You are on page 1of 7

Fracture Type Transition under Mixed Mode I/II Quasi-static and Fatigue

Loading Conditions

Shu Liu, Yuh J. Chao and Rishiraj R. Gaddam


Department of Mechanical Engineering, University of South Carolina
Columbia, SC 29208
Tel: (803)777-5869, Fax: (803)777-0106, e-mail: chao@sc.edu

ABSTRACT

Mixed mode I/II fracture criteria based on monotonic loading conditions [1-6] were applied to the fatigue loading conditions.
Preliminary experiments using mild steel and Al2024-T3 under mixed mode I/II monotonic and proportional fatigue loads were
conducted. Fracture types and crack growth angles were examined. It is shown that the transition of fracture type from tensile
to shear for ductile materials, as observed in monotonic loadings , was not generally observed under proportional fatigue
loadings . To promote the shear type fracture, non-proportional cyclic loading might be needed, i.e. static mode I loading
superimposed on the mixed mode loading to open the crack face and overcome the friction effects (closure effects) of fatigue
crack.

1 INTRODUCTION

Many structural failures take place from cracks subjected to mixed mode loadings. A characteristic of mixed mode I/II quasi-
static and fatigue cracks is that they usually propagate in a non-self similar manner. Therefore, under mixed mode static
(fatigue) loading conditions, not only the onset fracture load (fatigue load and crack growth rate) is of importance, but also the
crack growth direction.

Under monotonic loading of ductile materials, typical experimental and theoretical results for crack initiation direction are shown
in Fig.1[1], which shows two fundamentally distinct types of crack growth. The first is the tensile dominated fracture that can be
described by the MHSC (maximum hoop stress criterion) [5], where crack turns to the direction perpendicular to the maximum
hoop stress around the crack tip. The second is the shear dominated fracture described by the MSSC (maximum shear stress
criterion) [6], where crack turns to the direction parallel to the maximum shear stress.

In papers [1-4], the authors extended the basic concept of MHSC and MSSC to incorporate a material strength ratio into a new
unified failure criterion for prediction of fracture type transition from tensile to shear. Similarly, under cyclic mixed mode
loadings of ductile materials, even though specimens of very different geometry were used, both tensile and shear stress
controlled stable fatigue crack growth could happen [7-9]. The major results in fatigue crack propagation under proportional
mixed mode I+II loading are summarized by Bold et al. [7]. They found that crack mostly propagates in tensile type. And, under
large shear load, after shear type crack initiates to about 1 mm in length, either crack is arrested or tensile type fracture for
large amplitude of applied loads follows . Pook [10] stated that mixed mode I/II fatigue crack generally propagates in mode I
dominated fracture and only under special conditions will it undergo shear fracture. The special conditions [9] could be that (1)
shear bands develop at the initial crack tip; (2) the net section stress approaches or exceeds the shear yield strength; (3) a
static compressive stress is applied; (4) under non-proportional loadings ; and (5) in the presence of a geometric constraint.

Various criteria for crack growth direction under mixed mode fatigue loadings have been proposed [10]. Among the various
criteria, the MHSC and MSSC from static mixed mode loading conditions are most widely applied for fatigue loading. Note that
the two criteria do not distinguish between static and cyclic loadings, although the micromechanics of crack initiation and
growth may be significantly different for the two loading conditions. Several authors have suggested some justifications for this
[11].

Since the first study of mixed mode I/II fatigue crack growth by Iida and Kobayashi [12], much progress has been made in this
area. As already indicated, there are two major aspects in mixed mode fatigue crack growth: crack growth direction and crack
growth rate. In addition, it is found that crack propagation rates depend on the fracture type, i.e., crack propagation rates in
shear are up to ten times higher than that in tensile type [12]. It is therefore important to distinguish the crack propagation
types in fatigue crack growth.
Experimental evidence suggests that crack growth type depends on the material, load magnitude, initial crack tip condition,
ratio among different mode loads and mean stress. Studies dealing with the contributions of these effects are very limited, and
the question of when the fracture type changes from tensile to shear is far from clear.

