Sei sulla pagina 1di 18

Engineering Fracture Mechanics 72 (2005) 691708 www.elsevier.

com/locate/engfracmech

A cohesive model for fatigue failure of polymers


Spandan Maiti, Philippe H. Geubelle
*
Department of Aerospace Engineering and Beckman Institute, University of Illinois at Urbana-Champaign, 306 Talbot Laboratory, 104 S. Wright Street, Urbana, IL 61801, USA Received 19 April 2004; received in revised form 25 June 2004; accepted 27 June 2004 Available online 17 September 2004

Abstract A cohesive failure model is proposed to simulate fatigue crack propagation in polymeric materials. The model relies on the combination of a bi-linear cohesive failure law used for fracture simulations under monotonic loading and an evolution law relating the cohesive stiness, the rate of crack opening displacement and the number of cycles since the onset of failure. The fatigue component of the cohesive model involves two parameters that can be readily calibrated based on the classical loglog Paris failure curve between the crack advance per cycle and the range of applied stress intensity factor. The paper also summarizes a semi-implicit implementation of the cohesive model into a cohesive-volumetric nite element framework, allowing for the simulation of a wide range of fatigue fracture problems. 2004 Elsevier Ltd. All rights reserved.
Keywords: Cohesive model; Fatigue failure; Cohesive nite element; Polymeric materials

1. Introduction Polymeric materials subjected to cyclic loading are highly susceptible to fatigue failure, which often precedes monotonic or creep failures at similar applied load levels [19]. Substantial eort has thus been devoted to understand the mechanisms underlying the fatigue failure of this class of materials and to obtain predictive capabilities for their service life. Early attempts to characterize the fatigue failure in polymers were inspired by similar works for metals and were generally empirical in nature. Fatigue life test results were generally presented in the form of SN curves. But these curves are of limited use in practical situations as, unlike metals, the crack initiation life for these materials are often short compared to the fatigue crack propagation life [15]. Subsequently, dierent relationships were proposed to relate the per cycle crack

Corresponding author. Tel.: +1 217 244 7648; fax: +1 217 244 0720. E-mail address: geubelle@uiuc.edu (P.H. Geubelle).

0013-7944/$ - see front matter 2004 Elsevier Ltd. All rights reserved. doi:10.1016/j.engfracmech.2004.06.005

692

S. Maiti, P.H. Geubelle / Engineering Fracture Mechanics 72 (2005) 691708

advance with various loading and fracture parameters, such as the range of energy release rate, the range of the stress intensity factor (DK), or the mean stress intensity factor [21]. It has since been realized that the Paris power relation between the crack advance rate per cycle (da/dN) and DK, which is extensively used to represent the fatigue failure of metals, can also correlate the experimental data well for dierent polymers for various loading and material parameters. But the slope of the Paris loglog curve, i.e., the exponent m of the power relationship between da/dN and DK in da/dN = CDKm, is generally much higher than that for metals, typically exceeding a value of 6, indicating a much higher sensitivity to the applied loading. Extensive experiments performed on dierent polymers established further dierences between the mechanisms responsible for the crack growth in metals and glassy polymers. While it is mostly plasticity and distributed damage ahead of the crack tip in the case of metals, the primary mechanism for crack growth in polymers is the crazing of the material in a thin strip ahead of the crack tip. Though brittle to ductile transition is widely reported for this class of materials [22], most polymers tend to fail in a brittle manner under tension, and almost all of them fail in a macroscopically brittle way under cyclic loading provided the stress level is above some minimum value [18]. It has also been noted that, in the case of cyclic loading, in addition to continuous per cycle crack growth as in metals, an additional crack propagation mechanism is at work. In many polymers, the craze zone may grow over a number of cycles before the brils fail and the crack jumps through two thirds of the craze length, giving rise to discontinuous crack growth [20,8]. One of the most successful techniques used to model crack propagation in polymers relies on the cohesive modeling of fracture, which assumes the fracture process to take place in a vanishingly thin region ahead of the crack tip. This approach seems indeed particularly appropriate to capture the failure of polymeric systems where, as mentioned earlier, the appearance of a thin crazing zone is a key component [7]. Models based on cohesive technique have appeared in the literature for quasi-static [14] and dynamic crack propagation [23] in polymers. Though there is a wealth of experimental results on fatigue crack propagation in polymers, very few attempts have been undertaken to model fatigue crack propagation by cohesive technique in this class of materials. Cohesive modeling technique has been attempted for fatigue crack growth in metals [3,13,5,4], along interfaces [17] and in quasi-brittle materials [24]. In one of the early models proposed in [3], no distinction was made in the model between the loading and unloading paths, but a damage parameter was assumed. The evolution of this parameter with the number of load cycles was prescribed explicitly in the model. The presence of plasticity in the bulk material around the crack tip inuenced the crack closure and hence the failure of the material. But it was found that the crack ceases to grow after a few cycles due to plastic shakedown [13]. It has since been identied that a distinction needs to be made between the loading and unloading paths allowing for hysteresis so that subcritical crack growth becomes possible. Nguyen and co-workers has worked out a one-parameter cohesive model for metals [13], which is able to capture experimental Paris curves quite well and, in particular, the slope m of the curves (equal to approximately 3 for most metals). However, this one parameter model cannot yield steeper Paris curves, thus precluding the glassy polymers where the slope is at least 6. Another study deals with near threshold crack growth along a metal-rigid substrate interface and in a single crystal [5]. In this work, the unloading path is assumed to be parallel to the previous loading path, thus leaving certain amount of residual separation in the cohesive zone after each cycle. It is argued that the environment-assisted oxidation of the crack faces can give rise to this kind of behavior. But this type of crack closure eect is not very common in polymers. The fatigue model for interface cracks presented by Roe and Siegmund [17] is based on damage mechanics, where a history-dependent damage parameter gives rise to irreversibility. In this model also, the unloading path is not toward the origin. Values of the slope m of the Paris curves up to 3.1 have been reported in that study. Finally, fatigue crack growth in quasi-brittle materials has also been studied by cohesive techniques in [24], where the irreversibility of the loading and unloading paths is taken into account. A polynomial expression for the cyclic behavior is postulated in that study. A special case of Paris law, where the multiplicative constant is functionally dependent on the maximum loading, has been reported by these authors.

