On crack path evolution rules

On crack path evolution rules

Engineering Fracture Mechanics 77 (2010) 1781–1807 Contents lists available at ScienceDirect Engineering Fracture Mechanics journal homepage: www.el...

3MB Sizes 2 Downloads 53 Views

Engineering Fracture Mechanics 77 (2010) 1781–1807

Contents lists available at ScienceDirect

Engineering Fracture Mechanics journal homepage: www.elsevier.com/locate/engfracmech

On crack path evolution rules K.P. Mróz *, Z. Mróz ´ skiego 5b, 02-106 Warsaw, Poland Institute of Fundamental Technological Research, Pawin

a r t i c l e

i n f o

a b s t r a c t

Article history: Received 15 December 2009 Received in revised form 9 March 2010 Accepted 23 March 2010 Available online 27 March 2010

The models of crack growth in mixed mode conditions are reviewed for the plane and three-dimensional (3D) states of stress. Both critical load value and crack path or surface growth are predicted by different criteria in terms of elastic singular stress states and Tstress component. Monotonic and cyclic loading induced crack growth is considered. The energy and critical plane criteria based on local or non-local measure of stress and strain are most useful in developing predictive crack growth simulation. The finite critical distance from the crack tip should be specified to provide averaged or local stress and strain states. The application of MK-criterion of crack growth expressed in terms of volumetric and deviatoric strain energies is presented for several specific cases of monotonic and cyclic loading. The concepts of smooth and rough crack surface are discussed with application to 3D crack surface growth. Ó 2010 Elsevier Ltd. All rights reserved.

Keywords: Fracture criteria T-stress Fatigue crack growth 3D crack Smooth and rough cracks Crack path Critical plane Non-local models

1. Introduction The prediction of crack growth path under imposed loading and of its rate of growth constitutes the most important problem in fracture mechanics. Most papers devoted to this problem are based on the assumption of linear elastic fracture mechanics (LEFM), but recent studies also include the effects of plastic deformation and also damage accumulation at the crack tip. So the term elasto-plastic fracture mechanics (EPFM) can be applied for these studies. The present paper will apply LEFM to predict crack growth but the effects of plasticity and damage will be accounted for by postulating the relative fracture criteria. The crack growth can be simulated as a succession of straight segments in plane cases or linear surface elements in 3D cases. The effects of crack curvature can also be incorporated in the analysis. The stress and strain states near the crack tip (Fig. 1) can be expressed within the assumption of LEFM in terms of the stress intensity factors (SIF) KI, KII, KIII and the asymptotic singularity solution. For the plane strain or stress cases the stress state is expressed in the Cartesian reference system as follows [1]

K

h



h

 3h

K

h



h

3h



I II ffi cos rxx ffi pffiffiffiffiffiffiffiffi 1  sin sin  pffiffiffiffiffiffiffiffiffi sin 2 þ cos cos þ T; 2 2 2 2 2 2 2pr 2pr   KI h h 3h K II h h 3h ffi cos 1 þ sin sin þ pffiffiffiffiffiffiffiffiffi cos sin cos ; ryy ffi pffiffiffiffiffiffiffiffi 2 2 2 2 2 2 2pr 2pr  

K

h

h

3h

K

h

h

3h

I II ffi cos sin cos þ pffiffiffiffiffiffiffiffiffi cos ; rxy ffi pffiffiffiffiffiffiffiffi 1  sin sin 2 2 2 2 2 2 2pr 2pr

* Corresponding author. Tel.: +48 22 826 12 81. E-mail addresses: [email protected] (K.P. Mróz), [email protected] (Z. Mróz). 0013-7944/$ - see front matter Ó 2010 Elsevier Ltd. All rights reserved. doi:10.1016/j.engfracmech.2010.03.038

ð1Þ

1782

K.P. Mróz, Z. Mróz / Engineering Fracture Mechanics 77 (2010) 1781–1807

σ∞

r

θ

α

kσ ∞

kσ ∞

σ∞ Fig. 1. The geometry of the main crack.

and in the polar coordinate system:

K

 5

1

h

3h





K

5

 3h

3

h

I II ffi cos  cos þ pffiffiffiffiffiffiffiffiffi  sin þ sin þ T cos2 h; rrr ffi pffiffiffiffiffiffiffiffi 2 4 2 4 2 4 2 2pr 4 2p r     KI 3 h 1 3h K II 3 h 3 3h 2 ffi cos þ cos þ pffiffiffiffiffiffiffiffiffi  sin  sin þ T sin h; rhh ffi pffiffiffiffiffiffiffiffi 2 4 2 4 2 4 2 2pr 4 2p r    

K

1

1

h

3h

K

1

h

3

ð2Þ

3h

I II ffi þ pffiffiffiffiffiffiffiffiffi  T cos h sin h: sin þ sin cos þ cos rrh ffi pffiffiffiffiffiffiffiffi 2 4 2 2 4 2 2pr 4 2p r 4

The nonsingular term T (T-stress) is added in (1) as it affects the crack orientation. When ry0 ¼ r1 and rx0 ¼ kry0 ¼ kr1 are 0 0 0 the principal stresses and the crack tip is inclined at the angle a to the y -axis then T = r(1  k)cos 2a. The x ,y -reference system follows the axes of principal stresses and the x,y-system represents the crack plane and its normal vector, Fig. 1. The higher order terms of the expansion with respect to the distance r from the crack tip are not introduced in the analysis. 2. Two-dimensional (2D) mixed mode fracture criteria The fracture criteria should be based upon physical models accounting for damage processes of the material in front of the crack tip. However, since the exact description of these processes is difficult, the problems are usually treated within the framework of LEFM by using representative values such as stress components or specific strain energy specified at some finite distance from the crack or sharp notch tip. This characteristic distance specifying the core region affected by damage and plastic deformation is an unknown parameter to be specified experimentally or analytically. 2.1. Monotonic crack growth The first fracture criteria related to the angled crack problem and the orientation of crack growth used the simplest assumption of the core region bounded by a circle of radius rC from the crack tip. The criteria such as: maximum circumferential tensile stress (MTS), minimum strain energy density (SED), maximum energy release rate (MERR) and local symmetry (LS) became popular and were analyzed in the literature. The maximum tensile hoop stress criterion (MTS) formulated by Erdogan and Sih [2] in 1963 was one of the first conditions predicting critical stress and crack growth orientation. Expressing the hoop stress in the form following from (2) without T-stress effect:

K

 3

h

1

3h



K



3

h

3

 3h

I II ffi cos þ cos þ pffiffiffiffiffiffiffiffiffiffiffi  sin  sin ; rhh ffi pffiffiffiffiffiffiffiffiffiffi 2 4 2 4 2 4 2 2pr C 4 2pr C

it is postulated that crack will grow when

ð3Þ

K.P. Mróz, Z. Mróz / Engineering Fracture Mechanics 77 (2010) 1781–1807

@ 2 rhh

@ rhh ¼ 0; @h

@h2

< 0 at h ¼ hPR ;

K

IC ffi rhh ðrC ; hPR Þ ¼ pffiffiffiffiffiffiffiffiffiffi 2prC

1783

ð4Þ

and the stationary condition provides

K I sin hPR þ K II ð3 cos hPR  1Þ ¼ 0:

ð5Þ

For mode I there is hPR = 0 and for pure mode II there is hPR = 70.5° and KIIC = 0.866KIC. The SED-criterion by Sih [3] in 1974 expressed the specific strain energy density at the radius rC in terms of the energy density factor S expressed as follows:

S ¼ a11

K 2I

þ 2a12

p

K I K II

p

þ a22

K 2II

p

ð6Þ

;

where

1 ð1 þ cos hÞðj  cos hÞ; 16l 1 ¼ sinð2 cos h  ðj  1ÞÞ; 16l 1 ¼ ððj þ 1Þð1  cos hÞ þ ð1 þ cos hÞð3 cos h  1ÞÞ; 16l

a11 ¼ a12 a22

ð7Þ

and l is the elastic shear modulus, j = 3–4m for the plane strain case and j = (3  m)/(1 + m) for the plane stress. The crack growth criteria now are

@S ¼ 0; @h

@2S @h

2

< 0 at h ¼ hPR ;

SðhPR Þ ¼

ð1  2mÞK 2IC 4pl

ð8Þ

and the crack growth orientation hPR corresponds to minimum of S reaching its critical value. For pure mode II there is hPR = 79.2° and KIIC/KIC = 1.054 for m = 0.2, where m is the Poisson ratio. For larger ratios KII/KI this criterion provides predictions differing significantly from those of MTS-criterion. The maximum energy release rate criterion (MERR) follows from the Griffith condition and states that crack growth follows the orientation of maximum energy release G = @ P/@s, where P is the potential energy of the element and s denotes the length of the crack path, thus

@G ¼ 0; @h

@2G 2

@h

6 0;

at h ¼ hPR ;

GðhPR Þ ¼

K 2IC ; E

ð9Þ

where E* = E for plane stress cases and E* = E/(1  m2) for plane strain cases. Here GðhÞ ¼ lims!0 ½PðsÞ  Pð0Þ where P(s) is the potential energy for the kinked crack of length s and P(0) is the energy for the initial crack. For a straight crack with a kink of infinitesimal length the SIFs at the kink tip are related to the intensity factor of original crack, namely:

h 3 h  K II cos sin h; 2 2 2 1 h 1 h K II ¼ K I cos sin h þ K II cos ð3 cos h  1Þ; 2 2 2 2

ð10Þ

K i ¼ F ij K j ;

ð11Þ

K I ¼ K I cos3

or

i; j ¼ 1; 2:

The energy release along the kinked crack now is



1 2 ðK þ K 2 II Þ ¼ GðK I ; K II ; hÞ ¼ GC : E I

ð12Þ

The MERR-criterion was analyzed by Hussian et al. in 1974 [4] and by Wu [5] for modes I and II loading. For mode II this criterion predicts hPR = 75.6° and KIIC/KIC = 0.817. The local symmetry conditions (LS) proposed by Goldstein and Salganik [6] in 1974 requires that crack propagation occurs along the path of vanishing KII, thus

K II ¼ 0;

on s-path

ð13Þ

and the crack path orientation is the same as that predicted by Eq. (5) resulting from MTS-criterion. To analyze the relation between the conditions MTS, MERR, LS and max KI, let us express the stationary condition of G as follows

       @G 1 @K I @K  1 @F 11 @F 12 @F 21 @F 22 ¼  K I þ K II II ¼  K I K I þ K II þ K II K I þ K II ¼ 0: @h E E @h @h @h @h @h @h

ð14Þ

1784

K.P. Mróz, Z. Mróz / Engineering Fracture Mechanics 77 (2010) 1781–1807

Assume now that the coefficients of Fij satisfy the conditions

@F 11 =@h @F 12 =@h ¼ ¼ gðhÞ or F 21 F 22

@K I ¼ gðhÞK II : @h

ð15Þ

Then K II ¼ 0 implies @G/@h = 0 and @K I =@h ¼ 0, so these three criteria are equivalent. It can be verified that the components (10) of Fij satisfy the conditions (15). However, they are obtained for vanishing length of the kink crack. The expression of the SIFs with respect to the crack length s in the form

