of mixed-mode loading, each of these parameters needed to be defined both for mode I and mode II.

3. Overall Approach

The development of the final modified T–SL for the co-consolidated laminates consisted of experimental, analytical, and numerical parts. Initially, the required tests, mentioned in Section 2, were conducted; specifically, double cantilever beam (DCB) specimens' tests for mode I and edge notched flexure (ENF) specimens' tests for mode II. The data acquired from this first step are the fracture toughness values, as well as the resistance curves, which are indicative of the bridging behavior of the specimens. However, the latter could not be generalized to any coupon for the modeling of fiber bridging as they are geometry dependent. To overcome this issue, a bridging law for each loading mode needed to be defined.

The technique involved extracting the bridging law and, subsequently, the bridging stress σb is described in [[23]] and comprises a simplified procedure for measuring the crack opening displacement (COD) in the DCB and the crack sliding displacement (CSD) in the ENF specimens, thus eliminating the need for complex and time-consuming digital image correlation monitoring techniques during testing. The R-curves extracted from the two mechanical tests were introduced in two FE models simulating the two tests, respectively. The critical energy release rate or fracture toughness values obeyed the R-curve behavior and the COD could be measured numerically. In this manner the bridging stress–crack propagation curve could be plotted using the following equation:

(4) σb=Gδ

where δ stands for the crack opening or sliding displacement. With this last step, all the parameters needed to construct the trilinear T–SL became available.

Finally, two different options for the numerical implementation of the modified mixed-mode CZM were followed, including the analysis through an already available LS Dyna's material model and the development of a user defined subroutine. The latter two methods were employed for the simulation of SLS specimens, and the results were compared and validated upon available test data. Furthermore, the numerical results that were obtained from the modified models were compared with the output of a simple bilinear T–SL model.

Overall, the procedure described above was divided in three stages: development, implementation, and validation (Figure 2). In the first step, the mandatory experimental, numerical, and analytical data were acquired. Subsequently these were used as input in the numerical models of the implementation stage and finally, the results were compared and validated throughout the third stage. In the following sections, the three stages are analyzed in detail.

4. Stage 1: Development

4.1. Experimental

4.1.1. Mode I—DCB

The mechanical testing of laminates in mode I loading conditions was performed according to the ASTM D5528-01 standard [[24], [26]]. A total of eight co-consolidated specimens were cut in 150 mm length, 25 mm width and 1.8 mm thickness for each substrate. The plies followed a quasi-isotropic stacking [0°/45°/90°/45°/−45°]s and the predefined crack was applied with a 63 mm Kapton film insert between the laminates. The prepreg material of the laminates was a LM-PAEK-based matrix, reinforced with TC1225 carbon fibers, provided by Toray. The experimental set-up of the mode I test is illustrated in Figure 3.

Piano hinge tabs were bonded to the edges of the cantilever beam in order to apply the opening load, according to the standard. The tests were held in room temperature conditions, prescribing a constant crosshead displacement rate of 1 mm/min. During the loading of the specimens, the crack propagation was optically monitored through a digital microscope.

The fracture toughness values for each state of the propagation were obtained through the modified beam theory method. Finally, the R-curves and the GI,ini and GI,prop properties of the specimens were computed.

The mechanical testing conducted on DCB specimens showed similar response and close crack initiation load values for all the specimens; indicatively, in Figure 4, the load–displacement curves for the eight specimens (DCB02–DCB09) are presented. Figure 5 depicts a characteristic R-curve for the co-consolidated laminates. The initiation and propagation fracture toughness values in this case were roughly 1.1 N/mm and 2.1 N/mm, respectively, while the length of the crack propagation for the stabilization of the fracture toughness to a steady value, also defined as the bridging length, was about 40 mm.