S. Maiti, P.H. Geubelle / Engineering Fracture Mechanics 72 (2005) 691708

693

In the present study, we formulate a new cohesive model specialized for polymers with higher fatigue crack growth sensitivity on the range of cyclic loading. Special emphasis is placed on the mode I fatigue failure of quasi-brittle glassy polymers such as PMMA, PC, epoxy etc. Our approach is inspired by the work presented in [13,24,17], where all the nonlinear eects associated with the failure process are accounted for in the cohesive zone model. We also describe the implementation of the cohesive model in a nite element framework (referred to as the Cohesive Volumetric Finite Element or CVFE scheme) allowing for the solution of a large set of structural problems. For simplicity, the examples presented in the present study involve its incorporation in a 1-D EulerBernoulli beam bending model of a double cantilever beam (DCB) specimen. The paper is organized as follows: To describe the cohesive model for fatigue failure, we start in Section 2 with a brief review of its foundation, i.e., a bi-linear cohesive model used to simulate quasi-static crack propagation under monotonic loading, and its implementation in the cohesive nite element scheme. We then discuss in Section 3 the formulation and implementation of the proposed cohesive model for fatigue crack propagation in polymeric materials. Finally, we present in Section 4 a detailed discussion of the results associated with the proposed model, including a comparison with some experimental results obtained on epoxy.

2. Cohesive model for monotonic loading 2.1. Cohesive failure law The cohesive model that serves as the foundation for the proposed cohesive modeling of fatigue crack propagation is characterized by a bi-linear, rate-independent, damage-dependent failure law between the cohesive traction vector T and the displacement jump vector D acting across the cohesive surfaces Cc. This particular cohesive model has been successfully used to simulate various fracture events in brittle media, namely, the impact-induced delamination of composite materials [6], quasi-static ber pushout in a model composite [10], the instability of dynamic crack path in ceramic materials [11], and the dynamic fragmentation of ceramics [12]. In the tensile (mode I) case, which is the focus of the work presented hereafter, this cohesive model takes the simple form Tn S Dn rmax ; 1 S Dnc Sinit 1

where Tn and Dn respectively denote the normal component of the cohesive traction and crack opening displacement vectors, rmax is the tensile cohesive failure strength, and Dnc the critical opening displacement jump (Fig. 1). The evolution of the damage process is quantied by the monotonically decreasing damage parameter S dened as S min Sp ; h1 Dn =Dnc i ; 2 where hai = a if a > 0 and =0 otherwise, and Sp denotes the previously achieved S value. As the material starts failing along its cohesive interface, the value of S gradually decreases from an initial value Sinit (chosen close to unity) to zero, point at which complete failure is achieved. The monotonicity of S is ensured by (2): the accumulated damage is preserved and no healing of the cohesive zone occurs due to unloading. As schematically indicated in Fig. 1, upon reloading, the cohesive stiness maintains its most recent value until further failure is achieved (i.e., when the cohesive failure envelope is once again reached). A contact algorithm based on the penalty method has been incorporated in the numerical model to prevent overlapping of the crack faces and thus simulate the crack closure eects. To that eect, the damage

694

S. Maiti, P.H. Geubelle / Engineering Fracture Mechanics 72 (2005) 691708

Tn

max

unloading loading

GIc

nc

Fig. 1. Cohesive tractionseparation law for tensile failure described by (1). After an unloading phase, reloading occurs along the unloading path rather than the initial loading path, thereby preserving the previously achieved damage level.