 3 1 pffiffi KðsÞ ¼ K  þ K 2 s þ K ð1Þ s þ 0 s2 ;

ð16Þ

developed by Amestoy and Leblond [9,8,7] indicates that conditions (15) are not satisfied when the higher order terms are included. The apparent crack extension force criterion (CEF) proposed by Strifors [10] in 1973, next applied by Hellen and Blackburn [11] was expressed in terms of path-independent J-integral vector

Jk ¼

Z

ðWnk  ui;k T i Þdl

k ¼ I; II;

ð17Þ

C

where C is an open path connecting two points on opposite crack sides, W is the strain energy density, nk is the k component of the unit outward normal vector to the C, ui,k is the displacement gradient, Ti is the traction on the path C. The J-vector is then projected onto kink growth direction, thus

J p ðhÞ ¼ J 1 cos h þ J2 sin h

ð18Þ

and the crack propagation condition is expressed: max Jp(h) = JPCR. Chang [12] in 1981 proposed the maximum tangential strain criterion (MTSN) specifying the hoop strain ehh ¼ ðrhh  v rrr Þ=E along the crack tip core region r = rC. The crack angle then corresponds to a maximum value of ehh reaching its critical value eT = rT/E, where rT is the brittle tensile fracture strength. Papadopoulos [13] in 1987 proposed the stress invariant criterion (Det-criterion), assuming the third stress invariants as a local fracture condition, reaching its maximum at r = rC. Koo and Choy [14] in 1991 postulated that crack propagates along the direction of maximum tangential strain energy density component Sh = 1/2rheh at r = rC (MTSE). Then, the tangential strain energy density factor C is equal:

C ¼ r 0 Sh ¼ b11

K 2I

p

þ b12

K I K II

p

þ b22

K 2II

p

ð19Þ

;

where

1 ðð1 þ cos hÞðj þ 2 þ cos hÞÞ 64l 1 ¼ ðsin h  3=2  j  3 cos hÞÞ; 64l 1 2 ¼ ð3 sin hðj þ 3 cos hÞÞ: 64l

b11 ¼ b12 b22

ð20Þ

Kong et al. [15] in 1995 formulated the Mt-criterion assuming the crack to grow along the line of maximal stress triaxiality ratio M = rH/rEQ, where rH is the hydrostatic stress, and rEQ denotes the Huber–Mises effective stress, thus

1

2ð1 þ v Þ



h

h



rH ¼ ðrxx þ ryy þ rzz Þ ¼ pffiffiffiffiffiffiffiffiffi K I cos  K II sin ; 3 2 2 3 2pr

  1=2 1 ðrxx  ryy Þ2 þ ðryy  rzz Þ2 þ ðrzz  rxx Þ2 þ 6r2xy 2     1 3 2 9 2 h 2 ¼ pffiffiffipffiffiffiffiffiffiffiffiffi K I  K II sin h þ 2ð1  2v Þ2 K 2I þ 6K 2II cos2 2 2 2 2pr 2 1=2 2 h þ K I K II ð3 sin 2h  2ð1  2v Þ2 sin hÞ : þ8ð1  v þ v 2 ÞK 2II sin 2

rEQ ¼

ð21Þ

However, there exists the possibility to improve accuracy of the proposed fracture criteria by including the T-stress effect into the criterion. By adding the nonsingular terms to singular stress, the value of r specifying the core region becomes essential. There have been numerous studies of extended fracture criteria accounting for T-stress, starting from the analysis of Ewing and Williams [16] in 1972 related to MTS-criterion. The subsequent paper by Swedlow [17], Maiti and Smith [18,19], Wong [20], Wu and Li [21] demonstrated that the crack growth orientation depends on the radius rC and better agreement with experiment can be attained by assuming the value of rC to depend on the orientation angle h.

K.P. Mróz, Z. Mróz / Engineering Fracture Mechanics 77 (2010) 1781–1807

1785

A simple extension can be obtained by assuming the process zone to coincide with the localized plastic zone at the crack tip. This idea was applied by Theocaris and Andrianopoulos [22] who in 1982 proposed T-criterion, modifying the earlier SED-criterion. The specific stress and strain energy S is split into distortional SD and hydrostatic portions SH, thus

S ¼ SD þ SH ; 1  2m ðrxx þ ryy þ rzz Þ2 ; SH ¼ 6E  1 þ m 2 rxx þ r2yy þ r2zz  rxx ryy  ryy rzz  rxx rzz þ 3r2xy : SD ¼ 3E

ð22Þ

As rzz = rzx = rzy = 0 for plane stress and rzz = m(rxx + ryy) for plane strain, there is for the plane stress

SD ¼

  1 þ m ðrxx þ ryy Þ2  3 rxx ryy  r2xy ; 3E

SH ¼

1  2m ðrxx þ ryy Þ2 6E

ð23Þ

and for the plane strain

  1 þ m ð1  m þ m2 Þðrxx þ ryy Þ2  3 rxx ryy  r2xy ; 3E ð1 þ mÞð1  2mÞ ðrxx þ ryy Þ2 : SH ¼ 6E

SD ¼

ð24Þ

Assuming the distortional energy to correspond to plastic flow and the hydrostatic stress energy to decohesion and fracture, the T-criterion postulates that the crack propagates along the direction corresponding to a maximum of total specific stress energy on the perimeter of varying radius, r = rC(h) specified by the condition of constant distortional energy at the yield point, thus

1þm 2 r ; for the plane stress; 3E YS ð1 þ mÞð1  m þ m2 Þ 2 ¼ rYS for the plane strain; 3E

SD ¼ SYS D ¼

ð25Þ

SD ¼ SYS D

ð26Þ

where rYS is the yield stress in uniaxial tension. The T-criterion can now be expressed as follows

@SH ðr; hÞ ¼ 0 for SD ðr; hÞ ¼ SYS D ¼ const: @h

ð27Þ

This criterion is formulated for the varying core region radius. In the case of brittle materials it tends to the SED-criterion as the size of plastic zone is very small and its shape can be assumed as circular. Wasiluk and Golos [23] in 2000 proposed W-criterion assuming that the crack growth angle is specified by the minimum value of Z-factor defined as Z = rp(h)/a where rp (h) is radius of plastic zone and a is half the crack length. The Huber–Mises yield conditions has been applied. The W-criterion is based on an assumption of the minimum energy consumed during the fracture process in the plastic zone. A similar criterion formulated by Khan and Khraisheh [24] in 2004 is based on the assumption that the crack growth orientation corresponds to the local or global minimum of radius of plastic core boundary at the crack tip stating that crack growth follows the direction of minimum distance from the crack tip to the plastic core perimeter. Yan et al. [25] in 1992 specified the plastic core region by applying the yield condition dependent on the first stress and second stress deviator invariants, thus

f ¼ aI1 þ ðJ 2 Þ1=2  c ¼ 0;

ð28Þ

where I1 ¼ rkk ; J 2 ¼ sij = rij  dijrkk/3, rij and sij are the stress tensor and the deviatoric stress tensor, respectively, I1 and J2 are the first stress invariant and the second stress deviator invariant, dij is the Kronecker symbol, a and c are the positive materials constants. In the case a = 0 the Huber–Mises yield condition is obtained. However, in the vicinity of the crack tip we may distinguish three zones of different damage states, presented schematically in Fig. 2. First, the process zone (PZ) of large damage affecting essentially material stiffness and strength. Second, the plastic slip zone and the third zone characterized by growth of microcracks induced by tensile microstress [26]. This damage is assumed to be related to the hydrostatic stress energy density SH. Basing on this assumption Mróz [27] in 2008 proposed the MK-fracture criterion postulating that size of damage zone is specified by the condition SH ¼ SCH and macrocrack propagation follows the direction of smallest plastic dissipation, that is corresponds to a minimum value of the distortional stress energy SD specified along the perimeter SH ðr; hÞ ¼ SCH ¼ const:, thus 1 s s , 2 ij ij

@SD ðr; hÞ ¼ 0; @h

for SH ðr; hÞ ¼ SCH ¼ const:

ð29Þ

or

SH j ! hpr SD max

at

SH ¼ SCH ¼ const:

!

SD ðrðSCH ; hÞÞjmin ! hpr

ð30Þ

1786

K.P. Mróz, Z. Mróz / Engineering Fracture Mechanics 77 (2010) 1781–1807

S (σ ) = const.

PZ

S (σ ) = const.

ρ

ρ (σ ) ρ (σ )

ρ (σ )

0

H

Fig. 2. The schematic plot of damage zones within the crack tip region (where q is the conceptual density of defects).

where

2 rðSCH ; h; aÞ ¼ r H jSH ¼const: ¼

2

p

32

h h 6 K I cos 2  K II sin 2 4qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi C 6E SH ð1mÞð1þ  Tð m Þ2

7 5



ð31Þ

and m* = 0 for plane stress, m* = m for plane strain cases. The term under square root can be replaced by the equivalent stress rC, so we may write

rðrC ; h; aÞ ¼

 2 2 K I cos 2h  K II sin 2h : p rC  TðaÞ

ð32Þ

The crack growth criterion can now be formulated as follows. For the specified loading the distance defined by (31) provides the size of damage zone associated with the assumed volumetric energy level. The critical value r(rC, hpr, a) = rC is related to the fracture strength parameters KI, KII. The crack growth does not occur when rC P rðrC ; hpr ; aÞ and crack growth occurs when r(rC, hpr, a) P rC with the crack growth onset specified by r(rC, hpr, a) = rC. This criterion will be discussed in the following. Consider the particular case of mode I when a = p/2 and the crack propagation occurs along the primary crack, h = 0. From (32), it follows that

K.P. Mróz, Z. Mróz / Engineering Fracture Mechanics 77 (2010) 1781–1807

p 2 rCI ðrC ; 0; Þ ¼ 2 p

2  2 K IC 2 K IC ¼ : rC  Tðp2Þ p rC  r1 ð1  kÞ

1787



ð33Þ

Assume now, that the critical state is reached on the radius rC = rCI = const. The crack propagation criterion can now be presented as follows

h h rC  Tð a Þ rC  r1 ð1  kÞ cos 2a ¼ K IC K I cos  K II sin ¼ K IC ; 2 2 rC þ r1 ð1  kÞ rC  T p2

ð34Þ

where T(p/2) = r1(1  k). This criterion is based on the assumption that mode I is prevailing along the crack path and KIC is the single crack growth parameter. However, in more general cases Eq. (34) should be written in following form:

  hPR jK II ¼0 rC  TðaÞ h h K I cos  K II sin ¼ K IC sin : 2 2 2 rC  TjK II ¼0

ð35Þ

The general criterion can be proposed by accounting for two fracture parameters of mode I and II. For pure mode II the critical value rCII equals

rCII ¼

 2  2 2 K IIC sinðhPR =2Þ 2 0:76 K IIC ¼ ; p rC  TðaÞ p rC

ð36Þ

as T(a) = 0 and sin(hPR/2) is equal to 0.76 for pure mode II loading. In more general cases Eq. (36) should be replaced by

rCII

h j 32 2 PR K I ¼0 2 2 4K IIC sin 5 : ¼ p rC  TjK I ¼0

ð37Þ

Assume that for a mixed mode loading the critical value of rC is specified by the relation

rC ¼ cos c r CI þ sin c r CII ;