In Figure 6, the fractured surfaces along the DCB specimens are shown. It is obvious that the propagation of the crack suffered strong stick-slip behavior, which is noticed both from the color inhomogeneity of the surface and the load–displacement curves. Additionally, loose fibers and bridged fiber bundles can be found by a closer surface inspection, confirming the bridging behavior presented by the extracted R-Curve.

4.1.2. Mode II—ENF

The mode II tests were performed according to the AITM 1.0006 standard [[27], [29]]. Six specimens, with adherents of 115 mm by 25 mm dimensions and a 45 mm predefined cracking, were studied. In the same principle as mode I tests, the crack propagation under the constantly increasing load was monitored through a digital microscope [[30]]. The testing machine used was an MTS universal with a load cell capacity of 100 kN (Figure 7).

Figure 8 depicts the load–displacement response of the ENF specimens, as obtained from mode II testing. In this case, the ENF-05 specimen was excluded from the study, as a considerable difference was recorded due to remaining cohesion at the Kapton region and, subsequently, falsely increased stiffness and interlaminar strength. Specimens ENF-04 and ENF-06 also presented a slightly higher maximum load prior to failure, possibly for the same cause as ENF-05, although the discrepancy was considered acceptable.

Figure 9 shows the characteristic fracture toughness–crack propagation curve, where the initiation GII value was measured quite low, at about 0.5 N/mm, and the steady state GII value at 2.6 N/mm. For this test, the bridging length was considerably lower from the mode I test, lying at 5 mm. It is noteworthy that due to the nature of the three-point bending test, the load–crack propagated length data recording proved to be quite difficult, rendering the acquisition of more data points on the R-curve impossible.

The fracture surfaces of the ENF specimens after testing, as shown in Figure 10, indicated a considerably lower amount of bridged and loose fibers, differing to the previous case. As expected, the crack stick-slip phenomenon was visually absent in mode II.

4.1.3. Mixed Mode SLS

The mixed mode-loaded SLS specimens used for the implementation and validation stage of the model are briefly described. The experimental procedure was based on the ASTM 5868-01 standard. The stacking sequence of the composite laminates was [0°/−45°/45°/90°/45°/−45°]s, joined via co-consolidation at the overlapped area. The dimensions of the substrates were 101.6 mm in length, 25.4 mm wide and a total number of six specimens were tested. The coupons were loaded in a quasi-static displacement rate of 1 mm/min until final rupture.

4.2. Numerical Part

The numerical aspect of the present stage incorporates FE simulations for the numerical estimation of the bridging law in DCB and ENF specimens. The numerical analyses were performed using the commercial FE suite, LS Dyna, using 8-noded reduced integrated solid elements for the composite laminates (Figure 11) [[31], [33]].

User Defined Subroutine—UMAT_RCurve

The numerical part also includes the development of the user defined routines in order to introduce the experimental R-curves in the DCB and ENF models (UMAT_RCurve). Simulations were performed in LS Dyna, while the proprietary scripting was written in FORTRAN 77 programming language and compiled through Intel Visual Fortran Compiler 2010. The developed subroutine was implemented through the LS Dyna's available material model UMAT43c, which is suitable for modeling three dimensional cohesive elements [[34], [36]].

The mechanical response of the user defined model was exactly the same as a simple bi-linear CZM, with the exception that the fracture toughness in each loading mode was not a constant value but was position-dependent, based on the experimental R-curves. In order to implement this, a cohesive element position reading algorithm had to be written.

As an input, the total coupon length ( lplate ) and the pre-cracked length ( a0 ) were required. Through an iterative procedure, the nodal coordinates of the cohesive elements were obtained and, subsequently, the distance from the element's center to the crack tip ( da(i) ) was calculated (Figure 12). By these means, the fracture toughness for each cohesive element was prescribed from the R-curve, which also constitutes a subroutine's input.

The objective of this numerical step was to bypass the need for complex experimental monitoring techniques for the crack opening displacement (COD) and crack sliding displacement (CSD) (Figure 13) by directly extracting the desired curve through FE models of the DCB and ENF specimens. This procedure has been proposed in [[23]]; however, it has only been applied for mode I loading.