parameter S is kept to its initial value Sinit close to unity when the normal displacement jump Dn becomes negative. As illustrated in Fig. 1, this approach results in very high repulsive normal traction across the contacting fracture surfaces. 2.2. Implementation The cohesive model described by (1) and (2) can readily be implemented in a nite element framework, referred to as the cohesive volumetric nite element (CVFE) scheme, using the following form of the principle of virtual work Z Z Z S : dE dX T ex du dCex T n dDn dCc 0; 3 X Cex Cc |{z} |{z} |{z}
internal external cohesive

where u is the displacement vector, S and E denote the internal stress and strain tensors, respectively, Tex is the externally applied traction, and X, Cc and Cex respectively denote the volume, cohesive boundary and exterior boundary of the deformable body. The last term in (3) corresponds to the virtual work done by cohesive traction Tn for a virtual separation dDn and is associated with the contribution of the cohesive elements. The expression of the internal component of the virtual work (rst term in (3)) depends on the type of volumetric element used in the analysis, which, in turn, aects the expression of the stress S and virtual strain dE. As mentioned in the introductory section, we use in the present study classical 2-node Euler Bernoulli beam elements as volumetric elements since the primary focus of this work is the simulation of fatigue delamination failure in a double cantilever beam (DCB) specimen. The corresponding nite element formulation can be found in Appendix A. To ensure the spatial convergence of the CVFE scheme, it is critical to capture the fracture process accurately, i.e., to have a sucient number of cohesive elements in the active cohesive zone. A rough estimate of the cohesive zone size Lcoh is given by [16] Lcoh p E GIc ; 8 1 m2 r 2 ave

S. Maiti, P.H. Geubelle / Engineering Fracture Mechanics 72 (2005) 691708

695

where E is the Youngs modulus, GIc the mode I fracture toughness (=rmaxDnc/2 for the bi-linear cohesive model described earlier), and rave the average stress in the cohesive zone (taken as rmax/2 for the bi-linear cohesive law). Our experience with cohesive element modeling suggests that at least 4 cohesive elements must be present in the active cohesive zone. 2.3. Verication and validation To verify the CVFE scheme for crack propagation in a DCB specimen subjected to monotonic loading, we compare our numerical results to the theoretical predictions obtained in the limiting case of a vanishingly small cohesive zone size. For reference, the analytical expression of the evolution of the crack length a and load P for a displacement-controlled DCB specimen (of initial uncracked length b0 and initial crack length a0) can be found in Appendix B. In the example shown hereafter, b0 = 100 mm, the Youngs modulus E = 3.4 GPa, the fracture toughness GIc = 88.97 J/m2, and the cohesive strength rmax = 35 MPa. In Fig. 2, the reaction force developed in the beam at the loading point is plotted against the applied displacement D for three values of the initial crack length a0, namely 50, 100 and 200 mm. Excellent agreement is achieved between numerical and theoretical curves, especially for the larger value of a0 for which the assumptions underlying the theoretical solution are increasingly applicable. This agreement between theoretical and numerical predictions is also observed in the evolution of the crack length presented in Fig. 3. As predicted by the theory, the crack remains stationary until sucient energy is stored in the deformed beam elements to allow for crack propagation. Once again, the longer the initial crack, the better this agreement. Despite its simplicity, the CVFE model of DCB delamination can be used very eciently to model the progressive failure of actual DCB specimens. To validate the CVFE scheme, we choose to model the mode I delamination experiments performed by Borg and co-workers [1] on HTA/6376C composite DCB specimens. The corresponding constitutive and failure properties are E = 146 GPa and GIc = 259 J/m2 as reported in their paper. For the simulation of this problem, the number of cohesive elements ahead of the pre-crack is chosen as 220. The cohesive strength rmax is taken to be 35 MPa, and, as shown in Fig. 4, the cohesive nite element model is able to capture quite well the evolution of both the reaction force and the crack length (a0 = 35 mm).

1000 900 800 700 600 P (N) 500 400 300 200 100 0 0 0.005 0.01 0.015 0.02 0.025 (mm) 0.03 0.035 0.04
a = 100 mm
0

a = 50 mm
0

theoretical

a0 = 200 mm

Fig. 2. Comparison between theoretical (dashed curve) and numerical (solid curves) evolution of the reaction force P at the tip of a DCB specimen under prescribed displacement D, for 3 values of the initial crack length a0.

696

S. Maiti, P.H. Geubelle / Engineering Fracture Mechanics 72 (2005) 691708


250
a0 = 200 mm

200

a (mm)

150
a0 = 100 mm

theoretical

100

a = 50 mm

50

10

15 20 (mm)

25

30

35

Fig. 3. Crack length a versus prescribed displacement D: comparison between numerical (solid curves) and theoretical (dashed curve) results obtained for 3 values of the initial crack length a0.

60

60

50 40
P (N)
computed reaction experimental reaction

50 Crack Extension (mm)

40

30
experimental crack extension

30

20
computed crack extension

20

10

10

3 (mm)

Fig. 4. Delamination of a composite DCB specimen: experimental measurements [1] and numerical predictions of the reaction and crack extension versus D curves.