ð38Þ

K II =K IIC K I =K IC

ð39Þ

where

tan c ¼

is a measure of mode mixity. Let us note that rC = rCI for mode I (sin c = 0 and cos c = 1) and rC = rCII for mode II (cos c = 0 and sin c = 1). The crack propagation criterion is now stated in the form (Fig. 1)

   2   h h rC  r1 ð1  kÞ cos 2a 2 rC  r1 ð1  kÞ cos 2a 2 K I cos  K II sin ¼ cos c K IC þ sin c 0:76 K IIC 2 2 rC þ r1 ð1  kÞ rC

ð40Þ

and in the general form:

!2 !2  2     hPR jK II ¼0 rC  TðaÞ hPR jK I ¼0 rC  TðaÞ h h K I cos  K II sin ¼ cos c K IC cos þ sin c K IIC sin 2 2 2 rC  TjK II ¼0 2 rC  TjK I ¼0

ð41Þ

or

2 !2 !2 3  2     hPR jK II ¼0 rC  TðaÞ hPR jK I ¼0 h h rC  TðaÞ 2 4 5; K I cos  K II sin ¼ K IC cos c cos þ sin c / sin 2 2 2 rC  TjK II ¼0 2 rC  uTjK II ¼0

ð42Þ

where KIIC = /KIC and TjK I ¼0 ¼ u  TjK II ¼0 . However, in many cases condition (34) which incorporates only KIC is sufficiently accurate, especially for brittle materials. In Fig. 3, the analysis of crack growth based on Eq. (34) incorporating only KIC is presented and the critical values of KI and KII are specified for varying mode mixity. Fig. 3 presents the results according to Eq. (41) in terms of mean values of TjK I;II ¼0 and also KI,II C. Fig. 4 presents the bounding curves based on Eq. (41) and extreme values of TjK I;II ¼0 and KI,II C. The analysis is compared with the experimental results of Ayatollahi et al. [28] who conducted mixed mode I/II fracture tests on polymethylmethacrylate (PMMA) and specified fracture loads for varying initial crack angles. The numerical analysis of [28] was based on MTS-criterion with account for T-stress. The constant radius r = rC was assumed in specifying the core region for accounting for damage material ahead of the crack tip. Assuming rC varying between 0.065 and 0.25 mm these authors obtained fair correlation between mixed mode fracture toughness prediction and experimental results. The present analysis applying MKcriterion (34) and varying r = rC(h) provides more accurate prediction of critical load locus in the KI, KII plane. Fig. 5 shows the Huber–Mises elastic–plastic boundary and the decohesion boundary for the nonsingular solution (with T-stress term) based on Eq. (31) for a = 90° cf. (Fig. 1), and Figs. 6 and 7 show the similar energy profiles for the singular stress field for a = 30°, 90°. Particularly in the case a = 90° the decohesion area is considerably larger than the plastic area and for the same value of load-

K.P. Mróz, Z. Mróz / Engineering Fracture Mechanics 77 (2010) 1781–1807

K II / K IC

1788

K I / K IC

K II / K IC

Fig. 3. Mixed mode fracture locus specified experimentally in [28]. The solid line represents the average test results, the dashed line represents results according to Eq. (34) and the dotted line represents results obtained from Eq. (41). The results according to Eq. (41) based on the mean values of TjK I ¼0 ; TjK II ¼0 and KIC, KIIC.

K I / K IC Fig. 4. Mixed mode fracture locus specified experimentally in [28]. The solid line represents the average test results and the dotted and the dashed lines represent results according to Eq. (41) based on the extreme values of TjK I ¼0 ; TjK II ¼0 and KIC, KIIC.

ing the shape and the size of both boundary curves vary, when the T-term is introduced. Fig. 8 presents the energy profiles for the plane stress and plane strain conditions in pure shear loading. Next, Fig. 9 presents graphically the MK-criterion: T D ðrðhÞ; T const: Þ calculated along the r ¼ rðT c: H H ; hÞ contour following form (29)–(31) and the nonsingular solution, where the minimum of the radius of T D ðrðT cH ; hÞÞ represents the direction of the crack propagation. Fig. 10 and Table 1 present the angle of crack propagation hpr predicted by MK-criterion for the various main crack inclinations a, for the uniaxial tension (k = 0) and various ratios of applied and critical stress. The most interesting are the differences of the crack propagation angles for the same value of a but different load values. This effect can be associated with the T-stress which affects significantly the propagation angles at varying load levels. The T-stress term is also essential in stability control of straight crack path under mode I loading conditions [29]. Also the shape and size of the plastic zone ahead of the crack tip depends on the value of Tstress [30]. The analysis of [31] demonstrates the importance of retaining the second term in asymptotic series representing the local stress.

1789

K.P. Mróz, Z. Mróz / Engineering Fracture Mechanics 77 (2010) 1781–1807 90

0.0008

120

90

0.0008

120

60

0.0006 150

150

30

0.0004

30

0.0004

0.0002

0.0002

180

0

180

330

210

0

330

210

300

240

60

0.0006

300

240

270

270

(a)

(b)

Fig. 5. The Mises elastic–plastic boundary, TD = const. (the dotted line) and the decohesion boundary, TV = const. (the solid line) Eq. (31), for Al 7075-T6: pffiffiffiffiffi pffiffiffiffiffi a ¼ 90 ; rYS ¼ 538 MPa; rC ¼ 586 MPa;K TH ¼ 1 MPa m; K C ¼ 25 MPa m; E ¼ 71 GPa; m ¼ 0:33; 2a ¼ 20 mm. The nonsingular solution (T – 0). (a) The case of plane stress. (b) The case of plane strain.

90 120

0.0015

90 60

0.0004

120

60

0.001 150

30

0.0002

150

30

0.0005

180

0

330

210

300

240

180

0

330

210

300

240

270

270

(a)

(b)

Fig. 6. The Mises elastic–plastic boundary, TD = const. (the dotted line) and the decohesion boundary, TH = const. (the solid line) Eq. (31), for Al 7075-T6: pffiffiffiffiffi pffiffiffiffiffi rYS ¼ 538 MPa; rC ¼ 586 MPa; K TH ¼ 1 MPa m; K C ¼ 25 MPa m; E ¼ 71 GPa; m ¼ 0:33; 2a ¼ 20 mm. The singular solution (T = 0) and the plane strain conditions. (a) a = 90°; (b) a = 30°.

In fact, the T-term introduced the correction to the core contour rC = rC(h), Eq. (31), on which the stress tensor components are specified. For the singular solution with T = 0 the ratio r/rC does not affect the orientation angle as the stress components are proportionally varying along any radius from the crack tip. Similarly, when the core region is specified by the constant radius r = rC, there is no effect of r/rC ratio on the crack propagation angle. The scattering of results can be associated with the inhomogeneity of material represented in the MK-criterion by scatter in the values of rC. However, it is interesting to note that the smallest scattering of these results occurs in neighborhood of a = 45° consistently with the MK-criterion predictions. The different predictions of the angle crack propagation by accounting for the T-stress term was observed already in several works. In 1995 Seweryn and Mróz [33,34] discussed the application of the non-local failure criterion to sharp notches and cracks and formulated the R-criterion (see Section 4). They showed that MTS-criterion is a particular case of non-local stress condition. In 1998 Seweryn [35] discussed the effect of nonsingular T-stress term on the crack propagation in R-criterion (Fig. 11) for different ratios d0/a, where d0 is the size parameter representing the size of damage zone generating different values of the propagation angles, but without any physical background. The size parameter d0 representing the size of damage zone can be equivalent to the Griffith condition in the case of tensile crack propagation. This provides

d0 ¼

 2 2 K IC

p rc

:

ð43Þ

1790

K.P. Mróz, Z. Mróz / Engineering Fracture Mechanics 77 (2010) 1781–1807 90 120

0.0015

90

0.0004

120

60

60

0.001 150

0.0002

150

30

30

0.0005

180

0

180

330

210

0

300

240

330

210

300

240

270

270

(a)

(b)

Fig. 7. The Mises elastic–plastic boundary, TD = const. (the dotted line) and the decohesion boundary, TH = const. (the solid line) Eq. (31), for Al 7075-T6: pffiffiffiffiffi pffiffiffiffiffi rYS ¼ 538 MPa; rC ¼ 586 MPa; K TH ¼ 1 MPa m; K C ¼ 25 MPa m; E ¼ 71 GPa; m ¼ 0:33; 2a ¼ 20 mm. The singular solution (T = 0) and the plane stress conditions. (a) a = 90°; (b) a = 30°.

90

90

0.0015

120

60

60

0.001

0.001

150

150

30

30

0.0005

0.0005

180

0

330

210

300

240 270

(a)

0.0015

120

180

0

330

210

300

240 270

(b)

Fig. 8. The Mises elastic–plastic boundary, TD = const. (the dotted line) and the decohesion boundary, TH = const. (the solid line) Eq. (31), for Al 7075-T6: pffiffiffiffiffi pffiffiffiffiffi rYS ¼ 538 MPa; rC ¼ 586 MPa; K TH ¼ 1 MPa m; K C ¼ 25 MPa m; E ¼ 71 GPa; m ¼ 0:33; 2a ¼ 20 mm. The singular solution (T = 0) and a = 90° for the pure shear loading. (a) Plane stress conditions. (b) Plane strain conditions.

He concluded that the approach applied to the materials with defects should be modified in order to account for stress redistribution due to plastic strain (or microcracks) in order to predict better crack initiation or growth. In the case of MK-criterion these differences result explicitly from varying geometrical configurations and loading conditions of the problem. In 2D cases the R-criterion is expressed in terms of the failure stress function averaged over the distance d0, but the distance rC = const. or r(rC, hpr, a) specifies the point ahead of the crack tip where the stress and specific energy are evaluated. In general there should be d0 P r C . However, for large stress gradient or the singular stress field and for a = p/2, k = 0, h = 0 (Fig. 1) r(rC, hpr, a) and d0 values can be equal, when KI = KIC. The present discussion of crack growth conditions and their path orientation in the plane strain or stress cases is based on linear-elastic stress and strain fields at the crack tip. For the stress or local specific energy conditions the characteristic distance rC from the crack tip constitutes a new parameter. In fact, the crack propagation condition now has a non-local character. The parallel formulation of critical plane models have the same character as they are expressed in terms of mean stress or strain components over plane face element of specified finite area. The use of distortional and volumetric specific energies to determine rC and the failure condition, such as (34) or (40), provides the simple methodology of prediction of crack path and the limit fracture locus. Let us note that crack growth orientation now depends on the stress ratio r/rC which affects the value of r specified by (32). Using the formulation of linear fracture mechanics, the size parameter d0, or critical distance r is specified from the volumetric energy condition as in the MK-criterion, cf. Eqs. (32) and (36). For the non-local critical plane criteria, such as (70),

1791

K.P. Mróz, Z. Mróz / Engineering Fracture Mechanics 77 (2010) 1781–1807 90 120

0.15

90

0.1

120

60

60

0.08 0.1 150

30

0.06

150

30

0.04

0.05

0.02 180

0

210

330

240

180

0

210

330

300

240

270

300 270

(a)

(b) rðT const: ; hÞ H

Fig. 9. The graphically represented MK-criterion. The decohesion boundary, nonsingular solution (T – 0) and the plane strain conditions: (a) a = 60°; (b) a = 30°.