Based on the procedure described above, the fracture toughness distribution along the cohesive elements at the interface of the DCB is shown in Figure 14. The lowest G value was applied at the crack-front elements, increasing according to the experimentally defined R-curve. Similarly, the same principle was applied to the ENF cohesive interface.

The COD and CSD computed from the two numerical analyses are presented in Figure 15. In both cases, the relative nodal displacements between the joined laminates were recorded for each timestep, in the appropriate directions for each loading case, in relation to the modeled crosshead displacement, imposed by the boundary conditions. The curves showed expected behavior both for mode I and mode II, as indicated by previous studies [[23], [37], [39]]. In Figure 15a, the relative nodal displacement reached a maximum value of about 23 mm, following a slightly parabolic increase. In Figure 15b, the relative nodal displacement had a maximum relative nodal displacement of 0.35 mm, presenting a plateau at the crack initiation point.

4.3. Analytical Part

In this subsection, the calculations for the mixed-mode tri-linear T–SL are presented. As the first requirement, the bridging stress could be calculated from the differentiation of the fracture toughness with respect to the COD or the CSD, depending on the loading mode (Equation (4)). In this way, the bridging stress–opening/sliding displacement curves could be drawn, as shown in Figure 16 and Figure 17. The first points in these two diagrams corresponded to mode I and II bridging strength, specifically σI,b=4.96MPa and σII,b=τb=22.4MPa .

According to the literature [[21]], the peak traction of the matrix interface contributions ( σ0,I , σ0,II ) can be approximated by a percentage of 55–65% of the interfacial matrix strength in normal and shearing directions, respectively. In this case, a 0.65 ratio was selected, resulting in:

(5) σ0,I=0.65T=55.9MPa

(6) σ0,II=0.65S=27.3MPa

K1 and K2 describe the stiffnesses for the two superimposed T–SLs to the corresponding loading mode. The K1 value can be calculated by the following equation

(7) K1=2EZh

where EZ was the through the thickness laminates' stiffness, and h half of the DCB or ENF specimen thickness. K2 was defined through simple trigonometric calculations from the T–SL schematic, resulting in K1=9.136MPa and K2=0.143MPa .

Furthermore, the displacement jumps for each loading mode could be calculated in the following manner:

(8) δ0,I=σ0,IK1,I=0.006mm

(9) δb,I=2GI,iniσ0,I=0.04mm

(10) δf,I=2(GI,propGI,ini)σb,I=0.389mm

Following the same procedure for mode II loading, δ0,II , δb,II and δf,II could be calculated as 0.004 mm, 0.049 mm, and 0.182 mm, respectively. Ultimately, the final form of the mixed-mode tri-linear T–SL is shown in Figure 18.

5. Stage 2: Implementation

Numerical analyses were held in this stage to determine the behavior for the co-consolidated laminates under mixed mode loading. The simple bilinear T–SL was implemented through the LS Dyna's MAT_138 material model [[40]]. The numerical implementation of the finally developed model of the modified tri-linear T–SL was studied in two ways.

5.1. User Defined Subroutine—UMAT_Tril

The numerical application of the final modified T–SL required the development of another subroutine, able to follow the trilinear traction–separation behavior of the cohesive elements. The developed subroutine, UMAT_Tril, required the user input data previously calculated from the experimental, numerical, and analytical procedures, namely the bridging strength, the interfacial strength, the initiation, and propagation fracture toughness values, the initial stiffnesses of the matrix and fiber bridging elastic response. These parameters needed to be defined prior to the analysis, both for mode I and II loading conditions, as the routine calculates the mixed-mode displacements and stresses of the cohesive elements. In Figure 19, the flowchart of the iterative calculation procedure for every element in UMAT_Tril is depicted.

