You are on page 1of 14

Plastic Zone Size Measurement Techniques for Metallic Materials

Agah U$juz and John W. Martin+


*Uludag University, Miihendislik-Mimarlik Fakiiltesi Gtiriikle-16059, Bursa, Turkey; and +Department ofMaterials, University of Oxford, Oxford OX1 3PH, England The size of the plastic zone (PZS) associated with a propagating crack in a material is re-

lated to its fracture toughness. The factors determining the crack tip PZS in both static and fatigue fracture are surveyed, and the importance of measuring this parameter is discussed. A number of plastic zone size measurement techniques have been reviewed, and their ap0 Elsevier Science Inc., 1996 plications discussed.

INTRODUCTION
CRACK TIP PLASTICITY

SHAPE OF THE PLASTIC ZONE

It is well known from linear elastic fracture mechanics (LEFM) that the stresses in the vicinity of a crack tip are of the form
0.

An estimate can be obtained of the distance rP over which the material is plastically deformed ahead of the crack (i.e., 0 = 0) by substituting the yield stress, uYs into eq. (l), giving

K Tfii(0) (2nr)

+ . ..

(2)
where (Yis a numerical constant. Irwin argued that the presence of a crack tip plastic zone makes the crack behave as if it were longer than its physical size, and the distribution of stress is equivalent to that for an elastic crack of length (a + r,); that is, K = o(x(a + rP))li2 (3)

where r, 0 are the polar coordinates of a point with respect to the crack tip and K is the stress intensity factor. These equations indicate that all the stresses tend to infinity at r = 0 (the crack tip). However, the theoretical very high elastic stresses in the environs of the crack tip exceed the yield strength of the real materials and induce a small plastically deformed volume called the crack tip plastic zone (PZ). This plastic zone shields the crack tip from the high stresses and plays an important role in the fracture of materials. If the PZ is large, a large amount of energy is dissipated during crack propagation; whereas, if the PZ is small, crack propagation requires less energy, so the plastic zone size (PZS) is directly related to the toughness of the material [l]. The mechanical properties of the material, together with the state of stress, govern the size and shape of the PZ, as well as the distribution of stress and strain within the PZ itself [2].
105
MATERIALS CHARACTERIZATION 37:105-118 (1996) 0 Elsevier Science Inc., 1996 655 Avenue of the Americas, New York, NY 10010

Dugdale assumed that all plastic deformation concentrates in a strip in front of the crack, again suggesting that the effective crack length is longer than the physical length, leading to a value for rP: (4) In fact, the situation is more complicated, and the size and shape of plastic zone also depend both on the yield criterion used and on the thickness of the material of interest, because this influences the state of stress at the crack tip.
1@44-5803/96/$15.00 PII S1044-5803(96)00074-5

106 The plastic zone shape can be derived by substituting the appropriate stress equations into the yield criterion under consideration. For this purpose, rP should be determined over the range -n < 8 d + n instead of just for 0 = 0. For example, according to the Von Mises yield criterion, yielding occurs when ((3,

A. U&u and J. W. Martin

-o,)*+

((J*

-03)* +
(5)

(03 - oi )* = 2oys2

where ~1, uz, and (~3 are the principal stresses. From a Mohrs circle construction, we can derive that
012 = yY&

tion of plastic zone shapes can be found in several references [3-61. Most solutions for the PZS at the tip of a crack assume that the crack is sharp-that is, that the crack tip radius of curvature, p, approaches zero. Kim et al. [7] consider the effect of a finite p on the two-dimensional plastic domain at the tip of a blunt crack tip. The actual rP at the tip of a blunted crack is given by the plastic zone of an imaginary sharp crack, r& subjected to an applied stress intensity factor, K, minus the geometric effect of the blunting, rP; therefore:
yP = rK-rp (11)

+t xy

(6)

Thus, if two cracks have differing p but the same length and applied K, rp decreases with increasing tip bluntness.
PLASTIC ZONE IN FATIGUE

Substituting eq. (1) for ox, uY, and T~,,in eq. (6), we obtain: o, = &cosi(I+sini) (7) *cosi(l o* = &J%G -sin:)

In plane stress, 03 = 0 and, in plane strain, u = v(cr1 + crz), where u is Poissons ratio. Then for plane strain: 2vK, e 0s = J27cy cos 2 Substituting stress (8)

(7) into (5), we obtain for plane I


G i

~Wplane stress =

4 2
0

At the tip of a growing fatigue crack subject to a stress intensity range AK (= Km,, Kmi,), each loading cycle generates a monotonic plastic zone whose size corresponds to Km,, in the appropriate equation above. AS K is reduced to Kmin, a much smaller, reversed plastic zone is generated (the reversed plastic zone is approximately onequarter the size of the monotonic plastic zone). There is thus a residual zone of plastic deformation consisting of monotonically stretched material. As the crack grows, the PZ forms a wake of monotonically stretched material along the flanks of the crack.
REASONS FOR PLASTIC SIZE MEASUREMENT ZONE

YS

1
1

(9)
I

1 + g sin20 + cost3

and for plane strain 1 Ki23 r(e)planestrain = G 0 2 sin*@ + ( YS)L


(10)

(1 - 2v12(l

+ co&)1

It can be seen that, assuming v = 0.33, the plane stress value of r(0) is about nine times as great as the plane strain value straight ahead of the crack tip (0 = 0). The evalua-

In the next section, we review the experimental techniques that have been employed for the measurement of the dimensions of crack tip plastic zones under either monotonic or fatigue conditions. Measurements confined to the specimen surface will essentially obtain the dimensions of the plane stress PZ (eq. 9). In relatively thick specimens, the crack front may be in a state of plane strain within the interior; and, to measure the plane strain PZ of such

Plastic Zone Size Measurement

Techniquesfor

Metallic Materials

107

a specimen (eq. lo), a sectioning technique may have to be employed. In many cases, these measurements were conducted to verify some of the relations outlined above. Again, a knowledge of the extent of crack plasticity may permit an estimate to be made of the fracture toughness of a material from a fundamental rather than an empirical approach. Other workers have studied the phenomenon to estimate the crack tip opening displacement or to identify the deformation micromechanisms of the fracture process. P&facto studies of the PZ size associated with the crack in an unexpectedly failed component may assist in identifying the cause of failure. For example, if the yield stress and crack geometry are known, the application of eq. (9) and (10) may permit the magnitude of the failure stress to be estimated.