3. Cohesive model for fatigue 3.1. Model description After describing, verifying and validating the cohesive model for mode I crack propagation under monotonic loading, we now turn our attention to the main topic of the present paper: the formulation and implementation of a cohesive model for fatigue crack propagation in polymers. As mentioned earlier, although it prevents healing of the fracture surfaces through the enforcement of monotonic decay of the damage parameter S, the cohesive model discussed in the previous section leads to similar unloading and reloading paths in the tractionseparation curve. This characteristic prevents crack growth under subcritical cyclic loading due to the progressive degradation of the cohesive properties in the cohesive failure zone. This lim-

S. Maiti, P.H. Geubelle / Engineering Fracture Mechanics 72 (2005) 691708

697

Tn max

nc

Fig. 5. Schematic evolution of the cohesive stiness during cyclic loading, assuming unloading towards the origin and degradation of the cohesive stiness during reloading.

itation suggests the need for an evolution law to describe the changes incurred by the cohesive strength under fatigue [24,17,13,5,4]. As mentioned earlier, complex physical phenomena take place in the craze (or cohesive) zone ahead of the advancing crack with the appearance of irreversible processes associated with the creation, extension and failure of brils. A phenomenological model of such processes involves the progressive degradation of the cohesive zone strength during reloading events as illustrated schematically in Fig. 5. The evolution law of the instantaneous cohesive stiness kc, i.e., the ratio of the cohesive traction Tn to the displacement jump Dn, during reloading can be expressed in the general form kc dT n FN f ; T n ; dDn 4

where Nf denotes the number of loading cycle experienced by the material point since the onset of failure, i.e., at the time the cohesive traction Tn rst exceeded the failure strength rmax. Simplifying (4), we adopt the separable form dT n cN f T n ; dDn 5

leading to an exponential decay of the cohesive strength, with the rate of decay controlled by the parameter c. In the present study, we use the following two-parameter power-law relation 1 c N b ; 6 a f where a and b are material parameters describing the degradation of the cohesive failure properties. The physical signicance of these two parameters is discussed in Section 4 in terms of their eect on the resulting Paris fatigue curve. Note that the parameter a has the dimension of length and b denotes the history dependence of the failure process. When it vanishes, we retrieve the one-parameter model of Nguyen et al. [13].

698

S. Maiti, P.H. Geubelle / Engineering Fracture Mechanics 72 (2005) 691708

The proposed evolution law for the cohesive model can also be expressed in terms of the rate of change _ of the cohesive stiness k c as follows: 1 _ _ _ k c N b k c Dn if Dn P 0; a f _ 0 if Dn 6 0; 7

_ where Dn is the rate of change of the normal separation. The second equation simply states that the cohesive stiness is assumed to remain constant during unloading. 3.2. Implementation issues During the reloading phase, (7) can be described in its discretized form as 1 b k j1 k j N j k j Dj1 Dj ; c c c n n a f where the superscripts j and j + 1 stand for loading steps j and j + 1, respectively. Recalling the expression for kc obtained from (1), the updated value of the cohesive stiness k j1 can be written as c ! j S 1 rmax 1 1 N j b Dj1 Dj ; Dj1 P Dj : 8 k j1 n n n n c j a f 1 S Dnc Sinit A quasi-implicit load stepping scheme is used in this investigation to form the equations of equilibrium. In this simple DCB implementation, the stiness matrix arising from the volumetric (beam) elements is constant throughout the simulation. The only nonlinear component of this problem is associated with the cohesive component of the stiness matrix, which is updated at each load step based on the displacement jump values achieved at the previous loading step. As shown in relation (8), during the reloading phase, the cohesive stiness at each material point along the cohesive zone gradually reduces proportionally to the increment in crack opening displacement. This proportionality factor c evolves with the cycles to failure Nf in accordance with (6) and thus gives a measure of the total accumulated damage in the failure process. Note that the second term in the square-bracketed quantity in (8) can be viewed as a damage parameter which depends on the increment of deformation and starts accumulating only after the parameter S goes below its initial value Sinit . Therefore, the material can cycle innitely without failure when the cohesive traction is below its critical value rmax. The basic tenets of damage evolution laws are thus also satised [9]. To further elaborate on the dierence between subcritical and critical failure, we consider in Fig. 6 the evolution of the failure process taking place in the cohesive zone ahead of a delamination crack propagating in a DCB specimen subjected to a prescribed sinusoidal displacement loading. The DCB specimen is 150 mm long and has an initial crack length a0 equal to 50.1 mm. The cyclic displacement D applied at the end of the beam has a maximum amplitude of 1.5 mm and the ratio of minimum to maximum amplitude of loading R = 0. The specimen is made of epoxy with a Youngs modulus E = 3.9 GPa, and a mode fracp I ture toughness GIc = 88.97 J/m2, or, equivalently, a critical stress intensity factor KIc = 0.55 MPa m. The cohesive strength rmax is taken as 35 MPa and the values of the two fatigue parameters a and b are chosen as 4 lm and 0.3 respectively. Since the estimated cohesive zone size is approximately 0.45 mm, the cohesive element size is chosen as 0.1 mm. Fig. 6 represents the evolution of the failure process at a point located ahead of the pre-existing crack, more precisely, at the location of the rst Gauss point of the tenth cohesive element. During the initial phase of the failure process, the stress concentration present at the crack tip leads to a critical fatigue failure, as the material degradation curve follows the envelope of the bi-linear failure law. The failure process then becomes subcritical as the degradation leads to peak cohesive traction values well below the critical value (denoted by the dashed curve). In the event of an overload, the fatigue