In the current paper, we attempt to use the MHSC and MSSC criteria plus the transition criterion proposed in [1-4] to predict the
fracture type transition in mixed mode I/II fatigue crack initiation and propagation. The present investigation used Arcan
specimens [13], made from two elastic-plastic materials, mild steel and aluminum 2024-T3 alloy, subjected to mixed tensile and
shearing proportional cyclic loads through specially designed apparatus (for comparison, static mixed mode loading were also
applied). Fracture types, their associated transition and crack growth directions were studied and compared for static and
fatigue loading conditions. Note that since shear type fracture might happen when the net section stress approaches or
exceeds the shear yield strength [9], high maximum loads causing net section yielding of specimens were also introduced for
fatigue loading.

2 PROBLEM DEFINITIONS

We consider a plate, containing a crack as shown in Fig.2, under far field loads such that mixed-mode I/II stress fields exist
near the crack tip. Within the context of linear elasticity, the crack tip hoop stress and shear stress, employing a polar
coordinate system (r,θ), can be written as
1 θ θ 3 (1)
σ θθ = cos( )( K I c o s2 − K II sin θ )
2π r 2 2 2
τ rθ
=
2
1
2π r
cos(
θ
2
) K sin θ + K
I
[
II
( 3 cos θ − 1 ) ] (2)

where KI and KII are symmetric (mode I) and skew-symmetric (mode II) components of the stress intensity factors that are
functions of the magnitude of the loads, loading configurations, and geometry of the plate.

For convenience the equivalent crack angle, defined as


KI (3)
β eq = tan−1 ( )
K II
is used for quantifying the relative amount of mode I to mode II loading.

Fracture is assumed to be governed by the competition of the maximum of the two stresses in (1) and (2) relative to their
corresponding critical stress. The type of fracture is controlled by (τmax/σmax) > (τc /σc) for a tensile type of fracture (MHSC) and
(τmax/σmax) > (τc/σc) for a shear type of fracture (MSSC) [1-3].

The crack deviation angles θ* (θ**) from MHSC (MSSC) are


 D − D 2
+ 8  (4)
θ *
= 2 tan −1 
 4 
 
 D  (5)
θ **
= 2 tan −1
d + 
 3 
where the mode mixity parameters D, d and ω are given by:
K I
D =
K II
4π  D 2
d = 2 cos  ω +
7
 +
 3  9 6
1  D 3 + 9D 
ω = cos −1
 
 27 (D + 7 / 6 )
3/ 2
3 2

Note that KI and KII are calculated using the fracturing loads from experiments and the following equations for the standard
Arcan specimen geometry [1].
P (6)
K I = sin ϕ ⋅ π a f I ( a / b )
bt
P (7)
K II = cos ϕ ⋅ π a f II ( a / b )
bt
where

f I (a / b) = 112 . (a / b) + 10.55(a / b)2 − 2127


. − 0231 . (a / b)3 + 3039
. (a / b)4
1.1 2 2 − 0.5 6 1( a / b ) + 0.0 8 5( a / b ) 2 + 0.1 8 0 ( a / b ) 3
f II ( a / b ) =
[1 − ( a / b )]1 / 2
where a is the crack length, b is the total width of the Arcan specimen, P is the far field tensile load, t is the thickness of the
specimen, and ϕ is the loading angle. For cyclic loading KI and KII are replaced by ?KI and ?KII , respectively.

3 EXPERIMENTAL PROCEDURE

3.1 Materials and material properties

Both mild steel and 2024-T3 aluminum alloy were used in the experiments. They were delivered in 2.0 mm thick rolled sheets.
Owing to the rolling process, the specimen sheet exhibited microstructural anisotropy. This was taken into account by taking
out the specimens in different orientation to the rolling direction, LT- or TL-orientation in accordance with the material test
standard ASTM E399. The engineering stress-strain curves of the two materials are shown in Fig.3. Both materials exhibit
typical elastic-plastic mechanical behavior, and the mild steel exhibits lower yield strength than that of 2024-T3 aluminum alloy.
(Note that the stress-strain curves in Fig.3 from TL or LT are about the same.)