Based on the element's mixed-mode displacement for each step, the algorithm selected its status: 1. elastic response; 2. matrix interface degradation; 3. fiber bridging contribution; 4. total failure. Each damage variable indicated the degradation percentage of the corresponding modeled part. The two main damage variables d1 and d2, corresponding to the matrix interface and bridging effect degradation, respectively, were constantly available throughout the computational procedure. This enabled the definition of the damage status for every cohesive element in each time step of the simulation.

5.2. LS Dyna's MAT_186

Another implementation technique followed for the application of the tri-linear T–SL was through the employment of LS Dyna's MAT_186 (MAT_COHESIVE_GENERAL) material model, which allowed the definition of user-defined T–SL by prescribing a normalized curve (Figure 20). Different cohesive responses were prescribed for the two loading modes, thus making the mixed-mode simulation possible. The previously developed subroutine UMAT_Tril worked in the exact same computational way as MAT_186, so identical results were to be expected.

6. Stage 3: Final Results and Validation

In this section, the final mechanical responses of the mixed-mode SLS specimens from the numerical simulations are presented. The numerical results included the output data for the trilinear T–SL modeled interfaces, and then they were compared to the bilinear modeled interface response and experimentally acquired data.

Figure 21 depicts the load displacement curves computed numerically. The experimentally obtained curve peaked at 16 MPa prior to failure. Significantly lower was the maximum load for the bilinearly modeled coupon. Such a difference was expected, as MAT_138 did not reckon in the fiber bridging effect and was extensively present on the failed areas of the tested SLS specimens (Figure 22). The mechanical response of the two numerical models (MAT_186 & UMAT_Tril) was virtually the same due to their identical computational behavior. The models based on the modified traction–separation law managed to approach the experimental behavior more accurately than the simple bilinear model. Although the load–displacement curves did not fully capture the real-world mechanical behavior, considerable improvement occurred.

The damage variables d1 and d2 obtained from the UMAT_Tril simulation for the final timestep prior to the failure initiation are plotted in Figure 23a,b. A damage region contour was also captured (Figure 23c), indicating each element's degradation region. Such a result led to the conclusion that the developed models presented quite an accurate representation of the complex fiber-bridging phenomenon, as the damage progression and failure of the tested specimens showed a similar pattern by examining the imprints of the fractured surfaces (Figure 22).

7. Conclusions

In the present work, a model for simulating fiber bridging along the co-consolidated interface of thermoplastic laminates was developed, based on a modified trilinear T–SL.

The conclusions emerging from the study are listed below:

  • Overall, the procedure described above was proven adequately accurate for simulating the fiber bridging behavior of mixed mode-loaded co-consolidated thermoplastic laminates. Although the numerical results were not totally identical with the mechanical tests' data, the improvement over the traditional bilinear cohesive zone modeling was quite considerable. The trilinear T–SL managed to simulate the increased fracture properties due to fiber bridging, following the mechanical aspects of the phenomenon.
  • The developed subroutine, UMAT_RCurve, proved to be an efficient way to numerically measure the crack opening/sliding displacements in both loading modes. A significant amount of experimental work is omitted and maintained though the validity and accuracy of the results.
  • In the final numerical step, where two implementation options were studied, the output was the same. However, in the case of the UMAT_Tril subroutine, additional flexibility in terms of results' monitoring was given. The status of every cohesive element's degradation could be recorded for each timestep, offering the opportunity for meticulous CZM behavior observation during failure.

Finally, it is evident that such a procedure added complexity to the numerical simulation when compared to the mostly used bilinear T–SL. With this in mind, this technique should be used in cases where the fiber bridging phenomenon is prominent, noticeably affecting the mechanical response of the joint.

Apart from the application on co-consolidated joints, this technique is feasible to be applied in any other type of joint for thermoplastic laminates. Thermoplastic welding, for example, constituting a fusion procedure, usually without the integration of third material into the joint, is highly likely to develop the ideal conditions for extensive bridging of the fibers.