They also argue that a local work hardening due to plastic strains as small as 1% can be detected by microhardness measurements. Plastic zone size measurements by the etching technique (below) were in close agreement with the microhardness measurements. Two types of austenitic stainless steels were studied by Pineau and Pelloux [9]. They measured the cyclic and monotonic plane strain plastic zone sizes and found that the monotonic PZS was given by
ri = 0.053

( OYS 1

+E

and was from 4.4 to 5.8 times as great as the cyclic PZS in different heat treatment conditions for the two alloys. The microhardness technique was employed in a detailed study by Saxena and Antolovich [lo] on fatigue crack propagations of four Cu-Al alloys. The proposed equation for the cyclic plastic zone size was

PLASTIC ZONE SIZE MEASUREMENTS TECHNIQUES


MICROHARDNESS MEASUREMENTS

The microhardness technique is applicable to materials that strongly work harden (like austenitic stainless steels) or work soften (like maraging steels). The procedure is simply taking hardness indentations ahead or around the crack tip and determining the elastic-plastic boundary as a sudden variation in hardness. Bathias and Pelloux [8] used this technique to determine the monotonic and cyclic plane strain plastic zone sizes of maraging and austenitic stainless steels as a function of AK. They used 25g and 50g hardness loads, and the size of the indentation was 10km. They argue that, because of the size of the indentation, very small plastic zone sizes could be measured only by an etching technique. They found the cyclic plastic zone sizes (r: and r&) to be onefourth of the monotonic plastic zone sizes (r): and ry) where

where, for midthickness PZS straight ahead of the crack tip (rh), (Y is between 0.011 and 0.014 and 2 + S is between 1.54 and 2.20 for the four materials studied. Plastic zone size measurements by microhardness technique have also been applied on aluminum alloys by Bathias [ll] and Petit et al. ]12], on a single crystal of pure copper by Purcell and Weertman [13], and on a-brass by Kwun and Park [14]. The technique is normally employed for surface (i.e., plane stress) PZS measurements: if applied to a polished section, care must be exercised not to introduce damage during the sectioning process. These comments also apply to the following techniques.

ETCHING

TECHNIQUE

r;

010

2 (A!?1
(5,

In this technique, sections of the specimens containing the crack are etched with an appropriate etchant. Because the deformed (high dislocation density) and the undeformed regions of the material respond differently to etching, the size and shape of

108 the plastic zone can be easily observed as a dark- or light-etched region. Hahn and Rosenfield [15] and Hahn et al. [16] applied the technique to reveal the plastic zone in plates of Fe-3 wt. % Si. Specimens having different thicknesses were fatigued to different numbers of cycles, and AK values varied from 21 to 35 MPa<m. After the fatigue cracking, the specimens were first aged to decorate the dislocations; then the sections were polished and etched. They observed a highly strained, light-etching region and a lightly strained, dark-etching region. They argued that the cyclic zone corresponds to the lightly etched region whose size was in fact one-fourth that of the monotonic plastic zone. Bathias and Pelloux [8] used the etching and microhardness techniques together. Although they found it time consuming, they could successfully conduct the measurements by using the etching method and found PZ sizes equivalent to those measured by the microhardness technique. Clavel et al. [17] developed an etching technique to measure the PZS of INCO 718 alloy. They used CT and SENB specimens, and measurements were made both in plane strain and in plane stress. They found that the plane strain plastic zone was half of the plane stress plastic zone and (Y was close to 1/37~ as calculated theoretically by Irwin.
OPTICAL INTERFERENCE TECHNIQUE

A. U@LZ and J. W. Martin

light path is disturbed and the interference fringe is displaced one way or the other. A total fringe displacement corresponds to a change in height of half the wavelength of the monochromatic light used. In materials deformation studies, the interference of the light with the slip line steps in the plastic zone (in plane stress) makes it possible to determine the elasticplastic boundary. Chanani [18] demonstrates the results of plastic zone size measurements after single overload cycles during fatigue testing on a series of aluminum alloys. A Zeiss Interference Microscope with a thallium spectrum lamp was used. To obtain satisfactory interference fringes, a mirrorlike surface had to be obtained with no surface distortion. Grinding and mechanical polishing was suggested for his aluminum alloys (electropolishing was found to be unsuitable) to give a suitable surface to measure the plastic zone size in plane stress. The point where a change was observed in the fringe patterns at the crack tip was taken as the plastic zone boundary. He found a good agreement with the calculated plastic zone sizes. Pettit and Hoeppner [19] used this technique with Na light to do calibrations for their image distortion technique (discussed later in this section).
FOIL STRAIN-GAUGE TECHNIQUE

In interference microscopy, a parallel beam of monochromatic light normal to the surface of the specimen is used. An optical flat with approximately 95% reflectivity is placed in contact with the specimen and slightly tilted to produce a thin wedge between the two. The light is repeatedly reflected between the specimen and the reference flat, and interference, which produces thin and sharp dark fringes, takes place. However, to obtain sharp fringes, the reflectivity of the specimen surface should be the same as that of the reference plate (i.e., about 95%). If a variation such as a step is present on the surface of the specimen, the