(the dotted line) and the T D ðrðT cH ; hÞÞ line (the solid line) for the

σ =1 σC

θ [° ]

Experimental results

T =0

β = (π / 2 − α ) [ ° ] Fig. 10. The angle of crack propagation hpr based on the MK-criterion for various initial crack orientations b = p/2  ain uniaxial tension (k = 0) as compared to experimental results of [32] (bounded by dotted lines) and the varying ratios of applied to critical stress (according to Table 1).

Table 1 The angle of crack propagation hpr for varying initial crack orientation a in the case of uniaxial tension (k = 0) and varying ratios of applied and critical stress.

a r/rC

15

30

45

60

80

T=0 0.1 0.2 0.4 0.6 0.8 1

85.7 86.7 87.7 89.7 91.2 92.7 93.7

72.7 73.7 74.7 76.7 78.7 80.2 81.2

59.6 59.6 59.6 59.6 59.6 59.6 59.6

45.1 43.1 40.6 37.6 35.6 34.1 33.1

19.3 15.5 13.5 11.5 11.0 10.5 10.0

the size parameter d0 is specified from the equivalence condition with the Griffith condition for the tensile crack growth, cf. Eq. (43). The strain energy density criterion for a non-linear Ramberg–Osgood material was assumed by Shlyannikov [36,37] in order to specify the size of fracture damage zone. The combined non-local stress and energy release conditions were proposed by Seweryn [38] and next by Taylor [39], Carpinteri et al. [40] and Leguillon [41]. The averaged stress conditions on a

1792

K.P. Mróz, Z. Mróz / Engineering Fracture Mechanics 77 (2010) 1781–1807

θ pr

β = π /2 −α Fig. 11. The angle of crack propagation, based on [35] for different ratio d0/a and experimental results [32].

critical plane element of size d0 is applied and the value of d0 is next specified from the mean energy release value for a crack growing from a to a + d0, thus

Z

aþd0

GðaÞda ¼ Gc d0 ;

ð44Þ

a

where Gc is the assumed critical value of the energy release. 3. Three-dimensional (3D) cracks and fracture criteria Structural elements are usually subjected to complex stress states inducing singular 3D regimes at sharp defects (cracks, notches, etc.). The analysis of 3D singular regimes is more complex and crack growth modes are not well understood. 3.1. Crack growth modes Any growth of the plane crack surface which is located through the thickness of a plate, can be obtained by superposition of three basic modes, namely: mode I (opening mode) where opposing crack surfaces move directly apart; mode II (sliding mode) where crack surfaces move over each other perpendicularly to the crack front; and mode III (tearing mode), where crack surfaces move over each other parallel to the crack front. Two fundamental concepts can now be assumed  the new crack growing surface is smooth and continuous at the edge of existing pre-crack  the new crack surface is composed of facets of modes I, II and III oriented according to local critical state conditions. The crack surface is composed of the pattern of facets and its evolution is governed by geometric and mechanical characteristics of the element. We shall refer to the first case as smooth crack surface models and to the second as rough crack surface models. To discuss these two approaches, consider the 3D square plane crack, Fig. 12 within the 3D element. It can be expected that new crack surface may grow from each edge. However, satisfying the continuity conditions the new crack surface may evolve in mode I or II or in combined mode I + II. But the mode III loading and the associated crack facets generated locally by the KI max condition violate the continuity at the existing crack edge. Thus the mode III loading may induce the rough crack surface combined with facets developed in mode II, Fig. 13. In fact, brittle materials exhibit rough crack surface in mode III composed of facets of different modes, cf. [42,43]. In ductile materials the shear band evolution may affect the crack roughness due to localized plastic deformation and damage growth. In fact Dyskin and Salganik [44] analyzed 3D cracks in compression and showed that 3D crack growth can result in a more complicated growth mechanism than 2D. They investigated possible mechanisms of wing crack growth induced also by rough crack surface and associated dilatancy effect [45]. A more general case of the 3D crack is the elliptic crack, which was analyzed in uniaxial compression tests by Adams and Sines [46] on PMMA samples. They showed that crack growth generally occurs in wing mode, but in addition at the lateral parts of crack edge a number of microcracks are developed. This can be interpreted as the mode III microcracks in brittle

K.P. Mróz, Z. Mróz / Engineering Fracture Mechanics 77 (2010) 1781–1807

1793

a a

I mode

II mode

Fig. 12. 3D crack and pure modes of smooth crack tip deformation.

I mode

II mode θ (s )

III mode

θ (s)

Fig. 13. Pure modes of rough 3D crack tip deformation.

materials, cf. Fig. 18. The extensively investigated 3D elliptical cracks embedded in brittle material under compression loading have also been presented by Dyskin et al. [45,47]. 3.2. Rough (dilatant) crack models The effect of crack surface contact and asperity interaction is essential for modes II and III loading. In fact, the growth of pre-crack generated in mode I and next subjected to shear is associated with evolution of roughness pattern due to process of microcracking and associated inclined facets in mode I with subsequent connecting microcracks, cf. Pook [48]. The interface sliding along formed asperities induced mode I stress and crack dilatancy. Crack tip shielding then occurs due to frictional resistance to sliding and asperity interaction. The cases of closed or partially closed cracks subjected to shear and exhibiting contact shielding are numerous and occur, for instance, in compression or shear induced fracture of rocks or ceramic materials, rolling contact induced sub-surface fatigue cracks, mixed mode fatigue crack growth, etc. The referenced papers [49,50,43,51–56] contain both analytical and experimental studies of asperity interaction modes, specification of the effective SIFs and prediction of crack growth rates. The assumption of the contact interface interaction at the crack front or at the whole cracked interface provides different modelling effects. The detailed review of literature is not presented here. A simplified model of a closed crack interface will only be discussed in the following, cf. Mróz and Seweryn [55]. Consider a crack surface in a form of wedge-shaped asperities, inclined at the angle cz 2 (0, p/2) to the nominal crack l l plane, Fig. 14. Denote the local stresses acting on wedge flanks Cl by rn and sn , so that the local friction condition on Cl is

sln ¼ gl rln ¼ tan cl rln ;

ð45Þ

where gl is the local friction coefficient and cl is the friction angle. The stress rzn and szn acting on the nominal crack segment are expressed from the equilibrium equations for a single asperity, thus

1794

K.P. Mróz, Z. Mróz / Engineering Fracture Mechanics 77 (2010) 1781–1807

μ σ nμ τ n

Γ0

γ

Γμ

s1

z

τ nz

Γμ

γz s2

σ nz

Fig. 14. Crack with wedge-shaped asperities and microstresses acting on asperity facets Cl and macrostress acting on the nominal crack plane segment C0.

s s2

s s2

rzn ¼ rln 1 ðcos cz  gl sin cz Þ;

szn ¼ sln 1 ðsin cz þ gl cos cz Þ;

ð46Þ

where s1 denotes the wedge flank length and s2 is the asperity length within the plane C0. The conditions on the nominal plane C0 can be expressed as follows

szn ¼ rzn tanðcz þ cl Þ;

u# ¼ jur j tan cz ;

# ¼ p:

ð47Þ K dI ; K dII

These conditions can be expressed in terms of the asymptotic fields (1) with neglect of T-stress. Denoting by the effecl l tive SIFs for the dilatation crack model and by K I ; K II the SIFs resulting from the stress on the wedge asperity flanks Cl, we obtain l

l

K dI ¼ jK dII j tan cz ; jK II j ¼ K I tanðcz þ cl Þ:

ð48Þ

Denote by KI, KII the SIFs for a smooth and plane crack. Applying the superposition principle for the external loading r, s and the wedge asperity loading rzn ; szn , the effective SIFs are

sin cz ðK I sinðcz þ cl Þ þ jK II j cosðcz þ cl ÞÞ; cos cl cos cz l jK dII j ¼ K II þ K II ¼ ðK I sinðcz þ cl Þ þ jK II j cosðcz þ cl ÞÞ: sin cl l

K dI ¼ K I þ K I ¼

ð49Þ

These relations are valid when the contact occurs on Cl, thus

K I tanðcz þ cl Þ 6 jK II j 6 K I cotðcz Þ:

ð50Þ z

These inequalities specify the sliding domain B, Fig. 15. When jKIIj
K II K

= K cot( γ )

K

= − K tan( γ + γ )

B

A C B

A

γ

KI

B

C

Fig. 15. Domains of crack loading for the dilatant crack model: open crack (A), slip at contact interface (B) and closed crack (C).

K.P. Mróz, Z. Mróz / Engineering Fracture Mechanics 77 (2010) 1781–1807

KI

KI

K II

K II

K Id

K Id

K IId K K

K K

K IId

K K

K K

γ z = 0° , γ μ = 10°

1795

γ z = 20 °, γ μ = 10°

ψ [° ]

ψ [° ]

Fig. 16. Values of SIFs specified for the frictional and dilatant crack models, based on [55]. The angle w specifies the combined mode loading, tan w = KII/KI.

z

β −ω

x

z

σ

y

z

Crack plane

B

t

r θ

x

y

n

b A σ Fig. 17. Flat crack with linear front.

Mode III a>b

Mode II a

b Mode II

Fig. 18. 3D elliptic crack growth in compression.

circular or elliptical flat crack, Fig. 18. The case of a linear crack edge can be reduced to a plane case. In fact, assuming the uniaxial loading by the principal stress r oriented at angles b and x to the coordinate axes x, y, z, we can write, cf. Fig. 17

1796

K.P. Mróz, Z. Mróz / Engineering Fracture Mechanics 77 (2010) 1781–1807

pffiffiffiffiffiffi 2 K I ¼ r pb sin b; pffiffiffiffiffiffi K II ¼ r pb sin b cos b sin x; pffiffiffiffiffiffi K III ¼ r pb sin b cos b cos x

ð51Þ

and the stress intensity factors (SIF) remain constant on the crack edge. However, they should vary on a curvilinear front AB. In general, SIF depends on the curvature of the crack front. The stress analysis at the elliptical crack front was presented by Sih et al. [61–63]. Firstly in [62,61] it was shown that the three-dimensional stress state (Fig. 18) in a certain plane is identical to the two-dimensional case. However the stress intensity factor in general depends upon the curvature of the crack edge for three-dimensional problems. The same applies to the displacement field. Thus introducing the local reference system n, t, z at the crack front, Fig. 19, there is

K ðsÞ

K ðsÞ

K ðsÞ

K ðsÞ

I II ffi A1  pffiffiffiffiffiffiffiffiffi A2 rn ffi pffiffiffiffiffiffiffiffi 2p r 2pr I II ffi B1 þ pffiffiffiffiffiffiffiffiffi B2; rz ffi pffiffiffiffiffiffiffiffi 2p r 2pr 

K ðsÞ

K ðsÞ



I ffi C1  pIIffiffiffiffiffiffiffiffiffi C2 ; rt ffi 2m pffiffiffiffiffiffiffiffi 2pr 2pr

K ðsÞ

K ðsÞ

I II ffi D1 þ pffiffiffiffiffiffiffiffiffi D2; rnz ffi pffiffiffiffiffiffiffiffi 2pr 2pr

K ðsÞ

rnt ffi  pIIIffiffiffiffiffiffiffiffiffi E1; 2pr K ðsÞ

rzt ffi pIIIffiffiffiffiffiffiffiffiffi E2 2p r where

    h h 3h h h 3h 1  sin sin ; A2 ¼ sin 2 þ cos cos ; 2 2 2 2 2 2   h h 3h h h 3h 1 þ sin sin ; B2 ¼ cos sin cos ; B1 ¼ cos 2 2 2 2 2 2 h h C1 ¼ cos ; C2 ¼ sin ; 2 2   h h 3h h h 3h D1 ¼ cos sin cos ; D2 ¼ cos 1  sin sin ; 2 2 2 2 2 2 h h E1 ¼ sin ; E2 ¼ cos 2 2 A1 ¼ cos

Fig. 19. The elliptical crack with global and local coordinate systems x, y, z and n, t, z.

