Sei sulla pagina 1di 11

Void Nucleation by Inclusion Cracking

M.N. SHABROV, E. SYLVEN, S. KIM, D.H. SHERMAN, L. CHUZHOY, C.L. BRIANT, and A. NEEDLEMAN Void nucleation is studied both experimentally and computationally with the aim of identifying a macroscopic criterion for nucleation by particle cracking. Three types of circumferentially notched cylindrical specimens made of a low-alloy steel were used, in order to vary the stress triaxiality in the notch region. The tensile tests were interrupted at various loads below the fracture load. The specimens were sectioned parallel to the loading axis, and the locations of cracked and uncracked titanium-nitride inclusions were identified. No evidence was found of void nucleation by inclusion debonding. Finite-element calculations were carried out for each specimen geometry using conventional isotropic-hardening plasticity theory. The ability of various potential void-nucleation criteria to predict the onset of void nucleation by inclusion cracking is explored.

I.

INTRODUCTION

ROOM-TEMPERATURE ductile fracture in structural metals is a problem of technological importance and scientific interest that has received extensive attention from a variety of perspectives. The voids generally nucleate by decohesion or fracture of second-phase particles and grow by plastic deformation of the surrounding matrix. Void coalescence occurs either by necking down of the matrix material between adjacent voids or by localized shearing between wellseparated voids. The role played by void nucleation, growth, and coalescence was identified by Tipper,[1] and, subsequently, the phenomenology of this process was documented by Puttick,[2] Rogers,[3] Beachem,[4] and Gurland and Plateau.[5] Mechanics modeling of the ductile-fracture process originated with the pioneering work of McClintock[6] and Rice and Tracey.[7] Reviews of various aspects of the ductile-fracture literature with extensive references can be found in the articles by Goods and Brown,[8] Garrison and Moody,[9] Tvergaard[10] and in a series of reviews articles in a volume honoring F.A. McClintock.[1114] Although much progress has been made, development of quantitative descriptions of ductile fracture continues to be an active area of research; recent studies and further references to the recent literature can be found in References 15 through 19. Of particular focus here is ductile failure in steels, which has been a research topic of great interest for many years; refer, for example, to References 20, 21, and 24 through 26, in addition to references in the reviews already cited. Although the main body of this work is aimed at providing a description of microvoid growth and coalescence (e.g., References 1 through 3, 5, and 27 through 31), the process of void nucleation also received attention in, for example,
M.N. SHABROV, Graduate Student, and C.L. BRIANT and A. NEEDLEMAN, Professors of Engineering, are with the Division of Engineering, Brown University, Providence, RI 02912. S. KIM, formerly Post Doctoral Researcher, Division of Engineering, Brown University, Providence, RI, is a Researcher in the Plate and Rod Group, POSCO, Pohang, 790-785, South Korea. Contact e-mail: clyde_briant@brown.edu E. SYLVEN, formerly Graduate Student, Division of Engineering, Brown University, is with ZAPP USA, Dartmouth, MA 02745. D.H. SHERMAN and L. CHUZHOY are with Caterpillar, Inc. TCK/Div 854, Peoria, IL 61656-1875. Manuscript submitted June 23, 2003.
METALLURGICAL AND MATERIALS TRANSACTIONS A

References 8, 21 through 23, 32 through 34, 37, and 38. In steels, most studies found that microvoids first form around ellipsoidal MnS inclusions in the steels through debonding of the steel-sulfide interface. Two reasons have been suggested to explain why these interfaces are so weak. One is that the difference in the coefficient of thermal expansion between these inclusions and the matrix is such that this interface is in tension.[35] The other is that sulfur segregated to this interface weakens it.[36] A significant amount of research has also addressed the issue of how these large voids link up to cause fracture. Most results indicate that small voids form around the carbides very late in the fracture process, and these small voids help link the large voids formed around the sulfides.[3,28,30,36] In most of these steels, the sulfides are arrayed in the material so that the fracture plane contains the long axis of the ellipsoidal inclusions. Relatively little work has been done for materials in which the long axis is perpendicular to the fracture plane, which would presumably make these inclusions less effective as void nuclei. In addition, there has been relatively less work investigating the role that inclusions other than sulfides play in nucleating the initial microvoids (e.g., References 8 and 23). Cox and Low[30] reported results on an 18-Ni maraging steel in which the nucleating inclusions were titanium carbo-nitrides. They found that these inclusions cracked when a critical stress was reached and that these cracked inclusions nucleated microvoid formation. Various void-nucleation criteria have been proposed based on experimental observations or micromechanical analyses.[34,3943] In order to formulate a physically based nucleation criterion, careful attention must be paid to the mechanism by which this process occurs. As suggested earlier, the two most common mechanisms would be either debonding of the interface between the inclusion and the matrix, as has been reported for MnS, or cracking of the second-phase inclusions, which has been reported for the titanium carbo-nitrides in maraging steel or Fe-Si inclusions in aluminum alloys.[50] In this article, we report on a study in which we have examined microvoid nucleation in a quenched and tempered steel. In contrast to the behavior in most low-alloy steels, the microvoids in this material nucleated from cracked TiN inclusions. The MnS inclusions had their long axis parallel to the tensile axis and did not appear to participate in the
VOLUME 35A, JUNE 20041745

fracture process. As in many previous ductile-fracture studies, e.g., References 44 through 49, circumferentially notched bars were used, and analyses were carried out to determine the stress state under which inclusion cracking occurred. We pulled these samples to various load levels, stopped the test, and then sectioned the sample to determine if nucleation of microvoids had occurred. With this information we can suggest a critical stress condition for inclusion cracking in these steels. The data also show that the onset of void nucleation by inclusion cracking depends on the inclusion size.