In the foil strain-gauge technique, a tiny strain-gauge is situated on the specimen so that it is parallel to the loading direction ahead of the crack-tip. During loading, the output from the strain-gauge is recorded as a stress-strain diagram. When strain becomes nonlinear, the distance from the crack tip to the strain-gauge center is determined as the plastic zone size. Being a surface measurement, this method is only applicable to the plane stress plastic zone. Izumi and Fine [20] applied this method on a series of Al alloys and on a Ni-Al alloy and measured the plastic zones as a function of AK. They found a monotonic plastic zone that was two times as large as the cy-

Plastic Zone Size Measurement

Techniques for Metallic Materials

109

clic plastic zone. They also mapped the strain distributions within the butterfly wing shaped plastic zones. Kwun and Fine [21] also used the straingauge method to measure the plastic zone size of powder metallurgy aluminum alloy MA87 fatigued at two AK values. They found a 1.6 times larger plastic zone size at AK of 17 MPa<m than at 10 MPadm, which was theoretically expected to be 2.9 times larger. Pettit and Hoeppner [19] tested some specimens with SR-4 phenolic glass straingauges to calibrate their Al alloy for image distortion technique. Fine and Davidson [22] also measured the crack tip strains and compared this method with other techniques.

less-deformed areas. The recrystallization technique thus reveals local areas of high deformation in an inhomogeneously strained material. 3. Annealing may cause decarburization (e.g., in high-carbon steels) on the surface of the material, resulting in large grains. Iino 1231 studied the recrystallization of deformed grains in 304 stainless steel and stated that recrystallization occurred when this material was strained >2%. He also observed the plastic zone around fatigue cracks in the same material and found that the relation between daldn (fatigue crack growth rate) and the cyclic plastic zone size was in good agreement with the theoretical conclusions. Kwun and Park [14] also applied the recrystallization method on notched a-brass plates. They emphasized that the advantages of this method are that one can see the real shape and size of the plastic zone and one may also deduce the plastic strain distribution within the zone. It was also pointed out that the limit of plastic strain for which recrystallization can occur was 2%, so the method tends to underestimate the PZS. Petit et al. [12] observed the plastic zone of fatigued Al alloys after recrystallization and found their results consistent with those obtained with microhardness measurements.

RECRYSTALLIZATION

METHOD

It is well known that the recrystallization temperature depends on the amount of stored energy of deformation in a material, and the recrystallized grain size is a function of the amount of this energy [23] and of the annealing temperature. The deformed region around a crack or at a crack tip can therefore be revealed as a large grain region by annealing the material at appropriate temperatures. Tschegg et al. [24] measured plastic zones around fatigue cracks in mild steel with the help of the recrystallization and X-ray microbeam techniques. They enumerated some difficulties and limitations of the recrystallization technique: 1. The minimum size of the plastic zone must be greater than that of the newly formed grains. Very small plastic strains do not cause recrystallization, and local strains corresponding to less than 2% deformation cannot be detected by this method. 2. In a fatigued material, there can be zones with high and low dislocation densities close to each -other. The zones of high dislocation density initiate the recrystallization process, and the recrystallized grains then start to grow into the

X-RAY MICROBEAM

TECHNIQUE

An X-ray diffraction pattern of an undeformed crystal appears as a series of sharp spots. However, if the material is strained, owing to a change in diffraction conditions, Debye-Scherrer (DS) rings replace the sharp single reflections. Tschegg et al. [24] employed the X-ray microbeam (XMB) technique as well as the recrystallization method to measure plastic zones around fatigue cracks in mild steel. They used an X-ray tube with a Cr-anode at 30kV, and the collimator tube had a diameter of 250 km. The irradiated area on the

110 specimen was 300 pm in diameter. They managed to calibrate this method by measuring the integral line breadth of samples with known amounts of deformation. They took a number of diffraction patterns along the crack and perpendicular to the crack edge and, by measuring the DS line broadening, they mapped the plastic zone around the fatigue cracks. Latiere [25] used a tungsten target at 60kV with a 400km diameter collimator to measure the plastic zone size of aluminum alloys. Yokobori et al. [26] also used this technique at the tips of fatigue cracks in low-carbon steel by employing a Co target at 30kV. They found the following relation between the plastic zone size and AK:
yP =

A. U@z and 1. W. Martin

the plastic zone to the shear lip size found on the fracture surface of a material. He suggests a similar approximation to Lai and Fergusons [29] and uses this approach to estimate the stress level.

SELECTED-AREA CHANNELING PATTERN TECHNIQUE Electron channeling can be briefly explained as a mechanism of contrast obtained when either backscattered or forward-scattered electrons, with approximately the same energy as those in the beam, are used in an imaging mode in the scanning electron microscope (SEM). When an impinging electron beam interacts with the planes of an illuminated crystal, electron diffraction occurs. When this beam is swept across the crystal, diffraction conditions change with the change of the incidence angle. As a result, a diffraction pattern may be observed [31]. Among many theoretical treatments of this diffraction process, that of Spencer and Humphreys is the most quantitative [32]. Channeling patterns may be produced from a very small area in the form of selected-area electron channeling patterns (SACIs). In the SACP technique, the electron beam is made to rock about a spot on the specimen surface. This means physically that the scanning electron beam is directed so as to meet the electron-optical axis of the microscope at the specimen surface. The size of the spot on the specimen surface from which the SACP is produced can be varied (over the range lo-1000pm in diameter) by changing the electron-optical conditions in the microscope. This makes the technique very advantageous for material-deformation studies, because it allows averaging of information over an area of chosen dimensions. The minimum spot size can be as small as 2km, with the use of a scanning transmission electron microscope (STEM), because of the lower aberrations of coils and lenses in a STEM [33]. A large rocking angle means that a maximum angular coverage is available in the pattern;

9.7X 10P1AK32

Matsumoto and Kitagawa [27] applied an XMB to the fracture surfaces of a low carbon-low alloy steel. They measured the plastic zone depth (perpendicular to the fracture surface) by using a Co target and an illuminated area of 400 X 800pm2. By using a similar technique called X-ray fractography, Biognonnet et al. [28] also measured the in-depth PZS of a Cr-alloyed steel.