ð52Þ

K.P. Mróz, Z. Mróz / Engineering Fracture Mechanics 77 (2010) 1781–1807

1797

and

 12

r sin2 b b

1

2

2

ða2 sin a þ b cos2 aÞ4 ; " #  2 # 12 " r sin b cos b b b a sin a cos x b cos a cos x K II ðsÞ ¼  1

þ 2 ; 1 2 02 02 02 2 2 a a ðk þ mk ÞEðkÞ  mk KðkÞ ðk  mÞEðkÞ þ k KðkÞ ða2 sin a þ b cos2 aÞ4 " # " #  2  12 ð1  mÞr sin b cos b b b a sin a cos x b cos a sin x  2 1 

K III ðsÞ ¼ : 1 2 02 02 02 2 2 a a 2 2 4 ðk  mÞEðkÞ þ k KðkÞ ðk þ mk ÞEðkÞ  mk KðkÞ ða sin a þ b cos aÞ

K I ðsÞ ¼

EðkÞ

a

ð53Þ

The coefficients K(k) and E(k) are complete elliptic integrals of the first and second kind associated with the argument 02 2 k ¼ ½1  ðb=aÞ2 1=2 where k ¼ 1  k and a specifies the location along the crack border. In the compression case r should be replaced by r. In the subsequent paper Hartranft and Sih [63,64] provided the local singular 3D stress field at the crack front for small scale yielding. It can be expressed in the local spherical coordinates (r, h, /) in the form













K I ðsÞ K II ðsÞ ffi cos 2h 1  sin 2h sin 3h ffi sin 2h 2 þ cos 2h cos 3h rn ffi pffiffiffiffiffiffiffiffiffiffiffiffiffi  pffiffiffiffiffiffiffiffiffiffiffiffiffi ; 2 2 2pr cos / 2pr cos / K I ðsÞ K II ðsÞ ffi cos 2h 1 þ sin 2h sin 3h ffi cos 2h sin 2h cos 3h rz ffi pffiffiffiffiffiffiffiffiffiffiffiffiffi ; þ pffiffiffiffiffiffiffiffiffiffiffiffiffi 2 2 2pr cos / 2pr cos /





2m ffi K I ðsÞ cos 2h  K II ðsÞ sin 2h ; rt ffi pffiffiffiffiffiffiffiffiffiffiffiffiffi 2pr cos /





K I ðsÞ K II ðsÞ ffi cos 2h sin 2h cos 3h ffi cos 2h 1  sin 2h sin 3h rnz ffi pffiffiffiffiffiffiffiffiffiffiffiffiffi þ pffiffiffiffiffiffiffiffiffiffiffiffiffi ; 2 2 2pr cos / 2pr cos /

ð54Þ

1 ffi K III ðsÞ sin 2h ; rnt ffi pffiffiffiffiffiffiffiffiffiffiffiffiffi 2pr cos / 1 ffi K III ðsÞ cos 2h : rzt ffi pffiffiffiffiffiffiffiffiffiffiffiffiffi 2pr cos /

However, the asymptotic series expansion in 3D can be extended by adding more terms to the singular solution, cf. [64]. The most interesting is the second term, namely the T-stress which for 3D cases was not sufficiently analyzed. Based on the linespring method, Wang and Parks [65] evaluated the T-stress distribution along a semi-elliptical surface crack front in 3D plate. By introducing an interaction integral, Nakamura and Parks [66] developed method to obtain T-stress distribution along 3D crack fronts from finite element solutions. Effects of plate thickness and Poisson’s ratio on the T-stress were also explored in [65]. Though there are numerous fracture criteria proposed for 2D stress states under I/II mixed mode conditions, only several 3D fracture criteria have been proposed and the related experimental work is limited. The first approach is based on the assumption that a new increment of fracture surface developing at the crack front is specified by locus of critical points and linear segments connecting these points to the crack front. The smooth crack surface is then generated. The other approach is based on the critical plane concept by specifying orientation of the critical plane element of a new fracture surface near the crack front. The critical plane approach allows for the rough crack surface evolution. Using the singular stress components of Eq. (54), Sih and Cha [67] extended the S-criterion to 3D by specifying the strain energy density near the crack front as follows

dW 1 Sðh; sÞ ¼ þ Oð1Þ; dV r cos /

ð55Þ

where

Sðh; sÞ ¼ a11

KðsÞ2I

p

þ 2a12

K I ðsÞK II ðsÞ

p

þ a22

K 2II ðsÞ

p

þ a33

K 2III ðsÞ

p

ð56Þ

and

16la11 ¼ ð3  4m  cos hÞð1 þ cos hÞ; 16la12 ¼ 2 sin hðcos h  1 þ 2mÞ; 16la22 ¼ 4ð1  mÞð1  cos hÞ þ ð3 cos h  1Þð1 þ cos hÞ;

ð57Þ

16la33 ¼ 4: In the case of / = 0° Eq. (57) are the same as Eq. (7) in the plane case. Similarly, as in the plane case, a minimum of S is searched on a sphere centred at each point on the crack front [67,68]. The crack growth occurs when S = Smin reaches critical value, SC, and the size of rC along a three-dimensional crack front is assumed to vary such that Smin (h,s)/rC (s) = (d W/dV)cr remains constant. The continuous formulation S(h)/cos / exhibits a local minimum at (h0, /0) in the region ðp 6 h 6 pÞ; ðp=2 6 / 6 p=2Þ, provided Sðh; /Þ P Sðh0 ; /0 Þ. It should be noted that SðhÞ= cos / attains a local minimum always in the normal plane to the crack front, so / = 0 in our case and direction of the crack growth does not depend on KIII. So,

1798

K.P. Mróz, Z. Mróz / Engineering Fracture Mechanics 77 (2010) 1781–1807

a minimum of S is searched on a circle centered at each point on the crack front. However, as this radius does not affect the value of S-factor, then rC(s) = Smin(hpr, s)/(dW/dV)cr and generates the initial segment of the fracture surface. The extension of stress triaxiality condition (Mt-criterion) to 3D cases was presented by Kong et al. [15]. Introducing the stress triaxiality parameter



rH F 1 ¼ ; rEQ F 2

ð58Þ

where

pffiffiffi F 1 ¼ 2 2ð1 þ v Þ K I cos 2h  K II sin 2h ; 2 31=2  2 3 2 K I  92 K 2II sin h þ ð2ð1  2v Þ2 K 2I þ 6K 2II Þ cos2 2h 2 5 ; F 2 ¼ 34 2 þ8ð1  v þ v 2 ÞK 2II sin 2h þ K I K II ð3 sin 2h  2ð1  2v Þ2 sin hÞ þ 6K 2III

ð59Þ

the value of the ratio M is maximized with respect to the angle h, as it does not depend on /. The optimal orientation segments on the existing crack front then generate a new incremental crack surface and its size is specified by setting the critical value of KI = KIC or of the effective SIF measure. In similar way we can extend the MK-criterion to 3D problem using Eqs. (52) and (53) in the following form m ðr þ r þ r Þ; T H ¼ 12 n z t 6E   2 1þm T D ¼ 6E ðrt  rz Þ þ ðrz  rn Þ2 þ ðrn  rt Þ2 þ 6ðr2nz þ r2nt þ r2zt Þ ;

ð60Þ

then

rH ðrC ; h; aÞ ¼ rH jT H ¼const ¼

 2 2 K I cos 2h ð1 þ mÞ  K II sin 2h ð1 þ mÞ

p

rC

ð61Þ

:

The results of applied MK-criterion to the 3D problem of growth of elliptical crack under mixed loading conditions according to Fig. 19 are presented in Fig. 20. The crack growth for pure tensile and shear loadings is shown in Fig. 21. The sizes of the crack wings are calculated from Eq. (61). It is seen that for pure tensile loading the elliptical crack grows in its plane to a circular shape and for shear loading the wing cracks emanate from two boundary portions along the shear stress direction. 4. Critical plane criteria Let us now present an alternative formulation of crack growth criteria based on the concept of critical plane. Such criteria are expressed in terms of traction or strain components on the material plane element. The unit vector n specifies the orientation of plane on which the normal and shear stresses are specified as follows:

ω=

π 4

,β=

π

3 π π ω= ,β = 4

4

ω=

π 4

,β=

π 6

Fig. 20. 3D elliptical smooth crack growth based on the MK-criterion for the mixed mode loading.

1799

K.P. Mróz, Z. Mróz / Engineering Fracture Mechanics 77 (2010) 1781–1807

rn ¼ rij ni nj ¼ r11 n21 þ r22 n22 þ r33 n23 þ 2r12 n1 n2 þ 2r23 n2 n3 þ 2r31 n3 n1 ; sn ¼

qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi rij nj rik nk  ðrij ni nj Þ2

ð62Þ

and the components of any radius vector r or normal vector n are expressed in the form (Fig. 22)

x ¼ r cos h cos /;

y ¼ r cos h sin /;

z ¼ r sin h:

ð63Þ

Similarly, the normal and shear strain en, cn can be expressed. The simplest critical plane orientation is formulated by assuming that crack will follow along the critical plane on which the normal tensile stress rn is maximal. This criterion was applied by Scholmann [69] and used to specify the orientation angles h0, /0 for the 3D stress state. In fact, formulating the problem

max n

rn ¼ max ðrij ni nj Þ subject to nk nk ¼ 1 n

ð64Þ

and considering the Lagrangian L ¼ rij ni nj  kðnk nk  1Þ, the optimality condition is as follows

dL ¼ ðrij nj  k ni Þ d ni ¼ 0 or

rij nj ¼ k ni ;

ð65Þ

so the maximal normal stress is the principal stress and the critical plane is the maximal principal stress plane and the angles h0, /0 are specified by solving the eigenvalue problem (65). A more general critical plane condition can be formulated as follows

F ¼ max f ðrn ; sn ; en ; cn Þ  fC ¼ 0;

ð66Þ

n

where fC represents the critical value reached by the failure condition generally depending on both stress and strain components associated with the respective planes. Here en ¼ ðn  e nÞn and cn ¼ ð1  n nÞen where 1 is the unit tensor. A particular form (66) applied to fatigue problems is obtained by applying the strain energy density associated with the amplitudes of stress and strain components acting on the critical plane, cf. Glinka et al. [70].