3.2 Mixed mode I/II experiments

Both materials were made into the Arcan specimens. A single-edge notch is cut on the Arcan specimen by a saw blade.
Fatigue precracking under Mode I loading conditions is then performed and provides the specimen with a sharp fatigue
precrack which initiates from the notch. The residual plasticity from the procedure of precracking is assumed to be small and
neglected. Specimens had the initial fatigue crack length a =4.0~10.0 mm and width w = 38.1 mm. The geometry and
dimensions of the Arcan specimen and the fixture are shown in Figs.4 (a) and (b) [13]. The fixtures were made of 15-5PH
stainless-steel, which are 19 mm thick, with an outside radius 114.5 mm.

The Arcan specimen was loaded with a MTS tensile testing machine under displacement (load) control for monotonic (cyclic)
loading. The opposite sides of the fixture were pulled apart at a pair of grip holes along a radial line. The Arcan specimen fixture
provides in-plane loading of different mode mixities. By changing the loading angle, as shown in Figs. 4, different mode mixities
can be obtained. Only pure mode II or near mode II (ϕ=75 degree loading angle) were adopted in the current tests since shear
type of fracture is expected at these two loading conditions [1-4]. Specifically, for mild steel, one specimen under static loading
of pure mode II, two specimens under static loading at ϕ=75 degree, and one specimen under fatigue loading at ϕ=75 degree
were tested. For Al2024-T3, one specimen under static loading of pure mode II, and two specimens under fatigue loading of
pure mode II, and one specimen under fatigue loading at 75 degree were conducted. During investigation of fatigue crack
propagation, harmonic loading was applied at constant loading ranges. The load frequency is 10Hz, and the loading ratio is
R=0.1.

During the tes ting, digital images of the crack tip region were taken. The crack tip locations were determined by examining the
consecutive images. The amount of crack extension was then measured from the images. The load, displacement and cycle
data were obtained from the machine readout. During the crack propagation, both the images, loads and cycles were recorded.
The cycles-crack extension curves were then established.

4 EXPERIMENTAL RESULTS AND DISCUSSIONS

For mild steel specimens under mixed mode static loading, crack went straight (shear fracture) initially, then it stoped and the
whole specimen warped (Fig.5 (a) and (b)) due to the large out-of-plane distortion of the specimen. For mixed mode fatigue
loading (75 degree loading), initially crack also went almost straight (shear fracture) for about 2 mm; it then paused for about
100,000 cycles having large plastic blunting built-up at the crack tip; and finally it deviated from the original precrack direction
and propagated in tensile type of fracture (see Fig.6).

For Al2024-T3 specimens under mixed mode static loading, crack went straight (shear fracture, see Fig.7(a)) which matches
the previous results in monotonic loading [1-4]. Under mixed mode fatigue loading, in two cases the crack initially went straight
for less than 1 mm; paused for about 100,000 cycles having large plastic blunting built-up at the crack tip; and finally deviated
from the original precrack direction and propagated in tensile type of fracture (see Fig.7(b)). In the other two cases, crack
curved from its original direction and turned into tensile type of fracture.

The experimental results of precrack length, static crack initiation load (P0), cyclic maximum load (Pmax), equivalent loading
angle (β eq ) from Equ.(3), stress intensity factors ratio (KI/KII, and KI, KII are from Equ.(6) and (7)), and crack growth angle from
MHSC (Equ.(4)), MSSC (Equ.(5)) and experiments, respectively, are shown in Table 1 (a) and (b).

When comparing the crack propagation angles with the theoretically expected values from [1-4] for monotonic loading
conditions , it appears that for static (fatigue) loading under mode II or near mode II conditions, the fracture type is shear
(tensile). The discrepancy is quite interesting because it indicates that tensile fracture is the only dominating type in proportional
fatigue loading conditions.