SHEAR LIP SIZE MEASUREMENTS Lai and Ferguson [29] measured the shear lip size of steel and 7075-T6 aluminum alloy compact tension specimens after they had been broken apart. They found the following relation between the plane strain fracture toughness and the size of the shear
lip (BsL): 2

B SL = 0.41 5

i OYS 1

They argue that this equation is similar to Dugdales theoretical plane stress plastic zone (eq. 4); therefore, BSL = rp. The shear lip size is not only a representation of the plastic zone, but also a measure of its size. However, only the ultimate plastic zone size (at KI = Ki,) in plane stress was estimated by this approach. Hertzberg [30] also correlates the size of

Plastic Zone Size Measurement

Techniques for Metallic Materials

111

that is, the maximum amount of available information is being derived from the material. It also means that the beam is being moved a considerable distance off the optical axis of the microscope, and that condition leads to a large spot size, because lens and coil aberrations increase with off-axis distance. In plastic zone size measurements, smaller spot sizes (i.e., smaller rocking angles) are preferred to minimize the error in determining the deformed region. The principles of channeling and the theory and applications have been reviewed by Schulson [34], Joy et al. [35] and Davidson Deformation changes the intensity and width of channeling pattern bands. Dislocations have a small core region (-lnm) of large atomic distortion, in addition to a much larger region (m30nm) of elastic atomic distortion. The incident electron beam interacts with these regions of distortion, leading to a decrease in channeling line acuity and intensity. These effects are the result of elastic bending of crystal planes by the dislocations. Detailed information of channeling patterns in deformation studies is given in the references [36-381. Tekin and Martin [33] illustrate the degradation of SACP with strain, the means of obtaining the best channeling patterns, and the procedure for plastic zone size mapping; the advantages and disadvantages of the technique are summarized elsewhere [31,33 and 39411. By using this method, one can map the plastic zone at the crack tip and around the crack of any crystalline material exhibiting clear channeling patterns both in plane stress, from observations on the external surface, and in plane strain conditions, by preparing sections taken through the specimen thickness. Many workers have applied this technique in PZS determination for metallic materials [33,41-551. Davidson [42] and Lankford and Davidson [43] applied this technique in order to measure the plastic zone size of 6061 Al alloy both in fatigued and overloaded conditions. In another detailed study by Davidson and Lankford 1441, the PZS of fatigue cycled 304 stainless steel, Fe-3 wt. % Si steel

1311.

and 6061 Al alloy were measured. They combined the positive replica technique (see below) with SACP and found that the strains are proportional to r-i/* in the cyclic plastic zone-see Post [60] and to In r in the monotonic plastic zone. Davidson and Lankford [45] also measured the plastic zone sizes and the strain distributions within the plastic zones of a series of Al alloys by using calibration standards obtained from tensile tests. They found that the plastic zone sizes were not proportional to (K/o&* but proportional to (dW/dA)E/u& where dW/dA is a measure of the work expended in forming new crack surfaces. The plastic zone sizes at the tips of small fatigue cracks in 7075 Al alloys were measured by the SACP technique [46, 471. Lankford et al. [46] found a AK* dependence of YP, whereas Zhang and Edwards [47] developed a r-,/a ratio (ratio of plastic zone size to half surface crack length), being 0.5 for very small cracks and decreasing with increasing crack length to 0.25 when a = 200km. Selected-area channeling patterns have been used to measure the plane strain plastic zone sizes in CT fracture toughness specimens of Al-Mg-Si alloys containing different volume fractions of dispersoid phase by Prince and Martin [48, 491 and Blind and Martin [50, 511. They found that dispersoid-free alloys had low toughnesses and very small plastic zones at fracture. The PZSs of all the alloys obeyed eq. (2), although the magnitude of the constant (Yincreased with increasing dispersoid content because of the slip homogenization characteristics of the dispersoids, leading to crack tip blunting (eq. 11). They calculated CK in eq. (2) to be between 0.12 and 0.034. Crompton and Martin [52,53] and Tekin and Martin [33, 541 measured the plastic zones sizes of a series of fatigued superal10~s. The increase in PZS with the increasing test temperature was observed. Tekin and Martin [33] also established the relation
Yp =

A. U@LZand J. W. Martin

Electron channeling contrast image of a fatigue crack tip in SRR99. The diffraction conditions used are shown in the inset electron-channeling pattern. The plastic zone is seen as dark and bright regions surrounding the crack, and the outer edge has been marked as have a number of boundary points determined by using a conventional SACP technique (from Wilkenson [68]).
FIG. 1.

where da/&zis the fatigue crack growth rate. Uguz and Martin [41, 551 also measured the plastic zone sizes of Al-Zn-Mg alloys by means of SAG. They mapped the plastic zones at different stress intensity factors and found the relation
rP =

IMAGE

DISTORTION

0.033 K ( OYs 1

The technique is based on the observation of superimposed photographs of undeformed and deformed crack tips (previously marked). Pettit and Hoeppner [19] used center-cracked thin specimens, and their measurements were in close agreement with the theoretical plane stress plastic zone size:
0.55
rY = 7

which was in close agreement with the theoretical calculations by Rice and Johnson [56], who suggested a value of 0.032 for 01in plane strain.

4
( (Jys 1

POSITIVE SLIP LINES METHOD

REPLICA

This method essentially depends on the observation of slip lines around and at the crack tip under an optical microscope [12, 261. It is only applicable to plane stress plastic zones.