F ¼ W ¼ max n



 Dcn Dsn Den Drn þ  wC ¼ 0: 2 2 2 2

ð67Þ

This parameter represents only a fraction of the strain energy. However, it does not account for the effect of mean stress. An alternative energy condition was proposed by Chu [71] who combined maximum normal and shear stress with the corresponding strain amplitudes, thus max F ¼ W ¼ max ð2smax Den Þ  wC ¼ 0: n Dcn þ 2rn

ð68Þ

n

We shall not provide the review of the critical plane models proposed by numerous authors in order to simulate fatigue damage and crack growth. The comprehensive review was recently presented by Karolczuk and Macha [72]. In this section, the local critical plane models will be extended by introducing non-local failure criteria applicable to both regular and singular stress regimes and also for monotonic and cyclic loading cases. Consider an arbitrary physical plane D and the local coordinate system (n1, n2, n3), Fig. 23. In the global coordinate system (x1, x2, x3) the origin of the local system is specified by the position vector x0(x01, x02, x03) and the unit normal vector n(n1, n2, n3) specifies the plane orientation, where ni = cos(n3, xi). The resulting shear stress and strain in the plane D are exqffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi pressed as follows: sn ¼ s2n1 þ s2n2 , cn ¼ c2n1 þ c2n2 . Assume that crack initiation and propagation process depends on the 0.4 0.3 0.2 0.1 0

x 10 1.5 1 0.5 0 −0.5

−0.1 −1 −0.2 −0.3

0.2

0.1

0

(a)

−0.1

−0.4 −0.2

−1.5 0.1

0.05

0

−0.05

−0.1

−0.3

−0.2

−0.1

0

0.1

0.2

0.3

(b)

Fig. 21. 3D elliptical smooth crack growth based on the MK-criterion: (a) Pure tensile loading (b = p/2), (b) shear loading (x = 0, b?0).

1800

K.P. Mróz, Z. Mróz / Engineering Fracture Mechanics 77 (2010) 1781–1807

z

n MC r

θ

y

φ

x Fig. 22. Orientation of the critical plane MC.

x3

x0

x2

x1

Fig. 23. The 3D system in R-criterion, from [73].

contact stress and strain components and also on the damage accumulation on the physical plane. Consider an elliptical condition for rn > 0 and the Coulomb condition for rn 6 0, thus



8   > rn 2



9 >

 2 1=2 þ ssnc  1;

rn > 0; =

sc ðjsn j þ rn tan uÞ  1;

> rn < 0; ;

< rn sn  1 ¼ max F ¼ Rr ; n > rc sc :1

rc

ð69Þ

where rc, sc denote the failure stress of material in tension and shear. For large stress gradients or singular stress regimes such as those occurring at vertices of wedge shaped notches, the non-local stress failure conditions is applied by averaging the failure stress function over an area d0 d0, thus

" F ¼ Rr  1 ¼ max ðn;x0 Þ

Z

1 2

d0

0

d0

Z

#

d0

Rr dn1 dn2  1 ¼ 0:

ð70Þ

0

An alternating, non-local condition can be obtained by averaging normal and shear stress components on the plane D, thus

r n ¼

1 2 d0

Z

d0 0

Z

d0 0

1 rn dn1 dn2 ; sn ¼ 2

Z

d0

0

d0

Z

d0

sn dn1 dn2

ð71Þ

0

and substituting to (70)

F ¼ Rr  1 ¼ max Rr n





r n sn  1 ¼ 0: ; rc sc

ð72Þ

The size parameter d0 representing the size of damage zone can be specified by requiring the non-local conditions (72) to be equivalent to the Griffith condition in the case of tensile crack propagation (43). The extensive application of the non-local failure criterion to monotonically loaded elements witch sharp notches and cracks was discussed by Seweryn and Mróz [33,34]. The application to fatigue crack initiation and growth was discussed in [74,75] by applying the non-local criterion with account for local damage growth. To formulate the non-local criterion in the case of fatigue crack growth, introduce the damage initiation function Rr 0(rn/ r0, sn/s0) which is assumed to have the same form as the failure function Rr(rn/r0, sn/s0), thus

K.P. Mróz, Z. Mróz / Engineering Fracture Mechanics 77 (2010) 1781–1807

r0 s0 ¼ ¼ f: rc sc

1801

ð73Þ

Assume the stress values r0 and s0 to depend on the accumulated damage, so that

r0 ¼ r00 ð1  xn Þq ;

s0 ¼ s00 ð1  xn Þq ;

ð74Þ

where 0 6 xn 6 1 is the damage measure on the physical plane, and q is the material parameter. Consider the domain of damage accumulation bounded by the curves F 0 ¼ 0 and F ¼ 0. The actual stress point lies on the loading surface  n =r0 ; s n =s0 Þ  f ¼ 0. The damage growth can now be expressed as follows: F l ¼ Rr ðr

( dxn ¼

for dRn > 0 and f P f ; for dRn < 0 or f < f ;

Wr dRn ; 0;

ð75Þ

r ds and W is the damage growth function assumed in the form where dRr ¼ @@Rrrn drn þ @R n r @ sn

Wr ðRr Þ ¼ A

f  f 1f

!n

dRr : 1f

ð76Þ

The term in brackets represents the ratios PP0/PCP0 for the stress state corresponding to point P in Fig. 24, and A, n are the material parameters. The accumulated damage is now expressed as follows

xn ¼

Z

Rr

WðrÞ dRr

ð77Þ

0

and the crack initiation condition is satisfied when xn = 1. The crack propagation condition in Mode I due to cyclic loading can now be generated from (76), namely

da ¼C dN

" nþ1  nþ1 # K max  DK th K min  DK th ;  K C  DK th K C  DK th

ð78Þ

where Kmax and Kmin are the maximal and minimal values of the stress intensity factor in consecutive loading cycles, DKth is the threshold value of DK corresponding to the onset of fatigue and KC is the critical value of SIF, C and n are the material parameters. The application of the non-local critical plane approach to modelling crack initiation stress and orientation of crack growth was presented in papers by Seweryn et al. [27,74], and the application to fatigue crack growth simulation in [43,75]. The evolution of damage zones at the crack tip can then be specified for any loading history, and the orientation of crack growth can be predicted. The polar diagram of damage distribution in a tubular element subjected to combined bending and torsion with varying stresses rz(t) = ra sin(x t) + rm, szh(t) = sa sin(xt  d) + sm is shown in Fig. 25, cf. [34]. When the maximal tensile conditions is used in (69), then the Novozhilov [76] condition is obtained. Fig. 26 presents the angle of the crack pffiffiffi propagation hpr for the various main crack orientation angle in the uniaxial tension (k = 0) and the various values of s ¼ 3rc =sc between the tension and shear critical failure stress. The most interesting are the differences of the angle of the crack propagation for the same value of w but different s values. In fact, for s = 0 the mode I crack grows at vary-

σ τ Rσ n n = σc τc

τn ∂Rσ ∂σ

τc τ

σf

P P

σu Rσ

σn τn = σ τ

Pc

σ

σc σn

Fig. 24. Damage initiation F0 = 0, stress failure, F = 0 and loading curve Fl = 0 in the plane rn, sn.

1802

K.P. Mróz, Z. Mróz / Engineering Fracture Mechanics 77 (2010) 1781–1807

σc τc = R fσ =

σm σa = k τm τa = k δ= f = σ a = τ a nσ =

ω nω θ

θ=

Rd

θ

θ= θ= k=-1

k=-0.5 k=0-1

Fig. 25. Polar diagram of damage distribution for proportional bending and torsion, after [34].

θpr

ψ Fig. 26. The angle of crack propagation, based on [33] for different values of the ratio s ¼

pffiffiffi r 3 scc and KII/KI = tan w.

ing orientations depending on the mode mixity factor. Similarly, for s = 1, the mode II crack grows. Between the lines representing growth orientation in modes I and II shown in Fig. 26 there are discontinuity points represented by vertical segments of transition from mode I to mode II (cf. paths for s = 1.0, 1.5, 2.0) and bifurcation points corresponding to continuous transition of orientation between these modes (cf. paths for s = 2.5, 2.75). 5. Fatigue crack growth In fact, the mixed mode fracture models are unable to predict accurately the crack growth rate or the initiation angle, especially for mode II dominant loading (low value of a, k = 0, Fig. 1 ). Usually, this prediction is difficult for several initial cycles for which mode II condition dominates (DKII > DKI). Similar effect is observed also for monotonic loading cases. The other important factor is prediction of the fatigue crack growth rate. The history of rate of crack growth modeling starts from the Paris law [77,78] formulated in the form

1803

K.P. Mróz, Z. Mróz / Engineering Fracture Mechanics 77 (2010) 1781–1807

da ¼ CðDKÞn ; dN

ð79Þ

where C and n are the material quasi-parameters, and DK = Kmax  Kmin. This equation predicts the fatigue crack growth in one cycle for the case of small scale yielding in terms of the amplitude DK in mode I. In actuality, the rate of crack growth depends on many factors, such as mean and maximal stress, crack closure effect, and mode mixity. There have been numerous extensions of (79) to account for other effects. Tanaka [79] in 1974 introduced the concept of the effective stress intensity factor DKeff for mixed mode conditions, thus

da ¼ CðDK eff Þn ; dN

ð80Þ

DK eff ¼ ðDK I þ 8DK II Þ0:25 :

ð81Þ

where

Another form of the DKeff resulting from the MTS-criterion was proposed by Yan et al.[80], namely

DK eff ¼ 0:5 cos

hpr ðDK I ð1 þ cos hpr Þ  3DK II sin hpr Þ; 2

ð82Þ

where hpr is obtained from MTS-criterion. There are also other parameters used to correlate fatigue crack growth under mixed mode loading. Sih and Barthelemy [81] used the strain energy density factors DS replacing DK in the Paris type equation. They compared the predicted crack path used S parameter with experimental data [82] for specimen made of Ti–6Al–4V with inclined cracks cf. Table 2. However, these predictions are unsatisfactory, especially for a = 30°. The other parameter used frequently to predict the crack growth rate is the crack tip opening displacement (CTOD). It is assumed that the crack growth in the loading cycle is proportional to CTOD [83]. This parameter is often combined with other factors affecting the crack growth, such as the crack closure parameter [84]. In 1989 Li [85] developed the method of simulation of crack growth in terms of crack tip displacement (CTD) treated as a vector composed of CTOD and CTSD associated with mode I and II, thus allowing for specification of the effective CTD. Usually, the fatigue crack growth rate conventionally is correlated with SIFs by Paris type equation. However, experimental results show that specimen geometry seems to play an important influence on fatigue behaviour. Picard et al. [86] and Tong et al. [87] showed that fatigue crack growth rates differ significantly under the same testing conditions for different specimens, CN (corner notch) and CT (compact tension). Analysis based on the DK alone does not provide satisfactory results. Therefore, the T-stress could also be used to explain this phenomenon with satisfactory results [88]. The application of MK-fracture criterion to the case of cyclic loading was presented in 2008 by Mróz [27]. This criterion predicts the crack growth orientation depending on the load level. For the stress cycle with stress level varying between rmin and rmax, the crack growth initiation stress rpr was introduced with the assumption that rpr = rop, where rop is the crack opening stress associated with crack closure effect [89]. Then it is assumed that rpr = rmin if rmin > rop when T-stress effect is associated. The following relation specifying the crack growth per one cycle was proposed in [27]:

da ¼ C  DðSD ðr pr Þ  rpr ðSconst: ÞÞn ; V dN

ð83Þ

where rrp is the length of decohesion zone along the crack growth direction and C, n are the fatigue parameters obtained from uniaxial tests (a = 90°). The relation (83) can be also written in the following form