From our test results, it is observed that typically there are four types of static and fatigue crack growth features (see Table.2).
(1) If the fatigue load is lower than the fatigue threshold load, the crack will not grow for several millions cycles. (2) Crack grows
almost straight in the plane of the extension of the precrack with shear type of fracutre (or called coplanar growth). This
appears only occur under static loading conditions, but not under fatigue loading condition. (3) Shear type fracture occurs at
fracture initiation for one or a few millimeters and then crack changes its type to tensile for stable crack growth. This appears to
happen under fatigue loading only. (4) Tensile type fracture occurs as the crack deviates from the direction of the precrack
from the very beginning of the crack propagation. It happens for fatigue loading only.

Therefore, even under pure mode II or close to pure mode II (shearing) loading from our current fatigue testing of mild steel and
Al2024-T3 materials, we found that the final failure of the specimen is always dominated by tensile type. Shear type fracture, if
happens, only occurs at the crack initiation phase.

The question posed is “why cannot the mixed mode fracture criteria under monotonic loading conditions be used to predict the
fracture type under fatigue loading?”. The most possible reason is that after crack initiates in shear type under fatigue loading,
the crack surface remains flat and they scratch each other which causes friction and closure. The friction traces can generally
be seen on the fractured faces from the broken specimens. The consequence from the rubbing of the crack surfces is to
prohibit the crack growth in shear type and makes it easy to grow in tensile (opening) type.

To generate stable shear type fracture under mixed mode fatigue loading, a static mode I loading might be necessary to open
the crack and eliminate the friction of the crack surfaces .

5 CONCLUSIONS AND FUTURE WORKS

The present paper attempts to extend the mixed mode I/II fracture criteria in monotonic loading conditions to the fatigue loading
conditions. It is shown that under proportional mixed mode fatigue loadings, even under pure mode II or near mode II loading,
the final failure of crack is always tensile type. Or, no fracture type transition occurs . In order to promote the shear type
fracture, non-proportional cyclic loading might be necessary, i.e. a superposition of a static mode I loading to the mixed mode
fatigue loading to open the crack face and overcome the friction effects (closure effects) of fatigue crack. To achieve this, an
additional apparatus was designed which can be adapted to the mixed mode arcan loading device (see Fig.8). More detailed
experiments, numerical analysis and discussions will be conducted and reported later.

REFERENCES

[1] Chao, Y.J. and Liu, S., “On the Failure of Cracks under Mixed-Mode Loads”, International Journal of Fracture, Vol.87,
No.3, pp. 201-223, 1997.

[2] Chao, Y.J. and Zhu, X. K., “A Simple Theory for Describing the Transition between Tensile and Shear Mechanisms in
Mode-I, II, III and Mixed-Mode Fracture,” Mixed-Mode Crack Behavior, ASTM STP 1359, 1999.

[3] Chao, Y.J., Liu, S. and Zhu, X.K., “Failure of Materials: Shear Fracture vs. Tensile Fracture,” Proceedings of the Society for
Experimental Mechanics Annual Conference, pp. 51-54, June 7-9, Cincinnati, Ohio, 1999.

[4] Liu, S., Chao, Y.J., and Zhu, X.K., “Tensile-shear transition in mixed mode I/III fracture,” to appear in International Journal
of Solids and Structures, 2004.

[5] Erdorgan, F. and Sih, G.C.,“On the Crack Extension in Plates under Plane Loading and Transverse Shear”, ASME Journal
of Basic Engineering, Vol.85D, pp.519-527, 1963.

[6] Maccagno, T.M. and Knott, J.F., “The mixed modes I/II fracture behavior of lightly tempered HY130 steel at room
temperature”, Engineering Fracture Mechanics, 41, pp. 805-820, 1992.

[7] Plank, R., and Kuhn, G., “Fatigue crack propagation under non-proportional mixed mode loading”, Engineering Fracture
Mechanics, 62, pp. 203-229, 1999.