S. Maiti, P.H. Geubelle / Engineering Fracture Mechanics 72 (2005) 691708

699

1 0.9 0.8 0.7


1 15 20 25 x 10 2
5

(m)

1 0

Tn / max

0.6 0.5 0.4 0.3 0.2 0.1 0 0 0.2 0.4 0.6 n / nc

Cycles

0.8

Fig. 6. Subcritical versus critical failure: evolution of the tractionseparation curve for a point located in the path of a fatigue crack. This plot illustrates the initial critical failure followed by subcritical degradation of the cohesive properties, with a transient critical cycle associated with an overload (inset) in applied displacement.

failure may become critical again and follow the failure envelope until the next unloading phase, as illustrated in Fig. 6. In all cases, the unloading portion of the cohesive failure curve points to the origin.

4. Results and discussion 4.1. Characteristics of fatigue crack propagation In this section, we comment on some basic characteristics of the fatigue crack propagation results obtained from the numerical model developed in Section 3. As a basis for this discussion, we consider the fatigue failure of the DCB specimen whose geometry and constitutive and failure properties have been described at the end of Section 3.2. The cyclic displacement applied at the tip of the beam has the same amplitude (1.5 mm) but its minimum value is chosen as 0.15 mm, resulting in an amplitude ratio R = 0.1. In this initial simulation, the parameters a and b are taken as 4 lm and 0.3 respectively. Fig. 7 presents the evolution of the crack tip and cohesive zone tip locations as a function of the number of loading cycles. As expected, in this prescribed displacement problem, the crack driving force decreases with crack advance, and the crack extension rate decreases with the number of loading cycles giving rise to a stable crack growth. The cohesive zone tip follows a similar evolution, but, as the crack extension rate decreases, the cohesive zone length increases signicantly, from an initial size of 1.1 mm to a nal size of 2.4 mm. Using the analytical solution of the DCB fracture problem summarized in Appendix B, the crack length can be related to the applied stress intensity factor K, in terms of its maximum applied value (Kmax) or its range (DK). As mentioned earlier, the proposed model leads to subcritical material degradation characterized with energy levels well inferior to the (monotonic) fracture toughness GIc. This fact is illustrated in Fig. 8, which presents the variation of the expended energy G (i.e., the area under the fatigue cohesive failure curve) with the applied Kmax value (normalized by KIc). As apparent in that gure, the fracture toughness

700

S. Maiti, P.H. Geubelle / Engineering Fracture Mechanics 72 (2005) 691708


80 Crack and cohesive zone tip location (mm)

75

70

65 crack tip cohesive zone tip 55

60

50 0

1000

2000

3000 Cycles

4000

5000

Fig. 7. Evolution of crack and cohesive zone tip locations with number of loading cycles, a = 4 lm and b = 0.3.

0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0.4 0.5 0.6 0.7 Kmax /KIc 0.8 0.9 1

7000 6000 5000 4000 3000 2000 1000 0 Cycles to failure

Fig. 8. Variation of the fatigue fracture toughness G and the number of cycles to failure with the applied Kmax, for a = 4 lm and b = 0.3. G and Kmax both are normalized by their monotonic counterparts, GIc and KIc, respectively.

under fatigue loading can represent a very small fraction of its monotonic counterpart, with values inferior to 5% of GIc for Kmax < 0.4KIc. For higher values of the applied cyclic load, the required energy for failure increases rapidly, especially as the applied Kmax approaches 0.9KIc. The gure also illustrates the number of cycles to failure as a function of Kmax/KIc for the same fatigue crack propagation problem, showing a rapidly decaying trend down to zero for Kmax = KIc. 4.2. Eect of the parameters on fatigue crack propagation We now turn our attention on the eect of the parameters a and b on the predicted fatigue failure. Fig. 9 presents the evolution of the total crack extension Da with the number of loading cycles for three values of a, with the parameter b kept constant at 0.3. As apparent there, the crack growth rate decreases with increasing a: for a = 4 lm, the crack extends by 26 mm after 4000 loading cycles, whereas it has extended by only 20 mm and 17 mm for a = 6 lm and 8 lm, respectively. Dierentiating these curves, we can characterize the eect of a on the fatigue response in terms of the evolution of the crack advance per cycle (da/dN)

G/GIC

S. Maiti, P.H. Geubelle / Engineering Fracture Mechanics 72 (2005) 691708


80

701

75 Crack extension (mm)

= 4 m
f

70

f = 6 m = 8 m
f

65

60

55

50

1000

2000

3000

4000 Cycles

5000

6000

7000

8000

Fig. 9. Evolution of the crack extension with the number of cycles for three values of a, with b = 0.3.

versus the applied range of stress intensity factor, DK (normalized with KIc), better known as the Paris fatigue failure law (Fig. 10). As apparent in that gure, all the curves have the same slope but dierent intercepts, suggesting that a unit change in DK produces an equal amount of change in the crack growth rate, though the actual value of da/dN can be dierent.