This technique consists of making a plastic replica of the crack-tip region with methyl acetate. The replica is coated with palladium by vapor deposition and then plated with nickel to produce a strong positive replica after the plastic is peeled. This replica is then examined under the SEM [44,57].

Plastic Zone Size Measurement

Techniques for Metallic Materials

113

PHOTOELECTRON

MICROSCOPY

Photoelectron microscopy, explained in Hammond [58], is based on the observation of plastic deformation zones of a metal as regions of enhanced electron emission, which is called photostimulated exoelectron emission. This enhancement is due to the rupturing of the brittle natural oxide layer of the deformed metal 1591. MOIRE INTERFEROMETRY This technique provides whole-field measurements of in-plane displacement fields on a specimen surface with submicron sensitivity. In this technique, a grid with lines parallel to the crack is placed on the specimen surface; after deformation, the change in the fringe spacing is used to calculate the amount of strain, and consequently the plastic zone size [60-641. A recent paper by Poon et al. [65] demonstrate that Moire interferometry combined with an automated image-processing technique provides a robust and highly efficient method for measuring strain fields on various types of materials, in particular composites that are difficult to model in terms of strength and failure mechanisms. ELECTRON CHANNELING CONTRAST IMAGING Electron channeling contrast imaging (ECCI) can be used to reveal the formation of subgrains due to deformation adjacent to the crack. Therefore, the plastic zone can be measured [66, 671. In very recent work by Wilkinson et al. [68] the technique was applied successfully by using a commercially available SEM fitted with a field emission gun. The size and shape of the plastic zone associated with a fatigue crack in a nickelbased superalloy was revealed (Fig. l), the measurements having been completed more swiftly and completely than with the method using the degradation of electron channeling patterns (ECPs). The shift in the position of zone axes in electron backscatter diffraction (EBSD) patterns was also used to measure the extent of the lattice plane tilt-

ing, thus providing quantitative information about the strains within the plastic zone. DEFORMED GRAINS

This method is applicable to materials having very fine grains. The deformed grains are observed under an optical microscope and the change in the aspect ratios of the grains is used to calculate the amount of strain and the boundary of the plastic zone 1691. TRANSMISSION ELECTRON MICROSCOPY

In this technique, the dislocation substructure around and at the tip of the crack is observed under a transmission electron microscope (TEM). Klesnil and Lucas 1701 observed tiny cell structure of dislocations in the plastic zone of fatigued Cu single crystals, different from the undeformed material; see also Purcell and Weertman [13]. CAUSTICS METHOD

An optical shadow method known as caustics was first introduced by Manogg [71]. In plastic zone size measurements, a set of parallel light rays obtained from a laser beam source are incident on a polished specimen surface. The reflected light rays at the crack tip deviate from parallelism and an image is obtained on a screen away from the specimen. By using this method, Lee and Hong [72] directly observed and measured the plastic zone sizes of an aluminum alloy (Al 5086 H24) and polycarbonate plates.

CONCLUSIONS

Table 1 lists the various plastic zone size measurement techniques described herein, together with some of the references and the materials studied. Table 2 presents a summary of the advantages and disadvantages of the different techniques. Over time, a substantial number of techniques have been developed for measuring the plastic zone size at the tip and around the cracks of metallic materials. All have

114
Table
TdWlipe

A. Ll@z and 1. W. Martin


1 References on Plastic Zone Size Measurement Reference and etching Bathias and Pineau and Saxena and Techniques Material [7] Maraging and austenitic stainless steels Austenitic stainless steels Cu-Al alloys Al-alloys and stainless steel Al-alloys Cu-single crystals o-brass Fe3Si INCO 718 2024,7075 Al-alloys Al-alloys and Ni-Al alloy MA87 Al P/M alloy 304 stainless steel Mild steel Al-alloys Low C steel Low C-low alloy steel Cr-alloyed steel Steel and 7075 Al-alloy 6061 Al-alloy 6061 Al-alloy 304 SS, 6061 Al-alloy, Fe-3Si 2024,6061, and 7075 Al-alloys alloy 7075 Al-alloy Al-Zn-Mg-Cu Al-Mg-Si alloys Al-Mg-Si alloys Nimonic 90 and MAR-M246 superalloys MAR-MOO2 single crystal superalloy MA6000 superalloy Al-Zn-Mg alloys 7075 Al-alloy Stainless steel SAE 1018 and 1015 steel 2024 Al-alloy and 4340 steel 7075 Al-alloy 7075 Al-alloy and steels Low-C steel SRR99 sinle crystal superalloy HSLA steels Cu-single crystals Al alloy and polycarbonate ]

Microhardness Microhardness Microhardness Microhardness

Pelleoux .............. Pelleoux .............. Antolovich ............

[S]
[9] [lo] [ll] [12]

Bathias .......................... Petit et al.

Microhardness, slip lines and recrystallization Microhardness and TEM Microhardness and recrystallization Etching Etching Optical interference Foil strain-gauges Foil strain-gauges Recrystallization Recrystallization X-ray microbeam X-ray microbeam X-ray microbeam X-ray microbeam Shear lip size Selected area channeling patterns (SACP) SACP SACP and positive replica SACP SACP SACP SACP SACP SACP SACP SACP SACP Image distortion Positive replica Photoelectron microscopy Moire interference Moire interferometry Moire interferometry Electron channeling Electron channeling Deformed grains TEM Caustics contrast contrast and XMB (XMB) and slip lines

Purcell and

....................... Weertman...........

................. ................. Clavel et al. ...................


Hahnetal. Chanani........................[17

Kwun and Park

[131
[14,15] .[16]

................. Kwun and Fine. ................. lino .............................