II.

EXPERIMENTAL PROCEDURE

The steel used for this study was a quenched and tempered SAE 4330 modified steel that contained 0.040 wt pct Ti. The steel was heat treated by first austenitizing the material at 975 C for 1 hour in a neutral atmosphere, quenching in highly agitated water, and then tempering at 215 C for 1 hour. Sample with dimensions of 19 19 150 mm were cut from a single plate for heat treatment and testing. After heat treatment, the samples were machined into circumferentially notched cylindrical tensile specimens, as shown in Figure 1. The radius of the notch was machined to either 1,

Fig. 1Photographs of the three circumferentially notched bars used in this work. The upper bar has a notch radius of 1 mm, the middle bar has a notch radius of 2 mm, and the lower bar has a notch radius of 4 mm.

2, or 4 mm. This variation in notch radius gave different states of stress triaxiality. The diameter of all samples at the apex of the notch was 6.35 mm. The notched specimens were 114.3-mm long and 12.7 mm in diameter, with notches introduced in the middle. The mechanical tests were performed in the following way. A sample of each notch geometry was pulled to failure on an Instron testing machine at a constant crosshead speed of 0.2 mm/min, which corresponds to a strain rate of the order of 105. That test allowed us to determine the failure load for a given notch geometry. We then pulled samples to various fractions of that load and stopped the test. The load was removed at a crosshead speed of 0.6 mm/min. A longitudinal section of these samples was then prepared by electrodischarge machining, so that we could view the midplane. This surface was prepared metallographically and examined in the scanning electron microscope. From this examination, we could make maps of the polished surfaces that showed the coordinates of all the inclusions. For most samples, we took traverses across the sample from one side of the notch to the other. These were spaced so that on any given sample, there were ten to fifteen traverses in the notch region. We recorded the position of all inclusions in those traverses and determined whether or not the inclusion was cracked. For two samples, we determined the coordinates of every inclusion that we viewed, determined whether the inclusion was cracked, and measured its size. These two maps gave us detailed information about the interplay between inclusion size and its tendency to crack. We also used energy-dispersive spectroscopy in the scanning electron microscope to determine the composition of the inclusions. One concern was that sample preparation caused inclusion cracking. We expect that the particle cracking seen in our experiments was not an artifact of the sample-preparation procedure for two reasons. First, for some samples tested to low loads, no particle cracking was observed (Table I). These samples were cut and polished in the same way as those where particle cracking was seen. Second, in the samples where particle cracking occurred in the notch region, no particle cracking was observed outside the notch region.

Table I. Notch Radius B0 (mm) 4.0 4.0 4.0 4.0 4.0 2.0 2.0 2.0 2.0 2.0 2.0 1.0 1.0 1.0 1.0

Specimens Tested Peak Load (kN) 60 68 72 72 70 (fractured) 50 60 70 78 80 80.76 (fractured) 70 80 88 90 Particles Cracked yes yes yes yes yes no no yes yes yes yes no no yes yes

III.

NUMERICAL CALCULATION PROCEDURE

Quasi-static calculations are carried out using a finitedeformation Lagrangian formulation with the initial unstressed state taken as a reference, as described, for example, by Needleman,[51] Tvergaard,[52] and Shabrov and Needleman,[53] where more complete descriptions of the formulation and further references can be found. The material is characterized as an isotropic-hardening elastic-viscoplastic solid. The elastic strain is assumed to remain small, and the total rate of deformation (D) is written as the sum of an elastic part (De) and a plastic part (D p), with D e given by a rate form of Hookes law and # 3e Dp s [1] 2se where is the stress deviator. In Cartesian tensor notation, s ij sij sh dij [2]

1746VOLUME 35A, JUNE 2004

METALLURGICAL AND MATERIALS TRANSACTIONS A

with ij being the Kronecker delta and with 3 1 s2 s ij s ij and sh skk e 2 3 [3]

Here, e is the effective stress and h is the hydrostatic tension. With the elastic strains small, ij is the Cauchy (true) stress. The elastic response is assumed to be isotropic, characterized by the Youngs modulus# (E ) and Poissons ratio (). The effective plastic strain rate (e) in Eq. [1] is given by
N se 1/m e # # e 0 a b and g(e) s0 a 1 b [4] 0 g(e) # and e e dt. The function g(e) represents the effective stress vs effective strain response # # in a tensile test carried out at a strain rate such that e 0. Also, 0 0.01 0 /E, 0 is a reference strength, and N and m are the strain-hardening exponent and strain ratehardening exponent, respectively. A schematic of the specimen geometry analyzed is shown in Figure 2(a). In all cases, the initial length and radius of the specimens, respectively, are specified by L0 57.15 mm and R0 6.35 mm. The internal diameter at the minimum section is di. Specimens with notch radii (B0) of 1, 2, and 4 mm were analyzed. For comparison with displacements measured in the experiments, the gage length 2G0, where 2G0 25.4 mm, is also shown in Figure 2(a). A reference cylindrical coordinate system (r, , z) is used, and attention is confined to axisymmetric deformations so that all field quantities are independent of . Moreover, symmetry about z 0 is assumed. In addition to the symmetry conditions imposed at z 0 and # traction-free lateral surfaces, a uniform displacement rate (U ) is prescribed along # z L0 that gives rise to a constant nominal strain rate (). Hence, # # # [5] uz U (L0 U ) on z 0 # In all calculations here, is taken to be equal to the reference # strain rate 0. The region analyzed numerically is meshed by quadrilateral finite elements. Each quadrilateral element is made up of four crossed linear-displacement triangular subelements. Figure 2(b) shows the finite-element mesh used for the 2 mm notched specimen. This mesh consists of 1008 quadrilateral elements and 1080 nodes. The discretizations for the 1 and 4 mm notched specimens have similar resolutions, and the near-notch meshes for all three specimen geometries are shown in Figure 3. Numerical convergence studies were carried out for all three notch geometries. For example, for