= 4 m

10

= 6 m

da/dN (mm/cycle)

= 8 m

10

10

0.4

10

0.3

K/KIC

10

0.2

10

0.1

Fig. 10. Eect of the parameter a on the fatigue curve, showing that a aects the intercept but not the slope of the curves (b = 0.3).

702

S. Maiti, P.H. Geubelle / Engineering Fracture Mechanics 72 (2005) 691708


85 80 Crack extension (mm) 75 70
= 0.3

= 0.0

65 60 55 50

= 0.5

2000

4000

6000

8000 10000 12000 14000 16000 Cycles

Fig. 11. Eect of b on the evolution of the crack extension (a = 8 lm).

The eect of the parameter b on the fatigue crack propagation is more pronounced, as illustrated by Fig. 11, which presents the crack extension versus loading cycles for b = 0.0, 0.3 and 0.5, with a kept constant at 8 lm. As apparent in that gure, b aects the crack growth rate, leading to Paris failure curves showing very dierent slopes (Fig. 12). This parameter plays therefore a key role in capturing the fatigue response of polymeric materials, where the rate of crack growth per cycles may vary substantially with the crack driving force from material to material. The eect of the parameter b on the slope m of the Paris fatigue failure curve is summarized in Fig. 13: for lower values (b < 0.4), the predicted value for the exponent m is about 3, which is characteristic of the fatigue failure of metals. As b increases, increasing values of the exponent m of the Paris fatigue law are obtained. 4.3. Fatigue failure of epoxycomparison with experiments To conclude this result discussion, we simulate in this last section the fatigue crack growth in epoxy and compare it with experimental results [2]. As many polymeric materials, epoxy has a very high sensitivity of the crack growth rate with the change in stress intensity factor, hence a large slope of the Paris curve. The slope m of the Paris fatigue curve for epoxy is about 10 compared to about 3 for most metals. So the change in fatigue crack growth rate is strongly correlated with the crack driving force. The experimental values of the crack growth rate da/dN were obtained by Brown [2] with a tapered cantilever beam specimen made of neat EPON 828 epoxy resin with 12 pph Ancamine DETA (diethylenetriamine) curing agent subjected to a range of cyclic loading conditions. Owing to the tapered conguration, the stress intensity factor remained constant during the experiments. Therefore, every test led to a single data point on the da/dN versus DK plot. In our numerical simulations, we use a cantilever beam with constant cross section so that the stress intensity factor varies continuously with the propagating crack, allowing us to obtain in one simulation the entire fatigue failure curve. The maximum amplitude of the applied displacement papplied is taken to be D 1.8 mm resulting in an initial value of the maximum stress intensity factor 0.6 MPa m, so that the crack quickly establishes itself. Once the crack starts growing, the stress intensity factor goes down and fatigue crack propagation is established. The load ratio R is taken to be 0.1. Fig. 14 presents the comparison between the experimental points and numerical simulations. The oscillations apparent in the da/dN versus DK curve for low values of DK are associated with the diculty of dierentiating the crack evolution curve for very small crack advances. The two parameters of the cohesive fatigue model are taken to be a = 8 lm and b = 0.65. The value of b has been chosen based on the results

S. Maiti, P.H. Geubelle / Engineering Fracture Mechanics 72 (2005) 691708


1

703

10

= 0.0

10 da/dN (mm/cycle)

= 0.3

10

= 0.5

10

10

0.4

10

0.3

10

0.2

K/KIC

Fig. 12. Eect of b on the Paris fatigue curve, with a = 8 lm.

22 20 18 16 Slope m 14 12 10 8 6 4 2 0 0.1 0.2 0.3 0.4 Parameter 0.5 0.6 0.7

Fig. 13. Variation of slope of the Paris fatigue curve m with the parameter b.

presented in Fig. 13. The proposed two-parameter model captures the entire fatigue failure response curve quite well, including in the higher crack growth rate regime, where the numerical results show the rapid acceleration of the fatigue crack for DK/KIc > 0.8.

704

S. Maiti, P.H. Geubelle / Engineering Fracture Mechanics 72 (2005) 691708


10
3

da/dN (mm/cycle)

10

10

0.4

K/KIC

1.0

Fig. 14. Comparison between experimental (symbols) and computed (solid curve) fatigue response curves for epoxy. The experimental results are taken from [2].

For low values of the applied displacement, the proposed model naturally introduces a threshold value of DK below which no fatigue crack propagation is possible. For the simple DCB problem described earlier, this threshold value can be obtained analytically by solving the linear problem of a beam on an elastic foundation over a portion (L a0) of its length L, with a0 denoting the initial crack length. In the absence of any failure, the stiness of the elastic foundation is constant and is equal to kf 1 rmax : 1 Sinit Dnc 9

The displacement d at the initial crack tip is related to the half applied displacement D at the tip of the beam by d Ka0 ; k f ; L; E; ID; where the coecient of proportionality K is given in Appendix C. The onset of fatigue failure will be reached when d 1 Sinit Dnc =2; 11 which, combined with (9) and (B.2), leads to the following expression of the threshold value of the applied DK at which the rst cohesive element enters the softening portion of the cohesive model r 3E1 R I 1 Sinit Dnc =2K: 12 DK th a2 b 10

S. Maiti, P.H. Geubelle / Engineering Fracture Mechanics 72 (2005) 691708


10
0

705

10

da/dN (mm/cycle)

10

10

10

K/KIC

10

Fig. 15. Paris plot showing the upper and lower thresholds of the fatigue crack propagation with a = 4 lm and b = 0.3. The right vertical arrow denotes DK normalized with KIc corresponding to Kmax = KIc and the left vertical arrow is the theoretical lower threshold.