Tschegg et al. .................... Latiere

Izumi and Fine

,[19]
[20] [22] [23] [24] [25] [26]

..........................
and

Yokobori et al. ................... Matsumoto

Kitagawa ........ Biognonnet..................... Lai and Ferguson................


Davidson

[2i]
[28] [43]

....................... ........ Davidson and Lankford ......... Davidson and Lankford .......... Lankford et al. ................. Zhang and Edwards ............
Lankford and Davidson. Prince and Martin. Blind and Martin. ...........

[44]
[45] [46] 1471 .[48] [49,50]

Crompton and Martin.

............ [51,52] .......... [53] ..........


[54] .[34,55] .[42,56] [18] [58] [60] [62] [64] [65] [66] [67]

Crompton and Martin. Tekin and Martin Uguz and Martin. Pettit and

............
...........

Hoeppner ............. Lankford and Barbee. ........... Baxter and Rouze................ Liu and lino ..................... Nicoletto ........................ Nicoletto ........................
Davidson et al. Lankford et al. Wilkinson et al. Luoetal.

................. ...............

.[68]
[68] [69] [71]

......................

Klesnil and Lucas ................ Lee and Hong.

..................

Plastic Zone Size Measurement Table Method Microhardness 2 Advantages

Techniques for Metallic Materials of Selected PZS Measurement Advantages Practical and rapid Techniques Disadvantages Applicable to strongly work hardening or softening materials: Sensitivity depends on the indentation size

115

and Disadvantages
Plarze strain/

P2ane stress Both

Etching

Both

Direct observation

of the

Etchants and etching conditions should be determined for each material: Time consuming Interference microscope and perfect polishing of the specimen surface needed Delicacy of dealing with tiny strain-gauges Less than 2% deformation cannot be measured: In carbon-containing materials decarburization may result Sensitivity is limited owing to probe size

plastic zone is possible Optical interference Plane stress Direct observation of the plastic zone is possible: Very sensitive Foil strain-gauges Plane stress Strain distributions within the plastic zone can be determined Recrystallization Both Real shape and size of the plastic zone can be observed, strain distribution can X-ray microbeam Both be constructed The amount of plastic deformation can be determined in depth: PZS can be measured on fracture surfaces Shear lip size measurements Selected area channeling patterns (SAG) Image distortion Photoelectron microscopy Moire interferometry Plane stress Plane stress Plane stress Plane stress Both Practical and rapid Measurements can be repeated: Sensitivity is good: Applicable to any crystalline material Direct observation of the plastic zone is possible Direct observation of the plastic zone is possible Submicron displacements

Sensitive measurements cannot be made Specimen preparation may be time consuming: Selected area size should be smaller than the grain size Not very sensitive Photoemission microscope is needed and not very sensitive Vibration-free environment is needed Applicable to very fine grained materials With EBSD can measure strains within the PZ

and

Deformed grains Electron channeling contrast imaging (ECCI)

Both Both

the amount of deformation can be determined Strain distribution can be determined More rapid than SACP

116 been able to perform measurements for the specific conditions under which they were developed and applied, but no single method can be recommended as being superior to all others for all materials and conditions. Thus, some of the techniques are applicable to materials of certain properties. For example, microhardness measurements would be appropriate for workhardening materials, whereas the deformed grains method is specifically directed toward very finely grained materials. The applicability of some surface techniques, such as foil strain-gauging and photoelectron microscopy, is limited to plane stress conditions. Consequently, each combination of materials properties and plastic zone size must be considered separately before a decision can be made as to which technique might be the most appropriate one to use.

A. U@z

and f. W. Marfin

11. C. Bathias, Plastic Zone Formation and Fatigue Crack Grotitk, Proc. 4th Int. Conf. on Fracture II (ICF4), pp. 1307-1312 (1977). 12. J. Petit, B..Bouchet, C. Gasc, and J. de Fouquet, A Contribution to the Study of the Influence of Environment on the Crack Growth Rate ofHigh-Strength Aluminum Alloys in Fatigue, Proc. 4th Int. Conf. on Fracture II (ICF4), pp. 867-872 (1977). 13. A. H. Purcell and J. Weertman, Crack tip area in fatigued copper single crystals. Metall. Trans. 5: 1805-1809 (1974). 14. S. I. Kwun and S. H. Park, Plastic zone size measurement by critical grain growth method. Scriptu Metall. 21:797-800 (1987). 15 G. T. Hahn and A. R. Rosenfield, Sources of fracture toughness: the relation between K,, and the ordinary tensile properties of metals, in Applications Related Pkenomenn in Titanium Alloys, ASTM STP 432, American Society for Testing and Materials, Philadelphia, pp. 5-32 (1968). 16 G. T. Hahn, R. G. Hoagland, and A. R. Rosenfield, Local yielding attending fatigue crack growth. Metall. Trans. 3:1189-1202 (1972). 17 M. Clavel, D. Fournier, and A. Pineau, Plastic zone sizes in fatigued specimens of INCO 718. Mefali. Trans. A 6:2305-2307. 18 G. R. Chanani, Determination of plastic-zone sizes at fatigue-cracks by optical interference technique. ht. 1. Fruct. 13:394-399 (1977). 19 D. E. Pettit and D. W. Hoeppner, The influence of specimen geometry on crack tip plasticity. Eng. Fract. Meek. 5:923-934 (1973). 20. Y. Izumi and M. E. Fine, Role of plastic work in fatigue crack propagation in metals. Eng. Fract. Meek. 11:791-804 (1979). 21. S. I. Kwun and M. E. Fine, Dependence of cyclic plastic work of fatigue crack propagation on AK in MA87 Al P/M alloy. Scripfn Mefall. 14~155158 (1980). 22. M. E. Fine and D.,L. Davidson, Quantitative measurement of energy associated with a moving crack, in Fatigue Mechanisms: Advances in Quuntitatiue Measuremeht of Physical Damage, ASTM STP 811, American Society for Testing and Materials, Philadelphia, pp. 350-370 (1983). 23. Y. Iino, Accumulated plastic zone around fatigue crack in type 304 stainless steel. Mefu2 Sci. 10:159164 (1976). 24. E. Tschegg, C. Faltin, and S. Stanzl, X-ray microbeam and recrystallization studies of plastic deformation around fatiguk cracks. I. Muter. Sci. 15:131138 (1980). 25. H. J. Latiere, X-ray diffraction study of the plastic zone. Eng. Fract. Meek. 8:691-700 (1976). 26. T. Yokobori, K. Sato, and H. Yaguchi, Observations of microscopic plastic zone and slip band zone at the tip of Fatigue Crack, in Report of the Research Institutefor Strength and Fracture of Materials, vol. 9, Tohoku University, pp. l-10 (1973).