(a)

(b)

Fig. 2Circumferentially notched cylindrical specimens: (a) axisymmetric notched cylinder geometry analyzed numerically and (b) finite-element mesh used for the 2-mm notch specimen.

(a)

(b)

(c)

Fig. 3Finite-element meshes in the notch region: (a) 1-mm notch specimen, (b) 2-mm notch specimen, and (c) 4-mm notch specimen.
METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 35A, JUNE 20041747

in the matrix around the inclusion or second-phase particle was found to precede void nucleation, an early criterion was that void nucleation was associated with a critical value of the accumulated plastic strain. For monotonic loading and with a characterization of the plastic-flow properties as in Eq. [4], this is equivalent to a critical value of the matrix flow strength. This criterion can be expressed as se N [6]

where N is the nucleation stress. The work of Argon and Im[34] showed that a superposed hydrostatic stress contributed to the onset of void nucleation, and they proposed a criterion that can be expressed as se sh sa N [7]

Fig. 4A fit of the numerical force-displacement response curve to the experimentally determined one. Uniaxial tension of a notch-free cylindrical specimen. The W is the value of the axial displacement at the gage length 2G0.

The rationale for this criterion is that the stress across the matrix/particle interface increases in the presence of a hydrostatic tension. In a subsequent micromechanical analysis of void nucleation from a rigid particle in an elastic-viscoplastic matrix, Needleman[40] suggested a modification of the Argon and Im[34] criterion that has the form se csh sn N [8]

the 1 mm notch, meshes of 1048, 4256, and 9624 elements were used, and the maximum load levels obtained using these meshes were 95.98, 95.86, and 95.84 kN, respectively. Calculations were also carried out for smooth cylindrical tensile specimens in order to obtain values of the elastic and the plastic material parameters. Based on these calculations, the values E 200 GPa, 0.3, 0 1340 MPa, N 0.045, m 0.02, 0 0.01 0/E 6.7 105, and # 0 0.002 s1 were chosen and are the values used in the analyses of the notched specimens. Figure 4 shows a comparison between the computed and two measured (labeled t555 and t561) force vs deflection curves for smooth cylindrical tensile specimens. As in Figure 2(a), 2G0 is the gage length and w is the average displacement at the gage length. The calculated peak stress in the smooth specimen (which occurs in the center of the specimen after the onset of necking) is below the nucleation stress for w/G0 less than a value between 0.06 and 0.09, depending on the precise value of the nucleation stress. Hence, the properties chosen are representative of the undamaged material in the regime before the onset of particle fracture. IV. VOID-NUCLEATION CRITERIA

where c 1. The analyses in Reference 40 indicated that only some fraction of a far-field hydrostatic tension was effective in increasing the stress acting across the matrix/ particle interface. The value of c depends on the particle volume fraction, morphology, and the material properties.[40,53] The criteria [6] and [7] are limiting cases of criterion [8]; [6] corresponds to c 0 and [7] to c 1. A criterion that has been used to model cleavage fracture is that the maximum principal stress reaches a critical value, i.e., s1 N [9] One rationale for this criterion is that it accounts for the possibility that there may be an initial crack in the brittle particle or inclusion (perhaps due to processing) that is unfavorably oriented with respect to the stress state. The opening stress-intensity factor for this crack is proportional to the maximum principal stress. There is, in general, no simple relationship between Eqs. [9] and [8]. However, to illustrate the connection, consider an axisymmetric stress state corresponding to 1 , 2 3 , where for the purposes of discussion we will take 0. Then, from Eq. [3], e |1 | and h (1 2)/3. The maximum principal stress criterion is that N when 1 and N when 1. The Criterion [8] becomes

In structural metals, void nucleation primarily occurs either by debonding or cracking of second-phase particles and/or inclusions. In phenomenological theories of ductile fracture, e.g., those of Gurson[54] and Rousselier,[55] the voids are represented in terms of a single parameter, the void volume fraction. The evolution equation for the void volume fraction can include a term that accounts for the increase in void volume fraction due to void nucleation. The form of the void-nucleation criterion can have a strong effect on the predicted stability of plastic flow.[56] Phenomenological void-nucleation criteria have mainly been proposed for void nucleation by inclusion debonding, e.g., References 34, 39, and 40. Since plastic deformation
1748VOLUME 35A, JUNE 2004

1 |1 r| 3 (1 2r) 2

s N

[10]