Figures

Graph: Figure 1 A schematic of the modified T–SL (ODE: Matrix interface T–SL; OBC: Fiber bridging T–SL; OABC: Combined trilinear T–SL).

Graph: Figure 2 Modeling procedure flowchart.

Graph: Figure 3 Mode I loading experimental set up.

Graph: Figure 4 Mode I experimental load–displacement curves.

Graph: Figure 5 Characteristic R-curve extracted from Mode I tests.

Graph: Figure 6 DCB specimens' fractured surfaces after testing.

Graph: Figure 7 Mode II loading experimental configuration.

Graph: Figure 8 Mode II experimental load–displacement curves.

Graph: Figure 9 Characteristic R-curve extracted from Mode II tests.

Graph: Figure 10 ENF specimens' fractured surfaces after testing.

Graph: Figure 11 Typical FE mesh of the DCB coupon.

Graph: Figure 12 Element position calculation principle for the UMAT_RCurve routine.

Graph: Figure 13 Measurement of the COD and CSD from DCB and ENF specimens, respectively.

Graph: Figure 14 Mode I fracture toughness distribution at the DCB coupon's cohesive interface (UMAT_RCurve).

Graph: Figure 15 (a) Numerically computed COD—displacement curve for the DCB specimen; (b) Numerically computed CSD—crosshead displacement curve—for the ENF specimen.

Graph: Figure 16 Mode I bridging stress—COD curve.

Graph: Figure 17 Mode II bridging stress—CSD curve.

Graph: Figure 18 The final mixed-mode tri-linear T–SL.

Graph: Figure 19 UMAT_Tril subroutine flowchart.

Graph: Figure 20 Normalized traction-separation curve used in MAT_186 material model.

Graph: Figure 21 Experimental and numerical load–displacement curves of the SLS specimen.

Graph: Figure 22 SLS specimens' characteristic fractured surface investigated through optical microscopy.

Graph: Figure 23 (a) d1 damage variable and (b) d2 damage variable monitored through UMAT_Tril analysis; (c) CZM degradation region (0: elastic region; 1: matrix interface degradation; 2: fiber bridging degradation).

Author Contributions

Conceptualization, I.S. and K.T.; methodology, I.S. and K.T.; software, I.S.; validation, I.S. and K.T.; formal analysis, I.S.; investigation, I.S. and K.T.; resources, I.S.; data curation, I.S.; writing—original draft preparation, I.S.; writing—review and editing, I.S. and K.T.; visualization, I.S.; supervision, K.T.; project administration, K.T.; funding acquisition, K.T. All authors have read and agreed to the published version of the manuscript.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Acknowledgments

The authors would like to thank KVE composites for manufacturing the test specimens and RESCOLL for conducting part of the experimental work.