C  ðSD  Dr pr þ DSD  r pr Þn ;

ð84Þ

where DSD = SD max  SD min, Drpr = rD pr max  rD pr min and SD ; rpr are the mean values of SD and rpr on one cycle, respectively. Let us note that the crack growth rate now depends on both deviatoric and volumetric energy amplitudes. In view of (25), (26) and (31) the growth rule (83) can be presented in the following form expressed in terms of the singular stress field (1) (T = 0):

0 1n DK 2I ððm2 þ m þ 1ÞðC þ AÞ2  3ðAC  F 2 ÞÞþ da B C ¼ C 2 @ þDK 2II ððm2 þ m þ 1ÞðB  DÞ2 þ 3ðBD  G2 ÞÞþ A ; dN þDK I DK II ð2ðm2 þ m þ 1ÞðC þ AÞðB  DÞ  3ðCB  ADÞÞ

ð85Þ

Table 2 Characterization of specimen A and B. Initial crack length a0 (mm)

Specimen A: a0 = 7.11 mm

Specimen B: a0 = 6.73 mm

a (°) rmin (MPa) rmax (MPa)

a = 30

a = 43 17.24 172.38

20.69 206.85

1804

K.P. Mróz, Z. Mróz / Engineering Fracture Mechanics 77 (2010) 1781–1807

Table 3 Mixed mode fatigue results for specimen A. Prediction [81]

Prediction Eq. (83)

Exp. results According to [82]

N

da/dN

N

da/dN

Cycle numb.

[10–7 m]

Cycle numb.

7.05 7.01 8.52 8.53 10.1 10 11 13 16 16

1390 2611 3340 4260 4771 5208 5682 6141 6648 7111

16,000

22,000

25,000

Prediction Eq. (87)

N

da/dN

N

da/dN

[10–7 m]

Cycle numb.

6.77 6.25 7.4 8.53 9.78 10.86 12.14 13.71 15.78 18.38

1390 2430 3110 4000 4550 4900 5400 5900 6400 7000

[10–7 m]

Cycle numb.

[10–7 m]

6.54 7.59 7.35 9.1 8 11.1 12 12.2 16 18

3583 5416 6352 7365 7872 8279 8695 9075 9471 9813

2.64 4.15 5.77 7.74 9.87 16.67 13.83 16.55 20.19 24.89

Table 4 Mixed mode fatigue results for specimen B. Prediction [81]

Prediction Eq. (83)

Exp. results According to [82]

Prediction Eq. (87)

N

da/dN

N

da/dN

N

da/dN

N

da/dN

Cycle numb.

[10–7 m]

Cycle numb.

[10–7 m]

Cycle numb.

[10–7 m]

Cycle numb.

[10–7 m]

0.64 2.84 3.76 3.82 5.43 6.27 5.43 7.59 8.46 10

2954 4331 6523 8106 9804 10,569 11,268 12,988 13,605 14,319

2.91 3.12 3.6 4.17 4.91 5.62 6.15 7.29 8.58 9.54

4660 6100 8250 10,000 11,490 12,080 13,000 14,660 15,310 15,960

1.85 3.12 3.58 3.71 5.77 8.3 3.91 7.53 7.84 10.92

3546 5732 8859 10,873 12,862 13,711 14,462 16,192 16,780 17,436

1.64 1.96 2.53 3.27 4.19 5.06 5.72 7.25 9.0 10.36

10,000 12,080

19,100

where





; D ¼ sin 2h 2 þ cos 2h cos 3h ; A ¼ cos 2h 1 þ sin 2h sin 3h 2 2  n 1þm 1 h h 3h h h 3h C2 ¼ C ; B ¼ sin 2 cos 2 cos 2 ; F ¼ sin 2 cos 2 cos 2 ; 3E 2p



C ¼ cos 2h 1  sin 2h sin 3h ; G ¼ cos 2h 1  sin 2h sin 3h 2 2

ð86Þ

and m* = 0 for plane stress, m* = m for plane strain cases. When T-stress is included, the crack growth rule is expressed in terms of both stress intensity amplitudes and their mean values over loading cycle. Following the rule (69), the simplified equation to determine the crack increment in mixed mode per cycle can be proposed

2sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi3n1  2  2 da rn sn 5 ¼ C 1 D4 þ ; dN rf sf

rn > 0;

ð87Þ

pffiffiffi where C1, n1 are the fatigue constants for the uniaxial test (a = 90°), rf is the fatigue limit and rf =sf ¼ 3. The comparison of results is presented in Figs. 27 and 28, also in Tables 3 and 4, where the last two columns present the results according to Eq. (87), the second the results according to MK-criterion and Eq. (83), and the third the experimental results of Pustejowsky [82] conducted for titanium alloy specimens. The comparison with predictions of Sih model [81] is also presented in tables in the first two columns. It is seen that the relation (83) provides much more accurate prediction of the growth rate. It should be noted that T-stress should be retained in the analysis of stress state near the crack tip and in prediction of rate and orientation of crack growth.

1805

K.P. Mróz, Z. Mróz / Engineering Fracture Mechanics 77 (2010) 1781–1807

3110

Ti-6Al-4V

4900

5900 7000

3340 5208 6141 7111

Prediction Exp.results

0

R=

= 7 11 mm

σ min 20.6 = = 0.1 σ max 206

[m] Fig. 27. Mixed mode fatigue trajectory for specimen A.

8250

Ti-6Al-4V 6523

12080

15960

10569 14319

Prediction Exp. results

R=

0

σ min 17.2 = = 0.1 σ max 172

= 6 73 mm

[m] Fig. 28. Mixed mode fatigue trajectory for specimen B.

6. Concluding remarks The present paper provides a synthetic review of different criteria of crack growth for 2D and 3D cases. The criteria are based on the elastic singular stress distribution and the regular T-stress component. The stress criteria, energy concepts and critical plane approaches are discussed and their prediction compared for several specific cases. The objective of this paper is

1806

K.P. Mróz, Z. Mróz / Engineering Fracture Mechanics 77 (2010) 1781–1807

to indicate also the major topics requiring further study, such as specification of size parameters at the crack front, generation of rough crack patterns, and crack growth in 3D stress condition. The non-local character of the criteria is incorporated by specifying a critical distance segment from the crack tip on which the averaged stresses and strain states or their local values at the end segment point are determined and used in formulating the damage and crack growth criteria. Such critical segment length can be interpreted as representing the fracture or localized damage zone. The forms and sizes of plastic and distributed damages zones are illustrated in Fig. 2. Considering only elastic singular stress field with T-stress, it can be assumed that the deviatoric and volumetric portions of the specific strain energy may represent the forms and sizes of plastic and distributed damage zones. The crack growth criterion can then be formulated and the critical distance rc specified in terms of the stress intensity factors and the critical stress value rc representing the critical volumetric energy level. The critical plane approaches require also specification of the averaging size parameter d0 which can be expressed by considering non-local energy release for a crack growth through the finite distance d0. For modes II and III the interaction of rough crack interfaces with account for frictional slip and dilatancy effects is of fundamental importance. For compressive normal stress the crack interface slip then induces shielding of KII due to friction and growth of KI due to dilatancy. The generation of rough crack pattern in mode III can be modeled by assuming the facets of mode I developing periodically at the existing crack edge and creating the crack microsurface with facets of mode II and III. The rough crack pattern violates the continuity condition at the existing crack edge on the local scale of facets size but preserves the continuity for a crack macrosurface. For 3D-cracks both smooth and rough crack surfaces can be generated at different locations of the initial crack boundary. The predictive models of rough cracks are not yet available and require further development. The growth of 3D cracks is still not well investigated as more complex crack patterns can grow at the existing crack edge. Both experimental observations and analytical models are limited. It is expected that in a near future the substantial progress in this area will be achieved. Acknowledgments The results presented in this paper have been obtained within the project KomCerMet (Contract No. POIG.01.03.01-00013/08 with the Polish Ministry of Science and Higher Education) in the framework of the Operational Programme Innovative Economy 2007–2013. The authors wish to thank reviewers for their critical remarks and improvement suggestions. References [1] [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23] [24] [25] [26] [27] [28] [29] [30] [31]

Eftis J, Subramonian N. The inclined crack under biaxial load. Engng Fract Mech 1978;10:43–67. Erdogan F, Sih GC. On the crack extension in plates under plane loading and transverse shear. J Basic Engng 1963;85:519–27. Sih GC. Some basic problems in fracture mechanics and new concepts. Engng Fract Mech 1973;5:365–77. Hussain MA, Pu SL, Underwood JH. Strain energy release rate for a crack under combined mode I and mode II. Fract Anal ASTM STP 1974;560:2–28. Wu CH. Fracture under combined loads by maximum-energy-release-rate-criterion. ASME J Appl Mech 1978;45:553–8. Goldstein RV, Salganik RL. Brittle fracture of solids with arbitrary cracks. Int J Fract 1974;10:507–23. Amestoy M, Leblond JB. Curvilinear cracks in plane situation. Proc 6th Eur Conf Fract 1986;2:1377–401. Leblond JB. Crack paths in plan situation – I. General form of the expansion of the stress intensity factors. Int J Solids Struct 1989;25:1311–25. Amestoy M, Leblond JB. Crack path in plane situation – II. Detailed form of the expansion of the stress intensity factors. Int J Solids Struct 1992;29:465–501. Strifors HC. A generalized force measure of conditions at crack tips. Int J Solids Struct 1973;10:1389–404. Hellen TK, Blackburn WS. The calculation of stress intensity factors for combined tensile and shear loading. Int J Fract Mech 1975;11:605–17. Chang KJ. On the maximum strain criterion. A new approach to the angled crack problem. Engng Fract Mech 1981;14:107–24. Papadopoulos GA. The stationary value oh the third stress invariant as a local fracture parameters (Det-criterion). Engng Fract Mech 1987;27:643–52. Koo JM, Choy YS. A new mixed mode fracture criterion: maximum tangential strain energy density criterion. Engng Fract Mech 1991;39:443–9. Kong XM, Schluter N, Dahl W. Effect of triaxial stress on mixed-mode fracture. Engng Fract Mech 1995;52:379–88. Ewing PD, Williams JG. Fracture under complex stress – the angled crack problem. Int J Fract Mech 1972;8:441–6. Swedlow JL. Criteria for growth of the angled crack. In: Cracks and fracture. ASTM STP 601; 1976. p. 506–21. Maiti SK, Smith RA. Comparison of the criteria for mixed mode brittle fracture based on the preinstability stress–strain field, part I: slit and elliptical cracks under uniaxial tensile loading. Int J Fract 1983;23:281–95. Maiti SK, Smith RA. Criteria for brittle fracture in biaxial tension. Engng Fract Mech 1984;19:793–804. Wong AK. On the application of the strain energy density theory in predicting crack initiation and angle of growth. Engng Fract Mech 1987;27:157–70. Wu X, Li X. Analysis and modification of fracture criteria for mixed-mode crack. Engng Fract Mech 1989;34:55–64. Theocaris PS, Andrianopoulos NP. The Mises elastic–plastic boundary as the core region in fracture criteria. Engng Fract Mech 1982;16:425–32. Wasiluk B, Golos K. Prediction of crack growth direction under plane stress for mixed-mode I and II loading. Fatigue Fract Engng Mater Struct 2000;23:381–6. Khan SMA, Khraisheh MK. A new criterion for mixed mode fracture initiation based on the crack tip plastic core region. Int J Plast 2004;20:55–84. Yan X, Zhang Z, Du S. Mixed-mode fracture criteria for the materials with the different yield strength in tension and compression. Engng Fract Mech 1992;42:109–16. Williams JG. Fracture mechanics of polymers. Chichester: Ellis Horwood; 1984. Mróz KP. Fatigue cracks growth in the bimaterial system. The mathematical model and numerical solution. PhD thesis IPPT PAN (IFTR PAS), Warsaw; 2008. Ayatollahi MR, Aliha MRM, Hassani MM. Mixed mode brittle fracture in PMMA – an experimental study using SCB specimen. Mater Sci Engng 2006;A417:348–56. Cotterell B, Rice JR. Slightly curved or kinked cracks. Int J Fract 1980;16:155–69. Larsson SG, Carlsson AJ. Influence of non-singular stress term and specimen geometry on small-scale yielding at crack tips in elastic–plastic material. J Mech Phys Solids 1973;21:263–78. Haefele PM, Lee JD. Combination of finite element analysis and analytical crack tip solution for mixed mode fracture. Engng Fract Mech 1995;50:849–68.