For 0, uniaxial tension, the term in brackets in Eq. [10] is 1 c/3, whereas for 1, a state of hydrostatic tension, this term is c. For 3, the term in brackets in Eq. [10] is 2 7c/3. For example, with c 0.6, as used subsequently, and 3, the maximum principal stress criterion predicts nucleation at 3 N, whereas the criterion in Eq. [8]
METALLURGICAL AND MATERIALS TRANSACTIONS A

predicts nucleation at 3.4 N. Such an axisymmetric stress state is attained in the central section of the notched specimens (near r 0, z 0), but not closer to the notch surfaces. Both n and 1 were calculated along z 0 for all three notch geometries at various load levels ranging from 60 to 80 kN. The values of n and 1 were within 10 pct of each other, with the greatest deviation occurring for the 4 mm notch, i.e., the lowest triaxiality. Also, the peak value of n tended to occur slightly closer to the notch surface than the peak value of 1. The difference between the various nucleation criteria depends strongly on the stress triaxiality, the ratio h/e. For the notched bars in the present investigation, the stress triaxiality varies throughout the specimen and, at a given material point, varies during the deformation history. To give some representative values, for the 1 mm notched bar, in the region near where n is maximum and at loads in the range where void nucleation occurs, h/e 1.0 to 1.2; for the 2 mm notched bar and for the 4 mm notched bar, the corresponding representative values of h/e are 0.85 to 0.95 and 0.75 to 0.85, respectively. Although there is really no firm micromechanical basis that any of these four criteria appropriately characterize void nucleation by the cracking of second-phase particles or inclusions, we will explore their effectiveness in rationalizing the dependence of void nucleation by particle cracking on the triaxiality of the stress state, the ratio of h/e, as seen in the experiments. V. RESULTS

section of one of these inclusions is also seen in Figure 6(a). These nitrides would crack during the tensile test, as shown in Figure 6(b). We, thus, carefully documented the mechanical conditions under which this cracking would occur. In the remaining discussion, references to inclusions refer only to these nitrides. Table I lists all of the mechanical tests that we performed. As described in Section II, three different notch radii were examined, which gave rise to different stress states at the notch. For each notch geometry we pulled one sample to failure and then pulled the others to loads below the failure load. Each of these unfailed samples were sectioned, and the location of cracked inclusions was determined. From Table I, we can observe that the fracture stress decreased with increasing notch radius. We also found from

A micrograph of the steel used in this study is shown in Figure 5. The basic microstructure is a tempered martensite, and two types of inclusions were found. One set were elongated MnS inclusions, an example of which is shown in Figure 6(a). We did not observe any evidence of debonding of these inclusions or their participation in the ductile-fracture process. We assume that this result occurred because the sulfides were parallel to the tensile axis. The other set of inclusions were the cuboidal titanium nitrides. A triangular

(a)

(b)
Fig. 6(a) An optical micrograph showing the two types of inclusions observed in this study. The elongated inclusions are MnS and the triangular inclusion is titanium nitride. (b) A titanium nitride inclusion that was cracked as a result of a tensile test.
VOLUME 35A, JUNE 20041749

Fig. 5The tempered martensite microstructure of the steel used for this study.
METALLURGICAL AND MATERIALS TRANSACTIONS A

examination of the sections that, initially, we would not detect any cracked inclusions, but that, over a load range of approximately 10 kN, the sample would change from one in which no inclusions were cracked to one in which almost all inclusions were cracked. For example, for the 1 mm notch, at 70 and 80 kN loads, no inclusions were cracked, whereas at 90 kN, almost all were. Also, as the notch radius increased, we obtained more cracked inclusions for a given load. To make a detailed comparison of the location of the cracked inclusions and the stress state in the notch, we made maps of the inclusion location and the stress. These maps required that we select a proper description of the stress state, and we now describe this selection. Calculations were carried out for the three specimen geometries used in the experiments. The good agreement obtained between the experimental and computed loaddisplacement curve with the material properties chosen was shown in Briant et al.[57] Since the calculations give stress and deformation values at each point in the specimen, the predictability of various phenomenological void-nucleation criteria can be explored. It is important to note that the quantities obtained from the

calculations presume an isotropic classical plastic solid, thus ignoring local effects such as grain texture, stress concentrations due to any elastic mismatch between the inclusion and matrix, and the influence of organized dislocation structures. Figures 7 and 8 show distributions near the minimum section of the field quantity governing void nucleation according to the four criteria in Section IV for the notches with B0 1 mm and B0 4 mm, respectively. Overlayed on the contour plots are symbols showing the locations of cracked and uncracked inclusions identified in the experiments. The contour plots are shown for a specimen cross section. Due to the assumed symmetries, the region analyzed is only onequarter of that shown. The measured inclusion distributions do not satisfy the symmetries assumed in the calculations. In Figures 7(d) and 8(d), n is calculated with c 0.60. This value of c was chosen to optimize the fit to the observed distribution of cracked inclusions. Micromechanical analyses of void nucleation by debonding suggest values of c in the range of 0.25 to 0.45.[40,53] For brevity, corresponding plots for the intermediate notch acuity B0 2 mm are not shown. The plot in Figure 7(a) shows that there are many cracked particles in the center of the specimen, where the effective stress

(a)

(b)

(c)

(d )

Fig. 7Plots of the stress around the notch in a stressed bar. For all of these plots, the notch radius was 1 mm and the load was 90 kN. Black squares indicate locations of inclusions that had cracked and open squares indicate locations of inclusions that had not cracked. (a) The contours indicate the effective stress. (b) The contours indicate the maximum principal stress. (c) The contours indicate the nucleation stress, defined in Ref. 40, with c 1.0. (d ) The contours indicate the nucleation stress, defined in Ref. 40, with c 0.6.
1750VOLUME 35A, JUNE 2004 METALLURGICAL AND MATERIALS TRANSACTIONS A

(a)

(b)

(c)

(d)