Footnotes 1 Publisher's Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations. References Pantelakis S., Tserpes K.. Revolutionizing Aircraft Materials and Processes; Springer International Publishing: Cham, Switzerland. 2020. 978-3-030-35345-2 2 Khan R. Fiber bridging in composite laminates: A literature review. Compos. Struct. 2019; 229: 111418. 10.1016/j.compstruct.2019.111418 3 Budiansky B., Amazigo J.C., Evans A.G. Small-scale crack bridging and the fracture toughness of particulate-reinforced ceramics. J. Mech. Phys. Solids. 1988; 36: 167-187. 10.1016/S0022-5096(98)90003-5 4 Hutchinson J.W., Jensen H.M. Models of fiber debonding and pullout in brittle composites with friction. Mech. Mater. Elsevier. 1990; 25: 139-163. 10.1016/0167-6636(90)90037-G 5 Cox B.N., Marshall D.B. The determination of crack bridging forces. Kluwer Acad. Publ. 1991; 18: 159-176. 10.1007/BF00035040 6 Bao G., Song Y. Crack bridging models for fiber composites with slip-dependent interfaces. J. Mech. Phys. Solids. 1993; 41: 1425-1444. 10.1016/0022-5096(93)90033-C 7 Sørensen B.F., Jacobsen T.K. Large-scale bridging in composites: R-curves and bridging laws. Compos. Part A Appl. Sci. Manuf. 1998; 29: 1443-1451. 10.1016/S1359-835X(98)00025-6 8 Lindhagen J.E., Berglund L.A. Application of bridging-law concepts to short-®bre composites Part 1: DCB test procedures for bridging law and fracture energy. Compos. Sci. Technol. 2000; 13: 871-883. 10.1016/S0266-3538(00)00004-X 9 Lindhagen J.E., Berglund L.A. Application of bridging-law concepts to short-®bre composites Part 2: Notch sensitivity. Compos. Sci. Technol. 2000; 9: 885-893. 10.1016/S0266-3538(99)00109-8 Lindhagen J.E., Gamstedt E.K., Berglund L.A. Application of bridging-law concepts to short-®bre composites Part 3: Bridging law derivation from experimental crack pro®les. Compos. Sci. Technol. 2000; 12: 2883-2894. 10.1016/S0266-3538(99)00177-3 Canal L.P., Alfano M., Botsis J. A multi-scale based cohesive zone model for the analysis of thickness scaling effect in fiber bridging. Compos. Sci. Technol. 2017; 139: 90-98. 10.1016/j.compscitech.2016.11.027 Pappas G., Botsis J. Intralaminar fracture of unidirectional carbon/epoxy composite: Experimental results and numerical analysis. Int. J. Solids Struct. 2016; 85–86: 114-124. 10.1016/j.ijsolstr.2016.02.007 Sohail A., Younas M., Bhatti Y., Li Z., Tunç S., Abid M. Analysis of Trabecular Bone Mechanics Using Machine Learning. Evol. Bioinform. Online. 2019; 15: 117693431882508. 10.1177/1176934318825084. 30936677 Bohner M., Tunç O., Tunç C. Qualitative analysis of caputo fractional integro-differential equations with constant delays. Comp. Appl. Math. 2021; 40: 214. 10.1007/s40314-021-01595-3 Russo A., Zarrelli M., Sellitto A., Riccio A. Fiber Bridging Induced Toughening Effects on the Delamination Behavior of Composite Stiffened Panels under Bending Loading: A Numerical/Experimental Study. Materials. 2019; 122407. 10.3390/ma12152407. 31357723 Riccio A., Russo A., Raimondo A., Sellitto A. A Numerical Assessment of Fibre Bridging Influence on Composite Panels Skin-stringer Debonding. Procedia Eng. 2016; 167: 56-63. 10.1016/j.proeng.2016.11.669 Riccio A., Russo A., Sellitto A., Toscano C., Alfano D., Zarrelli M. Experimental and Numerical Assessment of Fibre Bridging Toughening Effects on the Compressive Behaviour of Delaminated Composite Plates. Polymers. 2020; 12554. 10.3390/polym12030554 Fu C., Wang X. Simulating delamination in composite laminates involving large scale fiber bridging based on the mixed-mode three-linear cohesive zone model. Theor. Appl. Fract. Mech. 2022; 117: 103164. 10.1016/j.tafmec.2021.103164 Afshar A., Daneshyar A., Mohammadi S. XFEM analysis of fiber bridging in mixed-mode crack propagation in composites. Compos. Struct. 2015; 125: 314-327. 10.1016/j.compstruct.2015.02.002 Heidari-Rarani M., Shokrieh M.M., Camanho P.P. Finite element modeling of mode I delamination growth in laminated DCB specimens with R-curve effects. Compos. Part B Eng. 2013; 45: 897-903. 10.1016/j.compositesb.2012.09.051 Gong Y., Hou Y., Zhao L., Li W., Zhang J., Hu N. A modified mode I cohesive zone model for the delamination growth in DCB laminates with the effect of fiber bridging. Int. J. Mech. Sci. 2020; 176: 105514. 