For the geometry and material parameters used in the DCB problem described in Section 3, with R = 0.1 p and Sinit 0:98, the computed value of DK th =K Ic 0:09 MPa m. Due to the high fatigue sensitivity of epoxy (m % 10), this threshold value corresponds to an extremely small value of the crack extension per cycle (da/dN < 106 mm/cycle), making its numerical capture excessively expensive. However, if we consider a material system with a = 4 lm and b = 0.3 (i.e., m % 3), the cohesive modeling of the fatigue crack propagation yields the results shown in Fig. 15, which clearly illustrates the near-threshold behavior. As DK/KIc decreases and approaches the threshold value (DKth = 0.09KIc) indicated by a vertical arrow, the crack advance slows down to a halt. The dierence between theoretical and numerical values of DKth is due to numerical error associated with the nite element approximation of the beam on (nonlinear) elastic foundation.

5. Conclusion We have presented a cohesive failure model specially developed for the simulation of fatigue crack propagation in polymeric structures subjected to mode I cyclic loading. The phenomenological model is based on a two-parameter evolution relation between the cohesive stiness, the crack opening displacement rate and the number of cycles since the onset of failure. The fatigue evolution law has been combined with a bi-linear cohesive model for crack propagation under monotonic loading. The two parameters (a and b in (7)) entering the evolution law dene the irreversible degradation process during each reloading cycle

706

S. Maiti, P.H. Geubelle / Engineering Fracture Mechanics 72 (2005) 691708

and can be readily calibrated based on the slope and intercept of the Paris fatigue failure curve characterizing the material. The semi-implicit implementation of the cohesive model into a cohesive-volumetric nite element (CVFE) framework has also been described. Although the simulations presented in the present paper have used a simple 1-D CVFE scheme combining cohesive elements with conventional EulerBernouilli beam elements, the formulation presented here applied to other volumetric nite elements and therefore allows for the simulation of a wide range of fatigue failure problems. The cohesive model has been shown to be quite successful in capturing the fatigue response of an epoxy, not only within the range of loading amplitudes where the Paris fatigue law is applicable, but also with regards to the low and high amplitude regimes.

Acknowledgments The authors gratefully acknowledge funding for this project provided by AFOSR Aerospace and Materials Science Directorate (Grant # F49620-02-1-0080). They also wish to thank Dr. Eric Brown and Prof. Nancy Sottos for helpful discussions.

Appendix A. Cohesive element formulation The expression of the local stiness matrix kcoh of the cohesive element for this DCB problem is derived from the cohesive component of the principle of virtual work described by (3). Let d denote the vector containing the nodal deection (w1 and w2) and rotation (h1 and h2) degrees of freedom associated with the 2-node EulerBernouilli (C1 continuous) element. Only the upper half of the DCB specimen has been considered for computations due to symmetry. The normal displacement jump or crack opening displacement along that cohesive element (of length l) is approximated by Dn s 2d N; A:1

where the vector N contains the four cubic shape functions of the beam element. The introduction of the cohesive failure law described by (1) leads to kcoh 2 Z
0 l

N T k c N dCc ;

A:2

where kc denotes the cohesive stiness and, for the bi-linear cohesive relation (1), is given by kc S 1 rmax : 1 S Dnc Sinit A:3

A 3-point Gauss quadrature scheme is used to evaluate the integral appearing in (A.2).

Appendix B. LEFM solution for a DCB We summarize here the classical linearly elastic fracture mechanics (LEFM) solution for the delamination of a double cantilever beam (DCB) specimen with an initial crack length a0 and subjected to an end displacement 2D. This solution is used in Section 2 to verify the CVFE scheme in the limit of a very small

S. Maiti, P.H. Geubelle / Engineering Fracture Mechanics 72 (2005) 691708

707

cohesive zone length (Lc ( a0), and in Sections 3 and 4 to extract the range (DK) and maximum value (Kmax) of the applied stress intensity factor K during fatigue crack propagation. Based on the expression of the compliance C of the DCB specimen, C 2D=P 2a3 ; 3EI B:1

the energy release rate G associated with a crack of length a is given by G K 2 P 2 dC 9EI 2 I 4D: E 2b da ba B:2

In (B.1) and (B.2), E, I and b respectively denote the Youngs modulus, moment of inertia and width of the two beams. Under steady-state crack propagation, G = Gc, which leads to the classical relations between the crack length a, the reaction force P and the applied displacement D:  a 9EI bGc 0:25 D0:5 ; P EI0:25 bGc 0:75 3D
0:5