Fig. 8Plots of the stress around the notch in a stressed bar. For all of these plots, the notch radius was 4 mm and the load was 72 kN. Black squares indicate locations of inclusions that had cracked and open squares indicate locations of inclusions that had not cracked. (a) The contours indicate the effective stress. (b) The contours indicate the maximum principal stress. (c) The contours indicate the nucleation stress, defined in Ref. 40, with c 1.0. (d) The contours indicate the nucleation stress, defined in Ref. 40, with c 0.6.

is low (the correlation of cracked particles with e for the B0 2 mm notch is similar to that seen in Figure 7(a)). A value of N 2.4 GPa shows a reasonable correlation with the distribution of cracked particles for the three remaining potential void-nucleation quantities. Figure 7 shows the near-notch region for the 1 mm notch radius at 90 kN. As mentioned previously, inclusions had not cracked in the samples loaded to 70 and 80 kN. As seen in Figure 7, the majority of the inclusions examined in the near-notch region were cracked at 90 kN. For this notch radius, there is a region in the center of the specimen where the value of n is smaller than it is nearer to the notch surfaces. Nearly all the cracked inclusions occur in the region where the value of n exceeds 2.6 GPa. However, a few cracked inclusions fall within the central region where N is between 2.3 GPa and 2.6 GPa. In addition, not all inclusions in the region where n exceeds 2.6 GPa have cracked. Clearly, there is a distribution of inclusion strengths. At a load of 60 kN for the specimen with B0 4 mm, there were eight cracked inclusions in the near-notch region. The location of these inclusions did not correlate with high
METALLURGICAL AND MATERIALS TRANSACTIONS A

values of n. Contours of n are shown in Figure 8(d) for the 4 mm notch-radius specimen at a load of 72 kN. At this load, nearly all of the cracked inclusions are in the region where n exceeds 2.3 GPa, and, indeed, most are near the center of the specimen where n is above 2.6 GPa. The results described previously make several important points about ductile fracture in steel. The first is that there appears to be a critical value of n of approximately 2.3 GPa, above which it is likely that inclusions will crack. If the stress ahead of the notch does not reach this value, a few random inclusions may crack, but one does not observe the large number of cracked inclusions that are represented on the maps in Figures 7 and 8. The quantity n in Eq. [8], with c 0.6, was found to give a somewhat better correlation with the distribution of cracked particles than either a in Eq. [7] or 1 in Eq. [9]. This is not surprising considering that n contains an adjustable parameter, whereas neither a nor 1 do. In the following text, n with c 0.6 will be taken as the stress measure governing void nucleation. At least qualitatively, for the range of stress states considered, there would not be a large difference if a criterion using a or 1 were chosen.
VOLUME 35A, JUNE 20041751

With the choice of n as the quantity governing nucleation, we can now further interpret our results. Let us first compare maps from samples with a 1 mm notch radius pulled to 70, 80, and 90 kN. These are shown in Figure 9. We can see that cracked nitrides were only observed in the sample

pulled to 90 kN, and these cracked inclusions lay in the region where n was above 2.3 GPa. Figures 9(a) and 10 show examples of each notch radius taken from samples loaded to 70 kN. No cracked inclusions were observed in the sample that had a notch radius of 1 mm, six cracked inclusions were recorded for the sample that had a notch radius of 2 mm, and a significant number of inclusions had cracked in the sample that had a notch radius of 4 mm. Two samples were investigated in much greater detail. For these samples, every inclusion in the upper half of the sample cross section was recorded and its size determined. We will report the size as the maximum dimension of the cuboidal TiN inclusions. The first sample that we consider is one that had a 4 mm notch radius and was pulled to 72 kN. This load was very near the fracture load. Of the 189 inclusions counted, 146 were cracked and 43 were uncracked. Figure 11 shows a plot of the number of cracked and

(a)

(a) (b)

(c)
Fig. 9Contour plots of the nucleation stress, n, in samples with a 1-mm notch radius pulled to different loads. The contours were calculated using Needleman criterion[40] with c 0.6. Black squares indicate locations of inclusions that had cracked, and open squares indicate locations of inclusions that had not cracked. The loads were (a) F 70 kN, (b) F 80 kN, and (c) F 90 kN.
1752VOLUME 35A, JUNE 2004

(b)
Fig. 10Contour plots of the nucleation stress, n, in samples with different notch radii pulled to a load of 70 kN. The contours were calculated using Needleman criterion[40] with c 0.6. Black squares indicate locations of inclusions that had cracked, and open squares indicate locations of inclusions that had not cracked. The radii were (a) 2 mm and (b) 4 mm.
METALLURGICAL AND MATERIALS TRANSACTIONS A

uncracked inclusions plotted as a function of size. Of particular interest was the size of the uncracked inclusions in this sample. We see that all inclusions with a maximum dimension greater than 15 m were cracked and, in general, the number of uncracked inclusions observed increased with decreasing size. Even more instructive was to plot the location of the uncracked inclusions having various sizes. This result is shown in Figure 12. We see that the smallest uncracked inclusions lie near the centerline of the sample; all larger inclusions in that region had cracked. As one moves farther away from the centerline, inclusions of large size

remain uncracked. The small uncracked inclusions lie in the region with the highest value of n. A similar study was carried out for a sample that had a 4 mm notch and had been loaded to 60 kN. In this sample, 122 inclusions were counted; 15 were cracked and 107 were uncracked. As can be seen in Figure 13, the sizes of the cracked inclusions covered the same range as the uncracked inclusions. However, by looking at the location as well as the size of these inclusions, we again found a correlation between size and location. Figure 14 shows a map of the cracked inclusions, but with different inclusion sizes indicated by different symbols.