10.1016/j.ijmecsci.2020.105514 Zhao L., Gong Y., Zhang J., Chen Y., Fei B. Simulation of delamination growth in multidirectional laminates under mode I and mixed mode I/II loadings using cohesive elements. Compos. Struct. 2014; 116: 509-522. 10.1016/j.compstruct.2014.05.042 Gong Y., Chen X., Tao J., Zhao L., Zhang J., Hu N. A simple procedure for determining the mode I bridging stress of composite DCB laminates without measuring the crack opening displacement. Compos. Struct. 2020; 243: 112147. 10.1016/j.compstruct.2020.112147 D 5528-01. Standard Test Method for Mode I Interlaminar Fracture Toughness of Unidirectional Fiber-Reinforced Polymer Matrix Composites; ASTM International: West Conshohocken, PA, USA. 2009 Moutsompegka E., Tserpes K., Noeske M., Schlag M., Brune K. Experimental Investigation of the Effect of Pre-Bond Contamination with Fingerprints and Ageing on the Fracture Toughness of Composite Bonded Joints. Appl. Compos. Mater. 2019; 26: 1001-1019. 10.1007/s10443-019-09763-9 Floros I.S., Tserpes K.I., Löbel T. Mode-I, mode-II and mixed-mode I + II fracture behavior of composite bonded joints: Experimental characterization and numerical simulation. Compos. Part B Eng. 2015; 78: 459-468. 10.1016/j.compositesb.2015.04.006 Airbus Industrie. Test Method AITM 1.0006 Determination of Interlaminar Fracture Toughness Energy Mode II; Airbus Industrie: Blagnac Cedex, France. 1994 Moutsompegka E., Polydoropoulou P., Tserpes K.I., Tornow C., Schlag M., Brune K., Mayer B., Pantelakis S. The effects of pre-bond contamination with de-icing fluid on the mode-I and mode-II fracture toughness of composite bonded joints. Proceedings of the 6th EASN International. Porto, Portugal. 18–21 October 2016Available online: https://www.researchgate.net/publication/318672893(accessed on 26 June 2022) Leite Cavalcanti W., Brune K., Noeske M., Tserpes K., Ostachowicz W.M., Schlag M.. Adhesive Bonding of Aircraft Composite Structures: Non-Destructive Testing and Quality Assurance Concepts; Springer International Publishing: Cham, Switzerland. 2021. 978-3-319-92809-8 Anyfantis K.N., Tsouvalis N.G. Characterization of Fiber Bridging in Mode II Fracture Growth of Laminated Composite Materials. AMM. 2010; 24–25: 245-250. 10.4028/www.scientific.net/AMM.24-25.245 Hallquist J.O.. LS-DYNA: Theory Manual; Livermore Software Technology Corp: Livermore, CA, USA. 2006. 978-0-9778540-0-4 Kormpos P., Tserpes K., Floros G. Towards simulation of disassembly of bonded composite parts using the laser shock technique. IOP Conf. Ser. Mater. Sci. Eng. 2022; 1226: 012081. 10.1088/1757-899X/1226/1/012081 Tserpes K.I., Peikert G., Floros I.S. Crack stopping in composite adhesively bonded joints through corrugation. Theor. Appl. Fract. Mech. 2016; 83: 152-157. 10.1016/j.tafmec.2015.10.003 Floros I., Tserpes K. Numerical simulation of quasi-static and fatigue debonding growth in adhesively bonded composite joints containing bolts as crack stoppers. J. Adhes. 2021; 97: 611-633. 10.1080/00218464.2019.1690473 Tserpes K., Floros I. Fatigue crack growth simulation in adhesively bonded composite joints. Fatigue Fract. Eng. Mater. Struct. 2019; 42: 1430-1440. 10.1111/ffe.12969 Floros I., Tserpes K. Fatigue crack growth characterization in adhesive CFRP joints. Compos. Struct. 2019; 207: 531-536. 10.1016/j.compstruct.2018.09.020 Yi H., Jingjie C., Gang L., Hongbiao Y. A new method of crack-tip opening displacement determined based on maximum crack opening displacement. Eng. Fract. Mech. 2011; 78: 1441-1451. 10.1016/j.engfracmech.2011.01.008 Sourisseau Q., Lepretre E., Chataigner S., Chapeleau X., Mouton L., Paboeuf S. Use of high spatial resolution distributed optical fiber to monitor the crack propagation of an adhesively bonded joint during ENF and DCB tests. Int. J. Adhes. Adhes. 2022; 115: 103124. 10.1016/j.ijadhadh.2022.103124 Liu W., Chen P. Determination of the bridging law for mode I delamination via elastic restraint beam model and equivalent crack method. Eng. Fract. Mech. 2020; 226: 106867. 10.1016/j.engfracmech.2020.106867 Tserpes K., Barroso-Caro A., Carraro P.A., Beber V.C., Floros I., Gamon W., Kozłowski M., Santandrea F., Shahverdi M., Skejić D. A review on failure theories and simulation models for adhesive joints. J. Adhes. 2021; 97: 1-61. 10.1080/00218464.2021.1941903