[8] Bold, P.E, Brown, M.W., and Allen, R.J., “A review of fatigue crack growth in steels under mixed mode I and II loading”,
Fatigue & Fracture of Engineering Materials & Structures, 15, pp.965-977, 1992.
[9] Pook, L.P., “Crack paths”, published by WIT press 2002, Southampton, UK.

[10] Pook, L.P., “Comments on Fatigue Crack Growth under Mixed Modes I and III and Pure Mode III loading”, Multiaxial
Fatigue, ASTM STP 853, American Society for Testing and Materials , Philadelphia, pp. 249-263, 1985.

[11] Qian, J., and Fatemi, A., “Mixed mode fatigue crack growth: a literature survey”, Engineering Fracture Mechanics, 55,
No.6, pp.969-990, 1996.

[12] Iida, S., and Kobayashi, A.S., “Crack propagation trate in 7075-T6 plates under cyclic tensile and transverse shear
loadings”, Transaction of ASME, Serial D.91, pp.764-769, 1969.

[13] Sutton, M.A., Deng, X., Ma, F., Newman, J.C., and James, M., Development and application of a crack tip opening
displacement-based mixed mode fracture criterion, International Journal of Solids and Structures, 37, pp.3591-3618, 2000.

100
σ
80

60
n=15 τx
40

MSSC
n=1 σθ τθ
20
τr
Y θ r σr
r
θ ( degree )

0 θ
0 10 20 30 40 50 60 70 80 90 r
-20 Predicted transition
β eq-12 = 48.62 o
θ
-40 Crack X
MHSC
-60 n=1
-80
n=15
-100
τx

-120

Mode mixity βeq (degree) σ

Fig.1 Transition of the fracture initiation angle; theory and the Fig.2 Mixed mode I/II stress fields near the crack tip
experimental data [1]
500

400
Stress (MPa)

300

200
Mild Steel
100
Al 2024-T6-TL
0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
Strain (m/m)
Fig.3 Engineering stress-strain curves from simple tensile test Fig.5 (a) Static loading of mild steel (mode II & 750 loading)

Precrack tip

Fig.5 (b) Close-up of near crack tip field of Fig.5(a)

Precrack tip

Fig.4 Dimensions of (a) the Arcan fixture and (b) specimen[12] Fig.6 Mode II fatigue loading (75 degree loading)

Load

Load

Precrack tip Load Precrack tip

Load

Fig.7 (a) Mode II monotonic loading of Al2024-T3 specimen Fig.7 (b) Mode II fatigue loading of Al2024-T3 specimen
Static mode I
loading

Fig. 8. Static mode I apparatus mounted on an Arcan fatigue mixed mode test specimen

Table.1(a) Static mixed mode testing results

No. Precrack length Init. load β eq KI/KII θ* (deg) θ** (deg) θ (deg)
(mm) (KN) (deg) (exp)
Mild 1 4.0 9.01 0 0 -70.5 0 0
Steel 2 9.5 7.65 19.4 0.353 -64.0 6.7 7.0
3 10.0 7.65 19.9 0.361 -63.8 6.9 8.0
Al2024 1 4.7 9.56 0 0 -70.5 0 0

Table.1(b) Fatigue mixed mode testing results

No. Precrack Max load β eq ?KI /?KII θ* (deg) θ** (deg) θ (deg)
length (mm) (KN) (deg) (exp)
Mild 1 4.0 7.57 15.8 0.284 -65.2 5.2 -41.2
steel
1 4.0 13.08 0 0 -70.5 0 -69.2
Al2024 2 4.2 13.08 0 0 -70.5 0 -62.5
3 3.8 13.08 16.1 0.289 -65.1 5.3 -60.5

Table.2 Typical types of crack growth in mild steel and aluminum alloy 2024-T3

No growth Shear type growth Shear type initiation followed Tensile type growth
by tensile type growth

You might also like