References
1. S. Banejee, Influence of specimen size and configuration on the plastic zone size: toughness and crack growth. Eng. Fract. Meek. 15:343-390 (1981). 2. D. L. Davidson and J. Lankford, Fatigue crack propagation: new tools for the study of an old problem. J. Metals 31:11-16 (1979). of Fracture 3 J. F. Knott, Fundamentals Wiley, New York, pp. 119-120 (1979). Mechanics,

4 D. Broek, Elementary Engineering Fracture Meckanits, Sijthoff and Noordhoff, The Netherlands, pp. 96-111 (1978). 5 H. L. Ewalds and R. J. H. Wanhill, Fracture Mechanics, Edward Arnold, The Netherlands, pp. 63-74 (1985). 6 M. F. Kanninen and C. H. Popelar, Advanced Fracture Mechanics, Oxford University Press, New York, pp. 172-182, (1985). 7. Y. H. Kim, M. E. Fine, and T. Mura, Plastic yielding at the tip of a blunt notch during static and fatigue loading. Eng. Fract. Meek. 11:653-660 (1979). 8. C. Bathias and R. M. Pelloux, Fatigue crack propagation in martensitic and austenitic steels. MetaL Trans. 4:1265-1273 (1973). 9. A. G. Pineau and R. M. Pelloux, Influence of strain-induced martensitic transformations on fatigue crack growth rates in stainless steels. Metall. Trans. 5:1103-1112 (1974). 10. A. Saxena and S. D. Antolovich, Low cycle fatigue, fatigue crack propagation, and substructures in a series of polycrystalline Cu-Al alloys. M&U. Trans. A 6:1809-1828 (1975).

Plastic Zone Size Measurement

Techniques

for Metallic Materials

117

27. T. Matsumoto and H. Kiagawa, X-ray Investigation o,fFafigue Crack Growth OM Critical Strain for Fracture at fhe Crack Tip, Proc. Int. Conf. on Mechanical Behaviour of Materials II, 1971, The Society of Materials Science, Japan, pp. 59-66 (1972). 28. A. Biognonnet, A. Dias, and J. L. Lebrun, Fatigue Failure Analysis by X-ray Fractography, Proc. 7th Conf. on Fracture (ICF7) V, Pergamon Press, Oxford, pp. 3457-3463 (1989). 29. M. 0. Lai and W. G. Ferguson, Relationship between the shear lip size and the fracture toughness. Mater. Sci. Eng. 45:183-188 (1980). 30. R. W. Hertzberg, Deformation and Fracture Mechanics of Engineering Materials, Wiley, Singapore, pp. 608-610 (1989). 31. D. L. Davidson, Uses of electron channelling in studying material deformation. ht. Metals Rev. 29: 75-95 (1984). 32. J. I. Spencer and C. J. Humphreys, A multiple scattering transport theory for electron channelling patterns. Phil. Msg. A 42:433451 (1980). 33. A. Tekin and J. W. Martin, High-resolution measurement of crack-tip plastic zone sizes by selected area channelling patterns. Metallography 22:1-14 (1989). 34. E. M. Schulson, Electron channeling patterns in scanning electron microscopy. 1. Mater. Sci. 12: 1071-1087 (1977). 35. D. C. Joy, D. E. Newbury, and D. L. Davidson, Electron channelling patterns in the scanning electron microscope. 1. Appl. Phys. 53:R81-R122 (1982). 36. R. Stickler, C. W. Hughes, and G. R. Booker, Application of the SA-ECP method to deformation studies, in Scnnning Electron Microscopy/l971 (Part l), Proc. 4th Annual Scanning Electron Microscope Symposium, ITT Research Institute, Chicago, pp. 473-481 (1971). 37. D. L. Davidson, A method for quantifying electron channelling pattern degradation due to material deformation, in Scanning Electron Microscopy/l974 (Part 4) Proc. Workshop on Failure Analysis and the SEM, ITT Research Institute, Chicago, pp. 927934 (1974). 38. J. P. Spencer, G. R. Booker, C. J. Humphreys, and D. C. Joy, Electron channelling patterns from deformed crystals, in Scanning Electron Microscopy/ 1974 (Part 4) Proc. Workshop on Failure Analysis and the SEM, ITT Research Institute, Chicago, pp. 919-926 (1974). 39. G. R. Booker and R. Stickler, Scanning electron microscope selected area channelling patterns: dependence on area on rocking angle and working distance. J. Mater. Sci. Lett. 7712-713 (1972). 40. E. M. Schulson and C. G. van Essen, Optimum conditions for generating channelling patterns in the scanning electron microscope. 1. Sci. Instruments (1. Physics E) 2:247-251 (1969). 41. A. U&r and J. W. Martin, Plastic zone size mea-