By Ioannis Sioutis and Konstantinos Tserpes

Reported by Author; Author

Titel:
Development and Numerical Implementation of a Modified Mixed-Mode Traction-Separation Law for the Simulation of Interlaminar Fracture of Co-Consolidated Thermoplastic Laminates Considering the Effect of Fiber Bridging.
Autor/in / Beteiligte Person: Sioutis, I ; Tserpes, K
Link:
Zeitschrift: Materials (Basel, Switzerland), Jg. 15 (2022-07-22), Heft 15
Veröffentlichung: May 2010- : Basel, Switzerland : MDPI ; <i>Original Publication</i>: Basel, Switzerland : Molecular Diversity Preservation International, 2008-2010., 2022
Medientyp: academicJournal
ISSN: 1996-1944 (print)
DOI: 10.3390/ma15155108
Sonstiges:
  • Nachgewiesen in: MEDLINE
  • Sprachen: English
  • Publication Type: Journal Article
  • Language: English
  • [Materials (Basel)] 2022 Jul 22; Vol. 15 (15). <i>Date of Electronic Publication: </i>2022 Jul 22.
  • References: Evol Bioinform Online. 2019 Mar 24;15:1176934318825084. (PMID: 30936677) ; Materials (Basel). 2019 Jul 28;12(15):. (PMID: 31357723) ; Polymers (Basel). 2020 Mar 03;12(3):. (PMID: 32138254)
  • Grant Information: 101007494 European Commission
  • Contributed Indexing: Keywords: co-consolidated joints; cohesive zone modeling; finite element analysis; fracture toughness; thermoplastics
  • Entry Date(s): Date Created: 20220728 Latest Revision: 20220731
  • Update Code: 20240513
  • PubMed Central ID: PMC9329915

Klicken Sie ein Format an und speichern Sie dann die Daten oder geben Sie eine Empfänger-Adresse ein und lassen Sie sich per Email zusenden.

oder
oder

Wählen Sie das für Sie passende Zitationsformat und kopieren Sie es dann in die Zwischenablage, lassen es sich per Mail zusenden oder speichern es als PDF-Datei.

oder
oder

Bitte prüfen Sie, ob die Zitation formal korrekt ist, bevor Sie sie in einer Arbeit verwenden. Benutzen Sie gegebenenfalls den "Exportieren"-Dialog, wenn Sie ein Literaturverwaltungsprogramm verwenden und die Zitat-Angaben selbst formatieren wollen.

xs 0 - 576
sm 576 - 768
md 768 - 992
lg 992 - 1200
xl 1200 - 1366
xxl 1366 -