K.P. Mróz, Z. Mróz / Engineering Fracture Mechanics 77 (2010) 1781–1807

1807

[32] Kosai M, Kobayashi AS, Ramulu M. Tear straps in airplane fuselage. In: Alturi SN, Harris CE, Hoggard A, Miller N, Sampath SG, editors. Durability of metal aircraft structures. Atlanta: Atlanta Technology Publications; 1993. p. 443–57. [33] Seweryn A, Mróz Z. A non-local stress failure conditions for structural elements under multiaxial loading. Engng Fract Mech 1995;51:955–73. [34] Seweryn A, Mróz Z. On the criterion of damage evolution for variable multiaxial stress state. Int J Solids Struct 1998;35:1589–616. [35] Seweryn A. A non-local stress and strain energy release rate mixed mode fracture initiation and propagation criterion. Engng Fract Mech 1998;59:737–60. [36] Shlyannikov VN. Strain energy density and fracture process zone. Strength Mater 1995;27:569–79. [37] Shlyannikov VN. Modelling of crack growth by fracture damage zone. Theor Appl Fract Mech 1996;25:187–201,. [38] Seweryn A. A non-local stress and strain energy release rate mixed mode fracture initiation and propagation criteria. Engng Fract Mech 1998;59:737–60. [39] Taylor D. The theory of critical distance. Engng Fract Mech 2008;75:1696–705. [40] Carpinteri A, Cornetti P, Pugno N, Sapora A, Taylor D. A finite fracture mechanics approach to structures with sharp V-notches. Engng Fract Mech 2008;75:1736–52. [41] Leguillon D. Strength or toughness? A criterion for crack onset at a notch. Eur J Mech A/Solids 2002;21:61–72. [42] Knauss WG. An observation of crack propagation in anti-plane shear. Int J Fract 1970;1:183–5. [43] Tschegg EK, Suresh S. Mode III fracture of 4340 steel: effects of tempering temperature and fracture surface interference. Metall Trans 1988;19A:3035–43. [44] Dyskin AV, Salganik RL. Model of dilatancy of brittle materials with cracks under compression. Mech Solids 1987;22:165–73. [45] Dyskin AV, Sahouryeh E, Jewell RJ, Joer H, Ustinow KB. Influence of shape and locations of initial 3-D cracks on their growth in uniaxial compression. Engng Fract Mech 2003;70:2115–36. [46] Adams M, Sines G. Crack extension from flaws in brittle material subjected to compression. Tectonophysics 1978;49:97–118. [47] Dyskin AV, Germanovich LN, Ustinow KB. A 3-D model of wing crack growth and interaction. Engng Fract Mech 1999;63:81–110. [48] Pook LP. Linear elastic fracture mechanics for engineers: theory and application. Southampton: WIT; 2000. [49] Tschegg EK. Sliding mode crack closure and mode III fatigue crack growth in mild steel. Acta Metall 1983;31:1323–30. [50] Ballarini R, Plesha ME. The effects of crack surface friction and roughness on crack tip stress fields. Int J Fract 1987;34:195–207. [51] Evans AG, Hutchinson JW. Effects of non-planarity on the mixed mode fracture resistance of biomaterial interfaces. Acta Metall 1989;37:909–16. [52] Gross TS, Mendelsohn DA. Mode I stress intensity factors induced by fracture surface roughness under pure mode III loading: application to the effect of loading modes on stress corrosion crack growth. Metall Trans 1989;20A:1989–99. [53] Mendelsohn DA, Gross TS, Zhang Y. Fracture surface interference in shear – I. A model based an experimental surface characterization. Acta Metall Mater 1995;43:893–900. [54] Gross TS, Zhang Y, Watt DW. Fracture surface interference in shear – II. Experimental measurements of crack tip displacement field under mode II loading in 7075-T6Al. Acta Metall Mater 1995;43:901–6. [55] Mróz Z, Seweryn A. Non-local failure and damage evaluation rule: application to a dilatation crack model. J Phys IV France 1998;8:257–68. [56] Pook LP. Crack paths. WIT Press; 2002. [57] Benthem JP. State of stress at the vertex of a quarter-infinite crack in a half-space. Int J Solids Struct 1977;13:479–92. [58] Bazant ZP, Estenssoro LF. Surface singularity and crack propagation. Int J Solids Struct 1979;15:405–26. [59] Folias ES. Method of solution of a class of three-dimensional elastostatic problems under mode I loading. Int J Fract 1980;16:335–48. [60] Panasyuk VV, Andrejkiv AE, Stadnik MM. Three-dimensional static crack problem solution (a review). Engng Fract Mech 1981;14:245–60. [61] Kassir MK, Sih GC. Three dimensional crack problems. A new selection of crack problems in three dimensional elasticity. In: Sih GC, editor. Mechanics of fracture II. Noordhoff; 1974. [62] Kassir MK, Sih GC. Three dimensional stress distribution around an elliptical crack under arbitrary loadings. J Appl Mech 1966;33 [Trans ASME;88:601–611]. [63] Hartranft RJ, Sih GC. Stress singularity for a crack with an arbitrarily curved front. Engng Fract Mech 1977;9:705–18. [64] Hartranft RJ, Sih GC. The use of eigenfunction expansion in the general solution of three-dimensional crack problem. J Math Mech 1969;19:123–38. [65] Wang YY, Parks DM. Evaluation of elastic T-stress in surface-cracked plates using the line-spring method. Int J Fract 1992;56:25–40. [66] Nakamura T, Parks DM. Determination of elastic T-stress along three-dimensional crack fronts using an interaction integral. Int J Solids Struct 1992;29:1597–611. [67] Sih GC, Cha CK. A fracture criterion for three-dimensional crack problems. Engng Fract Mech 1974;6:699–723. [68] Zacharopoulos DA. Stability analysis of crack path using the strain energy density theory. Theor Appl Fract Mech 2004;41:327–37. [69] Schollmann M, Kullmer G, Fulland M, Richard HA. A new criterion for 3D crack growth under mixed-mode (I + II + III) loading. In: de Freitas MM, editor. Proceedings of the 6th international conference on biaxial/multiaxial fatigue and fracture II. Lisboa: Instituto Superior Tecnico; 2001. p. 589–96. [70] Glinka G, Wang G, Plumtree A. Mean stress effects in multiaxial fatigue. Fatigue Fract Engng Mater Struct 1995;18:755–64. [71] Chu CC. Fatigue damage calculation using the critical plane approach. J Engng Mater Technol 1995;117:41–9. [72] Karolczuk A, Macha E. A review of critical plane orientations in multiaxial fatigue failure criteria of metallic materials. Int J Fract 2005;134:267–304. [73] Mróz Z, Seweryn A. Multiaxial damage and fatigue conditions. In: Basista M, Nowacki WK, editors. Modeling of damage and fracture processes in engineering materials. Warszawa: Trends in mechanics of materials; 1999. p. 117–80. [74] Seweryn A, Poskrobko S, Mróz Z. Brittle fracture in plane elements with sharp notches under mixed-mode loading. J Engng Mech 1997;123:535–43. [75] Mróz Z, Seweryn A, Tomczyk A. Fatigue crack growth prediction accounting for the damage zone. Fatigue Fract Engng Mater Struct 2005;28:61–71. [76] Novozhilov VV. On necessary and sufficient criterion of brittle fracture. Prikl Mater Mech 1969;33:212–6 [in Russian]. [77] Paris P. A note on the variable affecting the rate of crack growth due to cyclic loading. The Boeing Company, Doc. No. D-17867; 1957. p. 12. [78] Paris P, Gomez MP, Anderson WE. A rational analytical theory of fatigue. Trend Engng 1961;13:9–14. [79] Tanaka K. Fatigue crack propagation from a crack inclined to the cyclic tensile axis. Engng Fract Mech 1974;6:493–507. [80] Yan X, Du S, Zhang Z. Mixed mode fatigue crack growth prediction in biaxially stretched sheets. Engng Fract Mech 1992;43:471–5. [81] Sih GC, Barthelemy BM. Mixed mode fatigue crack growth predictions. Engng Fract Mech 1980;13:439–51. [82] Pustejovsky MA. Fatigue crack propagation in titanium under general in-plane loading – I: Experiments. Engng Fract Mech 1979;11:9–15. [83] McEvily AJ. Current aspects of fatigue. Metal Sci 1977;11:274–84. [84] Shademan S, Soboyejo ABO, Knott JF, Soboyejo WO. A physically-based model for the prediction of long fatigue crack growth in Ti–6Al–4V. Mater Sci Engng 2001;A315:1–10. [85] Li C. Vector CTD criterion applied to mixed mode fatigue crack growth. Fatigue Fract Engng Mater Struct 1989;12:59–65. [86] Pickard AC, Brown CW, Hicks MA. The development of advanced testing and analysis techniques applied to fracture mechanics lifting of gas turbine components. In: Proceedings of the international conference on advanced in life prediction methods; 1983. p. 173–8. [87] Tong J, Byrne J, Hall R, Allabadi MH. A comparison of corner notched and compact tension specimens for high temperature fatigue testing. Engng Against Fatigue 1999:583–90. [88] Zhao LG, Tong J, Byrne J. Stress intensity factor K and the elastic T-stress for corner cracks. Int J Fract 2001;109:209–25. [89] Elber W. Fatigue crack closure under cyclic tension. Engng Fract Mech 1970;2:37–45.