Fig. 11The number of cracked and uncracked titanium nitride inclusions, N, plotted as a function of the maximum dimension of the inclusion, S. This sample had a 4-mm notch radius and had been loaded to a peak load of 72 kN. Note that there is a population of cracked inclusions with sizes larger than 15 m, whereas there are no uncracked inclusions with S greater than 15 m.

Fig. 13The number of cracked and uncracked titanium nitride inclusions, N, plotted as a function of the maximum dimension of the inclusion, S, for a sample with a radius of 4 mm that had been loaded to 60 kN. Note that there are many more uncracked inclusions than cracked inclusions for this sample, but that the cracked inclusions span the same size range as the uncracked inclusions.

Fig. 12Contour plots of the nucleation stress, n, defined in Ref. 40, with c 0.6 for a sample with a 4-mm notch radius loaded to 72 kN. Included on the plot are the location of uncracked titanium nitride inclusions. The symbols indicate the size of the inclusion. Note that in the region of highest stress, the uncracked inclusions tend to be in the smallest size category. Large inclusions that remain uncracked are in regions of lower stress.
METALLURGICAL AND MATERIALS TRANSACTIONS A

Fig. 14Contour plots of the nucleation stress, n, defined in Ref. 40, with c 0.6 for a sample with a 4-mm notch radius loaded to 60 kN. Included on the plot are the location of cracked titanium nitride inclusions. The symbols indicate the size of the inclusion.
VOLUME 35A, JUNE 20041753

Only one inclusion that had a diameter below 7 mm had fractured, and it was in the region of highest n. The other cracked inclusions were not in this region of high n, but were larger.

VI.

DISCUSSION

Previous studies of ductile fracture in steels have tended to emphasize the role that MnS inclusions play. The picture of the ductile-fracture process that has emerged is that voids form at the sulfide-steel interface at very low strains as a result of separation of this interface. These microvoids grow and often connect immediately before final fracture through the formation of sheets of fine voids that nucleate from the carbides, e.g., Senior et al.[59] In contrast, in our study of low-alloy steels, the primary nucleation site for microvoid nucleation was found to be cracking of titanium nitride inclusions. Furthermore, there was no significant participation of the sulfides in the fracture process. Ductile fracture not involving the primary sulfide particles has been seen previously at crack tips in very clean steels. However, in those cases, the sulfide spacing was large compared to the size scale of the high-stress region at the crack tip. In the notched bars here, the length scale associated with the high-stress region was 1 to 2 mm, whereas the sulfides are distributed along rows parallel to the tensile axis, which are roughly 200 to 500 m apart. Thus, even though the high-stress region will, in general, encompass several sulfide particles, the sulfide particles were never observed to nucleate voids. Possible reasons that the sulfides did not play a role in the fracture process include (1) their alignment with the tensile axis, which affected the load that the sulfides carried; and (2) the chemistry of the sulfides was such as to result in a strong bond with the matrix, as in References 37 and 38. Microvoid nucleation around cracked particles is very common in aluminum alloys,[50] where large iron- and siliconcontaining inclusions can be easily fractured. High-nickel maraging steels also fracture by microvoid coalescence, with the voids forming on cracked titanium carbo-nitride inclusions.[60] The latter material is analogous to the titaniummodified 4330 steel studied here. The most complete study of this fracture process in maraging steels has been provided by Cox and Low,[30] who considered the growth of microvoids in smooth tensile bars. Several of their observations are very similar to ours. They did not observe any cracking of titanium carbo-nitrides until they reached a tensile stress of approximately 1.4 GPa. The number of microvoids increased until most inclusions had cracked and formed a microvoid at a stress of approximately 2.1 GPa. They found that the largest inclusions cracked first and that, as the stress increased, smaller inclusions progressively cracked. They also noted that the average cross-sectional void area did not increase substantially until very near fracture, at which point it grew rapidly. These results on maraging steels parallel the results that we obtained on titanium-modified 4330 steel and help us establish a description of ductile fracture in iron-based alloys, where the microvoids form around cracked inclusions. It appears that a significant load must be applied before any inclusions crack, and that the largest inclusions will crack at the lowest loads. As the load increases above this initial value many more inclusions crack, and the load range over which the material goes from having no inclusions cracked to having
1754VOLUME 35A, JUNE 2004