B:3

Appendix C. Threshold limit for the fatigue crack propagation For the problem of an elastic beam on a partial elastic foundation shown in Fig. 16, Relation (10) in Section 4 can be restated as d KD A D; B C:1 C:2

  a a a  a a  a A 3e4b1 1 b 2e2b1 sin2b1  b 2bcos2 b1  1 b; a a

  a a a 2 a 3 a  a a  B e4b1 3 6b 6b 2 2b 3 e2b1 6 sin2b1  12b1 2bcos2 b1  a a

 a  a  a  a a a a 12b 2 sin2b1  4b 3 1 2cos2 b1  3 6b 6b 2 2b 3 ; with b k f =4EI0:25 ;  b bL and  a0 =L: a

C:3

kf

L a0

a0

Fig. 16. Schematic of a cantilever beam of length L partially on an elastic foundation of stiness kf subjected to a displacement D at the tip.

708

S. Maiti, P.H. Geubelle / Engineering Fracture Mechanics 72 (2005) 691708

References
[1] Borg R, Nilsson L, Simonsson K. Modeling of delamination using a discretized cohesive zone and damage formulation. Comp Sci Tech 2002;62:1299314. [2] Brown EN. Fracture and fatigue of a self-healing polymer composite material. PhD thesis, University of Illinois at UrbanaChampaign, 2003. [3] de-Andres A, Perez JL, Ortiz M. Elastoplastic nite element analysis of three-dimensional fatigue crack growth in aluminium shafts subjected to axial loading. Int J Solids Struct 1999;36:223158. [4] Deshpande VS, Needleman A, Van der Giessen E. A discrete dislocation analysis of near-threshold fatigue crack growth. Acta Mater 2001;49(16):3189203. [5] Deshpande VS, Needleman A, Van der Giessen E. Discrete dislocation modeling of fatigue crack propagation. Acta Mater 2002;50(4):83146. [6] Geubelle PH, Baylor J. Impact-induced delamination of composites: a 2-D simulation. Composites B 1998;29:589602. [7] Knauss WG. Time dependent fracture and cohesive zones. J Engng Mater Tech Trans ASME 1993;115:2627. [8] Konczol L, Doll W, Bevan L. Mechanisms and micromechanics of fatigue crack propagation in glassy thermoplastics. Coll Polym Sci 1990;268:81422. [9] Lemaitre J. A course on damage mechanics. 2nd ed. Springer; 1996. [10] Lin G, Geubelle PH, Sottos NR. Simulation of ber debonding with friction in a model composite pushout test. Int J Solids Struct 2001;38(4647):854762. [11] Maiti S, Geubelle PH. Mesoscale modeling of dynamic fracture of ceramic materials. Comp Meth Engng Sci 2004;5(2):91102. [12] Maiti S, Geubelle PH, Krishnan R. Fragmentation of ceramics in rapid expansion mode, vol. 1415. Fract Mech Ceram (The 8th International Symposium on Fracture Mechanics of Ceramics, Houston, USA), Plenum Press, in press. [13] Nguyen O, Repetto EA, Ortiz M, Radovitzky RA. A cohesive model of fatigue crack growth. Int J Fract 2001;110:35169. [14] Pandya KC, Ivankovic A, Williams JG. Cohesive zone modelling of crack growth in polymers: Part 2Numerical simulation of crack growth. Plast Rubber Compos 2000;29(9):44752. [15] Radon JC. Fatigue crack growth in polymers. Int J Fract 1980;16(6):53352. [16] Rice JR. Mathematical analysis in the mechanics of fracture. In: Liebowitz H, editor. Fracture, vol. 2. New York: Academic Press; 1968. p. 191311. [17] Roe KL, Siegmund T. An irreversible cohesive zone model for interface fatigue crack growth simulation. Engng Fract Mech 2003;70:20932. [18] Sauer JA, Hara M. Eect of molecular variables on crazing and fatigue of polymers. In: Kausch HH, editor. Advances in Polymer Science, vol. 9192. Berlin: Springer-Verlag; 1990. p. 69118. [19] Sauer JA, Richardson GC. Fatigue of polymers. Int J Fract 1980;16(6):499532. [20] Skibo MD, Hertzberg RW, Manson JA, Kim SL. On the generality of discontinuous fatigue crack growth in glassy polymers. J Mater Sci 1977;12:53142. [21] Sutton SA. Fatigue crack propagation in an epoxy polymer. Engng Fract Mech 1974;6:58795. [22] Van der Giessen E, Estevez R, Pijnenburg KGW, Tijssens MGA. Computational modeling of failure process in polymers. In: European Conference on Computational Mechanics, Munchen, Germany, 1999. [23] Xu X-P, Needleman A. Numerical simulation of fast crack growth in brittle solids. J Mech Phys Solids 1994;42:1397434. [24] Yang B, Mall S, Ravi-Chandar K. A cohesive zone model for fatigue crack growth in quasibrittle materials. Int J Solids Struct 2001;38:392744.

Potrebbero piacerti anche