surements in Al-Zn-Mg alloys by means of selected area channelling patterns. Z. Metalkunde 83: 270-275 (1992). 42. D. L. Davidson, Fatigue crack plastic zone and shape through the specimen thickness. Int. 1. Fract. 11:1047-1048 (1975). 43. J. Lankford and D. L. Davidson, Fatigue crack tip plasticity associated with overloads and subsequent cycling. 1. Eng. Mater. Tech. 98:17-23 (1976). 44. D. L. Davidson and J. Lankford, Plastic strain distributions at the tips of propagating fatigue cracks. I. Eng. Mater. Tech. 98:24-29 (1976). 45. D. L. Davidson and J. Lankford, Fatigue crack tip plastic strain in high-strength aluminum alloys. Fatigue Eng. Mater. Strucf. 3:289-303 (1980). 46. J. Lankford, D. L. Davidson, and K. S. Chan, The influence of crack tip plasticity in the growth of small fatigue cracks. Metall. Trans. A 15:1579-1588 (1984). 47. Y. H. Zhang and L. Edwards, The effect of grain boundaries on the development of plastic deformation ahead of small fatigue cracks. Scripta Metall. Mater. 26:1901-1906 (1992). 48. K. C. Prince and J. W. Martin, The study of local plastic strain distribution in aged Al alloys by selected area channelling patterns in the SEM. Metallography 10:107-113 (1977). 49. K. C. Prince and J. W. Martin,The effects of dispersoids upon the micromechanisms of crack propagation in Al-Mg-Si alloys. Acta Metall. 27: 1401-1408 (1979). 50. J. A. Blind and J. W. Martin, The effect of dispersoids on the ductile fracture toughness of Al-MgSi alloys. Mater. Sci. Eng. 57:49-54 (1983). 51. J. A. Blind and J. W. Martin, The effect of dispersoids on the micromechanisms of crack extension in Al-Mg-Si alloys. 1. Mater. Sci. 18:1224-1234 (1983). 52. J. S. Crompton and J. W. Martin, The study of local plastic strain in nickel-based superalloys by selected area channelling patterns in the STEM. Metallography 13:225-234 (1980). 53. J. S. Crompton and J. W. Martin, Crack tip plasticity and crack growth in a single crystal superalloy at elevated temperatures. Mater. Sci. Eng. 64:3743 (1984). 54. A. Tekin and J. W. Martin, Fatigue Crack Growth Bekaviuur ufMA6000, Proc. Conf. on High Temperature Alloys for Gas Turbines and Other Applications II, W. Betz et al., eds., Liege, Belgium pp. 1057-1065 (1986). 55. A. U&z and J. W. Martin, The E/j&d of Dzflerent Dispersoids on the Ductile Fracture Toughness of an Al-Zn-Mg Alloy, Proc. 7th Int. Conf. on Fracture (ICF7) II, Pergamon Press, Oxford, pp. 1045-1052 (1989). 56. J. R. Rice and M. A. Johnson, The role of large crack tip geometries in plane strain fracture, in In-

118
elastic Behavior of Solids, McGraw-Hill, pp. 641-672 (1970). New York,

A. U@.LZ and 1. W. Martin 66. D. L. Davidson, J. Lankford, T. Yokobori, and K. Sato, Fatigue crack tip plastic zones in low carbon steel. Int. J. Erect. 12579-585 (1976). 67. J. Lankford and D. L. Davidson, Environmental alteration of crack tip dislocation cell structure and mode of growth during fatigue crack propagation in ferritic steel. Int. J. Frucf. 12775-776 (1976). 68. A. J. Wilkinson, M. B. Henderson, and J. W. Martin, Examination of fatigue crack plastic zones using SEM-based electron diffraction techniques. Phil. Mag. Left. 74:145-151 (1996). 69. L. G. Luo, A. Ryks, and J. D. Embury, On the development of a metallographic method to determine the strain distribution ahead of a crack tip. Metallography 23:101-117 (1989). 70. M. Klesnil and I. Lucas, Dislocation Substructure Associated Propagating Fatigue Crack, Proc. 2nd Int. Conf. on Fracture (ICF2), Chapman and Hall, London, pp. 725-730 (1969). 71. I. Manogg, Investigation of the rupture of a Plexiglas plate by means of an optical method involving high-speed filming of the shadow originating around holes drilling in the plate. Int. J. Fruct. Mech. 2:604613 (1966). 72. 0. S. Lee and S. K. Hong, Determination of stress intensity factors and J-integrals using the method of caustics. Eng. Fruct. Mech. 44:981-989 (1993).

57. J. Lankford and J. G. Barbee, SEM characterization of fatigue crack tip deformation in stainless steel using a positive replica technique. J. Muter. Sci. 9: 1906-1908 (1974). 58. C. Hammond, The Future ofPhoto Emission Electron Microscopy, Proc. RMS, 19, pp. 39-44 (1984). 59. W. J. Baxter and S. R. Rouze, Photoelectron microscopy of plastic zone around a fatigue crack. Metall. Trans. A 7647-654 (1976). 60. I. Post, The Moire grid analyzer method for strain analysis. Exp. Mech. 5:368-377 (1965). 61. H. W. Liu and N. Iino, A Mechanical Model for Fatigue Crack Propagation, Proc. 2nd Int. Conf. on Fracture (ICFZ), Chapman and Hall, London, pp. 812-823 (1969). 62. H. W. Liu and J. S. Ke, Moire method. Exp. Techn. Fract. Me&. 2:111-165 (1975). 63. G. Nicoletto, Fatigue crack tip strains in 7075-T6 aluminum alloy. Fatigue Fracf. Eng. Mater. Struct. 10~37-49 (1987). 64. G. Nicoletto, Plastic zones around fatigue cracks in metals. Int. J. Fatigue 2107-115 (1989). 65. C. Y. Poon, M. Kujawinska, and C. Ruiz, Strain measurements of composites using an automated Moire interferometry method. Measurement 123.547 (1993).

You might also like