the majority cracked is relatively narrow (10 kN), which corresponds to a nucleation-stress range of 200 to 300 MPa. These voids then grow and connect to provide the fracture path. It is worth noting in this regard that analyses of the localization of deformation in porous plastic solids.[56,61] suggest that a narrow range of nucleation-stress levels, coupled with stressdependent void nucleation, can precipitate a localization instability in plane-strain deformation states. The susceptibility to localization is much reduced in axisymmetric deformation states, as in the notched bars used in this investigation. Many microstructural studies of ductile fracture have focused on a qualitative description of the fracture process. Although extremely useful, such descriptions are not predictive. A main aim of our investigation has been to provide a predictive void-nucleation criterion for particle fracture that can be used in phenomenological ductile-fracture models such as those of Gurson[54] and Rousselier.[55] Several assumptions and limitations of our analyses should be noted. We assume that the local stress state at a nucleation site is appropriately characterized by an analysis based on characterizing the material as a homogeneous isotropic-hardening solid. Also, although the experimental results show a clear effect of particle size on void nucleation, with larger particles cracking more easily than smaller particles, none of the criteria considered account for this size dependence directly, although, of course, the value of the critical nucleation stress can be taken as a function of inclusion size. The three notched bars used in the experiments provided data on stress triaxialities ranging from about 0.75 to about 1.2. Comparisons between the predications of several potential void-nucleation criteria and the experimental results showed that characterization of nucleation in terms of a critical effective stress, or, equivalently, a critical effective plastic strain, is not consistent with the data. The criterion proposed in Reference 40, where the nucleation stress has the form of a linear combination of the effective stress and the hydrostatic tension, with the proportion each contributes being an adjustable parameter, provided a reasonably good characterization of void nucleation by particle cracking. This criterion was obtained from a micromechanical analysis of debonding, and the value of the adjustable parameter that fit the data in this study is significantly different from the value from the micromechanical analyses of debonding. Whether there is any physical basis for the parameter value chosen or whether the correlation found here is simply data fitting remains to be determined. Also, for the range of stress states in the notched bars used in the experiments, the deviation between the voidnucleation stress in Eq. [8] and the maximum principal stress in Eq. [9] was only 10 pct. Hence, there are not sufficient data to choose between these two alternatives. However, the voidnucleation stress is easier to compute and fits more naturally into ductile-fracture frameworks such as those of Gurson[54] and Rousselier.[55] Thus, the combination of experiment and analysis in our study has provided a nucleation criterion for microvoids formed around cracked particles that can be used in predictive analyses of ductile fracture.

VII.

CONCLUSIONS

1. Experiments on notched tensile bars show that in the titanium-modified 4330 steel investigated, microvoids
METALLURGICAL AND MATERIALS TRANSACTIONS A

that lead to ductile fracture nucleate from cracked titanium nitrides. 2. Cracking of the titanium-nitride particles occurred over a relatively narrow load (and stress) range, with larger inclusions tending to crack at lower loads (and stresses). 3. We examined various void-nucleation criteria and found that the onset of nucleation was well correlated in terms of a nucleation stress that is a weighted sum of the hydrostatic tension, h, and the effective stress, e. 4. With this void-nucleation criterion, we found N 2.3 to 2.4 GPa to be the value of the nucleation stress in the titanium-modified 4330 steel. ACKNOWLEDGMENTS We are pleased to acknowledge support from the Materials Research Science and Engineering Center on Micro- and Nano-Mechanics of Electronic and Structural Materials at Brown University (NSF Grant No. DMR0079964) and from Caterpillar Inc. We thank Mr. Kenneth Burris for valuable discussions. REFERENCES
1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. C.F. Tipper: Metallurgica, 1949, vol. 39, pp. 133-37. K.E. Puttick: Phil. Mag., 1959, vol. 4, pp. 964-69. H.C. Rogers: Trans. TMS-AIME, 1960, vol. 281, pp. 498-506. C.D. Beachem: Trans. ASM, 1963, vol. 56, pp. 318-26. J. Gurland and J. Plateau: Trans. ASM, 1963, vol. 56, pp. 441-54. F.A. McClintock: J. Appl. Mech., 1968, vol. 35, pp. 363-71. J.R. Rice and D.M. Tracey: J. Mech. Phys. Solids, 1969, vol. 17, pp. 201-17. S.H. Goods and L.M. Brown: Acta Metall., 1979, vol. 27, pp. 1-15. W.M. Garrison and N.R. Moody: J. Phys. Chem. Solids, 1987, vol. 48, pp. 1035-74. V. Tvergaard: Adv. Appl. Mech., 1990, vol. 27, pp. 83-151. J.W. Hancock: in Topics in Fracture and Fatigue, A.S. Argon, ed., Springer-Verlag, New York, NY, 1992, pp. 99-144. A. Needleman, V. Tvergaard, and J.W. Hutchinson: in Topics in Fracture and Fatigue, A.S. Argon, ed., Springer-Verlag, New York, NY, 1992, pp. 145-78. R.M. McMeeking: in Topics in Fracture and Fatigue, A.S. Argon, ed., Springer-Verlag, New York, NY, 1992, pp. 179-96. A. Pineau: in Topics in Fracture and Fatigue, A.S. Argon, ed., SpringerVerlag, New York, NY, 1992, pp. 197-234. M. Gologanu, J.B. Leblond, G. Perrin, and J. Devaux: Int. J. Solids Struct., 2001, vol. 38, pp. 5581-94. A.A. Benzerga: J. Mech. Phys. Solids, 2002, vol. 50, pp. 1331-62. T. Pardoen and J.W. Hutchinson: Acta Mater., 2003, vol. 51, pp. 133-48. M.F. Horstemeyer, S. Ramaswamy, and M. Negrete: Mech. Mater., 2003, vol. 35, pp. 675-87. C.I.A. Thomson, M.J. Worswick, A.K. Pilkey, and D.J. Lloyd: J. Mech. Phys. Solids, 2003, vol. 51, pp. 127-46. R.O. Ritchie and R.M. Horn: Metall. Trans. A, 1978, vol. 9A, pp. 331-41. F.M. Beremin: Metall. Trans. A, 1981, vol. 12A, pp. 723-31. G. Le Roy, J.D. Embury, G. Edward, and M.F. Ashby: Acta Metall., 1981, vol. 29, pp. 1509-22.

23. J.L. Maloney and W.M. Garrison, Jr.: Scripta Metall., 1989, vol. 23, pp. 2097-100. 24. S. Lee, L. Majno, and R.J. Asaro: Metall. Trans. A, 1985, vol. 16A, pp. 1633-48. 25. R.O. Ritchie and A.W. Thompson: Metall. Trans. A, 1985, vol. 16A, pp. 233-48. 26. D.M. Goto, D.A. Koss, and V. Jablokov: Metall. Mater. Trans. A, 1999, vol. 30A, pp. 2835-42. 27. G. Green and J.F. Knott: J. Eng. Mech. Technol., 1976, vol. 98, pp. 37-46. 28. J.Q. Clayton and J.F. Knott: Met. Sci., 1976, vol. 10, pp. 63-71. 29. D.A. Curry and P.L. Pratt: Mater. Sci. Eng., 1979, vol. 37, pp. 223-35. 30. J.B. Cox and J.R. Low: Metall. Trans., 1974, vol. 5, pp. 1457-70. 31. P.E. Magnusen, E.M. Dubensky, and D.A. Koss: Acta Metall., 1988, vol. 36, pp. 1503-09. 32. A.S. Argon, J. Im, and A. Needleman: Metall. Trans., 1975, vol. 6, pp. 815-24. 33. A.S. Argon, J. Im, and R. Safoglu: Metall. Trans., 1975, vol. 6, pp. 825-37. 34. A.S. Argon and J. Im: Metall. Trans., 1975, vol. 6, pp. 839-51. 35. D. Brooksbank and K.W. Andrews: J. Iron Steel Inst., 1968, vol. 206, pp. 595-99. 36. C.L. Briant and S.K. Banerji: Metall. Trans. A, 1982, vol. 13A, pp. 827-36. 37. J.W. Bray, J.L. Maloney, K.S. Raghavan, and W.M. Garrison, Jr.: Metall. Trans. A, 1991, vol. 22A, pp. 2277-85. 38. L.E. Iorio and W.M. Garrison, Jr.: Scripta Mater., 2002, vol. 46, pp. 863-68 39. J.R. Fisher and J. Gurland: Met. Sci., 1981, vol 15, pp. 185-92. 40. A. Needleman: J. Appl. Mech., 1987, vol. 54, pp. 525-31. 41. C.I.A. Thomson, M.J. Worswick, A.K. Pilkey, D.J. Lloyd, and G. Burger: J. Mech. Phys. Solids, 1999, vol. 47, pp. 1-26. 42. M.F. Horstemeyer and A.M. Gokhale: Int. J. Solids Struct., 1999, vol. 36, pp. 5029-55. 43. D. Steglich, T. Siegmund, and W. Brocks: Comp. Mater. Sci., 1999, vol. 16, 404-13. 44. J.E. Neimark: Sc.D. Thesis, MIT, Cambridge, MA, 1959. 45. J.W. Hancock and A.C. Mackenzie: J. Mech. Phys. Solids, 1977, vol. 14, pp. 147-69. 46. A.C. Mackenzie, J.W. Hancock, and D.K. Brown: Eng. Fract. Mech., 1977, vol. 9, pp. 167-88. 47. F.M. Beremin: Metall. Trans. A, 1983, vol. 24A, pp. 2272-87. 48. R. Becker, A. Needleman, O. Richmond, and V. Tvergaard: J. Mech. Phys. Solids, 1988, vol. 36, pp. 317-51. 49. M.F. Horstemeyer, J. Lathrop, A.M. Gokhale, and M. Dighe: Theor. Appl. Frac. Mech., 2000, vol. 33, pp. 31-47. 50. R.H. Van Stone and T.B. Cox: in Fractography-Microscopic Cracking Processes, ASTMSTP 600, C.D. Beacham and W.R. Warke, eds., ASTM, Philadelphia, PA, 1976, pp. 5-29. 51. A. Needleman: J. Appl. Mech., 1972, vol. 20, pp. 111-20. 52. V. Tvergaard: J. Mech. Phys. Solids, 1976, vol. 24, pp. 291-304. 53. M.N. Shabrov and A. Needleman: Model. Simul. Mater. Sci. Eng., 2002, vol. 10, pp. 163-83. 54. A.L. Gurson: Ph.D. Thesis, Brown University, Providence, RI, 1975. 55. G. Rousselier: Nucl. Eng. Design, 1987, vol. 105, pp. 97-111. 56. A. Needleman and J.R. Rice: in Mechanics of Sheet Metal Forming, D.P. Koistinen and N.-M. Wang, eds., Plenum Press, New York, NY, 1978, pp. 237-65. 57. C.L. Briant, E. Sylven, M.N. Shabrov, D.H. Sherman, L. Chuzhoy, and A. Needleman: Mechanisms and Mechanics of Fracture, Symp. in Honor of Professor J.F. Knott, W.O. Soboyejo, J.J. Lewandowski, and R.O. Ritchie, ed., TMS, Warrendale, PA, 2002, pp. 169-74. 58. J.E. King and J.F. Knott: Met. Sci., 1981, vol. 15, pp. 1-6. 59. B.A. Senior, F.W. Noble, and B.L. Eyre: Acta Metall., 1986, vol. 34, pp. 1321-27. 60. S. Floreen and H.W. Haydern: Met. Sci. J., 1970, vol. 4, pp 77-80. 61. M. Saje, J. Pan, and A. Needleman: Int. J. Fract., 1982, vol. 19, pp. 163-82.

METALLURGICAL AND MATERIALS TRANSACTIONS A

VOLUME 35A, JUNE 20041755

Potrebbero piacerti anche